Vous êtes sur la page 1sur 12

Spatial variations of DCMD performance for desalination

through countercurrent hollow fiber modules


Li-Hua Cheng, Ping-Chung Wu, Cho-Kai Kong, Junghui Chen*
R&D Center for Membrane Technology and Department of Chemical Engineering,
Chung Yuan University, Chungli 320, Taiwan
Tel. 886(3)2654107; Fax 886(3)2654199; email: jason@wavenet.cycu.edu.tw
Received 1 July 2007; accepted revised 25 September 2007
Abstract
To understand the spatial variations of the desalination behavior, a mathematical model is systematically
formulated to simulate a direct contact membrane distillation (DCMD) with the hollow fiber module for desalina-
tion. The model equation for the entire module is derived by integrating the mass, momentum and energy balances
on both feed and permeate sides with the permeate flux across the membrane. The property variations of feed and
permeate sides along the length of the membrane module are thoroughly simulated. Sensitivity of the permeate
flux and the thermal efficiency along the fiber length to operating conditions is further investigated over a range of
temperature and flow rate. It is shown that both permeate flux and thermal efficiency decrease along the fiber
length as the feed gets cooled down. Although the velocity variations of both streams are not big, the flow rate
ratio of the hot feed to the cold solution has a dramatic impact on the distribution of the permeate flux and the
thermal efficiency along the fiber length. The effect analysis could potentially be applied to optimal design and
scale-up of hollow fiber DCMD modules.
Keywords: Direct contact membrane distillation; Hollow fiber module; Membrane distillation; Modeling;
Simulation
1. Introduction
Membrane distillation (MD) is a thermally
driven process, in which only vapor molecules
are transported through porous hydrophobic
membranes. The potential advantages of the MD
process in comparison with the conventional
processes lie in the lower operating temperature
and the hydrostatic pressure, insensitivity to
salt concentration, along with the merits in
utilizing low-grade waste and/or alternative energy
sources [1]. Although MD was introduced as a
*Corresponding author.
Presented at the Fourth Conference of Aseanian Membrane Society (AMS 4), 1618 August 2007, Taipei,
Taiwan.
Desalination 234 (2008) 323334
0011-9164/08/$ See front matter # 2008 Published by Elsevier B.V.
doi:10.1016/j.desal.2007.09.101
water desalination process in 1963, it is not
industrially implemented on any large scale
plants even today. Besides the development of
membrane, the commercialization of the MD
process will, to a large extent, depend upon the
appropriate design of module. Compared with
plate and frame modules, hollow fiber-based
membrane devices for MD are more likely to
be of considerable utility because they are
simple, potentially scalable, and they can often
pack a large membrane surface area per unit
device volume without any need for a support-
ing structure [2]. Hollow fiber module also
shows the least temperature polarization among
various types of membrane modules [3]. Some
research on MD has employed hollow fiber
membrane modules. For example, Gryta and
Tomaszewska presented and verified the heat
transport in MD capillary modules [4]. Lagana
et al. developed a model of tubular membrane
(with relative large fiber diameter) for effect
evaluation of membrane morphology, such as
thickness, elastic modules and pore radius distri-
bution on permeate flux [5]. Hollow fibers of
relatively larger wall thickness, considerable
porosity and increased internal diameter were
preferred for the performance improvement of
direct contact membrane distillation (DCMD) [6].
Preliminary experimental MD investigations
on a small scale using hollow fiber membranes
illustrated by Li and Sirkar showed a remarkably
high water vapor flux up to 79 kg/m
2
h high
module productivity [2].
Although past research on MD has adopted
hollow fiber module, modeling of hollow fiber
MD modules has not been seriously and system-
atically studied due to the complex nature of
the MD process itself and the flow through fiber
modules. The primary goal of currently available
MD models is to predict the values of the
permeate flux other than the estimation of the
temperature and the concentration polarization
coefficients [1]. El-Bourawi et al. pointed out
the issue that only a mean temperature and a
mean concentration were considered when
modeling MD processes [1]. This will increase
the risk of membrane pore wetting because of
the hydrostatic pressure drop along the length
of the membrane module. To simulate a hollow
fiber MD module, three sub-models are required.
Two sub-models are used to describe the flow on
each side of the membrane, and the third model
characterizes the separative properties of the
membrane. A general approach to modeling
membrane modules has been reported [7];
however, the case study on MD has not been
involved. For the commonly adopted type of the
hollow fiber DCMD module, spatial variations
of parameters, like temperature, velocity and
concentration, etc. along the fiber length are of
considerable importance for modeling and its
scale-up, which has not been thoroughly studied
to our best knowledge. Furthermore, thermal
efficiency, which depicts the fraction of energy
used for the vaporization of water, is also an
important factor for the evaluation of the MD
process with the non-isothermal characteristics.
It can be used as a criterion for improvement.
In this paper, systematical model formulations
of DCMD for desalination through countercur-
rent hollow fiber modules are derived from
rigorous mass, momentum and energy balances
of both the feed and permeate sides coupled to
the simultaneous mass and heat transfer across
the membrane. The simulations can help us esti-
mate the axial variation of properties, such as the
feed and permeate flow rates, retentate and
permeate temperatures, hydrostatic pressure, flux
and thermal efficiency, etc. An attempt can be
carried out to further understand the sensitivity
of the process efficiency to operating conditions
along the fiber length.
2. Hollow fiber DCMD module model
Hollow fiber modules contain a large number
of membrane fibers housed in a module shell.
Feed can be introduced on either the fiber or
324 L.-H. Cheng et al. / Desalination 234 (2008) 323334
the shell side and permeate is withdrawn in a
co-current or a countercurrent, where the latter
is generally more effective [8]. A schematic
representation of the flow pattern in a hollow
fiber module with the fiber side feed and the
countercurrent permeate stream is illustrated in
Fig. 1. The zoom of the fiber shown at the bottom
of Fig. 1 describes the transport of the water vapor
from the feed side to the permeate side through
the porous and hydrophobic membrane. The cor-
responding variables are labeled on the module.
To describe the hollow fiber module transport
behavior, the transport equations are derived for
the retentate as well as for the permeate side of the
module as the flux depends on both of the feed
and permeate side conditions. For the simulation
with reasonable computational expense, the fol-
lowing assumptions for the transport equations
are made to describe the flow on each side of the
membrane: (i) steady incompressible flow, (ii)
lubricating approximation at the feed side, (iii)
negligible axial diffusion compared to convection,
(iv) negligible channeling when the uniform flow
exists and (v) negligible heat loss to environment.
2.1. Feed side in tube
On the basis of the momentum, mass and
energy balances in the feed side, the following
equations in terms of pressure (P
F
), velocity
(v
F
), salt compositions (x
F
s
) and temperature
(T
F
) are derived
dP
F
dz

32j
F
d
2
i
v
F
1
1
V
F
dv
F
dz

v
F
V
F

2
M
s
,
s

M
w
,
w
_ _
dx
F
s
dz

4Jd
o
M
F
d
2
i
2
x
F
s
V
F
dv
F
dz

v
F
M
w
,
w
V
F

2
dx
F
s
dz
0 3
d,
F
v
F
C
F
p
T
F
dz

4Q
F
:Nd
2
i
4
where d
i
and d
o
are the inside and outside
diameters of the fiber, respectively. C
F
p
is the spe-
cific heat of the feed. ,
F
and ,
w
are the densities
of feed and water. V
F
is the feed molar volume.
J is the mass flux. M
s
, M
w
and M
F
are the mole-
cular weights of salt, water and feed. j
F
is the
viscosity of the feed.
2.2. Permeate side in shell
In the permeate side, the differential equa-
tions in terms of pressure (P
P
), velocity (v
P
) and
temperature (T
P
) are derived
dP
p
dz

32j
P
d
2
h
v
p
5
dv
P
dz

4V
P
NJd
o
M
P
d
2
s
Nd
2
o
_ _
6
d,
P
v
p
C
P
p
T
p
dz

Q
P
d
2
s
Nd
2
o
7
The composition term is neglected due to the
nearly 100% rejection of non-volatile ionic
r
z
v
F

m
d
i d
o
v
P
d
s
Retentate Feed inlet
Cooling water
Permeate
Permeate
Permeate
Retentate
Fig. 1. Hollow fiber module with tube side feed flow.
L.-H. Cheng et al. / Desalination 234 (2008) 323334 325
solutes in DCMD of desalination. d
h
is the
hydraulic diameter of the shell, which is a func-
tion of the packing density of the module c,
d
h

1c
c
d
o
and c N
d
o
d
s
_ _
2
. d
s
is the inside
diameter of the hollow module shell. N is the
number of fibers. All the other variables are
identical to those variables in the feed side, but
the superscript is for the permeate side.
2.3. Flux across the membrane
The above two side modules are completely
independent except for the coupling terms which
describe energy flux (Q) and mass flux (J)
through the membrane simultaneously occurring
in the radial direction shown in Fig. 2, where T
F
m
and T
P
m
are temperatures at both sides of the
membrane, respectively.
2.3.1. Heat transfer
The heat transfer process of DCMD includes
the heat transferred through the boundary of the
feed side (Q
F
) and the permeate side (Q
P
), and
across the membrane (Q
m
).
(i) Within the boundary of the tube feed side:
Q
F
h
F
A
F
r
c T
F
T
F
m
_ _
8
where A
F
r
1 and c N:d
i
. The heat transfer
coefficient in the tube side h
F
may be described
in a number of ways. Only one of them is chosen
here because of its relative simple formand wide
adoption [4]:
Nu
F

h
F
d
i
k
F

1.62Re
F

0.33
Pr
F

0.33
d
i
L
_ _
0.33
Re
F
< 2100
0.023Re
F

0.8
Pr
F

0.33
Re
F
! 2100
_

_
9
where Re
F

d
i
v
F
,
F
j
F
and Pr
F

C
F
p
j
F
k
F
.
(ii) Across the membrane:
Q
m
A
m
r
c JH
k
m
c
m
T
F
m
T
P
m
_ _
_ _
10
where A
m
r
d
lm
,d
i
and d
lm
is the log-mean
radius difference of the fiber. H is the enthalpy
of evaporation which is determined based on the
mean temperature of feed and permeate side
temperatures, T
F
m
T
P
m
_ _
,2 [9]. c
m
is the mem-
brane thickness. k
m
is the thermal conductivity.
k
m
k
mg
1 k
ms
, is the porosity and
k
mg
and k
ms
refer to the conductive coefficients
of air within the membrane pore and solid
membrane, respectively.
(iii) Within the boundary of the shell cold side:
Q
P
h
P
A
P
r
c T
P
m
T
P
_ _
11
where A
P
r
d
o
,d
i
and d
o
is the outside diameter
of the fiber. The heat transfer coefficient in the
shell side of the hollow fiber module h
P
could
be described by correlation similar to Eq. (9),

m
T
F
P
F
T
P
P
P
T
F
P
F
T
P
P
P
m
m
m
m
Membrane
Vapor
Vapor
Vapor
Fig. 2. Mass and heat transfers during MD.
326 L.-H. Cheng et al. / Desalination 234 (2008) 323334
but the heat transfer refers to the hydraulic
diameter of shell d
h
[10].
According to the energy conservation, at
steady state, the amount of heat transferred
through the above three phases is equal,
Q
F
Q
m
Q
P
12
2.3.2. Mass transfer
The mass transfer process in DCMD has two
steps. The volatile component passes through
the concentration boundary on the feed side of the
membrane and then through the porous mem-
brane itself. Since the mean free molecular path
of saturated water vapor under typical DCMD
operating conditions is comparable to the typical
membrane pore size used in MD, the Knudsen
diffusion Molecular diffusion Poiseuille
flowtransition [11] is chosen among the numerous
expressions of mass transfer of MD.
J R
1
K
R
1
M
_ _
1
R
1
P
_ _
P
F
m
P
P
m
_ _
13
where P
F
m
and P
P
m
are the vapor pressures on
the feed and permeate sides of the membrane.
The difference of both pressures is the driv-
ing force of the MD process. R
1
K
C
K
M
w
RT
m
_ _
0.5
,
R
1
M
C
M
DM
w
P
aM
RT
m
_ _
and R
1
P
C
P
M
w
P
m
jRT
m
_ _
. C
K
,
C
M
, and C
P
represent the individual contribu-
tion of Knudsen diffusion, molecular diffusion
and Poiseuille flow, respectively. The diffusion
coefficient D depends on temperature and pres-
sure [11]. Water vapor pressures are calculated
using the Antoine equation. The diffusion models
presented above have to be modified to account
for the reduction in vapor pressure caused by
dissolved species. For dilute salt solutions,
Raoults law can be used,
P
F
m
P
sat
1 x
F
s
_ _
14
where P
sat
is the saturated vapor pressure of
water at T
F
m
. Based on the relationship of the
mass and heat transfers through membrane,
the temperatures (T
F
m
and T
P
m
) and the flux (J) are
determined by solving the non-linear least
squares optimization problem using MATLAB.
2.4. Solution of model
The hollow fiber module parameters used for
the simulation are listed in Table 1. For full infor-
mation, please refer to [11]. The coupled ordinary
differential equations, including the flow on each
side of the membrane and the characterization of
hydrophobic membrane, are solved numerically
with the following boundary conditions
For the feed side
T0 T
F
in
. v0 v
F
in
. x
F
s
0 x
s.in
.
P
F
L P
0
15
For the permeate side
TL T
P
in
. vL v
P
in
. x
P
s
0 0.
P
p
0 P
0
16
P
0
is equal to the ambient atmospheric pressure
at both outlets of the streams. In order to apply
the numerical method to the operation in a
Table 1
Characteristics of the membrane module
Hollow fiber membrane module PVDF
Length of fibers (L, m) 0.34
Shell diameter (d
s
, m) 0.03
Number of fibers (N) 3000
Inner diameter of fibers (d
i
, mm) 0.30
Membrane thickness (c
m
, mm) 60
Nominal pore diameter (mm) 0.2
Packing density (c, %) 60
Porosity (, %) 75
L.-H. Cheng et al. / Desalination 234 (2008) 323334 327
countercurrent manner, the outlet condition for the
permeate side on a trial basis should be estimated.
It has to be found out if its entrance condition in
Eq. (16) is satisfied. If not, make another guess and
try again. The simulated system is repeated until
the entrance condition is satisfied.
3. Results and discussion
3.1. Spatial variations
3.1.1. Temperature variation
The variations of T
F
, T
P
, T
F
m
and T
P
m
as func-
tions of the fiber length are shown in Fig. 3,
where the corresponding operation condition is
added to the caption. v
F
in
and v
P
in
are the volume
flow rates, and w
F
in
is the mass fraction of salt
in the feed inlet. All these temperatures change
almost linearly along the fiber length. T
F
decreases from the inlet 343 K to outlet
312.5 K, while T
P
increases from the inlet
298 K to outlet 330 K. Because of the counter-
current flow fashion, nearly constant transmem-
brane temperature difference is maintained
throughout the membrane module. It can be seen
that with the temperature dropping by 30 K at
the feed side, the transmembrane temperature
ranges 36 K, whereas the temperature at the
permeate side increases by 30 K.
Fig. 3 also shows the obvious temperature
polarization as the transmembrane temperature
difference is almost equal to those at the bound-
aries of the feed side and the permeate side. It
might be alleviated by increasing the flow rates
of both feed and permeate.
3.1.2. Velocity variation
Fig. 4 shows the velocity variation of both
retentate and permeate along the fiber length
axial position. Because of the countercurrent flow
fashion and the cumulative changes of permeate
flux along the fiber length, the feed side velocity
v
F
decreases fromthe inlet 0.472 m/s to the outlet
0.455 m/s, while the permeate side velocity v
P
increases from the inlet 0.332 m/s to the outlet
0.343 m/s. The relative small velocity variation
at the permeate side is attributed to the larger
cross-sectional area when the packing density
is 0.6. The velocity variations of both retentate
and permeate are not significantly big enough;
however, they have great effect on the trans-
membrane hydrostatic pressure.
3.1.3. Hydrostatic pressure variation
The bulk pressure along the length of module
is a function of both fluid velocity and fiber
length as expressed in Eqs. (1) and (5). Fig. 5
0 0.05 0.1 0.15 0.2 0.25 0.3
300
305
310
315
320
325
330
335
340
345
z, m
T
,

K
T
F
T
P
T
F
T
P
m
m
Fig. 3. Temperature as a function of the fiber length (T
F
in

343 K, T
P
in
298 K, v
F
in
v
P
in
6 L/min, w
F
in
0.025.
0.45
0.46
0.47
0.48
z, m
0 0.05 0.1 0.15 0.2 0.25 0.3
0.33
0.335
0.34
0.345
v
P
,

m
/
s
v
F
,

m
/
s
v
F
v
P
Fig. 4. Velocity as a function of the fiber length (T
F
in

343 K, T
P
in
298 K, v
F
in
v
P
in
6 L/min, w
F
in
0.025).
328 L.-H. Cheng et al. / Desalination 234 (2008) 323334
shows the local hydrostatic pressure drop of both
feed and permeate sides along the fiber length.
When both outlets of retentate and permeate
are required to be the ambient atmospheric pres-
sure, the inlet feed pressure reaches 131 kPa
while the inlet permeate pressure is 125 kPa.
The relatively lower pressure drop at the perme-
ate side is attributed to the countercurrent flow
fashion, the cumulative permeate flux from the
feed side to the permeate side as well as the
relative larger cross-sectional area at the per-
meate side with packing density of 0.6. Nazish
et al. also reported the downstream pressure
build-up in case of the hollow fiber module for
pervaporation [12].
Variation of hydrostatic pressure has no
influence on flux since the DCMD driving force
is the function of the vapor pressure difference
between both membrane sides. However, it will
increase the risk of membrane pore wetting
because membranes used in MD requires being
non-wetting [1]. For a typical PVDF membrane
contact angle of 130

, the penetration pressure


of a cylindrical of the maximum pore diameter
1 mm is only 185 kPa. When longer fiber, 1.2 m,
is adopted, for instance, liquid entry pressure
of 185 kPa would have been reached when both
inlet flow rates are maintained at 6 L/min. The
effect of pressure drop may be reduced by
increasing the fiber diameter or shortening
the fiber length at the cost of high membrane
area packing density. Therefore, there is a
trade-off between the module volume and
attained separation, showing the necessity to
optimize the fiber dimensions for the hollow
fiber DCMD module.
3.1.4. Permeate flux and thermal
efficiency variation
As DCMD is a non-isothermal process, the
properties of interests, the permeate flux Jz
and the thermal efficiency jz along the length
of fiber should be computed locally. The thermal
efficiency is defined as
jz
JzHTz
JzHTz
k
m
c
m
T
F
m
z T
P
m
z
17
which depicts the fraction of energy that has
been used for the vaporization of water.
Fig. 6 shows the variations of Jz and jz
with the fiber length. Both of them decrease
0 0.05 0.1 0.15 0.2 0.25 0.3
100
105
110
115
120
125
130
135
z, m
P
,

k
P
a
P
F
P
P
Fig. 5. Hydrostatic pressure as a function of the fiber
length (T
F
in
343 K, T
P
in
298 K, v
F
in
v
P
in
6 L/min,
w
F
in
0.025).
0 0.05 0.1 0.15 0.2 0.25 0.3
0
5
10
15
20
J
,

k
g
/
m
2

h
z, m
0
0.2
0.4
0.6
0.8

Fig. 6. Permeate flux and thermal efficiency as a


function of the fiber length (T
F
in
343 K, T
P
in
298 K,
v
F
in
v
P
in
6 L/min, w
F
in
0.025).
L.-H. Cheng et al. / Desalination 234 (2008) 323334 329
along the fiber length because DCMD driving
force comes from the vapor pressure difference
between both membrane sides. Although the
transmembrane temperature difference keeps
almost constant as shown in Fig. 3, the vapor
pressure difference has significantly changed
since the vapor pressure with the Antoine
Equation is an exponential depending on tem-
perature. It can be seen that J(z) decreases from
the inlet 18 to the outlet 3 kg/m
2
h, which is
corresponding to transmembrane difference of
3 and 6 K as shown in Fig. 3, respectively.
Because of the almost constant transmembrane
temperature difference of 36 K, the conductive
heat transfer keeps nearly unchanged. Therefore,
with the decrease in Jz and the linear decrease
of H with the temperature along the fiber
length, it is expected that jz decreases linearly
according to Eq. (17) as shown in Fig. 6.
3.1.5. Concentration variation
The salinity of the retentate increases parabo-
lically along the fiber length as shown in Fig. 7,
in which w
F
S
is the mass fraction of salt in the
feed side. It is also attributed to the permeate
flux changed along the fiber. Nevertheless, the
change in salinity is very small because of single
pass through the fiber.
3.2. Effect of operating variables
3.2.1. Effect of temperature
In Fig. 8(a), the spatial variations of Jz
change significantly with T
F
in
, because the expo-
nential increase of the vapor pressure of the feed
0 0.05 0.1 0.15 0.2 0.25 0.3
0.025
0.0252
0.0254
0.0256
0.0258
0.026
0.0262
z, m
w
F
s
Fig. 7. Retentate salinity as a function of the fiber
length (T
F
in
343 K, T
P
in
298 K, v
F
in
v
P
in
6 L/min,
w
F
in
0.025).
(a)
0 0.05 0.1 0.15 0.2 0.25 0.3
0
5
10
15
20
25
30
z, m
J
,

k
g
/
m
2

h
T
F
in
T
F
in
=
313 K
323 K
333 K
343 K
353 K
(b)
0 0.05 0.1 0.15 0.2 0.25 0.3
0
0.2
0.4
0.6
0.8
1
z, m

=
313 K
323 K
333 K
343 K
353 K
Fig. 8. The effect of the inlet temperature of the feed
solution on the spatial variations of (a) the permeate
flux and (b) the thermal efficiency (T
P
in
298 K, v
F
in

v
P
in
6 L/min).
330 L.-H. Cheng et al. / Desalination 234 (2008) 323334
solution with temperature would increase the
transmembrane vapor pressure difference as all
the other involved parameters are maintained
invariables. It is consistent with the general
conclusion that the permeate flux increases as
T
F
in
increases when the whole fiber length is
investigated. The higher jz also occurs at
the front of the fiber length as T
F
in
increases
(Fig. 8(b)). This indicates that it is not worthy
to increase the fiber length without consider-
ing jz.
As the feed temperature is kept constant and
lower T
P
in
is adopted in Fig. 9(a), the transmem-
brane vapor pressure difference at larger T
T
F
m
T
P
m
and lower T
F
in
would be the same as the
one at smaller T and higher T
F
in
in the case of
higher T
P
in
. It is because the transmembrane
vapor pressure based on the Antoine equation
has the exponential change in T
F
in
. Therefore,
lowering T
P
in
may not increase permeate flux
when the whole fiber length is investigated.
Nevertheless, it decreases thermal efficiency as
shown in Fig. 9(b). Hence, there is a trade-off
between the permeate flux and the thermal effi-
ciency as the permeate temperature increases.
3.2.2. Effect of flow rate
In Fig. 10, the spatial variations of Jz and
jz change dramatically along the fiber length
at different v
F
in
when v
P
in
remains constant. It also
indicates that Jz and jz increase at the
higher feed flow rate with the increase of v
F
in
.
The effect of feed velocity can increase the heat
transfer coefficient and then reduce the tempera-
ture and the concentration polarization effect at
both boundary sides of bulk streams. On the other
hand, because of the shorter retention time of the
stream within the hollow fiber module, the higher
transmembrane difference is maintained along
the fiber length. Thus, this can increase the process
driving force and yield higher permeate flux.
The effect of stream velocity on the spatial
variations of Jz and jz demonstrates that the
ratio of the hot feed to the cold solution plays an
important role in the distribution of transmem-
brane temperature difference along the fiber
length. When the ratio is lower than 1, the effec-
tive fiber length reduces as shown in Fig. 10(a).
For example, Jz decreases from the inlet
22.5 kg/m
2
h to almost 0 at the fiber length of
0.15 m at the feed flow rate of 2 L/min and
(a)
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
0
5
10
15
20
25
z, m
J
,

k
g
/
m
2

h
=
290 K
293 K
298 K
308 K
318 K
(b)
0 0.05 0.1 0.15 0.2 0.25 0.3
0
0.2
0.4
0.6
0.8
1
z, m

=
290 K
293 K
298 K
308 K
318 K
T
P
in
T
P
in
Fig. 9. The effect of the inlet temperature of cold water
on the spatial variations of (a) the permeate flux and
(b) the thermal efficiency (T
F
in
343 K, v
F
in
v
P
in

6 L/min).
L.-H. Cheng et al. / Desalination 234 (2008) 323334 331
the permeate flow rate of 6 L/min. With the
increase of the ratio, the permeate flux drop along
the fiber length decreases gradually. When the
ratio increases to more than 1, the Jz distribu-
tion along the fiber length becomes even as
shown in Fig. 10 (a), indicating the full utilization
of the membrane area of the process.
On the contrary, although both Jz and jz
change along the fiber length at different v
P
in
when v
F
in
is maintained constant as shown in
Fig. 11, the distribution of Jz could be chan-
ged drastically throughout the whole fiber
length. When the flow rate of the hot feed is
much greater than that of the cold solution, more
permeate flux occurs at the end of the fiber
length. This is because the larger transmem-
brane temperature difference is maintained near
the inlet of permeate, and the driving force
reduces with the rapid temperature rise of
(a)
0 0.05 0.1 0.15 0.2 0.25 0.3
0
5
10
15
20
25
z, m
J
,

k
g
/
m
2

h
v
F
in
=
2 L/min
4 L/min
6 L/min
8 L/min
10 L/min
(b)
0 0.05 0.1 0.15 0.2 0.25 0.3
0
0.2
0.4
0.6
0.8
1
z, m

=
2 L/min
4 L/min
6 L/min
8 L/min
10 L/min
v
F
in
Fig. 10. Spatial variations of the permeate flux (a) and
the thermal efficiency (b) as a function of the inlet flow
rate of the feed solution (v
P
in
6 L/min, T
F
in
343 K,
T
P
in
298 K).
(a)
0 0.05 0.1 0.15 0.2 0.25 0.3
0
5
10
15
20
25
30
z, m
=
2 L/min
2.5 L/min
3 L/min
4 L/min
6 L/min
8 L/min
10 L/min
(b)
0 0.05 0.1 0.15 0.2 0.25 0.3
0
0.2
0.4
0.6
0.8
z, m

=
2 L/min
2.5 L/min
3 L/min
4 L/min
6 L/min
8 L/min
10 L/min
J
,

k
g
/
m
2

h
v
P
in
v
P
in
Fig. 11. Spatial variations of the permeate flux (a) and
thermal efficiency (b) as a function of the inlet flow
rate of cold water (v
F
in
6 L/min, T
F
in
343 K, T
P
in

298 K).
332 L.-H. Cheng et al. / Desalination 234 (2008) 323334
permeate along the fiber length. Similarly, when
the flow rate of the hot feed is much lower than
that of the cold solution, the lower Jz is
covered at the end of the fiber length. Therefore,
to obtain higher productivity, the membrane
area of the module should be fully utilized and
the flow rate ratio of the hot feed and the cold
solution should be appropriately adjusted.
4. Conclusions
The detailed DCMD model is developed by
integrating the permeate flux through the mem-
brane with rigorous mass, momentum and energy
balance on both feed and permeate sides. In this
paper, the use of the model is demonstrated to
investigate complex spatial property variations
of feed and permeate sides along the fiber length
in a desalination system during DCMD.
Even if nearly constant transmembrane tem-
perature difference is maintained in the counter-
current operation, both permeate flux and thermal
efficiency decrease along the fiber length as the
feed gets cooled down or the permeate is heated
up. Although the velocity variations of both
streams are not great, the risk of membrane
pore wetting increases dramatically because of
the hydrostatic pressure build-up. Although the
operating condition in DCMD processes can be
adjusted any time, it is difficult to fully under-
stand how well the process is operated. Further-
more, the simulated DCMD results depicting the
flow rate ratio of the hot feed to the cold solution
has a dramatic effect on distribution of permeate
flux and thermal efficiency along the fiber
length. With the increase of the ratio, the effective
fiber length increases, resulting in the increase of
both the flux and the thermal efficiency.
As shown in this paper, the detailed model is
essential to the design of the DCMD module
with the trade-off between the permeate flux and
the thermal efficiency as the permeate condition
varies. It is also necessary to optimize fiber
dimensions and operating conditions so that the
performance can be optimized in the hollow
fiber DCMD system for the desalination. More-
over, the channeling problem exists in the
practical application when the longer fiber is
applied. The model to be modified for optimal
design of DCMD systems will be investigated
in our future work.
Acknowledgements
This work was sponsored by the Center-
of-Excellence Program on Membrane Technol-
ogy, the Ministry of Education, Taiwan, R.O.C.
References
[1] M.S. El-Bourawi, Z. Ding, R. Ma and M. Khayet,
A framework for better understanding membrane
distillation separation process, J. Membr. Sci.,
285 (2006) 429.
[2] B. Li and K.K. Sirkar, Novel membrane and
device for direct contact membrane distillation-
based desalination process, Ind. Eng. Chem. Res.,
43 (2004) 53005309.
[3] R.W. Schofield, A.G. Fane and C.J.D. Fell, Heat
and mass transfer in membrane distillation,
J. Membr. Sci., 33 (1987) 299313.
[4] M. Gryta and M. Tomaszewska, Heat transport in
the membrane distillation process, J. Membr. Sci.,
144 (1998) 211222.
[5] F. Lagana, G. Barbieri and E. Drioli, Direct
contact membrane distillation: modeling and
concentration experiments, J. Membr. Sci., 166
(2000) 111.
[6] J. Gilron, L. Song and K.K. Sirkar, Design for
cascade of crossflow direct contact membrane
distillation, Ind. Eng. Chem. Res., 46 (2007)
23242334.
[7] J.I. Marriott and E. Sorensen, A general approach
to modeling membrane modules, Chem. Eng. Sci.,
58 (2003) 49754990.
[8] J.I. Marriott, E. Sorensen and I.D.L. Bogle,
Detailed mathematical modeling of membrane
modules, Comput. Chem. Eng., 25 (2001)
693700.
L.-H. Cheng et al. / Desalination 234 (2008) 323334 333
[9] J.M. Smith, H.C. Van Ness and M.M. Abbott,
Introduction to chemical engineering ther-
modynamics, seventh edition, McGraw-Hill,
Boston, 2005.
[10] R. Thanedgunbaworn, R. Jiraratananon and
M.H. Nguyen, Shell-side mass transfer of hollow
fibre modules in osmotic distillation process,
J. Membr. Sci., 290 (2007) 105113.
[11] Z. Ding, Study on the module performance and
dynamic process of membrane distillation, PhD
Thesis, University of Chemical Technology,
P.R. China, 2001.
[12] N. Nazish, S.V. Suggala and P.K. Bhattacharya,
Pervaporation of hydrazine-water through hollow
fiber module: modeling and simulation, Comput.
Chem. Eng., 30 (2005) 202214.
334 L.-H. Cheng et al. / Desalination 234 (2008) 323334

Vous aimerez peut-être aussi