Anda di halaman 1dari 12

Journal of Membrane Science 348 (2010) 298309

Contents lists available at ScienceDirect

Journal of Membrane Science


journal homepage: www.elsevier.com/locate/memsci

Effect of draw solution concentration and operating conditions on forward


osmosis and pressure retarded osmosis performance in a spiral wound module
Yuan Xu a,1 , Xiaoyu Peng a,1 , Chuyang Y. Tang b,c, , Q. Shiang Fu d, , Shengzhe Nie a
a

Environmental and Clean Technology Laboratory, Suzhou Institute of Sichuan University, Suzhou, China
School of Civil and Environmental Engineering, Nanyang Technological University, Singapore
Singapore Membrane Technology Centre, Nanyang Technological University, Singapore
d
Nanjing Institute of Geography and Limnology, Chinese Academy of Sciences, Nanjing, China
b
c

a r t i c l e

i n f o

Article history:
Received 22 July 2009
Received in revised form 8 November 2009
Accepted 9 November 2009
Available online 13 November 2009
Keywords:
Forward osmosis (FO)
Pressure retarded osmosis (PRO)
Reverse osmosis (RO)
Internal concentration polarization (ICP)
Osmotic driving force

a b s t r a c t
Forward osmosis (FO) and pressure retarded osmosis (PRO) are concentration-driven membrane processes. While they can be potentially used in water, wastewater, and energy applications, these processes
suffer from the concentration polarization inside the porous membrane support resulting in severe ux
decrease, a phenomenon known as internal concentration polarization (ICP). Researchers have investigated the effect of ICP both in theoretical and experimental studies. The current study extends the
existing ICP model to include the effect of draw solution dilution by membrane permeate ow in a spiral
wound FO module (SWFO). FO and PRO experiments were performed using a Hydrowell SWFO under
both submerged and cross-ow conditions. The effect of draw solution concentration, draw solution ow
rate, feed water ow rate, and membrane orientation on FO and PRO water ux performance was systematically investigated. Permeate ow increased with greater draw solution concentration in both FO and
PRO modes. ICP was found to drastically limit the available membrane ux in the concentration-driven
membrane processes, and its adverse effect was more severe at greater draw solution concentration.
Membrane ux was also affected by the dilution of draw solution when the permeate ow rate was comparable or greater than the draw solution ow rate. The submerged FO conguration performed nearly
as good as the cross-ow conguration with feed water circulating outside of the membrane envelope
(shorter ow path). In this case, the feed water ow rate only had limited effect on membrane ux likely
due to its low mass transfer resistance. In contrary, the membrane ux can be adversely affected at low
feed water ow rate when it was circulated inside of the membrane envelope (longer ow path).
2009 Elsevier B.V. All rights reserved.

1. Introduction
Membrane separation has gained increasing popularity in
water, wastewater, and many other industrial applications.
Traditionally, pressure-driven membrane processes, such as microltration (MF), ultraltration (UF), nanoltration (NF), and reverse
osmosis (RO), have received much attention. For example, lowpressure porous MF and UF membranes are widely used for surface
water treatment [1,2], in membrane bioreactors (MBRs) [3,4], as
well as for pretreatment of RO processes [5]. In parallel, the
application of reverse osmosis (RO) is fuelled by the increasing

Corresponding author at: Department of Civil and Environmental Engineering,


Nanyang Technological University, Blk N1, Rm #1b-35, Singapore 639798.
Tel.: +65 6790 5267; fax: +65 6791 0676.
Corresponding author.
E-mail addresses: cytang@ntu.edu.sg (C.Y. Tang), fu shiang@hotmail.com
(Q.S. Fu).
1
These author contributed equally to this study.
0376-7388/$ see front matter 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.memsci.2009.11.013

needs for seawater desalination, water treatment, and wastewater reclamation [5,6], along with the signicant improvements of
RO membrane properties and module design [6,7]. On the other
hand, although their full-scale implementation is still lacking, there
is also increased interest in concentration-driven forward osmosis
(FO) and pressure retarded osmosis (PRO) processes for water and
energy applications due to their potential for energy production (or
saving) [810].
During FO, a high concentration solution (draw solution) is separated from a low concentration solution by a semi-permeable
(water permeable but not solute permeable) membrane (Fig. 1(a)).
Due to the osmotic pressure difference across the membrane, water
ows spontaneously through the membrane from the low concentration side to the draw solution side. This water permeating
through the FO membrane is of high quality, with nearly complete
retention of organic matter, particulates, and microorganisms as
well as signicant retention of dissolved salts [1013]. Where a natural source of osmotic energy is available (e.g., seawater) [8,9,14],
FO can be highly attractive since the process requires very little additional energy input (except a small amount of energy is

Y. Xu et al. / Journal of Membrane Science 348 (2010) 298309

299

Fig. 1. Illustration of FO, PRO, and RO processes. (a) FO process where no pressure is applied on the high concentration solution. Water ows from low concentration side
to high concentration side. (b) PRO process where applied pressure on the high concentration solution is less than the osmotic pressure difference across the membrane.
Water ows from low concentration side to high concentration side. (c) RO process where applied pressure on the high concentration solution is greater than the osmotic
pressure difference across the membrane. Water ows from high concentration side to low concentration side. (d) Classication of FO, PRO, and RO in a ux versus pressure
plot. Adapted from Refs. [8,10].

Fig. 2. Modeling internal concentration polarization. (a) Active layer is facing feed water. Draw solution concentration in the support layer is diluted by the water permeate
ux, causing the available driving force (the concentration difference across the rejection layer) to be much lesser than the apparent driving force (concentration difference
between the draw solution and the feed water). (b) Active layer is facing draw solution. The solute concentration at the rejection-support layer interface is much higher than
the feed concentration, leading to reduced available driving force. Adapted from [8,10].

300

Y. Xu et al. / Journal of Membrane Science 348 (2010) 298309

required for recirculating the feedwater and draw solution) [10].


In other cases, a concentrated brine may be used as draw solution,
but the diluted brine after the FO process has to be regenerated to
complete a closed loop for the draw solution (e.g., via ltration or
thermal processes) [15,16], which can be highly energy intensive.
The PRO process is similar to the FO process, except an additional back pressure is applied on the draw solution (Fig. 1(b)).
Water ows from low concentration side to the high concentration side as long as the applied pressure is lower than the osmotic
pressure difference across the membrane [10]. When the applied
pressure is greater than the osmotic pressure difference, however,
the direction of water ux is reversed, which is the well known
reverse osmosis condition (Fig. 1(c)). PRO, along with RO and FO,
is summarized in a ux versus pressure plot as shown in Fig. 1(d).
Unlike RO desalination where mechanical energy (pumping) is used
to overcome the osmotic pressure of seawater, the PRO process
converts the chemical potential (osmotic energy) of a concentrated
draw solution into mechanical energy [8,17,18]. Loeb and coworkers [9,14,17,19] proposed the idea of PRO based osmotic power
plant for electricity generation by running the pressurized permeate ow through a turbine generator. A Norwegian company in the
clean energy sector, Statkraft, has built the worlds rst prototype
seawater osmotic power plant [20,21]. It is planning to commercialize osmotic power by year 2015 with a projected energy cost
competitive against other renewable energy sources such as solar
energy, offshore wind power, and hydroelectric power [20,21].
Despite the many potential applications [10], concentrationdriven FO and PRO processes are severely limited by internal
concentration polarization (ICP) in the porous membrane support,
i.e.: (i) the accumulation of rejected solute from low concentration feed when the dense active layer is oriented towards the draw
solution (i.e., support facing the feed water) or (ii) the dilution of
draw solution concentration when the active layer is facing the
feed water (Fig. 2 and Refs. [8,22,23]). A direct consequence of
ICP is a reduction of the membrane ux due to the signicantly
reduced concentration gradient across the membrane rejection
lm [23]. This was evident in view of the impractically low water
ux when conventional reverse osmosis membranes were tested
in FO and PRO modes, as a result of their thick membrane support
[8,24,25]. Recently, Hydration Technology Inc. (HTI, Albany, Oregon) has developed a commercial cellulose triacetate membrane
with a thin support (<50 m in thickness), which showed superior
ux performance and less ICP in the FO mode [12,13,25,26]. Many of
the existing studies on the HTI membrane were performed for small
at sheet membrane coupons in the FO mode, while few studies are
available in the PRO mode. To the authors best knowledge, systematic tests on spiral wound modules in both FO and PRO modes are
not available.
The purpose of the current study was to evaluate the effect of
draw solution concentration and operational conditions on the permeate ux in FO and PRO processes. Filtration tests were performed
with a spiral wound HTI membrane module using sodium chloride
brine as a draw solution. The ux performance of the spiral wound
module was also modeled via classical ICP models [8,24]. This study
helps us to understand the role of various operational parameters
on membrane ux performance in both FO and PRO modes.

of the dense rejection layer and the porous support layer (Csupport )
is signicantly lower than the bulk draw solution concentration
(Cdraw ) due to the convective transport of solute away from the
support back to the draw solution (dilutive internal concentration polarization (DICP) [23]). Consequently, the available driving
force (i.e., the concentration difference across the dense membrane
rejection layer) is much smaller than the apparent concentration
difference between the draw solution and the feed [8,10]. Similar
ICP effect has also been reported when the draw solution is oriented towards the dense rejection layer (Fig. 2(b)), where solute
from the feed solution and that transmitted through the membrane from the draw solution are accumulated and concentrated in
the support layer [12,22,27]. In this case, the concentration in the
support (Csupport ) is signicantly higher than the feed water concentration (Cfeed ), which in turn reduces the available driving force
(concentrative internal concentration polarization (CICP) [23]).
The effect of internal concentration polarization can be modeled
by adopting the classical solution-diffusion theory for the dense
rejection layer coupled with convection and diffusion transport of
the solute in the porous support layer [8,22,24]. Following Lee et
al. [8] and Leob et al. [24] (refer to Appendix A), the water ux in
FO can be expressed as

Jv = Km ln

Adraw + B
Afeed + Jv + B

(FO, active layer facing feed water, DICP)


and

Jv = Km ln

Adraw Jv + B
Afeed + B

(FO, active layer facing draw solution, CICP)

(2)

where Jv is the volumetric ux of water; A and B are the transport


coefcients for water and solute, respectively; draw and feed are
the osmotic pressure of the bulk draw solution and that of the bulk
feed water, respectively.
In Eqs. (1) and (2), Km is the mass transfer coefcient, which is
given by the ratio of the solute diffusion coefcient (D) over the
membrane structural parameter (S):
Km =

D
S

(3)

The S parameter is a property of the support structure. Its value


is proportional to the thickness (t) and the tortuosity () of the
support layer, but it is inversely proportional to the porosity ()
of the support [20]. In a way, the S parameter in ICP, which provides a length scale of the concentration polarization in the support
layer, is analogous to the boundary layer thickness in external concentration polarization. A larger structural parameter (e.g., thicker
membrane support) leads to a lower mass transfer coefcient,
which can signicantly reduce permeate ux due to the exponential
dependence of concentration polarization on Km .
Eqs. (1) and (2) may be extended to include the effect of the
applied pressure (P) in the PRO mode [8,28] (refer to the brief
derivation in Appendix A as well):

Jv = Km ln
2. Theory

Adraw + B((AP/Jv ) + 1)
(Afeed + Jv + AP) + B((AP/Jv ) + 1)

(PRO, active layer facing feed water)


A unique feature in osmotically driven membrane processes
(both FO and PRO) is that the direction of water ux is opposite
to the direction of solute ux. This causes internal concentration
polarization in the porous support layer. When the active rejection
layer is facing the feed water (i.e., the porous support facing the
draw solution, Fig. 2(a)), the solute concentration at the interface

(1)


Jv = Km ln

(Adraw AP Jv ) + B((AP/Jv ) + 1)
Afeed + B((AP/Jv ) + 1)

(PRO, active layer facing draw solution)

(4)

(5)

Y. Xu et al. / Journal of Membrane Science 348 (2010) 298309

301

Fig. 3. Experimental setup. (a) An illustration of an open Hydrowell spiral wound module. (b) A cross-ow ltration test setup where the spiral wound membrane module
was housed inside a pressure vessel. Cross-ows can be maintained both inside and outside of the membrane envelope. (c) Different test congurations adopted in the
current study.

For the special case where P = 0, Eqs. (4) and (5) become identical
to Eqs. (1) and (2), respectively.
Eqs. (1), (2), (4), and (5) are valid for small membrane coupons
where the membrane permeate ow (Qp ) is small compared to the
ow rate of draw solution (i.e., the cross-ow of draw solution, Qcf ),
e.g., in batch experiments with low recoveries. In a spiral wound
module, however, the permeate ow can signicantly dilute the
bulk draw solution concentration, which results in reduced ux
performance compared to small coupon tests. Such dilution effect
needs to be explicitly accounted for in a spiral wound module. Consider a module with a draw solution concentration Cdraw,0 and a
cross-ow of Qcf,0 at the module inlet. The bulk draw solution concentration Cdraw,x anywhere inside the module can be determined
by
Cdraw,x = Cdraw,0

Qcf,0
Qcf,0 +

x

(6)

J dAm
0 v

x

where Am is the membrane area, and the term 0 Jv dAm represents


the accumulative permeate ow rate from the module inlet to the
location of interest (x). Thus, the diluted bulk concentration Cdraw,x
is simply related to the original draw solution concentration Cdraw,0

x

via a volumetric dilution factor (Qcf,0 /(Qcf,0 + 0 Jv dAm )). At the outlet of the module, the draw solution is diluted to a concentration

Cdraw,exit :
Cdraw,exit = Cdraw,0

Qcf,0
Qcf,0 + Qp

(7)

where Qp is the total permeate ow rate through the entire membrane module, i.e.:


Qp =

exit

Jv dAm

(8)

The average permeate ux in the module Jv is given by


Jv =

Qp
Am

(9)

Eq. (6) can be used to determine the bulk draw solution concentration Cdraw anywhere in the module, and Cdraw can then be used
to determine the local membrane ux Jv via the FO and/or PRO ux
models (Eqs. (1), (2), (4), or (5)). Thus, Eq. (6), coupled with the
ICP ux models, can be used for ux performance modeling of spiral wound FO/PRO modules. A numerical example is provided in
Appendix A.
For PRO applications, the power output is of interest. The total
power available from the module (PRO ) can be determined as a

302

Y. Xu et al. / Journal of Membrane Science 348 (2010) 298309

function of membrane permeate ow rate [8]:


PRO = PQp

(10)

The power density, that is, the total power normalized by the
membrane area in the module, is given by
PRO = PJv

(11)

cross-ow rate at the feed solution side. A separate pump was used
to circulate the draw solution from a 40-L stainless steel pressure
tank. The pressure inside the draw solution tank was controlled
by adjusting a needle valve located at the top of the tank, and it
was measured by a mechanical pressure gauge (Fig. 3(b)). Where
no back pressure was required in the FO mode, the needle valve
was fully open so that the pressure inside the draw solution tank
reached atmospheric pressure.

3. Materials and methods


3.1. Chemicals
Unless otherwise specied, all reagents and chemicals were of
analytical grade. Deionized water was used throughout the experiments. Analytical grade sodium chloride was used to prepare draw
solutions of various concentrations.
3.2. Membrane material and module conguration
The commercially available Hydrowell membrane elements
were purchased from HTI [29]. This membrane was specially
developed for FO applications with an ultrathin polyester screen
mesh support (comprising of polyester bers 20 m in diameter
arranged orthogonally at a ber-to-ber spacing of 120 m, overall thickness < 50 m) to reduce internal concentration polarization
[10,12]. According to the membrane supplier, the active rejection
layer of the HTI membrane is based on cellulose triacetate, with
a sodium chloride rejection of 9395%. The water permeability
of HTI membrane is 2.2 1012 m/s Pa [12], which is signicantly lower than that for typical brackish water RO membranes
(12 1011 m/s Pa with NaCl rejection > 95%) but comparable to
typical seawater RO membranes (e.g., 3 1012 m/s Pa for SWC4)
[30,31]. Despite of the relatively poorer separation properties of the
HTI membrane in comparison to commercial RO membranes, the
HTI membrane demonstrated much superior ux performance in
the FO mode, which was attributed to the drastically lower ICP due
its special thin support layer design (Section 2 and Refs. [12,25]).
The HTI membrane elements were supplied in the spiral wound
conguration (Fig. 3(a)). Each element comprises of two pieces of
at sheet membranes (225 cm in length and 30 cm in width).
These membrane sheets are to be glued face to face on three
edges to form a membrane envelope (i.e., a membrane leaf). The
remaining edge of the envelope is connected to a central tube with
holes for water collection. The active layers of the membranes are
on the outer sides of the envelope. The membrane leaf is rolled
into a spiral wound conguration, and spacers (diamond pattern,
0.8 mm in thickness and 2.5 mm spacing) are used both inside
and outside of the membrane envelope for maintaining ow channels. Unlike a typical spiral wound RO element, the Hydrowell
element has additional glue lines along the central line of the membranes (Fig. 3(a)), and the center of the water collection tube is
also plugged. This design forces two independent cross-ows: (1)
one within the membrane envelope (along the length of the membrane leaf), and (2) the other one outside the envelope (along the
width of the membrane leaf). Each Hydrowell element has a total
active membrane area of 0.94 m2 (excluding glued membrane
area).
3.3. FO and PRO experiments
3.3.1. Test setup
To allow the application of hydraulic pressure as required for
PRO tests, the Hydrowell element was housed in a pressure vessel self-manufactured with a maximum pressure rating of 30 bar
(Fig. 3(b)). The pressure vessel was connected to a feed water tank
(unpressurized) and a circulation pump was used for controlling the

3.3.2. Modes of testing


Three modes of testing were compared in this study (Fig. 3(c)):
(1) submerged FO mode (s-FO), (2) FO mode with cross-ows for
both feed water and draw solution (x-FO), and (3) PRO mode with
cross-ows for both feed water and pressurized draw solution (xPRO).
In the submerged FO mode (s-FO-in), the pressure vessel was
not used and the Hydrowell module was submerged directly and
completely in the feed water tank. The high-concentration draw
solution was pumped into the central collection tube (Fig. 3(a))
and was circulated inside the membrane envelope. In this conguration, the draw solution was in direct contact with the
porous support layer, while the feed water reached the active
sides of the membranes via the spacer outside of the membrane
envelope. Feed water was drawn through the HTI membrane
under the concentration gradient across membrane. This conguration was somewhat similar to that in a submerged membrane
bioreactor in that no circulation pump was used on the feed
water side. Aeration was not used due to the thin membrane
spacer. The s-FO conguration has been used for water production for emergency uses (military applications, disaster relief, etc.)
[10], as originally advertised for Hydrowell elements by HTI
[29].
In the cross-ow FO mode (x-FO in Fig. 3(c)), the membrane
module was housed inside the pressure vessel although no pressure was applied in the module by fully open the pressure control
valve of the draw solution tank. Both feed water and draw solution were circulated via two circulation pumps (Fig. 3(b)), with the
draw solution either inside the membrane envelope in direct contact with the porous membrane support (x-FO-in conguration) or
outside of the envelope (x-FO-out conguration) in direct contact
with the active membrane layer. The cross-ow PRO mode (x-PRO)
had a similar conguration to that of x-FO, except a back pressure
was exerted on the draw solution side. The x-FO mode may nd
potential applications for water treatment and desalination, while
the x-PRO mode has been suggested for osmotic power harvesting
[10,32,33].
3.3.3. Experimental conditions
Most of the experimental work was performed using a single
SWFO element. However, several SWFO elements were used for
x-PRO-in mode due to the resulting damage of glue lines (refer
to Section 4.3) under the high pressure in this test mode. All
ltration tests were performed at room temperature (2224 C).
Before each experiment, the SWFO module was thoroughly rinsed
and ushed with deionized water on both sides of the membrane envelope. The volumetric ow rate through the Hydrowell
module was determined by measuring the time needed to ll up
a volumetric cylinder (20 ml or 50 ml). The ux was calculated
by normalizing the volumetric ow rate by the effective membrane area according to Eq. (9). Combination of different testing
modes and membrane orientation (s-FO-in, x-FO-in, x-FO-out, xPRO-in, and x-PRO-out, refer to Fig. 3(c)) were evaluated. For
each testing mode, different draw solution concentrations and the
cross-ow velocities were evaluated. In PRO tests, the effect of
applied pressure on membrane ux performance was also monitored.

Y. Xu et al. / Journal of Membrane Science 348 (2010) 298309

303

Fig. 4. Effect of draw solution concentration on permeate ow in s-FO-in conguration. The error bars shown in the gure represent one standard deviation based
on at least three different tests. Note: 1 L/m2 h = 2.78 107 m/s.

4. Results and discussions


4.1. s-FO conguration
Fig. 4 presents the permeate ow rate of the Hydrowell element in the submerged FO mode (s-FO-in) as a function of draw
solution concentration at different circulation rate of the draw
solution (50190 ml/min, corresponding to a cross-ow velocity
of 14 cm/s at the SWFO inlet). Clearly, greater permeate ow
rate was observed at higher draw solution concentration. This was
due to the increased driving force (osmotic pressure difference
in this case) across the membrane. However, unlike the classical solution-diffusion model which predicts that the water ux Jv
in the absence of applied pressure shall be directly proportional
to the osmotic pressure difference, i.e., Jv = A(draw feed ), the
experimental ux was highly non-linear with respect to the draw
solution concentration (Cdraw,0 ). For relatively dilute draw solutions (Cdraw,0 < 0.5 M), signicant ux enhancement was achieved
at increased concentration. A plateau was reached at higher Cdraw,0
(>0.6 M). The maximum permeate ow rate was 100 ml/min, corresponding to a permeate ux of 6 L/m2 h. Similar non-linear
ux behavior has been reported for FO by several research groups
[8,12,13,34], which was attributed to internal concentration polarization in the porous support layer. As explained in Section 2 and
Fig. 2(a), the solute concentration in the support layer (Csupport ) was
signicantly lower than the bulk concentration in the draw solution, which led to a drastic reduction of the available driving force.
This simple ICP model [8,10,12] also predicts, in consistency with
our experimental observation, that:
(1) Flux behavior is less affected by ICP at lower membrane ux
and
(2) At high permeate ux, ICP plays an increasingly dominant role
due to its exponential dependence on ux (Eq. (1)). Under this
condition, any further increase in the draw solution concentration is offset by a much more severe ICP, resulting in less
effective ux enhancement.
In addition to the ICP effect, external concentration polarization (ECP) may also play an important role at higher draw solution
concentrations in s-FO-in mode. This is due to the lack of forced
recirculation at the feed solution side in this mode. The ECP
in s-FO-in mode can be potentially controlled by bubbling (not

Fig. 5. Comparison of ux performances in FO and RO modes. (a) Permeate ux


of RO, FO coupon, and FO module (s-FO-in conguration). Results of RO and FO
coupon were obtained from reference [12]. The s-FO-in experimental results were
obtained at a draw solution cross-ow rate of 100 ml/min. The s-FO-in simulation
results were obtained from Eq. (6) coupled with ICP models (Section 2 and Appendix
A). (b) FO over RO ux ratio as a function of draw solution concentration. Note:
1 L/m2 h = 2.78 107 m/s.

studied in this work), similar to that for submerged membrane


bioreactors.
It is worthwhile to note that the ux achieved in s-FO-in
mode was relatively low (57 L/m2 h for a 0.5 M NaCl draw solution (corresponding to an osmotic pressure difference of 22 bar),
depending on the cross-ow rate). In contrast, signicantly higher
membrane ux can be achieved if the same HTI membrane were
tested in the RO mode (Fig. 5(a) and Ref. [12]). She et al. [12]
reported a water permeability of 2.2 1012 m/s Pa for the HTI
membrane. This corresponds to a ux of 35.6 L/m2 h in the RO
mode at an applied pressure of 45 bar, an order of magnitude larger
than a ux of 6.6 L/m2 h observed in the s-FO-in mode under the
same apparent driving force (1 M draw solution, osmotic pressure = 45 bar). Such comparison clearly conrms that the FO process
performed less efciently compared to RO under the same driving
force, as a result of internal concentration polarization in FO. The
ux efciency of FO may be dene by the ratio of FO ux over RO
ux at identical apparent driving force [34]. This ratio is shown in
Fig. 5(b) as a function of draw solution concentration. The FO ux
efciency was drastically reduced at greater draw solution concentration, which is a clear indication that ICP was more severe under
larger draw solution concentration. This observation is consistent

304

Y. Xu et al. / Journal of Membrane Science 348 (2010) 298309

Fig. 6. Effect of draw solution cross-ow on permeate ow in s-FO-in conguration. Both experimental (solid symbols) and simulation results (continuous lines)
were shown. The s-FO-in simulation results were obtained from Eq. (6) coupled
with ICP models (Section 2 and Appendix A). The error bars shown in the gure represent one standard deviation based on at least three different tests. Note:
1 L/m2 h = 2.78 107 m/s.

with the previous ndings by McGinnis and Elimelech [35], who


reported lower ux efciency at greater ammonia carbonate draw
solution concentrations.
It is also worthwhile to compare the ux performance of the
spiral wound FO module (in s-FO-in conguration) to that of a
at FO coupon (Fig. 5(a)). The latter was measured in a cross-ow
FO testing cell using a small HTI membrane coupon (60 cm2 ) cut
from a Hydrowell module [12]. In both cases, the porous support
was orientated towards the draw solution. Interestingly, the spiral
wound module (s-FO-in conguration) had much lower membrane
ux at comparable draw solution concentrations. This might be
explained by the major difference between testing in a spiral wound
conguration versus testing a small-area at couponthe bulk concentration of the draw solution Cdraw was signicantly diluted by
a sizable permeate ow in the spiral wound module (Eq. (6)). Consider a cross-ow rate of 50 ml/min at the inlet to the module
and a permeate ow of 100 ml/min, the total cross-ow rate was
increased to 150 ml/min with Cdraw diluted by a factor of 3 at the
module outlet. This is somewhat analogous to testing an RO spiral
wound module where brine concentration increases from module inlet to outlet as pure water permeate through the membrane,
which tends to reduce the ux performance of the module at high
recovery in comparison to that of a small coupon (recovery 0). In
the present case, the low ux in the s-FO-in conguration (Fig. 4
and Fig. 5(a)) was likely due to ICP as well as dilution of the bulk
draw solution concentration. Model simulation was performed in
accordance to Appendix A, where the dilution effect by the permeate ux was simulated by Eq. (6), and the ICP effect was determined
by Eq. (1). The experimental results agreed very well with the simulation results at lower draw solution concentrations (Fig. 5(a)). On
the other hand, the experimental ux was slightly lower than the
theoretical prediction for Cdraw,0 above 0.8 M. This discrepancy at
high draw solution concentration might be due to the mass transfer
resistance and ECP on the feed water side which became increasingly important at high permeate ow (i.e., higher Cdraw,0 ) as forced
feed water circulation was not employed in the submerged conguration (refer to Section 4.2 for additional discussion).
The effect of draw solution cross-ow rate on submerged FO
ux performance at a xed draw solution concentration is shown in
Fig. 6. At low draw solution ow rate, increasing cross-ow resulted
in substantial permeate ux enhancement. Such enhancement

effect became less prominent at relatively high cross-ow rate


(>100 ml/min). Flux enhancement at greater cross-ow velocity has
been well documented for pressure-driven membrane processes
(e.g., RO) thanks to the reduced external concentration polarization
and membrane fouling as a result of the increased mass transfer
coefcient [36,37]. However, the internal concentration polarization is less likely to be affected by cross-ow velocity as the salt
accumulation (or dilution) occurred in the porous membrane support which acts like an unstirred layer [38,39]. For example, She
et al. [12] tested at sheet HTI membrane coupons in the FO
mode, and they reported only a marginal increase (<5%) in FO ux
when the cross-ow rate was doubled. Therefore, the increased
FO ux at greater cross-ow rate in the current study cannot
be solely attributed to the enhancement of mass transfer at the
liquidmembrane interface on the draw solution side. A more plausible explanation is the dilution of bulk concentration of the draw
solution (Eq. (6)). As discussed in Section 2, the draw solution concentration in the spiral wound module became signicantly diluted
by the permeate ow. The dilution factor, i.e., the ratio of the inlet
draw solution concentration to the outlet concentration, is given
by (Qp + Qcf,0 )/Qcf,0 , where Qp and Qcf,0 are the total permeate ow
rate and the inlet cross-ow rate, respectively (refer to Eq. (7)).
Thus, excessive dilution of draw solution (large dilution factor)
may occur at low draw solution cross-ow rate, which was likely
responsible for the reduced FO ux. The dilution factor decreased
at greater cross-ow, corresponding to enhanced ux performance
(Fig. 6). However, at relatively high cross-ow rate (draw solution ow rate  permeate ow rate), further increase in cross-ow
became less effective in reducing the dilution factor, which is consistent with the experimental observation of reduced effectiveness
of cross-ow. Fig. 6 also presents the simulation results for various draw solution cross-ows and concentrations based on Eq.
(1) (ICP effect) and Eq. (6) (dilution effect). The simulated results
agreed well with the experimental ones, which conrms that the
dependence of the permeate ow of FO module on draw solution
cross-ow was mainly caused by the dilution of bulk draw solution
concentration.
Based on the above discussion, the cross-ow in a spiral wound
FO module needs to be carefully optimized. Permeate ux can be
adversely affected due to dilution effect if the draw solution ow
rate is relatively low in comparison to the permeate ow rate. On
the other hand, high cross-ow of draw solution demands greater
pumping energy, yet its effect on ux enhancement may be limited. A balance of pumping energy and ux performance shall be
achieved at the optimal cross-ow. In Fig. 6, the lines corresponding to dilution factors of 3, 2 and 1.5, respectively, are indicated.
For example, a dilution factor of 2 is represented by a line with
1:1 slope in a permeate ow rate versus cross-ow plot. It seems
that substantial ux reduction occurred when the dilution factor
was greater than 2. Such conditions shall be avoided to maintain
an acceptable FO ux.
4.2. x-FO congurations
Two cross-ow FO congurations (x-FO-in and x-FO-out) were
tested in the current study, and their ux performance is compared to the submerged FO conguration in Fig. 7. For all the three
congurations, permeate ow increased at higher draw solution
concentration as a result of increased driving force. In addition, the
slope of each curve reduced as concentration increased. Once again,
this is consistent with Eqs. (1) and (2) that ICP is more severe at
increased ux level and greater draw solution concentration due
to its exponential dependence on permeate ux.
In Fig. 7, the ux performance of x-FO-in was very similar to
that of s-FO-in, except slightly higher ux was achieved in the
cross-ow conguration at higher draw solution concentrations

Y. Xu et al. / Journal of Membrane Science 348 (2010) 298309

Fig. 7. Comparison between s-FO-in, x-FO-in, and x-FO-out congurations. Draw


solution ow rate at 75 ml/min, and feed water ow rate at 140 ml/min (where
applicable). Simulation results were obtained from Eq. (6) coupled with ICP
models (Section 2 and Appendix A). The error bars shown in the gure represent one standard deviation based on at least three different tests. Note:
1 L/m2 h = 2.78 107 m/s.

(Cdraw,0 > 0.5 M). In both congurations, the draw solution was circulated at a rate of 75 ml/min. While a cross-ow of 140 ml/min
pure water was circulated on the feed water side for the x-FO-in
conguration, no forced circulation was available for submerged
one. Thus, the slightly better ux in x-FO-in conguration at greater
Cdraw,0 was likely due to the reduced mass transfer resistance and
ECP on the feed water side, which is consistent with our earlier
discussion in Section 4.1. For comparison purpose, simulation was
performed based on Eqs. (1) and (6). It is worthwhile to note that
the simple theoretical model in the current study does not take into
consideration the feed water cross-ow by assuming that the membrane ux is not limited by the mass transfer and ECP on the feed
water side. Therefore, identical simulation results were obtained
for both x-FO-in and s-FO-in congurations. A comparison between
the simulation results and the experimental data shows (1) that
the model prediction agreed better with the experimental results
at lower draw solution concentration; and (2) that the model simulation worked better for the x-FO-in conguration. Both trends can
be explained by the lower mass transfer resistance and ECP on the
feed water side under these conditions.
The permeate ow of the x-FO-out conguration is also presented in Fig. 7. This conguration (where the dense active layer
was facing the draw solution) had higher ux compared to the
s-FO-in and x-FO-in congurations where active layer was facing the feed water. Similar results have been reported by various
research groups [12,13,22]. Such difference was likely due to the
more severe ICP in the s-FO-in and x-FO-in congurations, where a
dilutive concentration polarization occurred in the support (i.e., the
draw solution was diluted inside the porous support layer to cause
a drastic loss of driving force, Fig. 2(a)). This explanation was also
conrmed by the simulation results based on Eqs. (2) and (6)the
simulated permeate ow for x-FO-out is clearly superior to those
for x-FO-in and s-FO-in (Fig. 7).
Surprisingly, the experimental results in the x-FO-out conguration were signicantly lower compared to the simulated results
(Fig. 7). It is hypothesized that the mass transfer resistance on the
feed water side was much greater in this conguration due to its
much longer ow path. The ow path of feed water in the x-FOout conguration was 4.5 m (twice the length of the membrane
leaf (Fig. 3(a)) as compared to a 0.3 m ow path length for x-FO-in
where feed water owed outside of the membrane envelope. The x-

305

Fig. 8. Effect of feed water ow rate on permeate ow in the x-FO-out conguration.


Simulation results were obtained from Eq. (6) coupled with ICP models (Section 2
and Appendix A). Note: 1 L/m2 h = 2.78 107 m/s.

FO-out mode could result in more severe salt accumulation and ECP
on the feed water side inside the membrane envelope. The effect
of feed water ow rate on module performance in x-FO-out conguration is better illustrated in Fig. 8 for different feed and draw
ow rates. At a draw solution cross-ow of 75 ml/min, the permeate ow at a feed ow rate of 60 ml/min was drastically lower than
that at 140 ml/min feed water ow. Indeed, the maximum permeate ow rate (60 ml/min) was identical to the feed ow rate at the
spiral wound module inlet. When this maximum permeate ow
was achieved (which corresponds to 0.5 and 0.8 M draw solutions),
the feed water ow at the module outlet was almost zero. Under
this condition, nearly all the available feed water passed through
the membrane resulting in a limitation of the maximum permeate ow to this amount. The same was true for the combination
of a 140 ml/min draw solution and a 140 ml/min feed water. Fig. 8
clearly suggests that the permeate ow in the x-FO-out conguration was constrained by the feed water ow rate in addition to
the ICP effect and the draw solution dilution effect. In contrast, the
effect of feed water ow rate in the x-FO-in conguration was much
less important. The ux performance of s-FO-in (where there was
no forced circulation of feed water) was nearly identical to that of
x-FO-in, as a result of signicantly shorter ow path of feed water.
4.3. x-PRO congurations
The PRO performance of the Hydrowell spiral wound module was evaluated in both x-PRO-in and x-PRO-out congurations
(Fig. 9). While the permeate ow in the x-PRO-out conguration
seems to be more stable, the ux performance in x-PRO-in deteriorated drastically upon applying a back pressure. In the x-PRO-in
conguration, the draw solution that owed inside the membrane
envelope was pressurized. Major leakage of draw solution was
identied at a pressure of 3 bar which forced the experiment to
be stopped. The leakage was due to the failure of the glue between
the two pieces of membranes (Fig. 3(a)) under the applied pressure.
Similar failures were observed when additional Hydrowell modules were tested (data not reported here). It is concluded that the
Hydrowell module is not suitable for the x-PRO-in conguration,
except where applied pressure is relatively low. Where it is desirable to have the dense rejection layer facing the feed water (e.g.,
for fouling control [12,13]), the module may need to be redesigned
such that the rejection layers are inside of the membrane envelope and pressure is applied outside of the envelope to ensure the
integrity of the module.

306

Y. Xu et al. / Journal of Membrane Science 348 (2010) 298309

Fig. 9. Comparison of x-PRO-out and x-PRO-in congurations. Draw solution concentration at 0.5 M. The open symbols represent the respective uxes obtained in
the x-FO-out and x-FO-in congurations. Note: 1 L/m2 h = 2.78 107 m/s.

The ux performance at various draw solution concentrations


and applied pressures in the x-PRO-out conguration was shown in
Fig. 10. As predicted correctly by Eq. (5), membrane ux reduced at
higher pressure and/or lower draws solution concentration due to
the reduced driving force (the osmotic pressure difference minus
the applied pressure, refer to Fig. 1(b) and (d)). At a draw solution concentration of 0.5 M, the permeate ow was reduced by
approximately 50% when the applied pressure increased from 0
to 5 bar. However, the experimentally measured permeate ow
for 0.3 and 0.5 M draw solutions was signicantly lower than the
corresponding simulated results. As discussed in Section 4.2, this
difference was likely due to the mass transfer resistance and ECP
on the feed water side. Interestingly, it became much more difcult to pump feed water into the membrane module (i.e., inside
the membrane envelope) upon pressurizing on the draw solution.
In the Hydrowell module, identical spacers were used both inside
and outside the membrane envelope (Fig. 3(a)). These spacers were
formed by plastic strings arranged in orthogonal manner, with
spacing between adjacent strings 2.5 mm. This means that membranes that were unsupported over a 2.5 mm span had to withstand
the high pressure applied on the draw solution side, which is not

Fig. 10. Effect of applied pressure on permeate ow in the x-PRO-out conguration at different draw solution concentrations. Cross-ow rate of draw solution
at 75 ml/min. Simulation results were obtained from Eq. (6) coupled with Eq. (5)
(Section 2 and Appendix A). Note: 1 L/m2 h = 2.78 107 m/s.

Fig. 11. Available power density from the Hydrowell spiral wound module in xPRO-out conguration. A 0.5 M NaCl was used as the draw solution.

an ideal design for high pressure applications such as PRO. The


deformation of the membranes under the high pressure and the
corresponding narrowing of the ow channel inside the membrane
envelope can signicantly increase the mass transfer resistance on
the feed water side. A spacer design that is similar to an RO permeate collector may work better in this regard. At a draw solution
of 0.1 M, our simulated results agreed reasonably well with the
observed permeate ow. Consistent with the discussion in Section
4.2, the membrane ux in this case was sufciently low that it was
not constrained by the feed water ow.
A potential application of PRO is for harvesting osmotic power
from seawater [14,17,20,32,40]. The power density, or the power
output per membrane area from the PRO process, can be determined from Eq. (11). Fig. 11 presents the available power density
using a 0.5 M NaCl draw solution. The dotted line represent the simulated power density assuming a large ow rate of draw solution
such that dilution of the bulk draw solution by the permeate ow
is negligible. The maximum power density is 1.5 W/m2 membrane
area, which occurs at an applied pressure of 1011 bar. This agrees
well with earlier studies that the maximum power density occurs at
an applied pressure of 50% of the draw solution osmotic pressure
[8,18]. The power density can be signicantly lower if dilution of the
draw solution becomes signicant (refer to the solid line in Fig. 11).
For a draw solution ow rate of 75 ml/min, the maximum power
density is only 1 W/m2 . The corresponding optimal pressure is also
lower (9 bar) as a reection of the dilution effect. Compared to the
simulated data, the actual power performance of the Hydrowell
module was much worse due to the combination of dilution effect
as well as the feed water mass transfer resistance. Improved module/spacer design and proper management of draw solution and
feed water cross-ows are therefore required for optimized power
performance from a PRO module.
The maximum available PRO power density is presented as a
function of draw solution concentration in Fig. 12. The simulated
results without the dilution effect (dotted line) shows that the
power density is approximately proportional to (Cdraw,0 )2 . Similar
results have been reported by Ludwig et al. [41]. In contrast, the
power density is proportional to (Cdraw,0 )1.7 when the dilution effect
is considered for a draw solution of 75 ml/min (solid line), while the
experimental results showed a nearly linear dependence on the
draw solution concentration. Our results suggest that improved
module design and cross-ow management are more critical for
greater draw solution concentrations (higher ux levels).

Y. Xu et al. / Journal of Membrane Science 348 (2010) 298309

307

Acknowledgements
This research received nancial support from Suzhou Industrial
Park Technology Innovation Zone and Suzhou Institute of Sichuan
University. Xu and Peng were supported by Suzhou Institute of
Sichuan University.
Appendix A.
A.1. Derivation of ICP models
Based on the classical solution-diffusion model for the rejection
layer, we have
Jv = A(draw support P)

(A1)

Note: P = 0 for FO and P > 0 for PRO


Js = B(Cdraw Csupport )
Fig. 12. Effect of draw solution on maximum power density of PRO.

The current study did not perform any direct measurements on


the salt ux through the FO membrane and the pressure drops cross
the FO module (for both the feed water side and the draw solution
side). Such measurements may provide additional insights into the
mass transfer and ux behavior of the FO process. Therefore, it is
recommended that future studies may incorporate these measurements as well. In addition, the cascading of multiple FO modules in
a membrane array will also be of great interest, though beyond the
scope of the current study.
5. Conclusions
The effect of draw solution concentration and operating conditions on permeate ow rate of forward osmosis and pressure
retarded osmosis was investigated in the current study using a spiral wound Hydrowell module. The following conclusions can be
made from the current study:
In both FO and PRO modes, the permeate ow increased at greater
draw solution concentration. However, the ux behavior was
highly non-linear with respect to Cdraw,0 due to internal concentration polarization as well as the dilution of draw solution by the
permeate ow.
ICP, the concentration polarization in the porous membrane support, played an important role in membrane ux reduction in FO
and PRO modes, and it became increasingly dominant at greater
draw solution concentration.
Likewise, the dilution effect was more severe at greater draw
solution concentration. The cross-ow of the draw solution shall
be sufciently high (dilution factor < 2) to limit the adverse dilution effect.
The feed water ow circulation had limited effect on membrane
ux when it was circulated outside of the membrane envelope,
and the submerged FO conguration performed nearly as good
as the cross-ow x-FO-in conguration. However, the feed ow
imposed an upper limit for the membrane permeate when it was
circulated inside of the membrane envelope. The effect of feed
ow may be explained by the mass ow resistance and ECP on
the feed solution side.
A simple and useful model was developed to account for the ICP
as well as the dilution effects. The model prediction agreed well
with the experimental results when mass transfer resistance on
the feed water side was insignicant. However, the model overestimates membrane ux when ECP on the feed solution side is
signicant.

(A2)

For the solute transport in the support layer


Jv C + Js =

D dC
dx

(A3)

where C is the concentration in the support layer at a distance x


away from the interface of the support layer and the active layer,
and D/ is the effective diffusion coefcient of the solute in the
porous support with a porosity of .
Consider the active layer facing the draw solution orientation,
the boundary conditions of Eq (A3) are
C = Csupport at x = 0

(A4)

and
C = Cfeed at x = t

(A5)

By combining Eqs. (A2) and (A3), one can derive:

Csupport + (B(Cdraw Csupport )/Jv )


ln
Cfeed + (B(Cdraw Csupport )/Jv )

Jv
Km

(A6)

where
Km =
S=

D
D
=
t
S

(A7)

t

(A8)

If we assume that the osmotic pressure of a solution is proportional


to the solution concentration, Eq. (A6) becomes

support Jv + B(draw support )


ln
feed Jv + B(draw support )

Jv
Km

(A9)

By substituting Eq. (A1) into Eq. (A9) to illuminate support , we can


obtain the ICP equation for the active layer facing the draw solution
orientation in the PRO mode (Eq. (5) in the main paper):

Jv = Km ln

(Adraw AP Jv ) + B((AP/Jv ) + 1)
Afeed + B((AP/Jv ) + 1)

(5)

Eq. (2) can be obtained by setting P = 0 in Eq. (5). Eqs. (1) and (4) can
be derived in a similar fashion.
A.2. An example for modeling FO/PRO permeate ow in a spiral
wound module
The local membrane ux anywhere inside the spiral wound
module can be modeled by ICP ux models presented in Section 2
(Eqs. (1)(5)). Based on the at coupon FO test results and additional
independent RO tests by She [12], the HTI membrane used in the
current studies had the following properties: A = 2.2 1012 m/s Pa

308

Y. Xu et al. / Journal of Membrane Science 348 (2010) 298309

Table A1
An example for simulating spiral wound FO permeate uxa .
Stepb

Local draw solution


concentration, Cdraw,x
(M)c

Local membrane ux,


Jv,x (L/m2 h)d

Incremental membrane
area, Am (m2 )b

Incremental permeate
ow, Qp,x (ml/min)e

Accumulative
permeate ow, Qp,x
(ml/min)f

1
2
3
4

100

0.500
0.492
0.484
0.477

0.237

7.90
7.79
7.68
7.58

4.04

0.0094
0.0094
0.0094
0.0094

0.0094

1.24
1.22
1.20
1.19

0.63

1.24
2.46
3.66
4.85

83.87

Notes:
a
Simulation conditions: s-FO-in/x-FO-in congurations; total effective membrane area in the module of 0.94 m2 ; a 0.5 M draw solution with a ow rate of 75 ml/min at
the module inlet.
b
The total membrane area was divided into 100 equal small areas to compute the local incremental permeate ow.
c
The local draw solution was determined from the inlet draw solution concentration according to Eq. (6).
d
The local ux was obtained as a function of local draw solution concentration from Fig. A.1. Unit conversion factor: 1 L/m2 h = 2.78 107 m/s.
e
The incremental permeate ow was determined from Qp,x = Jv,x Am .
f

The accumulative permeate ow was obtained from Qp,x = (Qp,x ) or Qp,x =

and B = 1.7 107 m/s. The mass transfer coefcient Km inside the
membrane support was 3.3 106 m/s when the dense rejection
layer faced the draw solution, and it was 4.2 106 m/s in the alternative membrane orientation. These values agreed reasonably well
with Gray et al. [22]. Based on Eqs. (1) and (2), the local FO ux can
be determined as a function of local draw solution concentration
(Fig. A.1). Similarly, ux in PRO can be determined by Eqs. (4) and
(5) (results not shown here).
The FO/PRO permeate ow rate of a spiral wound module can be
modeled using both the ICP equations (e.g., the ux-concentration
equations in Fig. A.1 for FO mode) together with Eq. (6) which
accounts for the dilution effect. A worked example is shown in
Table A1. With a 0.5 M NaCl draw solution at a ow rate of
75 ml/min at the module inlet, the local membrane ux is as high as
7.9 L/m2 h according to Fig. A.1. As the draw solution ows through
the module, its concentration becomes increasingly diluted by the
permeate ow. A total of 84 ml/min permeate ow is available from
the entire membrane module. Consequently, the concentration
at the module outlet is determined by 0.5 75/(75 + 84) = 0.24 M,
which is much lower than the inlet concentration. The corresponding local ux at the outlet is only 4.0 L/m2 h. Clearly, the ux

x
0

Jv dAm .

performance is adversely affected due to the draw solution dilution


effect.
Nomenclature 2
A
Am
B
Cdraw
Cdraw,0
Cdraw,exit
Cdraw,x
Cfeed
Csupport
D
Jv
Jv
Km
P
Qcf
Qcf,0
Qp
S
t

draw
feed

PRO
PRO

transport coefcient for water (i.e., water permeability) (m/s Pa)


membrane area (m2 )
transport coefcient for solute (m/s)
draw solution concentration (M)
draw solution concentration at module inlet (M)
draw solution concentration at module outlet (M)
draw solution concentration at location x (M)
feed solution concentration (M)
solute concentration at the interface of the dense
rejection layer and the porous support layer (M)
solute diffusion coefcient (m2 /s)
permeate ux of water (m/s or L/m2 h)
average permeate ux in a spiral wound module
(m/s or L/m2 h)
mass transfer coefcient of the membrane support
(m/s)
applied pressure on the draw solution (Pa)
cross-ow rate of the draw solution (m3 /s or
ml/min)
cross-ow rate of the draw solution at module inlet
(m3 /s or ml/min)
membrane permeate ow (m3 /s or ml/min)
membrane structural parameter (m)
membrane support layer thickness (m)
porosity of the membrane support layer
osmotic pressure of the bulk draw solution (Pa)
osmotic pressure of the bulk feed water (Pa)
tortuosity of the membrane support layer
total power available from the module (W)
power density (W/m2 )

References
[1] E. Aoustin, A.I. Schafer, A.G. Fane, T.D. Waite, Ultraltration of natural organic
matter, Separation and Purication Technology 22 (3) (2001) 63.
Fig. A.1. Local permeate ux in FO mode as a function of local draw solution concentration. Data obtained from Ref. [12]. The permeate ux was approximated
by the respective equations in the gure for both membrane orientations. Note:
1 L/m2 h = 2.78 107 m/s.

Unit conversion factor: 1 L/m2 h = 2.78 107 m/s.

Y. Xu et al. / Journal of Membrane Science 348 (2010) 298309


[2] W. Yuan, A.L. Zydney, Humic acid fouling during ultraltration, Environmental
Science & Technology 34 (2000) 5043.
[3] A. Fane, Membrane bioreactors: design and operational options, Filtration and
Separation 39 (2002) 26.
[4] W. Yang, N. Cicek, J. Ilg, State-of-the-art of membrane bioreactors: worldwide
research and commercial applications in North America, Journal of Membrane
Science 270 (2006) 201.
[5] C.R. Bartels, M. Wilf, K. Andes, J. Iong, Design considerations for wastewater
treatment by reverse osmosis, Water Science and Technology 51 (2005) 473.
[6] A. Bennett, Recent developments in RO technology reduces whole life project
costs, Filtration and Separation 40 (2003) 20.
[7] C. Bartels, M. Hirose, H. Fujioka, Performance advancement in the spiral wound
RO/NF element design, Desalination 221 (2008) 207214.
[8] K.L. Lee, R.W. Baker, H.K. Lonsdale, Membranes for power generation by
pressure-retarded osmosis, Journal of Membrane Science 8 (1981) 141171.
[9] S. Loeb, One hundred and thirty benign and renewable megawatts from Great
Salt Lake? The possibilities of hydroelectric power by pressure-retarded osmosis, Desalination 141 (2001) 8591.
[10] T.Y. Cath, A.E. Childress, M. Elimelech, Forward osmosis: principles, applications, and recent developments, Journal of Membrane Science 281 (2006)
7087.
[11] R.W. Holloway, A.E. Childress, K.E. Dennett, T.Y. Cath, Forward osmosis for
concentration of anaerobic digester centrate, Water Research 41 (2007)
40054014.
[12] Q. She, Effect of Hydrodynamic Conditions and Feedwater Composition on
Fouling of Ultraltration and Forward Osmosis Membranes by Organic Macromolecules, M.Eng. Thesis. In School of Civil and Environmental Engineering,
Nanyang Technological University, Singapore, 2008.
[13] E.R. Cornelissen, D. Harmsen, K.F. de Korte, C.J. Ruiken, J.J. Qin, H. Oo, L.P.
Wessels, Membrane fouling and process performance of forward osmosis
membranes on activated sludge, Journal of Membrane Science 319 (2008)
158168.
[14] S. Loeb, Energy production at the Dead Sea by pressure-retarded osmosis: challenge or chimera? Desalination 120 (1998) 247262.
[15] M. Reali, Closed cycle osmotic power plants for electric power production,
Energy 5 (1980) 325329.
[16] J.R. McCutcheon, R.L. McGinnis, M. Elimelech, A novel ammonia-carbon dioxide
forward (direct) osmosis desalination process, Desalination 174 (2005) 111.
[17] S. Loeb, Osmotic power plants, Science 189 (1975) 654655.
[18] R.L. McGinnis, J.R. McCutcheon, M. Elimelech, A novel ammoniacarbon dioxide
osmotic heat engine for power generation, Journal of Membrane Science 305
(2007) 1319.
[19] S. Loeb, Production of energy from concentrated brines by pressure retarded
osmosis. I. Preliminary technical and economic correlations, Journal of Membrane Science 1 (1976) 4963.
[20] K. Gerstandt, K.V. Peinemann, S.E. Skilhagen, T. Thorsen, T. Holt, Membrane
processes in energy supply for an osmotic power plant, Desalination 224 (2008)
6470.
[21] S.E. Skilhagen, J.E. Dugstad, R.J. Aaberg, Osmotic powerpower production
based on the osmotic pressure difference between waters with varying salt
gradients, Desalination 220 (2008) 476482.
[22] G.T. Gray, J.R. McCutcheon, M. Elimelech, Internal concentration polarization in
forward osmosis: role of membrane orientation, Desalination 197 (2006) 18.
[23] J.R. McCutcheon, M. Elimelech, Inuence of concentrative and dilutive internal concentration polarization on ux behavior in forward osmosis, Journal of
Membrane Science 284 (2006) 237247.
[24] S. Loeb, L. Titelman, E. Korngold, J. Freiman, Effect of porous support fabric
on osmosis through a Loeb-Sourirajan type asymmetric membrane, Journal of
Membrane Science 129 (1997) 243249.

309

[25] J.R. McCutcheon, M. Elimelech, Inuence of membrane support layer hydrophobicity on water ux in osmotically driven membrane processes, Journal of
Membrane Science 318 (2008) 458466.
[26] W. Tang, H.Y. Ng, Concentration of brine by forward osmosis: performance and
inuence of membrane structure, Desalination 224 (2008) 143153.
[27] T.Y. Cath, D. Adams, A.E. Childress, Membrane contactor processes for wastewater reclamation in space: II. Combined direct osmosis, osmotic distillation,
and membrane distillation for treatment of metabolic wastewater, Journal of
Membrane Science 257 (2005) 111119.
[28] A. Achilli, T.Y. Cath, A.E. Childress, Power generation with pressure retarded
osmosis: an experimental and theoretical investigation, Journal of Membrane
Science 343 (2009) 4252.
[29] http://www.htiwater.com/.
[30] C.Y. Tang, Y.-N. Kwon, J.O. Leckie, Effect of membrane chemistry and coating
layer on physiochemical properties of thin lm composite polyamide RO and
NF membranes. I. FTIR and XPS characterization of polyamide and coating layer
chemistry, Desalination 242 (2009) 149167.
[31] C.Y. Tang, Y.-N. Kwon, J.O. Leckie, Effect of membrane chemistry and coating layer on physiochemical properties of thin lm composite polyamide
RO and NF membranes II. Membrane physiochemical properties and their
dependence on polyamide and coating layers, Desalination 242 (2009)
168182.
[32] S. Loeb, Large-scale power production by pressure-retarded osmosis, using
river water and sea water passing through spiral modules, Desalination 143
(2002) 115122.
[33] K.B. Petrotos, P. Quantick, H. Petropakis, A study of the direct osmotic concentration of tomato juice in tubular membranemodule conguration. I. The
effect of certain basic process parameters on the process performance, Journal
of Membrane Science 150 (1998) 99110.
[34] J.R. McCutcheon, R.L. McGinnis, M. Elimelech, Desalination by
ammoniacarbon dioxide forward osmosis: inuence of draw and feed
solution concentrations on process performance, Journal of Membrane Science
278 (2006) 114123.
[35] R.L. McGinnis, M. Elimelech, Energy requirements of ammoniacarbon dioxide
forward osmosis desalination, Desalination 207 (2007) 370382.
[36] P. Bacchin, D. Si-Hassen, V. Starov, M.J. Clifton, P. Aimar, A unifying model
for concentration polarization, gel-layer formation and particle deposition in
cross-ow membrane ltration of colloidal suspensions, Chemical Engineering
Science 57 (2002) 77.
[37] T.H. Chong, F.S. Wong, A.G. Fane, Implications of critical ux and cake
enhanced osmotic pressure (CEOP) on colloidal fouling in reverse osmosis: experimental observations, Journal of Membrane Science 314 (2008)
101111.
[38] E.M.V. Hoek, M. Elimelech, Cake-enhanced concentration polarization: a new
fouling mechanism for salt-rejecting membranes, Environmental Science &
Technology 37 (2003) 5581.
[39] T.H. Chong, F.S. Wong, A.G. Fane, Enhanced concentration polarization by
unstirred fouling layers in reverse osmosis: detection by sodium chloride tracer response technique, Journal of Membrane Science 287 (2007)
198210.
[40] R.J. Aaberg, Osmotic power: a new and powerful renewable energy source?
Refocus 4 (12) (2003) 4850.
[41] W. Ludwig, A. Seppala, M.J. Lampinen, Experimental study of the osmotic
behaviour of reverse osmosis membranes for different NaCl solutions and
hydrostatic pressure differences, Experimental Thermal and Fluid Science 26
(2002) 963969.

Anda mungkin juga menyukai