Anda di halaman 1dari 8

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/44653620

Reverse Draw Solute Permeation in Forward


Osmosis: Modeling and Experiments
Article in Environmental Science and Technology July 2010
DOI: 10.1021/es100901n Source: PubMed

CITATIONS

READS

240

1,243

3 authors, including:
William A Phillip

Menachem Elimelech

University of Notre Dame

Yale University

43 PUBLICATIONS 2,619 CITATIONS

412 PUBLICATIONS 36,563 CITATIONS

SEE PROFILE

All in-text references underlined in blue are linked to publications on ResearchGate,


letting you access and read them immediately.

SEE PROFILE

Available from: William A Phillip


Retrieved on: 22 July 2016

Environ. Sci. Technol. 2010, 44, 51705176

Reverse Draw Solute Permeation in


Forward Osmosis: Modeling and
Experiments
WILLIAM A. PHILLIP, JUI SHAN YONG,
AND MENACHEM ELIMELECH*
Department of Chemical Engineering, Environmental
Engineering Program, P.O. Box 208286, Yale University,
New Haven, Connecticut 06520-8286

Received March 20, 2010. Revised manuscript received


May 19, 2010. Accepted May 21, 2010.

Osmotically driven membrane processes are an emerging


set of technologies that show promise in water and wastewater
treatment, desalination, and power generation. The effective
operation of these systems requires that the reverse flux of draw
solute from the draw solution into the feed solution be
minimized. A model was developed that describes the reverse
permeation of draw solution across an asymmetric membrane
in forward osmosis operation. Experiments were carried out to
validate the model predictions with a highly soluble salt
(NaCl) as a draw solution and a cellulose acetate membrane
designed for forward osmosis. Using independently determined
membrane transport coefficients, strong agreement between
the model predictions and experimental results was observed.
Further analysis shows that the reverse flux selectivity, the
ratio of the forward water flux to the reverse solute flux, is a
key parameter in the design of osmotically driven membrane
processes. The model predictions and experiments demonstrate
that this parameter is independent of the draw solution
concentration and the structure of the membrane support
layer. The value of the reverse flux selectivity is determined
solely by the selectivity of the membrane active layer.

Introduction
Forward osmosis (FO) and pressure retarded osmosis (PRO)
are two emerging technologies that fall under the classification of osmotically driven membrane processes (1, 2). These
technologies take advantage of the osmotic pressure difference that is generated when a semipermeable membrane
separates two solutions of differing concentrations. By using
the osmotic pressure difference to drive the permeation of
water across the semipermeable membrane, osmotically
driven membrane processes may be capable of addressing
several of the shortcomings of hydraulically driven membrane
processes, such as reverse osmosis (RO).
Unlike RO, FO does not require a high applied hydraulic
pressure, thereby decreasing capital and energy costs (3).
Furthermore, recent investigations have demonstrated a
lower fouling propensity with FO (1, 4-6), implying lower
operating costs. Several studies have taken advantage of these
benefits and demonstrated the use of osmotically driven
membrane processes to desalinate seawater and brackish
water (1, 7-9), treat wastewater (4, 10), and reclaim wastewater using an osmotic membrane bioreactor (11).
* Corresponding author phone: (203)432-2789; e-mail: menachem.
elimelech@yale.edu.
5170

ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 44, NO. 13, 2010

A significant portion of the efforts to improve FO and


PRO operations has focused on tailoring the membrane
structure to decrease the effects of internal concentration
polarization (ICP) (12-19) or on developing new draw
solutions (20) that are capable of generating large osmotic
pressures, but are still relatively easy to separate from water.
Further developments in these areas are still needed for
successful commercialization of these technologies. However,
one area of research that has received limited attention, but
could be a significant impediment to the viability of osmotically driven membrane processes, is the reverse permeation
of draw solute from the draw solution into the feed solution
(11, 21).
An ideal semipermeable membrane would prevent any
dissolved draw solute from permeating into the feed solution.
However, no membrane is a perfect barrier, and a small
amount of dissolved solute will be transported across the
membrane. If an expensive draw solute is used, the cost of
replenishing the draw solute lost to the feed solution could
make the process less economical. Alternatively, if the draw
solute was detrimental to the aquatic environment, an
additional treatment step of the feed solution concentrate
would be required prior to discharge. Therefore, a thorough
understanding of the phenomenon of reverse solute permeation is critical to the effective development of osmotically
driven membrane technologies.
The objectives of this paper are (i) to formulate a model
that describes the reverse permeation of a single draw solute
across an asymmetric membrane in a forward osmosis
operation, (ii) to validate the model through laboratory
experiments with a well characterized forward osmosis
membrane and draw solution, and (iii) to use the model to
gain insights into the processes that govern draw solute
permeation in forward osmosis. In our model, the draw solute
reverse flux is described in terms of experimentally accessible
quantities and common transport parameters. The implications of our results for the future development of forward
osmosis are further evaluated and discussed.

Theory
A schematic of an asymmetric membrane operating in FO
mode (i.e., with the selective layer facing the feed solution)
is shown in Figure 1. For the draw solute to leak into the feed
solution, it must first diffuse through the support layer, where
its diffusion is opposed by the convective flow of solvent,
until it reaches the interface between the support layer and
the active layer. Once there, the draw solute partitions into
the active layer before diffusing across it. After diffusing
across the active layer, the draw solute partitions into the
feed solution, which has a negligible concentration of draw
solute. This process can be described by considering the
mass transfer through the support layer and then the active
layer in series.
Draw Solute Mass Balance in the Support Layer. For the
support layer, a steady-state mass balance can be written on
a differential volume
dJsS
d2c
dc
) -DS 2 + Jw
)0
dz
dz
dz

(1)

where JsS is the total flux of draw solute, c is the solute


concentration, DS is the solute diffusion coefficient in the
support layer, and Jw is the superficial fluid velocity, which
is equivalent to the solvent permeate flux. This mass balance
is, in principle, the same as that used to describe the
10.1021/es100901n

2010 American Chemical Society

Published on Web 06/07/2010

General Solution for Draw Solute Concentration Profile


and Flux in the Support Layer. Integrating eq 1 twice and
using boundary conditions 3a-3b give the following expression for the draw solute concentration profile in the support
layer
exp
c)

( )

JwtS z
JwtS s
(c - cis) + exp
c - cD
D tS D
D i
JwtS
exp
-1
D

( )

(4)

The concentration profile can then be used to find an


expression for the total draw solute flux by taking the sum
of the diffusive and convective components of the flux
JsS ) -DS

dc
+ Jwc
dz

(5)

Substituting eq 4 for c in eq 5 yields an expression for the


solute flux into the feed solution

JsS

FIGURE 1. A schematic of draw solute leaking into the feed


solution. The high concentration of solute in the draw solution,
cD, creates a chemical potential gradient that drives both the
forward water flux, Jw, and the reverse flux of solute, Js. For
the draw solute to permeate across the asymmetric membrane
into the feed solution, where its concentration cF is negligible,
it must be transported across the support layer of thickness tS,
and the active layer of thickness tA. ciS and ciA represent the
draw solute concentrations on the support layer side and
active layer side of the support layer-active layer interface,
respectively.
phenomenonofinternalconcentrationpolarization(12,14,22).
We also note that eq 1 implies that DS, JsS, and Jw are
independent of z and are constant across the entire support
layer.
Implicitly, we assume that the draw solute is a single entity,
even though it may consist of several chemical species that
are strongly associated, as is the case with strong electrolytes
(23). Also, the diffusion coefficient in this equation DS is an
effective diffusion coefficient, which can be related to the
bulk diffusion coefficient D by accounting for the porosity,
, and tortuosity, , of the support layer (24)
DS )

(2)

The porosity accounts for the fact that the experimental


measurements are based on the total membrane area and
not the cross-sectional area of the pores, while the tortuosity
accounts for the additional distance a solute molecule must
travel relative to the support layer thickness.
Because the coordinate system z points into the support
layer, eq 1 is subject to the following boundary conditions
z ) 0 c ) ciS

(3a)

z ) tS c ) cD

(3b)

Here, ciS is the draw solute concentration on the support


layer side of the support layer-active layer interface, cD is the
bulk draw solute concentration, and tS is the support layer
thickness.

( ( )
( )

JwtS s
c - cD
D i
JwtS
exp
-1
D

Jw exp

(6)

This expression is not readily compared to experiments


because the draw solute concentration at the interface, ciS,
cannot be measured. However, we do know ciS has a finite
value because the osmotically driven water flux, Jw, is not
zero.
Draw Solute Flux Across the Active Layer. In order to
express the reverse flux in terms of experimentally accessible
quantities, we begin by considering the flux of draw solute
across the active layer. This flux, JsA, can be written as
JsA ) -

DA A
(c - 0)
tA i

(7)

where DA is the draw solute diffusion coefficient in the active


layer, tA is the active layer thickness, and ciA is the draw solute
concentration on the active layer side of the support layeractive layer interface. An ideal draw solute should not
permeate the active layer (i.e., the reflection coefficient should
be equal to 1); therefore, we have ignored any multicomponent diffusion effects in writing eq 7 (25, 26). An additional
term accounting for the influence of the water flux on the
salt flux would need to be included in eq 7 only if
multicomponent diffusion effects were important (25).
Analytical Expression for the Reverse Flux of Draw
Solute. The flux across the active layer, JsA, can be related to the
flux across the support layer by examining the interface (z )
0) between the two layers. Because no draw solute is accumulating or reacting at the interface, a mass balance yields
z ) 0 JsA ) JsS

(8)

Additionally, the draw solute concentration on the active layer


side of the interface can be related to the concentration on the
support layer side by assuming that the chemical potential is
equal across the interface
ciA ) HciS

(9)

where H is the partition coefficient describing the relative


concentration in each phase.
Equating eqs 6 and 7 and substituting eq 9 for ciA, the
reverse draw solute flux is expressed in terms of the
experimentally accessible bulk draw solution concentration cD
VOL. 44, NO. 13, 2010 / ENVIRONMENTAL SCIENCE & TECHNOLOGY

5171

Js )

JwcD
JwtA

) ( )

JwtS
exp
1- 1+ A
D
D H

JwcD
Jw
JwS
exp
1- 1+
B
D
(10)

) ( )

Note that the derived equation for the draw solute flux
contains two important transport parameters
S)

B)

tS

(11)

DAH
tA

(12)

The first term is the membrane structural parameter, S,


characterizing the average distance a solute molecule must
travel through the support layer when going from the bulk
draw solution to the active layer. The structural parameter,
S, can be determined from FO and RO experiments as
described elsewhere (14, 19). The second parameter, B, is
the active layer salt permeability coefficient (3, 26). This
parameter can be determined from RO or PRO experiments
as described later in this paper. Written in this form, eq 10
can be used to predict the draw solute reverse flux and
compared with FO mode experiments.

Materials and Methods


Model Draw Solute. ACS reagent sodium chloride (NaCl
crystals, J.T. Baker) was used as a draw solute because it is
highly soluble in water and its properties in solution are
well-characterized. For reverse permeation experiments,
NaCl was dissolved in deionized water (DI) obtained from
a Milli-Q ultrapure water purification system (Millipore,
Billerica, MA) at concentrations ranging from 0.5 to 4 M. The
osmotic pressures of these solutions were calculated using
a software package from OLI Systems, Inc. (Morris Plains,
NJ), and the binary diffusion coefficient for sodium chloride
and water was assumed constant at a value of 1.61 10-9
m2/s (27, 28). NaCl concentration in the feed solution was
measured using a calibrated conductivity meter (Oakton
Instruments, Vernon Hills, IL).
Forward Osmosis Membrane and Crossflow Setup. A
commercial asymmetric cellulose triacetate (HTI-CTA) membrane (Hydration Technology Innovations, Albany, OR) was
used for the reverse permeation experiments. This proprietary
membrane consists of a woven fabric mesh embedded within
a continuous polymer layer. The HTI-CTA membrane has
been used extensively in prior research exploring osmotically
driven membrane processes (20, 29).
The experimental crossflow FO system employed is similar
to that described in our previous studies (14, 20, 29). The
unit was custom built with channel dimensions measuring
77 mm long, 26 mm wide, and 3 mm deep on both sides of
the membrane. Variable speed gear pumps (Cole-Parmer,
Vernon Hills, IL) were used to pump the feed and draw
solutions cocurrently. No mesh feed spacers were used, and
the solutes were pumped in closed loops at 1.0 L/min,
corresponding to a crossflow velocity of 21.4 cm/s. A water
bath (Neslab, Newington, NH) maintained the temperature
of both the feed and draw solutions at 20 ( 0.5 C. The draw
solution reservoir rested upon a balance from Ohaus (Pine
Brook, NJ), and the change in mass as a function of time was
used to determine the water flux across the membrane.
Measurement of Membrane Properties in Reverse
Osmosis. The pure water permeability coefficient, A, and
NaCl permeability coefficient, B, of the HTI-CTA membranes
were evaluated in a laboratory-scale crossflow reverse
osmosis test unit. The effective membrane area was 20.02
cm2, and the crossflow velocity was fixed at 21.4 cm/s. During
5172

ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 44, NO. 13, 2010

experiments, the temperature was kept constant at 20 ( 0.5


C. Initially, the membrane was equilibrated with DI at an
applied pressure, P, of 27.6 bar (400 psi) until the permeate
flux reached a steady value (after about 10 min). After
equilibration, the volumetric permeate rate was measured
at applied pressures ranging from 6.9 to 27.6 bar (100 to 400
psi) in 6.9 bar (100 psi) increments. The water flux, Jw, at each
pressure drop was calculated by dividing the volumetric
permeate rate by the membrane area. The water permeability
coefficient, A, was obtained from the slope of water flux
plotted versus pressure drop.
NaCl rejection, R, was also determined at applied pressures
ranging from 6.9 to 27.6 bar (100 to 400 psi) in 6.9 bar (100
psi) increments. Using a 50-mM NaCl feed solution, the
observed rejection was determined from the difference in
bulk feed (cb) and permeate (cp) salt concentrations, R ) 1
- cp/cb. The NaCl permeability coefficient was determined
after correcting for concentration polarization using (3, 30)

( 1 -R R ) exp(- k )
Jw

B ) Jw

(13)

where k, the crossflow cell mass transfer coefficient of NaCl,


was calculated from correlations for rectangular cell geometry
(14, 31). In this equation, Jw and R correspond to the permeate
flux and NaCl rejection observed at each applied pressure.
Measurement of Reverse Draw Solute Permeation in
Forward Osmosis. The reverse NaCl flux, Js, was determined
using the following protocol. An HTI-CTA membrane was
placed in the custom built cross-flow cell. Both FO and PRO
mode tests were run, depending on the purpose of the
experiments. FO mode tests were carried out to test our newly
developed model for reverse solute permeation, while PRO
mode tests were run to verify the accuracy of the salt
permeability coefficient, B, determined from RO experiments.
After loading the membrane, both the feed and draw
solution reservoirs were filled with DI. The inlet and outlet
to the cross-flow cell were closed, isolating the membrane
from feed and draw solutions, and the draw solution reservoir
was dosed with the proper amount of NaCl stock solution.
The draw solution was then mixed by pumping the solution
in a closed loop for 10 min. After the draw solution was mixed,
the membrane was exposed to the feed and draw solutions,
and data recording was initiated. The mass of the draw
solution was monitored as a function of time to determine
the water flux, and the NaCl concentration in the feed was
monitored by submerging the conductivity meter probe at
20-min time intervals. It usually took about 20 to 30 min for
the water flux to stabilize; once this occurred, the reverse
solute flux was assumed to be at a steady-state. The data
collected after the water flux had stabilized was used to
calculate the experimental reverse draw solute flux.
Because the initial NaCl concentration in the feed is zero,
a species mass balance yields
cF(VF0 - JwAmt) ) JsAmt

(14)

where cF is the NaCl concentration in the feed, VF0 is the


initial volume of feed solution, Jw is the measured water flux,
Am is the membrane area, and t is time. This equation can
be linearized such that the slope of a plot of 1/cF versus 1/t
is equal to VF0/(JsAm), allowing the experimental draw solute
flux, Js, to be calculated
VF0 1
Jw
1
)
cF
JsAm t
Js

()

(15)

The active layer salt permeability coefficient B was also


determined from PRO mode experiments using DI water as
a feed on the support layer side and NaCl (0.5, 1.0, and 2.0

M) as a draw solution on the membrane active layer side.


The salt permeability coefficient B is calculated by dividing
the measured salt flux Js by the draw solute concentration
at the membrane active layer surface. The latter is obtained
after correcting for dilutive external concentration polarization (14), yielding
Js

B)

( )

cD exp -

Jw
k

(16)

where k is the mass transfer coefficient calculated for a


rectangular cell (31).

Results and Discussion


Membrane Performance Parameters. Five samples of the
HTI-CTA membranes were used to determine the performance parameters A and B. The water flux for a pure DI feed
was measured at four different applied pressures in RO
operation. Linear regression was then used to determine the
water permeability coefficient A from a plot of water flux
versus applied pressure, giving a value for A of 0.44 ( 0.12
L m-2 h-1 bar-1 (1.23 ( 0.33 10-12 m s-1 Pa-1). The relatively
large standard deviation is attributed to cutting smaller
samples from a large nonuniform flat sheet membrane. For
our calculations, we use the average A value of 0.44 L m-2h-1
bar-1.
The salt permeability coefficient B was also determined
in RO operation. The observed salt rejection was used to
determine B after correcting for external concentration
polarization using eq 13. The experiments yielded observed
percent rejections that ranged from 89.1% to 96.1%. The value
of B determined from these measurements was 0.261 ( 0.061
L m-2 h-1 (7.25 ( 1.69 10-8 m s-1). PRO mode was also used
to determine the B parameter to ensure that the value
determined from RO operation (i.e., high applied pressure)
accurately reflected the value in FO operation (i.e., no applied
pressure). The B parameter for a single membrane was
calculated from eq 16 to be 0.269 ( 0.020 L m-2 h-1 (7.47 (
0.55 10-8 m s-1), in good agreement with the RO
experiments. The average of the B values obtained in RO and
PRO modes, 0.265 L m-2h-1, was used for our calculations.
Membrane Structural Parameter. The membrane structural parameter S was calculated from (12, 14, 15, 19-22)
S)

()

B + ADb
D
ln
Jw
B + Jw

(17)

where Db is the osmotic pressure in the bulk draw solution.


The experimentally determined A and B values were used
along with the bulk osmotic pressure for the 1 M NaCl draw
solution (47.3 bar or 686.3 psi), and the water flux was
determined in FO mode with a 1 M NaCl draw solution and
DI water as the feed. Higher concentration draw solutions
were not used for this calculation because the osmotic
pressure deviated from the ideal vant Hoff behavior, an
assumption underlying the model to calculate S. The
nonideality of these solutions was observed as a deviation
of the osmotic pressure calculated using the OLI software
from that calculated using the vant Hoff equation. The S
parameter determined from the 1 M data was 481 m.
Reverse Draw Solute Flux as a Function of Draw Solution
Concentration. As the concentration of the draw solution
increases, the measured water flux and reverse draw solute
flux should both increase. Experiments were run using NaCl
draw solutions ranging in concentration from 1 to 4 M to
examine the dependency of water flux and reverse NaCl flux
on NaCl concentration. The results of these experiments,
plotted in Figure 2, show that, as anticipated, both the water
flux and reverse NaCl flux increase with increasing NaCl

FIGURE 2. Experimental water flux and reverse flux of NaCl as


a function of NaCl draw solution concentration. The water flux
and reverse solute flux were measured from experiments using
DI water as a feed solution and draw solutions of varying
concentration as indicated. The experiments were conducted
with a crossflow velocity of 21.4 cm/s, a constant temperature
of 20 C, and an ambient (unadjusted) solution pH of 5.7. The
osmotic pressures of the 1, 2, 3, and 4 M NaCl draw solutions
are 47.3, 105.3, 172.3, and 246.2 bar, respectively. The draw
solution osmotic pressure was calculated using a software
package from OLI Systems, Inc. (Morris Plains, NJ).

FIGURE 3. A comparison of experimental results and model


predictions for the reverse draw solute flux. The measured
solute fluxes are those presented in Figure 2, and the predicted
solute fluxes are calculated using eq 10, using the
corresponding water fluxes, Jw, presented in Figure 2. Values
for the transport parameters in eq 10 were as follows: A ) 0.44
L m-2 h-1 bar-1, B ) 0.265 L m-2 h-1, S ) 481 m, and DNaCl )
1.61 10-9 m2/s. The solid line (slope ) 1) represents perfect
agreement between experimental data and predictions.
concentration. This is consistent with prior experimental
(14, 15, 29) and modeling efforts (12, 14, 22) that have
investigated the relationship between measured water flux and
draw solution concentration. The reverse flux of draw solute
has only recently been explored through experiments (21).
However, this phenomenon has not yet been modeled, which
motivated our efforts to develop the model described above.
The values of the predicted and measured NaCl fluxes are
compared on a linear-linear plot in Figure 3. The agreement
between the model and experiments is strong, as demonstrated by the experimental data, located near the solid line
(slope ) 1), that represents perfect agreement between
experimental data and model predictions. The predicted
values in Figure 3 were calculated using the B and S values
determined in the preceding sections and the experimentally
measured water fluxes. The experimental water fluxes were
VOL. 44, NO. 13, 2010 / ENVIRONMENTAL SCIENCE & TECHNOLOGY

5173

FIGURE 4. Experimental salt flux as a function of the draw


solute concentration at the interface between the support and
active layers of the membrane. The experimental conditions for
measuring the solute flux are presented in Figure 2. The draw
solute concentration at the active layer surface was calculated
using eq 18, using the experimental water fluxes from Figure 2,
a value of A ) 0.44 L m-2 h-1 bar-1, and T ) 293 K.
used instead of attempting to predict Jw a priori because of
the nonideal behavior observed at higher draw solution
concentration. Future, more complex models can account
for the nonideal behavior at higher concentrations, but here
we continue using the experimental value because of the
physical insights our model can provide.
Reverse Draw Solute Flux as a Function of Interfacial
Concentration. It is interesting to explore what form NaCl
takes when permeating the active layer of the asymmetric
HTI-CTA membrane. To do this, we compare how the
measured reverse salt flux varies with the NaCl concentration
at the surface of the active layer, ciS. The osmotic pressure
difference across the active layer can be calculated using the
experimental water flux and the water permeability coefficient. Because the feed is DI, the osmotic pressure difference
is equal to the draw side osmotic pressure at the support
layer-active layer interface. We use the vant Hoff equation
to calculate the NaCl concentration at the active layer surface
Di )

Jw
) nRgTciS
A

(18)

where Di is the osmotic pressure of the draw solution at the


active layer surface, n is the number of dissolved species
created by the draw solute (2 for NaCl), Rg is the ideal gas
constant, and T is the absolute temperature.
The measured NaCl flux is plotted versus the concentration calculated from eq 18 in Figure 4. When these data are
fit using linear regression, the slope equals 0.251 L m-2 h-1
(6.97 10-8 m s-1)sa value close to the salt permeability
coefficient determined from RO and PRO experiments (0.265
L m-2 h-1 or 7.36 10-8 m s-1). The fact that this inferred
B value is consistent with that measured in PRO mode
indicates that there is no significant change in membrane
structure when it is exposed to high NaCl concentrations.
Therefore, the nonlinear dependence of flux on osmotic
pressure observed in FO mode is primarily a result of internal
concentration polarization, not osmotic deswelling (32).
The linear relationship between the NaCl flux and the
NaCl concentration at the surface of the active layer, cSi , further
indicates that individual ions, not ion pairs, are permeating
the active layer. If the ions were to form ion pairs before
diffusing across the active layer, the flux would depend on
the NaCl concentration squared (33). It is intriguing that
even though the ions permeate the active layer as two separate
entities, the process can be described by the single transport
coefficient B. Prior experimental work examined this aspect
5174

ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 44, NO. 13, 2010

of salt transport by quantifying the flux of individual ions


(21). For NaCl, it was found that the flux of Na+ and Cl- ions
were nearly equimolar, which explains why a single transport
coefficient can be used to describe the flux of NaCl. In order
to maintain electroneutrality, the ion that permeates the
active layer more quickly drags a counterion across the
membrane.
Implications for FO Membrane Design and System
Performance. The model developed in this work can help
in selecting a membrane that minimizes the loss of draw
solute into the feed solution, thereby reducing operating
costs. A previous publication has defined the ratio of the
reverse solute flux to the forward water flux as the specific
reverse salt flux (21). Because the ratio has units of concentration rather than flux, we suggest that a more appropriate
quantity is the inverse of this ratiosa value analogous to the
membrane selectivity (3). This ratio of the water flux to the
reverse solute flux, which we term the reverse flux selectivity,
can be regarded as the volume of water produced per the
moles (or mass) of draw solute lost.
An expression for the reverse flux selectivity is developed
using our model. First, the water flux is expressed as a function
of the osmotic pressure of the bulk draw solution, Db. The
nonlinear relationship between water flux and the bulk
osmotic pressure, a result of internal concentration polarization, is given for a DI feed by (14)

( )

Jw ) ADb exp -

( )

JwS
JwS
) AnRgTcD exp D
D

(19)

where the bulk osmotic pressure has been written using the
vant Hoff equation. Next, the reverse flux selectivity is
calculated by taking the ratio of eq 19 to eq 10 to give

[ ( ) (

AnRgT
JwS
Jw
Jw
)
exp - 1+
Js
Jw
D
B

)]

(20)

While valid, eq 20 is not immediately useful because it does


not explicitly express the ratio Jw/Js. Examining the term in
brackets allows the equation to be significantly simplified.
The water flux can range from 0 to infinity; however, the
term, exp(-JwS/D) - 1, will always fall between 0 and -1,
respectively. Therefore, the -Jw/B term dominates, and the
absolute value of the reverse flux selectivity reduces to
Jw A
nR T
Js B g

(21)

This is an insightful result. It states that the reverse flux


selectivity is independent of the support layer structural
parameter S as well as the bulk draw solution concentration.
The second implication of eq 21 can be tested experimentally.
This is done in Figure 5, which plots the reverse flux selectivity
versus the bulk NaCl concentration. As predicted, the
experimental data do not vary with concentrationsan
observation consistent with another recent study (21).
Furthermore, as shown by the solid line in Figure 5, eq 21
accurately predicts the reverse flux selectivity using only the
measured values of A and B.
The reverse flux selectivity is determined solely by the
selectivity of the active layer A/B and the ability of the draw
solute to generate an osmotic pressure, nRgT. That the reverse
flux selectivity is independent of the membrane structural
parameter is consistent with our physical understanding of
the system. A high concentration of draw solute at the support
layer - active layer interface is necessary to generate a large
osmotic gradient, which drives a high water flux. However,
this higher concentration of draw solute also increases the
concentration gradient across the active layer, which increases the reverse salt flux. For an ideal solution, the osmotic

DS

FIGURE 5. Reverse flux selectivity as a function of draw solute


concentration. The experimental values were determined using
the measured water fluxes and reverse solute fluxes depicted
in Figure 2. The theoretical prediction was calculated using eq
21, with A ) 0.44 L m-2 h-1 bar-1, B ) 0.265 L m-2 h-1, n ) 2, T
) 293 K, and Rg ) 8.3145 10-2 L bar K-1 mol-1.
gradient is proportional to the concentration gradient, and,
therefore, the ratio of the two quantities remains constant
(34).
The analysis above highlights the need to select a
membrane with a highly selective active layer (i.e., high A
and low B) and a draw solute capable of generating a large
osmotic pressure, but it does not diminish the importance
of reducing the structural parameter S of forward osmosis
membranes. An FO process needs to achieve high water fluxes
at low draw solution concentrations to minimize the energy
required to separate fresh water from the diluted draw
solution and reconcentrate/recycle the draw solution. Reducing this energy can only be realized by having FO
membranes with small S so that the effects of ICP are
minimized.
Effectively operating an osmotically driven membrane
process requires selecting a membrane and draw solute
pair to maximize the forward water flux, minimize the
reverse draw solute flux, and achieve an efficient and
effective separation of the product water from the draw
solution among other considerations. The model developed
in this work provides a simple basis (eq 21) for selecting
a membrane and draw solute system that maximizes the
water flux while minimizing the loss of draw solute. In the
future, this criterion can be used in concert with other
design heuristics to help optimize a forward osmosis
process.

NOMENCLATURE
A
Am
B
c
cD
cF
ciA
ciS
D
DA

water permeability coefficient


membrane area
draw solute permeability coefficient
molar concentration of draw solute
molar concentration of draw solute in the bulk draw
solution
molar concentration of draw solute in the bulk feed
solution
molar concentration of draw solute on the active
layer side of the interface between the support
layer and active layer
molar concentration of draw solute on the support
layer side of the interface between the support
layer and active layer
bulk diffusion coefficient of draw solute in water
diffusion coefficient of draw solute in active layer

H
Js
Jw
n
R
Rg
S
t
T
tA
tS
VF
z

effective diffusion coefficient of the draw solute in


support layer
solute partition coefficient
total draw solute flux
water flux
number of dissolved species created by draw solute
solute rejection
ideal gas constant
membrane structural parameter
time
absolute temperature
thickness of active layer
thickness of support layer
feed solution volume
coordinate system
porosity of support layer
osmotic pressure
tortuosity of support layer

Acknowledgments
The work was supported by the WaterCAMPWS, a Science
and Technology Center of Advanced Materials for the
Purification of Water with Systems under the National Science
Foundation Grant CTS-0120978; and Oasys Water Inc.

Literature Cited
(1) Cath, T. Y.; Childress, A. E.; Elimelech, M. Forward osmosis:
Principles, applications, and recent developments. J. Membr.
Sci. 2006, 281, 7087.
(2) McGinnis, R. L.; Elimelech, M. Global Challenges in Energy and
Water Supply: The Promise of Engineered Osmosis. Environ.
Sci. Technol. 2008, 42, 86258629.
(3) Baker, R. W. Membrane technology and applications, 2nd ed.;
J. Wiley: Chichester, New York, 2004.
(4) Holloway, R. W.; Childress, A. E.; Dennett, K. E.; Cath, T. Y.
Forward osmosis for concentration of anaerobic digester
centrate. Water Res. 2007, 41, 40054014.
(5) Mi, B.; Elimelech, M. Chemical and physical aspects of organic
fouling of forward osmosis membranes. J. Membr. Sci. 2008,
320, 292302.
(6) Lay, W. C. L.; Chong, T. H.; Tang, C. Y. Y.; Fane, A. G.; Zhang,
J. S.; Liu, Y. Fouling propensity of forward osmosis: investigation
of the slower flux decline phenomenon. Water Sci. Technol.
2010, 61, 927936.
(7) Kravath, R. E.; Davis, J. A. Desalination of Sea-Water by Direct
Osmosis. Desalination 1975, 16, 151155.
(8) Tan, C. H.; Ng, H. Y. A novel hybrid forward osmosis nanofiltration (FO-NF) process for seawater desalination: Draw
solution selection and system configuration. Desalin. Water
Treat. 2010, 13, 356361.
(9) Choi, J. S.; Kim, H.; Lee, S.; Hwang, T. M.; Oh, H.; Yang, D. R.;
Kim, J. H. Theoretical investigation of hybrid desalination system
combining reverse osmosis and forward osmosis. Desalin. Water
Treat. 2010, 15, 114120.
(10) Cartinella, J. L.; Cath, T. Y.; Flynn, M. T.; Miller, G. C.; Hunter,
K. W.; Childress, A. E. Removal of natural steroid hormones
from wastewater using membrane contactor processes. Environ.
Sci. Technol. 2006, 40, 73817386.
(11) Achilli, A.; Cath, T. Y.; Marchand, E. A.; Childress, A. E. The
forward osmosis membrane bioreactor: A low fouling alternative
to MBR processes. Desalination 2009, 239, 1021.
(12) Loeb, S.; Titelman, L.; Korngold, E.; Freiman, J. Effect of porous
support fabric on osmosis through a Loeb-Sourirajan type
asymmetric membrane. J. Membr. Sci. 1997, 129, 243249.
(13) Lonsdale, H. K.; Merten, U.; Riley, R. L. Transport Properties of
Cellulose Acetate Osmotic Membranes. J. Appl. Polym. Sci. 1965,
9, 1341&.
(14) McCutcheon, J. R.; Elimelech, M. Influence of concentrative
and dilutive internal concentration polarization on flux
behavior in forward osmosis. J. Membr. Sci. 2006, 284, 237
247.
(15) McCutcheon, J. R.; Elimelech, M. Modeling water flux in forward
osmosis: Implications for improved membrane design. AIChE
J. 2007, 53, 17361744.
(16) Ng, H. Y.; Tang, W. L.; Wong, W. S. Performance of forward
(direct) osmosis process: Membrane structure and transport
phenomenon. Environ. Sci. Technol. 2006, 40, 24082413.
VOL. 44, NO. 13, 2010 / ENVIRONMENTAL SCIENCE & TECHNOLOGY

5175

(17) Yang, Q.; Wang, K. Y.; Chung, T. S. Dual-Layer Hollow Fibers


with Enhanced Flux As Novel Forward Osmosis Membranes for
Water Production. Environ. Sci. Technol. 2009, 43, 28002805.
(18) Wang, R.; Shi, L.; Tang, C. Y.; Chou, S.; Qiu, C.; Fane, A. G.
Characterization of novel forward osmosis hollow fiber membranes. J. Membr. Sci. 2010, 355, 158167.
(19) Yip, N. Y.; Tiraferri, A.; Phillip, W. A.; Schiffman, J. D.; Elimelech,
M. High Performance Thin-Film Composite Forward Osmosis
Membrane. Environ. Sci. Technol. 2010, 44, 38123818.
(20) McCutcheon, J. R.; McGinnis, R. L.; Elimelech, M. A novel
ammonia-carbon dioxide forward (direct) osmosis desalination
process. Desalination 2005, 174, 111.
(21) Hancock, N. T.; Cath, T. Y. Solute Coupled Diffusion in
Osmotically Driven Membrane Processes. Environ. Sci. Technol.
2009, 43, 67696775.
(22) Lee, K. L.; Baker, R. W.; Lonsdale, H. K. Membranes for PowerGeneration by Pressure-Retarded Osmosis. J. Membr. Sci. 1981,
8, 141171.
(23) Cussler, E. L. Diffusion: mass transfer in fluid systems, 3rd ed.;
Cambridge University Press: Cambridge, New York, 2009.
(24) Dullien, F. A. L. Porous media: fluid transport and pore structure,
2nd ed.; Academic Press: San Diego, 1992.
(25) Kedem, O.; Katchalsky, A. Thermodynamic Analysis of the
Permeability of Biological Membranes to Non-Electrolytes.
Biochim. Biophys. Acta 1958, 27, 229246.
(26) Paul, D. R. Reformulation of the solution-diffusion theory of
reverse osmosis. J. Membr. Sci. 2004, 241, 371386.

5176

ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 44, NO. 13, 2010

(27) Lobo, V. M. M. Mutual Diffusion-Coefficients in AqueousElectrolyte Solutions (Technical Report). Pure Appl. Chem. 1993,
65, 26142640.
(28) Chang, Y. C.; Myerson, A. S. The Diffusivity of PotassiumChloride and Sodium-Chloride in Concentrated, Saturated, and
Supersaturated Aqueous-Solutions. AIChE J. 1985, 31, 890894.
(29) McCutcheon, J. R.; McGinnis, R. L.; Elimelech, M. Desalination
by ammonia-carbon dioxide forward osmosis: Influence of draw
and feed solution concentrations on process performance. J.
Membr. Sci. 2006, 278, 114123.
(30) Mulder, M. Basic principles of membrane technology, 2nd ed.;
Kluwer Academic: Dordrecht, Boston, 1996.
(31) Hoek, E. M. V.; Kim, A. S.; Elimelech, M. Influence of crossflow
membrane filter geometry and shear rate on colloidal fouling
in reverse osmosis and nanofiltration separations. Environ. Eng.
Sci. 2002, 19, 357372.
(32) Mehta, G. D.; Loeb, S. Performance of Permasep B-9 and B-10
Membranes in Various Osmotic Regions and at High Osmotic
Pressures. J. Membr. Sci. 1979, 4, 335349.
(33) Reusch, C. F.; Cussler, E. L. Selective Membrane Transport. AIChE
J. 1973, 19, 736741.
(34) Tang, C. Y.; She, Q.; Lay, W. C. L.; Wang, R.; Fane, A. G. Coupled
effects of internal concentration polarization and fouling on
flux behavior of forward osmosis membranes during humic
acid filtration. J. Membr. Sci. 2010, 354, 123133.

ES100901N

Anda mungkin juga menyukai