Latest FO Paper

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Journal of Membrane Science xxx (xxxx) xxx

Contents lists available at ScienceDirect

Journal of Membrane Science


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

An improved model for membrane characterization in forward osmosis


Jeffrey T. Martin, Georgios Kolliopoulos, Vladimiros G. Papangelakis *
Department of Chemical Engineering and Applied Chemistry, University of Toronto 200 College Street, Toronto, Ontario, M5S 3E5, Canada

A R T I C L E I N F O A B S T R A C T

Keywords: As a promising technology for industrial water recovery, forward osmosis (FO) requires a rigorous and accurate
Forward osmosis transport model to determine its efficacy for water recovery from challenging effluents, yet current models
Permeability neglect the non-ideality of concentrated draw solutions, and mass-transfer boundary layers within the process.
Membrane modeling
This work builds upon recent FO transport models and presents an improved FO-based membrane character­
Desalination
ization method that addresses the non-ideality of concentrated draw solutions, physical properties that are not
based on the bulk draw solution, and all instances of concentration polarization. Using the FO-based charac­
terization method of the present work, consistent water permeability and structural parameter values are ob­
tained with R2 > 0.985 for multiple inorganic draw solutions. When compared to additional experimental FO
transport data, improvements of up to 107% were observed over the existing FO-based characterization method.
Using the modeling approach developed in this work, the assessment for a potential FO application is refined by
providing the means for accurate process modeling and membrane characterization.

1. Introduction polarization occurs at three interfaces: external concentration polari­


zation (ECP) at the feed solution and active membrane interface (FECP);
Industrial wastewater treatment poses a substantial economic and external concentration polarization at the draw solution and support
environmental challenge, for it is necessary to recover fresh water for membrane interface (DECP); and internal concentration polarization
reuse while reducing the volume of the effluent. Compared to traditional (ICP) within the porous membrane support layer [31]. These polariza­
water recovery technologies, forward osmosis (FO) is an attractive so­ tions are illustrated in Fig. 1 by the decline in the draw solute concen­
lution due to its reduced membrane fouling potential [1–8] and tration, cs, between that of the bulk draw-side solution and that of the
favourable energy consumption [9,10]. The potential for FO use is active layer-porous support interface. To accurately model the osmotic
widespread, with current research focusing in hydrometallurgical process, it is vital to estimate concentration dependent physical prop­
[11–16], petrochemical [17–24], and seawater desalination [25–29] erties (i.e., osmotic pressure, diffusivity, dynamic viscosity, and density)
applications. using the concentration at each respective interface.
Driven by an osmotic pressure differential, forward osmosis uses a The transport of water across the membrane in FO has previously
concentrated draw solution to spontaneously pull water across a semi­ been modelled to provide estimates of the power generation capacity in
permeable membrane from an effluent, rejecting and concentrating its pressure-retarded osmosis systems [32–34]. However, these early
solutes. Conjunctly, the draw solute leaks in the opposite direction of the models assumed that the reverse draw solute flux was negligible [32,
water flux at a much slower rate, driven by the concentration gradient 33], taking into account only the ICP, neglecting thereby any external
across the membrane [30]. The transport profile for FO is illustrated in concentration polarization. Subsequent improvements to FO modeling
Fig. 1. considered the reverse draw solute flux and additional forms of con­
Due to the selective transport of water across the membrane, solutes centration polarization [31,34–38]. However, these models did not take
are rejected by the membrane and accumulate at the membrane-solution into account the nonideality of concentrated draw solutions [39] and
interface. This results in a localized concentration gradient defined as were derived using a tacit assumption that Van’t Hoff’s law remains
concentration polarization. For traditional membrane processes like valid under all conditions. To derive and simplify the water and draw
reverse osmosis, concentration polarization occurs only at the feed so­ solute transport equations of these models, an assumption was made
lution side. However, for forward osmosis (FO), concentration that the osmotic pressure is linearly proportional to the draw solute

* Corresponding author.
E-mail address: vladimiros.papangelakis@utoronto.ca (V.G. Papangelakis).

https://doi.org/10.1016/j.memsci.2019.117668
Received 7 September 2019; Received in revised form 10 November 2019; Accepted 14 November 2019
Available online 16 November 2019
0376-7388/© 2019 Elsevier B.V. All rights reserved.

Please cite this article as: Jeffrey T. Martin, Journal of Membrane Science, https://doi.org/10.1016/j.memsci.2019.117668
J.T. Martin et al. Journal of Membrane Science xxx (xxxx) xxx

concentration, as outlined by Van’t Hoff’s law, which is only valid at equations, which is solved using an in-house developed algorithm and
infinite dilution [40]. As a result, existing FO models miscalculate the intrinsic membrane parameters obtained solely from FO experimental
osmotic pressure and inaccurately predict systems with draw solutions data. The corresponding membrane characterization method considers
of high concentrations [41]. Further improvements excluded the Van’t the non-ideal osmotic pressure and all forms of concentration polari­
Hoff assumption, requiring iterative models to determine the water flux zation. When validated against the results from a second set of experi­
[42,43] and reverse draw solute flux [44,45]. However, these updated mental data and compared to the results from previous FO modeling
models did not consider all forms of external concentration polarization attempts in literature, this improved membrane characterization
[42,43], and do not properly take into account the draw solute diffu­ method and modeling approach produced significantly better agreement
sivity corrected for the thermodynamic driving force [44,45]. Addi­ with experimental measurements for three inorganic draw solutions.
tionally, it was shown that the assumption of constant solution physical
properties can cause substantial errors when modeling the FO process 2. Development of improved forward osmosis model
[46].
The permeability values of both the water and the draw solute were The water flux across the membrane active layer Jw (L/m2/h) for an
estimated using a methodology under pressurised conditions via reverse osmotically-driven process at steady state, is calculated by Equation (1)
osmosis [38], followed by an osmotically-driven experiment to estimate [50,51]:
the membrane structural parameter, S (the ratio of the support layer �
Jw ¼ AðΔπÞ ¼ A πda πfa ; (1)
thickness, corrected by the tortuosity, to the support layer porosity). The
validity of this approach has been questioned [38,47,48], as FO mem­
where A is the permeability of water across the membrane (L/m2/h/
branes behave differently under pressurised conditions [49], thus
bar), π is the osmotic pressure (bar) of the feed or the draw solution at
affecting their transport properties. The only attempt to develop a spe­
the interface of the active layer solution side (Fig. 1). Subscripts d, f, a
cific to FO approach by Tiraferri et al. [38], consisted of four FO ex­
refer to the draw side of the system, feed side, and the active layer
periments at different draw solution concentrations, assumed constant
interface solution side.
draw solute diffusivity, assumed Van’t Hoff’s law holds across all draw
Similarly, the salt flux at steady state, Js (mol/m2/h), is calculated by
solute concentrations, and also assumed negligible external concentra­
Equation (2) [50,51]:
tion polarization.

Extending previous FO models [44,45], in this work we present an Js ¼ B cs;da cs;fa ; (2)
improved FO transport model, derived from first principles, that ad­
dresses both the inaccuracy of assuming bulk solution properties where B is the salt membrane permeability (L/m2/h), and cs,fa is the
throughout the membrane and all instances of concentration polariza­ concentration of solutes (mol/L) at the active layer interface solution/
tion. It follows similar methodology with the work presented previously feed side and cs,da at the interface support layer side (Fig. 1). A linear
in Refs. [44,45]. It consists of a system of non-linear isothermal transport gradient is assumed in the active layer according to the solution

Fig. 1. Steady state concentration profile across an asymmetric forward osmosis membrane.

2
J.T. Martin et al. Journal of Membrane Science xxx (xxxx) xxx

� dcs
Js ¼ B cs;da cs;fa ¼ Ds;da ðcs Þε Jw cs : (5)
dz
Rearranging Equation (4) results in:
dcs Js þ Jw cs
¼ : (6)
dz Ds;d ðcs Þε

Which can be solved using the following boundary conditions where the
membrane support layer length is corrected by its tortuosity, τ
(dimensionless), the ratio of the actual flow path length to that of the
thickness of the porous support:
cs ¼ cs;da at z ¼ ta

cs ¼ cs;dp at z ¼ ts τ þ ta

For instances where the porous support layer tortuosity is unknown,


a variable transform is required:
z ta
bz ¼ ; (7)
ε

resulting in:
dcs Js þ Jw cs
¼ ; (8)
dbz Ds;d ðcs Þ

with the following adjusted boundary conditions:


cs ¼ cs;da at bz ¼ 0

ts τ ts τ
cs ¼ cs;dp at bz ¼ þ ta ta ¼ ¼S
ε ε

where S is the membrane structural parameter (μm) defined by:


Fig. 2. Algorithm developed to determine the water flux, reverse draw solute ts τ
S¼ : (9)
flux, and solute concentrations at the active layer and porous support. ε
Similarly, a mass balance is performed for the DECP at the draw
diffusion model, as the active layer is a dense polymer where mass solution-porous support boundary layer (Peclet number 0.26 to 3), using
transport occurs only through the diffusion of the species through spaces the following boundary conditions to obtain cs,dp.
in the polymer chains [50,51]. This is further justified by the Peclet
number, shown in Equation (3) [50], which is less than 0.08, indicating cðzÞ ¼ cs;db at z ¼ ta þ ts þ δd
that diffusion forces dominate, and convection is negligible in the active
cðzÞ ¼ cs;dp at z ¼ ta þ ts
layer.
Jw dcs
Pe ¼ (3) Js ¼ Ds;d ðcs Þ Jw cs ; (10)
Ds t dz
To consider internal concentration polarization in the porous support dcs Js þ Jw cs
boundary layer, the draw solute flux can be described by the sum of the ¼ ; (11)
dz Ds;d ðcs Þ
concentration driven diffusion component, and the convection compo­
nent resulting from the water flux in the opposite direction (Peclet where δd is the thickness of the draw-side external boundary layer (m),
number of 0.06–1.4 for the conditions examined). This is represented by defined by:
Equation (4),
Ds;db
dcs δd ¼ ; (12)
Js ¼ Ds;d ðcs Þε Jw cs ; (4) kd
dz
and kd is the draw-side mass transfer coefficient obtained from Equation
where Ds represents the diffusivity of the draw solute in the draw solu­
tion (m2/s). By “solute” it is meant an average value for both anion and
cation in case of a strong electrolyte, corrected by the thermodynamic Table 1
driving force [52]. In turn, ε represents the support layer porosity Draw solution materials.
(dimensionless), and z represents the distance (m) across the support Chemical Name CAS No. Source Purity-
layer from the active layer to the draw solution. The support layer Assaya
porosity defines the fraction of the volume of voids (volume for the CaCl2⋅2H2O 10035-04- ACROS 99þ%
solution to flow through), over the total volume of the support layer, and 8 ORGANICS
is included with the tortuosity to detail the actual transport path through NaCl Certified ACS 7647-14-5 Fisher Sci. 99.6%
Crystalline
the porous support. At steady state the water and draw solute fluxes
MgCl2 (anhydrous) 7786-30-3 Afla Aesar 99%
across the entire membrane are constant, and therefore: DI water 7732-18-5 Milli-Q Reference 18.2 MΩ cm
a
As stated by the supplier.

3
J.T. Martin et al. Journal of Membrane Science xxx (xxxx) xxx

Table 3
Draw solution osmotic pressure empirical parameters.
Draw Solution X1,π X2,π

NaCl 6.42 39.26


CaCl2 52.55 13.01
MgCl2 61.00 6.73

Table 4
Draw solute diffusivity empirical parameters.
Draw Solution X1,D X2,D X3,D

NaCl 6.92E-12 9.95E-11 1.51E09


CaCl2 1.50E-11 3.13E-10 1.34E-09
MgCl2 3.14E-11 3.82E-10 1.24E-09

Table 5
Draw solution density empirical parameters.
Draw Solution X1,ρ X2,ρ
Fig. 3. Experimental apparatus used obtain the experimental data required for
NaCl 0.04 1
model parametrization and validation.
CaCl2 0.08 1
MgCl2 0.07 1
(16). The effect of the water flux was found to have little effect on the
boundary layer thickness as it is four orders of magnitude less than that
of the crossflow velocity (5 � 10 6 m/s vs. 0.021 m/s). Table 6
Draw solution dynamic viscosity empirical parameters.
Similarly, the external feed-side boundary layer is represented by the
following mass balance (Peclet number 0.25 to 0.86): Draw Solution X1,μ X2,μ X3,μ

NaCl 0.02 0.06 0.89


cðzÞ ¼ cs;fb at z ¼ 0
CaCl2 0.31 0.34 0.89
MgCl2 0.42 0.32 0.89
cðzÞ ¼ cs;fa at z ¼ δf

dcs characteristic length (m) of the respective channel. The Reynolds


Js ¼ Ds;f ðcs Þ Jw cs ; (13)
dz number is determined using the following equation:
ρvdh
dcs Js þ Jw cs
¼ ; (14) Re ¼ ; (19)
dz Ds;f ðcs Þ μ

where δf is the feed-side external boundary layer thickness (m), defined where μ represents the dynamic viscosity (kg/m/s), ρ the density (kg/
by: m3), and v the velocity (m/s) of the solution respectively. The Reynolds
numbers for the conditions used were <4.3. Additionally, the Sc Schmidt
δf ¼
Ds;f
: (15) number is calculated using Equation (20):
kf
μ
Sc ¼ : (20)
The mass transfer coefficients kd and kf (m/s) of a solute in the draw ρDs
solution or feed stream respectively, is obtained from the Sherwood
All draw solute and solution physical properties (μ, ρ, D, and π) are
number equation,
determined using concentration dependent empirical equations ob­
Shd Ds;db tained from OLI Studio 9.6 simulation data (as explained in Section 3.2).
kd ¼ ; (16)
dh To improve the accuracy of the model, the empirical correlation for the
draw solute diffusivity, Ds, is substituted into each of the three ODE’s
kf ¼
Shf Ds;fb
; (17) presented (Equations (8), (11) and (14)) and solved over their respective
dh boundary conditions.
The proposed model for this work is a system of three ODE’s, to
where Sh refers to the Sherwood number and dh the hydraulic diameter
obtain the active-layer interfacial draw solute concentrations from the
(m) of the respective channel through which the fluid is flowing. The
bulk solution conditions, and two additional equations for the water and
Sherwood number is obtained from the following correlation for laminar
draw solute flux across the active layer. The system of equations is as
flow [53]:
follows:
� �0:33 �
dh
Sh ¼ 1:85 ReSc (18) Jw ¼ AΔπ ¼ A πda πfa ;
L

Js ¼ BΔcs;a ¼ B cs;da cs;fa ;
where Re is the Reynolds number, Sc is the Schmidt number, and L is the

Table 2
Model regression and validation experiments.
Type of Experiment NaCl Draw Solution Concentrations (mol/L) MgCl2 Draw Solution Concentrations (mol/L) CaCl2 Draw Solution Concentrations (mol/L)

Model Parameterization 1, 2, 3, 3.5 1, 2, 3, 3.5 1, 2, 3, 3.5


Model Validation 0.5, 2.5, 4 0.5, 2.5, 4 0.5, 2.5, 4

4
J.T. Martin et al. Journal of Membrane Science xxx (xxxx) xxx

Fig. 6. Experimental reverse draw solute flux performance as a function of


Fig. 4. Water flux profile used to determine steady state in the forward osmosis draw solute concentration for NaCl, MgCl2, and CaCl2 draw solutes.
experiments (2 M NaCl Draw Solution and DI Water Feed at 25 � C).

obtained from Sterlitech (active area of 42 cm2). An asymmetric cellu­


lose triacetate (CTA) membrane was used, obtained from Fluid Tech­
ICP dcs Js þ Jw cs nology Solutions. The membrane was oriented with the active layer
¼ ;
dbz Ds;d ðcs Þ
facing the feed solution side, which is consistent with previous FO
FECP dcs Js þ Jw cs
dz
¼
Ds;d ðcs Þ
; membrane characterization work [38]. The FO cell was run
DECP dcs Js þ Jw cs counter-currently, and the flow rates were controlled by a
Thermo-Fisher Easy Load II peristaltic pump.
¼ ;
dz Ds;f ðcs Þ
A summary of the draw solution concentrations used for the model
parameterization and validation is shown in Table 2. The intrinsic
with additional empirical equations used to determine Ds,i and πi as membrane parameters were obtained using draw solute concentrations
functions of the solute concentration in the respective stream i (listed in of 1, 2, 3, and 3.5 mol/L. The subsequent model validation was per­
Section 3.2). This system of equations is solved using the algorithm formed using additional experimental data including two concentrations
presented in Fig. 2, with each ODE solved using the ode 45 function in outside of the fitting range (i.e. 0.5, 2.5, and 4 mol/L).
MATLAB R2019b, using initial guesses for Js and Jw. For each experiment DI water was used as the feed stream to
establish a baseline for FO performance and to maintain consistency
3. Materials and methods with previous FO membrane characterization work [38]. The draw so­
lutions were prepared by dissolving the appropriate amount of the draw
3.1. Membrane, FO cell and materials solute in DI water. The draw and feed solutions were recirculated each in
closed loops using a solution reservoir of 2 L to ensure an approximately
The properties of the draw solution chemicals used are shown in constant solution concentration during the experiment duration. A
Table 1. crossflow velocity of 2.1 cm/s was used for both the feed and draw so­
Fig. 3 provides a schematic of the forward osmosis membrane lutions. Every 30 s the mass of each reservoir was recorded digitally by
apparatus, featuring a rectangular stainless steel CF042-FO cell, Mettler-Toledo balances (NewClassic MF MS4002S). Samples of both the
feed and draw solutions were taken initially at steady state (achieved
within 15 min as explained in Section 4.1), and again at the conclusion
of the experiment (after 1 h). Metal ion concentrations in the feed and
draw solutions were measured using inductively coupled plasma optical
emission spectrometry (Agilent 700 series ICP-OES). Prior to analysis,
each sample was diluted with 5% nitric acid: 50x for the feed samples,
and 1000x for the draw solution samples. The water flux was calculated
based on the change of the draw solution mass, while the reverse salt
flux was determined based on the change in the solute concentration in
the feed. The experiments were carried out in triplicates to ensure the
reproducibility of the results.

Table 7
Membrane intrinsic parameter results.
Draw A (L/m2/h/ B (L/ S (μm) R2Jw R2Js
Solution bar) m2/h)

NaCl 0.724 0.651 397.9 0.999 0.999


CaCl2 1.092 0.376 221.21 0.992 0.993
MgCl2 0.739 0.385 215.99 0.985 0.989
Average 0.851 � N/A 278.4 � 0.992 � 0.994 �
Fig. 5. Experimental water flux performance as a function of draw solute
0.09 48.8 0.003 0.002
concentration for NaCl, MgCl2, and CaCl2 draw solutes.

5
J.T. Martin et al. Journal of Membrane Science xxx (xxxx) xxx

Fig. 7. Activity coefficient of water as a function of draw solute concentration


Fig. 8. Regressed draw solute permeability values as a function of the hydrated
for NaCl, MgCl2, and CaCl2 draw solutes. Data was obtained using the Mixed
cation radius of the respective draw solute (Naþ, Mg2þ, and Ca2þ).
Solvent Electrolyte (MSE) model in OLI Studio 9.6.

4.2. Experimental results


3.2. Draw solution physical properties

Comparing the experimental water flux results for each draw solute
The physical properties of each draw solution (osmotic pressure,
in Fig. 5, it is observed that using a MgCl2 draw solution provides the
diffusivity of the draw solute in the draw solution, density, and dynamic
greatest water flux, closely followed by CaCl2. This is expected due to the
viscosity) were determined using the Mixed-Solvent Electrolyte (MSE)
greater osmotic pressures exhibited by divalent electrolytes when
model in OLI Studio 9.6. For ease of calculation, empirical polynomial
compared to the monovalent NaCl on a molar basis.
equations were fitted to the OLI data as a function of the draw solution
Contrary to the water flux, the NaCl draw solution produces the
concentration (mol/L) for each physical property dataset. All data was
highest reverse draw solute flux (Fig. 6), followed by CaCl2 and MgCl2.
fitted with an R2 > 0.98 for draw solute concentrations of 0 mol/L to
This trend follows the established draw solute permeability trends in
their respective solubility limits at 25 � C and 1 atm. The parameters and
Section 4.3, which are explained by the hydrated ion size and hydration
equations for osmotic pressure, diffusivity, density, and dynamic vis­
enthalpy. It appears that MgCl2 produces the highest water flux with the
cosity are detailed below in Equations (21)–(24) and Tables 3–6
lowest reverse draw solute flux. Although to a minor degree, the higher
respectively. The coefficients of the polynomials are represented by Xn,p,
reverse draw solute flux exhibited by NaCl will further reduce the
where the subscripts n and p refer to the number of the coefficient and
driving force for water transport, therefore further reducing the water
the corresponding physical property respectively.
flux when compared to the divalent draw solutes.
π ¼ X1;π c2s þ X2;π cs (21) To ensure the quality of the experimental data, the coefficient of
variation (CV) for the ratio of the water flux to the reverse draw solute
D ¼ X1;D c2s þ X2;D cs þ X3;D (22) flux, Jw/Js, is determined. The ratio of the two fluxes is only dependent
on the selectivity of the active layer and is an indicator of the quality of
ρ ¼ X1;ρ cs þ X2;ρ (23) the experimental FO data. The CV is defined as the standard deviation of

μ ¼ X1;μ c2s þ X2;μ cs þ X3;μ (24)

4. Results and discussion

4.1. Determination of steady state

Prior to the sampling of the feed and draw solutions, steady state
across the membrane had to be achieved. This was determined by
plotting the change in the mass flow rate of the draw solution against
time, until a constant mass flow rate was observed (�5%), indicating a
constant water flux. As seen in Fig. 4, steady state was typically achieved
within 5 min, however the sampling was performed after 15 min to
ensure a stable steady state in all experiments.

Table 8
Hydrated radii and hydration enthalpy of draw solute cations.
Ion B (L/m2/ Hydrated Radius [61] Hydration Enthalpy [62]
h) (Å) (kJ/mol)

Naþ 0.396 3.58 416 Fig. 9. Regressed draw solute permeability values as a function of the absolute
Ca2þ 0.301 4.12 1602
hydration enthalpy of the cation of the respective draw solute (Naþ, Mg2þ,
Mg2þ 0.267 4.28 1949
and Ca2þ).

6
J.T. Martin et al. Journal of Membrane Science xxx (xxxx) xxx

Fig. 10. Comparing the predicted water flux as a function of the NaCl bulk Fig. 12. Comparing the predicted water flux as a function of the MgCl2 bulk
draw solute concentration using the model of this work and previous models of draw solute concentration using the model of this work and previous models of
Chowdhury & McCutcheon [45], Nagy [63], and Tiraferri [38]. Chowdhury & McCutcheon [45], Nagy [63], and Tiraferri [38].

the experimental ratio of Jw/Js divided by the arithmetic mean. The !2 !2


closer the value of CV to zero, the better the fit of data, with a CV < 10% X J Exp J Calc X J Exp J Calc
n n
(25)
w;i w;i s;i s;i
E ¼ EJw þ EJs ¼ þ ;
recommended [38]. The coefficient of variation values for NaCl, CaCl2, i¼1
Exp;n
J w;i i¼1
Exp;n
J s;i
and MgCl2 are 2.8%, 1.5%, and 3.9% respectively. These are far below
the recommended 10%, justifying the integrity of the presented exper­ where n refers to the number of FO experiments performed and super­
imental results. scripts Exp and Calc refer to calculated and experimental values
respectively. The error for both Jw and Js are scaled by the average
4.3. Membrane parameterization experimental value of Jw and Js respectively, to avoid a biased global
error. The calculated values were obtained using the model and algo­
Building upon the method of Tiraferri et al. [38], four FO experi­ rithm presented in Section 2.0. However, unlike the method in Tiraferri
ments were performed each at different draw solution concentrations, et al. [38], this method does not assume constant Ds, ideal osmotic
and from these experiments eight data points were obtained (four water pressure, and negligible external concentration polarization. The initial
flux and four reverse draw solute flux corresponding measurements). guesses for A, B, and S were 0.5 L/m2/h/bar, 0.4 L/m2/h, and 200 μm
These eight data points were then used to regress the three intrinsic respectively. To justify the wide-spread applicability of this model and
membrane parameters; water permeability (A), draw solute perme­ membrane characterization approach, intrinsic parameter regressions
ability (B), and membrane structural parameter (S) by a least-squares were performed for each draw solute. Table 7 displays the intrinsic
minimization of the global error (E): parameters obtained using this method:
Using the proposed model and FO experimental method, consistent

Fig. 11. Comparing the predicted reverse draw solute flux as a function of the Fig. 13. Comparing the predicted reverse draw solute flux as a function of the
NaCl bulk draw solute concentration using the model of this work and previous MgCl2 bulk draw solute concentration using the model of this work and pre­
models of Chowdhury & McCutcheon [45], Nagy [63], and Tiraferri [38]. vious models of Chowdhury & McCutcheon [45], Nagy [63], and Tiraferri [38].

7
J.T. Martin et al. Journal of Membrane Science xxx (xxxx) xxx

illustrated by the water activity data shown in Fig. 7 as a function of


draw solute concentration (obtained from OLI Studio 9.6). Additionally,
due to mass transfer limitations, the concentration of the draw solute at
the active layer interface is substantially lower, up to 84%, than that of
the bulk, and outside of the range where a significant change in the
water permeability would be observed. For the range of bulk draw solute
concentrations examined, the draw solute concentration at the active
layer interface ranges from 0 to 1 mol/L, at which point there is a <1%
change in the water activity coefficient. It is therefore confidently
assumed in this work that the water permeability obtained is averaged
across all concentration values with little error.
Dw Kw xw;d vw
A¼ ; (26)
ta RT

γw
Kw ¼ ; (27)
γwðaÞ

The draw solute permeability B values in this work follow the


established trends based on the hydrated radius of the draw solute cation
[55,56] in Table 8. Further, the draw solute permeability can also be
Fig. 14. Comparing the predicted water flux as a function of the CaCl2 bulk
correlated with the hydration enthalpy of the cation [57–59], which
draw solute concentration using the model of this work and previous models of represents the energy barrier for the cation to become fully or partially
Chowdhury & McCutcheon [45], Nagy [63], and Tiraferri [38]. dehydrated [60]. It stands that a more strongly hydrated draw solution
cation (more negative hydration enthalpy) will transition from an
water permeability and structural parameter values were obtained, with aqueous phase to the membrane phase less favourably, inhibiting its
all R2 > 0.985, varying by �11.5% and �17.5% respectively between transport through the membrane according to the solution-diffusion
multiple draw solutes, which is within the �7.5 to �19% seen in the model [30]. Figs. 8 and 9 illustrate the linear correlation between the
literature [62], [86]. With minor variations between the water perme­ draw solute permeability and the hydrated cation radius and cation
ability and structural parameters obtained using different draw solutes, hydration enthalpy respectively.
it is therefore sufficient to characterize the membrane using only a single
draw solute at a minimum of four different concentrations. However, it 4.4. Modeling results and validation
is recommended that multiple draw solutes be examined so a more
comprehensive characterization is obtained. While the benefits of using a solely FO-based membrane character­
It is acknowledged that the water permeability A has been shown to ization approach over the traditional FO-RO method have been
be dependent on the draw solute concentration [54], however the re­ demonstrated by Tiraferri et al. [38], improvements are needed to
ported degree to which the water permeability is concentration depen­ provide robustness to this approach when applied to concentrated draw
dent is questionable. Based on the definition of water permeability, solutions that deviate from thermodynamic ideality. To demonstrate the
Equations (26) and (1) [30], this concentration dependence is minor as improvements of the model and characterization approach of our work,
the mole fractions of water, xw,d and xw,f, vary less than 10% up to the the experimental water and reverse draw solute flux were compared to
point of draw solute saturation. Further, the sorption coefficient Kw calculated values from four models: that of this work, of Chowdhury &
defined in Equation (27), will vary only by up to 35% for less ideal draw McCutcheon [45], of Tiraferri et al. [38], and of Nagy [63]. Our model
solutions (CaCl2 and MgCl2) at concentrations up to 4 mol/L. This is and that of Chowdhury & McCutcheon [45] do not have an explicit
solution, but the latter two models do. The water and reverse draw so­
lute flux models of Tiraferri et al. [38] are presented in Equations (28)
and (29),
8 � � �� � � 9
>
> π exp J S
π exp Jw >
>
< db w Ds fb kf =
JW ¼ A � � � � � ��� ; (28)
>
> >
>
:1 þ JB exp Jkw exp Jw DSs ;
w f

8 � � �� � � 9
>
> c exp J S
c exp Jw >
>
< s;db w Ds s;fb kf =
Js ¼ B � � � � � ��� : (29)
>
> >
>
:1 þ JB exp Jkw exp Jw DSs ;
w f

The model presented by Nagy [63] (shown in the form presented by


Bui et al. [31]), improves on that of Tiraferri et al. [38] by taking into

Table 9
Summary of iterative FO transport model comparison.
This Work Chowdhury & McCutcheon [45]

Draw Solution Species R2Jw R2Js R2Jw R2Js


Fig. 15. Comparing the predicted reverse draw solute flux as a function of the NaCl 0.99 0.99 0.96 0.99
MgCl2 0.97 0.99 0.87 0.97
CaCl2 bulk draw solute concentration using the model of this work and previous
CaCl2 0.95 0.98 0.88 0.93
models of Chowdhury & McCutcheon [45], Nagy [63], and Tiraferri [38].

8
J.T. Martin et al. Journal of Membrane Science xxx (xxxx) xxx

account the external concentration polarization on the draw solution our model improves upon that of Chowdhury & McCutcheon [45] by
side, considering localized estimates of the draw solute diffusivity corrected
8 � � �� � � 9 for the thermodynamic driving force, at the membrane active layer and
>
< πdb exp
> Jw k1d þ DSs πfb exp Jkwf >
>
= in the porous support-bulk solution boundary layer (calculated using the
JW ¼ A � � � � � ��� ; (30) average concentration of the draw solute at the porous support layer and
> >
>
:1 þ JB exp Jkw exp Jw
1
kd
þ S
Ds
>
; that of the bulk solution). Further, the mass transfer coefficients, kd and
kf, are calculated as part of the system of equations, based on the local
w f

8 � � �� � � 9 draw solute diffusivities.


>
>
< cs;db exp Jw k1d þ DSs cs;fb exp Jkwf >
>
= For the reverse draw solute flux, both models have high agreement
Js ¼ B � � � � � ��� : (31) with the experimental results, with R2 values of 0.99, 0.97, and 0.93 for
> >
>
:1 þ JB exp Jkw exp Jw k1d þ DSs >
; NaCl, MgCl2, and CaCl2 respectively for the work of Chowdhury &
McCutcheon [33], compared to R2 values of 0.99, 0.99, and 0.98 for
w f

The intrinsic membrane parameters, A, B, and S for the latter two NaCl, MgCl2, and CaCl2 respectively for this work (See Table 9). How­
models were determined using the method described in Tiraferri et al. ever, Chowdhury & McCutcheon [33] deviates by 20% from the
[38], with all the corresponding assumptions and the intrinsic param­ experimental results of CaCl2 at high concentrations (4 M), and our
eters for the Chowdhury & McCutcheon [45] model were determined model provides better agreement with experimental measurements for
using the method of this work. The intrinsic membrane parameters used the water flux of all draw solutions. This indicates that the model pre­
for the model presented in our work are listed in Table 7. The intrinsic sented in this work is better able to take into account the concentration
membrane parameters obtained were then used with their respective polarization and provide accurate estimates of both the water and
models to predict the water flux and reverse draw solute flux for NaCl, reverse draw solute flux for multiple draw solutions.
MgCl2, and CaCl2 draw solutions. To properly address the non-ideal FO
system, it was assumed that the diffusivity of the draw solute is a 5. Conclusion
function of the draw solute concentration and that all forms of con­
centration polarization were present for all simulated results. All sub­ This work details a rigorous and improved forward osmosis water
sequent R2 values were calculated using the experimental data not used and reverse draw solute flux model as well as a corresponding mem­
in the membrane parameterization (model validation data), as repre­ brane characterization method, by addressing the non-ideality of the
sented in Table 2. concentrated draw solutions. Based solely on experimental forward
Figs. 10 and 11 display the simulated water flux and reverse draw osmosis measurements, the proposed membrane characterization
solute flux results and compare to previous works for a NaCl draw solute. method produces consistent water permeability and structural param­
It appears that the water flux values of Tiraferri et al. [38] start to eter estimates (�11.5% and �17.5%) for three inorganic draw solutions,
deviate at around 2 mol/L due to the assumption of Van’t Hoff’s law NaCl, CaCl2 and MgCl2. Conjunctly, the draw solute permeabilities ob­
validity for all draw solute concentrations and the neglect of any tained follow established theoretical trends, with draw solute perme­
external concentration polarization on the draw solution side of the ability increasing with decreasing ionic radius and increasing hydration
membrane. Specifically for NaCl, the Van’t Hoff osmotic pressure starts enthalpy (less negative). This method improves upon the existing FO-
to deviate substantially from that based on water activity at around 2 based membrane characterization method, up to 107% in certain
mol/L NaCl, coinciding with the onset of deviation between the model of cases, by taking into account the non-ideal behaviour of draw solutions
Tiraferri et al., 2013 and our model and experimental measurements. regarding their osmotic pressure, the concentration dependence of draw
Similarly, for the divalent MgCl2 and CaCl2 draw solutions, the Van’t solute diffusivity, as well as all forms of concentration polarization.
Hoff osmotic pressure starts to deviate substantially from that based on Further, by considering localized estimates of the draw solute diffusivity
water activity at around 1.2 mol/L. In the case of the Nagy model [63], corrected for the thermodynamic driving force and calculating the
although it shows good agreement with the experimental water flux external mass transfer coefficients as part of the system of equations the
values (R2 ¼ 0.94), it diverges substantially from the measured reverse FO transport model of this work improves upon existing iterative
draw solute flux at NaCl concentrations greater than 2 mol/L. This is transport models up to 18.7%. The proposed FO model and character­
likely due to a poor estimate of the draw solute permeability obtained by ization method allow a more accurate evaluation of draw solutes,
using the parameterization method in Tiraferri et al. [38]. That membranes, and FO applications, producing R2 > 0.95 for the water and
parameterization method gave a water permeability and structural reverse draw solute flux for three inorganic draw solutions from 0 to 4
parameter that minimized the global error, but as a result the draw so­ mol/L. As novel applications for FO emerge, it is essential that a stan­
lute permeability produced a poor estimate of the reverse draw solute dard and rigorous FO model be implemented to accurately assess the
flux. In contrast, our model is in excellent agreement with the experi­ efficacy of FO for concentrated effluents and draw solutions.
mental water flux and the reverse draw solute flux data (R2 of 0.96 and
0.99 for Jw and Js respectively), demonstrating a better fit across a wider
range of NaCl concentrations. Declaration of competing interest
In Figs. 12–15, the models of Tiraferri et al. [38] and Nagy [63]
diverge even further from the experimental data for the divalent draw The authors declare that they have no known competing financial
solutes, MgCl2 and CaCl2. This is likely due to their assumptions that the interests or personal relationships that could have appeared to influence
draw solute diffusivity is constant, that Van’t Hoff’s Law validity holds the work reported in this paper.
at high draw solute concentrations, and that the external concentration
polarization is negligible. By removing these assumptions in our work, Acknowledgements
the water and reverse draw solute flux predictions were improved by
27–107% and 24–87% respectively over that of Tiraferri et al. [38] and The authors acknowledge the Natural Sciences and Engineering
Nagy [63]. The largest improvement was observed for the MgCl2 draw Research Council of Canada (NSERC RPGIN 105655-2013), and the
solute. The coefficients of determination (R2) for the water flux and University of Toronto Connaught Innovation Fund (#506081) for the
reverse draw solute flux, using this model, are 0.97, 0.99 for MgCl2 and financial support in this project. Professor Arun Ramchandran of the
0.95, and 0.98 for CaCl2 respectively. University of Toronto is also acknowledged for providing useful insights
Additionally, a comparison of the two iterative models shows that and corrections to the model. The OLI Systems Inc. is acknowledged for
providing access to the OLI software.

9
J.T. Martin et al. Journal of Membrane Science xxx (xxxx) xxx

References performance, J. Membr. Sci. 488 (2015) 40–55, https://doi.org/10.1016/j.


memsci.2015.03.059.
[25] G. Blandin, A.R.D. Verliefde, C.Y. Tang, P. Le-Clech, Opportunities to reach
[1] E.W. Tow, D.M. Warsinger, A.M. Trueworthy, J. Swaminathan, G.P. Thiel, S.
economic sustainability in forward osmosis-reverse osmosis hybrids for seawater
M. Zubair, A.S. Myerson, J.H. Lienhard V, Comparison of fouling propensity
desalination, Desalination (2015), https://doi.org/10.1016/j.desal.2014.12.011.
between reverse osmosis, forward osmosis, and membrane distillation, J. Membr.
[26] D.L. Shaffer, N.Y. Yip, J. Gilron, M. Elimelech, Seawater desalination for
Sci. (2018), https://doi.org/10.1016/j.memsci.2018.03.065.
agriculture by integrated forward and reverse osmosis: improved product water
[2] B. Mi, M. Elimelech, Organic fouling of forward osmosis membranes: fouling
quality for potentially less energy, J. Membr. Sci. (2012), https://doi.org/10.1016/
reversibility and cleaning without chemical reagents, J. Membr. Sci. 348 (2010)
j.memsci.2012.05.016.
337–345, https://doi.org/10.1016/j.memsci.2009.11.021.
[27] V. Yangali-Quintanilla, Z. Li, R. Valladares, Q. Li, G. Amy, Indirect desalination of
[3] S. Lee, C. Boo, M. Elimelech, S. Hong, Comparison of fouling behavior in forward
Red Sea water with forward osmosis and low pressure reverse osmosis for water
osmosis (FO) and reverse osmosis (RO), J. Membr. Sci. 365 (2010) 34–39, https://
reuse, Desalination 280 (2011) 160–166, https://doi.org/10.1016/j.
doi.org/10.1016/j.memsci.2010.08.036.
desal.2011.06.066.
[4] E. Arkhangelsky, F. Wicaksana, C. Tang, A.A. Al-Rabiah, S.M. Al-Zahrani, R. Wang,
[28] R. Valladares Linares, Z. Li, S. Sarp, S.S. Bucs, G. Amy, J.S. Vrouwenvelder,
Combined organic-inorganic fouling of forward osmosis hollow fiber membranes,
Forward osmosis niches in seawater desalination and wastewater reuse, Water Res.
Water Res. (2012), https://doi.org/10.1016/j.watres.2012.09.003.
66 (2014) 122–139, https://doi.org/10.1016/j.watres.2014.08.021.
[5] C.Y. Tang, Q. She, W.C.L. Lay, R. Wang, A.G. Fane, Coupled effects of internal
[29] D. Attarde, M. Jain, P.K. Singh, S.K. Gupta, Energy-efficient seawater desalination
concentration polarization and fouling on flux behavior of forward osmosis
and wastewater treatment using osmotically driven membrane processes,
membranes during humic acid filtration, J. Membr. Sci. (2010), https://doi.org/
Desalination 413 (2017) 86–100, https://doi.org/10.1016/j.desal.2017.03.010.
10.1016/j.memsci.2010.02.059.
[30] J.G. Wijmans, R.W. Baker, The solution-diffusion model : a review 107 (1995)
[6] E.R. Cornelissen, D. Harmsen, K.F. de Korte, C.J. Ruiken, J.J. Qin, H. Oo, L.
1–21, https://doi.org/10.1016/0376-7388(95)00102-I.
P. Wessels, Membrane fouling and process performance of forward osmosis
[31] N.N. Bui, J.T. Arena, J.R. McCutcheon, Proper accounting of mass transfer
membranes on activated sludge, J. Membr. Sci. (2008), https://doi.org/10.1016/j.
resistances in forward osmosis: improving the accuracy of model predictions of
memsci.2008.03.048.
structural parameter, J. Membr. Sci. 492 (2015) 289–302, https://doi.org/
[7] Y. Kim, S. Lee, H.K. Shon, S. Hong, Organic fouling mechanisms in forward osmosis
10.1016/j.memsci.2015.02.001.
membrane process under elevated feed and draw solution temperatures,
[32] G.D. Mehta, S. Loeb, Internal polarization in the porous substructure of a
Desalination 355 (2015) 169–177, https://doi.org/10.1016/j.desal.2014.10.041.
semipermeable membrane under pressure-retarded osmosis, J. Membr. Sci. 4
[8] B. Mi, M. Elimelech, Chemical and physical aspects of organic fouling of forward
(1978) 261–265, https://doi.org/10.1016/S0376-7388(00)83301-3.
osmosis membranes, J. Membr. Sci. 320 (2008) 292–302, https://doi.org/
[33] S. Loeb, F. Van Hessen, D. Shahaf, Production of energy from concentrated brines
10.1016/j.memsci.2008.04.036.
by pressure-retarded osmosis: I. Preliminary technical and economic correlations,
[9] G. Kolliopoulos, J.T. Martin, V.G. Papangelakis, Energy requirements in the
J. Membr. Sci. 1 (1976) 49–63, https://doi.org/10.1016/S0376-7388(00)82271-1.
separation-regeneration step in forward osmosis using TMA–CO2–H2O as the draw
[34] K.L. Lee, R.W. Baker, H.K. Lonsdale, Membranes for power generation by pressure-
solution, Chem. Eng. Res. Des. 140 (2018) 166–174, https://doi.org/10.1016/j.
retarded osmosis, J. Membr. Sci. (1981), https://doi.org/10.1016/S0376-7388(00)
cherd.2018.10.015.
82088-8.
[10] J.R. McCutcheon, R.L. McGinnis, M. Elimelech, Desalination by ammonia-carbon
[35] J.R. McCutcheon, M. Elimelech, Influence of concentrative and dilutive internal
dioxide forward osmosis: influence of draw and feed solution concentrations on
concentration polarization on flux behavior in forward osmosis, J. Membr. Sci. 284
process performance, J. Membr. Sci. 278 (2006) 114–123, https://doi.org/
(2006) 237–247, https://doi.org/10.1016/j.memsci.2006.07.049.
10.1016/j.memsci.2005.10.048.
[36] N.Y. Yip, A. Tiraferri, W.A. Phillip, J.D. Schiffman, L.A. Hoover, Y.C. Kim,
[11] P. Dou, S. Zhao, J. Song, H. He, Q. She, X.M. Li, Y. Zhang, T. He, Forward osmosis
M. Elimelech, Thin-film composite pressure retarded osmosis membranes for
concentration of a vanadium leaching solution, J. Membr. Sci. 582 (2019)
sustainable power generation from salinity gradients, Environ. Sci. Technol.
164–171, https://doi.org/10.1016/j.memsci.2019.04.012.
(2011), https://doi.org/10.1021/es104325z.
[12] B.K. Pramanik, L. Shu, J. Jegatheesan, K. Shah, N. Haque, M.A. Bhuiyan, Rejection
[37] A. Achilli, T.Y. Cath, A.E. Childress, Power generation with pressure retarded
of rare earth elements from a simulated acid mine drainage using forward osmosis:
osmosis: an experimental and theoretical investigation, J. Membr. Sci. (2009),
the role of membrane orientation, solution pH, and temperature variation, Process
https://doi.org/10.1016/j.memsci.2009.07.006.
Saf. Environ. Prot. 126 (2019) 53–59, https://doi.org/10.1016/j.
[38] A. Tiraferri, N.Y. Yip, A.P. Straub, S. Romero-Vargas Castrillon, M. Elimelech,
psep.2019.04.004.
A method for the simultaneous determination of transport and structural
[13] G. Kolliopoulos, V.G. Papangelakis, Extraction 2018, Springer International
parameters of forward osmosis membranes, J. Membr. Sci. 444 (2013) 523–538,
Publishing, 2018, https://doi.org/10.1007/978-3-319-95022-8.
https://doi.org/10.1016/j.memsci.2013.05.023.
[14] G. Gwak, D.I. Kim, S. Hong, New industrial application of forward osmosis (FO):
[39] K.H. Mistry, J.H. Lienhard, Effect of nonideal solution behavior on desalination of a
precious metal recovery from printed circuit board (PCB) plant wastewater,
sodium chloride solution and comparison to seawater, J. Energy Resour. Technol.
J. Membr. Sci. 552 (2018) 234–242, https://doi.org/10.1016/j.
135 (2013) 042003, https://doi.org/10.1115/1.4024544.
memsci.2018.02.022.
[40] G.N. Lewis, The osmotic pressure of concentrated solutions, and the laws of the
[15] J. Li, M. Wang, Y. Zhao, H. Yang, Y. Zhong, Enrichment of lithium from salt lake
perfect solution, J. Am. Chem. Soc. (1908), https://doi.org/10.1021/ja01947a002.
brine by forward osmosis, R. Soc. Open Sci. 5 (2018) 1–8, https://doi.org/
[41] M. Park, J.J. Lee, S. Lee, J.H. Kim, Determination of a constant membrane structure
10.1098/rsos.180965.
parameter in forward osmosis processes, J. Membr. Sci. 375 (2011) 241–248,
[16] H. Zhou, Optimisation of crystallisation parameters for lithium carbonate
https://doi.org/10.1016/j.memsci.2011.03.052.
microcrystals based on forward osmosis (FO) process, Mater. Res. Innov. 21 (2015)
[42] C.H. Tan, H.Y. Ng, Modified models to predict flux behavior in forward osmosis in
1–9, https://doi.org/10.1179/1433075x15y.0000000029.
consideration of external and internal concentration polarizations, J. Membr. Sci.
[17] R.L. McGinnis, N.T. Hancock, M.S. Nowosielski-Slepowron, G.D. McGurgan, Pilot
324 (2008) 209–219, https://doi.org/10.1016/j.memsci.2008.07.020.
demonstration of the NH3/CO2 forward osmosis desalination process on high
[43] C.H. Tan, H.Y. Ng, Revised external and internal concentration polarization models
salinity brines, Desalination 312 (2013) 67–74, https://doi.org/10.1016/j.
to improve flux prediction in forward osmosis process, Desalination 309 (2013)
desal.2012.11.032.
125–140, https://doi.org/10.1016/j.desal.2012.09.022.
[18] M.S. Islam, S. Sultana, J.R. McCutcheon, M.S. Rahaman, Treatment of fracking
[44] M.R. Chowdhury, J. Ren, K. Reimund, J.R. McCutcheon, A hybrid dead-end/cross-
wastewaters via forward osmosis: evaluation of suitable organic draw solutions,
flow forward osmosis system for evaluating osmotic flux performance at high
Desalination 452 (2019) 149–158, https://doi.org/10.1016/j.desal.2018.11.010.
recovery of produced water, Desalination (2017), https://doi.org/10.1016/j.
[19] B.D. Coday, P. Xu, E.G. Beaudry, J. Herron, K. Lampi, N.T. Hancock, T.Y. Cath, The
desal.2016.08.021.
sweet spot of forward osmosis: treatment of produced water, drilling wastewater,
[45] M.R. Chowdhury, J.R. McCutcheon, Elucidating the impact of temperature
and other complex and difficult liquid streams, Desalination 333 (2014) 23–35,
gradients across membranes during forward osmosis: coupling heat and mass
https://doi.org/10.1016/j.desal.2013.11.014.
transfer models for better prediction of real osmotic systems, J. Membr. Sci. 553
[20] B.D. Coday, T.Y. Cath, Forward osmosis: novel desalination of produced water and
(2018) 189–199, https://doi.org/10.1016/j.memsci.2018.01.004.
fracturing flowback, J. Am. Water Work. Assoc. 106 (2014) 37–38, https://doi.
[46] T.V. Bartholomew, M.S. Mauter, Computational framework for modeling
org/10.5942/jawwa.2014.106.0016.
membrane processes without process and solution property simplifications,
[21] D.L. Shaffer, L.H. Arias Chavez, M. Ben-Sasson, S. Romero-Vargas Castrill� on, N.
J. Membr. Sci. (2019), https://doi.org/10.1016/j.memsci.2018.11.067.
Y. Yip, M. Elimelech, Desalination and reuse of high-salinity shale gas produced
[47] S.S. Manickam, J.R. McCutcheon, Understanding mass transfer through
water: drivers, technologies, and future directions, Environ. Sci. Technol. 47
asymmetric membranes during forward osmosis: a historical perspective and
(2013) 9569–9583, https://doi.org/10.1021/es401966e.
critical review on measuring structural parameter with semi-empirical models and
[22] G. Chen, Z. Wang, L.D. Nghiem, X.M. Li, M. Xie, B. Zhao, M. Zhang, J. Song, T. He,
characterization approaches, Desalination 421 (2017) 110–126, https://doi.org/
Treatment of shale gas drilling flowback fluids (SGDFs) by forward osmosis:
10.1016/j.desal.2016.12.016.
membrane fouling and mitigation, Desalination 366 (2015) 113–120, https://doi.
[48] B. Kim, G. Gwak, S. Hong, Review on Methodology for Determining Forward
org/10.1016/j.desal.2015.02.025.
Osmosis (FO) Membrane Characteristics: Water Permeability (A), Solute
[23] J. Minier-Matar, A. Hussain, A. Janson, R. Wang, A.G. Fane, S. Adham, Application
Permeability (B), and Structural Parameter (S), Desalination (2017), https://doi.
of forward osmosis for reducing volume of produced/Process water from oil and
org/10.1016/j.desal.2017.08.006.
gas operations, Desalination 376 (2015) 1–8, https://doi.org/10.1016/j.
[49] B.D. Coday, D.M. Heil, P. Xu, T.Y. Cath, Effects of transmembrane hydraulic
desal.2015.08.008.
pressure on performance of forward osmosis membranes, Environ. Sci. Technol. 47
[24] B.D. Coday, N. Almaraz, T.Y. Cath, Forward osmosis desalination of oil and gas
(2013) 2386–2393, https://doi.org/10.1021/es304519p.
wastewater: impacts of membrane selection and operating conditions on process

10
J.T. Martin et al. Journal of Membrane Science xxx (xxxx) xxx

[50] R.W. Baker, Membrane Technology and Applications, 2004, https://doi.org/ and perspective, J. Phys. Chem. C 0 (2019), https://doi.org/10.1021/acs.
10.1002/0470020393. jpcc.9b02178 null-null.
[51] H.K. Shon, S. Phuntsho, T.C. Zhang, R.Y. Surampalli, Forward Osmosis: [57] W. Cheng, C. Liu, T. Tong, R. Epsztein, M. Sun, R. Verduzco, J. Ma, M. Elimelech,
Fundamentals and Applications, 2015, https://doi.org/10.1061/9780784414071. Selective removal of divalent cations by polyelectrolyte multilayer nanofiltration
[52] J.A. Rard, D.G. Miller, The mutual diffusion coefficients of NaCl-H2O and CaCl2- membrane: role of polyelectrolyte charge, ion size, and ionic strength, J. Membr.
H2O at 25� C from Rayleigh interferometry, J. Solut. Chem. (1979), https://doi. Sci. 559 (2018) 98–106, https://doi.org/10.1016/j.memsci.2018.04.052.
org/10.1007/BF00648776. [58] L.A. Richards, A.I. Sch€
afer, B.S. Richards, B. Corry, The importance of dehydration
[53] S. Phuntsho, S. Hong, M. Elimelech, H.K. Shon, Osmotic equilibrium in the forward in determining ion transport in narrow pores, Small 8 (2012) 1701–1709, https://
osmosis process: modelling, experiments and implications for process performance, doi.org/10.1002/smll.201102056.
J. Membr. Sci. 453 (2014) 240–252, https://doi.org/10.1016/j. [59] H.H.P. Fang, E.S.K. Chian, Criterion of ion separation by reverse osmosis, J. Appl.
memsci.2013.11.009. Polym. Sci. 19 (1975) 2889–2895, https://doi.org/10.1002/app.1975.070191024.
[54] A. Sagiv, A. Zhu, P.D. Christofides, Y. Cohen, R. Semiat, Analysis of forward [60] G.A. Moldoveanu, V.G. Papangelakis, Recovery of rare earth elements adsorbed on
osmosis desalination via two-dimensional FEM model, J. Membr. Sci. (2014), clay minerals: I. Desorption mechanism, Hydrometallurgy (2012), https://doi.org/
https://doi.org/10.1016/j.memsci.2014.04.001. 10.1016/j.hydromet.2012.02.007.
[55] G.M. Geise, D.R. Paul, B.D. Freeman, Fundamental water and salt transport [61] E.R. Nightingale, Phenomenological theory of ion solvation. Effective radii of
properties of polymeric materials, Prog. Polym. Sci. 39 (2014) 1–42, https://doi. hydrated ions, J. Phys. Chem. (1959), https://doi.org/10.1021/j150579a011.
org/10.1016/j.progpolymsci.2013.07.001. [62] Y. Marcus, Ions in Solution and Their Solvation, 2015, https://doi.org/10.1002/
[56] S.J. Faucher, N.R. Aluru, M.Z. Bazant, D. Blankschtein, A.H. Brozena, J. Cumings, 9781118892336.
J. Pedro de Souza, M. Elimelech, R. Epsztein, J.T. Fourkas, A. Govind Rajan, H. [63] E. Nagy, A general, resistance-in-series, salt- and water flux models for forward
J. Kulik, A. Levy, A. Majumdar, C. Martin, M. McEldrew, R. Prasanna Misra, osmosis and pressure-retarded osmosis for energy generation, J. Membr. Sci.
A. Noy, T. Anh Pham, M.A. Reed, E. Schwegler, Z.S. Siwy, Y. Wang, M.S. Strano, (2014), https://doi.org/10.1016/j.memsci.2014.02.021.
Critical knowledge gaps in mass transport through single-digit nanopores: a review

11

You might also like