You are on page 1of 11

Journal of Membrane Science 520 (2016) 790–800

Contents lists available at ScienceDirect

Journal of Membrane Science


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

Modeling the water permeability and water/salt selectivity tradeoff in


polymer membranes
Huan Zhang, Geoffrey M. Geise n
Department of Chemical Engineering, University of Virginia, 102 Engineers' Way, P.O. Box 400741, Charlottesville, VA 22904, USA

art ic l e i nf o a b s t r a c t

Article history: A theoretical framework is proposed to model the observed water/salt permeability, diffusivity, and
Received 18 May 2016 sorption selectivity tradeoff relationships with water permeability, diffusion, and sorption coefficients.
Received in revised form Salt sorption is modeled using an electrostatic (or dielectric) exclusion approach. Water and salt diffusion
19 August 2016
are modeled using hindered diffusion and free volume theories. The modeling approach suggested here
Accepted 20 August 2016
Available online 24 August 2016
is compared to a set of water and salt transport property data measured using a wide variety of polymers.
The models describe the general trends observed in the tradeoff relationships but also situations where
Keywords: penetrant-polymer interactions and/or other system specific effects are likely significant. The approach
Upper bound also highlights the importance of convective frame of reference and thermodynamic correction factors in
Desalination
the analysis of water diffusion in hydrated polymers. These corrections are discussed within the fra-
Diffusion
mework of experimental data from a variety of hydrated polymer systems. The expanded tradeoff re-
Sorption
Transport lationship data set and theoretical framework help to expand insight into the fundamental phenomena
that govern water and salt transport in polymer membranes for desalination applications.
& 2016 Elsevier B.V. All rights reserved.

1. Introduction to seawater desalination by RO.


Some significant challenges still face membrane desalination
Growing demand for purified water resources and clean energy technologies. The polyamide-based materials that are commonly
coupled with interest in the nexus of food, energy, and water re- used as the active layers of reverse osmosis and nanofiltration
sources has driven significant interest in membrane-based tech- membranes are inherently susceptible to degradation by chlorine-
nologies that could help address global need for water and energy containing compounds that are widely used to disinfect water and
[1–10]. Membrane-based technologies, such as reverse osmosis control membrane biofouling [1,23–27]. Additionally, zero-liquid
(RO), currently dominate the desalination market, and increases in discharge requirements are subjecting membranes to concentra-
energy efficiency over several decades have positioned mem- tions of salt that exceed traditional design parameters [19,28,29].
branes to play an increasingly important role in addressing global Increasingly contaminated waters that contain a variety of ions
water needs [9,11,12]. Furthermore, emerging energy technolo- and other pollutants also challenge membrane materials that have
gies, such as pressure-retarded osmosis (PRO), are of interest for traditionally been designed to reject sodium chloride [1,24,30–36].
recovering energy, stored as chemical potential, in solutions of Finally, interest in designing membrane polymers that are both
different salt concentration [13–18]. Accordingly, there is sig- highly selective and simultaneously resistant to fouling presents a
significant engineering challenge to the membrane field [29]. In
nificant interest in optimizing, preferably inexpensive, polymer
many cases, the literature provides insufficient guidance on the
membranes for these applications.
impact of polymer membrane chemistry and structure on the
Seawater desalination by reverse osmosis is a relatively energy
fundamental transport properties that are critical for desalination
efficient process [9,19–21]. The energy required to desalinate
system performance.
seawater is roughly within a factor of two of the thermodynamic
One approach to benchmarking material properties and the
minimum [9]. This efficiency is largely a result of improvements in
influence of polymer chemical and/or structural modification on
membranes over several decades and pressure recovery devices
transport properties is to consider the impact of those modifica-
that can be incorporated into the process [9,11,22]. On the other
tions on established water/salt permeability, sorption, and/or dif-
hand, brackish water RO is typically less energy efficient compared fusivity selectivity tradeoff relationships [37]. For a wide range of
polymers, a tradeoff relationship is observed whereby materials
n
Corresponding author. with higher water/salt permeability selectivity (i.e., the ratio of
E-mail address: geise@virginia.edu (G.M. Geise). water permeability to salt permeability) tend to have lower water

http://dx.doi.org/10.1016/j.memsci.2016.08.035
0376-7388/& 2016 Elsevier B.V. All rights reserved.
H. Zhang, G.M. Geise / Journal of Membrane Science 520 (2016) 790–800 791

the upstream face of the membrane precedes diffusion down a


concentration gradient and is followed by desorption at the
downstream face of the membrane [39,40]. The solution-diffusion
model describes the overall permeability of a polymer to a given
penetrant in terms of sorption and diffusive contributions to
transport:
Pi = Ki × Di (1)

where Pi is permeability, Ki is the sorption coefficient, and Di is the


diffusion coefficient of component i in the polymer [39,40]. Eq. (1)
follows from analysis of the governing flux equations [1,3,24,40].
While the separation of permeability into sorption and diffusion
components using Eq. (1) appears relatively straightforward, de-
scribing the fundamental underpinnings of the sorption and dif-
fusive properties and their connections to polymer chemistry and/
or structure is often much more challenging [1,24,37].
The intrinsic permeability selectivity of a material can be de-
fined as a ratio of permeability values. For desalination, selectivity
is often defined as the ratio of the water permeability to the salt
permeability, and as shown in Eq. (1), permeability can be split
into sorption and diffusion components [37]. The water/salt per-
Fig. 1. Water/salt permeability selectivity plotted versus diffusive water perme- meability selectivity and corresponding sorption and diffusivity
ability for a series of reported polymer membrane transport property data, char-
selectivity expressions are related as:
acterized using sodium chloride [23,37,46–58].
Pw K D
= w × w
permeability and vice versa (Fig. 1). The water/salt permeability Ps Ks Ds (2)
selectivity tradeoff with water permeability can be split into
Permeability selectivity can be determined by independently
sorption and diffusivity tradeoff relationships because perme-
measuring water and salt permeability values (i.e., the so-called
ability is defined, in these materials, as the product of a sorption
ideal selectivity), or the selectivity can be determined from si-
coefficient and an average diffusion coefficient [1,24,37–40]. These
multaneous water and salt transport experiments. The latter can
empirical tradeoff relationships allow for comparison of a variety
be particularly important when osmotic de-swelling or other ion-
of materials based on fundamental transport properties (i.e., per-
polymer interactions influence the water content and, thus, the
meability and selectivity as opposed to flux and rejection [1]). As
water permeability of the polymer [61]. Presently, the tradeoff
has been recognized in other areas of the membrane field [41–45],
relationships do not distinguish between these experimental
they can be used to illustrate the current state-of-the-art and/or
methods.
emphasize needed property improvements [10].
One limitation of the fundamental tradeoff relationships for
2.1. The sorption component of transport
desalination membranes is that the intrinsic water and salt
transport properties, e.g., permeability, sorption, and diffusion
The sorption coefficient can be defined, at equilibrium, as the
coefficients, must be known. Accurate measurement of these
ratio of the concentration of a component i in the polymer
properties is difficult for heterogeneous thin film composite
membrane, cim , to the concentration of component i in the external
membranes such as the polyamide materials that are used in
solution, cis [39,40]:
commercial desalination processes [37,59], and this situation, in
many cases, frustrates efforts to establish fundamental structure- cim
property relationships in such materials. Tradeoff relationships,
Ki ≡
cis (3)
however, can also be developed based on water and salt per-
meance data that are available for such materials [10,37,60]. For desalination applications, sorption coefficients are typically
While the empirical water permeability and water/salt se- defined for water, Kw, and either individual ions (e.g., K þ or K  ) or
lectivity tradeoff relationship, introduced in 2011 [37], provides a overall salt, Ks [1,24,37]. The water sorption coefficient is often
starting point for benchmarking polymers that are of interest for approximately equal to the volume fraction of water sorbed in the
desalination applications, a theoretical framework would be polymer, and it is convenient, and often appropriate, to use Kw and
helpful in further understanding the fundamental origins of the the volume fraction of water interchangeably [1,62]. For salt
tradeoff relationship. Here we incorporate additional data into the sorption in hydrated polymers, the concentrations in Eq. (3) are
tradeoff relationship (relative to the initial report). Additionally, typically defined as mol of i per kg or L of water sorbed in the
we incorporate theory to identify the contributions of electrostatic polymer [63].
interactions, hindered diffusion, and free volume transport in The concentrations used to define the salt sorption coefficient,
polymers to the tradeoff relationships. Finally, we seek to gain Ks, depend on the nature of the electrolyte and whether the
insight into the fundamental phenomena that result in the tradeoff polymer is uncharged (e.g., poly(ethylene oxide)) or charged (e.g.,
relationship. cation or anion exchange materials) [1,3,24,63–65]. The salt sorp-
tion coefficient for binary monovalent electrolyte sorption in an
uncharged polymer can be calculated using either the cation or
2. Theoretical background anion concentrations [1,3,65]. In charged polymers, however, the
co-ion concentrations (i.e., anions in cation exchange materials
Transport of water and salt through polymeric materials that and cations in anion exchange materials) are used to define the
are of interest for desalination applications is typically described salt sorption coefficient, Ks [1,3,65]. Here we will focus on the
using the solution-diffusion model where penetrant sorption at overall salt sorption and water sorption coefficients, Ks and Kw,
792 H. Zhang, G.M. Geise / Journal of Membrane Science 520 (2016) 790–800

respectively.
The sorption coefficient describes the thermodynamic parti-
tioning of a compound from the external solution into the mem-
brane phase. Therefore, the sorption coefficient can be related to
the free energy change associated with moving the component
from one state (e.g., solution) to another (e.g., the membrane
phase):

cim ⎡ ΔGi, sorption ⎤


Ki ≡ ⎢
s = exp −

ci ⎣ RT ⎦ (4)

where ΔGi, sorption is the free energy change associated with moving
component i from the external solution into the membrane phase,
R is the gas constant, and T is the absolute temperature [66]. The
relationship between the sorption coefficient and the right-hand
side of Eq. (4) is predicated on the assumption of solution ideality
as the activity coefficients have been set equal to unity. Ad-
ditionally, Eq. (4) is analogous to expressions for the Donnan po-
tential that is often discussed within the framework of ion sorp-
tion in charged polymers [63]. Conversion of Eq. (4) into the
equation typically used to define the Donnan potential is accom-
plished by setting ΔGi, sorption = ziFEDonnan where zi is the charge Fig. 2. Static dielectric constants, measured at 298 K, for solutions of water and
either dimethyl sulfoxide (DMSO, ) [76] or 1-methylimidazole (MeIM, ) [75] as
number of ion i, F is Faraday's constant, and EDonnan is the Donnan
a function of the mole fraction of water in the solution. The static dielectric con-
potential (note that the activity coefficients are still assumed to be stants of pure DMSO, MeIM, and water were reported as 47.29, 39.78, and 78.89,
taken as unity) [63,67]. In desalination applications, Ki is typically respectively [75,76].
less than unity [1] suggesting that ΔGi, sorption is positive and re-
presents the free energy barrier to sorption.
The ΔGi, sorption term in Eq. (4) should include contributions from is not necessarily clear. The relative permittivity of solutions of
all of the phenomena that influence the sorption process. For the organic molecules and water can exhibit linear correlations be-
purpose of describing ion sorption, we have chosen to consider a tween the static dielectric constant (relative permittivity) and
simple case where ΔGi, sorption is taken as the solvation energy
water composition (Fig. 2) [75,76]. While the DMSO/water system
barrier, ΔWi (i.e., the change in free energy associated with the data exhibit a linear transition from the static dielectric constant of
change in the solvation environment of the ion between the so-
DMSO to that of water, the data for the 1-methylimidazole
lution and membrane phases). This approach is often used to de-
(MeIm)/water system suggest that a highly non-linear region ex-
scribe dielectric exclusion of ions from hydrated polymers [68–70].
ists between 0.9 and 1.0 mol fraction of water. Other binary solu-
This choice neglects other phenomena that may contribute (in
tions of organic solvents exhibit non-linear static dielectric con-
some cases perhaps significantly) to the sorption free energy, but
stant correlations with composition, suggesting that different in-
as will be discussed later, this simple approach appears to be ef-
fective for describing the ion sorption in a collection of different termolecular interactions can influence the relationship between
polymers. composition and the static dielectric constant of the solution
The solvation energy barrier, ΔWi, can be calculated using the [77,78].
Born model: Relative permittivity data as a function of polymer water con-
tent are not common in the literature, but data were reported for
zi2e2 ⎛ 1 1 ⎞ Nafion™ 117, a perfluorosulfonic acid polymer (Fig. 3) [79]. As it is
ΔWi = ⎜ − ⎟
8πε0ai ⎝ εm εsol ⎠ (5) very difficult to dry Nafion™ completely, the relative permittivity
of Teflon™ was reported as the zero water content material be-
where e is the elementary charge, ε0 is the permittivity of free
cause Nafion™ and Teflon™ share a common backbone [79]. The
space, εm is the relative permittivity of the membrane, εsol is the
Nafion™ 117 relative permittivity data (Fig. 3) correlate linearly
dielectric constant of the external solution (taken equal to 80 for
aqueous electrolytes [71]), and ai is the bare ion radius [72]. with water content in a manner that is bounded by the dry and
The salt sorption coefficient can be written in terms of ΔWi as: pure water limits similar to the DMSO/water system (Fig. 2).
Based on the linear data observed for the DMSO/water and
csm ⎡ ΔWs ⎤ ⎡ zi2e2 ⎛ 1 1 ⎞⎤ Nafion™ 117/water systems, we suggest that a linear relationship
Ks ≡ s = exp⎢− ⎥ = exp⎢− ⎜ − ⎟⎥
cs ⎣ kT ⎦ ⎣ 8πkTε0as ⎝ εm εsol ⎠⎦ (6) between εm and the volume fraction of water sorbed in the
polymer (effectively equal to Kw) may be reasonable in the absence
where k is Boltzmann's constant. In order to apply Eq. (6), values of more general data or specific guidance from models. The re-
are needed for as and εm. To describe overall salt sorption in
lative permittivity of many dry hydrocarbon polymers fall in the
polymers, as was taken as 0.194 nm, which is the average of the
range of 2-6 (the value for Teflon™ is 2 [79]), so using the linear
cavity radii for sodium (0.151 nm) and chloride (0.236 nm) ions
regression from Fig. 3 may be a reasonable assumption to arrive at
[73]. The cavity radius is defined as the distance from the center of
a general empirical relationship between εm and Kw:
the ion to the point where the dielectric permittivity changes from
the vacuum value to that of the solvent (water) in continuum εm = 1.04 + ( 78.3 × K w ) (7)
models [73,74].
Determining values for εm can be challenging given a lack of Using Eq. (7) to link the relative permittivity of hydrated
permittivity data in the literature for hydrated polymers. One polymers to water sorption, it is possible to construct a model for
would expect the value of εm to increase as the water content of the variation of the salt sorption coefficient with the water sorp-
the polymer increases [1], but the functional form of this increase tion coefficient:
H. Zhang, G.M. Geise / Journal of Membrane Science 520 (2016) 790–800 793

⎛ ⎞¯
Dw K w ⎜ ∂ ln K w ⎟ Vw Δp − Δπ
⎟ RT ( )
nw =
L( 1 − K w ) ⎜⎝ ∂ ln a wm m
aw ⎠
0

Pw V¯
= δ w ( Δp − Δπ ) = A( Δp − Δπ )
L( 1 − K w ) RT (10)

where a wm
is the activity of water in the membrane, a wm0is the
activity of water in the membrane at the upstream (i.e., high
pressure) face, V̄w is the partial molar volume of water, R is the gas
constant, T is absolute temperature, Δp is the applied pressure
difference, Δπ is the osmotic pressure difference across the
membrane, Pw is the diffusive water permeability, and A is the
water permeance [1,37]. The thermodynamic correction factor, δ,
relates the activity of water in the membrane to the water sorption
coefficient and is evaluated at the upstream face of the membrane
[1,80,81]. Eq. (10) can be derived from a Maxwell-Stefan multi-
component transport analysis, and much like the salt diffusivity,
the water diffusivity, Dw, in Eq. (10) is averaged over the thickness
of the membrane during integration of the differential water flux
equation [1].
Eq. (10) has been simplified for use with low water content
Fig. 3. Static dielectric constant data measured at 303 K for Teflon™ and Nafion™
117 as a function of the volume fraction of water sorbed in the polymer [79]. The
polymers, and much of the early desalination literature uses this
static dielectric constant of pure water is also included [71]. The dashed line was simplification [1,39,40]:
determined by linear regression of the data.
Pw V¯w
nw =
L RT
( Δp − Δπ ) = A( Δp − Δπ ) (11)

⎡ The version of the water flux equation shown in Eq. (11) in-
csm zi2e2 ⎛ 1 1 ⎞⎤⎥
Ks ≡ = exp⎢− ⎜ − ⎟ cludes two key assumptions that distinguish it from Eq. (10). First,
css ⎢⎣ 8πkTε0as ⎝ 1.04 + (78.3 × K w ) εsol ⎠⎥⎦ (8) the water content of the polymer (taken as Kw) is assumed to be
small, such that Kw { 1, and this assumption removes the con-
Since all of the parameters other than Kw in Eq. (8) are fixed,
vective frame of reference term, (1 Kw), in Eq. (10). Second, the
the model can be used to predict the water/sorption selectivity
thermodynamic factor, δ, is assumed to be equal to unity. While
(Kw/Ks) as a function of Kw, which can then be compared to ex- these assumptions may be reasonable for some low water content
perimental data on the sorption selectivity tradeoff plot. polymers, they generally fail for polymers where Kw is not very
small compared to unity [1].
Including the convective frame of reference term, (1 Kw) in Eq.
2.2. Constitutive equations for diffusion
(10), is relatively straightforward when relating a measured water
flux to the average water diffusivity because the water content of
In order to determine water and salt diffusivities for hydrated
the polymer is usually known or easily obtained. The thermo-
polymer membranes, constitutive equations are needed to relate dynamic correction factor, however, requires additional informa-
the experimentally measured fluxes (typically salt flux and water tion. Either water sorption data are needed as a function of water
flux) to the external driving forces for transport. These constitutive activity (i.e., a sorption isotherm) or a suitable equation of state
equations have been developed for water and salt transport in (e.g., Flory-Huggins or Flory-Rehner theories [82,83]) is needed to
hydrated membranes [24,40]. evaluate δ [1]. Both the convective frame of reference and ther-
Salt flux, ns, is related to an external salt concentration driving modynamic correction factors are often significant for highly
force by: swollen polymers, but the thermodynamic correction factor is
more significant than the convective frame of reference term in
Δcss Δc s
ns = KsDs = Ps s = BΔcss some polymers [1].
(9)
L L While rigorous treatment of δ often requires water sorption
where L is the membrane thickness, Δcss
is the salt concentration data measured as a function of water activity, which is often not
difference between the external salt solutions on either side of the available for many materials reported in the literature, one can
membrane, Ps is the salt permeability, and B is the salt permeance approximate the significance of the thermodynamic correction
[1,24,40]. Eq. (9) can be derived from ion flux expressions using factor in Eq. (10) by performing a one point fit of the Flory-Huggins
the Nernst-Planck equation and is equivalent to the expression model using pure water sorption data that is often reported to-
gether with water flux measurements [84]. In this case, the Flory-
obtained from Fick's Law [24]. It is important to note that the salt
Huggins interaction parameter is calculated from the one point fit,
diffusivity, Ds, in Eq. (9) is an average over the thickness of the
and then, this interaction parameter is used in the model to cal-
membrane, and this average is defined during the integration of
culate δ [1,84]. While this approach is limited in rigor, it can be
the differential salt flux equation that results during analysis used to highlight the influence of the thermodynamic correction
[1,24]. Additionally, the dependence of Ds on ion diffusivities (i.e., factor on the interpretation of the average water diffusivity.
D þ and D  ) and ion concentrations, which is important for Reported water diffusivity data for a variety of hydrated poly-
charged polymers, can be described using the Nernst-Planck mers are shown in Fig. 4 [37,48–50,52,54,57,85–95]. The influence
analysis [1,24]. of the convective frame of reference term and the thermodynamic
Water flux, nw, is related to an applied pressure difference, correction factor (in most cases estimated using the one point
using a thermodynamic approach, as: Flory-Huggins fit described above) is highlighted. The dashed line
794 H. Zhang, G.M. Geise / Journal of Membrane Science 520 (2016) 790–800

Turnbull approach to arrive at a relationship between the diffusion


coefficient of water and the water content of the polymer:

Dw = Dw° exp⎢−
⎢ β ( 1
Kw)(1 − α) ⎤⎥⎥
−1

1 + α( −1) ⎥⎦
1
⎢⎣ Kw (13)

α = VF , p/VF , w (14)

β = V */VF , w (15)

where Dw°is the self-diffusion coefficient of water (taken as


5
2.8  10 cm2/s [96]) and V* is a characteristic volume parameter
that is related to the cross-section and the diffusional jump length
of the diffusing penetrant [57,97,98].

2.3.2. Salt diffusivity


Salt diffusivity, like water diffusivity, is expected to generally
Fig. 4. Reported water transport data for a variety of polymers [37,48– increase as the water content of the polymer increases
50,52,54,57,85–95]. Values of Dw calculated using Eq. (11) ( ) are compared to [1,24,46,58,99–101]. Two relatively simple theories explain this
those values calculated using Eq. (10) with δ ¼1 ( ) and those values calculated phenomenological observation. The theory of Mackie and Meares
using Eq. (10) including thermodynamic non-ideality estimated, where necessary,
accounts for the effect of area and tortuosity on diffusive transport
by a one point Flory-Huggins fit (◆). The thermodynamic correction factor appears
to be more significant (i.e., compare ◆ and ) than the convective frame of re- (i.e., the polymer occupies space and hinders transport) using a
ference term (i.e., compare and ), particularly at high values of Kw. The dashed random walk approach [99,102]. Alternatively, the theory of Ya-
line represents the self-diffusion coefficient of water, Dw°
¼ 2.8  10  5 cm2/s [96]. suda is based on a free volume approach where the free volume is
taken to be proportional to the water content of the polymer (si-
milar to that described above for the water diffusion coefficient)
in Fig. 4 represents the self-diffusion coefficient of water
( Dw° ¼ 2.8  10  5 cm2/s [96]). Values of Dw calculated using Eq. (11) [58]. In both cases, the diffusion coefficient of salt in the polymer is
from data measured on highly swollen polymers often exceed the expected to approach the diffusion coefficient of salt in bulk
self-diffusion coefficient of water. This situation, in some cases, has aqueous solution as the volume fraction of water approaches unity
been ascribed to hydraulic (e.g., viscous) flow in the membrane. (i.e., the polymer acts to slow the diffusion process relative to
Properly accounting for the convective frame of reference term diffusion in solution). Here we consider sodium chloride transport,
and approximating the thermodynamic correction term (i.e., using so the diffusion coefficient of NaCl in bulk aqueous solution is
Eq. (10)), however, results in values of Dw that fall satisfactorily taken as 1.5  10  5 cm2/s at 25 °C [103].
below the self-diffusion coefficient of water suggesting that, as The Mackie and Meares model assumes that the swollen
expected, the polymer membrane slows water diffusion relative to polymer is generally homogeneous [99]. Additionally, the model
the rate of water diffusion in bulk water. assumes that the swollen polymer contains an internal solution
that is intermixed with a polymer matrix and that the character-
2.3. Diffusivity selectivity istic length that separates polymer junctions (e.g., cross-links or
entanglements) is large compared to the size of the diffusing ions.
To model the water/salt diffusivity selectivity as a function of In this theory, the presence of the polymer acts to reduce the
water diffusivity, expressions are needed to relate Dw and Ds to cross-sectional area available for transport and introduces tortu-
material properties. Diffusion in polymers is often modeled as osity [99]. Both factors slow diffusion relative to bulk solution
either an activated or hindered transport process, and several re- properties, and increases in polymer water content reduce the
lationships have been reported with varying degrees of generality effect of this transport resistance. The model depends only on the
and/or assumptions. Here we seek to avoid specific structural or volume fraction of water sorbed in the polymer and the bulk so-
morphological assumptions in an effort to capture the transport lution diffusion coefficient:
properties of a variety of different polymers.
⎛ K ⎞2
w
Ds = Ds°⎜ ⎟
2.3.1. Water diffusivity ⎝ 2 − Kw ⎠ (16)
Based on the experimental data shown in Fig. 4, one would
expect the water diffusivity to increase with increasing water where Ds° ¼ 1.5  10  5 2
cm /s [103]. The model ensures that the salt
sorption. This relationship can be captured using free volume diffusion coefficient in the membrane approaches the bulk solu-
theory to describe diffusive transport in hydrated polymers. An tion salt diffusion coefficient as Kw approaches unity.
approach used by Yasuda et al., assumes that the free volume The Yasuda et al. theory is based on free volume analysis of
available for transport in the swollen polymer, VF, can be de- diffusion and the assumption that the polymer free volume is
termined using volume additivity (recall that Kw is effectively proportional to water content, taken in this case as Kw [58]:
equivalent to the volume fraction of water sorbed in the polymer): ⎡ b ⎤
Ds = a × exp⎢ − ⎥
VF = K wVF , w + ( 1 − K w )VF , p (12) ⎣ Kw ⎦ (17)

where VF,w is the free volume of pure water and VF,p is the free where a and b are adjustable constants. The constant a can be
volume of the dry polymer [57]. This approach can be combined constrained such that the model ensures that Ds will approach Ds°
with an activated state model for diffusion using the Cohen and as Kw approaches unity, which is a physically desirable result:
H. Zhang, G.M. Geise / Journal of Membrane Science 520 (2016) 790–800 795

a = Ds° × exp⎡⎣ b⎤⎦ (18)

such that the model reduces to a single adjustable parameter


model:
⎡ ⎛ 1 ⎞⎤
Ds = Ds° × exp⎢ b⎜ 1 − ⎟⎥
⎣ ⎝ K w ⎠⎦ (19)

The constant b is related to the size of the diffusing penetrant


[58]. The constant b (and a if considering Eq. (17)) can be regressed
from salt diffusion coefficient data as a function of water content.

3. Results and discussion

Water and sodium chloride transport data [23,37,46–58] were


compiled and used to evaluate the effectiveness of the models
described in the previous section to describe the water and salt
transport property tradeoff relationships. The data set contains a
variety of different polymers including materials based on aro-
matic, acrylic, ethoxy, cellulosic, and hydrocarbon backbones.
Some of the materials contain fixed charge groups (e.g., cation or Fig. 5. Water diffusion coefficient data (determined using Eq. (10) and the ther-
anion exchange materials) while other materials are hydrophilic modynamic correction factor approximation described above) versus water sorp-
uncharged polymers. Also, while most studies were conducted at tion coefficient [23,37,46–58]. The solid line corresponds to the model described in
Eq. (13) with α ¼0.5 and β ¼ 4.5 [57].
ambient temperature, the salt solution concentrations used to
measure salt permeability and the use of salt solution or pure
water to characterize water permeability vary to some extent
across the different studies reported in the literature. The goal of of the dry polymer. Alternatively, PALS measurements made on
this analysis is to identify the general models that capture trends dry sulfonated polysulfone materials result in fractional free vo-
for a diverse set of materials and highlight opportunities and need lumes ranging from 0.030 to 0.034 [84]. Considering that ortho-
for developing specific structure property relationships for differ- positronium inhibition processes in sulfonated polymers are be-
ent classes and categories of polymers. The model predictions for lieved to result in under-predictions of fractional free volume
Dw and Ds as a function of Kw are discussed below within the [107,108], the sulfonated polysulfone results may suggest that the
context of the literature data set, and the models are combined dry polymer fractional free volume of some materials is close to
and compared to the tradeoff relationships. the fractional free volume of water, which would suggest a value of
α closer to unity. For the purpose of this work, the Yasuda et al.
3.1. Water diffusion parameters were used as reported [57].

Yasuda et al. determined values for the parameters α ¼0.5 and 3.2. Salt diffusion
β ¼ 4.5 from water transport data measured using a series of ac-
rylic, methacrylic, and cellulosic polymers [57]. The model de- The models in Eqs. (16), (17), and (19) are compared to ex-
scribed in Eqs. (13)–(15) qualitatively describes the increase of Dw perimental salt diffusion and water content data obtained from
with Kw (Fig. 5) for the literature data set (where Dw was analyzed the literature (Fig. 6). As both models predict Ds-0 as Kw-0, it is
using Eq. (10) and the thermodynamic correction procedure de- expected that the models will break down at low water content,
scribed above). The scatter in Fig. 5 should be expected given that and the increased scatter in the data at low water content further
data from a wide range of polymers are considered and that the suggest that other phenomena, such as specific polymer-penetrant
thermodynamic correction factor was approximated for many of interactions may become more significant at low water content.
the data points using a one-point Flory-Huggins model fit. The data are reasonably well bounded by the Mackie and Meares
It is instructive to consider the implications of the parameter model (Eq. (16)), which tends to over predict Ds at moderate water
values reported by Yasuda, et al. [57]. It is reasonable that β 4 1 content, and the single parameter free volume model (Eq. (19)),
given that a characteristic volume greater than that of the free which tends to under predict Ds at moderate water content.
volume of water might be needed for water transport. The value of The fit of the 2-parameter free volume-based model (Eq. (17))
α ¼0.5, however, suggests that generally the free volume of water appears to capture the moderate water content Ds data better than
is twice that of the dry polymer. Free volume data collected using the other two models. The model, however, fails to capture the
positron annihilation lifetime spectroscopy (PALS) data suggest physically significant result where the salt diffusion coefficient is
that this situation might be true for some polymers but not for expected to approach its bulk solution value as Kw-1. Therefore,
others. the Mackie and Meares model (Eq. (16)), which contains no ad-
PALS data (measured ortho-positronium lifetime and intensity) justable parameters, and the single parameter free volume model
can be used to determine the fractional free volume in a material (Eq. (19)) are preferable to Eq. (17) based on agreement with the
using the Tao-Eldrup approach [104,105]. The fractional free vo- high water content limit. The apparent improved agreement be-
lume of bulk water is reported to be 0.041 [106]. PALS data mea- tween the data and the free volume model compared to the
sured using dry cellulose ester films and analyzed using the same Mackie and Meares model (at least at moderate to low water
approach result in fractional free volumes that range from 0.013 to content) is consistent with a recent report suggesting that free
0.020 [88]. These results align relatively well with the α ¼0.5 volume transport theory may be more applicable compared to
parameter discussed above as they suggest that the free volume of hindered transport theory for describing diffusion in some com-
bulk water may be approximately a factor of two greater than that mercially available desalination membranes [109].
796 H. Zhang, G.M. Geise / Journal of Membrane Science 520 (2016) 790–800

Fig. 6. Salt diffusion coefficient data plotted versus water sorption coefficient for Fig. 7. Water/salt sorption selectivity versus water sorption coefficient for a series
several polymers [23,37,46–58]. Model results are shown for the Mackie and of reported data [37,46,50–56,110,111]. The single circle point represents an aqu-
Meares model (Eq. (16)), the 2-parameter Yasuda model (Eq. (17), a¼ 2.0( 70.5)  eous sodium chloride solution, and the line was calculated using Eq. (8) to de-
10  5 cm2/s and b¼ 1.17 70.13), and the single parameter Yasuda model (Eq. (19), termine Ks (and, thus, Kw/Ks) as a function of Kw.
b ¼2.39 7 0.15). The parameters and their uncertainties for the Yasuda model were
regressed using OriginPro software. membrane could be considered to be a homogeneous dielectric,
and this assumption may not be appropriate for all materials
3.3. Modeling the water/salt tradeoff relationships [69,70]. Additionally, all of the data presented here were measured
using sodium chloride as the electrolyte. While sorption data are
The modeling framework described in the previous sections reported for other electrolytes, additional systematic sorption
was applied to the water/salt sorption, diffusivity, and perme- studies with other electrolytes would likely highlight the balance
ability selectivity tradeoff relationships with water sorption, dif- of the importance of different exclusion mechanisms and may give
fusion, and permeability coefficients. The application of these rise to additional insight into the interplay of water content and
models to the tradeoff plots introduces a theoretical framework for sorption selectivity.
understanding the phenomena that result in the tradeoff re- The diffusion selectivity data reported in 2011 [37] were ana-
lationships for sorption, diffusion, and permeability. The approach lyzed using Eq. (11), and this method of analysis resulted in some
employed here is a refinement of the empirical relationships used data where the calculated value of Dw exceeded the self-diffusion
previously [37]. coefficient of water. In Fig. 8, those data (shown as diamonds)
The Born model-based approach to modeling the water/salt were adjusted for the frame of reference and thermodynamic
sorption selectivity versus water sorption coefficient tradeoff re- correction factor terms discussed in relation to Eq. (10), and the
lationship describes the high selectivity-high Kw frontier of the resulting points are shown as squares. Satisfactorily, the values of
experimental data remarkably well given the relative simplicity of Dw are now found to be less than the self-diffusion coefficient of
the model and the nature of the dielectric constant treatment that water (indicated by the single circle symbol). Perhaps due to the
was used in the model (Fig. 7). It is reasonable to expect that the nature of the approximation used to determine the thermo-
Born model might predict water/salt selectivity values that are dynamic correction factor, some diffusion selectivity values fall
high compared to the majority of the data as the Born model has below unity, which generally is not expected for transport of hy-
been suggested to over-predict the energy barrier for sorption drated ions in hydrated polymers as the hydrated ions are ex-
[72]. Such an over-prediction would result in lower salt sorption pected to be larger and, thus, diffuse slower than water.
coefficients and ultimately higher water/salt sorption selectivity. The use of Eq. (10) to correct for the convective frame of re-
While a few experimental data points meet or exceed the model ference and thermodynamic terms in the salt flux equation,
predictions, the majority of the polymers considered exhibit wa- however, allows for comparison of experimental data and the
ter/salt sorption selectivity values that are below the result pre- diffusion models described above. The free volume-based model
dicted based strictly on electrostatics. for water transport (Eqs. (13)–(15)) was used to calculate Dw as a
The potential influence of polymer chemistry on ion sorption function of Kw, and either the Mackie and Meares model (Eq. (16))
selectivity is highlighted by the wide range of water/salt sorption or the free volume model (Eq. (19)) were used to calculate Ds as a
selectivity values that can be experimentally accessed at a given function of Kw. The model for Dw was combined with each model
polymer water content. The data set in Fig. 7 contains a mix of for Ds to produce the curves shown in Fig. 8.
uncharged hydrophilic polymers and charged (i.e., ion exchange) Using the free volume model proposed by Yasuda et al. for Ds
materials where additional ion exclusion mechanisms, such as describes the frontier of the diffusivity tradeoff relationship while
Donnan exclusion and/or specific ion-polymer interactions, may using the model proposed by Mackie and Meares for Ds brackets
be quite significant. While considerable work has been done to much of the data on the low-selectivity side. This result is con-
understand fundamental ion exclusion mechanisms such as di- sistent with the observation that the Mackie and Meares model
electric and Donnan exclusion [53,64,65,68,70,101,112–130], many predicted higher values of Ds, at a given water content, compared
questions still remain about the interplay of these mechanisms in to the free volume model as shown in Fig. 6. The observation that
a wide range of polymers and polymer water content. The mod- the data tend to lie between the two models suggests that diffu-
eling approach used here also assumed that the hydrated polymer sive transport in many polymers may fall somewhere between the
H. Zhang, G.M. Geise / Journal of Membrane Science 520 (2016) 790–800 797

Fig. 9. Water/salt permeability selectivity versus diffusive water permeability for a


Fig. 8. Water/salt diffusion selectivity versus water diffusion coefficient for a series
series of reported data [23,37,46–58]. The short dashed line represents the em-
of reported data [23,37,46–58]. The short dashed line represents the empirical
pirical tradeoff relationship based on diffusion coefficient (and, thus diffusive
tradeoff relationship based on diffusion coefficient data calculated using Eq. (11)
permeability) data calculated using Eq. (11) and also shown in Fig. 1 ( ) [37].
( ) [37]. Data calculated using Eq. (10) and the thermodynamic correction factor
Diffusion coefficient (and, thus diffusive permeability) data calculated using Eq.
procedure described above ( ) are compared to the model results where water
(10) and the thermodynamic correction factor procedure described above ( ) are
diffusion was modeled using the Yasuda approach (Eqs. (13)–(15)) and salt diffusion
compared to the model results: where water diffusion was modeled using the
was modeled using either the Mackie and Meares approach (long dashed line, Eq.
Yasuda approach (Eqs. (13)–(15)) and salt diffusion was modeled using either the
(16)) or the single parameter free volume model (solid line, Eq. (19)). The single
Mackie and Meares approach (long dashed line, Eq. (16)) or the single parameter
circle point represents the ratio of the self-diffusion coefficient of water and the salt
free volume model (solid line, Eq. (19)). The single circle point represents an
diffusion coefficient in aqueous solution plotted at the value of the self-diffusion
aqueous salt solution.
coefficient of water.

with the discussion (earlier in this section) of the applicability of


free volume description of transport and the obstruction (area/ the models to the salt diffusion coefficients and the diffusivity
tortuosity) description of transport used by Mackie and Meares. tradeoff relationship.
The result is also consistent with a recent report suggesting that
diffusive transport in highly selective commercially available de-
salination membranes may be more accurately modeled using free 4. Conclusions
volume theory compared to hindered transport theory [109].
The diffusion and sorption data and the modeling analysis The reported modeling results can be used to better under-
shown in Figs. 6 and 7 can be combined to describe the water/salt stand the fundamental underpinnings of the water/salt perme-
permeability selectivity tradeoff with the diffusive water perme- ability, sorption, and diffusion selectivity tradeoff relationship
ability (Fig. 9). To combine the data, the permeability and se- with water permeability, sorption, and diffusivity. The effective-
lectivity values were combined according to Eqs. (1) and (2) [39]. ness of the tradeoff relationship modeling approaches were
As the method used to determine the diffusion coefficient ulti- gauged using a transport property data set based on reported data
mately influences the permeability, the analysis described above for a broad range of polymers, and the previously reported tradeoff
for the diffusion coefficients carries over into the analysis of the relationship plots were updated with additional data from the
water/salt permeability and diffusive water permeability tradeoff recent literature. The analysis presented here also refines the in-
relationship. itial empirical tradeoff relationship frontiers.
As discussed in regard to Fig. 8, correcting the water diffusion The water/salt sorption tradeoff relationship can be described
coefficient data for the convective frame of reference and ther- by electrostatic (or dielectric exclusion) forces using the Born
modynamic factors (i.e., using Eq. (10) as opposed to Eq. (11)) model, though specific interactions between penetrants and the
impacts the diffusive water permeability. The data shown in the hydrated polymer are clearly important for some polymers. The
water/salt permeability selectivity tradeoff relationship with water water/salt diffusivity tradeoff relationship can be described using a
permeability (Figs. 1 and 9 as diamonds) were calculated using Eq. model for water and salt diffusivity that is based on free volume
(11). When Eq. (10) is used during the analysis, the data shift theory, but salt diffusivity can also be modeled using the para-
generally to lower selectivity and diffusive water permeability meter-free Mackie and Meares model, which accounts for the
values upon applying the frame of reference and thermodynamic space filling and tortuosity contributions to salt diffusion. Suc-
correction factors. When combined, the sorption selectivity and cessful water diffusion modeling relies on properly accounting for
diffusion selectivity models (both based on the Mackie and Meares the convective frame of reference and thermodynamic correction
model and the free volume model for Ds) describe the perme- factors in the water flux equation (particularly for high water
ability tradeoff relationship frontier relatively well. This result content polymers). These approaches can be combined to describe
suggests that the contribution of the sorption selectivity tradeoff the water/salt permeability tradeoff relationship with diffusive
helps to offset some of the differences between the two ap- water permeability, and the influence of the sorption tradeoff
proaches used to model salt diffusion coefficients. The observation, seems to mitigate differences introduced in the selection of the
in Fig. 9, that the free volume model predicts higher selectivity salt diffusion coefficient model.
values compared to the Mackie and Meares model is consistent The tradeoff relationships and modeling analysis reported here
798 H. Zhang, G.M. Geise / Journal of Membrane Science 520 (2016) 790–800

only considers sodium chloride transport, and the set of experi- [28] S.G.J. Heijman, H. Guo, S. Li, J.C. van Dijk, L.P. Wessels, Zero liquid discharge:
mental data contains properties measured using a broad range of heading for 99% recovery in nanofiltration and reverse osmosis, Desalination
236 (2009) 357–362.
polymers. Specific polymer-penetrant interactions are expected to [29] A.G. Fane, R. Wang, M.X. Hu, Synthetic membranes for water purification:
be significant in some charged polymers and in polymers that sorb status and future, Angew. Chem. Int. Ed. 54 (2015) 3368–3386.
[30] R. Bernstein, S. Belfer, V. Freger, Toward improved boron removal in RO by
relatively little water. While the analysis reported here is generally membrane modification: feasibility and challenges, Environ. Sci. Technol. 45
applicable for the wide range of materials considered, this ap- (2011) 3613–3620.
proach is not expected to capture specific details of particular [31] A. Ben-David, Y. Oren, V. Freger, Thermodynamic factors in partitioning and
rejection of organic compounds by polyamide composite membranes, En-
polymer categories or material chemistry. viron. Sci. Technol. 40 (2006) 7023–7028.
[32] L.J. Banasiak, A.I. Schäfer, Removal of boron, fluoride and nitrate by elec-
trodialysis in the presence of organic matter, J. Membr. Sci. 334 (2009)
101–109.
Appendix A. Supporting information [33] N. Hilal, G.J. Kim, C. Somerfield, Boron removal from saline water: a com-
prehensive review, Desalination 273 (2011) 23–35.
[34] X. Jin, Q.H. She, X.L. Ang, C.Y.Y. Tang, Removal of boron and arsenic by for-
Supplementary data associated with this article can be found in ward osmosis membrane: influence of membrane orientation and organic
the online version at doi:10.1016/j.memsci.2016.08.035. fouling, J. Membr. Sci. 389 (2012) 182–187.
[35] P. Dydo, M. Turek, Boron transport and removal using ion-exchange mem-
branes: a critical review, Desalination 310 (2013) 2–8.
[36] B. Teychene, G. Collet, H. Gallard, J.P. Croue, A comparative study of boron
References and arsenic (III) rejection from brackish water by reverse osmosis mem-
branes, Desalination 310 (2013) 109–114.
[37] G.M. Geise, H.B. Park, A.C. Sagle, B.D. Freeman, J.E. McGrath, Water perme-
[1] G.M. Geise, D.R. Paul, B.D. Freeman, Fundamental water and salt transport ability and water/salt selectivity tradeoff in polymers for desalination, J.
properties of polymeric materials, Prog. Polym. Sci. 39 (2014) 1–42. Membr. Sci. 369 (2011) 130–138.
[2] M.A. Shannon, P.W. Bohn, M. Elimelech, J.G. Georgiadis, B.J. Marinas, A. [38] J.G. Wijmans, R.W. Baker, The solution-diffusion model: a review, J. Membr.
M. Mayes, Science and technology for water purification in the coming Sci. 107 (1995) 1–21.
decades, Nature 452 (2008) 301–310. [39] R.W. Baker, Membrane Technology and Applications, 3rd ed., Wiley, New
[3] J. Kamcev, B.D. Freeman, Charged polymer membranes for environmental/ York, 2012.
energy applications, Annu. Rev. Chem. Biomol. Eng. 7 (2016), null. [40] D.R. Paul, Reformulation of the solution-diffusion theory of reverse osmosis,
[4] M.E. Webber, Catch-22: Water vs. Energy, Sci. Am. Sp. Ed., 18, 2008, pp. 34– J. Membr. Sci. 241 (2004) 371–386.
41. [41] L.M. Robeson, B.D. Freeman, D.R. Paul, B.W. Rowe, An empirical correlation of
[5] The Water-Energy Nexus: Challenges and Opportunities, U.S. Department of gas permeability and permselectivity in polymers and its theoretical basis, J.
Energy, June 2014. Membr. Sci. 341 (2009) 178–185.
[6] Energy demands on water resources: Report to Congress on the inter- [42] L.M. Robeson, Correlation of separation factor versus permeability for poly-
dependency of energy and water, U.S. Department of Energy, 2006 (accessed meric membranes, J. Membr. Sci. 62 (1991) 165–185.
12.09.13). [43] D.R. Paul, L.M. Robeson, Polymer nanotechnology: nanocomposites, Polymer
[7] Water for Energy: is energy becoming a thirstier resource? in: World Energy 49 (2008) 3187–3204.
Outlook 2012, International Energy Agency, 2012. [44] G.M. Geise, M.A. Hickner, B.E. Logan, Ionic resistance and permselectivity
[8] B. Cai, B. Zhang, J. Bi, W. Zhang, Energy's thirst for water in China, Environ. tradeoffs in anion exchange membranes, ACS Appl. Mater. Interfaces 5 (2013)
Sci. Technol. 48 (2014) 11760–11768. 10294–10301.
[9] M. Elimelech, W.A. Phillip, The future of seawater desalination: energy, [45] A. Mehta, A.L. Zydney, Permeability and selectivity analysis for ultrafiltration
technology, and the environment, Science 333 (2011) 712–717. membranes, J. Membr. Sci. 249 (2005) 245–249.
[10] J.R. Werber, A. Deshmukh, M. Elimelech, The critical need for increased se- [46] H. Strathmann, A.S. Michaels, Polymer-water interaction and its relation to
lectivity, not increased water permeability, for desalination membranes, reverse osmosis desalination efficiency, Desalination 21 (1977) 195–202.
Environ. Sci. Technol. Lett. 3 (2016) 112–120. [47] B.J. Sundell, E.-S. Jang, J.R. Cook, B.D. Freeman, J.S. Riffle, J.E. McGrath, Cross-
[11] C. Fritzmann, J. Löwenberg, T. Wintgens, T. Melin, State-of-the-art of reverse linked disulfonated poly(arylene ether sulfone) telechelic oligomers. 2. Ele-
osmosis desalination, Desalination 216 (2007) 1–76. vated transport performance with increasing hydrophilicity, Ind. Eng. Chem.
[12] K.P. Lee, T.C. Arnot, D. Mattia, A review of reverse osmosis membrane ma- Res. 55 (2016) 1419–1426.
terials for desalination – development to date and future potential, J. Membr. [48] B.J. Sundell, K.-s Lee, A. Nebipasagil, A. Shaver, J.R. Cook, E.-S. Jang, B.
Sci. 370 (2011) 1–22. D. Freeman, J.E. McGrath, Cross-linking disulfonated poly(arylene ether
[13] B.E. Logan, M. Elimelech, Membrane-based processes for sustainable power sulfone) telechelic oligomers. 1. Synthesis, characterization, and membrane
generation using water, Nature 488 (2012) 313–319. preparation, Ind. Eng. Chem. Res. 53 (2014) 2583–2593.
[14] T. Thorsen, T. Holt, The potential for power production from salinity gra- [49] W. Xie, G.M. Geise, B.D. Freeman, C.H. Lee, J.E. McGrath, Influence of pro-
dients by pressure retarded osmosis, J. Membr. Sci. 335 (2009) 103–110. cessing history on water and salt transport properties of disulfonated poly-
[15] K.K. Reimund, J.R. McCutcheon, A.D. Wilson, Thermodynamic analysis of sulfone random copolymers, Polymer 53 (2012) 1581–1592.
energy density in pressure retarded osmosis: the impact of solution volumes [50] W. Xie, J. Cook, H.B. Park, B.D. Freeman, C.H. Lee, J.E. McGrath, Fundamental
and costs, J. Membr. Sci. 487 (2015) 240–248. salt and water transport properties in directly copolymerized disulfonated
[16] F. Helfer, C. Lemckert, Y.G. Anissimov, Osmotic power with pressure retarded poly(arylene ether sulfone) random copolymers, Polymer 52 (2011)
osmosis: theory, performance and trends – a review, J. Membr. Sci. 453 2032–2043.
(2014) 337–358. [51] H.B. Park, W. Xie, B.D. Freeman, J.E. McGrath, Unpublished Data, The Uni-
[17] C. Klaysom, T.Y. Cath, T. Depuydt, I.F.J. Vankelecom, Forward and pressure versity of Texas at Austin, Austin, TX, 2007.
retarded osmosis: potential solutions for global challenges in energy and [52] H. Ju, A.C. Sagle, B.D. Freeman, J.I. Mardel, A.J. Hill, Characterization of sodium
water supply, Chem. Soc. Rev. 42 (2013) 6959–6989. chloride and water transport in poly(ethylene oxide) hydrogels, J. Membr.
[18] Y.C. Kim, M. Elimelech, Potential of osmotic power generation by pressure Sci. 358 (2010) 131–141.
retarded osmosis using seawater as feed solution: analysis and experiments, [53] G.M. Geise, C.L. Willis, C.M. Doherty, A.J. Hill, T.J. Bastow, J. Ford, K.I. Winey, B.
J. Membr. Sci. 429 (2013) 330–337. D. Freeman, D.R. Paul, Characterization of aluminum-neutralized sulfonated
[19] S. Miller, H. Shemer, R. Semiat, Energy and environmental issues in desali- styrenic pentablock copolymer films, Ind. Eng. Chem. Res. 52 (2013)
nation, Desalination 366 (2015) 2–8. 1056–1068.
[20] R. Semiat, Energy issues in desalination processes, Environ. Sci. Technol. 42 [54] H.K. Lonsdale, U. Merten, R.L. Riley, Transport properties of cellulose acetate
(2008) 8193–8201. osmotic membranes, J. Appl. Polym. Sci. 9 (1965) 1341–1362.
[21] K.S. Spiegler, Y.M. El-Sayed, The energetics of desalination processes, Desa- [55] M.A. Frommer, J.S. Murday, R.M. Messalem, Solubility and diffusivity of
lination 134 (2001) 109–128. water and of salts in an aromatic polyamide film, Eur. Polym. J. 9 (1973)
[22] R.L. Stover, Seawater reverse osmosis with isobaric energy recovery devices, 367–373.
Desalination 203 (2007) 168–175. [56] J.R. Cook, Fundamental Water and Ion Transport Characterization of Sulfo-
[23] H.B. Park, B.D. Freeman, Z.-B. Zhang, M. Sankir, J.E. McGrath, Highly chlorine- nated Polysulfone Desalination Materials, Department of Chemical En-
tolerant polymers for desalination, Angew. Chem. 120 (2008) 6108–6113. gineering, University of Texas, Austin, TX, 2014.
[24] G.M. Geise, H.-S. Lee, D.J. Miller, B.D. Freeman, J.E. McGrath, D.R. Paul, Water [57] H. Yasuda, C.E. Lamaze, A. Peterlin, Diffusive and hydraulic permeabilities of
purification by membranes: the role of polymer science, J. Polym. Sci. Part B: water in water-swollen polymer membranes, J. Polym. Sci.: Part A-2. 9 (1971)
Polym. Phys. 48 (2010) 1685–1718. 1117–1131.
[25] T. Knoell, Municipal wastewater: chlorine's impact on the performance and [58] H. Yasuda, C.E. Lamaze, L.D. Ikenberry, Permeability of solutes through hy-
properties of polyamide membranes, Ultrapure Water 23 (2006) 24–31. drated polymer membranes Part I. Diffusion of sodium chloride, Makromol.
[26] A.C. Anderson, R.S. Reimers, P. deKernion, A brief review of the current status Chem. 118 (1968) 19–35.
of alternatives to chlorine disinfection of water, Am. J. Public Health 72 [59] V. Freger, Nanoscale heterogeneity of polyamide membranes formed by in-
(1982) 1290–1293. terfacial polymerization, Langmuir 19 (2003) 4791–4797.
[27] A.E. Allegrezza Jr., B.S. Parekh, P.L. Parise, E.J. Swiniarski, J.L. White, Chlorine [60] N.Y. Yip, M. Elimelech, Performance limiting effects in power generation
resistant polysulfone reverse osmosis modules, Desalination 64 (1987) from salinity gradients by pressure retarded osmosis, Environ. Sci. Technol.
285–304. 45 (2011) 10273–10282.
H. Zhang, G.M. Geise / Journal of Membrane Science 520 (2016) 790–800 799

[61] G.M. Geise, Water and salt separation characteristics of a sulfonated styrenic [92] C.E. Evans, R.D. Noble, S. Nazeri-Thompson, B. Nazeri, C.A. Koval, Role of
pentablock copolymer for membrane applications, in: Water and Salt conditioning on water uptake and hydraulic permeability of Nafion mem-
Transport Structure/property Relationships in Polymer Membranes for De- branes, J. Membr. Sci. 279 (2006) 521–528.
salination and Power Generation Applications, Department of Chemical [93] P. Meares, The mechanism of water transport in membranes, Philos. Trans. R.
Engineering, University of Texas, Austin, TX, 2012. Soc. B 278 (1977) 113–150.
[62] G.M. Geise, B.D. Freeman, D.R. Paul, Sodium chloride diffusion in sulfonated [94] C.E. Reid, E.J. Breton, Water and ion flow across cellulosic membranes, J. Appl.
polymers for membrane applications, J. Membr. Sci. 427 (2013) 186–196. Polym. Sci. 1 (1959) 133–143.
[63] F. Helfferich, Ion Exchange, Dover Publications, New York, 1995. [95] S.G. Kimura, Reverse osmosis performance of sulfonated poly(2,6-di-
[64] J. Kamcev, M. Galizia, F.M. Benedetti, E.-S. Jang, D.R. Paul, B.D. Freeman, G. methylphenylene ether) ion exchange membranes, Ind. Eng. Chem. Res. 10
S. Manning, Partitioning of mobile ions between ion exchange polymers and (1971) 335–339.
aqueous salt solutions: importance of counter-ion condensation, Phys. Chem. [96] J.H. Wang, C.V. Robinson, I.S. Edelman, Self-diffusion and structure of liquid
Chem. Phys. 18 (2016) 6021–6031. water. III. Measurement of the self-diffusion of liquid water with H2, H3 and
[65] G.M. Geise, L.P. Falcon, B.D. Freeman, D.R. Paul, Sodium chloride sorption in O18 as tracers, J. Am. Chem. Soc. 75 (1953) 466–470.
sulfonated polymers for membrane applications, J. Membr. Sci. 423–424 [97] M.H. Cohen, D. Turnbull, Molecular transport in liquids and glasses, J. Chem.
(2012) 195–208. Phys. 31 (1959) 1164–1169.
[66] J.M. Prausnitz, R.N. Lichtenthaler, E. Gomes de Azevedo, Molecular Ther- [98] A.T. DiBenedetto, D.R. Paul, An interpretation of gaseous diffusion through
modynamics Of Fluid-Phase Equilibria, 3rd ed., Prentice Hall PTR, Upper polymers using fluctuation theory, J. Polym. Sci. Part A: General Papers. 2
Saddle River, NJ, 1999. (1964) 1001–1015.
[67] F.G. Donnan, The theory of membrane equilibria, Chem. Rev. 1 (1924) [99] J.S. Mackie, P. Meares, The diffusion of electrolytes in a cation-exchange resin
73–90. membrane. I. Theoretical, Proceedings of the Royal Society of London. Series
[68] E. Glueckauf, On the mechanism of osmotic desalting with porous mem- A, Mathematical and Physical Sciences, 232 (1955) 498–509.
branes, in: Proceedings of the First International Symposium on Water De- [100] G.M. Geise, B.D. Freeman, D.R. Paul, Characterization of a novel sulfonated
salination, Oct. 3–9, 1965, U.S. Department of the Interior, Washington, D.C., pentablock copolymer for desalination applications, Polymer 51 (2010)
1967, pp. 143–150. 5815–5822.
[69] J.E. Anderson, W. Pusch, The membrane/water partition coefficients of ions: [101] E. Dražević, K. Košutić, V. Freger, Permeability and selectivity of reverse os-
electrostatic calculations of dielectric heterogeneity, Ber. Bunsen Ges. 80 mosis membranes: correlation to swelling revisited, Water Res. 49 (2014)
(1976) 846–849. 444–452.
[70] A. Yaroshchuk, Dielectric exclusion of ions from membranes, Adv. Colloid [102] J.S. Mackie, P. Meares, The diffusion of electrolytes in a cation-exchange resin
Interface Sci. 85 (2000) 193–230. membrane. II. Experimental, Proceedings of the Royal Society of London.
[71] W.M. Haynes, CRC Handbook of Chemistry and Physics, 95th ed., CRC Press/ Series A, Mathematical and Physical Sciences, 232 (1955) 510–518.
Taylor and Francis, Boca Raton, FL, 2014, Internet Version 2014. [103] C.J.D. Fell, H.P. Hutchison, Diffusion coefficients for sodium and potassium
[72] W.R. Bowen, J.S. Welfoot, Modelling the performance of membrane nano- chlorides in water at elevated temperatures, J. Chem. Eng. Data 16 (1971)
filtration–critical assessment and model development, Chem. Eng. Sci. 57 427–429.
(2002) 1121–1137. [104] S.J. Tao, Positronium annihilation in molecular substances, J. Chem. Phys. 56
[73] T.T. Duignan, D.F. Parsons, B.W. Ninham, A continuum solvent model of the (1972) 5499–5510.
multipolar dispersion solvation energy, J. Phys. Chem. B 117 (2013) [105] M. Eldrup, D. Lightbody, J.N. Sherwood, The temperature dependence of
9412–9420. positron lifetimes in solid pivalic acid, Chem. Phys. 63 (1981) 51–58.
[74] T.T. Duignan, D.F. Parsons, B.W. Ninham, A continuum solvent model of the [106] M. Eldrup, O. Mogensen, Positron lifetimes in pure and doped ice and in
partial molar volumes and entropies of ionic solvation, J. Phys. Chem. B 118 water, J. Chem. Phys. 57 (1972) 495–504.
(2014) 3122–3132. [107] G.M. Geise, C.M. Doherty, A.J. Hill, B.D. Freeman, D.R. Paul, Free volume
[75] J.-c Liu, G.-z Jia, Dielectric relaxation of 1-methylimidazole-ethanol mixtures characterization of sulfonated styrenic pentablock copolymers using positron
at the microwave frequency, Colloid Polym. Sci. 293 (2015) 2053–2059. annihilation lifetime spectroscopy, J. Membr. Sci. 453 (2014) 425–434.
[76] Z. Lu, E. Manias, D.D. Macdonald, M. Lanagan, Dielectric relaxation in di- [108] Y. Kobayashi, H.F.M. Mohamed, A. Ohira, Positronium formation in aromatic
methyl sulfoxide/water mixtures studied by microwave dielectric relaxation polymer electrolytes for fuel cells, J. Phys. Chem. Lett. 113 (2009) 5698–5701.
spectroscopy, J. Phys. Chem. A 113 (2009) 12207–12214. [109] E. Dražević, K. Košutić, V. Kolev, V. Freger, Does hindered transport theory
[77] P.B. Undre, P.W. Khirade, V.S. Rajenimbalkar, S.N. Helambe, S.C. Mehrotra, apply to desalination membranes? Environ. Sci. Technol. 48 (2014)
Dielectric relaxation in ethylene glycol – dimethyl sulfoxide mixtures as a 11471–11478.
function of composition and temperature, J. Korean Chem. Soc. 56 (2012) [110] A.C. Sagle, H. Ju, B.D. Freeman, M.M. Sharma, PEG-based hydrogel membrane
416–423. coatings, Polymer 50 (2009) 756–766.
[78] P.W. Khirade, A. Chaudhari, J.B. Shinde, S.N. Helambe, S.C. Mehrotra, Static [111] G.E. Zaikov, A.P. Iordanskii, V.S. Markin, Diffusion of electrolytes in polymers,
dielectric constant and relaxation time measurements on binary mixtures of in, VSP, Utrecht, The Netherlands, 1988, pp. 321.
dimethyl sulfoxide with ethanol, 2-ethoxyethanol, and propan-1-ol at 293, [112] A.E. Yaroshchuk, Non-steric mechanisms of nanofiltration: superposition of
303, 313, and 323 K, J. Chem. Eng. Data 44 (1999) 879–881. Donnan and dielectric exclusion, Sep. Purif. Technol. 22–23 (2001) 143–158.
[79] S.J. Paddison, D.W. Reagor, T.A. Zawodzinski, High frequency dielectric stu- [113] E. Glueckauf, R.E. Watts, The Donnan law and its application to ion-exchanger
s
dies of hydrated Nafion , J. Electroanal. Chem. 459 (1998) 91–97. polymers, Proceedings of the Royal Society of London, Series A: Mathema-
[80] D.R. Paul, Diffusive transport in swollen polymer membranes, in: H. tical, Physical and Engineering Sciences, 268 (1962) 339–349.
B. Hopfenberg (Ed.), Permeability of Plastic Films and Coatings: To Gases, [114] E. Glueckauf, The distribution of electrolytes between cellulose acetate
Vapors, and Liquids, Plenum Press, New York, 1974, pp. 35–48. membranes and aqueous solutions, Desalination 18 (1976) 155–172.
[81] D.R. Paul, Relation between hydraulic permeability and diffusion in homo- [115] E. Glueckauf, A new approach to ion-exchange polymers, Proceedings of the
geneous swollen membranes, J. Polym. Sci. Polym. Phys. Ed. 11 (1973) Royal Society of London, Series A: Mathematical, Physical and Engineering
289–296. Sciences, 268 (1962) 350–370.
[82] P.J. Flory, J.J. Rehner, Statistical mechanics of cross-linked polymer networks [116] M.E. Heyde, C.R. Peters, J.E. Anderson, Factors influencing reverse osmosis
II. Swelling, J. Chem. Phys. 11 (1943) 521–526. rejection of inorganic solutes from aqueous solution, J. Colloid Interface Sci.
[83] P.J. Flory, Thermodynamics of high polymer solutions, J. Chem. Phys. 10 50 (1975) 467–487.
(1942) 51–61. [117] M.E. Heyde, J.E. Anderson, Ion sorption by cellulose acetate membranes from
[84] W. Xie, H. Ju, G.M. Geise, B.D. Freeman, J.I. Mardel, A.J. Hill, J.E. McGrath, binary salt solutions, J. Phys. Chem. 79 (1975) 1659–1664.
Effect of free volume on water and salt transport properties in directly co- [118] J.E. Anderson, H.W. Jackson, Membrane-water partition coefficients of ions.
polymerized disulfonated poly(arylene ether sulfone) random copolymers, Calculated effects of membrane thickness, J. Phys. Chem. 78 (1974)
Macromolecules 44 (2011) 4428–4438. 2259–2262.
[85] Y.-J. Kim, K.-S. Lee, M.-H. Jeong, J.-S. Lee, Highly chlorine-resistant end-group [119] J. Kamcev, D.R. Paul, B.D. Freeman, Ion activity coefficients in ion exchange
crosslinked sulfonated-fluorinated poly(arylene ether) for reverse osmosis polymers: applicability of manning's counterion condensation theory, Mac-
membrane, J. Membr. Sci. 378 (2011) 512–519. romolecules 48 (2015) 8011–8024.
[86] C.H. Lee, D. VanHouten, O. Lane, J.E. McGrath, J. Hou, L.A. Madsen, J. Spano, [120] A. Szymczyk, P. Fievet, Investigating transport properties of nanofiltration
S. Wi, J. Cook, W. Xie, H.J. Oh, G.M. Geise, B.D. Freeman, Disulfonated poly membranes by means of a steric, electric and dielectric exclusion model, J.
(arylene ether sulfone) random copolymer blends tuned for rapid water Membr. Sci. 252 (2005) 77–88.
permeation via cation complexation with poly(ethylene glycol) oligomers, [121] S. Bandini, D. Vezzani, Nanofiltration modeling: the role of dielectric exclu-
Chem. Mater. 23 (2011) 1039–1049. sion in membrane characterization, Chem. Eng. Sci. 58 (2003) 3303.
[87] M.R. Hibbs, M.A. Hickner, T.M. Alam, S.K. McIntyre, C.H. Fujimoto, C. [122] A. Seidel, J.J. Waypa, M. Elimelech, Role of charge (Donnan) exclusion in re-
J. Cornelius, Transport properties of hydroxide and proton conducting moval of arsenic from water by a negatively charged porous nanofiltration
membranes, Chem. Mater. 20 (2008) 2566–2573. membrane, Environ. Eng. Sci. 18 (2001) 105–113.
[88] S. Zhang, R. Zhang, Y.C. Jean, D.R. Paul, T.-S. Chung, Cellulose esters for for- [123] S. Mafe, P. Ramirez, J. Pellicer, Activity coefficients and Donnan coion ex-
ward osmosis: characterization of water and salt transport properties and clusion in charged membranes with weak-acid fixed charge groups, J.
free volume, Polymer 53 (2012) 2664–2672. Membr. Sci. 138 (1998) 269–277.
[89] A.C. Sagle, PEG hydrogels as anti-fouling coatings for reverse osmosis [124] S. Mafe, P. Ramirez, J.A. Manzanares, Activity coefficients and coion exclusion
membranes, in: Chemical Engineering, The University of Texas, Austin, TX, in charged polymeric membranes with macroscopic inhomogeneities, Col-
2009, pp. 168. loid Polym. Sci. 275 (1997) 599–603.
[90] L. Ni, J. Meng, G.M. Geise, Y. Zhang, J. Zhou, Water and salt transport prop- [125] L. Jones, P.N. Pintauro, H. Tang, Coion exclusion properties of polypho-
erties of zwitterionic polymers film, J. Membr. Sci. 491 (2015) 73–81. sphazene ion-exchange membranes, J. Membr. Sci. 162 (1999) 135–143.
[91] Q. Duan, H. Wang, J. Benziger, Transport of liquid water through Nafion [126] R. Tandon, P.N. Pintauro, Solvent effects during multicomponent ion uptake
membranes, J. Membr. Sci. 392–393 (2012) 88–94. into a Nafion cation-exchange membrane, J. Membr. Sci. 341 (2009) 21–29.
800 H. Zhang, G.M. Geise / Journal of Membrane Science 520 (2016) 790–800

[127] R. Tandon, P.N. Pintauro, Divalent/monovalent cation uptake selectivity in a [129] J. Palomo, P.N. Pintauro, Competitive absorption of quaternary ammonium
Nafion cation-exchange membrane: experimental and modeling studies, J. and alkali metal cations into a Nafion cation-exchange membrane, J. Membr.
Membr. Sci. 136 (1997) 207–219. Sci. 215 (2003) 103–114.
[128] P.N. Pintauro, D.N. Bennion, Mass transport of electrolytes in membranes. 2. [130] J.R. Bontha, P.N. Pintauro, Water orientation and ion solvation effects during
Determination of sodium chloride equilibrium and transport parameters for multicomponent salt partitioning in a Nafion cation exchange membrane,
Nafion, Ind. Eng. Chem. Fundam. 23 (1984) 234–243. Chem. Eng. Sci. 49 (1994) 3835–3851.

You might also like