Nano Composite Reverse Electrodialysis (RED) Ion-Exchange Membranes For Salinity Gradient Power Generation - 2014

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Journal of Membrane Science 460 (2014) 139–147

Contents lists available at ScienceDirect

Journal of Membrane Science


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

Nanocomposite reverse electrodialysis (RED) ion-exchange


membranes for salinity gradient power generation
Jin Gi Hong, Yongsheng Chen n
School of Civil and Environmental Engineering, Georgia Institute of Technology, Atlanta, GA 30332, United States

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

Article history: Despite the important role of ion-exchange membranes (IEMs) in reverse electrodialysis (RED) systems,
Received 11 October 2013 the current absence of proper ion-exchange membranes delays the sustainable development of the RED
Received in revised form process for salinity gradient power generation. This research presents the preparation of a new type of
7 February 2014
organic–inorganic nanocomposite cation exchange membrane and its performance characteristics. The
Accepted 19 February 2014
Available online 27 February 2014
combination of functionalized iron (III) oxide ðFe2 O3 SO4 2  Þ as an inorganic filler with the sulfonated
poly (2,6-dimethyl-1,4-phenylene oxide) (sPPO) polymer matrix proved to have great potential for power
Keywords: generation by RED. The results showed that an optimal amount of Fe2 O3 SO4 2  (0.5–0.7 wt%) enhanced
Ion-exchange membranes the key electrochemical properties of the ion-exchange membranes including a permselectivity up to
Nanocomposite
87.65% and an area resistance of 0.87 Ω cm2. The nanocomposite membrane containing 0.7 wt%
Reverse electrodialysis
ðFe2 O3 SO4 2  Þ achieved a maximum power density (amount of power per unit membrane area) of
Salinity gradient power
Ion transport 1.3 W m  2, which is relatively higher than that of the commercially available CSO (SelemionTM, Japan)
membranes. The goal of the present work is to maximize the salinity gradient power generation by
developing RED-specific nanocomposite IEMs. The results show the potential of the new design of the
nanocomposite IEMs for viable energy generation by RED.
& 2014 Elsevier B.V. All rights reserved.

1. Introduction not yet received significant attention [6,7]. Furthermore, drinking


water treatment and desalination generate highly concentrated
The use of limited energy resources, such as fossil fuels, with- saltwater (i.e., brine) as a by-product. Application of SGP has a
out securing practically realistic alternatives has long been recog- great potential to solve these salt problems while producing clean
nized as an environmental and energy sustainability problem. and sustainable electrical energy. Salinity gradient energy remains
Renewable and sustainable energy sources are becoming extre- a large untapped resource, and greater development of an engi-
mely critical due to growing environmental problems such as neered and optimized energy-capturing effort is needed.
pollution and global warming [1]. In recent years, membrane- One of the most promising methods to capture salinity gradient
based salinity gradient energy has received much attention as a energy is reverse electrodialysis (RED) [7–9], which generates
potential renewable and nonpolluting sustainable energy source power via the transport of the positive and negative ions in the
[2,3]. Salinity gradient power (SGP) is the energy harvested from water through selective ion-exchange membranes (IEMs). A typi-
reversible mixing of two water streams with different salt con- cal RED device consists of an alternating series of anion exchange
centrations (e.g., sea and river water). The worldwide availability membranes (AEMs) and cation exchange membranes (CEMs), as
of SGP makes it potentially the second-largest marine-based shown in Fig. 1. Woven fabric spacers separate these membranes,
energy source, with an estimated global power production poten- forming thin compartments and fixing the intermembrane
tial of 2.4–2.6 TW [3–5]. A remarkable aspect of this membrane- distance. In RED, a highly concentrated salt solution (e.g., sea-
based process is that it can also be an adequate treatment for water) and a dilute salt solution (e.g., river water) are brought into
excessive salt accumulation, such as that in agricultural irrigation contact through arrays of polymeric AEMs and CEMs; this contact
return water and reverse osmosis (RO) reject waters. Proper directly generates an electrochemical potential gradient. The
disposal or, even better, utilization of these ‘leftover’ salts has salinity difference between the two adjacent compartments drives
cations toward the cathode, whereas anions move in the opposite
direction toward the anode. At the electrodes, the ionic species
n
Corresponding author. Tel.: þ 1 404 894 3089.
accumulate and induce electronic current via redox reactions. The
E-mail address: yongsheng.chen@ce.gatech.edu (Y. Chen). redox process is determined by the electrodes and electrode rinse

http://dx.doi.org/10.1016/j.memsci.2014.02.027
0376-7388 & 2014 Elsevier B.V. All rights reserved.
140 J.G. Hong, Y. Chen / Journal of Membrane Science 460 (2014) 139–147

Fig. 1. Simplified schematic view of an RED stack representing the fluid transport through the ion-exchange membranes (i.e., CEM and AEM).

used in the system. A system with the opposite electrode redox features of inorganic materials with those of organic materials and
reaction and electrode rinse recirculation will lead to no net have received increasing attention for more than a decade owing
chemical reaction (i.e., zero equilibrium voltage). This electronic to their remarkable synergized properties [29]. Here we focus on
current and the potential difference over the electrodes can be the development of organic–inorganic nanocomposite IEMs based
used to generate electricity, when an external load or any form of on sulfonated poly (2,6-dimethyl-1,4-phenylene oxide) (sPPO) and
energy consumer is connected to the system [10,11]. As a full-scale sulfonated iron oxide ðFe2 O3 SO4 2  Þ created via the blending
RED system, known as a stack, is composed of multiple membrane method. PPO is a commonly used polymer in membranes for gas
cell pairs, RED may become practical through the development of separation, RO, ultrafiltration, and ion exchange [30–33] because
an RED-specific membrane. of its excellent membrane-forming properties; good chemical,
IEMs have attracted great interest and much research across thermal, and hydrolytic stability; high glass transition temperature
fields ranging from the bioindustries (food, pharmaceuticals, and (tg ¼210 1C); and low cost [33–35]. However, the hydrophobicity of
biotechnology) [12–14], to fuel cells [15–17], to desalination PPO often hinders its dissolution in conventional dipolar solvents.
[18,19], Each application normally requires the membranes to To make the polymer more ion-exchangeable, a conventional
have specific physical and electrochemical properties. As in many electrophilic substitution is often applied to introduce charged
other membrane processes, the membrane characteristics primar- groups (i.e., anionic or cationic moieties), typically from a sulfonic
ily determine the performance of an RED system. In most reported acid. Introduction of sulfonate groups (–SO3  ) to the polymer
RED studies, various commercially available membranes were chain of PPO via a sulfonating agent such as chlorosulfonic acid
tested and evaluated within a RED stack [7,9,20–22], but relatively causes PPO to become highly hydrophilic [36].
little is known about the key properties that determine membrane Recently, several studies have incorporated sulfonated meso-
performance in power generation via RED. Several researchers porous iron (III) oxide ðFe2 O3 SO4 2  Þ into proton conductive
have concluded that the development of RED-specific membranes membranes (PCMs) for fuel cell applications [9,37,38]. The studies
with low electrical resistance, high transport number, and high revealed that sulfonated Fe2O3 facilitates not only conductivity but
selectivity is necessary [4,23–26]. However, it remains challenging also strong hydrophilicity and large specific surface area. Sulfo-
to tailor membranes with such properties [27,28]. Thus, it is nated Fe2O3 is expected to be a good filler material because of its
worthwhile to investigate the key membrane properties necessary high versatilities such as (1) high specific surface area (decreases
to maximize overall RED performance both to advance our under- crystallites size when functionalized), (2) wide use and cost-
standing and to take steps toward the development of a new RED- effectiveness (easy to produce), (3) high ability to adsorb various
specific membrane. inorganic and organic materials, (4) easy coatability on various
The important membrane properties for power generation, surfaces, and (5) high hydrophilicity [37,39–43]. Moreover,
particularly those by electrochemical processes such as RED, are embedding such a functionalized inorganic filler in the structure
ion-exchange capacity (IEC), swelling degree (SD), membrane can be considered as a homogeneous ion-exchange resin up to
resistance, and permselectivity. Composite membranes created certain level of loading [19]. It is important to have a homogeneous
by introducing inorganic materials into an organic polymer matrix ion exchange resin, because the fixed charge groups are more
are viable candidates for RED, as they allow carrying of extra ion- evenly distributed over the entire membrane matrix in homo-
exchangeable functional groups by modulating the membrane geneous membranes, while heterogeneous membranes have
structure. Also, organic–inorganic composites combine the unique distinct region of uncharged polymer of ion exchange resin in
J.G. Hong, Y. Chen / Journal of Membrane Science 460 (2014) 139–147 141

the membrane matrix. Therefore, charge density in heterogeneous solution. The resulting mixture was cast on a glass substrate using
ion exchange membranes is relatively low compared to homo- the doctor blade method to obtain the desired thickness of the
geneous membranes and resistance is higher owing to its mem- membrane (100 mm). The obtained solid composite membrane was
brane structure where uncharged domain exists in the polymer then dried in a vacuum oven at 60 1C for 24 h and at 80 1C for
matrix [23]. Therefore, the sulfonated Fe2O3 particles are primarily another 24 h to remove residual solvents. The prepared membrane
introduced to increase ion-exchangeable functional groups in the was treated in warm water (40–50 1C) and then in 1 M HCl for 24 h.
membrane, which results in enhanced ion-exchange capacity, Finally, the membranes were rinsed with DI water and stored in
hydrophilicity and ion conductivity. Note that the introduction of 0.5 M NaCl solution.
negatively charged groups (thus, increase in charge density)
increases proton conductivity as well as sodium conductivity, but
2.3. Membrane characterization
increase in sodium conductivity may not necessarily be the same
as that of proton conductivity. However, until now, no research
2.3.1. FTIR spectroscopy
exploring the use of Fe2 O3 SO4 2  in the PPO polymer matrix has
The chemical structures of composite membranes were inves-
been reported in any field. In this study, we introduce our design
tigated using Fourier transform infrared (FTIR) spectra. FTIR
strategy for the development of a new type of composite RED
spectra of sPPO membranes were acquired with FTIR Spectrometer
membrane: incorporation of Fe2 O3 SO4 2  and sPPO polymer.
(Digilab FTS7000) equipped with a microscope (Digilab UMA600),
The effect of Fe2 O3 SO4 2  loading is also investigated to optimize
collecting 32 scans per sample at a 4 cm  1 resolution and a
the membrane performance in RED.
spectral range of 4000–700 cm  1. The spectrum of ambient air
was used as background.
2. Experimental
2.3.2. Membrane morphology
2.1. Materials Scanning electron microscopy (SEM, Zeiss Ultra60 FE-SEM) was
used to observe the morphology and structure of the prepared
Poly (2,6-dimethyl-1,4-phenylene oxide) (PPO) (Aldrich, analy- membranes. For the cross-sectional surface, the samples were
tical standard) was used as received. Chloroform (Aldrich, anhy- prepared in liquid nitrogen and cut manually to obtain sharp
drous, 99%) was used as a solvent. Chlorosulfonic acid (Aldrich, cross-section. Membrane samples were then dried overnight to
99%) and sulfuric acid (Aldrich, 98%) were used without purifica- preserve their structure.
tion. Dimethylsulfoxide (DMSO) (ACS grade, 99.9%) was obtained
from VWR. Iron (III) oxides (Aldrich, o 50 nm) were used as
received for sulfonation. 2.3.3. Swelling degree (SD)
The swelling not only determines mechanical properties and
2.2. Preparation of ion-exchange membranes dimensional stability of the membrane, it also influences its ability
of ion selectivity and electrical resistance. The SD simply refers to
2.2.1. Sulfonation of PPO (sPPO) the amount of water content in the membrane per unit weight of
Poly (2,6-dimethyl-1,4-phenylene oxide) (PPO) was sulfonated dry membrane. To determine the SD of a membrane, a sample was
with chlorosulfonic acid as the sulfonating agent and chloroform equilibrated in DI water for at least 24 h. The membrane was then
as the solvent. PPO of 6 wt% was dissolved in chloroform by taken out and surface water was removed from the sample with a
mechanical mixing at room temperature for 30 min. 8 wt% solu- blotting paper. The wet weight of the swollen membrane was
tion of chlorosulfonic acid in 50 ml of chloroform was slowly measured. The membrane was then dried at 50 1C in a vacuum
added to the PPO solution over a period of 30 min while stirring oven until a constant dry weight was obtained. The SD of a
the solution vigorously at 25–30 1C until the sulfonated PPO was membrane was calculated in wt% by the following equation:
precipitated. The sulfonated precipitate was washed multiple W wet  W dry
times with DI water until pH became approximately 5–6. The SD ¼  100% ð1Þ
W dry
polymer was then filtered and dissolved in methanol. The solution
was poured into a Pyrex glass tray forming a thin film of about where Wwet (g) and Wdry (g) are the weight of the wet and the dry
1–2 mm thickness, and allowed to dry in air for 24 h at room membrane, respectively.
temperature. The dried polymer sheet was shredded into small
particles and washed with DI water. The sPPO was then filtered 2.4. Electrochemical properties of ion-exchange membranes
and dried at room temperature.
2.4.1. Ion-exchange capacity (IEC)
2.2.2. Sulfonation of Fe2O3 ðFe2 O3 SO4 2  Þ The ion-exchange capacity, which is the number of fixed
The powder of Fe2O3 nanoparticles was sulfonated by concen- charges per unit weight of the membrane, was measured using a
trated H2SO4 (98%). The required amount of Fe2O3 was dissolved in titration method [44]. First, the cation exchange membrane was
0.25 M concentrated H2SO4 at room temperature for 24 h. The equilibrated in 1 M HCl for at least 15 h, and then rinsed free from
concentrated H2SO4 treated Fe2O3 was then filtered and dried at chloride with DI water. After that, the membrane was immersed in
80 1C in a vacuum oven. Finally, the powder was calcined at 500 1C 1 M NaCl solution for another 6 h. The displaced hydrogen ions
for 3 h to obtain the red sulfonated Fe2O3 powder. from the membrane were then titrated with 0.01 M NaOH solution
using phenolphthalein as an indicator. The IEC of the membranes
2.2.3. Membrane preparation was then calculated using the following equation:
Ion-exchange membranes were prepared by the solvent evapora- C NaOH  V NaOH
tion technique. Sulfonated PPO polymer was dissolved in Dimethyl- IEC ¼ ð2Þ
W dry
sulfoxide (DMSO) to obtain 19 wt% solutions. Then 0–2 wt% of
sulfonated Fe2O3 was mixed with the polymer solution at 40 1C for where CNaOH is the concentration of NaOH solution (mol L  1),
24 h while stirring. The solution was vibrated in an ultrasonic bath VNaOH is the volume of NaOH solution (L) used and Wdry is the dry
for 10 min to obtain an optimal dispersion of the particle in the weight (g) of the membrane.
142 J.G. Hong, Y. Chen / Journal of Membrane Science 460 (2014) 139–147

2.4.2. Fixed-charge density (ΔVtheoretical) as shown in the following equation:


Fixed-charged groups are attached to the polymer backbone in
ΔV measured
the ion-exchange membranes. The counter ion transport and ion α ð%Þ ¼  100 ð4Þ
ΔV theoretical
permselectivity through the membrane are determined by the
fixed-charge density in the membrane. The fixed-charge density, where α is the membrane permselectivity (%), ΔVmeasured is the
expressed in milliequivalent of fixed groups per volume of water measured membrane potential (V) and ΔVtheoretical is the theore-
in the membrane (meq L  1) can be determined by the IEC over the tical membrane potential. Note that the theoretical membrane
SD of the membrane. potential is the membrane potential for an ideal 100% permselec-
tive membrane, which was estimated to be 0.0379 V from the
IEC
C fix ¼ ð3Þ Nernst equation [46].
SD

2.5. RED membrane performance

2.4.3. Electrical resistance The RED performance of the synthesized membranes was
The electrical resistance of the membrane was measured in tested in a RED stack described in the previous study [47]. The
a two-compartment cell using 0.5 M NaCl aqueous solutions. RED stack used in this study initially consists of three cell pairs,
Titanium electrodes, coated with Ru–Ir mixed metal oxides were each containing a cation exchange membrane (CSO) and an anion
used (Jing Run Beijing Science and Technology Research Institute exchange membrane (ASV) (Selemion, Japan) and the CSO mem-
Company, China). The resistance of the membranes was measured branes were replaced later with the synthesized membranes for a
at room temperature by impedance spectroscopy (IS) using Auto- comparison. Each membrane has an effective area of 4 cm  9 cm.
lab PGSTAT302N, Metrohm in a frequency range from 10–105 Hz A CEM was placed at each end of the stack as a shielding
with an oscillating voltage of 0.1 V amplitude [45]. The resistance membrane. Titanium mesh end electrodes coated with iridium
was measured based on the intercept of x-axis (Z0 , the real part of plasma were used as the anode and cathode. For the electrode
the transfer function) in the Nyquist plot. Then the membrane rinse solution, we used a solution of NaCl (0.25 M) with K4Fe(CN)6
resistance (Rmem) was determined by subtraction of the resistance (0.05 M) and K3Fe(CN)6 (0.05 M). The intermembrane distance was
measured in the blank test (Rsol) from the resistance measured determined by woven fabric spacers (porosity 60%) at a thickness
with the membrane under investigation (Rcell), based on the total of 250 mm, which was placed between each membrane to form
cell resistance composition (i.e., Rcell ¼Rsol þRmem). water compartments for the synthetic seawater and river water
solutions to pass through. The synthetic feed water had a con-
centration of 0.5 M NaCl for seawater and 0.017 M NaCl for river
2.4.4. Permselectivity water. The feed solutions for both waters were delivered using
The apparent permselectivity of an ion-exchange membrane Masterflex peristaltic pumps (Cole-Parmer, USA) through the RED
was determined using a static membrane potential measurement. stack at various flow rates.
The experimental setup consists of a two-compartment cell A four electrodes configuration allows performance evaluation
separated by the testing membrane with an effective area of of the RED stack using an Autolab potentiostats (Metrohm, The
4.8 cm2. Two aqueous solutions of 0.1 M and 0.5 M NaCl were Netherlands) in the galvanostatic mode. At each flow rate, the
placed in each cell along the sides of the test membrane. The voltage (E) as function of the electrical current (I) was measured.
potential difference over the membrane was measured at room From the measured E–I curves, the gross power generation was
temperature with two Ag/AgCl reference electrodes (Hanna Instru- calculated. The corrected gross power was determined by subtrac-
ments, USA) using a multimeter (Tektronix, USA). The permselec- tion of the power generated in a blank run, which was tested with
tivity can be calculated from the ratio of the measured membrane only one commercial anion exchange membrane in the stack
potential (ΔVmeasured) over the theoretical membrane potential (Fig. 2).

Fig. 2. Scheme of the RED stack used in the performance test. Note that the number of cell pairs used in the performance test is different from what is presented in this
illustration.
J.G. Hong, Y. Chen / Journal of Membrane Science 460 (2014) 139–147 143

3. Results and discussion via the blending method and solvent evaporation. SEM was used
to observe the morphology of the synthesized Fe2O3–PPO cation
3.1. FTIR spectra study exchange membranes. Fig. 5 shows the cross-sectional surface
morphologies of the obtained nanocomposite membranes. It is
FTIR spectra were obtained from nanocomposite membrane clearly recognized that the cross-sections were all dense structures
film samples. All the spectra show C–H stretch of CH2 and CH3 without apparent large porosities. Note that settling of the particle
throughout 2868 and 2970 cm  1 and C–O–C stretch at 942 cm  1, in the membrane was not observed.
which represent the original PPO polymer. As shown in Fig. 3, the
characteristic absorption band for the presence of the –SO3H 3.3. Electrochemical properties of the nanocomposite ion-exchange
substitute to PPO aromatic rings appeared by the FTIR peak at membranes
1060 cm  1 that is formed from the sulfonation reaction. Also a
broadband associated with hydrogen reaction between –OH sPPO membranes were hybridized with Fe2 O3 SO4 2  at var-
groups and –SO3H was visible throughout 3300 and 3500 cm  1 ious concentrations between 0.2 and 2.0 wt%. Table 1 lists the
for all the sPPO membranes. IR vibration bands for the character- characteristics of the IEMs that are key to investigating their
istics of S¼ O stretching for the sulfate ðSO4 2  Þ groups attached to potential performance in a RED system. Prepared nanocomposite
the iron oxide were observed at 1400 cm  1 (Fig. 3 b–f). Overall, membranes were evaluated for permselectivity, area resistance,
these confirm the successful sulfonation of the PPO polymer and IEC, SD, and CD. Each electrochemical property test was performed
iron oxide nanoparticles and the presence of ion exchangeable three times to ensure reliable data, and the values are listed in
groups in these organic and inorganic components. Table 1. The errors are less than 5% and omitted in the table to
avoid confusion. In addition, a ratio, P2/R (also referred to as the
performance potential ratio), which derives simply from perms-
3.2. Morphology of the nanocomposite ion-exchange membranes electivity (P2) over area resistance (R), is also presented. IEMs are
often selected on the basis of their high permselectivity and low
The sulfonated iron oxide ðFe2 O3 SO4 2  Þ was prepared based resistance; thus, we assume that P2/R is a useful indicator of the
on previously reported procedure [37]. The nanoparticles were best-performing membranes for power generation. Considering
treated in sulfuric acid before it was mixed with the prepared that which membrane property dominates RED system perfor-
polymer solution, sPPO. The image of the sulfonated Fe2O3 mance is not yet fully known and the relationships between these
particles was observed on a scanning electron microscopy properties are not clearly understood, P2/R can be an indication
(Fig. 4), which indicates the particle sizes of 10–50 nm. Nanocom- only of power generation potential associated with permselec-
posite membranes were formed by mixing the sulfonated iron tivity and resistance. The properties of the commercial CEM, CSO
oxide ðFe2 O3 SO4 2  Þ nanoparticles and the sPPO polymer matrix (Selemion, Japan), were also evaluated for comparison.

3.3.1. Effect of nanoparticle loading on IEC and SD


The ion-exchange capacity (IEC) and swelling degree (SD) of
charged membranes are the key parameters that affect the degree
of charge density (CD). The IEC and SD of the membranes
increased as the Fe2 O3 SO4 2  concentration increased within
the range of 0–0.7 wt% (Fig. 6). The increase in swelling degree
was expected with the increase of inorganic filler loadings.
However, further increase of Fe2 O3 SO4 2  concentration ( 4
0.7 wt%) did not continue to increase the swelling degree. This
may be attributed to the interaction among the particles that lead
particle aggregation, and thus interrupt the functional groups of
the polymer matrix. An increasing trend of IEC with nanoparticle
loading can be explained by the increase of negatively charged
functional groups in the polymer matrix, attracting more cationic
species while repelling anionic species. The increased IEC leads to
the enhanced ion-conducting character of the nanocomposite. The
Fig. 3. FTIR spectra of nanocomposite membranes: (A) pristine, (B) 0.2 wt% greater number of fixed charges (negative charges in case of CEM)
Fe2 O3 SO4 2  , (C) 0.5 wt% Fe2 O3 SO4 2  , (D) 0.7 wt% Fe2 O3 SO4 2  , (E) 1.0 wt% in the membrane gives rise to enhanced transport properties,
Fe2 O3 SO4 2  , (F) 2.0 wt% Fe2 O3 SO4 2  , (G) Fe2 O3 SO4 2  . which increase the hydrophilic nature of the membrane. Thus, the

Fig. 4. SEM micrographs of (a) sulfonated iron oxide particles, Fe2 O3 SO4 2  and (b) composite membrane with Fe2 O3 SO4 2  .
144 J.G. Hong, Y. Chen / Journal of Membrane Science 460 (2014) 139–147

Fig. 5. SEM micrographs of cross section surfaces: (a) pristine sPPO membrane, (b) composite sPPO–0.2 wt% Fe2 O3 SO4 2  , (c) composite sPPO–0.5 wt% Fe2 O3 SO4 2  ,
(d) composite sPPO–0.7 wt% Fe2 O3 SO4 2  , (e) composite sPPO–1.0 wt% Fe2 O3 SO4 2  , and (f) composite sPPO-2.0 wt% Fe2 O3 SO4 2  .

membrane swells more [48]. In general, the greater swelling can


lead to lower mechanical strength and stability; however, several
studies have reported that organic–inorganic nanocomposites
enhance the mechanical properties of the material [29,49,50].
Moreover, the IEC did not continue to increase as more
Fe2 O3 SO4 2  ( 40.7 wt%) was incorporated. This phenomenon
may stem from the tendency of inorganic nanoparticles to aggre-
gate and thus to disrupt the functional groups of the polymer
matrix when excess inorganic nanoparticles are present. Specifi-
cally, when inorganic fillers form a large cluster, the surface of the
functional groups reduces, thus starting to interfere with the
charged functional groups of the polymer matrix. Moreover, the
ion-exchangeable functional groups contributed from sPPO could
also be reduced, as more inorganic nanoparticles agglomerated
(4 0.7 wt% Fe2 O3 SO4 2  ) in the composite. Consequently, the IEC
as well as other charge transport properties could be affected by Fig. 6. Ion-exchange capacity (IEC) and swelling degree (SD) of sPPO membranes as
excess nanoparticle loading to decrease the overall membrane a function of Fe2 O3 SO4 2  wt%.
performance [51,52].
Referring to the result that the IEC increased with the increase
of Fe2 O3 SO4 2  concentrations to a certain point (0.7 wt %), the
fixed-charge density also increased in a similar manner (Table 1).
3.3.2. Area resistance and permselectivity of nanocomposite IEMs As more ionic groups are present (fixed charges), the more
The charge density of the membrane has a great effect on its swelling occurs, so low area resistances are normally expected
ion permselectivity and area resistance [23]. The number of with a high swelling degree. However, it is not always the case
charged functional groups per water molecule absorbed in the with a high IEC that results in low resistance [26]. 0.7 wt%
membrane determines the degree of charge density. At a high Fe2 O3 SO4 2  with the highest IEC and SD showed higher resis-
charge density, more effective ion discrimination is possible (co- tance than that of 0.5 wt% Fe2 O3 SO4 2  (by 11%). The pristine
ion exclusion). In our experiment, we observed a decrease of area (0 wt%) and 0.2 wt% membranes have reasonable IEC and SD, but
resistance with increasing charge density (Table 1). However, we with a high area resistance; this can be explained by the relatively
often observe other random behavior that the resistance increases low charge density of these membranes. Similarly for the com-
with increasing fixed charge density [26,53]. This phenomenon mercial CSO membrane, having high IEC did not result in low area
may be attributed to a high degree of cross-linking, which resistance, which again implies that the relationship of IEC and
improves the mechanical strength of the membranes and thus resistance is not always inversely correlated [26,53].
increases in its resistance. Moreover, the pore formation and The ion permselectivity and area resistance of a membrane are
surface roughness of the composite have also been reported to often considered as the most important parameters for the
affect the transport of ions, permselectivity and conductivity [54]. selection of effective IEMs in RED power generation. Although
Thus, such membrane surface properties may contribute to this charge density has a great impact on membrane permselectivity
random tendency. However, the effect of these properties will be and resistance, charge density’s relationship with each of these
further investigated in future studies. parameters is not always linear, but rather varies according to the
J.G. Hong, Y. Chen / Journal of Membrane Science 460 (2014) 139–147 145

Table 1
Characteristics of the prepared membranes compared with a commercial membrane, CSO for RED performance.

Membranes Permselectivity [%] Area resistance [Ω cm2] Performance potential ratio P2/Ra IECb [meq g  1] SDc [%] CDd [meq L  1]

0 Fe2 O3 SO4 2  79.13 2.05 3054.4 0.96 21 4.6


0.2 Fe2 O3 SO4 2  84.41 1.09 6536.7 1.00 20 5.0
0.5 Fe2 O3 SO4 2  87.21 0.87 8742.1 1.12 22 5.1
0.7 Fe2 O3 SO4 2  87.65 0.97 7920.1 1.40 26 5.4
1.0 Fe2 O3 SO4 2  87.04 1.18 6420.3 1.03 20 5.2
2.0 Fe2 O3 SO4 2  68.58 1.58 2976.7 0.87 25 3.4
CSO 92.32 2.26 3771.2 1.04 16 6.4

a
Performance potential ratio presented by permselectivity (P) over area resistance (R).
b
Ion-exchange capacity.
c
Swelling degree.
d
Charge density.

Fig. 7. Membrane permselectivity and area resistance of sPPO membranes as a


function of Fe2 O3 SO4 2  wt%.

material [23,53]. The permselectivity of the composite membranes Fig. 8. Performance potential ratio (P2/R) (solid bars) vs. maximum gross power
tended to increase up to 0.7 wt% Fe2 O3 SO4 2  , as expected based density (shaded bars) of nanocomposite membranes.
on the increasing charge density (Fig. 7). The stronger co-ion
exclusion with a high charge density increases the membrane
permselectivity. Conversely, membranes with a lower charge
density have weaker co-ion exclusion and thus lower permselec-
tivity. The membranes with the highest permselectivity were
those containing 0.5, 0.7, and 1 wt% Fe2 O3 SO4 2  . Even though
1 wt% Fe2 O3 SO4 2  had a comparable permselectivity, the high-
est area resistance resulted with a low IEC level among these three
membranes. In Fig. 8, P2/R is illustrated using the relationship of
permselectivity and area resistance. On the one hand, the mem-
branes with the highest performance potential were those con-
taining 0.5 and 0.7 wt% Fe2 O3 SO4 2  , which have the lowest area
resistance as well as high permselectivity. On the other hand, the
pristine and 2.0 wt% Fe2 O3 SO4 2  membranes were rated as the
least effective owing to their high area resistance. Note that this
parameter (P2/R) may only be used as a reference indicator for
Fig. 9. Gross power density as a function of flow rate for the prepared nanocom-
power generation and cannot fully demonstrate the overall power
posite membranes in comparison with the commercial CSO membrane.
generation with such a limited information.

3.4. RED performance of nanocomposite IEMs Fig. 9. The membranes containing 0.5 and 0.7 wt% Fe2 O3 
SO24  exhibited the best performance at various flow rates. A
All the CEMs under investigation were employed in a RED stack maximum gross power density value of 1.3 W m  2 is achieved
containing an ASV membrane as the AEM. We tested the perfor- with 0.7 wt% Fe2 O3 SO4 2  , which exceeded the level of power
mance of nanocomposite membranes in RED and compared it with density acquired with the commercial CSO membrane. This result
that of the commercial CSO membrane. As mentioned earlier, aligns well with the electrochemical properties; the high perms-
several researchers stressed the importance of resistance and electivity of 87.65% and relatively low area resistance of
permselectivity and a direct correlation of area resistance to RED 0.97 Ω cm2 may play a crucial role in determining the power
power density [26,53]. The RED power generation performance output. The pristine membrane (0 wt% Fe2 O3 SO4 2  ), however,
was reasonably consistent with the electrochemical properties performed the worst owing to its high resistance (2.05 Ω cm2) and
discussed above. Membrane performance in RED is illustrated low permselectivity (79.13%). In addition, the amount of Fe2
using gross power density (W m  2) versus flow rate (cm3 s  1) in O3 SO4 2  incorporated influences the polymer matrix and its
146 J.G. Hong, Y. Chen / Journal of Membrane Science 460 (2014) 139–147

intrinsic membrane properties (IEC, SD, and CD), and as such may [10] D.A. Vermaas, E. Guler, M. Saakes, K. Nijmeijer, Theoretical power density from
also be a key factor in achieving such a power density. Although salinity gradients using reverse electrodialysis, in: G. Tranell (Ed.), Technoport
2012—Sharing Possibilities and 2nd Renewable Energy Research Conference,
the trend of these properties agreed well in determining the level 2012, pp. 170–184.
of RED power density, which factor is most important and the [11] J. Veerman, J. Post, M. Saakes, S. Metz, G. Harmsen, Reducing power losses
mutual relationship of electrochemical (IEC, SD, CD, permselec- caused by ionic shortcut currents in reverse electrodialysis stacks by a
validated model, J. Membr. Sci. 310 (2008) 418–430.
tivity and area resistance) and physical properties (porosity, sur- [12] M. Minagawa, A. Tanioka, P. Ramirez, S. Mafe, Amino acid transport through
face roughness) for RED power generation are still unclear. A cation exchange membranes: Effects of pH on interfacial transport, J. Colloid
further investigation on various membrane materials is necessary Interface Sci. 188 (1997) 176–182.
[13] V.K. Shahi, S.K. Thampy, R. Rangarajan, Chronopotentiometric studies on
to provide a persuasive conclusion on this matter. Nevertheless, dialytic properties of glycine across ion-exchange membranes, J. Membr. Sci.
the current practice of designing IEMs is a great step toward more 203 (2002) 43–51.
practical viability of the RED system to fill the growing energy [14] J.H. Choi, S.H. Kim, S.H. Moon, Recovery of lactic acid from sodium lactate by
ion substitution using ion-exchange membrane, Sep. Purif. Technol. 28 (2002)
needs of the world.
69–79.
[15] C.H. Lee, K.A. Min, H.B. Park, Y.T. Hong, B.O. Jung, Y.M. Lee, Sulfonated poly
(arylene ether sulfone)-silica nanocomposite membrane for direct methanol
fuel cell (DMFC), J. Membr. Sci. 303 (2007) 258–266.
4. Conclusions
[16] Y.H. Su, Y.L. Liu, Y.M. Sun, J.Y. Lai, D.M. Wang, Y. Gao, B.J. Liu, M.D. Guiver,
Proton exchange membranes modified with sulfonated silica nanoparticles for
This work emphasizes the importance of membrane role and direct methanol fuel cells, J. Membr. Sci. 296 (2007) 21–28.
its characteristics toward the performance of the RED system. [17] D.W. Seo, Y.D. Lim, S.H. Lee, M.M. Islam, H.M. Jin, K.H. Lee, H.H. Jang, W.G. Kim,
Nano composite membranes of sulfonated amine-poly(ether sulfone)s and SiO
A new type of organic–inorganic nanocomposite cation exchange (2) for fuel cell application, in: H.S. Kim, J.F. Yang, T. Sekino, M. Anpo, S.W. Lee
membranes was designed and evaluated for electrochemical (Eds.), Eco-Materials Processing and Design Xi, 2007, pp. 392–395.
properties and further tested in a RED stack to investigate its [18] C. Klaysom, R. Marschall, L.Z. Wang, B.P. Ladewig, G.Q.M. Lu, Synthesis of
composite ion-exchange membranes and their electrochemical properties for
power density generation. We incorporated a small amount of desalination applications, J. Mater. Chem. 20 (2010) 4669–4674.
sulfonated Fe2O3 in sPPO polymer to improve the properties of [19] C. Klaysom, S.H. Moon, B.P. Ladewig, G.Q.M. Lu, L.Z. Wang, The influence of
ion-exchange membranes. In addition, the effect of Fe2 O3 SO4 2  inorganic filler particle size on composite ion-exchange membranes for
desalination, J. Phys. Chem. C 115 (2011) 15124–15132.
loading was investigated to optimize the membrane performance [20] D.A. Vermaas, M. Saakes, K. Nijmeijer, Doubled power density from salinity
in RED. The ion-exchange capacity, swelling degree, area resis- gradients at reduced inter-membrane distance, Environ. Sci. Technol. 45
tance and permselectivity of the membranes were determined for (2011) 7089–7095.
[21] O. Scialdone, C. Guarisco, S. Grispo, A. D’ Angelo, A. Galia, Investigation of
various Fe2 O3 SO4 2  loadings and were best performed at the electrode material—redox couple systems for reverse electrodialysis processes.
range of 0.5–0.7 wt%. By controlling the inorganic filler amount, Part I: Iron redox couples, J. Electroanal. Chem. 681 (2012) 66–75.
the nanocomposite membrane was found to be optimized at [22] J. Veerman, R. De Jong, M. Saakes, S. Metz, G. Harmsen, Reverse electrodialysis:
0.7 wt% Fe2 O3 SO4 2  for RED performance, which enables us to comparison of six commercial membrane pairs on the thermodynamic
efficiency and power density, J. Membr. Sci. 343 (2009) 7–15.
generate higher power output than the commercially available [23] P. Dlugolecki, K. Nymeijer, S. Metz, M. Wessling, Current status of ion
CSO membrane. This newly-developed nanocomposite membrane exchange membranes for power generation from salinity gradients, J. Membr.
shows the potential to be a feasible candidate for use as ion- Sci. 319 (2008) 214–222.
[24] C. Forgacs, Generation of electricity by reverse electrodialysis, R&D Authority,
exchange membranes in RED power generation. Further investiga- Ben Gurion University, Beer Sheva, Israel, 1978.
tion to better understand the effect of inorganic fillers on organic [25] R.E. Lacey, Energy by reverse electrodialysis, Ocean Eng. 7 (1980) 1–47.
polymer as well as the correlation among key membrane proper- [26] E. Guler, R. Elizen, D. Vermaas, M. Saakes, K. Nijmeijer, Performance-
determining membrane properties in reverse electrodialysis, J. Membr. Sci.
ties even in other compositions should be followed in future (2013).
studies. [27] J.W. Post, C.H. Goeting, J. Valk, S. Goinga, J. Veerman, H.V.M. Hamelers, P. Hack,
Towards implementation of reverse electrodialysis for power generation from
salinity gradients, Desalination Water Treat. 16 (2010) 182–193.
[28] F.G. Wilhelm, N.F.A. van der Vegt, H. Strathmann, M. Wessling, Comparison of
Acknowledgments bipolar membranes by means of chronopotentiometry, J. Membr. Sci. 199
(2002) 177–190.
[29] X.T. Zuo, S.L. Yu, X. Xu, R.L. Bao, J. Xu, W.M. Qu, Preparation of organic-
This research was partially supported by the Litree Purification inorganic hybrid cation-exchange membranes via blending method and their
Company and the U.S. National Science Foundation (NSF Grant no. electrochemical characterization, J. Membr. Sci. 328 (2009) 23–30.
CBET-1235166). [30] D. Wu, L. Wu, J.J. Woo, S.H. Yun, S.J. Seo, T.W. Xu, S.H. Moon, A simple heat
treatment to prepare covalently crosslinked membranes from sulfonated poly
(2,6-dimethyl-1,4-phenylene oxide) for application in fuel cells, J. Membr. Sci.
348 (2010) 167–173.
References [31] T.W. Xu, W.H. Yang, B.L. He, Effect of solvent composition on the sulfonation
degree of poly(phenylene oxide) (PPO), Chin. J. Polym. Sci. 20 (2002)
[1] National Research Council (NRC), Advancing the Science of Climate Change, 53–57.
National Academy Press, Washington, DC, 2010. [32] T.W. Xu, D. Wu, L. Wu, Poly(2,6-dimethyl-1,4-phenylene oxide) (PPO)—A
[2] B.E. Logan, M. Elimelech, Membrane-based processes for sustainable power versatile starting polymer for proton conductive membranes (PCMs), Prog.
generation using water, Nature 488 (2012) 313–319. Polym. Sci. 33 (2008) 894–915.
[3] G.Z. Ramon, B.J. Feinberg, E.M.V. Hoek, Membrane-based production of [33] H. Yu, T.W. Xu, Fundamental studies of homogeneous cation exchange
salinity-gradient power, Energy Environ. Sci. 4 (2011) 4423. membranes from poly(2,6-dimethyl-1,4-phenylene oxide): membranes pre-
[4] J.N. Weinstein, F.B. Leitz, Electric power from differences in salinity: the pared by simultaneous aryl-sulfonation and aryl-bromination, J. Appl. Polym.
dialytic battery, Science 191 (1976) 557. Sci. 100 (2006) 2238–2243.
[5] G.L. Wick, W.R. Schmitt, Prospects for renewable energy from sea, Mar. [34] B. Kruczek, T. Matsuura, Development and characterization of homogeneous
Technol. Soc. J. 11 (1977) 16–21. membranes de from high molecular weight sulfonated polyphenylene oxide, J.
[6] J.L. Schnoor, Salt: the final frontier, Environ. Sci. Technol. (2013). Membr. Sci. 146 (1998) 263–275.
[7] J.W. Post, H.V.M. Hamelers, C.J.N. Buisman, Energy recovery from controlled [35] A. Khodabakhshi, S. Madaeni, T. Xu, L. Wu, C. Wu, C. Li, W. Na, S. Zolanvari,
mixing salt and fresh water with a reverse electrodialysis system, Environ. Sci. A. Babayi, J. Ghasemi, Preparation, optimization and characterization of novel
Technol. 42 (2008) 5785–5790. ion exchange membranes by blending of chemically modified PVDF and SPPO,
[8] P. Dlugolecki, A. Gambier, K. Nijmeijer, M. Wessling, Practical potential of Sep. Purif. Technol. 90 (2012) 10–21.
reverse electrodialysis as process for sustainable energy generation, Environ. [36] G. Magnacca, G. Cerrato, C. Morterra, M. Signoretto, F. Somma, F. Pinna,
Sci. Technol. 43 (2009) 6888–6894. Structural and surface characterization of pure and sulfated iron oxides, Chem.
[9] N. Hasanabadi, S.R. Ghaffarian, M.M. Hasani-Sadrabadi, Magnetic field aligned Mater. 15 (2003) 675–687.
nanocomposite proton exchange membranes based on sulfonated poly (ether [37] L.L. Sun, S.L. Wang, W. Jin, H.Y. Hou, L.H. Jiang, G.Q. Sun, Nano-sized
sulfone) and Fe2O3 nanoparticles for direct methanol fuel cell application, Int. Fe2 O3 SO4 2  solid superacid composite Nafion (R) membranes for direct
J. Hydrogen Energy. 36 (2011) 15323–15332. methanol fuel cells, Int. J. Hydrogen Energy. 35 (2010) 12461–12468.
J.G. Hong, Y. Chen / Journal of Membrane Science 460 (2014) 139–147 147

[38] M.T. Colomer, K. Zenzinger, Mesoporous alpha-Fe2O3 membranes as proton [47] J.G. Hong, W. Zhang, J. Luo, Y. Chen, Modeling of power generation from the
conductors: Synthesis by microwave-assisted sol-gel route and effect of their mixing of simulated saline and freshwater with a reverse electrodialysis
textural characteristics on water uptake and proton conductivity, Microporous system: the effect of monovalent and multivalent ions, Appl. Energy 110
Mesoporous Mater. 161 (2012) 123–133. (2013) 244–251.
[39] B. Gu, J. Schmitt, Z. Chen, L. Liang, J.F. McCarthy, Adsorption and desorption of [48] M. Wijers, M. Jin, M. Wessling, H. Strathmann, Supported liquid membranes
natural organic matter on iron oxide: mechanisms and models, Environ. Sci. modification with sulphonated poly (ether ether ketone): permeability,
Technol. 28 (1994) 38–46. selectivity and stability, J. Membr. Sci. 147 (1998) 117–130.
[40] P. Sabbatini, F. Yrazu, F. Rossi, G. Thern, A. Marajofsky, M. Fidalgo de Cortalezzi, [49] T.W. Xu, Ion exchange membranes: State of their development and perspec-
Fabrication and characterization of iron oxide ceramic membranes for arsenic tive, J. Membr. Sci. 263 (2005) 1–29.
removal, Water Res. 44 (2010) 5702–5712. [50] R.K. Nagarale, G.S. Gohil, V.K. Shahi, R. Rangarajan, Organic-inorganic hybrid
[41] B.S. Karnik, S.H. Davies, M.J. Baumann, S.J. Masten, Fabrication of catalytic
membrane: thermally stable cation-exchange membrane prepared by the sol–
membranes for the treatment of drinking water using combined ozonation
gel method, Macromolecules 37 (2004) 10023–10030.
and ultrafiltration, Environ. Sci. Technol. 39 (2005) 7656–7661.
[51] Y.I. Park, Y. Piao, N. Lee, B. Yoo, B.H. Kim, S.H. Choi, T. Hyeon, Transformation of
[42] W. Honghai, L. Yiying, W. Jiayi, Z. Lixuan, Z. Dingcai, D. Juan, Surface
hydrophobic iron oxide nanoparticles to hydrophilic and biocompatible
adsorption of iron oxide minerals for phenol and dissolved organic matter,
Earth Sci. Front. 15 (2008) 133–141. maghemite nanocrystals for use as highly efficient MRI contrast agent, J.
[43] B.I. Harman, H. Koseoglu, N.O. Yigit, M. Beyhan, M. Kitis, The use of iron oxide- Mater. Chem. 21 (2011) 11472–11477.
coated ceramic membranes in removing natural organic matter and phenol [52] C. Klaysom, R. Marschall, S.H. Moon, B.P. Ladewig, G.Q.M. Lu, L.Z. Wang,
from waters, Desalination 261 (2010) 27–33. Preparation of porous composite ion-exchange membranes for desalination
[44] H. Strathmann, Ion-Exchange Membrane Separation Processes, Elsevier application, J. Mater. Chem. 21 (2011) 7401–7409.
Science Limited, Amsterdam, 2004. [53] P. Długoł˛ecki, K. Nymeijer, S. Metz, M. Wessling, Current status of ion
[45] R.K. Nagarale, G.S. Gohil, V.K. Shahi, Recent developments on ion-exchange exchange membranes for power generation from salinity gradients, J. Membr.
membranes and electro-membrane processes, Adv. Colloid Interface Sci. 119 Sci. 319 (2008) 214–222.
(2006) 97–130. [54] J.H. Choi, S.H. Moon, Pore size characterization of cation-exchange membranes
[46] E. Guler, Y.L. Zhang, M. Saakes, K. Nijmeijer, Tailor-made anion-exchange by chronopotentiometry using homologous amine ions, J. Membr. Sci. 191
membranes for salinity gradient power generation using reverse electrodia- (2001) 225–236.
lysis, Chemsuschem 5 (2012) 2262–2270.

You might also like