Você está na página 1de 18

Rev Chem Eng 2016; 32(5): 533550

Avner Ronen, Sharon L. Walker and David Jassby*

Electroconductive and electroresponsive


membranes for water treatment
DOI 10.1515/revce-2015-0060
Received October 12, 2015; accepted March 18, 2016; previously
published online May 11, 2016

Abstract: In populated, water-scarce regions, seawater


and wastewater are considered as potable water resources
that require extensive treatment before being suitable for
consumption. The separation of water from salt, organic,
and inorganic matter is most commonly done through
membrane separation processes. Because of permeate flux and concentration polarization, membranes are
prone to fouling, resulting in a decline in membrane performance and increased energy demands. As the physical
and chemical properties of commercially available membranes (polymeric and ceramic) are relatively static and
insensitive to changes in the environment, there is a need
for stimuli-reactive membranes with controlled, tunable
surface and transport properties to decrease fouling and
control membrane properties such as hydrophilicity and
permselectivity. In this review, we first describe the application of electricity-conducting and electricity-responsive
membranes (ERMs) for fouling mitigation. We discuss
their ability to reduce organic, inorganic, and biological
fouling by several mechanisms, including control over
the membranes surface morphology, electrostatic rejection, piezoelectric vibrations, electrochemical reactions,
and local pH changes. Next, we examine the use of ERMs
for permselectivity modification, which allows for the
optimization of rejection and control over ion transport
through the application of electrical potentials and the
use of electrostatically charged membrane surfaces. In
addition, electrochemical reactions coupled with membrane filtration are examined, including electro-oxidation
and electro-Fenton reactions, demonstrating the capability of ERMs to electro-oxidize organic contaminates
with high efficiency due to high surface area and reduced
mass diffusion limitations. When applicable, ERM applications are compared with commercial membranes in
terms of energy consumptions. We conclude with a brief
*Corresponding author: David Jassby, University of California,
Riverside, Bourns Hall A241, Riverside, CA 92521, USA,
e-mail: djassby@engr.ucr.edu
Avner Ronen and Sharon L. Walker: University of California,
Riverside, Bourns Hall A241, Riverside, CA 92521, USA

discussion regarding the future directions of ERMs and


provide examples of several applications such as pore size
and selectivity control, electrowettability, and capacitive
deionization. To provide the reader with the current state
of knowledge, the review focuses on research published
in the last 5 years.
Keywords: electro-Fenton; electro-oxidation; electroreactive conductive membranes; electrostatic repulsive force;
hydrogen peroxide; piezoelectric membranes; pore size;
wettability.

1 Introduction
1.1 M
 embrane-based water treatment
processes
Long-range predictions of water supply and demand
predict that more than 40% of the worlds population is
likely to be living under severe water stress by 2050, with
global water demand projected to increase by 55% between
2000 and 2050 (Marchal etal. 2012). In populated, waterscarce regions limited by natural water resources, seawater and wastewater are used as potable water resources
but require extensive treatment before being suitable for
drinking (Shannon et al. 2008). The separation of water
from salt and other contaminants is most commonly done
through membrane separation processes (van der Bruggen
etal. 2003). Membranes are semipermeable barriers that
provide selectivity in response to a physical or chemical
potential gradient (e.g. pressure, concentration gradients,
and electric fields) (Buonomenna 2013). Membranes can
be divided into dense materials (reverse osmosis [RO]
and nanofiltration [NF]) or porous materials (microfiltration [MF] and ultrafiltration [UF]) (Kennedy et al. 2008).
Although UF and MF are mainly used for the filtration of
submicron-sized particles and organic material, NF and
RO are used for the desalination and removal of small
charged organic molecules (van der Bruggen etal. 2003).
The fastest growing desalination technology is RO, where
water is transported through a dense membrane that is
impermeable to ions, in response to a pressure gradient.

Brought to you by | The Flinders University


Authenticated
Download Date | 11/14/16 8:33 AM

534A. Ronen etal.: Electroconductive responsive membranes


Because of its energy efficiency in comparison with other
desalination technologies, approximately half of worldwide installed desalination capacity is based on RO
technology (Liu et al. 2008, Elimelech and Phillip 2011).
As membrane processes have become more widespread,
research efforts drove improvements in salt rejection, high
permeate flux, and antifouling properties (Lee etal. 2011,
Lenntech 2013). Despite these improvements, membranes
are still prone to fouling, which results in a decline in membrane performance and increased salt passage through
the membrane (Goosen etal. 2005, Al-Amoudi and Lovitt
2007). Fouling is characterized based on its cause, including particulate fouling, inorganic fouling, organic fouling,
and biological fouling (Field 2010). P
articulate fouling
is a result of inorganic or organic particles or colloids
physically depositing on the membrane surface (Guo etal.
2012). Inorganic fouling occurs when dissolved components (e.g. iron, manganese, calcium) precipitate onto the
membrane surface (Shirazi etal. 2010). Dissolved organic
molecules (e.g. humic acid, hydrophilic and hydrophobic materials, and proteins) are responsible for organic
fouling (Guo et al. 2012). Biological fouling (also known
as biofouling) is considered the Achilles heel of the RO
membrane process, as the transport of organic matter,
nutrients, and bacteria toward the membrane surface
typically leads to bacterial adhesion and a buildup of a
biofilm layer (Flemming et al. 1997, Baker and Dudley
1998, Flemming 2002, Eshed etal. 2008). Many bacterial
species form biofilms, which are extremely difficult to
remove by conventional chemical and physical treatments
(Jin etal. 2006, Lee etal. 2007).
Organic and biological fouling inhibition in conventional membranes is typically based on the removal of
most of the suspended organic matter and bacteria before
reaching the dense membrane (i.e. RO/NF) by a pretreatment step that typically consists of a coarser filtration
process (i.e. micro/ultrafiltration and granular filters). In
addition, the feed water is chlorinated to eliminate live
bacteria after a dechlorinated step to prevent damage
to the membranes active layer. Inorganic fouling (i.e.
mineral scaling) is reduced by treating the feed water with
additional chemical (i.e. antiscalants and pH changes)
to eventually decrease the concentration of precipitating
ions (i.e. phosphate, magnesium, etc.) below critical level
(Katz and Dosoretz 2008, Prihasto etal. 2009).
The physical and chemical properties of most commercially available polymeric membranes are relatively
static and do not significantly respond to changes in the
environment (other that changes in surface charge in
response to pH). Recently, it was shown that RO polyamide
membranes can reversibly swell as a function of the feed

solutions pH, leading to changes in flux and fouling


tendency (Ying etal. 2014, 2015).
The separation mechanism responsible for the membranes performance is based on size exclusion, electrostatic charge, or differences in diffusion rates (or a
combination of all), with separation performance remaining virtually unchanged under normal operating conditions. In addition, commercial polymeric membranes
are prone to both organic and biological fouling and are
dependent on periodic cleaning (e.g. chlorination, backwashing, and air scouring) (Ang etal. 2006, Hijnen etal.
2012). Fouling prevention through the modification of
membrane surface properties is continuously explored,
although only limited results have been published for
long and continuous filtration periods (Asatekin et al.
2006, Yao etal. 2008, Zhang etal. 2009, Wei etal. 2010,
Feng etal. 2011, Qiu etal. 2011). All of the above information has driven the need for reactive membranes (RMs),
with controllable surface and transport properties. There
has been significant progress in RM research in the last
decade, demonstrating the development of membranes
that respond to changes in the membranes proximal
environment and external stimulation (e.g. temperature,
pH, ionic strength, light, pressure, and magnetic and
electric potentials) (Wandera et al. 2010, Yu et al. 2011,
Bhattacharyya etal. 2012, Nicoletta etal. 2012, Wang etal.
2012a,b, Frost and Ulbricht 2013, Gugliuzza 2013, Katsuno
etal. 2013, Li etal. 2013, Qiu etal. 2013, Yang etal. 2013,
Chen etal. 2014, He etal. 2014, Ma etal. 2014, Xiao etal.
2014). Currently, there are no commercially available RMs.

1.2 E
 lectricity-conducting and electricityresponsive membranes
Electricity-conducting and electricity-responsive membranes (ERMs) are of interest as electrical potentials and
fields can be easily and quickly applied to the membrane
surface, making their use as drop-in replacements to
existing membrane modules a plausible prospect. ERMs
are defined as membranes in which surface properties
(i.e. surface charge and potential, hydrophobicity, roughness, and porosity) and mass transfer to and from the
membrane (i.e. electroosmosis, electrostatic attraction/
repulsion) can be modified using an external electrical
field or potential. Electrically conducting membranes
have been used for several application such as sensors,
electronic devices, and biomimetic devices (Meng and Hu
2010, Otero etal. 2012, Ates 2013, Guo etal. 2013). When
using membranes based on polymers for water treatment,
the electrically conducting polymers must be stable under
Brought to you by | The Flinders University
Authenticated
Download Date | 11/14/16 8:33 AM

A. Ronen etal.: Electroconductive responsive membranes535

aqueous and harsh conditions such as high pressures and


exposure to chemicals. Therefore, only a limited number
of conducting polymers have been used for the fabrication
of ERMs (e.g. polypyrrole and polyaniline). Most ERMs are
made as composite membranes through the deposition
of conducting nanomaterials on a porous support (i.e.
carbon nanotubes [CNTs] [Vecitis etal. 2011a,b, de Lannoy
etal. 2012, 2013, Dudchenko etal. 2014, Gao etal. 2015] and
graphene nanoparticles [GNPs] [Li etal. 2014, Hegab and
Zou 2015]). Additional methods include casting of a polymeric film containing nanoparticles, synthesizing a polymeric film over a conductive mesh, dual layer deposition
of CNT polymer, platinized titanium electrodes embedded
on the membrane surface, chemical polymerization of
polypyrrole on a polyester filter, and vapor phase polymerized pyrrole (Liu etal. 2012, Liu etal. 2013c,d,e, Holder
et al. 2013, Zhang and Vecitis 2014, Huang et al. 2015,
Wang etal. 2015a,b, Xu etal. 2015). In addition, ceramic
membranes are also used as ERMs where the membranes
are made from inorganic conducting material (i.e. Ti4O7,
lead zirconate titanate, and more) (Zaky and Chaplin 2013,
2014, Krinks etal. 2015).
As fouling is a critical issue in membrane separation processes and a crucial factor in commercial applications, the majority of the literature is related to using
ERMs for fouling mitigation. In addition, research has
been done on the controlled selectivity, flux, and electrochemical oxidation of organic contaminants and bacteria. In this review, we cite recent examples of ERMs as
well as provide detailed examples of ERMs developed
for different separation applications and purposes. To
provide the reader with the current state of knowledge,
this review focuses on research done in the last 5years
although the overall background includes previous
research as well.

2 Fouling mitigation
Fouling is the main reason for permeate flux decline and
increased energy consumption in membrane systems
(Goosen et al. 2005). Multiple examples of the use of
ERMs for fouling prevention have been published in the
last 5years, with mechanisms such as electrostatic repulsion, hydrogen peroxide (HP) production, electrochemical oxidation, morphological changes to the surface, and
piezoelectric vibrations. In the following sections, we
describe recent efforts in fouling mitigation using ERMs
by subdividing these efforts based on their antifouling
mechanism.

2.1 Membrane fouling


2.1.1 Electrostatic repulsion
Most foulants are negatively charged (Hong and Elimelech
1997; Park etal. 2005). Negatively (cathodic) charged conducting membranes induce electrostatic repulsive forces
between the charged membrane surface and these foulants, which mitigates fouling and increases permeate
fluxes. The Derjaguin, Landau, Verwey, and Overbeek
(DLVO) theory allows for the calculation of interaction
forces between the charged foulants and the membrane
surface (Brant and Childress 2002; Hermansson 1999;
Wang etal. 2005); however, DLVO is not accurate in predicting forces when large electrical potentials are applied
to an ERM surface (Dudchenko et al. 2014). High repulsive forces of up to several nanonewtons at a distance of
510 nm from the charged membrane surface has been
reported when applying -3.4 V versus a Ag/AgCl reference
electrode (Dudchenko etal. 2014). Huang etal. (2015) prepared a conducting polyvinylidene fluoride (PVDF) MF
membrane by casting a homogeneous PVDF solution on
a stainless steel mesh (pore size 96 m, thickness 43 m)
assembled on a polyester nonwoven fabric. The composite
membrane was formed after an immersion precipitation
process in a nonsolvent bath. The authors report high pure
water flux and low electrical resistance (66 L/m2/h [LMH]
and ~200 ). Batch filtration tests using bovine serum
albumin, sodium alginate, humic acid, and silicon dioxide
as model foulants were used to test the membranes antifouling properties. Experiments were performed by applying an electrical field of 2 V cm-1 to the system with the
membrane used as a cathode. Results indicate that the
application of an electrical potential to the membrane
decreased fouling rates for all model foulants tested. The
enhanced antifouling ability was explained by a combination of electrostatic repulsive forces between the foulants and the charged membrane, as well as by organic
foulant oxidation by electrochemically generated HP,
which increased their charge and enhanced electrostatic
repulsion.
Dudchenko et al. (2014) used a pressure deposition
process to create robust and electrically conducting thin
films made of cross-linked poly(vinyl alcohol) (PVA) and
carboxylated multiwalled CNTs (PVA-CNT-COOH) on a
polysulfone UF support. The separation properties of the
PVA-CNT-COOH-modified membranes were similar to the
UF support, and the membrane exhibited high electrical conductivity (2500 S/m). The influence of an applied
potential on organic fouling was examined in a crossflow filtration cell (Figure 1A) using high concentrations
Brought to you by | The Flinders University
Authenticated
Download Date | 11/14/16 8:33 AM

536A. Ronen etal.: Electroconductive responsive membranes

B 24

Conductive thin film

22
Pressure (psi)

Cross flow

Membrane surface
Permeate flow

0.0 VDC PS-35


0.0 VDC PVA-CNT
-3.0 VDC PVA-CNT
-5.0 VDC PVA-CNT

20

1.5
-3.412 V
-1.519 V
-0.1 V

1.0
Force (nN)

Counter electrode
Electro force

18
16

0.5
0.0

14
12
0

20

40
60
Time (min)

80

100

-0.5

15
10
Distances (nm)

20

Figure 1:(A) Electrofiltration with conducive membrane/thin film setup. (B) Fouling of 5 g/l AA on PS-35 and PVA-CNT membranes. (C)
Overall force acting on an AA particle in a DIW solution containing 5 g/l AA at various distances from the surface and applied potentials at
ionic strength of 25.36 mm. (Reproduced from Dudchenko etal., Copyright 2014, with permission from Elsevier.)

(30005000 ppm) of alginic acid. It was demonstrated


that after 100min of operation when the membrane was
used as a cathode (applying -3.4V vs. Ag/AgCl reference
electrode), the change in operating pressure was reduced
by 51% compared with the control membrane without
voltage (Figure 1B). Fouling mitigation was explained
using a modified Poisson-Boltzmann equation and a
DLVO-type theory, revealing that electrostatic interactions
produced significant repulsive forces between the membrane surface and the charged organic foulant (Figure 1C).
Other recent research on the use of enhanced electrostatic repulsive forces to reduce membrane fouling
includes work by Zhang and Vecitis (2014), who used
electrically conducting and cathodically charged membranes fabricated from a CNT-PVDF composite to prevent
fouling by natural organic matter, and work by Wang etal.
(2015a,b), who used electrically conducting cathodically
charged membranes fabricated from a duel layer CNTPVDF to prevent fouling by several organic compounds
(humic acid and sodium alginate) and inorganic compounds. In addition, fouling prevention in membrane bioreactors was explored by Liu etal. (2012, 2013a), who used
an electrically conducting membrane made of polypyrrole
while applying low electrical potentials (2 V), demonstrating reduced fouling and flux decrease compared with the
noncharged membranes.

2.1.2 Electroresponsive surface morphology


Changes in membrane surface morphology were used to
prevent organic fouling by cross-linking and de-cross-linking
the surface of a polymeric membrane, with the antifouling
mechanism attributed to changes in surface hydrophobicity and membrane pore size (Chuo et al. 2013). Here,
poly(glycidylmethacrylate) (PGMA) chains with a ferrocene
(Fc) end group were grafted on polytetrafluoroethylene

(PTFE) membranes (PTFE-g-PGMA-Fc). The surface layer


was cross-linked with a -cyclodextrin (-CD) compound
by means of the Fc/-CD complexation reaction (PTFEg-PGMA-Fc/CD), with the cross-linking and de-cross-linking
of the surface layer being controlled through applying an
external electric field of 9 V using two aluminum sheets
on either side of the membrane as electrodes. Structural
changes arising from the cross-linking and de-cross-linking of the membrane result in a change of surface hydrophilicity. The water contact angle of the non-cross-linked
PTFE-g-PGMA-Fc membrane is approximately 108, which
decreases to approximately 96 after cross-linking with
-CD. De-cross-linking the surface recovered the hydrophobicity of the sample, increasing the water contact angle to
113. In addition, the PTFE-g-PGMA-Fc membranes exhibited morphological changes in response to the electric field,
with pore sizes and surface areas decreasing from 450nm
and 2.5m2g-1 to 390nm and 2.3 m2 g-1 when the surface was
cross-linked in response to the electric field. Organic fouling
mitigation experiments were done using a model protein
(fibrinogen) under static conditions. Electrical treatment of
the fouled membrane resulted in protein detachment from
the membrane surface, with a 3-h exposure leading to only
45% of the protein remaining on the membrane compared
with the control. Because hydrophobic membranes are
more prone to fouling in comparison with hydrophilic membranes, as most foulants tend to be hydrophobic (Fan etal.
2001), the de-cross-linking reaction of the PTFE- g-PGMAFc/CD membrane is assumed to provide the driving force for
the protein release behavior due to the change in hydrophilicity and surface morphology.

2.2 Piezoelectric vibration


Control of permeate flux and fouling using a piezoelectric
ERMs was reported by Coster et al. (2011) and Darestani
Brought to you by | The Flinders University
Authenticated
Download Date | 11/14/16 8:33 AM

A. Ronen etal.: Electroconductive responsive membranes537

particles of calcium carbonate (CaCO3), sodium alginate


(SA), and a mixture of both materials. A larger decrease in
flux was noticed when the membranes were fouled with
sodium alginate due to internal pore blocking as well as the
formation of a cake layer on the surface. Flux decline from
CaCO3 fouling was significantly lower and is assumed to be
mainly a result of cake formation, as the particles are larger
than the membrane pore size. When a mixture of both SA
and CaCO3 was tested, flux decline was closer to that of
the pure sodium alginate solution, indicating that internal
pore blocking dominates the fouling process. When an AC
bias was applied across the membrane, the flux increased
regardless of the foulant. Thus, the vibration of the membrane was able to remove both internal blocking and cake
formation from the membrane (Darestani etal. 2013a,b).
Another unique piezoelectric material was investigated by Krinks etal. (2015), who fabricated piezoelectric
ceramic MF membranes. The membranes were synthesized using a commercially available lead zirconate titanate powder consisting of 50150-m spray-dried granules
composed of 0.5-m primary particles. Using dry pressing
(100 MPa) and sintering (950C for 10 h), circular membrane disks with 20 mm diameter and 1.3 mm thickness
were synthesized. The lead zirconate titanate membrane
disks contained ferroelectric domains with a random orientation. Selected sintered disks were poled by a strong
direct electric field so that the domain orientations
became aligned, with a net polarization vertical to the
disk surface. Experiments were performed in a dead-end
filtration cell, where a stainless steel bolt, installed at a
height of 1mm above the feed side of the membrane, acted

etal. (2013a,b). Both groups used the ability to alter PVDF


membranes by applying a strong electrical field, high
temperatures, and pressure to pole the membranes, transforming PVDF from the to phase, which has piezoelectric properties (Martins etal. 2014).
PVDF membranes were made piezoelectric by electrically poling a commercial MF membrane. Electrical poling
is a process used to change the semicrystalline structure
of PVDF to the phase, which is responsible for piezoelectric properties. It involves the application of an intense
electric field while the polymer sample is held at an elevated temperature (just below the melting temperature)
and then cooling it while still in the intense electric field.
Poling was performed using a 2-kV potential (corresponding to the field strength of 16.3106 V/m) applied across
the membrane at temperatures close to the melting temperatures of the polymer (90C10C). The poling process
converted the PVDF membranes to piezoelectric -PVDF
(Figure 2A) (Darestani etal. 2013a,b), which vibrates when
an AC signal is applied. Filtration experiments were performed over a wide range of frequencies (5 Hz50 kHz),
with maximum flux observed at 500 Hz; beyond 500
Hz, the flux steadily declined with increasing frequency
(Darestani et al. 2013a,b). In addition to frequency, the
influence of amplitude was investigated. Permeate flux
was tested as a function of time for various AC amplitudes
(010 V at 500 Hz), during the filtration of a solution of 1%
w/w polyethylene glycol (Coster etal. 2011, Darestani etal.
2013a,b). Results demonstrate a strong flux increase with
increasing amplitude of the applied AC signal (Figure 2B).
Additional fouling experiments were done using 10-m

Figure 2:(A) SEM micrographs of cross sections of a PVDF membranes before electrical poling and after poling at different magnifications
(the electrical field direction was the same and is shown in the diagram). (B) Effect of cross flow on normalized flux of a piezoelectric membrane excited by a 10-V AC signal at 500 Hz, an unexcited poled membrane, and an unpoled membrane after 35min filtration of a solution of
1 g/l PEG at 50 kPa. (Reproduced from Darestani etal., Copyright 2013, with permission from Elsevier.)

Brought to you by | The Flinders University


Authenticated
Download Date | 11/14/16 8:33 AM

538A. Ronen etal.: Electroconductive responsive membranes


as one electrode, and a stainless steel mesh electrode was
in direct contact with the permeate side of the membrane
(second electrode). To prevent fouling, the two electrodes
on the membrane cell were connected to an RF power
generator tuned to 72.6 kHz at an amplitude of 100V. The
electrical field generated vibrations of the piezoelectric
membrane material, which generated ultrasonic waves
and vibrations at the membrane surface. Voltage was
applied across the membrane in a cycle of 30 s on, followed by 270s off. The membranes were characterized as
MF with a 365-nm average pore size and permeability of
14 LMH/psi using deionized water. Fouling experiments
were done using latex particles (0.210 mg/L) in an electrolyte solution. The flux decreased by more than 20%
within 3 h of filtration when no alternating voltage was
applied and is assumed to be a result of a cake layer, which
was developed on membrane surface. By contrast, no
flux decline was observed when the intermittent voltage
was applied, indicating an intrinsic defouling (fouling
removal) functionality of the membrane. The physical
mechanism for defouling is assumed to be due to surface
vibration, cavitation, or a combination of both.

2.3 Mineral scaling


Duan et al. (2014) investigated mineral scaling prevention and removal using an electrically conducting
CNT-polyamide (PA) composite RO membrane (Figure3A).
A range of anodic potentials (1.52.5 V) were applied to
the membrane while filtering model scaling solutions
with high saturation indices. CaCO3 scale was efficiently
removed from the membrane surface through the intermittent application of a 2.5-V anodic potential to the

membrane surface. The applied potential led to proton


formation through water splitting, and the locally elevated proton concentrations led to the dissolution of the
deposited CaCO3 crystals (Figure 3B). In addition, CaSO4
scale formation was significantly inhibited through the
continuous application of 1.5 V to the membrane surface;
this phenomena was explained by the formation of a
layer of counterions along the membrane surface, which
pushed CaSO4 crystal formation away from the membrane
surface, transforming the nucleation process from heterogeneous to homogeneous, allowing the formed crystals to
be carried away by the cross-flow stream (Figure 3C).
Additional research regarding the prevention of
mineral scaling was done by Hashaikeh etal. (2014), who
used a vacuum to deposit multiwall CNTs on a membrane
surface, with the membranes resulting electrical resistance reported as 500 cm-1. The antiscaling efficiency of
the membrane was demonstrated using a high concentration of CaCO3 (100 ppm) in the feed solution. Control
membranes without applied electrical potential showed
membrane fouling and permeate flux decrease (80%
reduction in 170 min), whereas periodic electrolysis using
2V for 23min prevented most membrane scaling (30%
reduction in 170 min). The mechanism for scaling prevention was explained by the evolution of microgas bubbles
at the membrane surface (hydrogen and chlorine), which
removed the foulant away from the membrane surface.

2.4 B
 iological fouling and pathogen
inactivation
Biofouling is defined as the undesirable accumulation
of bacterial communities at the interface of the solid

Figure 3:(A) SEM image of surface of PA-CNT RO membrane. (B) (a) Electrically conducting RO membrane surface area after the application
of 2.5 V; the white area, corresponding to areas still covered by CaCO3 is greatly reduced. (b) Membrane area fully covered by CaCO3 scale. (c)
SEM of membrane surface after the application of 2.5 V displaying few CaCO3 crystals. (d) Thick CaCO3 deposits on fouled membrane surface
(no voltage applied); (C) As the applied surface potential increases, the region near the membrane where only SO42- ions exist expands out
into the solution, which pushes the nucleation zone away from the membrane surface. (Reproduced from Duan etal., Copyright 2014, with
permission from the Royal Society of Chemistry publishing.)

Brought to you by | The Flinders University


Authenticated
Download Date | 11/14/16 8:33 AM

A. Ronen etal.: Electroconductive responsive membranes539

(i.e.membrane) and water phase. Fouling occurs by deposition, attachment and growth of bacterial communities
on the membranes (Guo etal. 2012). Although biofouling
is usually referred to in regard to bacteria, it is not limited
to it and could also develop from the deposition of yeast
and marine organisms (Callow and Callow 2002, Libana
etal. 2015).
de Lannoy etal. (2013) prepared a CNT-PA (polyamide)
composite material through an interfacial polymerization process that covalently bound the CNTs to the (PA)
through ester bonds. The membrane exhibited high electrical conductivity (~400 S/m), good monovalent ion
rejection (>95%), and water permeability compared with
commercial NF membranes. The authors demonstrated
that the application of an alternating block current (1.5 V)
to the surface of an electrically conducting NF membrane
prevented the formation of microbial biofilms (Figure4A).
Biofouling mitigation experiments were conducted using a
feed flow containing 10% Luria broth and high concentrations (108 CFU/ml) of Pseudomonas aeruginosa bacteria.
Biofilm-induced nonreversible flux decline was observed
in all control experiments. By contrast, when a potential
was applied to the surface, flux was fully recoverable following a short rinse with the feed solution and no added
cleaning agents (Figure 4B).
Singh et al. (2015) prepared a CNT/PVDF composite
membrane by using vacuum filtration, where the deposition of CNTs on a PVDF substrate was followed by a curing
treatment at temperatures higher than the melting point
of PVDF. These composite membranes had an electric conductivity of 0.41 S/m and functioned as MF membranes
(pore size ranging from 15 to 415 nm). Biofouling mitigation was evaluated in a dead end filtration cell using yeast
as a model biofouling agent while the membrane was used

B
Normalized flux

2
1.1
1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0

as a cathode (2 V). Results indicate that the permeate flux


was restored to a higher value than the control (without
potential) but lower than the initial value. Biofouling
prevention was assumed to be related to water electrolysis and the formation of microbubbles on the membrane
surface, physically removing the deposited organisms.
Ronen et al. (2015) examined the effect of electric
potentials on bacterial attachment to electrically conducting CNT-PVA cross-linked UF membranes (electrical
conductivity of 2500 S/m). The membranes were tested as
both anodes and cathodes. Using a direct imaging system
coupled to fluorescent microscopy, it was observed that
microbial viability and cell membrane integrity were
highly disturbed by the application of an electrical potential to the membrane surface. Results revealed a connection between the applied potential and the bacterial
attachment even when a relatively low potential was
applied, with 90% of the bacteria detaching from the
membrane when a 1.5-V potential was applied. Low concentrations of HP, electrochemically generated on the
membrane surface, were determined to be responsible for
the observed microbial detachment and reduced viability.
Using conducting membranes as anodes is a promising field that couples membrane separation with an oxidative degradation process. In particular, the high surface
area and electrical conductivity of CNTs makes them
ideally suited to form electroactive, porous membranes.
Electrochemical oxidation is assumed to involve both
direct and indirect oxidation processes.
Vecitis etal. (2011a,b) used multiwalled CNT vacuum
filtered onto a 5-m PTFE membrane to simultaneously
remove and inactivate viral and bacterial pathogens. The
conductive membrane was used as an anode with relatively low potentials (13 V) applied for a short duration

Applied voltage to ECPNC membrane


Applied voltage to non-conducting PNC membrane
No voltage Applied to ECPNC membrane

50

Time (h)

100

150

Figure 4:(A) TEM image of the composite PA-CNT membrane. Three layers are evident: the polyether sulfone support, the deposited CNT
layer, and the PA-CNT composite layer. (B) Control experiments without applied voltage and with highly resistive membranes with applied
voltage suffered from irrecoverable biofouling. CNT membranes, with applied voltage, demonstrated much longer resistance to flux decline
and flux was completely recoverable with 1min of cross-flow flushing. Red circles represent membrane flushing points. (Reproduced from
de Lannoy etal., Copyright 2013, with permission from American Chemical Society.)

Brought to you by | The Flinders University


Authenticated
Download Date | 11/14/16 8:33 AM

540A. Ronen etal.: Electroconductive responsive membranes


(30 s). Results indicate that when applying potentials
of 2 and 3 V, no culturable bacteria were detected in the
effluent (Figure 5B). The membrane was also washed to
remove adsorbed viruses revealing a significant reduction
in culturable viruses adsorbed to the membrane after the
application of the electrical potentials (Figure 5A). Electron and fluorescence microscopy images suggest that
the application of 1 and 2 V interrupts specific, localized
regions of the cell membrane but does not disrupt the
macroscopic cell membrane structure. At higher potentials (3 V), cells were significantly degraded and had lost
all cell membrane integrity. Bacterial and viral inactivation is attributed to the direct oxidation of the pathogens
in contact with the CNT anode and the indirect oxidation
of pathogens via anodic production of oxidants (Cl2-, OH)
(Figure 5C).
Apart from direct and indirect oxidation, another
mechanism should be taken under consideration, as
described by Guo etal. (2016) Here, the authors demonstrated that by using tubular asymmetric TiO2-based UF
membranes and by applying anodic potentials (up to
5.8 V), the local pH near the membranes was reduced
as a function of the applied electrical potential, eventually reaching an acidic pH (about two pH units reduced).
This pH reduction could influence the activity of bacteria attached to the membranes surface. Therefore, both
direct and indirect oxidation coupled with a pH change
could be influencing bacterial cell viability under anodic
potentials.
Liu etal. (2013a) fabricated a conducting nanosponge
filtration device using a commercial polyurethane sponge
modified by CNTs and silver nanowires (AgNWs). The
modification was done by dipping the polyurethane into
a CNT ink and AgNWs solution several times followed by

7
6

B
No viruses in effluent (filtrate)

5
4
3
2
1
0

0V

2V
Applied potential

3V

0s
10 s
30 s

100
Loss of viability (%)

Log MS2 removal

drying at ~90C. The unique properties of the polyurethane sponge are due to the conductive CNTs leading to a
low electrical resistance of 1 /cm and from the presence
of AgNWs, which produce a large number of nanoscale
sharp tips distributed on the sponges surface. The
pore size of the sponge was characterized as very large
(~500m). The ability of the sponge filter to remove bacteria and viruses was evaluated under flowing conditions
(15,000 LMH) by applying an external voltage (020V) to
two parallel sponge electrodes contained inside an in-line
filter holder. Disinfection performance was evaluated
using four different model bacteria representing Gramnegative and Gram-positive species. Results show that
bacterial inactivation was enhanced with increasing external voltage for all four bacteria, indicating that a stronger
electric potential results in more efficient disinfection. For
voltages above 10 V, all four bacterial species were inactivated at over 6 log (99.9999%) removal, corresponding
to no culturable bacteria detected in the effluent, with a
retention time of only 1 s. In addition, virus inactivation
was evaluated using a model virus (MS2) and a bacteriophage (F+), which is related to enteric pathogenic viruses
in groundwater and raw wastewater. Results demonstrate
that inactivation increased with greater external voltage,
with 99.4% inactivation at a voltage of 20 V. The bacterial
disinfection mechanism was claimed to be based on electroporation of the cell membrane. Reversible electroporation was detected when lower potentials were applied. The
disinfection process was enhanced by the release silver
ions from the AgNWs, which could enter the bacterial
cell after the cell membrane was damaged, reacting with
inner cellular components to disrupt cell function. Lower
applied potential demonstrated lower removal rates, with
5V being the lowest voltage where removal was observed.

Direct

Indirect

80
Ox

60

Ox

40
20
0

2
Potential (V)

e-

e-

Figure 5:(A) Electrochemical MS2 removal vs. potential, Log MS2 removal as a function of applied potential during filtration. Influent was
10ml of 10 mm NaCl (pH 5.7) and 106 viruses ml-1. (B) Electrochemical loss of Escherichia coli viability versus potential and time. E. coli
suspension (107 cells, [NaCl]=10 mm, pH 5.7). (C) Electrolytic inactivation mechanisms. (Left) Depiction of direct electrochemical oxidation
of bacteria adhered to CNT surface and of indirect (right) electrochemical production of aqueous oxidant that subsequently inactivates the
bacteria in solution. (Reproduced from Vecitis etal., Copyright 2011, with permission from American Chemical Society.)

Brought to you by | The Flinders University


Authenticated
Download Date | 11/14/16 8:33 AM

A. Ronen etal.: Electroconductive responsive membranes541

It should be noted that in addition to direct and indirect oxidation, applying anodic potential on bacteria
colonies and biofilms has been shown to induce oxidative
stress because of the presence of hydroxyl radicals and
HP. Although these could lead to bacterial detachment,
damage to the cell membrane and increased susceptibility to antibiotics (Baatout etal. 2006, Barraud etal. 2006,
Narendra and Ramesh 2012, Munna etal. 2013), research
shows that they also function as a signaling molecule (p
et al. 2012) and could increase genetic variants, eventually increasing the diversity of bacteria, which may lead to
adaptability in biofilm communities and a more sustainable biofilm (Boles and Singh 2008).

2.5 F ouling inhibition in conventional


membranes vs. ERMs
Fouling inhibition in conventional membranes is typically
based on the removal of most of the suspended organic
matter and bacteria before reaching the RO/NF membrane
by a pretreatment step. This step consists of a coarser filtration process (i.e. micro/ultrafiltration, granular filters),
and the used filters are frequently replaced. Before reaching the RO/NF membranes, the feed water is also chlorinated to eliminate all live bacteria following a dechlorinated
step to prevent damage to the membranes active layer. For
the prevention of inorganic mineral scaling, feed water is
treated with additional chemical (i.e. antiscalants and
pH changes), eventually decreasing the concentration of
harmful ions (i.e. phosphate, magnesium, etc.) below critical level (Katz and Dosoretz 2008, Prihasto etal. 2009). In
addition, multiple surface modifications for conventional
membranes have been suggested, specifically for biofouling prevention. These include methods that decrease
initial bacterial attachment and/or embed antimicrobial
chemical and nanoparticles in the membranes active
surface. All modification typically require manipulating
the membranes active layer, influencing transport and
selectivity properties and are temporary due to the release
of the antibacterial chemical to the surrounding environment and the attachment of dead bacteria or organic
matter to the surface (Yao et al. 2008, Yang et al. 2009,
Liu et al. 2010, 2013f, Malaisamy et al. 2010, Qiu et al.
2011, Saeki etal. 2013). As the ERM field is relatively new,
no long-term results have been published yet, and most
experiments described were done at laboratory scale only.
It could be assumed that high-density ERMs (i.e. RO and
NF) are also in need for pretreatment processes, but the
membranes used for it (i.e. UF and MF ERM) will need less
frequent replacements as they can decrease the organic

and biological fouling buildup. In addition, bacteria that


will reach the membranes surface are at low concentrations and the ERM have been shown to be effective against
higher bacterial concentrations and harsh conditions.
Inorganic scaling experiments have proven to be efficient
at relatively high concentration of ions and, therefore, are
expected to reduce the amount of antiscalants and chemicals used.
In terms of energy and economic analysis, pretreatment is considered an extensive part of RO/NF separation
processes; therefore, significantly reducing the fouling
of UF and MF pretreatment processes is expected to lead
to simpler operational conditions and lower operating
costs. As for the ERMs, the fabrication of electrical conductivity NF and RO membranes would add ~2.5 $ m-2 in
material costs, representing only a 1.5% increase in the
total price of traditional membrane modules (de Lannoy
etal. 2013). The power consumption for ERMs is low, with
a maximum of 10 W m-2 for an applied potential of 5 V,
which is <10-4$m-2 (Dudchenko etal. 2014). In addition,
cost efficiency is also affected by the decrease in pumping
costs (i.e. backwash, increased pressure needed to maintain a constant flux) and cleaning processes, which are
reduced due to the decrease in fouling. This factor is not
calculated here as it is varies according to pump specification and is specific to each membrane separation process.

3 Permselectivity
In conventional polymeric membranes, selectivity is
mainly a function of pore sizes, although modified
membranes (with nanoparticles or zwitterionic groups
attached to their surface) have shown tunable permselectivity. However, permselectivity is limited in terms
of control, which relies on changes in the surrounding
pH (Chan etal. 2014, Geise etal. 2014, Ji etal. 2016). By
contrast, ERMs can be tailored to optimize rejection and
to control ion selectivity (permselectivity) by applying
electrical potential and electrostatically charging the
membrane surface (Kuo et al. 2001). Under these conditions, counter ions are attracted to the membrane surface,
whereas co-ions are rejected.
Weidlich and Mangold (2011) coated PVDF and polyethersulfone membranes with an electrically conductive
polypyrrole (PPy) film (Figure 6A) using a diffusion-limited
oxidation process. Membrane properties (film thickness
and permeate flux) were correlated to the polymerization time and the counter ions incorporated into the PPy
coating during polymerization. Permselectivity experiments were conducted by applying a potential (800 mV)
Brought to you by | The Flinders University
Authenticated
Download Date | 11/14/16 8:33 AM

542A. Ronen etal.: Electroconductive responsive membranes

Figure 6:(A) SEM cross section of a PPy coated MF membrane (polyethersulfone, 0.22 m) after 15min polymerization. (B) Ca2+ concentration in permeate passed a MF membrane (PVDF, 0.22 m) coated with PPy with PSS counter ions in dependence on cathodic (-800mV or
-0.7mA cm-2) or anodic (+800mV or +0.8mA cm-2) potentiostatic or galvanostatic polarization. (Reproduced from Weidlich etal., Copyright
2011, with permission from Elsevier.)

to the membrane while filtering a solution containing


0.005m CaCl2, with a Pt mesh used as counter electrode
(Weidlich and Mangold 2011). Cathodic (negative) polarization lad to a decrease in the Ca2+ permeate concentrations because of the adsorption of Ca2+ cations on the PPy
coating. When this step was followed with anodic (positive) polarization, the Ca2+ concentration in the permeate
increased because of the desorption of Ca2+ cations from
the membrane (Figure 6B). It should be noted that PPy
is sensitive to anodic oxidation, with repeated exposure
leading to a decrease in its electrical conductivity. Therefore, the applicability of PPy-based membranes in water
treatment application is limited (Pyo etal. 1994).
Chen and Chuang (2013) fabricated a conductive
composite nanoporous CNT-chitosan-gelatin membrane to demonstrate charge-selective transport of
charged molecules for selective separation of proteins.
Three molecules were tested: carboxyl-fluorescein (CF)
a small fluorescent dye (MW=376.32 g/mol), tumor
necrosis factor (TNF-) a cell-signaling protein
(MW=17,400 g/mol), and substance P (SP) a neuropeptide (MW=1347.63 g/mol).
The membrane was made porous and conductive
(2.540.29 k cm-1) via a gelatin nanoparticle leaching
technique and the addition of CNT. Average pore size
was found to be 40100 nm, which is in the range of a UF
membrane (UF has a typical molecular weight cutoff in the
range of 50,000 g/mole, significantly larger than the tested
charged molecules). (Dudchenko etal. 2014). Experiments
were conducted using a buffer solution (pH 7.4) with the
charge of CF, SP, and TNF- at this pH being -3, +1.76, and
-0.89 mV, respectively. Negatively charged molecules (CF
and TNF) diffused more rapidly through the membrane
at positive potentials (5 V) than at negative potentials. By

contrast, SP cations preferentially passed the membrane


when a negative potential was applied. In addition, it was
demonstrated that molecule size also plays a role, with
the membrane being 30 times more selective towards CF
than SP (anodic conditions), while only 2.5 times more
selective for SP than CF under cathodic conditions.
Guo etal. (2016) showed that by using tubular asymmetric TiO2-based UF membranes and applying high
cathodic potentials (up to -10 V), nitrate and perchlorate
(both negatively charged) were rejected from the charged
membrane and did not pass into the permeate. Here, ion
removal was assumed to be a result of electrostatic repulsion forces between the ions and the charged membrane
surface. No evidence for electrochemical reduction was
found in the permeate and concentrate streams. Removal
rates were found to be affected by the applied potential
and the permeate flux, with higher rejection evident with
increasing potential and decreasing flux. When applying
10 V, removal rates of 68.02%0.60% and 95.10%1.08%
were detected for nitrate and perchlorate, respectively.
The difference in removal rates was assumed to be due to
different permeate fluxes, which counteract the electrostatic rejection.

3.1 P
 ermselectivity in conventional
membranes vs. ERMs
Permselectivity is considered when charged membranes
allow specific species to be transported while restricting
the passage of other species. For water treatment aspects,
ion exchange membranes and charged NF membranes are
considered as permselective as they have fixed charges
on the membrane surface (Xu 2005, Nagarale etal. 2006,
Brought to you by | The Flinders University
Authenticated
Download Date | 11/14/16 8:33 AM

A. Ronen etal.: Electroconductive responsive membranes543

Yaroshchuk et al. 2009, Geise et al. 2014). Ion exchange


and NF membranes are charged with low potential only
(several millivolts) (Yaroshchuk et al. 2009) whereas
ERMs can be charged to a potential of several volts. As
the applied potentials and currents are relatively low, the
overall energy consumption is not high (similar to values
described for fouling prevention in the previous section).

4 E
 lectrochemical reactions
coupled with membrane filtration
Electro-oxidation and electro-Fenton (E-Fenton) are processes that can be used for the oxidation and degradation
of aqueous contaminants (e.g. pesticides, pharmaceuticals, and personal care products). Using ERMs allows
for these degradation reactions to occur during filtration,
with the benefit of a large membrane active surface area
and high mass transport rates (Gao etal. 2015). E-Fenton
reactions are initiated by the in situ electrochemical production of HP via the cathodic reduction of O2, followed
by the reaction of HP with iron (Fe2+) to generate hydroxyl
radicals (Ayadi et al. 2013, Zhao et al. 2013a, Gao et al.
2015). There are several types of E-Fenton reactions, where
the most promising include ferric ion reduction to ferrous
ion at the cathode and the production of hydroxyl radicals
in an electrolytic process. However, ferric ion regeneration
is slow even at an optimal current density, and the production of HP is slow because oxygen has low solubility in
water (Babuponnusami and Muthukumar 2012). Electrooxidation (Vecitis etal. 2011a,b, Liu etal. 2014) is based on
an anodic reaction on the membrane surface and can be
attributed to two mechanisms: (1) direct oxidation driven
by contact with the conductive membrane and (2) indirect
oxidation by anodic production of an aqueous oxidant
(e.g. chlorine and hydroxyl radicals).

4.1 Electro-Fenton
Gao et al. (2015) used a multiwall CNT membrane stack
for sequential electro-Fenton reactions. In the CNT membrane stack (overall thickness ~200 m), the feed initially
goes through a cathodic CNT network that reduces O2 to
HP. Next, the feed passes through a CNT-COOFe2+ cathode,
used to chemically reduce HP to hydroxyl radicals and
to regenerate Fe2+. In the final stage, the feed is passed
through a CNT filter anode for oxidation of any remaining intermediates. Results indicate high organic removal
at low cell potentials (<3.0 V) because of the thickness of

the stack, which allowed for a long contact time, and convective mass transport. In addition, as Fe was bound to
the electrode, the E-Fenton reaction was able to operate at
neutral pH and recycle the Fe2+ by in situ electrogeneration.
Zhao et al. (2013a) modified a polyester membrane
by impregnating it with graphene and by coating it with
polypyrrole using a vapor phase polymerization method.
After modification, the composite membrane had an electrical resistance of 0.7 k/cm. The membrane was used as
a cathode (-1 V DC), and a stainless iron mesh was used
as an anode. The modified membrane was evaluated for
the removal of methylene blue (MB) as a model compound
through an E-Fenton reaction under membrane filtration conditions. The feed solution contained 0.2mmolL-1
ferrous ion, 0.05 mol L-1 Na2SO4 as electrolyte, and
5 mg ml-1 MB. The membrane proved to be stable under
multiple cycles, with a constant removal rate of approximately 90%. In addition, the applied potential and the pH
level were determined to play key roles in the E-Fenton
reactions, with potential increases leading to enhanced
H2O2 generation,, and the pH affecting the stability of Fe2+
ions, which were found to be stable between a pH of 2.5
and 4. Higher pH values increase iron precipitation in
the Fe(OH)3 form. When applying a potential of -1 V to the
membrane with a flux of ~100 LMH at pH 4, 95% of MB
was removed after 90 min.

4.2 Electrochemical oxidation


Vecitis et al. (2011a,b) used an electrochemically active
multiwalled CNT filter to degrade organic dyes. The filter
was prepared by the vacuum filtration of multiwall CNTs
onto a 5-m PTFE membrane. Scanning electron microscope (SEM) image analysis of the filter determined an
average pore size of 115 nm and a thickness of 41 m
(Figure 7A). The experiment was done in a single pass,
with a 1.2-s residence time. The filter reacted with the dyes
in two steps, initially adsorbing the dyes and then oxidizing and desorbing them (Figure 7B). Adsorption was the
dominant removal mechanism at low anodic potentials
(<1 V), but when the anodic potential exceeded 2 V, a high
oxidation rate was observed (>98% oxidation) (Figure 7C).
In addition, the electrochemical filter was able to oxidize
aqueous chloride and iodide to reactive halide species.
Zaky and Chaplin (2013, 2014) showed electrochemical oxidation for several model organic compounds (i.e.
benzoquinone, oxalic acid, ferrocyanide, and p-methoxyphenol [p-MP]) by using a porous substoichiometric titanium dioxide (Ti4O7), ceramic membrane. The ERM was
an Ebonex one-channel tubular electrode with a bimodal
Brought to you by | The Flinders University
Authenticated
Download Date | 11/14/16 8:33 AM

544A. Ronen etal.: Electroconductive responsive membranes

Figure 7:(A) SEM of the CNT filter. (a) Aerial image of the CNT filter with an average pore size of 115 nm. (b) Cross-sectional image of the CNT
filter with an average height of 41 m. (B) Mechanisms of interaction for dyes with CNT filter. (C) Electrochemical desorption and/or oxidation of adsorbed methyl orange dye as a function of applied potential. Dye adsorption was completed in the absence of applied potential.
Negative time points are the time prior to application of potential at t=0 noted by the vertical dashed line-adsorbed methyl orange oxidation
at potentials of 1 V (black squares), 2 V (red circles), and 3 V (blue triangles). (Reproduced from Vecitis etal., Copyright 2011, with permission
from American Chemical Society.)

pore size distribution. Pores <10nm accounted for >90%


of the surface area, and pores between 1 and 2 m contributed to nearly the entire pore volume. Electrochemical
anodic oxidation was done at potentials up to approximately 3 V versus standard hydrogen electrode with relatively low current densities (i.e. 0.51.0mA cm-2). Results
indicate advection-enhanced mass transfer rates, on the
order of a 10-fold increase, when the membrane was operated in filtration mode, relative to a flow-through mode.
All organic compounds showed high removal efficiency
by anodic oxidation, where the removal mechanism was
determined to be by both direct oxidation reactions and
formation of hydroxyl radicals (OH). In addition, at low
potentials (<2 V), p-MP was removed by electroassisted
adsorption and subsequent oxidation, whereas at higher
potentials (>2 V), where OH concentrations were significant, p-MP was removed by a combination of electroassisted adsorption and OH oxidation. Overall, 99.9% p-MP
was removed in the permeate stream.
Santos et al. (2016) used a simultaneous electrospinning/electrospraying technique to fabricate a composite membrane made of poly(sulfone) fibers and Ti4O7
nanoparticles. The composite membranes were highly
porous (i.e. 1645 m) and electrically conductive. Using
cyclic voltammetry, it was detected that that the kinetics for water electrolysis reactions and the ferrocyanide
redox were enhanced by nanoparticles deposition in the
composite membrane. Anodic oxidation experiments
in filtration mode using phenol as a model contaminant showed a 2.6-fold enhancement in the observed
first-order rate constant for phenol oxidation relative to

cross-flow operation at a current density of 1.0mAcm-2.


The enhancement was attributed to the increased mass
transfer rate in filtration mode. In addition, phenol oxidation rate constants in filtration mode were found to
be close to the pore diffusion mass transfer limit and
were six to eight times higher than measured for similar
ceramic ERM, indicating that further rate enhancements
could be achieved by synthesizing membranes with
smaller pore diameters, which will decrease the time
scale for pore diffusion.
Liu et al. (2014) showed organic fouling degradation by anodic oxidation during filtration. Conducting
filters were prepared by dispersing different weight ratios
(100:050:50) of GNPs and CNTs in N-methyl-2-pyrrolidone, with the stable suspension vacuum filtered onto a
PTFE membrane. The added CNTs were intertwined with
each other to serve as binders for GNPs. The GNP-CNT
electrochemical filter was evaluated using ferrocyanide (Fe(CN)64-) as a model electron donor. When the
filter was used in batch mode, electro-oxidation rates
increased linearly with applied potential and plateaued
because of mass transfer limitations. When the GNP-CNT
filter was evaluated as part of a flow-through system, no
plateau was observed for high concentrations of Fe(CN)64(10mmol L-1) as a result of increased mass transfer rates.
Overall, electro-oxidation rates increased up to 15-fold by
increasing influent concentration and internal electrode
convection. Next, three model compounds (tetracycline,
phenol, and oxalate) were used to evaluate oxidative
degradation. For all three organic compounds, electrooxidation kinetics increased with increasing anode
Brought to you by | The Flinders University
Authenticated
Download Date | 11/14/16 8:33 AM

A. Ronen etal.: Electroconductive responsive membranes545

potential until maximum removal rates (0.010, 0.064,


and 0.050molh-1m-2 for tetracycline, phenol, and oxalate,
respectively) were achieved at 0.8 V.
Although CNT and carbon-based ERM show promising abilities, they have been found unstable to oxidize
anodic potentials in water environments because of the
oxidation of the CNT and their breakdown when exposed
to hydroxyl radicals generated on the CNT surface
(Ohmori and Saito 2012, Tominaga et al. 2014). Ohmori
and Saito (2012) showed that high anode potentials
(>4V) caused severe CNT degradation in deionized water
for single-wall CNTs. Therefore, the development of novel
anode materials with high stability is of great importance. Attempts to increase stability of carbon-based ERM
to anodic oxidative damage were mainly done by anchoring or coating stable metal nanoparticles (i.e. cobalt
oxide and bismuth-doped tin oxide) on the surface on the
CNT. Although results showed some stability at potentials
up to 2.2 V (Liu etal. 2013b, Lu and Zhao 2013) (against
Ag/AgCl reference electrode), ERMs based on CNT are
still considered unstable at anodic electrical potentials.
However, it should be noted that CNT-based membranes
are stable when used as cathodes with no notable change
in conductivity over time.

4.3 E
 lectrochemical reactions in ERM vs.
advanced oxidation processes
Because electrochemical reactions coupled with ERM
are not applicable with conventional membranes, a comparison between ERM and common advanced oxidation
process (AOP) is discussed in terms of energy consumption. For the removal of a model compound (MB), it was
demonstrated by Legrini etal. (1993) that using a photocatalytic treatment requires high energy consumptions
and a long contact time (i.e. 40 kW m-3 and a typical
contact time of 3060 min) (Lachheb et al. 2002). A
similar electrochemical process using ERM has significantly lower energy requirement and lower contact time
(i.e. 0.16kWm-3 and a typical contact time <1.2 s) (Vecitis
etal. 2011a,b).

5 Future directions
Three promising future applications of ERM are presented
below. The first allows control of the membranes flux by
changing the pore size of a metallic grid coated by a conducting polymer (Jeon etal. 2011, Abelow etal. 2014). This

will increase the control over membrane selectivity and


will allow better separation properties tailored to a specific need. In addition, selective control over the flux may
allow better cleaning of the membranes (i.e. backwash)
and fouling mitigation. The second application relates to
the electronic control of a membranes wettability (Tsai
etal. 2011, Teng etal. 2015). Controlled membrane wettability (i.e. hydrophilicity) has many applications, ranging
from controlling flux and fouling to allowing better separation of oil from water. The third topic is capacitive deionization (CDI), which is an alternative technology for water
desalination, which has been demonstrated to be useful
in the desalination of relatively low salinity water (such
as agricultural drainage) with significantly lower energy
demands in comparison with RO (Welgemoed and Schutte
2005, Anderson etal. 2010, Zhao etal. 2012, 2013b, Wang
etal. 2015a,b).

5.1 Flux and selectivity control


Jeon etal. (2011) electrodeposited an electrically responsive nanoporous membrane based on polypyrrole doped
with the dodecylbenzenesulfonate anions (PPy/DBS)
on an anodized aluminum oxide membrane. The membrane had a regular pore size (380 nm) and a very high
pore density. Pore size control was achieved through large
volume changes of the PPy/DBS in response to an applied
electrical potential. The pore diameter of the membrane
was tuned by applying a potential of <1.1 V, with the pore
size decreasing at the reduced state (200 nm) and increasing when oxidized (380 nm), with a very fast reaction time
(<10 s). The change in pore size is attributed to the expansion of the PPy/DBS thin film at the reduced state. As both
PPy and DBS are negatively charged at cathodic potentials, hydrated positive counter ions (e.g. Na+) enter the
PPy chains to provide charge neutrality, which swells the
polymer. When a positive potential (anodic) is applied,
the PPy becomes positively charged, with the charge balanced by the negative DBS molecule, which leads to the
expulsion of the positive counter ions and subsequent
polymer shrinking. As a result of the change in pore size,
permeate flux changed as well, with flux increasing from
250 to 800 LMH, when the membrane polarity is switched
from a cathode to an anode.

5.2 Electrowettability
Teng etal. (2015) developed a conductive polymer porous
film with tunable wettability. The film was fabricated from

Brought to you by | The Flinders University


Authenticated
Download Date | 11/14/16 8:33 AM

546A. Ronen etal.: Electroconductive responsive membranes


the chloroform solution of poly(3-hexylthiophene) (P3HT)
and [6,6]-phenyl-C61-butyricacid-methyl-ester (PCBM)
using a freeze drying method, with pore diameters ranging
from 50 to 500 nm. The film was not evaluated as a membrane, but the porosity is in the range MF membranes. The
hydrophobic porous surface with a water contact angle
of 144.7 could be transferred into a hydrophilic surface
with contact angle of 25 by applying a voltage. Contact
angles decrease slowly at applied potentials of 220 V and
rapidly decrease when applying 25 V.
In an additional study, Tsai et al. (2011) presented
a mechanism for changes in surface wettability (i.e.
contact angle) of a polypyrrole film subjected to a lowvoltage electric field. A PPy/DBS film was fabricated
using electropolymerization from an aqueous pyrrole
solution on a coated Si wafer. The film was scanned
across a range of low electric potentials (-1.0 to 0.7 V)
while measuring the contact angle of water and an
organic solvent, dichloromethane (DCM) droplet. It was
found that the state of PPy/DBS can be tuned from
less hydrophobic (107) to more hydrophobic (133) via
the reorientation of its dopant molecules (DBS). In oxidized PPy/DBS, DBS molecules are coupled to the PPy
chains through ionic bonding with the polar sulfonic
acid group causing the dodecyl chains to point away
from the polymer chains. Since the strongly hydrophilic
and polar sulfonic acid group is attracted to the PPy
backbone, the surface becomes more hydrophobic, and
the contact angle of a water droplet on an oxidized PPy/
DBS film is increased. On the other hand, a nonpolar
liquid such as DCM shows opposite wetting behavior on
the oxidized PPy/DBS surface. In addition, because of
the low solubility of ions in DCM, surface tension gradients and Marangoni stress are induced because of the
lack of ions for PPy/DBS reduction underneath the DCM
droplet. Thus, when a PPy/DBS substrate is reduced, the
substrate directly beneath the DCM droplet remains in
the oxidized form.

5.3 Capacitive deionization


CDI is a water desalination technology using a small
applied potential to drive charged species (e.g. ions) in
saline water to the porous surface of the charged electrodes. The adsorption of ions is done on electrodes
based on carbon-based material (i.e. carbon black, CNT,
activated carbon, graphene, etc.) with a high surface
area which is connected to a current collecting electrode
(cathode or anode). Typical applied electrical potential
is <2 V (Zhao etal. 2012, Wang etal. 2015a,b, Yong etal.

2014). Membrane CDI (MCDI) has a similar configuration


as a conventional CDI, but in addition contains anion and
cation exchange membranes, which are positioned before
the carbon material and prevent transport of counter ions
to the electrode (Zhao etal. 2013b,c). Although this is not
a typical membrane process but an adsorption process,
composite electrodes (polymer-CNT) have been demonstrated to be efficient in replacing conventional carbon
electrodes. Liu et al. (2014) used modified electrodes
based on CNT and ion exchange polymers to achieve high
removal rates of NaCl (93% in comparison with 74% for
a typical MDCI). As MCDI involves the use of anion and
cation exchange membranes and the electrodes could be
designed as ERM, it is assumed that ERM will be incorporated in the process in the future.

6 Conclusions
Membrane separation processes have been part of the
water treatment industry for the last half century, with
major improvements in separation, fouling prevention,
and transport properties achieved over the last decade.
Although commercial membranes are efficient, fouling
of the membrane is still an inevitable issue. Polymeric
and inorganic ERMs show promising abilities for membrane separation processes. The ability to change the
membranes properties (surface charge, flux, wettability, and redox potential) and to prevent fouling through
electrostatic repulsion and electrochemical reactions
makes them ideal for wastewater treatment and water
desalination. Although a significant part of the literature deals with fouling prevention, new and advanced
treatments such as electro-oxidation and E-Fenton show
encouraging results in the removal of organic compounds
and pollutants, which are not fully removed by commercial membranes (e.g. some pharmaceuticals and small
uncharged molecules). ERM separation is still facing
some challenges regarding implementation, as each ERM
module must include a counter electrode, which increases
the cost of the initial setup and also calls for engineering
solutions in terms of a different flow cell structure. For
example, ERMs have been demonstrated to be efficient
in a tubular and flat sheet setup, but more complicated
spiral-wound ERM modules have not been fabricated yet.
In addition, at anodic conditions ceramic membranes
have been demonstrated to be stable, whereas polymeric
membranes suffer from stability limitations. Therefore,
additional research is needed in this field to allow cheap
and stable polymeric ERMs.
Brought to you by | The Flinders University
Authenticated
Download Date | 11/14/16 8:33 AM

A. Ronen etal.: Electroconductive responsive membranes547

It is expected that in the following years, the ability


to control pore size, surface wettability, and charge by
external applied electrical potentials will favor the use
of ERM-based separation processes in new fields such
as biotechnology (protein separation) and biomedicine
(controlled release of drugs) as well as in water treatment
processes.
Acknowledgments: This research was supported by BARD,
the United States-Israel Binational Agricultural Research
and Development Fund, Vaadia-BARD Postdoctoral Fellowship Award No. FI-497-2014.

References
Abelow AE, Persson KM, Jager EWH, Berggren M, Zharov I.
Electroresponsive nanoporous membranes by coating
anodized alumina with poly(3,4-ethylenedioxythiophene) and
polypyrrole. Macromol Mater Eng 2014; 299: 190197.
Al-Amoudi A, Lovitt RW. Fouling strategies and the cleaning system
of NF membranes and factors affecting cleaning efficiency.
JMembr Sci 2007; 303: 428.
Anderson MA, Cudero AL, Palma J. Capacitive deionization as an
electrochemical means of saving energy and delivering clean
water. Comparison to present desalination practices: will it
compete? Electrochim Acta 2010; 55: 38453856.
Ang WS, Lee S, Elimelech M. Chemical and physical aspects of
cleaning of organic-fouled reverse osmosis membranes.
JMembr Sci 2006; 272: 198210.
Asatekin A, Mennitib A, Kangd S, Elimelechd M, MorgenrothbE,
Mayes AM. Antifouling nanofiltration membranes for
membrane bioreactors from self-assembling graft copolymers.
J Membr Sci 2006; 285: 8189.
Ates M. A review study of (bio)sensor systems based on conducting
polymers. Mater Sci Eng C 2013; 33: 18531859.
Ayadi S, Jedidi I, Rivallin M, Gillot F, Lacour S, Cerneaux S, Cretin M,
Amar RB. Elaboration and characterization of new conductive
porous graphite membranes for electrochemical advanced
oxidation processes. J Membr Sci 2013; 446: 4249.
Baatout S, Boever P, Mergeay M. Physiological changes induced in
four bacterial strains following oxidative stress. Prikl Biokhim
Mikrobiol 2006; 42: 369377.
Babuponnusami A, Muthukumar K. Advanced oxidation of
phenol: a comparison between fenton, electro-fenton,
sono-electro-fenton and photo-electro-fenton processes. Chem
Eng J 2012; 183: 19.
Baker JS, Dudley LY. Biofouling in membrane systems a review.
Desalination 1998; 118: 8189.
Barraud N, Hassett DJ, Hwang SH, Rice SA, Kjelleberg S, Webb JS.
Involvement of nitric oxide in biofilm dispersal of Pseudomonas
aeruginosa. J Bacteriol 2006; 188: 73447353.
Bhattacharyya D, Schfer T, Wickramasinghe SR, Daunert S.
Responsive membranes and materials. New York: Wiley, 2012.
Boles BR, Singh PK. Endogenous oxidative stress produces diversity
and adaptability in biofilm communities. Proc Natl Acad Sci
USA 2008; 105: 1250312508.

Brant J, Childress AE. Membrane-colloid interactions: comparison


of extended dlvo predictions with AFM force measurements.
Environ Eng Sci 2002; 19: 413427.
van der Bruggen B, Vandecasteele C, van Gestel T, Doyen W,
Leysen R. A review of pressure-driven membrane processes in
wastewater treatment and drinking water production. Environ
Prog 2003; 22: 4656.
Buonomenna MG. Membrane processes for a sustainable industrial
growth. RSC Adv 2013; 3: 5694.
Callow ME, Callow JA. Marine biofouling: a sticky problem. Biologist
2002; 49: 1014.
p M, Vchov L, Palkov Z. Reactive oxygen species in
the signaling and adaptation of multicellular microbial
communities. Oxid Med Cell Longev 2012; 2012: 113.
Chan EP, Mulhearn WD, Huang YR, Lee JH, Lee D, Stafford CM.
Tailoring the permselectivity of water desalination membranes
via nanoparticle assembly. Langmuir 2014; 30: 611616.
Chen PR, Chuang YJ. The development of conductive nanoporous
chitosan polymer membrane for selective transport of charged
molecules. J Nanomater 2013; 2013: 6.
Chen X, Zhao B, Zhao L, Bi S, Han P, Feng X, Chen L. Temperatureand pH-responsive membranes based on poly (vinylidene
fluoride) functionalized with microgels. J Membr Sci 2014; 469:
447457.
Chuo T-W, Wei T-C, Chang Y, Liu Y-L. Electrically driven biofouling
release of a poly(tetra fl uoroethylene) membrane modi fied
with an electrically induced reversibly cross-linked polymer.
ACS Appl Mater Interfaces 2013; 5: 99189925.
Coster HGL, Farahani TD, Chilcott TC. Production and
characterization of piezo-electric membranes. Desalination
2011; 283: 5257.
Darestani MT, Coster HGL, Chilcott TC. Piezoelectric membranes for
separation processes: fabrication and piezoelectric properties.
J Membr Sci 2013a; 434: 184192.
Darestani MT, Coster HGL, Chilcott TC. Piezoelectric membranes
for separation processes: operating conditions and filtration
performance. J Membr Sci 2013b; 435: 226232.
Duan W, Dudchenko A, Mende E, Flyer C, Zhu X, Jassby D.
Electrochemical mineral scale prevention and removal on
electrically conducting carbon nanotube polyamide reverse
osmosis membranes. Environ Sci Processes Impacts 2014; 16:
13001308.
Dudchenko AV, Rolf J, Russell K, Duan W, Jassby D. Organic fouling
inhibition on electrically conducting carbon nanotube
polyvinyl alcohol composite ultrafiltration membranes.
JMembr Sci 2014; 468: 110.
Elimelech M, Phillip WA. The future of seawater desalination:
energy, technology, and the environment. Science (New York)
2011; 333: 712717.
Eshed L, Yaron S, Dosoretz CG. Effect of permeate drag force on
the development of a biofouling layer in a pressure-driven
membrane separation system. Appl Environ Microbiol 2008;
74: 73387347.
Fan L, Harris JL, Roddick FA, Booker NA. Influence of the
characteristics of natural organic matter on the fouling of
microfiltration membranes. Water Res 2001; 35: 44554463.
Feng R, Wang C, Xu X, Yang F, Xu G, Jiang T. Highly effective
antifouling performance of N-vinyl-2-pyrrolidone modified
polypropylene non-woven fabric membranes by ATRP method.
JMembr Sci 2011; 369: 233242.

Brought to you by | The Flinders University


Authenticated
Download Date | 11/14/16 8:33 AM

548A. Ronen etal.: Electroconductive responsive membranes


Field R. Fundamentals of fouling in membrane technology. In:
Klaus-Viktor P, Suzana PN, editors. Membranes for water
treatment, vol. 4. Weinheim: Wiley-VCH Verlag GmbH & Co,
2010: 123.
Flemming HC. Biofouling in water systems-cases, causes and
countermeasures. Appl Microbiol Biotechnol 2002; 59:
629640.
Flemming HC, Schaule G, Griebe T, Schmitt J, TamachkiarowaA.
Biofouling the Achilles heel of membrane processes.
Desalination 1997; 113: 215225.
Frost S, Ulbricht M. Thermoresponsive ultrafiltration membranes for
the switchable permeation and fractionation of nanoparticles.
JMembr Sci 2013; 448: 111.
Gao G, Zhang Q, Hao Z, Vecitis CD. Carbon nanotube membrane
stack for flow-through sequential regenerative electro-fenton.
Environ Sci Technol 2015; 49: 23752383.
Geise GM, Cassady HJ, Paul DR, Logan BE, Hickner MA. Specific
ion effects on membrane potential and the permselectivity of
ion exchange membranes. Phys Chem Chem Phys 2014; 16:
2167321681.
Goosen MFA, Sablanib SS, Al-Hinaia H, Al-Obeidania S, Al-Belushib
R, Jackson D. Fouling of reverse osmosis and ultrafiltration
membranes: a critical review. Sep Sci Technol 2005; 39:
22612297.
Gugliuzza A. Intelligent membranes: dream or reality? Membranes
2013; 3: 151154.
Guo B, Glavas L, Albertsson A-C. Biodegradable and electrically
conducting polymers for biomedical applications. Prog Polym
Sci 2013; 38: 12631286.
Guo L, Jung Y, Chaplin BP. Development and characterization of
ultrafiltration TiO2 Magnli phase reactive electrochemical
membranes. Environ Sci Technol 2016; 50: 14281436.
Guo W, Ngo HH, Li J. A mini-review on membrane fouling. Bioresour
Technol 2012; 122: 2734.
Hashaikeh R, Lalia BS, Kochkodan V, Hilal N. A novel in situ
membrane cleaning method using periodic electrolysis.
JMembr Sci 2014; 471: 149154.
He Y, Chen X, Bi S, Fu W, Shi C, Chen L. Conferring pH-sensitivity on
poly (vinylidene fluoride) membrane by poly (acrylic acid-co-butyl
acrylate) microgels. React Funct Polym 2014; 74: 5866.
Hegab HM, Zou L. Graphene oxide-assisted membranes: fabrication
and potential applications in desalination and water
purification. J Membr Sci 2015; 484: 95106.
Hermansson M. The DLVO theory in microbial adhesion. Colloids
Surf B Biointerfaces 1999; 14: 105119.
Hijnen WAM, Castillo C, Brouwer-Hanzens AH, Harmsen DJ,
Cornelissen ER, van der Kooij D. Quantitative assessment of
the efficacy of spiral-wound membrane cleaning procedures to
remove biofilms. Water Res 2012; 46: 63696381.
Holder A, Weik J, Hinrichs J. A study of fouling during long-term
fractionation of functional peptides by means of cross-flow
ultrafiltration and cross-flow electro membrane filtration. J
Membr Sci 2013; 446: 440448.
Hong S, Elimelech M. Chemical and physical aspects of natural
organic matter (NOM) fouling of nanofiltration membranes. J
Membr Sci 1997; 132: 159181.
Huang J, Wang Z, Zhang J, Zhang X, Wu Z. A novel composite
conductive microfiltration membrane and its anti-fouling
performance with an external electric field in membrane
bioreactors. Sci Rep 2015; 5: 18.

Jeon G, Yang SY, Byun J, Kim JK. Electrically actuatable smart


nanoporous membrane for pulsatile drug release. Nano Lett
2011; 11: 12841288.
Ji Y-L, An Q-F, Guo Y-S, Hung W-S, Lee K-R, Gao C-J. Bio-inspired
fabrication of high perm-selectivity and anti-fouling
membranes based on zwitterionic polyelectrolyte
nanoparticles. J Mater Chem A 2016; 4: 42244231.
Jin Y-L, Lee WN, Lee CH, Chang IS, Huang X, Swaminathan T. Effect
of DO concentration on biofilm structure and membrane
filterability in submerged membrane bioreactor. Water Res
2006; 40: 28292836.
Katsuno C, Konda A, Urayama K, Takigawa T, Kidowaki M, Ito K.
Pressure-responsive polymer membranes of slide-ring gels
with movable cross-links. Adv Mater 2013; 25: 46364640.
Katz I, Dosoretz CG. Desalination of domestic wastewater effluents:
phosphate removal as pretreatment. Desalination 2008; 222:
230242.
Kennedy MD, Kamanyi J, Salinas Rodrguez SG, Lee NH, Schippers
JC, Amy C. Water treatment by microfiltration and ultrafiltration.
In: Matsuura T, Li NN, Fane AG, Winston Ho WS, editors.
Advance membrane technology and application. New Jersey:
John Wiley, 2008: 131150.
Krinks JK, Qiu MH, Mergos IA, Weavers LK, Mouser PJ, Verweij H.
Piezoceramic membrane with built-in ultrasonic defouling. J
Membr Sci 2015; 494: 130135.
Kuo TC, Sloan LA, Sweedler JV, Bohn PW. Manipulating molecular
transport through nanoporous membranes by control of
electrokinetic flow: effect of surface charge density and debye
length. Langmuir 2001; 17: 62986303.
Lachheb H, Puzenata E, Houasb A, Ksibib M, Elalouib E, Guillarda
C, Herrmann J-M. Photocatalytic degradation of various types
of dyes (alizarin S, crocein orange G, methyl red, congo red,
methylene blue) in water by UV-irradiated titania. Appl Catal B
Environ 2002; 39: 7590.
de Lannoy CF, Jassby D, Davis DD, Wiesner MR. A highly electrically
conductive polymermultiwalled carbon nanotube
nanocomposite membrane. J Membr Sci 2012; 415416: 718724.
de Lannoy CF, Jassby D, Gloe K, Gordon AD, Wiesner MR. Aquatic
biofouling prevention by electrically charged nanocomposite
polymer thin film membranes. Environ Sci Technol 2013; 47:
27602768.
Lee W-N, Changb I-S, Hwanga B-K, Parka P-K, Leea C-H, Huangc X.
Changes in biofilm architecture with addition of membrane
fouling reducer in a membrane bioreactor. Proc Biochem 2007;
42: 655661.
Lee KP, Arnot TC, Mattia D. A review of reverse osmosis membrane
materials for desalination-development to date and future
potential. J Membr Sci 2011; 370: 122.
Legrini O, Oliveros E, Braun AM. Photochemical processes for water
treatment. Chem Rev 1993; 93: 671698.
Lenntech. Product Information: Dow Filmtec BW30FR-365 Fouling
Resistant RO Element, 2013.
Li H, Liao J, Xiang T, Wang R, Wang D, Sun S, Zhao C. Preparation and
characterization of pH- and thermo-sensitive polyethersulfone
hollow fiber membranes modified with P(NIPAAm-MAA-MMA)
terpolymer. Desalination 2013; 309: 110.
Li N, Liu L, Yang F. Highly conductive graphene/PANi-Phytic acid
modi Fi Ed cathodic Fi Lter membrane and its antifouling
property in EMBR in neutral conditions. Desalination 2014;
338: 1016.

Brought to you by | The Flinders University


Authenticated
Download Date | 11/14/16 8:33 AM

A. Ronen etal.: Electroconductive responsive membranes549


Libana R, Arregui L, Belda I, Gamella L, Santos A, Marquina D,
Serrano S. Membrane bioreactor wastewater treatment plants
reveal diverse yeast and protist communities of potential
significance in biofouling. Biofouling 2015; 31: 7182.
Liu M, Cote P, Siverns S. Water reclamation and desalination by
membranes. In: Matsuura T, Li NN, Fane AG, Winston Ho WS,
editors. Advanced membrane technology and applications.
Hoboken, NJ: John Wiley & Sons, Inc., 2008: 17187.
Liu CX, Zhang DR, He Y, Zhao XS, Bai R. Modification of
membrane surface for anti-biofouling performance: effect
of anti-adhesion and anti-bacteria approaches. J Membr Sci
2010; 346: 121130.
Liu L, Liu J, Gao B, Yang F, Chellam S. Fouling reductions in a
membrane bioreactor using an intermittent electric field and
cathodic membrane modified by vapor phase polymerized
pyrrole. J Membr Sci 2012; 394395: 202208.
Liu C, Xie X, Zhao W, Liu N, Maraccini PA, Sassoubre LM, Boehm AB,
Cui Y. Conducting nanosponge electroporation for affordable
and high efficiency disinfection of bacteria and viruses in
water. Nano Lett 2013a; 13: 42884293.
Liu H, Vajpayee A, Vecitis CD. Bismuth-doped tin oxide-coated
carbon nanotube network: improved anode stability and
efficiency for flow-through organic electrooxidation. ACS Appl
Mater Interfaces 2013b; 5: 1005410066.
Liu J, Liu L, Gao B, Yang F. Integration of bio-electrochemical cell in
membrane bioreactor for membrane cathode fouling reduction
through electricity generation. J Membr Sci 2013c; 430: 196202.
Liu L, Liu J, Bo G, Yang F, Crittenden, Chen Y. Conductive and
hydrophilic polypyrrole modified membrane cathodes and
fouling reduction in MBR. J Membr Sci 2013d; 429: 252258.
Liu L, Zhao F, Liu J, Yang F. Preparation of highly conductive cathodic
membrane with graphene (oxide)/PPy and the membrane
antifouling property in filtrating yeast suspensions in EMBR. J
Membr Sci 2013e; 437: 99107.
Liu Y, Rosenfield E, Hu M, Mi B. Direct observation of bacterial
deposition on and detachment from nanocomposite
membranes embedded with silver nanoparticles. Water Res
2013f; 47: 29492958.
Liu Y, Lee JHD, Xia Q, Ma Y, Yu Y, Yung LYL, Xie J, Ong CN, Vecitis CD,
Zhou Z. A graphene-based electrochemical filter for water
purification. J Mater Chem A 2014; 2: 1655416562.
Lu X, Zhao C. Highly efficient and robust oxygen evolution catalysts
achieved by anchoring nanocrystalline cobalt oxides onto
mildly oxidized multiwalled carbon nanotubes. J Mater Chem A
2013; 1: 12053.
Ma S, Meng J, Li J, Zhang Y, Ni L. Synthesis of catalytic polypropylene
membranes enabling visible-light-driven photocatalytic
degradation of dyes in water. J Membr Sci 2014; 453: 221229.
Malaisamy R, Berry D, Holder D, Raskin L, Lepak L, Jones KL.
Development of reactive thin film polymer brush membranes to
prevent biofouling. J Membr Sci 2010; 350: 361370.
Marchal V, Dellink R, van Vuuren D, Clapp C, Chteau J, Lanzi E,
Magn B, van Vliet J. OECD environmental outlook to 2050:
theconsequences of inaction. Paris: OECD Publishing, 2012.
Martins P, Lopes AC, Lanceros-Mendez S. Electroactive phases
of poly(vinylidene fluoride): determination, processing and
applications. Prog Polym Sci 2014; 39: 683706.
Meng H, Hu J. A brief review of stimulus active polymers
responsive to thermal, light, magnetic, electric, and water/
solvent stimuli. J Intell Mater Sys Struct 2010; 21: 859885.

Munna MS, Nur I, Rahman T, Noor R. Influence of exogenous


oxidative stress on Escherichia coli cell growth, viability and
morphology. Am J Biosci 2013; 1: 59.
Nagarale RK, Gohil GS, Shahi VK. Recent developments on
ion-exchange membranes and electro-membrane processes.
Adv Colloid and Interface Sci 2006; 119: 97130.
Narendra G, Ramesh V. Effect of hydrogen peroxide induced oxidative
stress on increasing the susceptibility of gram negative bacteria
against ciprofloxacin. Int J Future Biotechnol 2012; 1: 16.
Nicoletta FP, Cupelli D, Formoso P, Filpo GD, Colella V, Gugliuzza A.
Light responsive polymer membranes: a review. Membranes
2012; 2: 134197.
Ohmori S, Saito T. Electrochemical durability of single-wall carbon
nanotube electrode against anodic oxidation in water. Carbon
2012; 50: 49324938.
Otero TF, Martinez JG, Arias-Pardilla J. Biomimetic electrochemistry
from conducting polymers. a review: artificial muscles, smart
membranes, smart drug delivery and computer/neuron
interfaces. Electrochim. Acta 2012; 84: 112128.
Park N, Kwon B, Kim IS, Cho J. Biofouling potential of various nf
membranes with respect to bacteria and their soluble microbial
products (SMP): characterizations, flux decline, and transport
parameters. J Membr Sci 2005; 258: 4354.
Prihasto N, Liu Q-F, Kim S-H. Pre-treatment strategies for seawater
desalination by reverse osmosis system. Desalination 2009;
249: 308316.
Pyo M, Reynolds JR, Warren LF, Marcy HO. Long-term redox switching
stability of polypyrrole. Synthetic Metals 1994; 68: 7177.
Qiu J-H, Zhang Y-W, Zhang Y-T, Zhang H-Q, Liu J-D. Synthesis and
antibacterial activity of copper-immobilized membrane
comprising grafted poly(4-vinylpyridine) chains. J Colloid
Interface Sci 2011; 354: 152159.
Qiu X, Yu H, Karunakaran M, Pradeep N, Nunes SP, Peinemann K-V.
Selective separation of similarly sized proteins with tunable
nanoporous block copolymer membranes. ACS Nano 2013; 7:
768776.
Ronen A, Duan W, Wheeldon I, Walker SL, Jassby D. Microbial
attachment inhibition through low voltage electrochemical
reactions on electrically conducting membranes. Environ Sci
Technol 2015; 49: 1274112750.
Saeki D, Nagao S, Sawada I, Ohmukai Y, Maruyama T, Matsuyama
H. Development of antibacterial polyamide reverse osmosis
membrane modified with a covalently immobilized enzyme. J
Membr Sci 2013; 428: 403409.
Santos MC, Elabd YA, Jing Y, Chaplin BP, Fang L. Highly porous
Ti4O7 reactive electrochemical water filtration membranes
fabricated via electrospinning/electrospraying. AIChE J 2016;
62: 508524.
Shannon MA, Bohn PW, Elimelech M, Georgiadis JG, Marias BJ,
Mayes AM. Science and technology for water purification in the
coming decades. Nature 2008; 452: 301310.
Shirazi S, Lin CJ, Chen D. Inorganic fouling of pressure-driven
membrane processesa critical review. Desalination 2010;
250: 236248.
Singh BL, Ejaz F, Shah T, Hilal N, Hashaikeh R. Electrically conductive
membranes based on carbon nanostructures for self-cleaning
of biofouling. Desalination 2015; 360: 812.
Teng Y, Zhang Y, Heng L, Meng X, Yang Q, Jiang L. Conductive
polymer porous film with tunable wettability and adhesion.
Materials 2015; 8: 18171830.

Brought to you by | The Flinders University


Authenticated
Download Date | 11/14/16 8:33 AM

550A. Ronen etal.: Electroconductive responsive membranes


Tominaga M, Yatsugi Y, Watanabe N. Oxidative corrosion potential
vs. pH diagram for single-walled carbon nanotubes. RSC Adv
2014; 4: 27224.
Tsai YT, Choi CH, Gao N, Yang EH. Tunable wetting mechanism of
polypyrrole surfaces and low-voltage droplet manipulation via
redox. Langmuir 2011; 27: 42494256.
Vecitis CD, Gao G, Liu H. Electrochemical carbon nanotube filter for
adsorption, desorption, and oxidation of aqueous dyes and
anions. J Phys Chem C 2011a; 115: 36213629.
Vecitis CD, Schnoor MH, Rahaman S, Schi JD, Elimelech M.
Electrochemical multiwalled carbon nanotube filter for viral and
bacterial removal and inactivation. Environ Sci Technol 2011b;
45: 36723679.
Wandera DS, Wickramasinghe R, Husson SM. Stimuli-responsive
membranes. J Membr Sci 2010; 357: 635.
Wang S, Guillen G, Hoek EMV. Direct observation of microbial
adhesion to membranes. Environ Sci Technol 2005; 39:
64616469.
Wang R, Xiang T, Yue W, Li H, Liang S, Sun S, Zhao C. Preparation and
characterization of pH-sensitive polyethersulfone hollow fiber
membranes modified by poly(methyl methylacrylate-Co-4-vinyl
pyridine) copolymer. J Membr Sci 2012a; 423424: 275283.
Wang Z, Yao X, Wang Y. Swelling-induced mesoporous block
copolymer membranes with intrinsically active surfaces for
size-selective separation. J Mater Chem 2012b; 22: 20542.
Wang C, Song H, Zhang Q, Wang B, Li A. Parameter optimization
based on capacitive deionization for highly efficient desalination
of domestic wastewater biotreated efficient and the fouled
electrode regeneration. Desalination 2015a; 365: 407415.
Wang S, Liang S, Liang P, Zhang X, Sun J, Wu S, Huang X. In-situ
combined dual-layer CNT/PVDF membrane for electricallyenhanced fouling resistance. J Membr Sci 2015b; 491: 3744.
Wei X, Wang Z, Chen J, Wang J, Wang S. A novel method of surface
modification on thin-film-composite reverse osmosis
membrane by grafting hydantoin derivative. J Membr Sci 2010;
346: 152162.
Weidlich C, Mangold KM. Electrochemically switchable polypyrrole
coated membranes. Electrochim Acta 2011; 56: 34813484.
Welgemoed TJ, Schutte CF. Capacitive deionization TechnologyTM:
an alternative desalination solution. Desalination 2005; 183:
327340.
Xiao L, Isner A, Waldrop K, Saad A, Takigawa D, Bhattacharyya D.
Development of bench and full-scale temperature and pH
responsive functionalized PVDF membranes with tunable
properties. J Membr Sci 2014; 457: 3949.
Xu T. Ion exchange membranes: state of their development and
perspective. J Membr Sci 2005; 263: 129.
Xu L, Zhang G-Q, Yuan G-E, Liu H-Y, Liu J-D, Yang F-L. Anti-fouling
performance and mechanism of anthraquinone/polypyrrole
composite modified membrane cathode in a novel MFCaerobic
MBR coupled system. RSC Adv 2015; 5: 2253322543.
Yang HL, Lin JC, Huang C. Application of nanosilver surface
modification to RO membrane and spacer for mitigating

biofouling in seawater desalination. Water Res 2009; 43:


37773786.
Yang Q, Himstedt HM, Ulbricht M, Qian X, Wickramasinghe
SR. Designing magnetic field responsive nanofiltration
membranes. J Membr Sci 2013; 430: 7078.
Yao C, Xinsong L, Neoh KG, Shi Z, Kang ET. Surface modification
and antibacterial activity of electrospun polyurethane fibrous
membranes with quaternary ammonium moieties. J Membr Sci
2008; 320: 259267.
Yaroshchuk A, Boiko Y, Makovetskiy A. Electrochemical
perm-selectivity of active layers and diffusion permeability
of supports of an asymmetric and a composite NF membrane
studied by concentration-step method. Desalination 2009;
245: 374387.
Ying W, Kumar R, Herzberg M, Kasher R. Diminished swelling of
cross-linked aromatic oligoamide surfaces revealing a new
fouling mechanism of reverse-osmosis membranes. Environ Sci
Technol 2015; 49: 68156822.
Ying W, Siebdrath N, Uhl W, Gitis V, Herzebrg M. New insights
on early stages of RO membranes fouling during tertiary
wastewater desalination. J Membr Sci 2014; 466: 2635.
Yong L, Pan L, Xu X, Lu T, Sun Z, Chu DHC. Enhanced desalination
efficiency in modified membrane capacitivedeionization
by introducing ion-exchange polymers in carbonnanotubes
electrodes. Electrochim. Acta 2014; 130: 619624.
Yu S, Liu X, Liu J, Wu D, Liu M, Gao C. Surface modification of thin-film
composite polyamide reverse osmosis membranes with thermoresponsive polymer (TRP) for improved fouling resistance and
cleaning efficiency. Sep Purif Technol 2011; 76: 283291.
Zaky AM, Chaplin BP. Porous substoichiometric TiO2 anodes as
reactive electrochemical membranes for water treatment.
Environ Sci Technol 2013; 47: 65546563.
Zaky AM, Chaplin BP. Mechanism of P-substituted phenol oxidation
at a Ti4O7 reactive electrochemical membrane. Environ Sci
Technol 2014; 48: 58575867.
Zhang M, Nguyen QT, Ping Z. Hydrophilic modification of poly
(vinylidene fluoride) microporous membrane. J Membr Sci
2009; 327: 7886.
Zhang Q, Vecitis CD. Conductive CNT-PVDF membrane for capacitive
organic fouling reduction. J Membr Sci 2014; 459: 143156.
Zhao R, Biesheuvel PM, van der Wal A. Energy consumption
and constant current operation in membrane capacitive
deionization. Energy Environ Sci 2012; 5: 9520.
Zhao F, Liu L, Yang F, Ren N. E-Fenton degradation of MB during
filtration with Gr/PPy modified membrane cathode. Chem Eng J
2013a; 230: 491498.
Zhao R, Porada S, Biesheuvel PM, van Der Wal A. Energy
consumption in membrane capacitive deionization for different
water recoveries and flow rates, and comparison with reverse
osmosis. Desalination 2013b; 330: 3541.
Zhao R, Satpradit O, Rijnaarts HHM, Biesheuvel PM, van der Wal A.
Optimization of salt adsorption rate in membrane capacitive
deionization. Water Res 2013c; 47: 19411952.

Brought to you by | The Flinders University


Authenticated
Download Date | 11/14/16 8:33 AM

Você também pode gostar