c03

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 44

69

Filtration Membranes with Responsive Gates

3.1 Membrane Separation for Water Purification


and Desalination
In the last century, with the booming of food, biotechnology, and pharmaceutical
industries, the conventional separation technologies, such as distillation, subli-
mation, crystallization, and adsorption, became either less efficient or impossible
to meet some special needs from these industries. The membrane-based separa-
tion, in which porous membrane serves as a barrier to separate undesired species
out of a mixture, came into being first in the early 1900s to aid the development
in these industries and has since dominated and expanded its presence in many
aspects of major industrial processes.
Membrane technology is a key component of an integrated water treatment
and reuse paradigm [1]. It is safe to state that the membrane technology has
contributed significantly to the steadily improved living standards of mankind in
the past. Particularly, in the water treatment sector, membrane-based separation
has its niche benefits: (i) Membrane separation is largely physical, and both per-
meate and retentate can be collected and utilized, which has a special meaning
nowadays in wastewater treatment as there is a growing interest in recovering
valuable resources, including water, nutrient, energy, etc., from municipal and
industrial wastewaters. As resource recovery will be an integral part of wastew-
ater treatment in the coming decades, the importance of membrane in water
treatment cannot be overstated. (ii) By employing membranes with different pore
sizes or separation mechanisms, membrane separation provides fit-for-purpose
products, which offers flexibility and precise separation at the lowest energy cost
and thus vastly boosts the separation energy efficiency. (iii) From the engineer-
ing point of view, membrane-based separation system requires less space than
most of the conventional technologies for the same purpose of separation [2]. All
these advantages make membranes essential tools to the current and future water
treatment industry.
The performance of a membrane largely depends on the membrane material,
which bears an inherent trade-off between solvent permeability and solute selec-
tivity or rejection, both of which are unalterable in conventional membranes
and cannot be tuned during their operations [3]. Based on the cutoff size of
membrane separation, filtration membrane can be classified into microfiltration

Artificially Intelligent Nanomaterials for Environmental Engineering, First Edition.


Peng Wang, Jian Chang, and Lianbin Zhang.
© 2019 Wiley-VCH Verlag GmbH & Co. KGaA. Published 2019 by Wiley-VCH Verlag GmbH & Co. KGaA.
70 3 Filtration Membranes with Responsive Gates

0.1 nm 1 nm 10 nm 100 nm 1 μm 10 μm

Hydrated ions Viruses


Micropollutants Bacteria
Natural organic matter Protozoa (for example, Cryptosporidium, Giardia)
Algal toxins

Reverse osmosis,
Ultrafiltration
forward osmosis

Nanofiltration Microfiltration

Solution–diffusion Size exclusion

Figure 3.1 Several common membrane processes for water purification and desalination
separate largely on the basis of solute size. Source: Werber et al. 2016 [1]. Reprinted with
permission of Springer Nature.

(MF), ultrafiltration (UF), nanofiltration (NF), reverse osmosis (RO), and forward
osmosis (FO) (Figure 3.1).
MF membranes are used to sieve suspended particles and microbial pathogens,
while UF membranes are designed to separate macromolecules with smaller
scales than MF, such as natural organic matter and smaller pathogens (viruses).
In general, the molecular weight cutoff of UF membranes ranges from approx-
imately 5 to 500 kDa [1]. NF membranes can remove scale-forming ions, like
calcium and magnesium, as a function of reducing salinity. Typically, molecular
weight cutoff of NF membranes is between 100 and 300 Da. Filtration in NF
is based on a combination of sieving and solution–diffusion mechanisms. RO
and FO membranes are designed for desalination to remove nearly all ions in
addition to uncharged solutes of molecular weight greater than around 100 Da
[1]. Furthermore, the filtration process of FO is driven by an osmotic pressure
difference between the feedwater and draw solution of high osmotic pressure
[4, 5].
Over the years, membrane separation has been an important playground
for innovative nano-designs, which enable many conventional membranes to
steadily improve their separation performances [6, 7].

3.2 Emerging Design and Concept of Filtration


Membranes with Responsive and Intelligent Gates
Generally, the separation performance of a membrane is determined by
trans-membrane flux and the rejection of specific substrate, and such perme-
ability and selectivity are in turn influenced by the membrane pore size, surface
properties, and the interactions between the permeating substances and the
pore surface. Unfortunately, membrane performance is largely pre-fixated in
3.3 Fabrication Methods of Intelligent Gating Membranes 71

conventional membranes owing to the lack of molecular-level design [8]. The


structural control of the selecting layer is limited, thus hampering dispersing
medium and substrate selectivity and increasing high fouling propensity [3].
Novel materials and scalable molecular-level design approach for membrane
fabrication would be imperative for overcoming these limitations and further
endow the capability of adjusting pore size, pore channel chemistry, etc. in
response to environmental conditions. Thus, membranes with tunable perme-
ability and selectivity would clearly outperform the conventional ones and offer
considerable promise for performing complex tasks and substantially advancing
water purification and desalination technologies.
In looking for inspiration to make better members, nature offers us a won-
derful model of intelligent gating: cell membrane. The cell membrane possesses
extremely selective ion channels that allow only targeted ions to pass through
with very high rate. More importantly, these ion channels can be switched on
and off on demand by modulation charge/concentration gradient between the
sides of the cell membrane, which alters the conformation of the channel pro-
teins. In a sense, cell membrane works as stimuli-responsive intelligent “gates,”
regulating their opening and closing in response to chemical or electrical signals,
temperature, or mechanical force [9, 10].
The cell membrane is always a great inspiration to scientist to develop artificial
intelligent gating membranes toward more versatile and effective separation, but
the stimuli-responsive membrane field is still at its infantry stage and is vastly far
from being close to the delicacy and precision nature offers us. Valuable but lim-
ited efforts have been made in making membranes with stimuli responsiveness
and tunable pore size/pore chemistry at laboratory scale, and there has not been
any successful application of these membranes at large scale.
Synthetic intelligent gating membranes have emerged since the 1960s when the
stimuli-responsive behaviors of some polymers were first revealed by Heskins
and Guillet [11]. As we discussed in Chapter 2, polymers are the most com-
monly investigated stimuli-responsive materials. The general design principle of
artificial intelligent gating membranes is to incorporate stimuli-responsive mate-
rials, dominantly polymers, into the pores of membranes and these polymers,
in response to appropriate stimuli, such as temperature, pH, and light, ion, ion
strength, etc. The change in conformations and chemistry would in turn adjust
the pore sizes and/or the surface properties of the membranes, leading to mod-
ulation of permeability and selectivity of the membranes (Figure 3.2) [12–14].
The artificial intelligent gating membranes combine the advantages of porous
substrates and intelligent gates for distinct performances responsive to environ-
mental triggers and thus perform more complex tasks [15–19].

3.3 Fabrication Methods of Intelligent Gating


Membranes
Methods to produce intelligent gating membranes can be generally divided into
two categories: (i) post-modification of existing porous membranes, which is
72 3 Filtration Membranes with Responsive Gates

Closed state Open state

Triggers

Temperature
pH
Light

Swollen state Shrunken state

Figure 3.2 Schematic representation of the different gating states of intelligent gating
membrane in response to appropriate triggers. Source: Chang et al. 2018 [2]. Reprinted with
permission of Royal Society of Chemistry.

the most popular method to fabricate polymer-based stimuli-responsive gating


membranes and where the polymeric species are bound onto the pore surface
via covalent bonding [20–23], van der Waals forces [24], electrostatic interaction
[25, 26], and so on, and (ii) one-step formation of stimuli-responsive gating
membranes.

3.3.1 Post-Modification Method


The post-modification method means incorporating the gates on existing porous
membrane substrates using covalently or noncovalently bonding.
Covalent grafting techniques, namely, grafting-to and grafting-from methods,
are that functional polymeric materials or responsive small molecules are cova-
lently modified onto the pore surface of an existing membrane. Both methods
allow the fabrication of gating membranes with steady gating structures and
highly efficient gating performances.
For grafting-from method, gating membranes are fabricated by first inducing
active sites on the pore surface and then polymerizing functional monomers
from the active sites to constitute linear polymers or cross-linked networks in
the pores as the intelligent gates. The typical grafting polymerization meth-
ods include atom transfer radical polymerization (ATRP) [27–37], reversible
addition–fragmentation chain transfer (RAFT) [23, 38], UV-induced grafting
[21, 39–43], and plasma-induced grafting [22, 44–60]. With various grafting
polymerization methods, the functional gates can be incorporated into a wide
range of membrane substrates for forming the intelligent gating membranes.
For the grafting-to method, the membrane surface can be functionalized by
chemically grafting preformed functional polymeric materials or responsive
small molecules based on covalent bonding between the functional group
of membrane materials and a reactive group from polymers [20]. Therefore,
the intelligent gating membranes are fabricated by incorporating preformed
functional gates onto the pore surfaces.
Moreover, since polymer chains with well-controlled length or size can
be pre-synthesized using well-established methods, the grafting-to method
possesses improved controllability and flexibility for the gate microstructures.
However, in comparison, the grafting-from method can lead to higher grafting
3.4 Application of Intelligent Gating Membranes to Environmental Separation 73

densities than grafting-to method due to the existence of steric hindrance by


surrounding bonded chains during the grafting-to process.
Comparing with the bonding between the gates and pore surface through
non-covalent interactions, such as van der Waals forces [24], and electrostatic
interaction [25, 26], covalent bonding surface modification is more robust for
application.
In addition, various existing membranes, organic and inorganic ones, have been
utilized as a matrix to fabricate intelligent gating membranes. The organic ones
include polypropylene (PP) [52, 61], polycarbonate (PC) [25, 44, 51, 55], polyethy-
lene (PE) [21, 22, 41], polytetrafluoroethylene (PTFE) [62, 63], Nylon-6 [32, 49,
64], polyvinylidene fluoride (PVDF) [23, 33, 49, 50, 57, 65], polyimide (PI) [30],
polyamide (PA) [66–68], poly(ethylene terephthalate) (PET) [28, 37, 40], polysul-
fone (PSF) [42], poly(viny1 chloride) (PVC) [69], and polyethersulfone (PES) [43,
70, 71], while the inorganic membranes mainly include anodic aluminum oxide
(AAO) [34, 72], nanoporous silica [73], and nanoporous silicon nitride mem-
branes [74]. Generally, inorganic membranes possess great performance of high
thermal and chemical stability, inertness to microbiological degradation, and ease
of cleaning after fouling compared with organic membranes, while they have
high cost.

3.3.2 One-Step Formation Method


Generally, polymeric membranes can be produced by a phase inversion method,
in which solvent is removed from a liquid-polymer solution, leaving a porous,
solid membrane due to the polymer transformation from a liquid phase to solid
phase [75]. Although the one-step formation of stimuli-responsive gating mem-
branes by the phase inversion method have also been reported in the literature,
it is applicable to only a small group of responsive polymers [71, 75, 76].

3.4 Application of Intelligent Gating Membranes


to Environmental Separation
Generally, membrane pore size modulation by external triggers heavily dom-
inates the intelligent membrane research thus far [24, 75, 77]. However, new
designs and concepts are emerging in combining responsive chemistry with
membranes toward better membrane performance. For example, for membrane
fouling, the membrane surface properties can be switched by changing the
wettability of the gates under specific stimuli; thus the affinity between the
contaminants and the membrane surface can be weakened. On the other
hand, stimuli-responsive materials can be combined with RO membranes to
improve their antifouling properties and be incorporated into the membranes
to endow the membrane with self-healing capability [66, 68, 78–82]. In addition,
there is increasing interest in combining membrane distillation (MD) with
photothermal materials, which, in response to solar light, generates heat locally
with high energy efficiency [83]. All-in-one membrane has been reported to
74 3 Filtration Membranes with Responsive Gates

integrate chemical reactions and physical separation in one system, where the
trigger-initiated sequential reactions selectively and on demand degraded and
separated water pollutants [65].
Overall, this chapter presents the state of the art of the intelligent gating
membranes for environmental separation and is organized according to the type
of environmental triggers, namely, temperature (i.e. heat), pH, light, ions, ion
strength, redox, etc., or multiple stimuli.

3.5 Thermoresponsiveness
Filtration membranes responsive to temperature is the main source of intelligent
gating membranes that have found wide interest among researchers. As men-
tioned in Chapter 2, thermoresponsiveness of polymers is mainly based on its
conformation changes below and above critical solution temperature (LCST or
UCST). PNIPAM has been extensively investigated as a thermoresponsive pore
controller in the field of intelligent gating for water filtration [84–86]. Given that
the typical LCST of PNIPAM is around 32 ∘ C, the temperature range of the feed-
water was usually 20–40 ∘ C for PNIPAM-based membranes in literature. At tem-
peratures below the LCST, PNIPAM chains exhibit swollen state and hydrophilic-
ity owing to the hydrogen bonding between the amide groups from PNIPAM and
water molecules, and the gates of the membrane close. With an increase in the
temperature above the LCST, the PNIPAM chains switch to shrunk conforma-
tion and exhibit hydrophobicity due to hydrogen-bonding cleavage, which leads
to the gates open (Figure 3.3) [50, 88].
It was in 1986 when the first intelligent thermoresponsive gating membrane
was constructed. The membrane was made by grafting PNIPAM onto a nylon
membrane, and the grafted polymer acted as permeation valve regulating
water flux by changing water temperature below and above LCST [89]. Ever

Polymer
network

OFF

Microporous
Temp > LCST Temp < LCST support

ON

Figure 3.3 Schematic illustration of thermoresponsiveness mechanism of PNIPAM polymer


grafted onto porous membranes. Source: Lue et al. 2007 [87]. Reprinted with permission of
Elsevier.
3.5 Thermoresponsiveness 75

since, there had been an explosive growth of using PNIPAM in gating mem-
branes [30, 37, 40, 50, 51, 90]. Moreover, the gating membranes with PNIPAM
block copolymers and PNIPAM-based inorganic/organic composites, such
as PNIPAM-g-PET [28, 91], PNIPAM-g-PE [21, 22], PNIPAM-g-PP [52],
PNIPAM-g-nylon [90], PNIPAM-g-PC [87], PNIPAM-modified SiO2 sphere
[31, 45], and PNIPAM-co-glycidyl methacrylate [92], were fabricated by UV or
plasma-induced graft polymerization, ATRP, etc.
In addition, valuable efforts were made to systematically increase and
decrease the LCST by introducing hydrophilic or hydrophobic monomers into
N-isopropylacrylamide (NIPAM) monomer solution in the fabrication of ther-
moresponsive gating membranes, which offers more flexibility in applying intelli-
gent gating membranes to environmental separations. In 2007, Chu et al. grafted
thermoresponsive polymers, poly(N-isopropyl-acrylamide-co-acrylamide)
(PNA) and poly(N-isopropylacrylamide-co-butyl methacrylate) (PNB), as
functional gates onto porous PVDF or nylon membranes via plasma-induced
grafting polymerization method (Figure 3.4) [49]. The response temperature
of copolymers was raised to 40 ∘ C when 7 mol% of hydrophilic acrylamide was
added into the NIPAM co-monomer solution but was reduced to 17.5 ∘ C as
10 mol% of hydrophobic butyl methacrylate was added into the same solution.
However, directly changing the temperature of a bulk water during continu-
ous separation is not trivial and more importantly is considered as energy inef-
ficient. Instead of changing the bulk water temperature to induce the membrane
pore size change, in 2014, Gajda and Ulbricht reported an in situ local heat gen-
eration scheme using an external magnetic field to excite Fe3 O4 nanoparticles
co-imbedded into the membrane pore along with PNIPAM by pre-adsorbing

H2C CH CH2 CH x CH2 CH y


CH2 CH
C O C O C O
+ C O
NH NH NH2
NH2
CH CH
H3C CH3 H3C CH3
(a)

CH3 CH3

H2C CH CH2 C CH2 CH x CH2 C y

C O C O C O C O

NH + O NH O

CH (CH2)3 CH (CH2)3
H3C CH3 H3C CH3
CH3 CH3
(b)

Figure 3.4 The polymerization reactions for grafting thermoresponsive copolymers (PNA and
PNB) onto membrane substrates. Source: Xie et al. 2007 [49]. Reprinted with permission of
Elsevier.
76 3 Filtration Membranes with Responsive Gates

ΔH on

ΔH off

(a)

O O
H H
N N Fe3O4
N N N
(b) H H H

Figure 3.5 (a) Schematic illustration of pore size control through the stimulation of
superparamagnetic Fe3 O4 nanoparticles in PNIPAAM-functionalized membrane pores under
an external magnetic field and (b) the peptide bond between carboxyl groups on the Fe3 O4
nanoparticles surface and prefunctionalized membrane surface. Source: Gajda and Ulbricht
2014 [93]. Reprinted with permission of Royal Society of Chemistry.

a cationic macroinitiator and subsequent photo-initiated graft copolymerization


(Figure 3.5) [93]. Such thermoresponsive polymer–nanoparticle hybrid mem-
branes were endowed with remote control in the water flux via electromagnetic
field “on” and “off.” The pore size and water flux of the membrane could be tuned
between 290 nm, 42 l m−2 h−1 and >400 nm, 240 l m−2 h−1 in the absence and
presence of the magnetic field, respectively.
Apart from PNIPAM, other thermoresponsive polymers with suitable LCST
were also introduced in intelligent gating membranes recently.
In 2005, Petrov et al. utilized thermoresponsive poly(vinyl alcohol-co-vinyl
acetal) copolymer to modify asymmetric porous poly(acrylonitrile) UF mem-
brane [94]. This modification resulted in the reversible opening and closing of
the membrane pores due to the thermoresponsive conformational switch from
shrinking to swelling of the copolymer; thus the membrane permeability and
selectivity can be intelligently regulated by varying the working temperature
from 25 to 45 ∘ C.
In the same year, poly(N-vinylcaprolactam) (PVCL), with an LCST around
32∼35 ∘ C, was also involved in the application of thermoresponsive gating
membranes. Prez and coworkers fabricated a PVCL-modified PET track-etched
membranes via the photochemical immobilization method [95]. The perme-
ability of a mixture of dextrane molecules was investigated as a function of
temperature. The concentration of the dextrane molecules with Mn = 100 000
in the filtrate increased with increasing temperature above its LCST, owing to
the increase in pore diameter for temperatures above its LCST. In addition, the
concentration of the lower molecular weight dextranes (Mn = 6000) remained
constant during temperature change between 20 and 45∘ C, indicating that the
original pore size of the membranes was too high to hinder their permeability
during the filtration process.
3.5 Thermoresponsiveness 77

In 2014, Wessling and coworkers also prepared PVCL microgel-modified


membranes, which exhibited a reversible thermoresponsive behavior [24]. The
membrane’s permeability can be alternatively controlled with water at 20 and
45 ∘ C. Similarly, in 2017, Wessling and coworkers also fabricated an electrically
conductive SiC–C hollow fiber membrane on which PVCL microgels were
immobilized via filtration coating [96]. Differing conventional thermoresponsive
behavior of gating membrane through externally tuning the feed stream temper-
ature, the concept they proposed was to adjust the permeability and selectivity
of the membrane by controlling the applied electrical power to heat membrane
(Figure 3.6a). Therefore, the temperature of the membrane itself directly enabled
to initiate a response of the membrane surface. Direct electric heating was more
energy efficient compared with heating of the whole feed stream, saving 14% of
the consumed energy. The electrical heating caused the microgels to collapse,
increasing the pore size for water permeation. Permeability was around 5–10
times greater in the heated state than without the applied voltage (Figure 3.6b).

(a)
100
P (Lm–2h–1bar–1)

100
100 Wm–1

50 80
Retention 200 kDa (%)

0 Wm–1
Cooling
0 60
0 10 20 30 40 50 60 70

50
Temperature (°C)

40
40 Heating

20
30

20 0
0 10 20 30 40 50 60 70
20 25 30 35 40 45 20
Time (min)
(b) (c) Temperature (°C)

Figure 3.6 (a) Schematic illustration of the electrically heated SiC–C hollow fiber membrane
modified by PVCL microgels. (b) Permeability over time for each 10 min cycles of heating with
30 W m−1 (upper) and 100 W m−1 (bottom) and no applied voltage. (c) Selectivity of 200 kDa
dextrane as a function of applied temperature on the membrane. The temperature was
increased to 50 ∘ C and then cooled back to room temperature. Source: Lohaus et al. 2017 [96].
Reprinted with permission of Elsevier.
78 3 Filtration Membranes with Responsive Gates

The selectivity can be reversibly tuned in a range of 10–80% for a 200 kDa
dextrane by heating the membrane (Figure 3.6c).
In 2016, Jin and coworkers grafted pyrene-terminated poly-(MEO2 MA-co-
OEGMA) (LCST ∼ 32 ∘ C) onto single-walled carbon nanotube (SWCNT) mem-
brane via π−π interaction [15]. The membrane’s pore size varied between 12 and
14 nm when the water temperature was changed between 25 and 40 ∘ C, leading
to a stable flux variation between 3730 and 6430 l m−2 h−1 over several cycles.
In comparison, the use of UCST response mode polymers in intelligent
gating membrane is very limited. Thus, it is not a surprise that the first UCST
behavior of the gating membrane with practical responsive temperature was not
reported until 2005 when Chu et al. demonstrated interpenetrating polymer
networks with poly(acrylamide) (PAAM) and poly(acrylic acid) (PAAC) as
negatively thermoresponsive gates [97]. As shown in Figure 3.7, the intelligent

Pore
Grafted PAAM PAAM/PAAC IPN
Substrate Substrate

(a) (b) (c)

T > UCST

T < UCST

CH2 CH CH2 CH CH2 CH CH2 CH CH2 CH CH2 CH


PAAC
C C C C C C
O O O O O O O OH O OH O OH
H H H H H H
HN O HN O HN O H2N O H 2N O H 2N O
C C C C C C
CH2 CH CH2 CH CH2 CH PAAM CH2 CH CH2 CH CH2 CH

(d) (e)

Figure 3.7 A schematic illustration of the functioning of the negative thermoresponsive


membrane. The functional gates of the membrane are thermoresponsive interpenetrating
polymer networks consisting of poly-acrylamide (PAAM) and poly(acrylic acid) (PAAC), in
which the volume phase transition is driven by the hydrogen bonding interactions between
the molecules. (a) Porous membrane substrates. (b) Membranes with grafted PAAM gates.
(c) Membranes with PAAM/PAAC-based gates. (d) The pores of the membranes are open since
the functional gates are under their shrunken state at temperatures below the UCST as a result
of PAAM/PAAC complex formation by hydrogen-bonding interactions. (e) The membrane
pores are closed because the gates are under their swollen state at temperatures above the
UCST as a result of complex dissociation by the breakage of hydrogen bonds. Source: Chu et al.
2005 [97]. Reprinted with permission of John Wiley and Sons.
3.5 Thermoresponsiveness 79

gates exhibited open state due to the shrinkage of PAAM/PAAC complex via
hydrogen-bonding formation at temperatures below UCST of the complex.
While at temperatures above the UCST, the PAAM/PAAC complex would
swell owing to its dissociation through the breakage of hydrogen bonds, leading
to pore “closing.” Thus, the membrane pores can switch from an “open” to a
“closed” state once the temperature increases above the UCST. The synthesized
membrane exhibited a sharp transition of water permeability in a practical
temperature range from 20 to 25 ∘ C.
Inspired by the stomatal closure feature of plant leaves at relatively high temper-
ature, in 2017, Zhao and coworkers constructed a negative temperature-response
nano-gating membrane by covalently grafting PNIPAM chains on GO sheets via
free radical polymerization (Figure 3.8a) [98]. By virtue of the temperature tun-
able lamellar spaces of GO sheets, the water permeation of this membrane was
12.4 l m−2 h−1 bar−1 at 25 ∘ C and 1.8 l m−2 h−1 bar−1 at 50 ∘ C. Moreover, such
membrane was capable of separating multiple molecules with different sizes by
regulating the temperature. The rejection rates of the five ions/molecules, Cu2+
(0.8 nm), [Fe(CN)6 ]3− (0.9 × 0.9 nm), rhodamine B (RB, 1.8 × 1.4 nm), coomassie
brilliant blue (CBB, 2.7 × 1.8 nm), and cytochrome c (Cyt. c, 2.5 × 2.5 × 3.7 nm),
through the gating membrane from 25 to 50 ∘ C were measured (Figure 3.8b).
Cu2+ only existed in the filtrate obtained at 50 ∘ C, while trace RB and Cyt. c were
obtained in the filtrate at 50 ∘ C. For RB, it enriched in the filtrate obtained at

PNIPAM and water molecule Intramolecular/intermolecular


hydrogen bonding hydrogen bonding

> LCST

PNIPAM < LCST

Open Closed
(a)
100
Cu2+
Feed Retentate Filtrate Filtrate Retentate
[Fe(CN)6]3–
Rejection rate (%)

80 solution at 50 °C at 50 °C at 25 °C at 25 °C
RB
CBB
60 Cyt. c
Mixed RB + Cyt. c Cu2+ RB Cyt. c
40

20

0
25 30 35 40 45 50
(b) Temperature (°C) (c)

Figure 3.8 (a) Fabrication process of the thermoresponsive membrane and its water gating
property. The water permeance of membrane could be regulated by changing temperature
(T). (b) Temperature-dependent rejection rates of membrane to Cu2+ , [Fe(CN)6 ]3− , RB, CBB, and
Cyt. c, implying smaller channel size of the membrane with the increasing temperature.
(c) Separation of mixed molecules solution, containing Cu2+ , RB, and Cyt. c. Source: Liu et al.
2017 [98]. Reprinted with permission of Springer Nature.
80 3 Filtration Membranes with Responsive Gates

25 ∘ C, and only the largest Cyt. c remained in the retentate obtained at 25 ∘ C


(Figure 3.8c). This nano-gating membrane expands the scope of intelligent gating
systems and molecular separation.

3.6 pH Responsiveness
Among the various categories of intelligent responsive materials, pH-responsive
polymeric materials are one of the most widely investigated. The polyelectrolytes
with weak acidic or weak basic groups are typically pH-responsive polymers.
Depending on solution pHs, the weak acidic and basic groups undergo reversible
protonation and deprotonation, leading to a reversible swollen and shrunken
conformation transition due to on-and-off switch of electrostatic repulsion
between these functional groups. The pH-dependent conformation changes
of the polymers have been widely used as functional gates in the intelligent
gating membranes for controllable separation. Such gating membranes have
been used toward adjustable water flux and molecular size selectivity for a vari-
ety of substances, including proteins [99], fluorescein isothiocyanate–dextran
(FITC–dextran) [71], macromolecules [25, 100], vitamin B12 [41, 101], riboflavin
[42], Au nanoparticles [102], etc.

3.6.1 Polybase Gating Membranes


The polymers with weak basic groups that have been applied to intelligent
gating membranes include poly(4-vinylpyridine) (P4VP) [16, 17], polystyrene-
b-poly(4-vinylpyridine) (PS-b-P4VP) [99, 102–104], poly(methyl methylacrylate-
co-4-vinyl pyridine) (P(MMA-4VPy)) [105], poly(2-vinylpyridine) (P2VP) [106],
PDMAEMA [33, 107, 108], and poly(allylamine hydrochloride) (PAH) [25].
Under suitable and generally acidic pH conditions, the weak basic groups
located in the side chains of these polymers accept protons leading to the
electrostatic repulsion among positively charged basic groups, exhibit a swollen
state in an acidic environment, and thus result in pore size reduction. In a basic
condition, the same basic groups deprotonate and are charge neutral, and the
polymers go back to their shrunken state, leading to the increased pore size of
the membranes.
In 1984, Okahata et al. fabricated the first pH-responsive gating membrane,
and in this work, P4VP was grafted onto a porous nylon membrane to act as
NaCl permeation valve between pH 2 and 9 [17]. In 1995, Childs and coworkers
further grafted P4VP onto microporous PP and PE substrates for pH-responsive
gating membranes. The P4VP was stabilized on the PE or PP porous substrate
by UV-initiated grafting. Such membrane showed the pH valve effect and a large
permeability change, as well as the selectivity of small inorganic ions, which mod-
erately rejected NaCl (40–50%) at pH < 3 and showed no salt rejection in neutral
or basic conditions [16].
Although the surface graft polymerization provides a useful method to create
pH-responsive gating membranes, its imprecise control of the density of grafted
3.6 pH Responsiveness 81

800 1
Pore diameter (nm)

0.8

Cfiltrate/Cfeed
600
0.6
400 0.4

200 0.2
0
0 Feed PEO PEO Acidic PEO
0 8.5 14.5 20.5 solution solution solution solution
Number of bilayers filtered in filtered in filtered in
open closed closed state
(a) (b) state state

Figure 3.9 (a) Changes in the average pore diameters in the pores of LbL membrane.
(b) Filtration of high molecular weight PEO (0.01 g dl−1 ) using 18.5 bilayer (PAH/PSS) LbL
membrane in different conditions. Open and closed states were attained by the pretreatment
of multilayer-modified membranes at pH 10.5 and 2.5, respectively. Source: Lee et al. 2006 [25].
Reprinted with permission of American Chemical Society.

chains and inability to control the pore size in the confined geometries of
the existing membranes are still challenges [109, 110]. In 2006, Rubner and
coworkers used layer-by-layer (LbL) method and assembled multilayers of PAH
and poly(sodium 4-styrenesulfonate) (PSS) as intelligent gates in a nanoporous
PC membrane. The method of LbL offered an advantageous capability of easy
and precise control over the pore diameters of the modified porous membrane
[25] (Figure 3.9a), and the functionalized membrane showed approximately 80%
poly(ethylene oxide) (PEO) rejection at pH 2.5 and no rejection at all at pH 10.5
(Figure 3.9b).
In 2007, Peinemann’s group reported a fast one-step procedure to prepare a
block copolymeric (PS-b-P4VP) membrane with nanometer-sized ordered pores
by using non-solvent-induced phase separation [75]. In 2011, the same group
developed metal-block PS-b-P4VP NF membrane with monodisperse asymmet-
ric pH-responsive nanochannels. This NF membrane possessed high pore density
(greater than 2 × 1014 per m2 ) and large-scale reproducibility in m2 scale. The
reported pH-sensitive range of the polymer nanochannels with synthetic pores
in the nm scale was from 2 to 8 (Figure 3.10), leading to more than 2 orders of
magnitude water flux increase and selectivity of the mixture of PEG with different
molecular weights [103].
Later, some follow-up works on PS-b-P4VP-based UF membranes were
conducted for controllable separation of protein and inorganic/organic
molecules [99, 104, 111, 112]. In 2017, Walker and coworkers fabricated
phosphotriesterase-functionalized poly(isoprene-b-styrene-b-4-vinylpyridine)
nanoporous membranes via self-assembly and non-solvent-induced phase
separation [113]. This gating membrane showed changeable permeation from
1522 l m−2 h−1 bar−1 (at pH = 7) to 11 l m−2 h−1 bar−1 (at pH = 3) as a result
of conformation transition of the P4VP chains between a collapsed state and
swollen state (Figure 3.11). In addition, integrating enzymatic recognition
capability into pH-responsive gating membranes may offer an intriguing outlook
82 3 Filtration Membranes with Responsive Gates

pH 2 pH 10

300 nm 300 nm
(a)

pH 2 pH 10

1 μm 1 μm Dry
(b)

Figure 3.10 (a) Cryo-field emission scanning electron microscopy and (b) environmental
scanning electron microscopy of PS-b-P4VP membranes at pH ∼ 2 and 10; (b, right) dry
membrane was observed by environmental scanning electron microscopy. Source: Nunes
et al. 2011 [103]. Reprinted with permission of American Chemical Society.

z 2500
z Neat ISV117
+
Average permeability (lm–2 h–1 bar–1)

+H Supported ISV117
2000 Neat ISV119
–H+ N Supported ISV119
N H
1500

1000

Collapsed neutral P4VP 500

Swollen charged P4VP


0
Pl-b-PS membrane matrix
2 4 6 8
(a) (b) pH

Figure 3.11 (a) pH-responsive behavior of the P4VP-based membranes. (b) The permeability
of membranes in a buffer solution as a function of pH. Source: Poole et al. 2017 [113].
Reprinted with permission of John Wiley and Sons.
3.6 pH Responsiveness 83

for engineered biomimetic materials responding to a complex range of external


parameters, providing sensing, protection, and remediation capabilities.

3.6.2 Polyacid Gating Membrane


The typical weak acidic polymers that have been reported in the intelligent
gating membranes include PAA [42, 55, 57, 65, 100], poly(methacry1ic acid)
(PMAA) [17, 41, 108, 114], poly(glutamic acid) (PGA) [62, 115], poly(l-glutamic
acid) (PLGA) [116], polystyrene-block-poly(acrylic acid) (PS-b-PAA) [71], and
poly(methyl methacrylate-co-acrylic acid) (P(MMA-AA)) [105], among others.
The weak acidic polymers have a pH-responsive behavior opposite to the weak
basic polymers.
The intermolecular hydrogen bonding among the carboxylic groups of the
polymers leads to the shrunken conformation of polymer chains under acidic
conditions, while under basic conditions, the carboxylic groups deprotonate and
become charged, and the electrostatic repulsion among deprotonated carboxylic
groups swells the polymer chains, leading to the reducing membrane pore size.
The first weak acidic polymeric (i.e. PMAA) gating membrane was reported
in 1984 by Okahata et al. [17]. Between 1996 and 1999, Lee et al. fabricated
PAA-grafted polymeric membranes by plasma [57] and UV-irradiated graft
polymerization method [42], respectively, all showing a decreased riboflavin
permeability in pH 4–5 compared with lower pH values.
Since 1992, Ito et al. had demonstrated several methods in making weak
basic polymer-based intelligent gating membranes by self-assembly or surface
graft polymerization of weak polyacids, including PMAA, PAA, and PLGA
onto PC, PTFE, and gold-coated membranes [62, 100, 109, 114, 115, 117]. In
2001, Zhang and Ito assembled PAA weak polyelectrolyte with thiol-modified
chains on a gold-coated PC porous membrane [100], and the water flux of the
membrane with surface densities of self-assembled PAA-SH of 18 pmol cm−2
decreased approximately from 5 to 1 ml cm−2 h−1 when pH increased from
2 to 7. The transport of PEG (Mw ∼ 8000) was also dependent on pH value
(polymer retention: 0% at pH 2 and 33% at pH 7).
In 2006, Qu et al. designed a pH-responsive intelligently controlled release sys-
tem that contained weak polyacid, PMAA-g-PVDF as pH-responsive valve, and a
cross-linked PDMAEMA hydrogels as a pump to pump the substances out [108].
As pH was decreased from 7 to 2, the membrane gates opened and the hydrogels
expanded, leading to a rapid release of a substance into the membrane system
by pumping out effect (Figure 3.12), which has potentials to be used as chemi-
cal carriers and environmental sensors as well as environmental separation. In
2014, Chu and coworkers reported the fabrication of pH-responsive PES com-
posite membranes blended with PS-b-PAA copolymers, and the membrane pores
opened at pH = 3 (<pK a ) and closed at pH = 8 (>pK a ), which showed controlled
separation of FITC–dextran [71].
In 2009, Li and coworkers reported pH-dependent MF membranes that were
cast by PES-grafted poly(methacrylic acid) (PES-g-PMAA) via phase inversion
method. The pore sizes and filtration ability of the membranes can be tuned by
changing pH [118]. In the most extreme case, the flux of aqueous acidic solution
84 3 Filtration Membranes with Responsive Gates

CH3 PDM CH3 PMAA


CH2 C CH2 C
n n
C O C O

O CH2 CH2 N CH3 OH

CH3
Swelling ratio

Swelling ratio
pH pH

pH < min(pKPMAA,pKPDM)

pH > max(pKPMAA,pKPDM)

Figure 3.12 Pumping systems with gating membranes containing pH-responsive gates for
enhanced controlled release. Source: Qu et al. 2006 [108]. Reprinted with permission of John
Wiley and Sons.

(pH = 1) was about four times as that of aqueous basic solution (pH = 9) through
the PES-g-PMAA membrane. Similarly, PES-g-PMAA was also cast into UF
membranes via the phase inversion process, showing reversible pH-sensitive
permeability as the pH value of feed solution was varied between 2.0 and
10.0 (Figure 3.13a) [119]. Furthermore, by using the same fabrication method,
PS-b-PAA was blended with PES membranes with nanoscale pores [71]. Such
membranes enabled selectively sieving of FITC–dextran molecule mixtures
with different molecular weights of 10, 40, and 70 kDa, at ambient pH values of
3 and 8 (Figure 3.13b).
In 2011, Lewis et al. applied a pH-responsive intelligent gating membrane
in a multilayered and all-in-one Fenton-reaction-active filtration system for
advanced oxidation [65]. The top layer of the membrane contained glucose
oxidase (GOx) for in situ H2 O2 generation by reacting with deliberately added
glucose in the raw water, which allowed for the flexibility of on-demand initiation
of the Fenton reaction. The bottom porous PVDF layer was functionalized with a
pH-responsive PAA network, and iron species was immobilized in the PAA layer
as catalysis for Fenton reaction. The H2 O2 generated in the top layer decomposed
3.7 Photo-responsiveness 85

100 100
pH 2.0 PSA280-1.5 membrane
90 90
pH = 3
80 80 pH = 8

Solute rejection (%)


70 70
Flux (I (m2h–1))

60 60

50 50
40
40
30
30
20
20
pH 10.0 10
10
0
0 1 2 3 4 5 6 7 8 9 10 11 12 10 40 70
(a) Cycle time (b) Molecular weight of FITC-dextran (kDa)

Figure 3.13 (a) Reversible change of water permeation through PES-g-PMAA membrane as a
function of pH. Source: Shi et al. 2010 [119]. Reprinted with permission of Elsevier.
(b) FITC-dextran rejection of pH-responsive membrane under pH = 3 and pH = 8 conditions.
Source: Luo et al. 2014 [71]. Reprinted with permission of Elsevier.

and generated free radical oxidants with the help of iron species to oxidize the
organic containments in the feedwater and the degradation-generated alkali
ions as by-products. The alkali ions, in turn, increased the pH and stimulated
the expansion of PAA, leading to a decrease in water flux and thus a longer
residence time for pollutant degradation. On the other hand, in the case with
the feedwater being free of organic contaminants, the water passed through the
membrane with a large flux. Therefore, this all-in-one reactive filtration system
possesses responsive and self-initiated intelligent behaviors.

3.7 Photo-responsiveness
In photo-responsive gating membrane system, photo-responsive polymers
act as functional gates, tuning membrane permeability and selectivity mainly
through the reversible conformational and/or structural changes of the polymers
in response to photoexcitation. Among various photoresponsive polymers,
azobenzene and spiropyran derivatives have been highly investigated as
photosensitive gates.
For azobenzene-based intelligent membrane, upon UV irradiation, the dis-
tance between the para carbon atoms from the azobenzene functional groups
decreases, leading to increased membrane pore size, while the membrane pore
size decreases under visible light irradiation. On the other hand, the spiropyran
groups undergo a transition from nonpolarity to polarity upon exposure to UV
light, leading to changes from shrunken to swollen state and corresponding
to a reduced pore size of the intelligent membrane. The polar state returns
to the nonpolar and hydrophobic state via either a thermal or visible light
treatment [43].

3.7.1 Azobenzene-Based Gating Membranes


Depending on the trans–cis isomerization transition under UV or visible light
irradiation, the synthetic azobenzene-based photo-responsive membrane can
86 3 Filtration Membranes with Responsive Gates

Responsive molecules + inorganic hosts Figure 3.14 Organization of


stimuli-responsive ligands
within a 3D porous
framework imparts is useful
for opening and closing a
nanoscale valve. Source: Liu
et al. 2003 [73]. Reprinted
with permission of John
Wiley and Sons.

Stimuli

effectively tune its pore size (commonly nanoscale size), which has been applied
in the controllable separation of ions [19, 26, 120], dye molecules [121], and
macromolecules [18].
In 1983, Anzai et al. pioneered azobenzene-functionalized PVC membrane
for the photo-controlled K+ ion permeation [69]. In 2003, Liu et al. organized
azobenzene-containing ligands, 4-(3-triethoxysilylpropylureido)azobenzene
(TSUA), into an ordered and rigid inorganic silica framework by an
evaporation-induced self-assembly procedure (Figure 3.14). The photo-induced
trans–cis conformation change of azobenzene ligands transduces solar energy
into mechanical work to further promote the operational stability, enabling
photo-control over its pore size [73].
In 2014, Shi et al. designed a photo-responsive system based on the host–guest
complex between azobenzene (Azo) and 𝛽-cyclodextrin (𝛽-CD) [18]. Under
irradiation at 450 nm, trans-azobenzene inserted into 𝛽-CD and formed 1 : 1
inclusion complex, resulting in stably closed pores. Under irradiation at 365 nm,
collapsed cis-Azo slid out of the cavity of 𝛽-CD, and thus the pores were opened,
showing high permeability for pure water and PEG solutions (Figure 3.15).
In 2015, Fujiwara and Imura reported a special responsive gating membrane for
water desalination by grafting azobenzene on an AAO membrane and using UV
and visible light as a gate switch [19]. The membrane blocked the water passage in
darkness but allowed water vapor passing through when simultaneously exposed
under UV and visible light. The simultaneous irradiation of UV and visible lights
onto the azobenzene induced its consecutive motion between the trans and cis
isomers, which promoted the water vapor permeation. Since only water perme-
ated through the membrane, the membrane was utilized for water treatment to
remove dye and protein. This water membrane separation process could be also
3.7 Photo-responsiveness 87

= PEG-600
365 nm
= PEG-4000
450 nm

= PEG-10 000

Figure 3.15 Schematic of PEG molecules (Mn = 600, 4000, 10 000) permeated through the
membranes. Source: Shi et al. 2014 [18]. Reprinted with permission of Elsevier.

Figure 3.16 A seawater


desalination system using
azobenzene-modified AAO
membrane and solar light Direct use of solar light
energy. Source: Fujiwara and
Imura 2015 [19]. Reprinted Solar light
with permission of American Seawater
Seawater
Chemical Society.
pool Azobenzene-
modified
anodized
alumina
membrane
Drain

Freshwater

applied to seawater desalination (Figure 3.16). As 3.5% sodium chloride solution


was employed as model seawater, the salt content of the permeated water through
the membranes was less than 0.01%, substantially reaching the level of drinkable
freshwater.
In 2017, Fujiwara further improved the membrane by using a visible light-
responsive dye, disperse red 1 (DR1), to replace azobenzene, which permitted
solely visible light responsiveness [122]. By simultaneously grafting DR1 and blue
14 (DB14) on a PTFE membrane, the membrane was responsive to a wider range
of light spectrum [123]. The flux of the double-dye-modified PTFE membrane
was higher than the single-dye-modified PTFE membrane.

3.7.2 Spiropyran-Based Gating Membranes


Compared with azobenzene, spiropyran-based polymer chains exhibit greater
volume change upon UV irradiation and thus are investigated more with intelli-
gent gating membrane for controllable permeation of water [124], organic solvent
[63, 125], protein molecules [43], and caffeine [59, 60].
88 3 Filtration Membranes with Responsive Gates

In the mid-1920s, Fisher and Hirshberg observed the photochromic character-


istics and the reversible reaction of spiropyrans for the first time and thus set in
motion the research on spiropyrans [126]. In 1994, Ito and coworkers pioneered a
spiropyran-containing methacrylate and acrylamide-functionalized PTFE mem-
brane, whose pore size and permeability for H2 O/CH3 OH were tuned by UV and
visible light irradiation [63]. Four years later in 1998, the same group fabricated
spiropyran-containing PMMA-grafted glass filter for controllable permeation of
toluene liquid, which showed increased flux by UV irradiation and decreased flux
under visible irradiation because the copolymer chains in toluene shrank under
UV irradiation but swelled under visible irradiation [125].
In 2006, Belfort and coworkers prepared spiropyran-grafted PES UF membrane
through a UV-induced graft polymerization method (Figure 3.17a). As-prepared
membrane exhibited an optically reversible switching behavior due to the switch-
able conformation of spiropyrans in response to light (Figure 3.17b). The mem-
brane had 17% lower permeation flux of phosphate-buffered saline (PBS) solution
under the visible light as compared with UV light (254 nm) irradiation [43].
Very recently, Padeste and coworkers demonstrated a two-step approach to
prepare a photo-responsive gating membrane: grafting PMAA polymer brushes
onto porous PP membrane, followed by covalently attaching spiropyran moieties
to the grafted PMMA polymer brushes with the lowest grafting level of PMMA

UV300 Visible light UV254


Vinyl
monomer

UV254

Vis

Radical sites Poly(ether sulfone) membrane Graft polymer Graft polymer


(closed) (open)
(a)
NO2

UV
N O NO2
Vis N + –O

NH
O NH
O

NH
NH
O
O

(b) Spiropyran (closed) Merocyanine (open)

Figure 3.17 (a) Schematic of graft polymerization and the switch of spiropyran. (b) The
chemical structure of the vinyl spiropyrans in two configurations as a function of UV–Vis
irradiation. Source: Nayak et al. 2006 [43]. Reprinted with permission of John Wiley and Sons.
3.8 Metallic Ion Responsiveness 89

polymer brushes (17% ± 6%). The water flux of photo-responsive membrane


decreased by 40% from 405 (visible light with 30 minutes) to 289 l m−2 h−1 (UV
light with 30 seconds) [124].

3.8 Metallic Ion Responsiveness


Since the crown ethers are able to recognize their matching ions, and amazingly,
this selective ion capture process causes a positive and negative shift in the LCST
of poly(NIPAM-co-AAB18 C6 ) and poly(NIPAM-co-AAB15 C5 ), respectively, this
type of copolymers could be integrated into porous membranes for controllable
“closed” or “open” pore switching, which has been employed in intelligent gating
membranes for controllable separation [127–131].
In 1993, poly(NIPAM-co-AAB18 C6 ) was first synthesized [132], which, in
response to K+ in aqueous solution, underwent conformation transition at
32 ∘ C. In 1999, Yamaguchi et al. reported Ba2+ -responsive gating membrane by
grafting poly(NIPAM-co-AAB18 C6 ) onto PE membrane [133]. Two years later,
the same group further demonstrated that the pore size of the membrane could
be changed from 5 to 27 nm with varying Ba2+ concentrations between 0 and
0.014 M, leading to the tunable rejection of dextran molecules with size ranging
between 2.7 and 27.2 nm by the membrane [129].
In 2013, Chu and coworkers developed an ion-responsive membrane that
selectively detected and removed Pb2+ from wastewater [46]. Poly(NIPAM-co-
AAB18 C6 ) copolymer chains as functional gates were grafted onto Nylon-6 mem-
brane substrate via a two-step method combining plasma-induced pore-filling
grafting polymerization and chemical modification (Figure 3.18a–c). By simply
tuning the operation temperature, the effective removal of Pb2+ and membrane
regeneration could be realized (Figure 3.18d–i). It was found that, in the presence
of 10−5 M Pb2+ , the pore size of the membrane decreased from 159 to 94 nm,
leading to a reduced permeation as low as 0.35 g min−1 cm−2 . The phenomenon
of an isothermal reduction of solution flux could be used as an informational
signal for detection of trace Pb2+ ions in the environment. This gating membrane
offers a promising opportunity in industrial and agricultural applications, such
as online detection and timely treatment of trace Pb2+ ions in wastewater
discharge, analysis for water quality, and remediation and protection of soil.
The pendent 15-crown-5 moieties recognize and capture majorly K+ ions,
which leads to a negative shift in the LCST values of the copolymers. Namely,
once K+ ions are removed from the crown ether receptors, the copolymer chains
swell and the functional gates close. Before 2008, the K+ -responsive gating
membranes were dominantly constructed from poly(NIPAM-co-AAB18 C6 ) as
discussed earlier [128–130, 133–137].
In 2008, poly(NIPAM-co-AAB15 C5 ) was developed as a new candidate material
for ion-responsive intelligent membrane system [127]. Later, Chu et al. grafted
poly(NIPAM-co-AAB15 C5 ) onto Nylon-6 membranes via plasma-induced
pore-filling grafting polymerization and chemical modification and found that
the grafted gates in the membrane pores spontaneously tuned the solution flux
90 3 Filtration Membranes with Responsive Gates

(a) (b) (c)


H2C CH x H2C CH y H2C CH x H2C CH y
O C O C O C O C
NH OH NH NH
CH CH
H3C CH3 H3C CH3 O
O O
O O
O

(d)

T = T1

(f) Pb2+ detection Pb2+ separation (h)

(e)
T = T2 T = T1
Swelling ratio

Pb2+ Pb2+
presence removal

(g) (i)
T1 LCSTa T2 LCSTb T3
Temperature

T = T2 T = T3

Figure 3.18 Schematic illustration of the preparation process (a–c) and the proposed concept
of the smart membrane with functional gates for detection and removal of trace Pb2+ ions
(d–i). Source: Liu et al. 2013 [46]. Reprinted with permission of Royal Society of Chemistry.

in response to K+ ion in permeate. As shown in Figure 3.19, in the presence


of 0.1 M K+ in permeate at 25 ∘ C, the flux increased to 245.4 kg m−2 h−1 (pore
size ∼ 118 nm) in comparison with the pure water flux of 4.5 kg m−2 h−1 (pore
size ∼ 43 nm) [138]. Other coexisting ions, such as Na+ , Ca2+ , and Mg2+ , did not
impose any influence on the membrane flux.
3.9 Redox Responsiveness 91

Mn+: K+ Na+ Ca2+ Mg2+ 120 d = 118 nm


300
Water 0.1M Mn+ Water 0.1M Mn+ Water 0.1M Mn+
100
250

Pore diameter (nm)


80 d
J (kg m–2 h–1)

200
d = 58 nm d = 59 nm d = 58 nm
150 60
d = 43 nm
100 40

50 20

0
0 60 120 180 240 300 360 0
Time (min) Water Na+ Ca2+ Mg2+ K+
(a) (b)

Figure 3.19 (a) Isothermally dynamic changes in solution flux of the poly(NIPAM-co-AAB15 C5 )-
grafted membrane in pure water and aqueous solutions containing different metal ions and
(b) the changes of estimated pore size of the grafted membrane in pure water and aqueous
solutions containing different metal ions. Source: Liu et al. 2012 [138]. Reprinted with
permission of John Wiley and Sons.

3.9 Redox Responsiveness


Redox-sensitive groups such as dithienylethenes [139], ferrocene [140], pyridine
[141], or disulfides [142] could be incorporated into polymer structures, initiating
the redox reactions and thus resulting in swelling/shrinking conformation tran-
sition of the polymers. In the oxidized state, the functional molecules are ion-
ized and water soluble and exhibit a swollen conformation due to their charges,
while, in the reduced state, they are deionized and insoluble, corresponding to a
shrunken conformation.
The first redox-responsive gating membrane was reported in 1997 by Ito
et al. [141], who fabricated a poly(3-carbamoyl-1-(p-vinylbenzyl)pyridinium
chloride) (PCVPC)-grafted PE/PTFE microporous membrane for controlling
the water permeation. In the oxidized state, pyridine groups were ionized, and
thus the grafted polymer chains stretched inside the membrane pores, leading
to decreased water permeation.
In 2014, Brunsen and coworkers reported a hybrid membrane with redox-
controlled gates by combining a mesoporous silica template and two different
redox-responsive ferrocene-containing polymers, polyvinylferrocene (PVFc) and
poly(2-(methacryloyloxy)ethyl ferrocenecarboxylate) (PFcMA), via grafting-to
and grafting-from method, respectively [143].
In the same year, Vancso and coworkers developed the polyelectrolyte mem-
brane containing redox-active PFS-based poly(ionic liquid) (PIL) and PAA [144].
The porous membrane, in the presence of Fe(ClO4 )3 or ascorbic acid, showed
significantly porous structure changes in the oxidized and reduced states, lead-
ing to a reversible switch of permeability. The average flow rate of pure water was
92 3 Filtration Membranes with Responsive Gates

0.10

0.08
Flux (ml cm–2 s–1)

0.06

0.04

0.02 Re

Ox
0.00
1 2 3 4 5 6
Number of cycles

Figure 3.20 The reversible switching of the flow of pure water for the oxidized and reduced
porous membranes. Source: Zhang et al. 2014 [144]. Reprinted with permission of John Wiley
and Sons.

0.092 ± 0.004 ml cm−2 s−1 in the oxidized state and 0.064 ± 0.005 ml cm−2 s−1 in
the reduced states, respectively (Figure 3.20).

3.10 Ion Strength Responsiveness


Among the categories of responsive gating membranes, the design of intelligent
materials that respond to the nature and strength of the salt ions in their sur-
rounding environment is promising as filtration membranes for water desali-
nation and wastewater treatment. Zwitterionic polymers are popularly used as
intelligent responsive gates due to the conformational changes of their structures
on the pore surface depending on the strength of ions such as NaCl. As shown
in Figure 3.21, the “open” and “close” of the membrane pore can be effectively
controlled through changing the ion strength. At a low ion strength, responsive
membrane possesses an open pore state owing to inter- and/or intrachain asso-
ciation of oppositely charged groups, leading to a collapsed chain conformation.
Upon addition of sodium chloride salt, sodium and chloride ions disrupt these
electrostatic interactions, and thus the membrane pore changes into closed state
as the polymeric chains expand [145].
In 1984, Okahata et al. firstly reported the ion strength-responsive ultrathin
nylon capsule membranes by using zwitterionic polymer [146]. In addition, Kang
and coworkers synthesized a new graft zwitterionic copolymer, through the
graft copolymerization of the zwitterionic N,N ′ -dimethyl(methylmethacryloyl
ethyl)ammonium propanesulfonate (DMAPS), with PVDF backbone [147].
Then the responsive MF membranes were cast by phase inversion in aqueous
media of different electrolyte concentration and temperature. The permeability
of aqueous solutions through the PDMAPS-g-PVDF-based MF membranes
3.10 Ion Strength Responsiveness 93

Salt solution

Water

+ – + Cl–
– Na+ Cl– – + Na+
– + Cl–Na+
– + + – + –
+ –
– + – Cl
+ – Na+ Cl– – Cl– Na+
Na+ Cl
– + – + – + – + + –
Interchain Intrachain
association association
“Open” pore “Closed” pore

Figure 3.21 Schematic illustration of the conformational states of zwitterionic polymeric


chains with increasing salt ion strength. Source: Zhao et al. 2016 [145]. Reprinted with
permission of Elsevier.

exhibited a dependence on the strength of salt ion, which decreased with


increasing the aqueous NaCl solution concentration from 10−7 to 10−1 mol l−1 .
Furthermore, ion strength-responsive membranes also exhibited antifouling
and self-cleaning properties in Meng’s research (2014). A commercial RO
membrane was modified by zwitterionic polymer, poly(4-(2-sulfoethyl)-1-(4-
vinylbenzyl) pyridinium betaine) (PSVBP), and this sodium chloride-responsive
membrane could reject the protein foulants with a 90% flux restoration by the
cleaned membrane [148].
In 2016, Zhu and coworkers fabricated a block copolymeric blend membrane
consisting of PES and amphiphilic PES-block-poly(sulfobetainemethacrylate)
(PES-b-PSBMA) via non-solvent-induced phase separation process [145]. In
addition, the water permeability and molecule sieving of such blend membranes
could be tuned by changing the salt concentration in aqueous solution. The BSA
rejection increased from 75% to 92% with the increasing NaCl concentration
from 0 to 0.1 M.
Poly(ionic liquids) (PILs) have also been applied in an ion strength-responsive
gating membranes. In 2017, Zhao and coworkers fabricated a PILs-modified PES
membrane via an in situ cross-linking copolymerization method [149]. The flux
of the aqueous NaCl solution (0.6 M) through the membrane was 26 times higher
than the flux of pure water through the membrane. Meanwhile, this membrane
exhibited responsiveness toward both anion species, including PF6− , BF4− , and
SCN− , and their strength.
In 2018, Huang and coworkers constructed intelligent supramolecular
mesoporous materials with salt ion strength-induced functional gates for
94 3 Filtration Membranes with Responsive Gates

selective adsorption and release of organic pollutants from water [150]. Hence,
propeller-shaped aromatic amphiphiles that aggregated into hollow spheres
were obtained by a self-assembly method. These pores from propeller-shaped
aromatic assembly provide an excellent hydrophobic characteristic with the
high surface area to remove organic micro-pollutants from wastewater. Corre-
spondingly, the removal efficiency was 92% and 90% for ethinyl estradiol (EE)
and bisphenol A (BPA), respectively. Additionally, as shown in Figure 3.22,
the folded architecture of propeller tends to be flattened by the salt addition,
resulting in a transition from porous to nonporous behavior. Accordingly, most
of the removed pollutants (EE and BPA) are able to be released by the dynamic

OR RO RO N OR
N N

OR RO
1

OR RO

RO OR

N N
N N

OR 2 RO OR 3 RO

(a) R=

NaCl

(b)

Figure 3.22 (a) Molecular structures of 1, 2, and 3. (b) Schematic of porous and nonporous
materials from conformation-tunable propeller assembly. Source: Xie et al. 2018 [150].
Reprinted with permission of John Wiley and Sons.
3.11 Dual and Multi-Stimuli Responsiveness 95

porous assembly, and subsequently, dialysis triggers the porous materials to be


recovered.

3.11 Dual and Multi-Stimuli Responsiveness


So far, the discussed intelligent gating membranes only respond to a single stimu-
lus. Actually, there have been intelligent materials that can respond to more than
one stimulus either simultaneously or independently.

3.11.1 pH and Temperature Dual Responsiveness


Among the multi-stimuli-responsive materials, the dual stimuli mode of pH and
temperature has been investigated and introduced into intelligent gating mem-
branes. Commonly, PNIPAM was utilized as the temperature-sensitive compo-
nent of dual pH- and temperature-responsive membranes [38, 151–158].
In 2009, Yu et al. fabricated thermo- and pH-responsive PP microporous mem-
brane, which was modified by PNIPAM and PAA via RAFT graft polymerization
[38]. It was found that PAA and PNIPAM grafting chains exhibited both pH-
and temperature-dependent water flux, owing to pH responsivity around the pK a
value of PAA (4.72) and temperature sensitivity between 20 and 55 ∘ C. Similarly,
in 2010, the copolymer PAA-b-PNIPAM was grafted onto the surface of regen-
erated cellulose membranes [151].
In 2014, the copolymer microgels (PAA-b-PNIPAM) were blended with
PVDF membrane [152], which acted as a sensor of temperature and pH
to reversibly regulate the permeation of aqueous solution. Meanwhile, the
membrane was superior due to its good fouling resistance for bovine serum
albumin (BSA) and self-cleaning property (Figure 3.23). In a separate study,
poly(N-vinylcaprolactam-co-acrylic acid) was modified onto PSF membranes to
have pH- and temperature-induced separation of BSA and self-cleaning [159].

T > LCST or pH < pKa

T < LCST or pH > pKa

PVDF/PNA microgel membranes PVDF/PNA microgel membranes

Poly(N-isopropylacrylamide) chain Poly(N-isopropylacrylamide-co-acrylic acid)


(PNA) microgels

Poly(acylic acid) chain BSA H2O Direction of water flow

Figure 3.23 Filtration mechanism of microgel membranes under different temperature and
pH conditions. Source: Chen et al. 2014 [152]. Reprinted with permission of Elsevier.
96 3 Filtration Membranes with Responsive Gates

In 2013, Zhao and coworkers blended a terpolymer of poly(N-isopropylacryl-


amide-co-methacrylic acid-co-methyl methacrylate) (P(NIPAM-MAA-MMA))
with PES hollow fiber membranes [153]. The modified membranes exhibited pH
sensitivity at the pH value between 7.0 and 10.0, while the thermoresponsiveness
induced significant changes in pore sizes [154]. When pH was decreased from
10.0 to 2.0, the membrane flux increased by 70%, while the flux got enhanced by
150% with increasing temperature from 20 to 45 ∘ C.
In 2017, Hu and coworkers synthesized a binary graft copolymer composed of
PAA-b-PNIPAM and PMMA and grafted it on the surface of commercial PSf
membrane to serve as thermo- and pH-responsive on/off switches [155]. The
water flux of the membranes increased with increasing temperature from 20 to
50 ∘ C and decreased when the pH increased from 2.0 to 8.5.
In 2018, Shen and coworkers prepared dual thermo/pH-responsive
intelligent gating membranes by in situ assembly of stimuli-responsive
poly(N-isopropylacrylaminde-co-methylacrylic acid) P(NIPAM-co-MAA)
microgels as gates [160]. As presented in Figure 3.24, the intelligent gates would
open as the microgels contract at 70 ∘ C or pH 3, since the NIPAM groups and
MAA groups are shrunken and hydrophobic due to the formation of intermolec-
ular hydrogen bonding between amide/carboxyl and carboxyl. On the contrary,
the intelligent gates would close due to the expanding volume of microgels
with decreasing temperature (30 ∘ C) or increasing pH (11). Correspondingly,
the NIPAM groups and MAA groups become swollen and hydrophilic owing
to the hydrogen bonding between amide groups and water molecules and the
electrostatic repulsion between the protonated carboxylic groups. Moreover,
such intelligent gating membranes showed self-cleaning property as the vol-
ume of microgels changed in response to pH stimulus and thus removed the
contaminants adhered on the membrane surface.
Another pH- and temperature-responsive polymer, polystyrene-block-
poly(N,N-dimethylaminoethyl methacrylate) (PS-b-PDMAEMA), also served
as intelligent gates for dual stimuli-responsive MF and UF membranes via the
non-solvent-induced phase separation process [156, 157]. In 2013, a nylon
membrane was modified with PDMAEMA, PNIPAM, and their diblock brushes,

T pH

Microgel T pH

70 °C/pH 3 30 °C/pH 3 30 °C/pH 11

Figure 3.24 The stimuli-responsive mechanism of intelligent gates and microgels in response
to temperature and pH. Source: Liu et al. 2018 [160]. Reprinted with permission of Royal
Society of Chemistry.
3.11 Dual and Multi-Stimuli Responsiveness 97

which exhibited tunable permeation flux in response to temperature between 30


and 35 ∘ C and pH between 6 and 8 [158].
In Abetz’s study, PS-b-P4VP membranes were first functionalized with poly-
dopamine coating and then reacted with an amine-terminated PNIPAM-NH2 via
Michael addition [20]. The modified membranes exhibited pH and thermo dou-
ble sensitivities, which was proven by measuring the water flux under different
temperature (30–35 ∘ C) and pH conditions (pH 3.8–3.4).
Similarly, in 2017, Chu and coworkers designed a dual layer dual thermo-
and pH-responsive composite membrane, consisting of PS-b-P4VP copoly-
mer and PS bended with PNIPAM nanogels as top layer and bottom layer of
composite membranes, respectively [161]. Under different temperatures and
pH values, the water fluxes of the composite membranes were changed by
several orders of magnitude from 778.03 kg m−2 h−1 bar−1 (45 ∘ C, pH 6.8) to
1.13 kg m−2 h−1 bar−1 (20 ∘ C, pH 2.5), and the dual thermo- and pH-responsive
permeation performances of the composite membranes were satisfactorily
reversible and reproducible (Figure 3.25).

3.11.2 Temperature and Ion Strength Dual Responsiveness


In 2016, PNIPAM and the ion strength-responsive poly-N,N-dimethyl-N-
methacryloyloxyethyl-N-(3-sulfopropyl) ammonium betaine (PSPE) were
grafted onto the pore walls of PET track-etched membranes by sequential
surface-initiated ATRP process [162]. This membrane exhibited switchable
modulations of pore sizes and had controlled UF permeability in response to
temperature (between 25 and 40 ∘ C) and salt ion strength (KClO4 ).
Similarly, in the same year, Meng and coworkers demonstrated ion
strength and temperature dual responsive gating membrane, which was
fabricated by a commercial PA-based RO membrane modified with PNI-
PAM, poly(4-(2-sulfoethyl)-1-(4-vinylbenzyl)pyridinium betain) (PSVBP), and
poly(sulfobetaine methacrylate) (PSBMA) [163]. With CaCO3 as the foulant, the
modified membranes possessed fouling resistance in the whole operations as
long as 320 hours, while with BSA as the foulant, the antifouling performance
existed in short term.

3.11.3 pH and Ion Strength Dual Responsiveness


Furthermore, cross-linked BSA-modified track-etched PC membranes was
prepared for pH- and ion strength-responsive gating membrane [164]. This
is because the repulsive electrostatic interactions between amine groups and
carbonyl groups, and the attractive dipole electrostatic interactions among zwit-
terionic amino acid residue groups among BSA molecules, could be controlled by
tuning pH values and ionic concentration in feed solutions. Figure 3.26a demon-
strated reversible permeability as pH value was alternated between 2.0 and 10.0
in feed solutions, which changed from 1100 ± 20 to 1566 ± 20 l (m−2 h−1 MPa−1 ).
Meanwhile, the water permeability was also controlled by the concentration of
NaCl/CaCl2 in aqueous solutions (Figure 3.26b).
98 3 Filtration Membranes with Responsive Gates

ClosepH – OpenT OpenpH – OpenT

45

40
700

Temperature (°C)
35 600
500
ClosepH – CloseT OpenpH – CloseT
30 400
30
2 300
4
6 50 200
25
8 150
10
70 90
20
2.5 3.0 3.5 4.0 4.5 5.0 5.5 6.0 6.5
(a) pH
1000 1600

1400 pH = 6.8/T = 45 °C
pH = 6.8/T = 20 °C
Flux (kg m–2 h–1 bar –1)

1200 pH = 2.5/T = 45 °C
Flux (kg m–2 h–1)

100 pH = 2.5/T = 20 °C
1000

800

600
10
400

200

1 0
45 °C 20 °C 45 °C 20 °C 0 0.05 0.1 0.15 0.2
pH 6.8 pH 6.8 pH 2.5 pH 2.5 Pressure (MPa)
(b) Environment condition (c)

Figure 3.25 (a) The water flux contour map of the composite membrane and four schemes
near the four vertices of the contour map correspond to four switching states of the composite
membrane at different pH values and temperatures. (b) Water fluxes of the composite
membrane at four switching states. (c) Effect of trans-membrane pressure on the water flux of
the composite membrane. Source: Ma et al. 2017 [161]. Reprinted with permission of
American Chemical Society.
1800

pH = 10.0 1500
Water permeability coefficient

Water permeability coefficient

1600
1200
(l m−2 h−1 MPa−1)

(l m−2 h−1 MPa−1)

NaCl
1400 900

1200 600

300
1000 pH = 2.0 CaCl2
0
0 2 4 6 8 10 12 0.00 0.02 0.04 0.06 0.08 0.10
(a) Cycle time (b) Salt concentration in feed solution (mol/L)

Figure 3.26 (a) Reversible change of water permeability of the composite membrane under
pH = 2.0 and 10.0 in feed solutions. (b) The effect of inorganic salt concentration in feed
solution on water permeability of the composite membrane. Source: Zhao et al. 2013 [164].
Reprinted with permission of Elsevier.
3.12 Conclusions 99

3.11.4 Temperature, pH, and Ion Strength Multi-responsiveness


In 2013, Chu and coworkers developed PNIPAM-b-PMAA-based responsive
gates in the membrane pores, which were able to respond to temperature, pH,
salt ion strength, and anion species [32]. The pore size of membrane and the
hydraulic permeability of buffer solution through membranes changed reversibly
in response to environmental temperature from 25 to 40 ∘ C, pH stimuli (pH = 3
or pH = 8), sodium chloride concentration between 0.1 and 0.3 M, and changing
the salt from sodium chloride to sodium sulfate in the buffer solution.

3.12 Conclusions
In conclusion, valuable efforts have been made in making intelligent gating mem-
branes for water treatment applications. The responsive pore size modulation has
been the major focus in the past, while trigger-initiated responsiveness of other
membrane performance parameters is emerging.
Some of the challenges regarding the state of the art of the intelligent mem-
branes are summarized as follows:
(1) All of the intelligent membranes in the literature were proven; their utilities
at bench scales with simplified testing conditions to provide proofs of con-
cept and so far efforts in scaling up these membranes and challenging them
with more realistic testing conditions are rare. Appropriate steps toward scal-
ing up of intelligent gating materials need to realize real-life applicability and
commercial viability in large-scale environmental applications [86].
(2) The size modulation of these intelligent membranes is far from being pre-
cise. This is so also partially due to the fact the current fabrication methods
for filtration membranes largely lack molecular-level design [8], which limits
the value of the intelligent gating in the overall improvement of membrane
performances, especially selectivity. The precise pore size modulation of the
intelligent membranes at nanometer or even sub-nanometer range would be
a significant target in future development.
(3) The responsiveness of the intelligent membranes is typically induced by
changes in the bulk water chemistry, such as pH and temperature, which
involves high chemical and/or energy consumption [86]. Generally, pH
switching in a scaled-up operation is always associated with the cost of
consumption of chemicals. Despite no external chemicals are required for
temperature responsiveness, it is no doubt that heat exchangers of fluid
(heat generation or cooling) require the cost of fuel.
(4) The previous research on the intelligent membranes was dominantly focused
on the chemistry and conformation changes of the responsive materials, and
little attention was paid to the detailed mechanisms for mass transfer and
separation within the intelligent gating membranes [12]. While important
efforts have been made to utilize responsive chemistry to improve perfor-
mance parameters of membranes other than pore size, their importance
should be further strengthened.
100 3 Filtration Membranes with Responsive Gates

Intelligent membranes have the potential to make some difference in the


following areas:
(1) Membrane fouling is always a major challenge in all kinds of membrane-based
separation, and it worsens along with increasing water flux. The self-initiated
conformations or chemistry changes in response to changing environmental
conditions by stimuli-responsive materials can be a good platform to design
membranes with improved antifouling and fouling-resistant performance.
Stimuli-responsive materials have been combined with and helped produce
a number of filtration membranes with self-healing performance, and more
efforts should be invested in this interesting topic.
(2) Light, especially UV light, as a remote and clean trigger, has been used to
induce performance adjustment of the intelligent membranes. However,
direct utilization of solar light to induce the same performance has been rare.
Solar light is the most renewable energy source, and thus the integration
of photothermal component into conventional and intelligent membranes
would result in more energy-efficient membrane separation with better
performance.
(3) Synergistically multifunctional and all-in-one membranes in the format of
point-of-use devices can be a niche area for the intelligent membrane to
thrive.
Therefore, addressing these critical issues from design, scalability, efficiency,
reproducibility, and self-cleaning, intelligent gating membranes would
provide the promising potential for a paradigm shift in the field of water
treatment.

References
1 Werber, J.R., Osuji, C.O., and Elimelech, M. (2016). Materials for
next-generation desalination and water purification membranes. Nature
Reviews Materials 1 (5): 16018.
2 Chang, J., Zhang, L., and Wang, P. (2018). Intelligent environmental nanoma-
terials. Environmental Science: Nano 5 (4): 811–836.
3 Elimelech, M. and Phillip, W.A. (2011). The future of seawater desalination:
energy, technology, and the environment. Science 333 (6043): 712–717.
4 Shaffer, D.L., Werber, J.R., Jaramillo, H. et al. (2015). Forward osmosis:
where are we now? Desalination 356: 271–284.
5 Geise, G.M., Paul, D.R., and Freeman, B.D. (2014). Fundamental water and
salt transport properties of polymeric materials. Progress in Polymer Science
39 (1): 1–42.
6 Qu, X., Alvarez, P.J., and Li, Q. (2013). Applications of nanotechnology in
water and wastewater treatment. Water Research 47 (12): 3931–3946.
7 Chen, D., Zhu, H., Yang, S. et al. (2016). Micro-nanocomposites in environ-
mental management. Advanced Materials 28 (47): 10443–10458.
8 Gin, D.L. and Noble, R.D. (2011). Designing the next generation of chemical
separation membranes. Science 332 (6030): 674–676.
References 101

9 Jiang, Y., Lee, A., Chen, J. et al. (2002). The open pore conformation of
potassium channels. Nature 417 (6888): 523–526.
10 Koros, W.J. and Zhang, C. (2017). Materials for next-generation molecularly
selective synthetic membranes. Nature Materials 16 (3): 289–297.
11 Heskins, M. and Guillet, J.E. (1968). Solution properties of
poly(N-isopropylacrylamide). Journal of Macromolecular Science-Chemistry 2
(8): 1441–1455.
12 Liu, Z., Wang, W., Xie, R. et al. (2016). Stimuli-responsive smart gating
membranes. Chemical Society Reviews 45 (3): 460–475.
13 Hou, X. (2016). Smart gating multi-scale pore/channel-based membranes.
Advanced Materials 28 (33): 7049–7064.
14 Tokarev, I. and Minko, S. (2010). Stimuli-responsive porous hydrogels at
interfaces for molecular filtration, separation, controlled release, and gating
in capsules and membranes. Advanced Materials 22 (31): 3446–3462.
15 Zhu, Y., Gao, S., Hu, L., and Jin, J. (2016). Thermoresponsive ultrathin mem-
branes with precisely tuned nanopores for high-flux separation. ACS Applied
Materials & Interfaces 8 (21): 13607–13614.
16 Mika, A., Childs, R., Dickson, J. et al. (1995). A new class of
polyelectrolyte-filled microfiltration membranes with environmentally con-
trolled porosity. Journal of Membrane Science 108 (1-2): 37–56.
17 Okahata, Y., Ozaki, K., and Seki, T. (1984). pH-sensitive permeability con-
trol of polymer-grafted nylon capsule membranes. Journal of the Chemical
Society, Chemical Communications (8): 519–521.
18 Shi, W., Deng, J., Qin, H. et al. (2014). Poly (ether sulfone) membranes with
photo-responsive permeability. Journal of Membrane Science 455: 357–367.
19 Fujiwara, M. and Imura, T. (2015). Photo induced membrane separation for
water purification and desalination using azobenzene modified anodized
alumina membranes. ACS Nano 9 (6): 5705–5712.
20 Clodt, J.I., Filiz, V., Rangou, S. et al. (2013). Double stimuli-responsive iso-
porous membranes via post-modification of pH-sensitive self-assembled
diblock copolymer membranes. Advanced Functional Materials 23 (6):
731–738.
21 Peng, T. and Cheng, Y.L. (1998). Temperature-responsive permeability of
porous PNIPAAm-g-PE membranes. Journal of Applied Polymer Science 70
(11): 2133–2142.
22 Chu, L.-Y., Niitsuma, T., Yamaguchi, T., and Nakao, S.-i. (2003). Thermore-
sponsive transport through porous membranes with grafted PNIPAM gates.
AIChE Journal 49 (4): 896–909.
23 Ying, L., Yu, W., Kang, E., and Neoh, K. (2004). Functional and surface-active
membranes from poly(vinylidene fluoride)-graft-poly(acrylic acid) prepared
via RAFT-mediated graft copolymerization. Langmuir 20 (14): 6032–6040.
24 Menne, D., Pitsch, F., Wong, J.E. et al. (2014). Temperature-modulated water
filtration using microgel-functionalized hollow-fiber membranes. Angewandte
Chemie International Edition 53 (22): 5706–5710.
25 Lee, D., Nolte, A.J., Kunz, A.L. et al. (2006). pH-induced hysteretic gating
of track-etched polycarbonate membranes: swelling/deswelling behavior of
102 3 Filtration Membranes with Responsive Gates

polyelectrolyte multilayers in confined geometry. Journal of the American


Chemical Society 128 (26): 8521–8529.
26 Kumar, S.K. and Hong, J.-D. (2008). Photoresponsive ion gating func-
tion of an azobenzene polyelectrolyte multilayer spin-self-assembled on a
nanoporous support. Langmuir 24 (8): 4190–4193.
27 Ran, J., Wu, L., Zhang, Z., and Xu, T. (2014). Atom transfer radical poly-
merization (ATRP): a versatile and forceful tool for functional membranes.
Progress in Polymer Science 39 (1): 124–144.
28 Alem, H., Duwez, A.-S., Lussis, P. et al. (2008). Microstructure and
thermo-responsive behavior of poly(N-isopropylacrylamide) brushes grafted
in nanopores of track-etched membranes. Journal of Membrane Science 308
(1-2): 75–86.
29 Friebe, A. and Ulbricht, M. (2007). Controlled pore functionalization of
poly(ethylene terephthalate) track-etched membranes via surface-initiated
atom transfer radical polymerization. Langmuir 23 (20): 10316–10322.
30 Yameen, B., Ali, M., Neumann, R. et al. (2009). Ionic transport through
single solid-state nanopores controlled with thermally nanoactuated macro-
molecular gates. Small 5 (11): 1287–1291.
31 Schepelina, O. and Zharov, I. (2007). PNIPAAM-modified nanoporous col-
loidal films with positive and negative temperature gating. Langmuir 23 (25):
12704–12709.
32 Chen, Y.-C., Xie, R., and Chu, L.-Y. (2013). Stimuli-responsive gating mem-
branes responding to temperature, pH, salt concentration and anion species.
Journal of Membrane Science 442: 206–215.
33 Xue, J., Chen, L., Wang, H. et al. (2008). Stimuli-responsive multifunc-
tional membranes of controllable morphology from poly(vinylidene
fluoride)-graft-poly[2-(N,N-dimethylamino)ethyl methacrylate] prepared
via atom transfer radical polymerization. Langmuir 24 (24): 14151–14158.
34 Lee, B.-Y., Hyun, S., Jeon, G. et al. (2016). Bioinspired dual
stimuli-responsive membranous system with multiple on-off gates. ACS
Applied Materials & Interfaces 8 (18): 11758–11764.
35 Friebe, A. and Ulbricht, M. (2009). Cylindrical pores responding to two dif-
ferent stimuli via surface-initiated atom transfer radical polymerization for
synthesis of grafted diblock copolymers. Macromolecules 42 (6): 1838–1848.
36 Lin, F., Li, Q., Jiang, D. et al. (2009). Thermoresponsive microfiltration mem-
branes prepared by atom transfer radical polymerization directly from
poly(vinylidene fluoride). Iranian Polymer Journal 18 (7): 561–568.
37 Frost, S. and Ulbricht, M. (2013). Thermoresponsive ultrafiltration mem-
branes for the switchable permeation and fractionation of nanoparticles.
Journal of Membrane Science 448: 1–11.
38 Yu, H.-Y., Li, W., Zhou, J. et al. (2009). Thermo-and pH-responsive
polypropylene microporous membrane prepared by the photoinduced
RAFT-mediated graft copolymerization. Journal of Membrane Science 343
(1-2): 82–89.
39 Deng, J., Wang, L., Liu, L., and Yang, W. (2009). Developments and new
applications of UV-induced surface graft polymerizations. Progress in Poly-
mer Science 34 (2): 156–193.
References 103

40 Geismann, C., Yaroshchuk, A., and Ulbricht, M. (2007). Permeability and


electrokinetic characterization of poly(ethylene terephthalate) capillary pore
membranes with grafted temperature-responsive polymers. Langmuir 23 (1):
76–83.
41 Peng, T. and Cheng, Y.L. (2000). pH-responsive permeability of PE-g-PMAA
membranes. Journal of Applied Polymer Science 76 (6): 778–786.
42 Shim, J.K., Lee, Y.B., and Lee, Y.M. (1999). pH-dependent permeation
through polysulfone ultrafiltration membranes prepared by ultraviolet
polymerization technique. Journal of Applied Polymer Science 74 (1): 75–82.
43 Nayak, A., Liu, H., and Belfort, G. (2006). An optically reversible switch-
ing membrane surface. Angewandte Chemie International Edition 45 (25):
4094–4098.
44 Xie, R., Chu, L.-Y., Chen, W.-M. et al. (2005). Characterization of
microstructure of poly(N-isopropylacrylamide)-grafted polycarbonate
track-etched membranes prepared by plasma-graft pore-filling polymer-
ization. Journal of Membrane Science 258 (1-2): 157–166.
45 Meng, T., Xie, R., Chen, Y.-C. et al. (2010). A thermo-responsive
affinity membrane with nano-structured pores and grafted
poly(N-isopropylacrylamide) surface layer for hydrophobic adsorption.
Journal of Membrane Science 349 (1-2): 258–267.
46 Liu, Z., Luo, F., Ju, X.-J. et al. (2013). Gating membranes for water treat-
ment: detection and removal of trace Pb2+ ions based on molecular
recognition and polymer phase transition. Journal of Materials Chemistry
A 1 (34): 9659–9671.
47 Chu, L.-Y., Li, Y., Zhu, J.-H. et al. (2004). Control of pore size and perme-
ability of a glucose-responsive gating membrane for insulin delivery. Journal
of Controlled Release 97 (1): 43–53.
48 Kuroki, H., Ohashi, H., Ito, T. et al. (2010). Isolation and analysis of a grafted
polymer onto a straight cylindrical pore in a thermal-responsive gating
membrane and elucidation of its permeation behavior. Journal of Membrane
Science 352 (1-2): 22–31.
49 Xie, R., Li, Y., and Chu, L.-Y. (2007). Preparation of thermo-responsive
gating membranes with controllable response temperature. Journal of Mem-
brane Science 289 (1-2): 76–85.
50 Li, Y., Chu, L.-Y., Zhu, J.-H. et al. (2004). Thermoresponsive gating char-
acteristics of poly(N-isopropylacrylamide)-grafted porous poly(vinylidene
fluoride) membranes. Industrial & Engineering Chemistry Research 43 (11):
2643–2649.
51 Ohashi, H., Chi, X., Kuroki, H., and Yamaguchi, T. (2016). Response sen-
sitivity of a gating membrane related to grafted polymer characteristics.
Industrial & Engineering Chemistry Research 55 (6): 1575–1581.
52 Choi, Y.-J., Yamaguchi, T., and Nakao, S.-i. (2000). A novel separation sys-
tem using porous thermosensitive membranes. Industrial & Engineering
Chemistry Research 39 (7): 2491–2495.
53 Yang, M., Xie, R., Wang, J.-Y. et al. (2010). Gating characteristics of
thermo-responsive and molecular-recognizable membranes based on
104 3 Filtration Membranes with Responsive Gates

poly(N-isopropylacrylamide) and 𝛽-cyclodextrin. Journal of Membrane


Science 355 (1-2): 142–150.
54 Hou, X., Liu, Y., Dong, H. et al. (2010). A pH-gating ionic transport nanode-
vice: asymmetric chemical modification of single nanochannels. Advanced
Materials 22 (22): 2440–2443.
55 Wu, J., Wang, N., Zhang, H. et al. (2013). Acrylic acid grafted porous poly-
carbonate membrane with smart hydrostatic pressure response to pH.
Journal of Materials Chemistry A 1 (15): 4642–4646.
56 Zhang, H., Tian, Y., Hou, J. et al. (2015). Bioinspired smart
gate-location-controllable single nanochannels: experiment and theoretical
simulation. ACS Nano 9 (12): 12264–12273.
57 Lee, Y.M. and Shim, J.K. (1996). Plasma surface graft of acrylic acid onto a
porous poly(vinylidene fluoride) membrane and its riboflavin permeation.
Journal of Applied Polymer Science 61 (8): 1245–1250.
58 Hou, X., Yang, F., Li, L. et al. (2010). A biomimetic asymmetric responsive
single nanochannel. Journal of the American Chemical Society 132 (33):
11736–11742.
59 Baumann, L., de Courten, D., Wolf, M. et al. (2013). Light-responsive caf-
feine transfer through porous polycarbonate. ACS Applied Materials &
Interfaces 5 (13): 5894–5897.
60 Baumann, L., Schöller, K., de Courten, D. et al. (2013). Development of
light-responsive porous polycarbonate membranes for controlled caffeine
delivery. RSC Advances 3 (45): 23317–23326.
61 Guo, H. and Ulbricht, M. (2011). Preparation of thermo-responsive
polypropylene membranes via surface entrapment of
poly(N-isopropylacrylamide)-containing macromolecules. Journal of Mem-
brane Science 372 (1-2): 331–339.
62 Ito, Y., Ochiai, Y., Park, Y.S., and Imanishi, Y. (1997). pH-sensitive gating
by conformational change of a polypeptide brush grafted onto a porous
polymer membrane. Journal of the American Chemical Society 119 (7):
1619–1623.
63 Chung, D.J., Ito, Y., and Imanishi, Y. (1994). Preparation of porous mem-
branes grafted with poly(spiropyran-containing methacrylate) and photocon-
trol of permeability. Journal of Applied Polymer Science 51 (12): 2027–2033.
64 Li, P.-F., Ju, X.-J., Chu, L.-Y., and Xie, R. (2006). Thermo-responsive mem-
branes with cross-linked poly(N-isopropyl-acrylamide) hydrogels inside
porous substrates. Chemical Engineering & Technology 29 (11): 1333–1339.
65 Lewis, S.R., Datta, S., Gui, M. et al. (2011). Reactive nanostructured mem-
branes for water purification. Proceedings of the National Academy of
Sciences 108 (21): 8577–8582.
66 Wu, D., Liu, X., Yu, S. et al. (2010). Modification of aromatic polyamide
thin-film composite reverse osmosis membranes by surface coating of
thermo-responsive copolymers P(NIPAM-co-Am). I: Preparation and charac-
terization. Journal of Membrane Science 352 (1-2): 76–85.
67 Yu, S., Lü, Z., Chen, Z. et al. (2011). Surface modification of thin-film
composite polyamide reverse osmosis membranes by coating
References 105

N-isopropylacrylamide-co-acrylic acid copolymers for improved membrane


properties. Journal of Membrane Science 371 (1-2): 293–306.
68 Yu, S., Chen, Z., Liu, J. et al. (2012). Intensified cleaning of organic-fouled
reverse osmosis membranes by thermo-responsive polymer (TRP). Journal of
Membrane Science 392-393: 181–191.
69 Anzai, J.i., Ueno, A., Sasaki, H. et al. (1983). Photocontrolled permeation of
alkali cations through poly(vinyl chloride)/crown ether membrane. Macro-
molecular Rapid Communications 4 (11): 731–734.
70 Wang, G., Xie, R., Ju, X.-J., and Chu, L.-Y. (2012). Thermo-responsive
polyethersulfone composite membranes blended with
poly(N-isopropylacrylamide) nanogels. Chemical Engineering & Technology
35 (11): 2015–2022.
71 Luo, T., Lin, S., Xie, R. et al. (2014). pH-responsive poly(ether sulfone) com-
posite membranes blended with amphiphilic polystyrene-block-poly(acrylic
acid) copolymers. Journal of Membrane Science 450: 162–173.
72 Li, P.-F., Xie, R., Jiang, J.-C. et al. (2009). Thermo-responsive gating mem-
branes with controllable length and density of poly(N-isopropylacrylamide)
chains grafted by ATRP method. Journal of Membrane Science 337 (1-2):
310–317.
73 Liu, N., Chen, Z., Dunphy, D.R. et al. (2003). Photoresponsive nanocompos-
ite formed by self-assembly of an azobenzene-modified silane. Angewandte
Chemie International Edition 42 (15): 1731–1734.
74 Bojko, A., Andreatta, G., Montagne, F. et al. (2014). Fabrica-
tion of thermo-responsive nano-valve by grafting-to in melt of
poly(N-isopropylacrylamide) onto nanoporous silicon nitride membranes.
Journal of Membrane Science 468: 118–125.
75 Peinemann, K.-V., Abetz, V., and Simon, P.F.W. (2007). Asymmetric super-
structure formed in a block copolymer via phase separation. Nature
Materials 6 (12): 992–996.
76 Yu, H., Qiu, X., Nunes, S.P., and Peinemann, K.-V. (2014). Biomimetic block
copolymer particles with gated nanopores and ultrahigh protein sorption
capacity. Nature Communications 5: 4110.
77 Yang, X., Deng, B., Liu, Z. et al. (2010). Microfiltration membranes prepared
from acryl amide grafted poly(vinylidene fluoride) powder and their pH sen-
sitive behaviour. Journal of Membrane Science 362 (1-2): 298–305.
78 Yu, S., Liu, X., Liu, J. et al. (2011). Surface modification of thin-film compos-
ite polyamide reverse osmosis membranes with thermo-responsive polymer
(TRP) for improved fouling resistance and cleaning efficiency. Separation
and Purification Technology 76 (3): 283–291.
79 Tyagi, P., Deratani, A., Bouyer, D. et al. (2012). Dynamic interactive mem-
branes with pressure-driven tunable porosity and self-healing ability.
Angewandte Chemie International Edition 51 (29): 7166–7170.
80 Kim, S.-R., Getachew, B.A., Park, S.-J. et al. (2016). Toward
microcapsule-embedded self-healing membranes. Environmental Science
& Technology Letters 3 (5): 216–221.
106 3 Filtration Membranes with Responsive Gates

81 Getachew, B.A., Kim, S.-R., and Kim, J.-H. (2017). Self-healing hydrogel
pore-filled water filtration membranes. Environmental Science & Technology
51 (2): 905–913.
82 Kim, S.-R., Getachew, B.A., and Kim, J.-H. (2017). In situ healing of compro-
mised membranes via polyethylenimine-functionalized silica microparticles.
Environmental Science & Technology 51 (21): 12630–12637.
83 Dongare, P.D., Alabastri, A., Pedersen, S. et al. (2017).
Nanophotonics-enabled solar membrane distillation for off-grid water
purification. Proceedings of the National Academy of Sciences 114 (27):
6936–6941.
84 Rama Rao, G., Krug, M.E., Balamurugan, S. et al. (2002). Synthesis and
characterization of silica-poly(N-isopropylacrylamide) hybrid membranes:
switchable molecular filters. Chemistry of Materials 14 (12): 5075–5080.
85 Fu, Q., Rama Rao, G., Ward, T.L. et al. (2007). Thermoresponsive transport
through ordered mesoporous silica/PNIPAAm copolymer membranes and
microspheres. Langmuir 23 (1): 170–174.
86 Dutta, K. and De, S. (2017). Smart responsive materials for water purifica-
tion: an overview. Journal of Materials Chemistry A 5 (42): 22095–22112.
87 Lue, S.J., Hsu, J.-J., Chen, C.-H., and Chen, B.-C. (2007). Thermally on-off
switching membranes of poly(N-isopropylacrylamide) immobilized in
track-etched polycarbonate films. Journal of Membrane Science 301 (1-2):
142–150.
88 Li, S. and D’Emanuele, A. (2001). On-off transport through a thermorespon-
sive hydrogel composite membrane. Journal of Controlled Release 75 (1-2):
55–67.
89 Okahata, Y., Noguchi, H., and Seki, T. (1986). Thermoselective permeation
from a polymer-grafted capsule membrane. Macromolecules 19 (2): 493–494.
90 Yang, M., Chu, L.Y., Li, Y. et al. (2006). Thermo-responsive gating char-
acteristics of poly(N-isopropylacrylamide)-grafted membranes. Chemical
Engineering & Technology 29 (5): 631–636.
91 Alem, H., Jonas, A.M., and Demoustier-Champagne, S. (2010).
Poly(N-isopropylacrylamide) grafted into nanopores: thermo-responsive
behaviour in the presence of different salts. Polymer Degradation and Stabil-
ity 95 (3): 327–331.
92 Wu, C.-J., Xie, R., Wei, H.-B. et al. (2016). Fabrication of a
thermo-responsive membrane with cross-linked smart gates via a
‘grafting-to’ method. RSC Advances 6 (51): 45428–45433.
93 Gajda, A.M. and Ulbricht, M. (2014). Magnetic Fe3 O4 nanoparticle heaters
in smart porous membrane valves. Journal of Materials Chemistry B 2 (10):
1317–1326.
94 Petrov, S., Ivanova, T., Christova, D., and Ivanova, S. (2005). Modi-
fication of polyacrylonitrile membranes with temperature sensitive
poly(vinylalcohol-co-vinylacetal). Journal of Membrane Science 261 (1-2):
1–6.
References 107

95 Lequieu, W., Shtanko, N., and Du Prez, F. (2005). Track etched membranes
with thermo-adjustable porosity and separation properties by surface immo-
bilization of poly(N-vinylcaprolactam). Journal of Membrane Science 256
(1-2): 64–71.
96 Lohaus, T., de Wit, P., Kather, M. et al. (2017). Tunable permeability and
selectivity: heatable inorganic porous hollow fiber membrane with a
thermo-responsive microgel coating. Journal of Membrane Science 539:
451–457.
97 Chu, L.-Y., Li, Y., Zhu, J.-H., and Chen, W.-M. (2005). Negatively ther-
moresponsive membranes with functional gates driven by zipper-type
hydrogen-bonding interactions. Angewandte Chemie International Edition 44
(14): 2124–2127.
98 Liu, J., Wang, N., Yu, L.-J. et al. (2017). Bioinspired graphene membrane
with temperature tunable channels for water gating and molecular separa-
tion. Nature Communications 8 (1): 2011.
99 Qiu, X., Yu, H., Karunakaran, M. et al. (2012). Selective separation of simi-
larly sized proteins with tunable nanoporous block copolymer membranes.
ACS Nano 7 (1): 768–776.
100 Zhang, H. and Ito, Y. (2001). pH control of transport through a porous
membrane self-assembled with a poly (acrylic acid) loop brush. Langmuir 17
(26): 8336–8340.
101 Wang, M., An, Q.-F., Wu, L.-G. et al. (2007). Preparation of pH-responsive
phenolphthalein poly (ether sulfone) membrane by redox-graft pore-filling
polymerization technique. Journal of Membrane Science 287 (2): 257–263.
102 Yu, H., Qiu, X., Nunes, S.P., and Peinemann, K.-V. (2014). Self-assembled
isoporous block copolymer membranes with tuned pore sizes. Angewandte
Chemie International Edition 53 (38): 10072–10076.
103 Nunes, S.P., Behzad, A.R., Hooghan, B. et al. (2011). Switchable
pH-responsive polymeric membranes prepared via block copolymer micelle
assembly. ACS Nano 5 (5): 3516–3522.
104 Hilke, R., Pradeep, N., Madhavan, P. et al. (2013). Block copolymer hol-
low fiber membranes with catalytic activity and pH-response. ACS Applied
Materials & Interfaces 5 (15): 7001–7006.
105 Han, Z., Cheng, C., Zhang, L. et al. (2014). Toward robust pH-responsive
and anti-fouling composite membranes via one-pot in-situ cross-linked
copolymerization. Desalination 349: 80–93.
106 Motornov, M., Sheparovych, R., Katz, E., and Minko, S. (2007). Chemical
gating with nanostructured responsive polymer brushes: mixed brush versus
homopolymer brush. ACS Nano 2 (1): 41–52.
107 Schepelina, O., Poth, N., and Zharov, I. (2010). pH-responsive nanoporous
silica colloidal membranes. Advanced Functional Materials 20 (12):
1962–1969.
108 Qu, J.-B., Chu, L.-Y., Yang, M. et al. (2006). A pH-responsive gating
membrane system with pumping effects for improved controlled release.
Advanced Functional Materials 16 (14): 1865–1872.
108 3 Filtration Membranes with Responsive Gates

109 Ito, Y., Inaba, M., Chung, D.J., and Imanishi, Y. (1992). Control of water
permeation by pH and ionic strength through a porous membrane having
poly(carboxylic acid) surface-grafted. Macromolecules 25 (26): 7313–7316.
110 Hester, J., Olugebefola, S., and Mayes, A. (2002). Preparation of
pH-responsive polymer membranes by self-organization. Journal of Mem-
brane Science 208 (1-2): 375–388.
111 Shevate, R., Karunakaran, M., Kumar, M., and Peinemann, K.-V. (2016).
Polyanionic pH-responsive polystyrene-b-poly(4-vinyl pyridine-N-oxide)
isoporous membranes. Journal of Membrane Science 501: 161–168.
112 Shevate, R., Kumar, M., Karunakaran, M. et al. (2018). Surprising trans-
formation of a block copolymer into a high performance polystyrene
ultrafiltration membrane with a hierarchically organized pore structure.
Journal of Materials Chemistry A 6 (10): 4337–4345.
113 Poole, J.L., Donahue, S., Wilson, D. et al. (2017). Biocatalytic
stimuli-responsive asymmetric triblock terpolymer membranes for local-
ized permeability gating. Macromolecular Rapid Communications 38 (19):
1700364.
114 Ito, Y., Park, Y.S., and Imanishi, Y. (1997). Visualization of critical
pH-controlled gating of a porous membrane grafted with polyelectrolyte
brushes. Journal of the American Chemical Society 119 (11): 2739–2740.
115 Ito, Y., Park, Y.S., and Imanishi, Y. (2000). Nanometer-sized channel gating
by a self-assembled polypeptide brush. Langmuir 16 (12): 5376–5381.
116 Ito, Y. (1998). Signal-responsive gating by a polyelectrolyte pelage on a
nanoporous membrane. Nanotechnology 9 (3): 205–207.
117 Ito, Y., Park, Y.S., and Imanishi, Y. (1997). Imaging of a pH-sensitive polymer
brush on a porous membrane using atomic force microscopy in aqueous
solution. Macromolecular Rapid Communications 18 (3): 221–224.
118 Deng, B., Yang, X., Xie, L. et al. (2009). Microfiltration membranes with pH
dependent property prepared from poly(methacrylic acid) grafted polyether-
sulfone powder. Journal of Membrane Science 330 (1-2): 363–368.
119 Shi, Q., Su, Y., Ning, X. et al. (2010). Graft polymerization of methacrylic
acid onto polyethersulfone for potential pH-responsive membrane materials.
Journal of Membrane Science 347 (1-2): 62–68.
120 Kumar, S.K., Pennakalathil, J., Kim, T.-H. et al. (2008). Photoregulation of
ion permeation through a polyelectrolyte multilayer membrane by manipu-
lating the chromophore orientation. Langmuir 25 (3): 1767–1771.
121 Kumeria, T., Yu, J., Alsawat, M. et al. (2015). Photoswitchable membranes
based on peptide-modified nanoporous anodic alumina: toward smart mem-
branes for on-demand molecular transport. Advanced Materials 27 (19):
3019–3024.
122 Fujiwara, M. (2017). Water desalination using visible light by disperse red 1
modified PTFE membrane. Desalination 404: 79–86.
123 Fujiwara, M. and Kikuchi, M. (2017). Solar desalination of seawater using
double-dye-modified PTFE membrane. Water Research 127: 96–103.
124 Dübner, M., Naoum, M.-E., Spencer, N.D., and Padeste, C. (2017). From
pH-to light-response: postpolymerization modification of polymer brushes
References 109

grafted onto microporous polymeric membranes. ACS Omega 2 (2):


455–461.
125 Park, Y.S., Ito, Y., and Imanishi, Y. (1998). Photocontrolled gating by polymer
brushes grafted on porous glass filter. Macromolecules 31 (8): 2606–2610.
126 Itoh, K., Okamoto, T., Wakita, S. et al. (1991). Thin films of peroxopoly-
tungstic acids: applications to optical waveguide components. Applied
Organometallic Chemistry 5 (4): 295–301.
127 Mi, P., Chu, L.Y., Ju, X.J., and Niu, C.H. (2008). A smart polymer with
ion-induced negative shift of the lower critical solution temperature for
phase transition. Macromolecular Rapid Communications 29 (1): 27–32.
128 Chu, L.Y., Yamaguchi, T., and Nakao, S.-i. (2002). A molecular-recognition
microcapsule for environmental stimuli-responsive controlled release.
Advanced Materials 14 (5): 386–389.
129 Ito, T., Hioki, T., Yamaguchi, T. et al. (2002). Development of a molecular
recognition ion gating membrane and estimation of its pore size control.
Journal of the American Chemical Society 124 (26): 7840–7846.
130 Ito, T., Sato, Y., Yamaguchi, T., and Nakao, S.-i. (2004). Response mechanism
of a molecular recognition ion gating membrane. Macromolecules 37 (9):
3407–3414.
131 Yu, H.-R., Hu, J.-Q., Lu, X.-H. et al. (2015). Insights into the effects of 2: 1
“sandwich-type” crown-ether/metal-ion complexes in responsive host-guest
systems. The Journal of Physical Chemistry B 119 (4): 1696–1705.
132 Irie, M., Misumi, Y., and Tanaka, T. (1993). Stimuli-responsive polymers:
chemical induced reversible phase separation of an aqueous solution of
poly(N-isopropylacrylamide) with pendent crown ether groups. Polymer 34
(21): 4531–4535.
133 Yamaguchi, T., Ito, T., Sato, T. et al. (1999). Development of a fast response
molecular recognition ion gating membrane. Journal of the American Chemi-
cal Society 121 (16): 4078–4079.
134 Ito, T. and Yamaguchi, T. (2004). Osmotic pressure control in response to a
specific ion signal at physiological temperature using a molecular recognition
ion gating membrane. Journal of the American Chemical Society 126 (20):
6202–6203.
135 Ito, T. and Yamaguchi, T. (2006). Nonlinear self-excited oscillation of a syn-
thetic ion-channel-inspired membrane. Angewandte Chemie International
Edition 45 (34): 5630–5633.
136 Ito, T. and Yamaguchi, T. (2006). Controlled release of model drugs through
a molecular recognition ion gating membrane in response to a specific ion
signal. Langmuir 22 (8): 3945–3949.
137 Ito, T., Oshiba, Y., Ohashi, H. et al. (2010). Reentrant phase transition
behavior and sensitivity enhancement of a molecular recognition ion gat-
ing membrane in an aqueous ethanol solution. Journal of Membrane Science
348 (1-2): 369–375.
138 Liu, Z., Luo, F., Ju, X.J. et al. (2012). Positively K+ -responsive membranes
with functional gates driven by host–guest molecular recognition. Advanced
Functional Materials 22 (22): 4742–4750.
110 3 Filtration Membranes with Responsive Gates

139 Logtenberg, H. and Browne, W.R. (2013). Electrochemistry of


dithienylethenes and their application in electropolymer modified photo-
and redox switchable surfaces. Organic & Biomolecular Chemistry 11 (2):
233–243.
140 Mazurowski, M., Gallei, M., Li, J. et al. (2012). Redox-responsive polymer
brushes grafted from polystyrene nanoparticles by means of surface initiated
atom transfer radical polymerization. Macromolecules 45 (22): 8970–8981.
141 Ito, Y., Nishi, S., Park, Y.S., and Imanishi, Y. (1997). Oxidoreduction-sensitive
control of water permeation through a polymer brushes-grafted porous
membrane. Macromolecules 30 (19): 5856–5859.
142 Phillips, D.J. and Gibson, M.I. (2012). Degradable thermoresponsive poly-
mers which display redox-responsive LCST behaviour. Chemical Communi-
cations 48 (7): 1054–1056.
143 Elbert, J., Krohm, F., Rüttiger, C. et al. (2014). Polymer-modified mesoporous
silica thin films for redox-mediated selective membrane gating. Advanced
Functional Materials 24 (11): 1591–1601.
144 Zhang, K., Feng, X., Sui, X. et al. (2014). Breathing pores on command:
redox-responsive spongy membranes from poly(ferrocenylsilane)s. Ange-
wandte Chemie International Edition 53 (50): 13789–13793.
145 Zhao, Y.-F., Zhang, P.-B., Sun, J. et al. (2016). Electrolyte-responsive
polyethersulfone membranes with zwitterionic polyethersulfone-based
copolymers as additive. Journal of Membrane Science 510: 306–313.
146 Okahata, Y., Lim, H.-J., Hachiya, S., and Nakamura, G.-I. (1984).
Bilayer-coated capsule membranes. IV.: Control of NaCl permeability by
phase transition of synthetic bilayer coatings, depending on their hydrophilic
head groups. Journal of Membrane Science 19 (3): 237–247.
147 Zhai, G., Toh, S., Tan, W. et al. (2003). Poly(vinylidene fluoride) with grafted
zwitterionic polymer side chains for electrolyte-responsive microfiltration
membranes. Langmuir 19 (17): 7030–7037.
148 Meng, J., Cao, Z., Ni, L. et al. (2014). A novel salt-responsive TFC RO mem-
brane having superior antifouling and easy-cleaning properties. Journal of
Membrane Science 461: 123–129.
149 Zhang, X., Zhou, J., Wei, R. et al. (2017). Design of anion species/strength
responsive membranes via in-situ cross-linked copolymerization of ionic
liquids. Journal of Membrane Science 535: 158–167.
150 Xie, S., Wu, S., Bao, S. et al. (2018). Intelligent mesoporous materials for
selective adsorption and mechanical release of organic pollutants from
water. Advanced Materials 30 (27): 1800683.
151 Pan, K., Zhang, X., Ren, R., and Cao, B. (2010). Double stimuli-responsive
membranes grafted with block copolymer by ATRP method. Journal of
Membrane Science 356 (1-2): 133–137.
152 Chen, X., He, Y., Shi, C. et al. (2014). Temperature- and pH-responsive
membranes based on poly(vinylidene fluoride) functionalized with microgels.
Journal of Membrane Science 469: 447–457.
153 Li, H., Liao, J., Xiang, T. et al. (2013). Preparation and characterization
of pH- and thermo-sensitive polyethersulfone hollow fiber membranes
modified with P (NIPAAm-MAA-MMA) terpolymer. Desalination 309: 1–10.
References 111

154 Zhang, R., Su, Y., Peng, J. et al. (2014). pH and temperature responsive
porous membranes via an in situ bulk copolymerization approach. Polymer
55 (6): 1347–1357.
155 Miao, L., Tu, Y., Yang, Y. et al. (2017). Robust stimuli-responsive mem-
branes prepared from a blend of polysulfone and a graft copolymer bearing
binary side chains with thermo- and pH-responsive switching behavior.
Chemistry--A European Journal 23 (32): 7737–7747.
156 Schacher, F., Ulbricht, M., and Müller, A.H. (2009).
Self-supporting, double stimuli-responsive porous membranes from
polystyrene-block-poly(N,N-dimethylaminoethyl methacrylate) diblock
copolymers. Advanced Functional Materials 19 (7): 1040–1045.
157 Schacher, F., Rudolph, T., Wieberger, F. et al. (2009). Dou-
ble stimuli-responsive ultrafiltration membranes from
polystyrene-block-poly(N,N-dimethylaminoethyl methacrylate) diblock
copolymers. ACS Applied Materials & Interfaces 1 (7): 1492–1503.
158 Zhang, Z., Zhu, X., Xu, F. et al. (2009). Temperature- and pH-sensitive nylon
membranes prepared via consecutive surface-initiated atom transfer radical
graft polymerizations. Journal of Membrane Science 342 (1-2): 300–306.
159 Sinha, M. and Purkait, M. (2014). Preparation and characterization of
stimuli-responsive hydrophilic polysulfone membrane modified with
poly(N-vinylcaprolactam-co-acrylic acid). Desalination 348: 16–25.
160 Liu, H., Zhao, X., Jia, N. et al. (2018). Engineering of thermo-/pH-responsive
membranes with enhanced gating coefficients, reversible behavior and
self-cleaning performance through acetic acid boosting microgels assembly.
Journal of Materials Chemistry A 6 (25): 11874–11883.
161 Ma, B., Ju, X.-J., Luo, F. et al. (2017). Facile fabrication of composite mem-
branes with dual thermo- and pH-responsive characteristics. ACS Applied
Materials & Interfaces 9 (16): 14409–14421.
162 Gajda, M. and Ulbricht, M. (2016). Capillary pore membranes with grafted
diblock copolymers showing reversibly changing ultrafiltration properties
with independent response to ions and temperature. Journal of Membrane
Science 514: 510–517.
163 You, M., Wang, P., Xu, M. et al. (2016). Fouling resistance and cleaning effi-
ciency of stimuli-responsive reverse osmosis (RO) membranes. Polymer 103:
457–467.
164 Zhao, J., Zhang, Y., Su, Y. et al. (2013). Cross-linked bovine serum albu-
min composite membranes prepared by interfacial polymerization with
stimuli-response properties. Journal of Membrane Science 445: 1–7.

You might also like