You are on page 1of 27

Review

Membranes
www.advmatinterfaces.de

Specially Wettable Membranes for Oil–Water Separation


Yibin Wei,* Hong Qi, Xiao Gong, and Shuaifei Zhao

Additionally, accidental release of crude oil


The oil–water separation has attracted heightened attention because of the and petrochemical products that causes
ever-increasing amounts of oily water produced from the daily activities of considerable environmental damage and
humans and industrial processes. Membrane technology as an advanced economic losses have occurred frequently
in the past decades. Therefore, cost-
water purification approach has offered an indispensable option due to its
effective separation of oil–water mixtures
cost-effective, energy-efficient, and easy-to-operate characteristics. However, is an urgent need for water, energy and
traditional membrane materials suffer from severe fouling, which counteracts environmental securities.[4–6]
the superiority of applying membranes in oil–water separation applications. Generally, the difficulties in sepa-
Thanks to the emerging bioinspired interface research, a series of special rating oil–water mixtures predominantly
wettability not only endows membrane surfaces with outstanding antifouling depends on the size of the oil droplet
and the stability of the mixture.[7] Oil–
properties but also breaks though the long-standing tradeoff effect between water mixtures are commonly classi-
membrane permeability and selectivity. In this review, the recent advances fied into three categories in terms of the
of membranes used for oil–water separation with special wettability and diameter of the oil droplet: free floating
perspectives on the on-going research are presented. The authors first oil (>150  µm), dispersed oil (20–150 µm)
discuss the wetting phenomenon on membrane surface, and then highlight and emulsified oil (<20 µm).[8,9] The most
challenging oily water exists in the form
the gradually evolved specially wettable system and its coupling with
of oil–water emulsions that can be further
membrane materials. Next, relatively comprehensive preparation methods classified into oil-in-water (o/w) and water-
and applications of oil–water filtration membranes utilizing special wettability in-oil (w/o) emulsions.[10,11] In an o/w
are summarized. Finally, the authors conclude the current achievements and emulsion, the continuous phase is water
challenges in specially wettable membranes for oil–water separation and and the dispersed phase is oil, while in a
outlook the future in this field. w/o emulsion oil is the continuous phase.
Surfactant-stabilized oil–water emulsions
have low interfacial tension between the
oil and water phases, which deters the
1. Introduction coalescence of droplets, contributing to the difficulty in treating
oil–water emulsions.[12,13]
Sustainable supply of clean water is one of the most important Numerous physical and chemical techniques have been
challenges for human societies. Since the industrial revolution, applied to separate oil–water mixtures, such as gravity sepa-
oily water has been produced in ever-increasing amounts due to ration, air floatation, adsorption, coagulation, centrifugation,
the daily activities of humans and industrial (e.g., petrochem- electrochemical and photocatalytic treatment.[10,14,15] However,
ical, oleochemical, and metallurgical) processes (Figure 1).[1–4] these conventional methods have their own limitations, e.g.,
undesirable separation efficiency, large footprint, secondary
pollution or high energy consumption.[16] Therefore, seeking
efficient approaches for oil–water separation has attracted
Y. Wei, Dr. S. Zhao
Department of Environmental Sciences growing research interest.[4,17–19]
Macquarie University Membrane technology as an advanced separation and puri-
Sydney, NSW 2109, Australia fication method is gaining heightened attention due to its high
E-mail: yibin.wei@hdr.mq.edu.au efficiency, low energy cost, ease for continuous operation, and
Prof. H. Qi low footprint.[20,21] Membrane separation technology has been
State Key Laboratory of Material-Oriented Chemical Engineering
Membrane Science and Technology Research Center
successfully used in desalination, wastewater treatment, food
Nanjing Tech University processing, chemical and pharmaceutical industries for a few
Nanjing 210009, Jiangsu, China decades, demonstrating its significant technical and economical
Prof. X. Gong merits.[22] For oily water treatment, pressure-driven membrane
State Key Laboratory of Silicate Materials for Architectures processes including microfiltration (MF), ultrafiltration (UF),
Wuhan University of Technology nanofiltration (NF), and reverse osmosis are widely used, espe-
Wuhan 430070, China
cially in produced water treatment.[23–26] Other membrane pro-
The ORCID identification number(s) for the author(s) of this article
can be found under https://doi.org/10.1002/admi.201800576.
cesses such as thermal-driven membrane distillation (MD) and
osmotically driven forward osmosis have also been attempted
DOI: 10.1002/admi.201800576 for oily wastewater purification.[27] However, the major problem

Adv. Mater. Interfaces 2018, 1800576 1800576  (1 of 27) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

in these membrane processes for oil–water separation is


membrane fouling, because surfactant adsorption or pore plug- Yibin Wei received his B.Eng.
ging by oil droplets can lead to diminished flux.[28] Membrane degree in chemical engi-
cleaning requires extra maintenance downtime, chemicals neering jointly offered by
and energy, which increases operation costs and may counter- Nanjing Tech University and
balance the advantages of membrane separation. Membrane University of Strathclyde in
fouling resulting from oily organic matters remains one of the 2014, and then he com-
greatest challenges in the treatment of oily wastewater.[29,30] pleted his M.Eng. degree
Ideal membrane materials should offer high selectivity, from Gachon University in
permeate flux, stability and fouling resistance.[21] In the 21st 2016. After working as a
century, nanotechnology has developed rapidly and led to a visiting scholar in Prof. Hong
material science revolution, which also offers new possibilities Qi’s group at Nanjing Tech
for oil–water separation membranes.[31] Learning from nature, University (2016–2017), he
materials with special wettability constructed by well-defined moved to Macquarie University for pursuing his Ph.D.
nanohierarchies have received increasingly extensive atten- degree. His areas of research interest include novel mem-
tions, and many novel functional interfacial nanomaterials brane materials preparation and application.
have been fabricated and applied in industry.[32] Recently,
new materials with special wettability (e.g., superlyophobic or Hong Qi is a professor in
superlyophilic) have been developed via surface engineering the State Key Laboratory of
(mainly by adjusting the surface chemical composition or Material-Oriented Chemical
topography).[33–35] Based on this new theory, a large number Engineering, Membrane
of interfacial or porous materials with special wettability have Science and Technology
been fabricated and used for oil–water separation.[8,17,36] It is Research Center, Nanjing
worth mentioning that some of these materials are even bi- Tech University. He received
functional that they could not only removal insoluble oils but his Ph.D. from Nanjing Tech
also soluble dyes, based upon different mechanism such as University in 2001 with Prof.
molecular sieving, adsorption, photodegradation, etc.[37–40] Nanping Xu. He was a vis-
Membranes, as the typical filtration materials, have been iting scholar (2007–2008) at
designed and improved with special surface wettability for oil– Inorganic Membranes Group,
water separation.[41] Up to now, material scientists have made University of Twente. His current research interest focuses
remarkable achievements in selective membranes for sepa- on porous ceramic supports and ceramic membranes with
rating oil–water mixtures to improve separation efficiency and applications in liquid and gas separation.
antifouling properties of the membranes.
Recently, many research and review articles[32,42–46] have Xiao Gong is a full professor
been published regarding applications of various wettable in the State Key Laboratory
materials for oil–water separation. However, few reviews on of Silicate Materials for
novel membrane materials with special wettability and their Architectures, Wuhan
corresponded improvements in membrane processes have University of Technology,
been summarized and published. Membrane processes are China. He received his
mature practical separation and purification technologies, Ph.D. in materials science
thus the innovation in membrane materials and their effects from Zhejiang University in
on membrane oil–water separation applications deserve to be 2009. Prior to joining Wuhan
reviewed from the view of both materials science and prac- University of Technology, he
tical engineering applications. In this review, we discuss fun- was a postdoctoral associate
damental knowledge on membrane materials with special in Prof. Lei Li’s group at the
wettability for oily water treatment. A brief introduction of University of Pittsburgh, USA. His research interests are on
the essence behind specially wettable interfacial materials for self-assembly and polymeric nanomaterials, polymer thin
oil–water separation, the wetting phenomenon, is in Section 2. films, colloid surface and interface chemistry, fabrication
We then summarize the latest advances of functionalized and characterization of nanostructured materials; interfa-
membranes with special wettability involving definitions and cial behavior of ionic liquid on solid surfaces.
general preparation procedures in Section 3. In Section 4,
membrane materials with special wettability are classified into
metallic-based, polymeric-based, ceramic-based and emerging 2. Wetting Phenomenon
nanomaterials-based membranes according to the sup-
port. The relatively comprehensive preparation methods and Wetting is a common phenomenon that can be widely found
applications of these membranes in oil–water separation are anywhere, generally defined as the ability of a liquid get-
discussed. Finally, conclusions and outlook on future develop- ting in contact with a solid surface through intermolecular
ment of membrane materials with special wettability for oily interactions between the two phases.[47] The oil–water separa-
water treatment are given. tion, in essence, is that wetting phenomenon occurs at the

Adv. Mater. Interfaces 2018, 1800576 1800576  (2 of 27) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

Figure 1.  Schematic illustration of oil–water mixtures produced in various industrial processes.

interface of the solid, air, water and oil phases.[17] Owing to a spherical droplet on the surface in air (Figure 2a). The
massive efforts on the research into a series of special wet- Young’s CA, θ is a function of the interfacial tension between
tabilities inspired from nature, interfacial materials with the solid–liquid (SL), solid–vapor (SV), and liquid–vapor (LV)
super-wettability or anti-wetting properties have made great interfaces:
breakthrough for the application of oil–water separation. In
recent years, wetting phenomenon in membrane applications γ SV − γ SL
cosθ = (1)
has aroused significant interest.[48,49] Mass transfer resist- γ LV
ance and separation efficiency are influenced by wetting phe-
nomenon that is related to membrane surface energy, pore where γSV, γSL, and γLV, are the surface tensions at the solid–
size and porosity. Thus, a comprehensive understanding of vapor, solid–liquid, and liquid–vapor interfaces, respectively.
wetting phenomenon is of great significance in advancing The Young’s equation is based on the assumption that the
membrane materials and improving membrane separation surface is chemically homogenous and topographically smooth.
performance. This is, however, not always correct because real surfaces do
not have perfect smoothness, rigidity, or chemical homogeneity.
The relationship between surface roughness and wettability was
2.1. Contact Angles defined in 1936 by Wenzel,[51] who stated that increasing surface
roughness would enhance the wettability caused by the chem-
Wettability can be studied by characterizing the contact istry of the surface. When homogenous wetting occurs on a solid
angle of a surface with a given liquid. Contact angle (CA), θ, rough surface (Figure 2b), the Wenzel CA (θW) is determined by
is a quantitative measure of wetting, which is geometrically
defined as the angle formed by a liquid at the three-phase cosθ W = r cosθ (2)
boundary where a liquid, gas and solid intersect. According
to the well-known Thomas Young equation,[50] when a sessile where θW is the apparent contact angle which corresponds to
drop is placed on an ideally smooth surface, it normally forms the stable equilibrium state, r is the roughness factor, defined

Adv. Mater. Interfaces 2018, 1800576 1800576  (3 of 27) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

2.2. Wetting States in Oil–Water Separation

Wetting is an intrinsic property of a surface,


governed by the chemical composition and
morphology of the surface.[49,56,57] Chemical
composition of the surface is the internal char-
acter of the material, which determines the
surface tension. Generally, oil has smaller
surface tension than water, which is the
reason why oils could be separated from
water through porous interfacial mate-
rials with special wettability. In latest a few
studies,[58–60] the concept of local texture angle
(LTA), ψ, was proposed as a supplement to
describe wetting on a hierarchical textured
Figure 2.  Wetting state of a liquid droplet on different solid surfaces in air and corresponded surface, which offers new clues for designing
contact angle (CA) models: a) Young’s, b) Wenzel’s, and c) Cassie–Baxter’s. d) Schematics of oil–water separation interfacial materials.
advancing CA, receding CA, and droplet moving along an inclined surface with a tilting angle.
Those researchers believed that formation of
the stable Cassie–Baxter state is possible only
as the ratio of the actual area of the solid surface to the if the Young’s CA is greater than the LTA. As shown in Figure 3,
apparent surface area (r  = 1 for a smooth surface and r  > 1 two types of surface texture are compared. When texture angle
for a rough one), and θ is the Young’s CA. Notably, The ψ > 90° and θ ≥ ψ, the surface texture could result in the forma-
Wenzel’s equation is based on the assumption that the liquid tion of a composite interface. If θ  < 90°, which is common for
penetrates into the roughness grooves. In cases where the oil with low surface tension on most surfaces, then the surface
liquid does not penetrate into the grooves, the Wenzel’s texture cannot maintain a stable composite interface, regardless
equation does not apply. of its surface energy or composition. However, for the same low
When dealing with a heterogeneous rough surface, a more surface tension oils with θ < 90°, it is possible to fabricate a het-
complicated model was proposed by Cassie–Baxter[52] (Figure 2c) erogeneous interface as long as θ  ≥  ψ. Such surface geometry
with ψ  < 90° is known as re-entrant texture. Re-entrant texture
allows formation of a heterogeneous interface with low surface
cosθ CB = rf f SL cosθ + f SL − 1 (3)
tension oils, which may lead to oleophobic properties. Further
concepts and design strategies for oil–water mixtures separation
where fSL represents the solid–liquid fraction under the con- membranes with special wettability are introduced in Section 3.
tact area. Distinguished from the total roughness factor r, rf
is defined as the roughness ratio of the wet part of the solid
surface and it always shows a lower value than r. Besides r, 2.3. Wetting with Membranes
the Cassie–Baxter equation suggests that the solid–liquid frac-
tion (solid–air fraction) is also an influencing parameter to the With the rapid development of research on membrane separa-
apparent CA. tion technologies, a comprehensive understanding of wetting
The difference in the three above-mentioned wetting states phenomenon with membranes plays a more and more impor-
is their CA hysteresis, which reflects the surface neterogeny.[53] tant role in improving membrane performance. For pressure-
The CA hysteresis is quite common on rough and chemical driven separation membranes, two parameters associated with
heterogeneous surfaces. CA hysteresis, i.e., the difference membrane pores (i.e., pore size and porosity) and breakthrough
between the advancing and receding CAs (Figure 2d) for the pressure should be considered in wetting study.[48]
Wenzel’s and Cassie–Baxter’s wetting states are totally different. The first critical physical characteristic is the breakthrough
On a given surface, advancing CA is the maximum value while pressure (ΔPc), defined as the maximum pressure applied on
receding CA is the minimum value, and
both of them arises due to the presence of
multiple metastable energy states on rough
and heterogeneous surfaces.[54] Typically,
CA hysteresis in the Wenzel’s state is larger
than that in the Cassie–Baxter’s state. This
is because the solid–liquid interface is
pinned on the textured surface. By contrast,
a composite solid–liquid–air interface in
the Cassie–Baxter’s state leads to lower CA
Figure 3. Composite interfaces and hierarchical texture based on Cassie–Baxter’s model.
hysteresis and higher apparent CAs when a) A concave texture with ψ > 90° and θ > 90°. b) A lower surface tension liquid (θ < 90°) on
the contact area between the solid and the convex, re-entrant texture (ψ < 90°). Reproduced with permission.[7] Copyright 2015, Cambridge
liquid is small.[55] University Press.

Adv. Mater. Interfaces 2018, 1800576 1800576  (4 of 27) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

Figure 4.  Summary of membrane materials with special wettability for oil–water separation.

a membrane before the membrane pores are wetted by a given review. These membrane materials with special wettability for
liquid (also called liquid entry pressure, LEP). It can be meas- oil–water separation are summarized in Figure 4.
ured by detecting the formation of the first liquid drop on the
permeate side of the membrane. For small uniform pores with
cylindrical geometry, the breakthrough pressure (ΔPc) can be 3.1. Superhydrophobic–Superoleophilic Membranes
determined by the Laplace–Young equation[61]:
Inspired by SHB surfaces from plants and animals in nature
2γ cosθ Y (e.g., lotus leaf, legs of the water strider, and the eyes of mos-
∆Pc = L (4)
rp quitos), a number of surfaces with similar wettability have been
applied for selectively separating oil–water mixtures.[62–65] Porous
where γL and rp represent the surface tension of the liquid membranes with both SHB and SOL abilities are typical oil–
and the maximum membrane pore radius, respectively. As water separation materials, which can allow oil permeate while
discussed above, Young’s equation is not suitable to describe repelling water. Such functional membranes can separate oil
a real rough membrane surface and the relationship between from the oil–water mixture and water-in-oil emulsions. The first
breakthrough pressure and surface wettability could be further artificial SHB-SOL membrane was fabricated by Feng et al.[66]
corrected, but Laplace–Young equation clearly points out in 2004 by spraying a polytetrafluoroethylene (PTFE) coating
that wettability of the membrane surface could significantly layer on a stainless-steel mesh (SSM). The prepared membrane
influence pressure-driven membrane filtration processes. possessed micro- and nanoscale surface roughness, exhibiting a
water CA (WCA) of 156° and a diesel oil CA (OCA) of 0 ± 1.3°
(Figure 5). The membrane surface showed the self-cleaning
3. Special Wettability on Membrane Surfaces property, similar wettability to lotus leaves.
Thereafter, many groups have been working toward fabri-
Various types of interfacial materials with special wettability cating this type of membranes with such a wettability for oil–
have emerged.[32] From the previous discussion, we can see water separation. Generally, such membranes are developed by
that chemical composition and architecture of the surface coating a film material with selective wettability onto porous
are two keys to developing selective membranes for oil–water supports. To prepare a SHB-SOL membrane, it is essential to
(emulsion) separation. In this section, we will systematically construct superhydrophobicity using materials with HB nature
introduce advanced membrane materials with special wetta- in chemical composition. In addition to chemical composition,
bility, including superhydrophobic–superoleophilic (SHB-SOL), the surface microstructure (roughness) is another key factor in
superhydrophilic–underwater superoleophobic (SHL-UWSOB), the construction of specially wettable surfaces. Therefore, SHB-
superhydrophilic–superoleophobic (SHL-SOB), superam- SOL membrane materials can be constructed by two routes: the
phiphilic (also known as superhydrophilic–superoleophilic, first is to introduce sufficient roughness onto a hydrophobic
SHL-SOL in-air), and superomniphobic (also known as supe- surface and the second is to tailor the surface chemical compo-
rhydrophobic–superoleophobic, SHB-SOB) membranes from sition with a material of low surface energy. A SHB-SOL mem-
their definitions, working principles to fabrication strategies. In brane usually has a water CA > 150° and a tilting angle <10°
addition to these membranes with fixed wetting abilities, smart (also known as sliding angle, SA).
membranes using switchable wettability and the novel Janus Although plenty of SHB-SOL membranes have been
membranes used for oil–water separation is also covered in this developed for oil–water separation, there are some intrinsic

Adv. Mater. Interfaces 2018, 1800576 1800576  (5 of 27) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

Figure 5.  Scanning electron microscopy (SEM) images of a) the modified SSM membrane with an average pore diameter of ≈ 115 µm; b) the coating
film ((I) ball-like and (II) block-like microstructured shapes); c,d) SEM images of the balls and blocks, and (III) the nanostructured craters with high
resolution; e) water droplet on the modified membrane surface with a WCA of 156.2° ± 2.8°, and f) spreading and permeating behaviors of a diesel oil
droplet on the membrane. Reproduced with permission.[66] Copyright 2004, Wiley.

limitations for membranes with this type of wettability. First efficiency (>99%) for separating the mixing systems of water-
of all, water with higher density may cause extra transfer crude oil, water-gasoline and water-diesel.
resistance between the membrane and oil, particularly in So far, there are two types of biomemimetic SHL-UWSOB
gravity-driven membrane separation processes. Another fatal membranes materials reported. The first one is inspired from
problem is membrane fouling caused by the oils adsorption rough fish scales that are covered by a thin layer, and the
onto membrane surface or membrane pores, especially when layer could consist of functional hydrogels or other coating
the mixture is viscous and sticky, resulting in severe flux films.[67,73] The other type is based on short clams whose shells
decrease. These challenges will raise operation downtime, are always clean under water.[74] It is believed that the inorganic
membrane cleaning, and membrane replacement costs when CaCO3 composition and hierarchical structure of the shells are
using SHB-SOL membranes. the key to maintaining the excellent underwater oil-repellent
property. To develop SHL-UWSOB membranes, there are two
typical strategies. The first is to determine the chemical com-
3.2. Superhydrophilic–Underwater Superoleophobic Membranes positions of the material surfaces (i.e., manipulate the surface
chemistry) that are adhesive to water and repulsive against oil.
It is known that fish scales can be protected from oil contamina- The second is to create sufficient roughness (i.e., construct the
tion in underwater environment. A study pointed out that fish surface architecture) on the membrane surfaces. These mate-
scales were SHL (WCA < 5°) in air and UWSOB (OCA ≈ 156°).[67] rials always exhibit an OCA > 150° when immersed in water.
Learning from this, a series of SHL-UWSOB interfacial mate- Although membranes with SHL-UWSOB properties can be
rials have been rapidly developed.[67–70] Compared with SHB- used for separation of oil–water mixtures by gravity, and are
SOL membranes (oil-removal), SHL-UWSOB membrane more resistant to fouling, they are unsuitable for the separa-
materials aim for water-removal, which overcomes the draw- tion of surfactant free o/w or w/o emulsions. This is because
backs (i.e., easily fouled and even clogged by oils) in using both oil and water can easily permeate through them, unless all
SHB-SOL membranes.[71] SHL-UWSOB membrane materials pores of the membrane are pre-wetted by water. Consequently,
can separate water from oil–water mixtures and oil-in-water oil permeates through the membrane if water dries out from
emulsions, where water with higher density tends to permeate even a single pore of the superhydrophilic membrane, which
through the surperhydrophilic membrane while oil is repelled. typically happens in a matter of minutes.
In 2011, Xue et al.[72] reported a SHL-UWSOB polyacrylamide
(PAM) hydrogel coated membrane for the first time. The PAM
hydrogel-coated SSM membrane had a rough surface with 3.3. Superhydrophilic–Superoleophobic (In-Air) Membranes
a discrete papillae structure (80–500 nm in size) and showed
an underwater superoleophobicity with 1,2-dichloroethane CA Considering to overcome the limitations of SHL-UWSOB
of 155.3 ± 1.8° (Figure 6). The as-prepared membrane also membranes discussed above, membranes with SHL-SOB
exhibited a low oil-adhesion under water and high separation properties in air seem more promising for removal water phase

Adv. Mater. Interfaces 2018, 1800576 1800576  (6 of 27) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

Figure 6.  a) SEM images of the PAM hydrogel-coated SSM membrane (average pore size ≈50 µm). b) Photos of an oil droplet (1,2-dichloroethane,
DCE, 2 µL) on the membrane surface in water with a CA of 155.3° ± 1.8° and c) an SA of 2.6° ± 0.5°. d) Optical photos of the dynamic UW oil-adhesion
measurements on the membrane. An oil droplet (DCE, 5 µL) was used to contact the surface. e) Oil–water separation tests: the coated membrane
was mounted between two glass tubes and a mixture of crude oil and water was put into the upper glass tube. Water selectively permeated through
the coated membrane, while the oil was blocked in the upper glass tube. Reproduced with permission.[72] Copyright 2011, Wiley.

from oil–water mixtures. A hydrophilic surface generally displays a challenge. Until 2012, Yang et al.[78] first developed a poly
oleophilicity because oil commonly has lower surface tension (diallyldimethylammonium chloride)–perfluorooctanoate/SiO2
(≈20–60 mN m−1) than that of water (≈72 mN m−1). Therefore, it (PDDA–PFO/SiO2) nanocomposite coating film with both
is very challenging to fabricate hydrophilic–oleophobic (HL-OB) superhydrophilicity and superoleophobicity. The abundant
interfacial materials. Several research groups found that HL-OB fluorinated groups together with carboxyl and quaternary ammo-
materials contain some stimuli-responsive substances, simul- nium groups from the PDDA–PFO complex film generated an
taneously exhibiting HL and OB properties based on a positive HL-OB surface. In addition, the embedded SiO2 nanoparticles
interaction with polar liquids and a negative interaction with (NPs) provided micro- and nanoscaled hierarchical structures,
nonpolar liquids.[75,76] For example, the stimuli-responsivity which enhanced the superhydrophilicity–superoleophobicity of
could be induced by a layer of poly­electrolyte-fluorosurfactant the membrane. When the PDDA–PFO/SiO2 film was deposited
complexes where the polyelectrolyte are HL while the fluorosur- on SSMs via spray-coating, the prepared metallic membranes
factant (an oppositely charged surfactant) are OB.[77] exhibited a water permeation and oil repellent behavior, and
Although some moderate HL-OB surfaces have been successfully achieved separation of water-oil mixture (Figure 7).
reported,[75,77] development of SHL-SOB membranes is still Besides, the coated-membranes are reusable after water rinsing.

Figure 7.  a,b) SEM images of the SSM membrane spray-coated with PDDA-PFO/SiO2 nanocomposite. c) Water droplet spreading on and permeating
through the membrane. d) A hexadecane droplet on the membrane surface with a CA of 157°. e,f) An oil–water separation test was conducted with
the modified membrane. Adapted with permission.[78] Copyright 2012, Royal Society of Chemistry.

Adv. Mater. Interfaces 2018, 1800576 1800576  (7 of 27) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

Figure 8.  Schematic illustration of the superhydrophobicity and superoleophobicity of the functionalized PET fabrics membrane. a) When pre-wetted
by oil, water droplet is blocked on the ZnO nanostructure. b) When pre-wetted by water, water can permeate the ZnO nanostructure while oil droplets
are blocked. Reproduced with permission.[82] Copyright 2015, Elsevier.

In the same year, Tuteja et al.[79] reported a hygroresponsive ability, the recent reported membranes are mainly based upon a
membrane with superhydrophilicity and superoleophobicity prewetting induction mechanism.[80]
both in air and underwater. They successfully coated fluorodecyl One of the first reports on superamphiphilic membranes was
polyhedral oligomeric silsesquioxane (POSS) and cross-linked from Tao et al.[81] in 2014. They fabricated porous superamphi-
polyethylene glycol diacrylate (x-PEGDA) complex onto porous philic poly (vinylidene fluoride) (PVDF) membranes using a
substrates, such as SSMs and polyester fabrics by dip-coating. nonwoven fabric (NWF) as a template via the phase inversion
In air, the membrane surface was superoleophobic (rapeseed method. Interestingly, the both UW oil and UO water droplets on
OCA = 152°) due to the fluorodecyl POSS aggregates. When the membrane surface have CAs of 157° and 156°, respectively.
the membrane was immersed in an aqueous environment, the It is believed that the special (UWSOB and UOSHB) wettability
fluorodecyl POSS aggregates disappeared due to surface recon- was mainly attributed to the hierarchical surface structure of the
figuration caused by water molecules. The membrane could membrane. Varying microstructures with four different sizes
separate surfactant-stabilized o/w and w/o emulsions with diam- and shapes, involving randomly aligned grooves, closely stacked
eters ranging from 10–20 µm by gravity with high separation spherulites, petaloid microstructures, and nanofibrils, lead to a
efficiency (>99%). Significantly, they utilized a coupling process rather raspy membrane surface (> 3500 nm in roughness) and
with a hygro-responsive membrane module and a conventional the corresponding superamphiphilicity.
HB-OL membrane module assembly in tandem for continuous Shortly after, Xiong et al.[82] reported an atomic-layer-
separation of oil–water emulsions, and it was found that the deposition-enabled nonwoven superamphiphilic membrane
fluxes of both water and oil did not reduce over a period of 100 h. with hierarchical ZnO nanostructures for switchable water-oil
For SHL-SOB (in-air) membranes, intermolecular force separation. The membranes could separate any phase in the
between membrane surface and water varies, while that oil–water mixture simply by pre-wetting the membrane surface
between membrane surface and oils stays the same. In compar- with the heavier component, showing excellent separation
asion with SHL-UWSOB, SHL-SOB membranes showed more efficiency and stability even under gravity driving force. The
technical merits in oil–water separation applications. However, superamphiphilicity was attributed to the prewetting process, in
fabrication methods of the SHL-SOB membranes are more which a thin layer of water or oil “cushion” was formed on the
complicated and challenging. Although a growing number of interface between the ZnO hierarchical nanostructure and water
research groups has been working on the development of SHL- or oil, resulting in UWSOB or UOSHB properties (Figure 8).
SOB membranes and the corresponding application of these The crucial role of prewetting was also identified in by another
membranes, membranes with such a special wettability are still group[83] and they successfully developed a switchable superam-
lack of investigation. phiphilic membrane for oil–water separation as well.
The superamphiphilic ability endows membranes with
switchable transport characteristics, which features unique
3.4. Superamphiphilic (In-Air) Membranes separation efficiency, permeability and antifouling property
for various o/w and w/o emulsion separations. However,
Superamphiphilic, literally, represents superhydrophlic and understanding of the structure-chemistry-superamphiphilicity
superoleophilic (in-air). Superamphiphilic membranes are mechanisms is still insufficient. Other emerging switchable
switchable in wettability, i.e., UWSOB or under oil (UO) SHB. superamphiphilic mechanisms, such as CO2-induced,[84] pH-
Compared with other types of membranes, superamphiphilic sensitive,[85] photoresponsive[86,87] superamphiphiles, may
membranes are capable for separating a wider scope of oil– provide new strategies to fabricate novel oil–water separation
water emulsions, including both o/w and w/o emulsions. When membranes.
placed in o/w emulsions, water in the rough surface structure
can form a barrier between oil and the membrane surface to let
water permeate but repel oil, leading to UW superoleophobicity. 3.5. Superomniphobic Membranes
In contrast, superamphiphilic membranes can block water and
let oil permeate showing UOSHB wettability when used for In contrast to membranes with selectively superwetting
w/o emulsion system. To achieve this superior superwetting ability for water against oil or vice versa, superomniphobic

Adv. Mater. Interfaces 2018, 1800576 1800576  (8 of 27) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

1H,1H,2H,2H-perfluorodecyltrichlorosilane
to reduce the surface energy. The prepared
superomniphobic membranes could keep
their super liquid-repellency (CAs > 150°)
for water, methyl diethanolamine (MDEA),
ethylene glycol (EG) and hexadecane.
Although the membranes are designed for
membrane contactor applications and the
preparation method seemed complex, it still
offers useful insights into developing and
using superomniphobic membranes.
In addition to MD and membrane
contactor applications, omniphobic mem-
branes have also been used for oil aerosol
removal. Feng et al.[101] fabricated an omni-
phobic 1H,1H,2H,2H-perfluorodecyl acrylate
Figure 9.  Schematic illustration of the synthetic routes of the modified membrane: a) pristine (PFDAE)-grafted ZnO@polytetrafluoroethylene
PTFE fibrils, b) coating with ZnO film, and c) grafting monomer under plasma. Reproduced (PTFE) membrane with enhanced antifouling
with permission.[101] Copyright 2016, American Chemical Society. functionality and high removal efficiency.
They used atomic-layer deposition (ALD) to
membranes are super anti-wetting, repelling both oil and coat a layer of ZnO nanoparticles onto porous PTFE matrix to
water. As discussed above, the surface tension of oil is gener- increase surface area, and then grafted PFDAE with plasma
ally smaller than that of water, SHB materials in air often suf- (Figure 9). Consequently, the membrane surface showed both
fers from oil-wetting.[88] In most superhydrophobic cases, the superhydrophobicity and oleophobicity with a WCA and an
intermolecular forces between oil molecules and solid surface OCA of 150° and 125°, respectively. The membrane exhibited
molecules/atoms are attractive forces, which allows oil to wet exceptional organic aerosol removal efficiencies (>99.5%).
the surface. However, recently materials with both SHB and According to above cases, modifying straight- or branched-
SOB properties have been fabricated.[89–94] SHB surfaces dis- chain molecules with fluoridated groups oriented outward on
play apparent WCA > 150° and low CA hysteresis, while SOB the surfaces, the intermolecular repulsive force between oil
surfaces display CA > 150° and low CA hysteresis to low sur- molecules and solid surface can be enhanced.[95] Omniphobic
face tension substances (i.e., oils). Superomniphobic surfaces membranes opens new opportunities for MD by expanding
display both superhydrophobicity and superoleophobicity.[95] its applications in treating oily wastewaters with low surface
Preventing wetting of MD membranes is particularly chal- tension contaminants. Other membrane contactors and direct
lenging in treating shale gas wastewater with high levels of membrane filtration process using omniphobic membranes
surfactants or low surface tension oily contaminants as these are also very promising for the treating media with low surface
impurities may significantly reduce the liquid entry pressure tension substances. It is still challenging to fabricate superom-
of the membrane.[96,97] To overcome this problem, novel MD niphobic membranes and further studies are needed to explore
membranes with omniphobic surfaces that repel not only water other fabrication techniques and correspondingly optimize the
but also oil with low surface tension were reported by Elimel- performance of superomniphobic membranes.
ech’s group in 2014.[98,99] They fabricated an omniphobic mem-
brane for MD by coating a hydrophilic glass fiber membrane
with silica NPs, followed by subsequent surface fluorination 3.6. Smart Wettability in Membranes
and polymer coating. The omniphobic membrane had a high
WCA near 150° and could not be wetted by mineral oil, eth- A surface sensitive to a certain external stimulus and then
anol, and decane (CA in the range of 70°–110°). Although the producing a peculiar response, is defined as smart surface.[102,103]
oleophobic property of the membrane was not superior because Artificial smart interfacial materials with switchable wetting
the WCA of their membrane was lower than 150° and the CAs behaviors have attracted growing interest, particularly in their
of the membrane with those low surface tension liquids are still fabrications and applications for oil–water separation.[104,105]
hard to surpass 140°, it still opened a new way to develop mem- Besides superamphiphilic (in-air) membranes that mainly
brane materials with special wettability. follow a special pre-wetting media-induced switchable
Recently, a real sense of superomniphobic membranes were wettability mechanism, switchable oil–water separation based
successfully fabricated and used in contacting membrane tech- on smart membrane surface can be realized by external stimuli,
nology, which can efficiently prevent wetting of membrane pores such as light,[106–108] pH,[109–112] temperature,[113,114] electrical
by the capturing media. Geyer et al. [100] field,[115,116] and magnetic field (Figure 4).[117–119] Most recently,
fabricated chemically and
mechanically robust superomniphobic flat-sheet membranes membrane materials with smart wettability based on these
based on a polyester fabric. Thick hydrophilic fabrics with inter- mechanisms, have been developed for controllable oil–water
woven networks of fibers were first immersed in toluene con- separation. These smart membranes have opened new strategies
taining trichloromethylsilane forming silicone nanofilaments, for the treatment of oil contaminated water to meet different
and then were activated in an oxygen plasma and modified with demands.

Adv. Mater. Interfaces 2018, 1800576 1800576  (9 of 27) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

Figure 10.  Photoinduced wettability of the aligned ZnO nanoarrays-coated mesh membrane: a) Images of a water droplet on the modified membrane
when placed in dark (left) and under UV irradiation (middle) in air with a CA of ≈155° and ≈0°, respectively. The water droplet can permeate through
the membrane (right). b) Images of an oil droplet (DCE) on the mesh membrane in air (left) and UW (middle) with a CA of ≈0° and ≈156°, respectively,
and an SA of ≈2.6° (right). c) The membrane was mounted between a glass funnel and glass tube for oil–water mixture separation. A small amount
of water was added into the glass funnel to form a water environment and ensure the stability of the device before the mixture of crude oil and water
was poured in. d) Once UV light irradiated the membrane, water selectively permeated through the membrane, while the oil was repelled and kept in
the glass funnel. Reproduced with permission.[121] Copyright 2012, Royal Society of Chemistry.

Light is a very common external stimulus of smart Polymers with acid or basic functional groups are always
materials. Many transition-metal oxides (e.g., ZnO, TiO2, sensitive to pH conditions showing pH-driven properties. By
WO3, and V2O5) have proved their photoresponsive wettability grafting a block copolymer comprising of pH-responsive poly
under light irradiation.[120] Tian et al.[121] created a photo­ (2-vinylpyridine) (P2VP) and HB-OL polydimethylsiloxane
induced oil–water separation membrane with switchable (PDMS), Zhang et al.[105] developed a series of pH-responsive
superhydrophobicity–superhydrophilicity and UW supero- smart membrane surfaces exhibiting switchable superoleophi-
leophobicity, based on a ZnO nanorod-deposited SSM. The licity and superoleophobicity for controllable oil–water separa-
as-prepared mesh membrane exhibited interesting UV light- tion. The P2VP block could alter its chemical conformation and
dominated wettability for separating oil–water mixtures the corresponding wettability via protonation and deprotona-
(Figure 10). Gao et al.[122] also prepared a double-layer TiO2- tion under varying pH environments. Using their membranes
based mesh membrane exhibiting switchable wetting behav- to separate a mixture of gasoline and water at pH 6.5, gasoline
iors under UV light irradiation. The upper layer was a TiO2 can pass though the membrane quickly while water is blocked.
coating film with hybrid micro- and nanostructures prepared However, when the membrane is pre-wetted in an acidic water
by a facile hydrothermal process, and the lower layer was the with a pH of 2.0, the opposite separation behavior is realized
same TiO2 mesh film but modified by octadecyl phosphonic (i.e., water removal). Xiang et al.[111] developed a smart PVDF
acid (ODP). The prepared membrane is SHB-SOL allowing membrane by incorporating pH-responsive N,N-dimethyl-
fast oil penetration, while the surface SHB property can be aminoethylmethacrylate (DMAEMA) hydrogels for oil–water
converted into SHL due to the photodecomposition of ODP by separation. The prepared membrane was UO HB and UW
TiO2 under UV light. OL at a pH of 7.4, but UW OB at a pH of 2.0 (Figure 11). The

Figure 11.  pH-responsive wettability of the PVDF membrane. a) A static WCA (4 µL water) at pH of 2.0 and 7.4 under n-hexane; b) an oil droplet
(chloroform, 3 µL) on the membrane surface in pure water at pH 7.4; c) an oil droplet (chloroform, 3 µL) on membrane surface in HCl-induced acidic
water with pH of 2.0. d) Schematic representation of the switchable oil permeability on the modified PVDF membrane under water at a pH of 7.4
(upper) or 2.0 (lower). Reproduced with permission.[111] Copyright 2015, Royal Society of Chemistry.

Adv. Mater. Interfaces 2018, 1800576 1800576  (10 of 27) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

pH-responsive wettability was caused by the protonation or water-oil on-off switch offer a promising tool for oil–water mix-
deprotonation of the tertiary amine side-groups (Figure 11d). tures separation.
The pH-responsive PVDF membrane shows excellent perme- Electric field can also endow membrane surfaces with switch-
ability, selectivity and antifouling performance in the separation able wettability. For example, Liu et al.[127] reported a reversible
of surfactant-stabilized w/o and o/w emulsions. Guo et al.[123] transition between SOB and SOL based on aligned polypyrrole
used a mixture of carboxyl- and methyl-terminated thiols to nanotube arrays, controlled by electrochemical driving poten-
modify a hierarchical structured copper mesh (CM) membrane. tial. Their discovery opens a new strategy to develop smart
By tuning the pH of aqueous media, protonation and deproto- membranes for oil–water separation. In 2012, Tuteja et al.[5]
nation phenomenon are also found on membrane surface. The applied electric filed to switch the wettability of a nontextured
membrane showed SHB-SOL to neutral and acid water yet SHL membrane and their unprecedented membrane process could
to basic water, make it a smart membrane for bi-directional separate all types of oil–water emulsions with high separation
oil–water separation. efficiency (>99.9%) by gravitiy and electric field (Figure 13).
Thermal (temperature) response has also been consid- Currently, magnetic field could not control the wettability of
ered promising for controlling the wettability of an interfacial membrane surfaces but it provides driving force to manipu-
material. Poly(N-isopropylacrylamide) (PNIPAAm) with a low late the transfer of magnetic fluid on interfacial materials with
critical solution temperature (LCST) of about 32–33 °C has special wettability for efficient oil–water separation. Recently,
been widely used to fabricate a thermal-sensitive surface with Feng et al.[128] fabricated a tri-phase water-oil-solid nanoarray
switchable wettability.[124] Xue et al.[125] reported a polymethyl- interface that are UWSOB (UW OCA = 156°). The prepared
methacrylate (PMMA)-b-PNIPAAm block copolymer modified ZnO nanoarrays consisted of uniformly aligned nanorods with
mesh membrane, which are reversible in wettability between diameter of 150–250 nm and length of ≈6 µm. After UV treat-
HL-OB and HB-OL characteristics at different temperatures ment, the resulted ZnO nanoarrays showed superhydrophilicity
(Figure 12). Interestingly, water could pass through the block and superoleophilicity in air (WCA 6° and OCA 9°). Oil transfer
copolymer-modified membrane but oils could not when the on the SOB ZnO nanoarrays could be controlled by a mag-
temperature < LCST; oil could penetrate the membrane but netic fluid via magnetic field. As shown in Figure 14, a ZnO
water could not when the temperature > LCST. Additional, nanoarray-coated glass substrate were used to investigate the
Yu et al.[126] also developed a thermal-responsive surface which top and the bottom interfaces in water. When an external mag-
shows great potential for oil–water separation applications. The netic field was put on the substrate top, the magnetic fluid
thermal-responsive silicon nanowire arrays (SiNWAs)-PNI- droplet became magnetized immediately and went toward the
PAAm surface can change its superhydrophilicity (WCA of 0°) substrate top, moving from the left to the right side by the mag-
to strong hydrophobicity (WCA of 120°) from 23 °C to 37 °C. netic field. Once the magnetic field was removed, the magnetic
Those thermal responsive membranes with controllable dual fluid was demagnetized and fell down to the oil droplet due to

Figure 12.  a) Temperature dependences of WCAs and OCAs of the PMMA-b-PNIPAAm membrane: (I) WCAs increase from 42° to 107° and (II) OCAs
drop from 137° to 36° with the temperature rise from 10 to 40 °C. Inset images (1–4) are the WCAs and OCAs obtained at 10 and 40 °C, respectively.
b) Reversible water and oil CA transition on the membrane surface at different temperatures (10 °C, < LCST; 40 °C, > LCST). c) Schematic diagram
of reversible formation of intermolecular hydrogen bonding between PNIPAAm chains and water molecules below the LCST, which results in hydro-
philicity/oleophobicity, and intramolecular hydrogen bonding between CO and NH groups in PNIPAAm chains above the LCST, which leads to
hydrophobicity/oleophilicity. Reproduced with permission.[125] Copyright 2013, Wiley.

Adv. Mater. Interfaces 2018, 1800576 1800576  (11 of 27) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

Figure 13.  a) CAs for water and hexadecane (HD) as a function of applied voltage on the nontextured substrate. b,c) The CA of HD keeps unchanged
with increasing voltage. d,e) The CA of water decreases with increasing voltage. f) A schematic representation of the sagging at the liquid–air interface.
g) An apparatus with a liquid column of free oil (dyed red) and water (dyed blue) above the membrane before applying an electric field. h) Water permeates
through while hexadecane is repelled above the membrane when a voltage V ≈ 2.0 kV is applied. Reproduced with permission.[5] Copyright 2012, Wiley.

gravity. The mixed large magnetic oil droplet was stable at the membranes with switchable wettability show great potential in
bottom ZnO nanoarray, which can be moved back to its initial various oil–water separation applications. Nevertheless, there
location driven by the external magnetic field. No losses of the are still very few reports on stimuli-responsive membranes for
oil and magnetic fluid droplets were found during the magnetic practical oil–water separation so far and the relevant research is
field-driven movement. still in its early stage. In addition, the existing types of special
In general, membrane surfaces exhibiting switchable wetta- stimuli-responsive materials are relatively rare which limits the
bility are often fabricated by top-down method (assembly rough development of stimuli-responsive membranes. Further efforts
microstructures on a stimuli-responsive material surface) or for fabricating controllable oil–water separation membranes
bottom-up method (constructing rough interface with stimuli- might be also concentrated on attempting new types of respon-
responsive molecular layers). These artificially intelligent sive materials or integrated dual-/multi-responsive materials.

3.7. Janus Membranes

The Janus concept is originated from the


double-faced Roman god Janus. Janus
materials have asymmetric properties (e.g.,
wetting and charge) on each side. Such
materials widely exist in nature. For example,
the upper side of a lotus leaf is SHB in air,
and the lower surface is UWSOB, endowing
the lotus leaf self-cleaning in air and UW oil-
repellent properties.[129] Inspired by nature,
much research efforts have been made into
Figure 14.  The manipulation and movement of a UW oil droplet by a magnetic fluid droplet
Janus materials regarding their fabrication,
on the aligned ZnO arrays. a–c) The magnetic fluid droplet movement and transfer process.
d–f) The magnetic fluid droplet combines with the oil droplet with a following movement and characterization and applications.[130] A Janus
transfer process. The red triangle in each figure is the oleophilic area for fixing the droplet. membrane generally has opposing properties
Reproduced with permission.[128] Copyright 2016, Wiley. on each side (e.g., wettability, morphology

Adv. Mater. Interfaces 2018, 1800576 1800576  (12 of 27) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

can penetrates easily through it. In contrast,


water is blocked at the HL side UO when the
membrane is reversely configured.
The directional liquid transport behaviors
make Janus membranes with integrated
selectivity, which are intriguing and useful
for development of various separation appli-
cations, especially advanced oil–water sepa-
ration. For instance, two differently wettable
surfaces of a Janus membrane can work
individually and the membrane becomes
switchable in treating oily water by changing
the side exposed to the feed. Gu et al.[138]
reported a Janus polymer/carbon nanotube
(CNT) composite membrane by grafting HL
poly(DMAEMA) and HB poly(styrene) onto
each side, respectively. They successfully
employed the Janus membrane to separate
both surfactant-stabilized o/w and w/o emul-
Figure 15.  a) Schematic illustration of synthetic routes of Janus-C membrane. b) Unidirectional sions. For the o/w emulsion separation, the
penetration demonstrated by dropwise addition of water droplets (≈ 20 µL per drop) onto HB HL side was exposed to the feed, while the
side (top) and HL side (bottom) of the membrane. c) Directional water droplet penetration membrane was overtunred for the separation
across the membrane in UO systems. The membrane allows water droplet penetration when of the w/o emulsion.
the HB side is exposed to oil and prevents water penetration when reversely aligned. Repro- Additionally, there are some stimuli-
duced with permission.[137] Copyright 2014, Wiley.
responsive Janus membranes showing
switchable wettability by adjusting the
and surface charge[131–133]), which can be fabricated by direct operating conditions. Because the asymmetric nanochannel can
asymmetric fabrication or asymmetric post-modification.[134] act as a selective ion channel dominated by pH conditions,[139]
Special transport behaviors and auxiliary functions in Janus a composite CaWO4/MnO2 nanowire membrane reported
membranes are useful for development of various advanced by Zhang et al. shows varying sets of wettability (i.e., HL/
oil–water separation processes. HL, HB/HL and HB/SHB) when immersed in different pH
The directional transport phenomenon inspires researchers aqueous environments and the pH adjustable channels and
to prepare Janus membranes for oil–water separation.[135] For could be used for oil–water separation.[140] It is worth noting
a Janus membrane with one hydrophilic layer and one hydro- that pressure-responsive wettability could be also realized by
phobic layer, water droplets could slowly penetrate through the Janus membrane due to its unique asymmetric structure. Jin’s
membrane from the hydrophobic side and rapidly spread to group[141] reported a pressure-responsive Janus membrane for
the hydrophilic side. In that case, water might be pulled to the ultrafast separation of oil–water emulsions by depositing an
hydrophilic layer by gravity when the hydrophobic layer is thin ultrathin layer (HL-UWOB) of polydopamine-coated single-
enough. Hence, the thickness of the two layers with oppo- walled CNTs (SWCNTs) on to a porous cellulose ester substrate
site wettability and the driving force play vital roles in liquid and subsequently coating with another ultrathin SWCNT
permeation. Most recently, Wang et al.[136] synthesized a Janus layer (HB-SOL) via the vacuum suction method (Figure 16).
membrane with both HL polyamine and SHB PDMS surfaces The applied pressure is crucial in the oil–water separa-
to separate oil from an o/w emulsion by exposing the HL side tion processes: when the applied pressure is higher than oil
to the emulsion. For the separation process, water could wet intrusion pressure but lower than water intrusion pressure,
the HL surface but not pass through the entire membrane due the membrane shows extremely high separation efficiency
to the SHB barrier. Meanwhile, the polyamine layer demulsi- (>99.95%) for w/o emulsion; similarly, when the applied pres-
fies the emulsion and assist the re-coalescence of oil droplets. sure is higher than water intrusion pressure but lower than oil
Followed by that, the oil fills into the SHB pores, which further intrusion pressure, the membrane also exhibits outstanding
prevents the water permeation and pushes the oil transport. separation performance (i.e., separation efficiency > 99.99%)
The work achieves oil demulsification and oil–water sepa- for o/w emulsion.
ration by the Janus membrane based on the unidirectional Preparation and application of Janus membranes is still an
transport of oil and water. Similarly, Tian et al.[137] fabricated emerging research area. Practical oil–water separation applica-
a thin 1H,1H,2H,2H-perfluorooctyltrichlorosilane (POTS)- tions may be beneficial from the unique (i.e., directional and
induced HB layer onto a HL cotton membrane (Figure 15). switchable) wettability offered by Janus membranes. However,
By exposing the HB side to water in air (positive direction), the like other membranes with special wettability, constructing
water droplets permeates spontaneously from the HB to the HL heterogeneous structures and surfaces remains challenging.
side. Whereas if the membrane is overturned (reverse direction), Unknown influencing factors such as stimuli-responsive
water droplets are blocked and spread on the HL side. Besides, effects on membrane configuration are required further
when the membrane is placed UO in a positive direction, water exploration.

Adv. Mater. Interfaces 2018, 1800576 1800576  (13 of 27) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

Figure 16.  a) SEM image of the as-prepared membrane. Inset: digital photo of the free standing bilayer membrane. b) The schematic structure of the
ultrathin bilayer membrane and a proposed pressure-responsive mechanism for on-demand oil–water separation. c) Water and oil CAs of the prepared
membrane. Reproduced with permission.[141] Copyright 2015, Royal Society of Chemistry.

4. Membranes for Oil–Water Separation zeolites, metals, etc.[146–149] Inorganic membranes can be classi-
fied into two major categories based on their structures: porous
Although a wide range of membrane processes with different inorganic membranes and dense (nonporous) inorganic mem-
wettability have been attempted in oil–water separation branes. Dense metallic membranes (e.g., Palladium mem-
field, membrane filtrations are the most predominant. In brane) have been dominantly considered for the separation of
comparison with Section 3 that focus on the recent advances hot gases for ages[150] while porous thin metallic membranes
of novel special wettability in membrane materials, this sec- (e.g., SSM, CM, and silver membrane) with pore sizes ranging
tion comprehensively summarizes the synthetic methods from micro- to nanometer level have also been widely used for
and applications of filtration membranes beneficial from water filtration. Recently, numerous bioinspired membranes
the special wettability for oil–water separation. Membrane with special wettability have been fabricated on porous metallic
filtration processes follows complicated principles, e.g., meshes, which can be used for separating all categories of oil–
adsorption, size-sieving, and electrostatic phenomenon.[3] water mixtures. It should be noted that the metallic mesh sub-
Conventional filtration membranes applied for separating strates always have larger pore sizes than the oil droplet sizes,
oily water are mainly followed by the “size-sieving” effect but the modified substrates with rough micro- and nanostruc-
driven by pressure, in which oil droplets with certain sizes tures show exceptional separation efficiency and high flux.
are not allowed to pass through the “pores” of the mem- The inevitable tradeoff relation (between selectivity and per-
brane. With the deeper understanding of surface chemistry, meability) in membrane material world seems to be broken.
functionalization of filtration membranes with special wet- In addition, those metallic mesh membranes have robust sur-
tability could not only endow the membranes with superior faces, harsh-environment resistance and unique repeatability,
antifouling property but also lead to favourable synergetic which make them suitable for practical oil–water separation
effects for improving the selectivity and separation efficiency application.[151–154] Metallic mesh membranes are fabricated by
in practical applications. various surface modification approaches including but not lim-
It is well-known that filtration membranes can be catego- ited to the layer-by-layer self-assembly (LbLSA) method,[155,156]
rized by pore sizes or material types. For a conventional MF hydrothermal method,[157,158] and other coating methods.[159–161]
process based on membranes with pore size ranging from Table 1 displays some typical metallic mesh-based membranes
50–1000 nm, a specific operation pressure (also known as for oil–water separation in terms of substrates, preparation
transmembrane pressure) is required. However, there are methods, oil–water mixing systems, types of wettabilities and
many oil–water separation membranes with special wettability separation efficiencies.[72,73,151,152,157,161–170]
breaking out of their dependence on transmembrane pres- Typically, Li et al.[161] prepared a series of coating layers with
sure. Thus, it is not appropriate to classify specially wet- hierarchical nanostructure and dual roughness onto SSMs
table oil–water filtration membranes according to their pore via LbLSA of SiO2 NPs. Followed by surface HB modification
sizes and this section is divided into four parts on the basis using chemical vapor deposition (CVD), their membranes
of membrane materials: metal-, polymer-, ceramics-, and could separate surfactant-free o/w emulsions through a non-
nanomaterial-based. Notably, functionalized membranes are sieving mechanism achieving an extremely high separation
generally fabricated with complicated fabrication procedures efficiency (99.4%). They found that increasing the deposition of
and the use of harmful chemicals. Green process or using SiO2 NPs can rise the surface roughness of the SSM resulting
environmentally benign materials at low-cost should be also in superhydrophobicity (WCA ≈ 158°), which was attributed to
concerned for practical applications.[142–145] the existence of air trap in the hierarchically structured surface.
After repeating o/w emulsion filtration for 20 cycles, the sepa-
ration efficiency remained as high as 99.2%, which suggested
4.1. Metal-Based Membranes the coating was highly stable and durable.
Additionally, CM is another good substrate for oil–water
Inorganic membranes refer to membranes made of materials separation because it is easy to be processed by electrochem-
such as oxides, nonoxides, carbon, metal organic frameworks, ical modification. Wang et al.[168] developed hierarchical CM

Adv. Mater. Interfaces 2018, 1800576 1800576  (14 of 27) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

Table 1.  Metallic mesh-based membranes for oil–water separation in terms of different substrates, modification methods, separation application,
CAs, types of wettability, and features.

Metallic membranes Modification methods WCA [°] OCA [°] Wettability Application Features Ref.
SiO2 modified SSM LbL, CVD 158 n-hexadecane, 115 SHB-OL o/w HSE [161]
MWCNTs deposited SSM CVD 150 Gasoline, 0 SHB-SOL w/o HP [164]
Glass nanofiber-coated SSM Hydrothermal 0 UW oils > 150 SHL-UWSOB o/w GD [157]
MFI zeolite-coated SSM Hydrothermal 0 UW oils > 151 SHL-UWSOB Layered o-w mixture GD [162]
PSBMA/PDA-modified SSM Dip coating 0 UW oils > 150 SHL-UWSOB w/n-hexane GD [151]
NDM/PDA-modified SSM Dip coating 144 0 HB-SOL w/diesel oil GD [163]
GO coated SSM Dip coating/O2 plasma 0 UW oils > 150 HL-UWSOB w/bean oil GD [73]
BiVO4 coated SSM Dip coating/calcination 0 UW oils > 150 SHL-UWSOB w/bean oil SC [152]
PAM hydrogel-coated SSM In situ polymerization 0 UW oils > 150 SHL-UWSOB Layered o-w mixture GD [72]
CNFs-PDMS modified SSM Vacuum suction 163 Toluene, 0 SHB-SOL w/o HP [166]
Fatty acid modified CM Electrodeposition >158 Diesel oil, ≈0 SHB-SOL w/n-hexadecane GD [168]
Cu NPs modified CM Electrodeposition 154.1 chloroform, 0 SHB-SOL w/chloroform GD [169]
Hexadecylthiol modified CM Dip coating 153 ± 1 Diesel oil, 0 SHB-SOL w/diesel oil GD [170]
AgNO3 modified CM Chemical modification 0 UW oils > 159 SHL-UWSOB w/n-hexane GD [167]
ZnO–Co3O4 overlapped CM Hydrothermal UO 157.9 UW, DCE, 159.2 SHL-SOL o/w and w/o GD [165]

Note: HSE, high separation efficiency; HP, high permeability; GD, gravity-driven; SC, self-cleaning; AF, antifouling; DCE, 1, 2-dichloroethane; DCM, dichloromethane; PW,
produced water; o/w, o/w emulsion; w/o, w/o emulsion.

membranes by electrochemically depositing copper micro- membrane could separate both surfactant-free and surfactant-
clusters on the mesh surface. After immersing the textured stabilized oil–water emulsions (including o/w and w/o) under
CM membranes in n-dodecanoic acid for 12 h, the mem- gravity with a high flux and separation efficiency (Figure 17c).
brane surface had a WCA of 158 ± 2° with an SA of 2° and In conjunction with excellent antifouling property, superior
a diesel OCA ≈0°. The developed SHB-SOL membrane was thermal stability as well as long-term use, this inorganic over-
found to be an effective diesel oil and water separator. Guo’s lapped membrane showed great potential in industrial oily
group[169] adopted the similar electrochemical approach and wastewater treatment.
deposited a 2 µm-thick copper NPs layer on the surface of a
CM. After n-octadecyl thiol grafting, the membrane surface
exhibits a WCA of 154.1° and an OCA of chloroform of 0°. 4.2. Polymer-Based Membranes
A mixture of chloroform and water could be efficiently sep-
arated by the prepared mesh membrane solely driven by Owing to the remarkable advantages including low cost, promi-
gravity. Besides of electrochemical deposition, Wang et al.[170] nent flexibility, excellent processability and ease of operation,
utilized a facile solution-immersion process to prepare SHB- polymeric MF and UF membranes made of polysulfone
SHL (WAC = 153 ± 1° with SA < 5°, diesel OCA = 0°) CM (PSf),[171–174] polyethersulfone (PES),[175–178] PVDF,[179–181]
surface. The commercial CM was first etched by nitric acid etc., have been widely used in oily water treatment.[182–184]
and then chemically modified with 1-hexadecanethiol (HDT). These membranes, however, are easy to trigger a cascade of
The membrane could not only be applied for diesel ultrafast events such as accumulation, adhesion, spreading, coales-
oil–water separation through a simple filtration process, but cence and migration of suspended oil foulants as well as inevi-
also showed excellent stability in corrosive acidic, basic, and tably getting clogged during treating oily wastewater, which
salt solutions. results in flux decline and rejection deterioration. Numerous
Hydrothermal synthesis also provides a facile tool to modify approaches have, thus, been developed improve hydrophi-
metallic mesh membranes for oil–water emulsion separa- licity and antifouling property of the polymeric membrane by
tion because it can easily grow large good-quality crystals either blending with beneficial additives or altering its surface
on membrane surface. For example, Liu et al.[165] prepared a properties via chemical or physical modification.[24] Moreover,
zinc oxide–cobalt oxide (ZnO–Co3O4) overlapped membrane polymer chemistry also provides more advanced methods such
based on a hydrothermal method. The mesh pores was first as atom transfer radical polymerization,[185] in situ polymeri-
filled with pine needle-like ZnO, which aggregated together to zation,[186] and interfacial polymerization,[187] which have also
form a cluster structures. After hydrothermal growth of Co3O4, been explored in order to prohibit membranes from fouling.
the flower-like Co3O4 was uniformly distributed and coated A summary of polymeric membranes for oily water treat-
on the ZnO-stuffed mesh surface. Figure 17a,b schematically ment is given in Table 2 and it is apparent that the special
shows that the overlapped membrane exhibits superoleopho- wettability plays a crucial role in enhancing their separation
bicity and UO superhydrophobicity at the same time. The performance.[172–174,179,181,188–203]

Adv. Mater. Interfaces 2018, 1800576 1800576  (15 of 27) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

Figure 17.  a) Schematic structure of the ZnO-Co3O4 overlapped membrane and its switchable wettability when immersed in different media, and the
corresponding separation capacities of oil–water emulsions. b) Wetting behavior of the membrane toward water in air and oil in water. c) Separation
performance to various emulsions: i) Tween 20-free diesel-in-water emulsion, ii) tween 20-stabilized diesel-in-water emulsion, iii) Span 80-free water-
in-diesel emulsion, and iv) Span 80-stabilized water-in-diesel emulsion. Reproduced with permission.[165] Copyright 2015, Nature Research.

Blending a polymer matrix with beneficial additives like the phase inversion process, dragged by the HL segments,
hydrophilic polymers, amphiphilic copolymers and inorganic the nonpolar HB segment tends to migrate to the membrane
nanoparticles during the phase inversion process is an effective surface to form microdomains. These microdomains are
and easy-to-use method to improve the selectivity and fouling advantageous to antifouling performance because of their
resistance of the porous polymeric filtration membranes.[9] An low surface free energy. Thus, the flux decline of the mem-
SHB-SOL PVDF membrane was fabricated by Zhang et al.[188] brane could be greatly reduced during the separation of the
by using an inert solvent-induced phase-inversion process o/w emulsion.
(Figure 18). The prepared PVDF membranes were, for the first In addition to polymeric additives, inorganic nanoparticles,
time, able to separate both micrometer and nanometer sized such as Al2O3, TiO2, ZrO2, and SiO2, have also been used to
surfactant-free and surfactant stabilized w/o emulsions solely blend with host polymers.[24] It is well known that nanopar-
driven by gravity, with a high flux and separation efficiency. ticles have large surface areas and abundant surface-active
Amphiphilic copolymers that are composed of both HL and groups. The addition of inorganic nanoparticles has been dem-
HB segments are also used as additives to blend with host onstrated to be helpful to membrane permeability and fouling
polymers.[204–206] Mayes and co-workers[207] first used a comb- resistance by either changing the pore structure or increasing
like copolymer P(MMA-r-POEM) (i.e., PMMA back bone and the hydrophilicity of the membranes. A simple and commonly
poly(ethylene oxide) side chain) as an additive. The amphi- used approach is to directly blend pre-prepared nanoparticles
philic polymer was blended into PVDF to form membranes with polymers in solution. Li and co-workers[212] blended Al2O3
via surface segregation. Their results demonstrated that the nanoparticles with PVDF UF membranes. It was observed that
antifouling property of the membrane was greatly improved, the addition of a certain amount of Al2O3 nanoparticles could
whereas the membrane structure could be maintained. Since slightly increase the hydrophilicity of the PVDF membrane
then, many amphiphilic copolymers, such as tri-block,[208] but had no effect on the effective pore size and porosity of the
comb-like,[209] and branched copolymers,[210] have been used membrane. Similarly, Yang and co-workers[213] used TiO2 NPs
as additives to blend into host polymers to produce composite to blend into a PSf UF membrane. The membrane was suc-
membranes with enhanced antifouling properties. Among cessfully used for separating kerosene-emulsified wastewater
these copolymers, ternary copolymers containing an HB with improved antifouling property. Despite these successes
“anchoring” segment, HL fouling resistance segment or non- of blending HL NPs into polymer matrixes, NPs are generally
polar HB segment with low surface free energy have received easily aggregated in polymer membranes due to their
much attention because of their structural merits.[211] During intrinsic poor dispensability. Therefore, the approach limits

Adv. Mater. Interfaces 2018, 1800576 1800576  (16 of 27) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

Table 2.  Polymer-based membranes for oil–water separation in terms of different substrates, modification methods, separation application, CAs,
types of wettability, and features.

Polymer-based membranes Preparation Methods WCA [°] OCA [°] Wettability Application Feature Ref.
Inert solvent induced-PVDF Phase inversion 158 <1 SHB-SOL w/o HSE [188]
ZrO2 blended PVDF Phase inversion 52 n.a. HL o/w AF [179]
PANI/TiO2 nanofiber blended PVDF Phase inversion 61 n.a. HL o/w AF [190]
Amphiphilic block-copolymer blended PVDF Phase inversion 96.1 ± 3.2 DCM, 56.9 HL-OB o/w AF [193]
TiO2/KH550 modified PVDF Phase inversion 0 UW diesel oil, 158.6 SHL-UWSOB o/w AF [199]
PVDF-g-PTA/TiO2 hybrid Phase inversion 0 UW hexadecane, 159.3 SHL-UWSOB o/w AF [195]
PDA coated PVDF Phase inversion 53 ± 2.3 UW chloroform, 152 SHL-UWSOB w/o AF [198]
TiO2 modified PLA Phase inversion 0 UW, DCM, 154 ± 1 SHL-UWSOB w/lubricating oil AF [194]
PAA-g-PVDF Phase inversion 0 UW, DCE, 160 SHL-UWSOB w/o GD [179]
Copolymer blended PVDF hollow fiber Phase inversion 26 n.a. HL-OB Real wastewater AF [181]
Amphiphilic zwitterionic copolymer blended PES Phase inversion 29.4 hexadecane, 28.5 HL o/w AF [196]
HMO blended PES Phase inversion 16.4 n.a. HL o/w AF [192]
PEG blended PSF Phase inversion n.a. n.a. HL o/w AF [202]
Bentonite NPs blended PSF Phase inversion 62 n.a. HL Synthetic PW AF [173]
HAO blended PSF Phase inversion 10 n.a. HL w/crude oil AF [172]
PVDF–HFP nanofibrous Electrospun 0 UW, DCM, 169 ± 3 SHL-UWSOB o/w AF [201]
N-perfluorooctyl-substituted Polyurethane Electrospun >150 >150 Switchable o/w AF [197]
P34HB blended polylactide Electrospun 0 after 10s n.a. HL o/w AF [216]
PEGDA@PG nanofibrous Electrospun 0 n-hexane, 108 SHL o/w GD [212]
CA–PI nanofibrous Electrospun 48 0 SOL w/DCM GD [202]
F-SNF/Al2O3 In situ polymerization 161 Rapeseed, 0 SHB-SOL w/o GD [186]
PSF nanofibrous Interfacial polymerization 3 n.a. SHL water/soybean oil GD [174]

Note: HSE, high separation efficiency; HP, high permeability; GD, gravity-driven; SC, self-cleaning; AF, antifouling; DCE, 1, 2-dichloroethane; DCM, dichloromethane; PW,
produced water; o/w, o/w emulsion; w/o, w/o emulsion.

Figure 18.  a) Schematic illustration of the formation of a superhydrophobic–superoleophilic PVDF membrane via a modified phase-inversion process.
b) Digital photo of an as-prepared PVDF membrane with a dyed water droplet (3 µL) dropped on it. The inset is a cross-sectional SEM image of the
membrane. c) Photographs of a water droplet on the membrane showing a CA of 158° (up) and an oil droplet on the membrane showing a nearly
0 CA (down). d) Separation performance of the PVDF membrane toward w/o emulsion. Reproduced with permission.[188] Copyright 2013, Wiley.

Adv. Mater. Interfaces 2018, 1800576 1800576  (17 of 27) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

the improvement in the membrane permeability and anti- HL polymers such as poly(ethylene glycol) methyl ether meth-
fouling performance and has the risk to cause defects in the acrylate (PEGMA),[217] poly(2-hydroxy-ethyl methacrylate),[218]
membranes. zwitterionic polyelectrolyte,[219] or small molecules[220] onto
In comparison with porous membranes prepared by the membrane surfaces via the formation of covalent bonds. The
phase inversion method, electrospun polymeric membranes also introduced HL materials form compact hydrated layers to
have outstanding characteristics, such as extremely high flux prevent the fouling of oil droplets on membrane surfaces and
under low operation pressure as well as relatively controllable allow the fouled oils to be easily removed during the following
pore size and porosity.[214] Traditional electrospun membranes cleaning process. Jiang and co-workers[220] conducted an inten-
with superhydrophobicity were used to separate oils from the sive study regarding surface grafting on filtration membranes
emulsions. For example, a electrospun fibrous membrane from and they recently grafted a low surface free energy molecule
the polymerization product of styrene and butyl acrylate was (pentadecafluorooctanoic acid) onto the surface of an aminated
reported.[215] Because of its obvious affinity for oils as well as its polyacrylonitrile UF membrane via an acrylated reaction. The
loose and porous structure, the oils could be completely absorbed modified membrane exhibited excellent antifouling property.
by membrane upon encountering the membrane surface, gen- In addition to chemical reactions, directly coating HL poly-
erating an oil film with good water repellency on the surface. mers and zwitterionic polymer onto membrane surfaces via
Zhang et al.[216] developed a polymer membrane composed of poly­­ physical absorption is also feasible. For example, Chu and
lactide (PLA) and poly (3-hydroxybutyrate-co-4-hydroxybutyrate) co-workers[221] reported the surface modification of the HL layer
(P34HB) using blend electrospinning. Taking advantages on nanofibrous UF membranes to achieve high flux for o/w
of the structural porosity of the electrospun membrane emulsion separation. A nanofibrous polyacrylonitrile weave
and the hydrophilicity of the P34HB, the electrospun mem- as a scaffold was first placed onto a nonwoven poly (ethylene
brane exhibited superior water permeability, resulting in highly terephathalate) microfilter substrate, and then, chitosan, as
efficient removal of water from a surfactant-free vacuum pump the hydrophilic layer, was spin-coated onto the polyacrylonitrile
oil–water emulsion under gravity (Figure 19). Recently Fang weave to form a three-tier nanofibrous membrane. Because of
et al.[197] first proposed a self-healing porous membrane concept the high porosity and thin and smooth barrier layer of the
by electrospinning fluorine-containing polyurethane (PU). The membrane, the membrane exhibited outstanding performance
obtained membrane can separate surfactant-stabilized water- including high flux and antifouling property. During the long
in-ether emulsion solely driven by gravity with high separation cross-flow filtration experiment using an o/w emulsion as the
efficiency and high flux. In this case, the hierarchical micro/ feed, the water flux maintained a much higher value than the
nanoscale structure of the membrane incorporated with low commercial prisitine membrane, and no obvious decline of oil
surface energy fluoroalkyl chains provided the membrane with rejection was found. Kim et al.[222] reported a PSf UF membrane
self-cleaning and self-healing properties. The fluorine-containing coated with a star-shaped block copolymer containing PEGMA
PU as the healing components can migrate to the outer surface and methyl methacrylate (MMA) moieties using a spinning
of the fiber to heal the superhydrophobicity of the membrane. method. The as-prepared membrane can separate surfactant-
In addition to this healing mechanism, the work give a founda- stabilized o/w emulsion. The amphiphilic membrane surfaces,
tion of self-healing for regeneration and reuse of w/o emulsion consisting of both HL poly (ethylene glycol) (PEG) and HB moi-
separation membranes, which matches well the requirements eties exhibited higher oil-fouling resistance than that of the sur-
for treating oily wastewater produced in industry and daily life. face with only HL PEG moieties. Interestingly, in this work, it
Surface wettability modification by either chemical or phys- is found that by controlling the surface morphology of polymer
ical methods is another efficient approach to enhance oil–water coatings with no oil-repellent moieties, the oil-fouling resist-
separation performance of polymeric membranes. Surface ance can also be increased. The favorable interactions between
modification through chemical reactions could firmly introduce the PMMA core blocks and the PSf membrane surface, caused
the larger content of denser PEG moieties
on the surfaces, which could weaken the
interactions between the oil foulants and the
membrane surface.
In situ polymerization, is a process in
which reactive monomers are filled in the
layers located in the lamellar and the relevant
polymerization is accordingly conducted. For
example, a novel flexible and hierarchically
porous silica nanofibrous (SNF) membrane
with superhydrophilicity and underwater
superoleophobicity was prepared by a facile
in situ polymerization method coupling with
electrospun,[186] which can effectively separate
microsize surfactant-stabilized o/w emul-
Figure 19.  a) SEM images of the electrospun membrane and the inset is the optical images of
the time-dependent water droplets on the membrane surface. b) Optical images of the separa- sions solely by gravity, with an extremely
tion performance of the as-prepared electrospun membrane under gravity. Reproduced with high flux of 2237 L m−2 h−1 (Figure 20). The
permission.[216] Copyright 2015, Elsevier. enhanced fouling resistance of the SNF

Adv. Mater. Interfaces 2018, 1800576 1800576  (18 of 27) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

Figure 20.  SEM images of a) SNF, b) fluorinated SNF (F-SNF), and c) F-SNF/Al2O3 membranes, respectively; the insets show the optical profiles of
water droplets on the relevant membrane surfaces. Droplets of rapeseed oil (dyed red) and water (dyed blue) on the a’) SNF, b’) F-SNF, and c’) F-SNF/
Al2O3 membranes, respectively. d) Schematic illustration of the CA hysteresis and SA of a liquid droplet. e) A water droplet (10 µL) sliding at a low
angle of 3.8° on the as-prepared F-SNF/Al2O3 membrane. f) The gravity-driven separation for oil–water emulsions using the F-SNF/Al2O3 membrane.
Photos and optical microscopic photos of the oil–water emulsion before and after separation. Reproduced with permission.[186] Copyright 2013, Royal
Society of Chemistry.

membrane arose from its outstanding surface wettability, which structure with certain thickness, the coating procedure must
resulted in water layer formation on the membrane surface and be well-controlled. Followed by coating and drying, separation
restrained from direct contact between the oil and membrane layer made of ceramic microcrystals is obtained by further
surface during separation. What is more, the SNF membrane calcination. These processes finally leave a fixed roughness on
also exhibits robust mechanical strength, high thermal stability, ceramic membrane surfaces. As discussed previously, surface
and easy cycling for long-term use. Therefore, as an efficient roughness is a key factor associated with membrane wettability
and diverse method, in situ polymerization could sturdily bind that affects membrane performance in oil–water separation.
some functional NPs (such as SiO2 NPs) on the polymeric sub- Over the past few years many works (Table 3) have been sig-
strates and create a nanoscale roughness with special wettability nificantly focusing on the development of high performance
and enhanced mechanical strength, which make those mem- ceramic MF and UF membranes for oily wastewater treat-
branes promising candidates for oil–water emulsion separation. ment.[223,227–235] Although very few reports[232,236] stated that
their enhancement in ceramic membranes for in oil–water
separation application is attributed to membrane surface wet-
4.3. Ceramic-Based Membranes tability, these work coincidently concentrated on the control of
roughness and HL modification of membrane surface. This is
In the family of membranes, porous ceramic membranes are consistent with our viewpoint that wettability plays a critical
an important category because of their high chemical, thermal role for ceramic membranes in the applications of oily water
and mechanical stability, which make them suitable for use in treatment.
harsh conditions such as corrosive and high temperature envi- It is well-known that ceramic membranes mainly consisting
ronments. These membranes are made from alumina, titania, of metal oxides always show hydrophilicity, thus nearly all
zirconia or silica oxides, silicon carbide, etc.[223] Typical ceramic membrane scientists have paid their attention on the design of
membranes are made in layers with different porous structure: HL-UW OB ceramic membranes for oily water treatment. For
a support (also known as substrate) layer and a thin separa- example, without introduction of any chemical modification,
tion (also known as top) layer. It is worth noted that sometimes Xing and co-workers[227] first investigated the effect of surface
an intermediate layer might exist in a ceramic membrane. roughness on HL ceramic membranes for oily water treatment.
Ceramic supports usually consist of by above-mentioned They prepared a series of ceramic membranes with controllable
metal oxide or inorganic powders, to which binders and plas- surface roughness using PMMA particles as the template with
ticizers may be added. The resulting mixture is then pressed, different diameters. It was found that surface roughness had
extruded or slip cast prior to a sintering process. To obtain a little effect on the pure water flux of membrane and oil rejec-
ceramic membrane, the ceramic support surface has to be tion, but it played an important role in the filtration flux of oily
smoothed first. Then the flat ceramic support is usually dip- wastewater. In addition, Chang and co-workers systematically
coated with a casting solution via capillary force and then dried studied the application of Al2O3 MF membranes for oily water
for a certain time.[224–226] In order to get a desired membrane treatment by modifying the hydrophilic HL nanosized ZrO2,[235]

Adv. Mater. Interfaces 2018, 1800576 1800576  (19 of 27) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

Table 3.  Ceramic-based membranes for oil–water separation in terms of different substrates, modification methods, separation application, CAs,
improvements of wettability, and features.

Ceramic membranes Modification WCA(°) OCA(°) Improvements Application Features Ref.


Al2O3 MF Template n.a. n.a. Surface roughness o/w HP [223]
Mullite-titania Sol–gel 11 n.a. HL modification o/w HSE [227]
TiO2 modified MF Hydrothermal 14.57 ± 0.5 n.a. HL modification o/w AF [231]
γ-Al2O3 modified MF Hydrothermal 19.4 ± 1.1 n.a. HL modification o/w AF [231]
NaA zeolite on ceramic Hydrothermal n.a. n.a. HL modification o/w AF [230]
Zeolite MCM-22/α-Al2O3 Hydrothermal n.a. n.a. HB modification o/w AF [232]
Nano-TiO2 on MF Sol–gel 8 n.a. HL modification o/w AF [228]
Nanosized ZrO2 on MF Sol–gel 20 n.a. HL modification o/w AF [235]
GO modified Al2O3 MF Vacuum suction n.a. n.a. HL modification o/w HP [233]
Silica NPs modified Sol–gel 3 UW, >150 SHL-UWSOB o/w AF [234]
PU-PDMS coated MF Sol–gel 161.2 kerosene, 0 SHB-SOL w/kerosene AF [229]

Note: HSE, high separation efficiency; HP, high permeability; GD, gravity-driven; SC, self-cleaning; AF, antifouling; DCE, 1,2-dichloroethane; DCM, dichloromethane; PW,
produced water; o/w, o/w emulsion; w/o, w/o emulsion.

nano-TiO2,[228] and graphene oxide nanosheets,[233] respectively. Completely different from conventional pore size sieving
Notably, all modified membranes have a better performance on mechanism (the average pore size is determined as 3.62 µm),
antifouling than that of the unmodified membrane because the they claimed that their membrane mainly depend on the
HL coatings prevents the oil droplets from adhering or pene- special wettability of the separation layer, where water spread
trating. These modified membranes also exhibited an enhanced on the membrane surface initiatively and form a layer of water
and stable flux for oil–water emulsion separation compared film to repel oil droplets due to the natural incompatibility
with the pristine Al2O3 MF membrane. between water and oil.
Very recently, Chen et al.[234] fabricated a SHL-UWSOB In contrast to SHL-UWSOB ceramic membranes, Su et al.[229]
hierarchically structured ceramic membrane for separating reported a porous ceramic membrane with a SHB-SOL surface
oily wastewater (Figure 21). A SHL-UWSOB silica sol was pre- which could be used for reclaiming oil from oily water (Figure 22).
pared via the well-known Stöber process and the hierarchically A commercial porous ceramic tube was first coated with a tetra-
structured ceramic membrane surface is constructed through ethoxysilane-derived silica sol and then the membrane surface
a dip-coating process. Their membrane can separate the was modified with a PU-PDMS film. The resulting ceramic
oily wastewater with a mean oil droplet diameter of 1.54 µm membrane shows a WCA of 161.2° while a kerosene OCA of 0°,
and the oil separation efficiency is extremely high (99.95%). which could be used to separate a mixture of water–kerosene–clay.

Figure 21.  SEM images of the silica NPs-modified membrane prepared at 60 °C. The inset is the shape of adding a water droplet on the membrane
surface. b) Spreading and adhesion behaviors of water and oil droplets. c) Separation performance of the membrane to oil field o-wo wastewater.
Photograph and optical microscopic images of the wastewater before and after filtration. Reproduced with permission.[234] Copyright 2016, Elsevier.

Adv. Mater. Interfaces 2018, 1800576 1800576  (20 of 27) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

Figure 22.  a) SEM images of the pore wall of the porous ceramic tube with SHB- SOL surface. The inset is the SEM image with magnification of
20 000 times. b) Optical images of water and kerosene droplets on the inner membrane surface (WCA is 161.2° and kerosene CA is 0°). c) Optical
images of the oily water containing clay particles and the filtrate collected from the separator. Reproduced with permission.[229] Copyright 2012, Elsevier.

It is found that the oil recovery efficiency, to a great extent, surface properties, 1D or 2D nanomaterials have made impres-
depends on the component of target mixtures and separa- sive progress in fabricating ultrathin filtration membranes with
tion velocity. Clay particles can clog the pore and adhere on the innovative structures which could exhibit low weight, high
surface to decrease the effective surface for separation, and there permeability and high rejection rate compared with traditional
are new common channels for kerosene and water between clay filtration membrane materials in oil–water separation applica-
particles Although the best oil recovery efficiency of their mem- tions (Table 4).[141,237–245]
brane is only 67.6% when the feed consists of 3840 g water and In the recent years, 1D nanomaterials, especially nanotube,
160 g kerosene, their work opens up the new world of making nanowires and nanofibers have attracted intensive attention
ceramic filtration membrane SHB-SOL for oil recovery from oil– for constructing free-standing membranes with special wetta-
water mixtures. bility for oily water treatment. It is of great interest to assemble
1D nanomaterials into fibrous macroscopic architectures
with separation performance. Among the existing 1D nano-
4.4. Nanomaterial-Based Free-Standing Membranes materials, carbonaceous nanomaterials including single- or
multiwalled carbon nanotube (SWCNT or MWCNT) are the
With tremendous interest in nanotechnology, various types of most attractive candidates for constructing free-standing mem-
metallic and several carbonaceous nanomaterials have drawn branes for oil–water separation. Jin and her co-workers did a
significant attention in recent years for water treatment.[31] In lot of pioneering work with CNTs,[238,246–249] for example, they
the field of oily water treatment membrane, nanomaterials prepared an ultrathin free-standing SWCNT membrane with
have been used not only for modifying membrane surfaces tunable thickness for the ultrafast separation of emulsified
but also for constructing nanomaterial-based free-standing oil–water mixtures by using a vacuum suction method. The
thin membranes. An ideal membrane is expected to have an SWCNT membrane are HB-SOL that could effectively separate
active separation layer, which should be as thin as possible both micrometer- and nanometer-sized surfactant-free and sur-
for pursuing low transport resistance and high flux. There- factant-stabilized w/o emulsions with both efficiencies higher
fore, taking advantages of their high surface area and special than 99%. Furthermore, after 20 recycling cycles, the separation

Table 4.  Nanomaterial-based free-standing membranes for oil–water separation in terms of different substrates, modification methods, separation
application, CAs, types of wettability, and features.

Free-standing membranes Fabrication WCA [°] OCA [°] Wettability Application Feature Ref.
SWCNT Vacuum suction 94 DCM, 0 OB-SOL o/w GD [237]
PDA-coated SWCNT Vacuum suction 104 0 HB-SOL o/w and w/o HSE [238]
PEI@CNT-TMC Vacuum suction 51 ± 0.7 UW oils > 152 HL-UWSOB o/w AF [241]
Cryptomelane nanowire Hydrothermal, CVD 172;0 at 390 °C n.a. SHL-SHB w/toluene Switchable [242]
SiO2/carbon composite Electrospun 144.2 ± 1.2 DCM, 0 HB-SOL w/hexane GD [243]
MnO2 nanowire Vacuum suction 32–143 n.a. HL-HB n.a Switchable [244]
PDA coated rGO Vacuum suction 0 UW oils > 150 SHL-UWSOB o/w AF [239]
Graphene–TiO2 composite Vacuum suction 0 UW oils > 151 SHL-UWSOB o/w SC [240]
GO-Palygorskite Vacuum suction 44 >150 HL-UWSOB o/w HP [250]

Note: HSE, high separation efficiency; HP, high permeability; GD, gravity-driven; SC, self-cleaning; AF, antifouling; DCE, 1,2-dichloroethane; DCM, dichloromethane; PW,
produced water; o/w, o/w emulsion; w/o, w/o emulsion.

Adv. Mater. Interfaces 2018, 1800576 1800576  (21 of 27) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

efficiency and flux remained same indicating that the mem- is promising building block for building laminate membrane
brane have excellent stability and antifouling properties for structure with tunable passageway for water molecules.[237] With
long-term w/o separations.[246] vacuum-assisted filtration self-assembly, Zhao et al.[250] utilized
Instead of carbon-dominated nanomaterials, Yuan et al.[242] palygorskite (PGS) nanorods to intercalate into adjacent GO
used a self-assembly method to construct thermally stable, nanosheets to form a laminate free-standing membrane (GOP)
free-standing cryptomelane nanowire membranes that exhibit through π−π stacking and cation cross-linking (Figure 24).
switchable superwetting behavior ranging from SHL to SHB. The palygorskite nanorods in the membranes could enlarge
The prepared membranes can selectively absorb oils including mass transfer channels, elevate hydration capacity, and create
cyclohexane, 1,2-dichlorobenzene, gasoline, motor oil, petro- hierarchical nanostructures of membrane surfaces. The hydra-
leum ether and toluene up to 20 times the membrane’s weight tion capacity and hierarchical nanostructures synergistically
in water through a combination of superhydrophobicity and endow the membranes with underwater superoleophobic and
capillary action. This result suggests the free-standing mem- low oil-adhesive water/membrane interfaces and the mem-
brane can removal oils from water and could be used for oil/ branes exhibit outstanding separation performance and anti-
water filtration separation and particularly for oil spill cleanup. fouling properties for various oil-in-water emulsion systems
Similarly, Lan et al.[244] reported a multifunctional free-standing (with different concentration, pH, or oil species). In contrast to
membrane from the self-assembly of ultralong MnO2 nano­ GO nanosheet, chemically reduced GO (rGO) has also widely
wires via the vacuum suction method. MnO2 nanowires with proposed for constructing free-standing filtration membranes.
50 nm in width and up to 100 µm in length were assembled Typically, Feng’s group[239] reported a facile route to fabricate
as a free-standing membrane with strong mechanical stability ultralight free-standing rGO membranes with superhydrophi-
and flexibility. Interestingly, a reversible wettability between licity-UW superoleophobicity. The membrane was obtained by
hydrophilicity and hydrophobicity of the membrane could be vacuum filtration a suspension of polydopamine-coated rGO
achieved by incorporating triethoxy(octyl)silane into the MnO2 nanosheets onto a commercial mixed cellulose ester filter. The
scaffold. As shown in Figure 23, it is found that when treated membrane is capable of separating multiple types of surfactant
by ethanol and heat-treatment at 60 °C for 5 min, the wetta- stabilized o/w emulsions with oil droplets of nano/sub-
bility of membrane could be switched between hydrophilicity micrometer size, as well as display high separation efficiency
(WCA of ≈32°) and hydrophobicity (WCA of 143°). The free- and excellent anti-fouling properties, endowing them high-
standing membranes also show a great potential for oil–water lighted alternatives for oily water treatment.
separation.
In addition to 1D nanomaterials, 2D nanomaterials especially
graphene-based nanomaterials have been extensively attempted 5. Conclusions and Outlook
to fabricate free-standing membranes with special wettability
for oil–water separation. Graphene oxide (GO) nanosheet is Membrane technology, as a widely utilized unit operation for
an oxidized derivative of graphene which can be regarded as a tackling the world-class challenge of treating oily water pro-
graphene scaffold with different oxygenous functional groups duced from various industrial processes and daily life, has
including epoxy, carboxyl, hydroxyl, and carbonyl. Due to the been considered as the most promising method due to its
hydrophilic nature, high surface area and abundant reaction high separation efficiency and relatively simple process. How-
sites, GO nanosheet allows a vast of chemical decorations that ever, traditional membrane materials suffer from variety of

Figure 23.  a) SEM image of the the membrane and the inset is the photographs of the free-standing membrane. b) Photographs of the water droplet
shape on the membrane before and after treatment with ethanol. c) Reversible HB-to-HL transformation of the as-prepared membrane under cycled
ethanol treatment and 60 °C heating. Reproduced with permission.[244] Copyright 2013, American Chemical Society.

Adv. Mater. Interfaces 2018, 1800576 1800576  (22 of 27) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

Figure 24.  a) Schematic illustration of fabricating of GOP membranes via vacuum-assisted filtration self-assembly. b) Apparent WCAs in air and
c) apparent OCAs UW of GOP membranes with different PGS/GO mass ratios. d) Schematic of the oil–water-membrane interfaces of GO and GOP
membranes. Reproduced with permission.[250] Copyright 2016, American Chemical Society.

problems, such as severe fouling, low separation efficiency and Despite enormous progress having been made of mem-
huge energy consumption. Significant advances in the field of branes with special wettability for oily water treatment, there
special wettability have opened new possibilities to improve are still some challenges and opportunities for both the scien-
membrane materials and their performances in separation pro- tific and industrial communities. For creating special wettability
cesses. This review offers a state-of-the-art assessment of the with membrane materials, some challenges are highlighted as
latest research work on selective membrane materials under the follows. 1) SHB-SOL membrane materials have been extensively
concept of special wettability for highly efficient oil–water sepa- developed, however gravity separation is not applicable for this
ration. Wetting on membrane surfaces has received increasing type of membrane if water contacts the membrane before oil,
attention among membranologists. Moreover, a comprehensive and fouling can still form when oils adsorb to the membrane
understanding of the emerging advanced membrane mate- surface. 2) SHL-UWSOB membrane materials can be utilized
rials with specific wettability is systematically presented in for gravity-driven separation of layered oil–water mixtures with
terms of membrane classifications, definitions, general design excellent fouling resistance but they are unsuitable for the
principles, characteristics and applications. The detailed dem- separation of emulsified oil–water mixtures. 3) SHL-SOB (in-
onstration of metallic-, polymeric-, ceramic- and nanomaterial- air) membranes are promising in various industrial processes
based filtration membranes in separating oil–water mixtures requiring oily water treatment and can be used for clean-up of
may offer practical insights into developing oil–water separa- oil spills, but the fabrication methods of the SHL-SOB mem-
tion membranes. This article also provides new insights into brane materials are complicated and still very challenging.
the construction of oil–water separation membrane materials 4) Superamphiphilic membranes exhibit excellent separation
with remarkably enhanced or newly derived functionalities/ performance for emulsified oil/water mixtures and are suitable
characteristics. for large-scale production process. However, the understanding

Adv. Mater. Interfaces 2018, 1800576 1800576  (23 of 27) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

of superamphiphilic mechanisms is still insufficient so far. Keywords


5) (Super) omniphobic ability provides enormous opportunities
membrane, oil–water separation, surface modification, wettability
for MD to extend its application for treating oily wastewaters
with low surface tension contaminants. Other membrane Received: April 16, 2018
contactors and direct membrane filtration processes may take Revised: July 5, 2018
advantages of omniphobicity as well. The fabrication techniques Published online:
for superomniphobic membranes remain challenges that
are still lack of investigation. 6) The research on membranes
with switchable wettability based on different mechanisms,
such as smart stimuli-responsibilities and Janus effects are still [1] W. Zhang, N. Liu, Y. Cao, X. Lin, Y. Liu, L. Feng, Adv. Mater. Inter-
in their early stages and further efforts are highly required. New face 2017, 1600029.
types of responsive or multi-responsive materials are prom- [2] D. J. Miller, D. R. Dreyer, C. W. Bielawski, D. R. Paul, B. D. Freeman,
Angew. Chem., Int. Ed. 2017, 56, 4662.
ising to construct membranes with tunable wettability. Janus
[3] M. Padaki, R. Surya Murali, M. S. Abdullah, N. Misdan,
membranes with opposing wettability may give a rise for the
A. Moslehyani, M. A. Kassim, N. Hilal, A. F. Ismail, Desalination
unidirectional or switchable separation of oil–water mixtures. 2015, 357, 197.
Although many membrane processes are feasible, the [4] Z. Xue, Y. Cao, N. Liu, L. Feng, L. Jiang, J. Mater. Chem. A 2014,
membrane filtration processes have played and will still 2, 2445.
play a predominant role in oil–water separation due to the [5] G. Kwon, A. K. Kota, Y. Li, A. Sohani, J. M. Mabry, A. Tuteja,
simplicity of the operation, benign experimental conditions Adv. Mater. 2012, 24, 3666.
and high separation efficiency. As to the possible futures [6] Y. Wang, X. Gong, Adv. Mater. Interface 2017, 4, 1700190.
for the filtration membranes with special wettability for [7] G. Kwon, E. Post, A. Tuteja, MRS Commun. 2015, 5, 475.
oil–water filtration, perspectives on technical advances and [8] H. Zhu, Z. Guo, J. Bionic Eng. 2016, 13, 1.
[9] T. A. Otitoju, A. L. Ahmad, B. S. Ooi, J. Water Process Eng. 2016, 14, 41.
market potential are highlighted as follows. 1) Chemically and
[10] M. Cheryan, N. Rajagopalan, J. Membr. Sci. 1998, 151, 13.
mechanically robust composite membrane materials with high
[11] C. Charcosset, I. Limayem, H. Fessi, J. Chem. Technol. Biotechnol.
oil fouling resistance should and will be developed with even 2004, 79, 209.
more facile methods at low costs. 2) Low pressure or gravity- [12] E. N. Tummons, V. V. Tarabara, J. W. Chew, A. G. Fane, Jo. Membr.
driven filtration processes integrated with smart membrane Sci. 2016, 500, 211.
surfaces and modules will lead to more advanced membrane [13] M. Hlavacek, J. Membr. Sci. 1995, 102, 1.
engineering, which may bring more solutions to treating oily [14] K. S. Ashaghi, M. Ebrahimi, P. Czermak, Open Environ. Sci. 2007, 1, 1.
water. 3) The on-going research of artificial intelligent respon- [15] S. Munirasu, M. A. Haija, F. Banat, Process Saf. Environ. Prot. 2016,
sive membrane, in essence, is about the membrane resistance 100, 183.
to convection and diffusion of solvent and solute when an [16] U. Daiminger, W. Nitsch, P. Plucinski, S. Hoffmann, J. Membr. Sci.
1995, 99, 197.
external stimuli changes physicochemical characteristics on
[17] B. Wang, W. Liang, Z. Guo, W. Liu, Chem. Soc. Rev. 2015, 44, 336.
membrane surface. Such membranes will be finally in wide
[18] Q. Ma, H. Cheng, A. G. Fane, R. Wang, H. Zhang, Small 2016, 12,
use based on full understanding of their mechanisms. 4) As 2186.
there is a growing tendency applying ceramic MF/UF mem- [19] K. Wang, W. Yiming, J. Saththasivam, Z. Liu, Nanoscale 2017, 9,
branes for oily water treatment, constructing a specially and 9018.
selectively wettable layer onto the membrane surface makes [20] M. G. Buonomenna, RSC Adv. 2013, 3, 5694.
ceramic membranes more attractive on a lifecycle cost basis. [21] A. Lee, J. W. Elam, S. B. Darling, Environ. Sci.: Water Res. Technol.
5) Nanotechnology plays a growing role in material science 2016, 2, 17.
and membrane fabrication techniques, and nanomate- [22] A. G. Fane, R. Wang, M. X. Hu, Angew. Chem., Int. Ed. 2015, 54, 3368.
rial-based membranes enrich filtration membrane family. [23] M. M. Pendergast, E. M. V. Hoek, Energy Environ. Sci. 2011, 4, 1946.
[24] J. Kim, B. Van der Bruggen, Environ. Pollut. 2010, 158, 2335.
Free-standing membranes constructed by low-dimensional
[25] L. Malaeb, G. M. Ayoub, Desalination 2011, 267, 1.
nanomaterials that break through the traditional membrane’s
[26] G. K. Anderson, C. B. Saw, M. S. Le, Environ. Technol. Lett. 1987,
tradeoff between permeability and selectivity may bring 8, 121.
continuous miracles in water treatment industry. [27] J. M. Estrada, R. Bhamidimarri, Fuel 2016, 182, 292.
[28] T. Mohammadi, M. Kazemimoghadam, M. Saadabadi,
Desalination 2003, 157, 369.
[29] J. Kong, K. Li, Sep. Purif. Technol. 1999, 16, 83.
Acknowledgements [30] G. D. Kang, Y. M. Cao, Water Res. 2012, 46, 584.
[31] C. H. Lee, B. Tiwari, D. Zhang, Y. K. Yap, Environ. Sci.: Nano 2017,
Y.W. thanks the Australian Government for providing the International
Research Training Program (iRTP) scholarship. X.G. acknowledges the 4, 514.
National Natural Science Foundation of China (No. 21774098) for the [32] B. Su, Y. Tian, L. Jiang, J. Am. Chem. Soc. 2016, 138, 1727.
financial support. [33] Z. Chu, Y. Feng, S. Seeger, Angew. Chem., Int. Ed. 2015, 54, 2328.
[34] T. A. Otitoju, A. L. Ahmad, B. S. Ooi, J. Ind. Eng. Chem. 2017, 47, 19.
[35] Y. Wang, X. Gong, J. Mater. Chem. A 2017, 5, 3759.
[36] Q. Wen, Z. Guo, Chem. Lett. 2016, 45, 1134.
[37] X. Gao, X. Wang, X. Ouyang, C. Wen, Sci. Rep. 2016, 6, 27207.
Conflict of Interest [38] F. Li, Z. Yu, H. Shi, Q. Yang, Q. Chen, Y. Pan, G. Zeng, L. Yan,
The authors declare no conflict of interest. Chem. Eng. J. 2017, 322, 33.

Adv. Mater. Interfaces 2018, 1800576 1800576  (24 of 27) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

[39] Z. Wan, D. Li, Y. Jiao, X. Ouyang, L. Chang, X. Wang, Appl. Mater. [80] J. Li, D. Li, Y. Yang, J. Li, F. Zha, Z. Lei, Green Chem. 2016, 18, 541.
Today 2017, 9, 551. [81] M. Tao, L. Xue, F. Liu, L. Jiang, Adv. Mater. 2014, 26, 2943.
[40] Z. Wan, Y. Liu, S. Chen, K. Song, Y. Peng, N. Zhao, X. Ouyang, [82] S. Xiong, L. Kong, J. Huang, X. Chen, Y. Wang, J. Membr. Sci. 2015,
X. Wang, Colloids Surf., A 2018, 546, 237. 493, 478.
[41] Y. Peng, Z. Guo, J. Mater. Chem. A 2016, 4, 15749. [83] X. Du, S. You, X. Wang, Q. Wang, J. Lu, Chem. Eng. J. 2017, 313,
[42] J. Yong, F. Chen, Q. Yang, J. Huo, X. Hou, Chem. Soc. Rev. 2017, 398.
46, 4168. [84] P. Xu, Z. Wang, Z. Xu, J. Hao, D. Sun, J. Colloid Interface Sci. 2016,
[43] A. K. Kota, G. Kwon, A. Tuteja, NPG Asia Mater. 2014, 6, e109. 480, 198.
[44] T. Jiang, Z. Guo, W. Liu, J. Mater. Chem. A 2015, 3, 1811. [85] Y. Chen, J. S. Huang, S. Y. Zhang, Z. W. Gu, Chem. Mater. 2017,
[45] R. K. Gupta, G. J. Dunderdale, M. W. England, A. Hozumi, 29, 3083.
J. Mater. Chem. A 2017, 5, 16025. [86] T. Sun, L. Shu, J. Shen, C. H. Ruan, Z. F. Zhao, C. Jiang, RSC Adv.
[46] Y. Rahmawan, L. Xu, S. Yang, J. Mater. Chem. A 2013, 1, 2955. 2016, 6, 52189.
[47] D. Myers, Surfaces, Interfaces, and Colloids: Principles and [87] Y. Chang, Y. Li, J. Wang, C. W. Wang, Mater. Lett. 2017, 187, 162.
Applications, 2nd ed., John Wiley & Sons, Inc., New York, USA 1999. [88] R. Hensel, C. Neinhuis, C. Werner, Chem. Soc. Rev. 2016, 45, 323.
[48] S. Mosadegh-Sedghi, D. Rodrigue, J. Brisson, M. C. Iliuta, [89] T. S. Wong, S. H. Kang, S. K. Tang, E. J. Smythe, B. D. Hatton,
J. Membr. Sci. 2014, 452, 332. A. Grinthal, J. Aizenberg, Nature 2011, 477, 443.
[49] J. Drelich, E. Chibowski, D. D. Meng, K. Terpilowski, Soft Matter [90] A. Grigoryev, Y. Roiter, I. Tokarev, I. Luzinov, S. Minko, Adv. Funct.
2011, 7, 9804. Mater. 2013, 23, 870.
[50] T. Young, Phil. Trans. R. Soc. London 1805, 95, 65. [91] S. P. Kobaku, A. K. Kota, D. H. Lee, J. M. Mabry, A. Tuteja,
[51] R. N. Wenzel, Ind. Eng. Chem. 1936, 28, 988. Angew. Chem., Int. Ed. 2012, 51, 10109.
[52] A. B. D. Cassie, S. Baxter, Trans. Faraday Soc. 1944, 40, 546. [92] A. R. Bielinski, M. Boban, Y. He, E. Kazyak, D. H. Lee, C. Wang,
[53] L. Gao, T. J. McCarthy, Langmuir 2006, 22, 6234. A. Tuteja, N. P. Dasgupta, ACS Nano 2017, 11, 478.
[54] R. Johnson, R. Dettre, Surf. Colloid Sci. 1969, 2, 85. [93] T. C. Rangel, A. F. Michels, F. Horowitz, D. E. Weibel, Langmuir
[55] G. McHale, N. J. Shirtcliffe, M. I. Newton, Langmuir 2004, 20, 2015, 31, 3465.
10146. [94] T. L. Liu, C. J. Kim, Science 2014, 346, 1096.
[56] A. Marmur, Langmuir 2003, 19, 8343. [95] S. Pan, A. K. Kota, J. M. Mabry, A. Tuteja, J. Am. Chem. Soc. 2013,
[57] X. Gong, A. Bartlett, A. Kozbial, L. Li, Adv. Eng. Mater. 2016, 18, 567. 135, 578.
[58] A. Marmur, Langmuir 2008, 24, 7573. [96] Z. Wang, S. Lin, Water Res. 2017, 112, 38.
[59] A. Tuteja, W. Choi, J. M. Mabry, G. H. McKinley, R. E. Cohen, [97] L. Eykens, K. De Sitter, C. Dotremont, W. De Schepper, L. Pinoy,
Proc. Natl. Acad. Sci. USA 2008, 105, 18200. B. Van Der Bruggen, Appl. Sci. 2017, 7, 118.
[60] M. Nosonovsky, Langmuir 2007, 23, 3157. [98] S. Lin, S. Nejati, C. Boo, Y. Hu, C. O. Osuji, M. Elimelech,
[61] B. S. Kim, P. Harriott, J. Colloid Interface Sci. 1987, 115, 1. Environ. Sci. Technol. Lett. 2014, 1, 443.
[62] W. Barthlott, C. Neinhuis, Planta 1997, 202, 1. [99] J. Lee, C. Boo, W. H. Ryu, A. D. Taylor, M. Elimelech, ACS Appl.
[63] X. Gao, L. Jiang, Nature 2004, 432, 36. Mater. Interface 2016, 8, 11154.
[64] X. F. Gao, X. Yan, X. Yao, L. Xu, K. Zhang, J. H. Zhang, B. Yang, [100] F. Geyer, C. Schonecker, H. J. Butt, D. Vollmer, Adv. Mater. 2017,
L. Jiang, Adv. Mater. 2007, 19, 2213. 29, 1603524.
[65] X. Q. Feng, X. Gao, Z. Wu, L. Jiang, Q. S. Zheng, Langmuir 2007, [101] S. Feng, Z. Zhong, F. Zhang, Y. Wang, W. Xing, ACS Appl. Mater.
23, 4892. Interface 2016, 8, 8773.
[66] L. Feng, Z. Y. Zhang, Z. H. Mai, Y. M. Ma, B. Q. Liu, L. Jiang, [102] S. H. Anastasiadis, Langmuir 2013, 29, 9277.
D. B. Zhu, Angew. Chem., Int. Ed. 2004, 43, 2012. [103] X. J. Liu, Y. M. Liang, F. Zhou, W. M. Liu, Soft Matter 2012, 8, 2070.
[67] M. J. Liu, S. T. Wang, Z. X. Wei, Y. L. Song, L. Jiang, Adv. Mater. [104] B. Xin, J. Hao, Chem. Soc. Rev. 2010, 39, 769.
2009, 21, 665. [105] L. Zhang, Z. Zhang, P. Wang, NPG Asia Mater. 2012, 4, e8.
[68] G. Q. Li, Y. Lu, P. C. Wu, Z. Zhang, J. W. Li, W. L. Zhu, Y. L. Hu, [106] Y. N. Zhou, J. J. Li, Q. Zhang, Z. H. Luo, Langmuir 2014, 30, 12236.
D. Wu, J. R. Chu, J. Mater. Chem. A 2015, 3, 18675. [107] H. S. Lim, J. T. Han, D. Kwak, M. Jin, K. Cho, J. Am. Chem. Soc.
[69] L. Li, Z. Liu, Q. Zhang, C. Meng, T. Zhang, J. Zhai, J. Mater. 2006, 128, 14458.
Chem. A 2015, 3, 1279. [108] D. Wang, P. W. Jiao, J. M. Wang, Q. L. Zhang, L. Feng, Z. Z. Yang,
[70] Z. Sun, T. Liao, W. Li, Y. Dou, K. Liu, L. Jiang, S.-W. Kim, J. Ho Kim, J. Appl. Polym. Sci. 2012, 125, 870.
S. Xue Dou, NPG Asia Mater. 2015, 7, e232. [109] B. Wang, Z. G. Guo, W. M. Liu, RSC Adv. 2014, 4, 14684.
[71] I. E. Palamà, S. D’Amone, V. Arcadio, D. Caschera, R. G. Toro, [110] C.-T. Liu, Y.-L. Liu, J. Mater. Chem. A 2016, 4, 13543.
G. Gigli, B. Cortese, J. Mater. Chem. A 2015, 3, 3854. [111] Y. Xiang, J. Shen, Y. Wang, F. Liu, L. Xue, RSC Adv. 2015, 5, 23530.
[72] Z. X. Xue, S. T. Wang, L. Lin, L. Chen, M. J. Liu, L. Feng, L. Jiang, [112] J. J. Li, Y. N. Zhou, Z. H. Luo, ACS Appl. Mater. Interface 2015, 7,
Adv. Mater. 2011, 23, 4270. 19643.
[73] Y. Q. Liu, Y. L. Zhang, X. Y. Fu, H. B. Sun, ACS Appl. Mater. Interface [113] R. Ou, J. Wei, L. Jiang, G. P. Simon, H. Wang, Environ. Sci. Technol.
2015, 7, 20930. 2016, 50, 906.
[74] X. Liu, J. Zhou, Z. Xue, J. Gao, J. Meng, S. Wang, L. Jiang, [114] H. Liu, X. Zhang, S. Wang, L. Jiang, Small 2015, 11, 3338.
Adv. Mater. 2012, 24, 3401. [115] L. Xu, W. Chen, A. Mulchandani, Y. Yan, Angew. Chem., Int. Ed.
[75] S. J. Hutton, J. M. Crowther, J. P. S. Badyal, Chem. Mater. 2000, 12, 2005, 44, 6009.
2282. [116] C. Ding, Y. Zhu, M. Liu, L. Feng, M. Wan, L. Jiang, Soft Matter
[76] J. A. Howarter, J. P. Youngblood, Adv. Mater. 2007, 19, 3838. 2012, 8, 9064.
[77] J. A. Howarter, K. L. Genson, J. P. Youngblood, ACS Appl. Mater. [117] C. Su, H. Yang, S. Song, B. Lu, R. Chen, Chem. Eng. J. 2017, 309,
Interface 2011, 3, 2022. 366.
[78] J. Yang, Z. Zhang, X. Xu, X. Zhu, X. Men, X. Zhou, J. Mater. Chem. [118] L. Wu, J. Zhang, B. Li, A. Wang, Polym. Chem. 2014, 5, 2382.
2012, 22, 2834. [119] X. Hong, X. F. Gao, L. Jiang, J. Am. Chem. Soc. 2007, 129, 1478.
[79] A. K. Kota, G. Kwon, W. Choi, J. M. Mabry, A. Tuteja, Nat. Commun. [120] J. Y. Huang, Y. K. Lai, L. N. Wang, S. H. Li, M. Z. Ge, K. Q. Zhang,
2012, 3, 1025. H. Fuchs, L. F. Chi, J. Mater. Chem. A 2014, 2, 18531.

Adv. Mater. Interfaces 2018, 1800576 1800576  (25 of 27) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

[121] D. L. Tian, X. F. Zhang, Y. Tian, Y. Wu, X. Wang, J. Zhai, L. Jiang, [158] Q. Wen, J. C. Di, L. Jiang, J. H. Yu, R. R. Xu, Chem. Sci. 2013, 4,
J. Mater. Chem. 2012, 22, 19652. 591.
[122] C. R. Gao, Z. X. Sun, K. Li, Y. N. Chen, Y. Z. Cao, S. Y. Zhang, [159] D. Deng, D. P. Prendergast, J. MacFarlane, R. Bagatin, F. Stellacci,
L. Feng, Energy Environ. Sci. 2013, 6, 1147. P. M. Gschwend, ACS Appl. Mater. Interface 2013, 5, 774.
[123] B. Wang, Z. Guo, Chem. Commun. 2013, 49, 9416. [160] Y. L. Yu, H. Chen, Y. Liu, V. Craig, L. H. Li, Y. Chen, Adv. Mater. Inter-
[124] T. Sun, G. Wang, L. Feng, B. Liu, Y. Ma, L. Jiang, D. Zhu, face 2014, 1, 1300002.
Angew. Chem., Int. Ed. 2004, 43, 357. [161] X. Y. Li, D. Hu, K. Huang, C. F. Yang, J. Mater. Chem. A 2014, 2,
[125] B. L. Xue, L. C. Gao, Y. P. Hou, Z. W. Liu, L. Jiang, Adv. Mater. 11830.
2013, 25, 273. [162] J. Zeng, Z. Guo, Colloids Surf., A 2014, 444, 283.
[126] Q. Yu, X. Li, Y. X. Zhang, L. Yuan, T. L. Zhao, H. Chen, RSC Adv. [163] Y. Cao, X. Zhang, L. Tao, K. Li, Z. Xue, L. Feng, Y. Wei, ACS Appl
2011, 1, 262. Mater. Interface 2013, 5, 4438.
[127] M. J. Liu, X. L. Liu, C. M. Ding, Z. X. Wei, Y. Zhu, L. Jiang, Soft [164] C. H. Lee, N. Johnson, J. Drelich, Y. K. Yap, Carbon 2011, 49, 669.
Matter 2011, 7, 4163. [165] N. Liu, X. Lin, W. Zhang, Y. Cao, Y. Chen, L. Feng, Y. Wei, Sci. Rep.
[128] H. Feng, X. Xu, W. Hao, Y. Du, D. Tian, L. Jiang, Phys. Chem. 2015, 5, 9688.
Chem. Phys. 2016, 18, 16202. [166] X. Lin, J. Heo, H. Jeong, M. Choi, M. Chang, J. Hong,
[129] Q. Cheng, M. Li, Y. Zheng, B. Su, S. Wang, L. Jiang, Soft Matter J. Mater. Chem. A 2016, 4, 17970.
2011, 7, 5948. [167] L. Liu, C. Chen, S. Yang, H. Xie, M. Gong, X. Xu, Phys. Chem.
[130] A. Walther, A. H. Muller, Chem. Rev. 2013, 113, 5194. Chem. Phys. 2016, 18, 1317.
[131] J. Hou, C. Ji, G. Dong, B. Xiao, Y. Ye, V. Chen, J. Mater. Chem. A [168] S. T. Wang, Y. L. Song, L. Jiang, Nanotechnology 2007, 18, 015103.
2015, 3, 17032. [169] B. Wang, Z. G. Guo, Appl. Phys. Lett. 2013, 103, 063704.
[132] H.-C. Yang, J. Hou, L.-S. Wan, V. Chen, Z.-K. Xu, Adv. Mater. Inter- [170] C. Wang, T. Yao, J. Wu, C. Ma, Z. Fan, Z. Wang, Y. Cheng, Q. Lin,
face 2016, 3, 1500774. B. Yang, ACS Appl. Mater. Interface 2009, 1, 2613.
[133] X. Yan, J. Li, L. Yi, Mater. Lett. 2017. [171] B. Chakrabarty, A. K. Ghoshal, M. K. Purkait, J. Membr. Sci. 2008,
[134] H. C. Yang, J. Hou, V. Chen, Z. K. Xu, Angew. Chem., Int. Ed. 2016, 325, 427.
55, 13398. [172] R. Jamshidi Gohari, F. Korminouri, W. J. Lau, A. F. Ismail,
[135] J. Wu, N. Wang, L. Wang, H. Dong, Y. Zhao, L. Jiang, Soft Matter T. Matsuura, M. N. K. Chowdhury, E. Halakoo, M. S. Jamshidi
2012, 8, 5996. Gohari, Sep. Purif. Technol. 2015, 150, 13.
[136] Z. Wang, Y. Wang, G. Liu, Angew. Chem., Int. Ed. 2016, 55, 1291. [173] S. Kumar, C. Guria, A. Mandal, Sep. Purif. Technol. 2015, 150,
[137] X. Tian, H. Jin, J. Sainio, R. H. A. Ras, O. Ikkala, Adv. Funct. Mater. 145.
2014, 24, 6023. [174] M. Obaid, N. A. M. Barakat, O. A. Fadali, S. Al-Meer, K. Elsaid,
[138] J. Gu, P. Xiao, J. Chen, J. Zhang, Y. Huang, T. Chen, ACS K. A. Khalil, Polymer 2015, 72, 125.
Appl. Mater. Interface 2014, 6, 16204. [175] J. A. Prince, S. Bhuvana, V. Anbharasi, N. Ayyanar, K. V. Boodhoo,
[139] X. Hou, Y. Liu, H. Dong, F. Yang, L. Li, L. Jiang, Adv. Mater. 2010, G. Singh, Water Res. 2016, 103, 311.
22, 2440. [176] I. Sadeghi, A. Aroujalian, A. Raisi, B. Dabir, M. Fathizadeh,
[140] J. Zhang, Y. Yang, Z. Zhang, P. Wang, X. Wang, Adv. Mater. 2014, J. Membr. Sci. 2013, 430, 24.
26, 1071. [177] W. Chen, J. Peng, Y. Su, L. Zheng, L. Wang, Z. Jiang, Sep. Purif.
[141] L. Hu, S. J. Gao, Y. Z. Zhu, F. Zhang, L. Jiang, J. Jin, J. Mater. Technol. 2009, 66, 591.
Chem. A 2015, 3, 23477. [178] C. Zhao, J. Xue, F. Ran, S. Sun, Prog. Mater. Sci. 2013, 58, 76.
[142] C. F. Wang, S. Y. Yang, S. W. Kuo, Sci. Rep. 2017, 7, 43053. [179] W. Zhang, Y. Zhu, X. Liu, D. Wang, J. Li, L. Jiang, J. Jin, Angew.
[143] J. Motuzas, C. Yacou, R. S. K. Madsen, W. Fu, D. K. Wang, A. Julbe, Chem., Int. Ed. 2014, 53, 856.
J. Vaughan, J. C. Diniz da Costa, J. Membr. Sci. 2018, 550, 407. [180] X. Lu, Y. Peng, L. Ge, R. Lin, Z. Zhu, S. Liu, J. Membr. Sci. 2016,
[144] J. Zhu, B. Liu, L. Li, Z. Zeng, W. Zhao, G. Wang, X. Guan, 505, 61.
J. Phys. Chem. A 2016, 120, 5617. [181] X. Zhu, W. Tu, K.-H. Wee, R. Bai, J. Membr. Sci. 2014, 466, 36.
[145] J. Wang, G. Geng, Environ. Technol. 2016, 37, 1591. [182] N. M. Kocherginsky, C. L. Tan, W. F. Lu, J. Membr. Sci. 2003, 220,
[146] K. S. Abdel Halim, M. Ramadan, A. Shawabkeh, A. Abufara, 117.
Beni-Suef Univ. J. Basic Appl. Sci. 2013, 2, 72. [183] M. Ulbricht, Polymer 2006, 47, 2217.
[147] W. Liu, N. Canfield, J. Membr. Sci. 2012, 409, 113. [184] G. Liu, X. Xie, Z. Liu, G. Cheng, E. C. Lee, Nanoscale 2018, 10,
[148] Y. H. Ma, J. Catalano, F. Guazzone, Encyclopedia of Membrane 11043.
Science and Technology, John Wiley & Sons, New York, USA 2013. [185] X. Zhu, H.-E. Loo, R. Bai, J. Membr. Sci. 2013, 436, 47.
[149] L. Chen, X. Y. Xie, Z. H. Liu, E. C. Lee, J. Mater. Chem. A 2017, 5, [186] M. L. Huang, Y. Si, X. M. Tang, Z. G. Zhu, B. Ding, L. F. Liu,
6974. G. Zheng, W. J. Luo, J. Y. Yu, J. Mater. Chem. A 2013, 1, 14071.
[150] S. Adhikari, S. Fernando, Ind. Eng. Chem. Res. 2006, 45, 875. [187] H.-C. Yang, K.-J. Liao, H. Huang, Q.-Y. Wu, L.-S. Wan, Z.-K. Xu,
[151] P. F. Ren, H. C. Yang, Y. N. Jin, H. Q. Liang, L. S. Wan, Z. K. Xu, J. Mater. Chem. A 2014, 2, 10225.
RSC Adv. 2015, 5, 47592. [188] W. Zhang, Z. Shi, F. Zhang, X. Liu, J. Jin, L. Jiang, Adv. Mater. 2013,
[152] S. Song, H. Yang, C. Zhou, J. Cheng, Z. Jiang, Z. Lu, J. Miao, 25, 2071.
Chem. Eng. J. 2017, 320, 342. [189] X. Shen, T. Xie, J. Wang, P. Liu, F. Wang, RSC Adv. 2017, 7,
[153] C. R. Crick, J. A. Gibbins, I. P. Parkin, J. Mater. Chem. A 2013, 1, 5262.
5943. [190] K. Venkatesh, G. Arthanareeswaran, A. C. Bose, RSC Adv. 2016, 6,
[154] J. Liu, L. Wang, F. Guo, L. Hou, Y. Chen, J. Liu, N. Wang, Y. Zhao, 18899.
L. Jiang, J. Mater. Chem. A 2016, 4, 4365. [191] W. Ma, Q. Zhang, S. K. Samal, F. Wang, B. Gao, H. Pan, H. Xu,
[155] H. Sun, A. Li, X. Qin, Z. Zhu, W. Liang, J. An, P. La, W. Deng, J. Yao, X. Zhan, S. C. De Smedt, C. Huang, RSC Adv. 2016, 6,
ChemSusChem 2013, 6, 2377. 41861.
[156] L. Zhang, Y. Zhong, D. Cha, P. Wang, Sci. Rep. 2013, 3, 2326. [192] R. Jamshidi Gohari, E. Halakoo, W. J. Lau, M. A. Kassim,
[157] Q. Ma, H. Cheng, Y. Yu, Y. Huang, Q. Lu, S. Han, J. Chen, R. Wang, T. Matsuura, A. F. Ismail, RSC Adv. 2014, 4, 17587.
A. G. Fane, H. Zhang, Small 2017, 13, 1700391. [193] Y. Liu, Y. Su, Y. Li, X. Zhao, Z. Jiang, RSC Adv. 2015, 5, 21349.

Adv. Mater. Interfaces 2018, 1800576 1800576  (26 of 27) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com
www.advmatinterfaces.de

[194] Z. Xiong, H. Lin, Y. Zhong, Y. Qin, T. Li, F. Liu, J. Mater. Chem. A [222] D. G. Kim, H. Kang, S. Han, H. J. Kim, J. C. Lee, RSC Adv. 2013,
2017, 5, 6538. 3, 18071.
[195] X. Zhao, Y. Su, J. Cao, Y. Li, R. Zhang, Y. Liu, Z. Jiang, [223] L. Zhu, M. Chen, Y. Dong, C. Y. Tang, A. Huang, L. Li, Water Res.
J. Mater. Chem. A 2015, 3, 7287. 2016, 90, 277.
[196] G. Zhang, F. Gao, Q. Zhang, X. Zhan, F. Chen, RSC Adv. 2016, 6, [224] S. H. Hyun, G. T. Kim, Sep. Sci. Technol. 1997, 32, 2927.
7532. [225] C. Yang, G. S. Zhang, N. P. Xu, J. Shi, J. Membr. Sci. 1998, 142, 235.
[197] W. Fang, L. Liu, T. Li, Z. Dang, C. Qiao, J. Xu, Y. Wang, Chemistry [226] M. Çakmakce, N. Kayaalp, I. Koyuncu, Desalination 2008, 222, 176.
2016, 22, 878. [227] Z. Zhong, W. Xing, B. Zhang, Ceram. Int. 2013, 39, 4355.
[198] Y. Xiang, F. Liu, L. Xue, J. Membr. Sci. 2015, 476, 321. [228] Q. Chang, J.-e. Zhou, Y. Wang, J. Liang, X. Zhang, S. Cerneaux,
[199] H. Shi, Y. He, Y. Pan, H. Di, G. Zeng, L. Zhang, C. Zhang, X. Wang, Z. Zhu, Y. Dong, J. Membr. Sci. 2014, 456, 128.
J. Membr. Sci. 2016, 506, 60. [229] C. Su, Y. Xu, W. Zhang, Y. Liu, J. Li, Appl. Surf. Sci. 2012, 258, 2319.
[200] A. Raza, B. Ding, G. Zainab, M. El-Newehy, S. S. Al-Deyab, J. Yu, [230] J. Cui, X. Zhang, H. Liu, S. Liu, K. L. Yeung, J. Membr. Sci. 2008,
J. Mater. Chem. A 2014, 2, 10137. 325, 420.
[201] F. Ejaz Ahmed, B. S. Lalia, N. Hilal, R. Hashaikeh, Desalination [231] K. Suresh, T. Srinu, A. K. Ghoshal, G. Pugazhenthi, RSC Adv. 2016,
2014, 344, 48. 6, 4877.
[202] A. Pagidi, R. Saranya, G. Arthanareeswaran, A. F. Ismail, [232] A. S. Barbosa, A. S. Barbosa, M. G. F. Rodrigues, Desalin. Water
T. Matsuura, Desalination 2014, 344, 280. Treat. 2015, 56, 3665.
[203] Y. Liu, Z. Liu, E.-C. Lee, J. Mater. Chem. C 2018, 6, 6705. [233] X. Hu, Y. Yu, J. Zhou, Y. Wang, J. Liang, X. Zhang, Q. Chang,
[204] J. F. Hester, A. M. Mayes, J. Membr. Sci. 2002, 202, 119. L. Song, J. Membr. Sci. 2015, 476, 200.
[205] A. Asatekin, S. Kang, M. Elimelech, A. M. Mayes, J. Membr. Sci. [234] T. Chen, M. Duan, S. Fang, Ceram. Int. 2016, 42, 8604.
2007, 298, 136. [235] J.-e. Zhou, Q. Chang, Y. Wang, J. Wang, G. Meng, Sep. Purif.
[206] A. Akthakul, R. F. Salinaro, A. M. Mayes, Macromolecules 2004, 37, Technol. 2010, 75, 243.
7663. [236] D. Lu, T. Zhang, J. Ma, Environ. Sci. Technol. 2015, 49, 4235.
[207] J. F. Hester, P. Banerjee, A. M. Mayes, Macromolecules 1999, 32, [237] S. J. Gao, H. L. Qin, P. P. Liu, J. Jin, J. Mater. Chem. A 2015, 3,
1643. 6649.
[208] Y. Q. Wang, T. Wang, Y. L. Su, F. B. Peng, H. Wu, Z. Y. Jiang, [238] S. J. Gao, Y. Z. Zhu, F. Zhang, J. Jin, J. Mater. Chem. A 2015, 3,
Langmuir 2005, 21, 11856. 2895.
[209] R. Revanur, B. McCloskey, K. Breitenkamp, B. D. Freeman, [239] N. Liu, M. Zhang, W. Zhang, Y. Cao, Y. Chen, X. Lin, L. Xu, C. Li,
T. Emrick, Macromolecules 2007, 40, 3624. L. Feng, Y. Wei, J. Mater. Chem. A 2015, 3, 20113.
[210] Y. H. Zhao, B. K. Zhu, L. Kong, Y. Y. Xu, Langmuir 2007, 23, 5779. [240] P. Gao, Z. Liu, D. D. Sun, W. J. Ng, J. Mater. Chem. A 2014, 2,
[211] W. J. Chen, Y. L. Su, J. M. Peng, Y. A. Dong, X. T. Zhao, Z. Y. Jiang, 14082.
Adv. Funct. Mater. 2011, 21, 191. [241] Y. Liu, Y. Su, J. Cao, J. Guan, L. Xu, R. Zhang, M. He, Q. Zhang,
[212] L. Yan, Y. S. Li, C. B. Xiang, Polymer 2005, 46, 7701. L. Fan, Z. Jiang, Nanoscale 2017, 9, 7508.
[213] Y. N. Yang, Zhang, H. X., Wang, P., Zheng, Q. Z.  & Li, J., [242] J. Yuan, X. Liu, O. Akbulut, J. Hu, S. L. Suib, J. Kong, F. Stellacci,
J. Membr. Sci. 2007, 288, 231. Nat. Nanotechnol. 2008, 3, 332.
[214] X. J. Chang, Z. X. Wang, S. Quan, Y. C. Xu, Z. X. Jiang, L. Shao, [243] M. H. Tai, P. Gao, B. Y. Tan, D. D. Sun, J. O. Leckie, ACS
Appl. Surf. Sci. 2014, 316, 537. Appl. Mater. Interface 2014, 6, 9393.
[215] L. Q. Ning, N. K. Xu, R. Wang, Y. Liu, RSC Adv. 2015, 5, 57101. [244] B. Lan, L. Yu, T. Lin, G. Cheng, M. Sun, F. Ye, Q. Sun, J. He, ACS
[216] P. Zhang, R. P. Tian, R. H. Lv, B. Na, Q. X. Liu, Chem. Eng. J. 2015, Appl. Mater. Interface 2013, 5, 7458.
269, 180. [245] Z. Liu, S. Y. Lee, E. C. Lee, Appl. Phys. Lett. 2014, 223306.
[217] S. Belfer, R. Fainshtain, Y. Purinson, J. Gilron, M. Nystrom, [246] Z. Shi, W. Zhang, F. Zhang, X. Liu, D. Wang, J. Jin, L. Jiang,
M. Manttari, J. Membr. Sci. 2004, 239, 55. Adv. Mater. 2013, 25, 2422.
[218] A. Rahimpour, S. S. Madaeni, S. Zereshki, Y. Mansourpanah, [247] S. J. Gao, Z. Shi, W. B. Zhang, F. Zhang, J. Lin, ACS Nano 2014,
Appl. Surf. Sci. 2009, 255, 7455. 8, 6344.
[219] Y. Z. Zhu, F. Zhang, D. Wang, X. F. Pei, W. B. Zhang, J. Jin, [248] L. Hu, S. Gao, X. Ding, D. Wang, J. Jiang, J. Jin, L. Jiang, ACS Nano
J. Mater. Chem. A 2013, 1, 5758. 2015, 9, 4835.
[220] X. T. Zhao, Y. L. Su, W. J. Chen, J. M. Peng, Z. Y. Jiang, [249] Z. Shi, X. Chen, X. Wang, T. Zhang, J. Jin, Adv. Funct. Mater. 2011,
J. Membr. Sci. 2012, 415, 824. 21, 4358.
[221] X. F. Wang, K. Zhang, Y. Yang, L. L. Wang, Z. Zhou, M. F. Zhu, [250] X. Zhao, Y. Su, Y. Liu, Y. Li, Z. Jiang, ACS Appl. Mater. Interface
B. S. Hsiao, B. Chu, J. Membr. Sci. 2010, 356, 110. 2016, 8, 8247.

Adv. Mater. Interfaces 2018, 1800576 1800576  (27 of 27) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

You might also like