You are on page 1of 16

REVIEWS

Materials for next-generation


desalination and water purification
membranes
Jay R. Werber, Chinedum O. Osuji and Menachem Elimelech
Abstract | Membrane-based separations for water purification and desalination have been
increasingly applied to address the global challenges of water scarcity and the pollution of aquatic
environments. However, progress in water purification membranes has been constrained by the
inherent limitations of conventional membrane materials. Recent advances in methods for
controlling the structure and chemical functionality in polymer films can potentially lead to new
classes of membranes for water purification. In this Review, we first discuss the state of the art of
existing membrane technologies for water purification and desalination, highlight their inherent
limitations and establish the urgent requirements for next-generation membranes. We then describe
molecular-level design approaches towards fabricating highly selective membranes, focusing on
novel materials such as aquaporin, synthetic nanochannels, graphene and self-assembled block
copolymers and small molecules. Finally, we highlight promising membrane surface modification
approaches that minimize interfacial interactions and enhance fouling resistance.

Addressing the lack of adequate and safe water is among The management of complex industrial wastewaters,
the most important challenges of our time1. Water is such as those from the oil and gas industries and coal-
Earth’s most precious resource for life and is becoming fired power plants, has become a major challenge in
perilously scarce and befouled1,2. Population growth, recent years6–8. Such wastewaters are characterized by a
industrialization and climate change constitute an high level of total dissolved solids, up to six times that of
expanding list of stressors to water resources that exacer- seawater, and a high fouling potential. These wastewaters
bate global water scarcity. To address this challenge, there are therefore difficult to treat 8. Furthermore, potential
is an important need for the development of efficient and new regulations may require zero liquid discharge for
sustainable technologies to tap unconventional sources industrial wastewaters (for example, from coal-fired
of water (for example, seawater, brackish groundwater power plants) or brines from inland desalination plants.
and wastewater) to augment water supply beyond what This will necessitate the development of low-cost, energy-
is obtainable from the hydrologic water cycle2. efficient technologies that effectively concentrate feed
A global emerging problem is the contamination of waters up to a very high level of total dissolved solids
drinking water sources, such as rivers and lakes, with and, eventually, to a solid salt 9,10.
micropollutants that originate from municipal, indus- In addition to mining fresh water from waste­
trial and agricultural wastewaters3. Micropollutants, water, there is a growing interest in recovering valuable
which include compounds such as hormones, pharma- resources from municipal and industrial wastewaters11,12.
Department of Chemical and ceuticals, pesticides, personal care products and indus- For example, municipal wastewater is now considered
Environmental Engineering, trial chemicals, can have detrimental effects even at low to be a renewable resource from which water, nutrients
Yale University, New Haven, concentrations. Furthermore, municipal wastewater (for example, phosphate and nitrogen), energy and sub-
Connecticut 06520–8286, effluents in water-stressed regions are often recycled for stances such as bioplastics can be extracted13,14. Resource
USA.
agriculture and even for indirect potable use, thereby recovery will be an integral part of wastewater treatment
Correspondence to M.E. 
introducing a wide range of micropollutants to fresh in the coming decades, and novel separation technologies
menachem.elimelech@yale.
edu water resources4,5. The development of efficacious tech- will be indispensable in this effort.
nologies for the removal of such micropollutants and Membrane-based treatment technologies can have a
Article number: 16018
doi:10.1038/natrevmats.2016.18 other contaminants, as listed in TABLE 1, is of paramount key role in water purification and desalination. These
Published online 5 Apr 2016 importance to ensure clean and safe drinking water. technologies produce water of superior quality, are

NATURE REVIEWS | MATERIALS VOLUME 1 | MAY 2016 | 1


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

less sensitive to feed quality fluctuations and have a and high fouling propensity 15. Novel materials and scal-
much smaller footprint compared with conventional able, molecular-level design approaches for membrane
‘Victorian’ water treatment technologies2. In addition, fabrication are imperative for overcoming these limita-
membrane-based desalination technologies are inher- tions and for substantially advancing water purification
ently more energy efficient than thermal approaches15. and desalination technologies.
For example, the energy required for seawater desali- Here, we review polymer-based membrane materi-
nation by current state-of-the-art reverse osmosis (RO) als for desalination and water purification, focusing on
technology is already within a factor of two from the molecular-level design approaches. We highlight the
limit of a reversible thermodynamic process, whereas main limitations and requirements of current membrane
thermal desalination technologies can consume as much materials, discuss how these limitations and needs can
as five times the energy used in RO16. be addressed using new materials, and identify where
Current materials and fabrication methods for mem- new approaches are needed to significantly advance
branes are largely based on empirical approaches and water purification and desalination technologies.
lack molecular-level design17,18. Consequently, control
of the structure in the selective layer of the membrane State-of-the-art membranes
is limited, thus hampering membrane performance and Membrane technologies for water purification and desal-
increasing the cost of water purification and desalina- ination with separation largely based on size include the
tion. Recent improvements in water purification mem- established microfiltration (MF), ultrafiltration (UF),
branes have been marginal because of inherent material nanofiltration (NF) and RO processes, as well as the
limitations, including the permeability–selectivity trade- emerging forward osmosis (FO) process. BOX 1 pro-
off, which limits the achievable water–solute selectivity, vides an introduction to these processes. A wide range

Table 1 | Contaminants of concern in source waters and membrane technologies that can meet treatment objectives
Source water Key contaminants Treatment objectives Membrane technologies Problems and challenges
Natural waters
Seawater Boric acid (affects crop health), • Reduce salinity • Pre-treatment: • Energy of desalination
divalent cations (cause scaling microfiltration (MF) and • Ecological impact of
issues) ultrafiltration (UF) seawater intake and brine
• Desalination: reverse discharge
osmosis (RO)
Brackish (saline) Divalent cations • Reduce salinity • Pre-treatment: MF and • Inland discharge of saline
groundwater UF brine
• Desalination: RO and
electrodialysis*
Surface waters Natural organic matter (precursor • Remove particles and • MF, UF, nanofiltration (NF) • High fouling potential
for disinfection by-products), microbial pathogens and RO • Formation of toxic
microbial pathogens, • Reduce natural organic disinfection by-products
micropollutants and algal toxins matter during oxidative
disinfection
Fresh groundwater Natural arsenic, nitrates, iron and • Reduce nitrate, iron, • MF, UF, NF and RO • Seawater intrusion for
manganese (staining issues) and manganese and/or aquifers near ocean
divalent cations scale-forming ions • Aquifer over-exploitation
Wastewaters
Municipal Microbial pathogens, • Degrade organic matter • MF, UF, NF and RO • Requirements for potable
wastewater micropollutants, phosphates • Remove or inactivate reuse
(algal bloom concerns) and pathogens • Large footprint
ammonia • Remove nutrients (that is, • Odour of conventional
nitrogen and phosphate) treatment plants
• High membrane-fouling
potential
Shale-gas produced Drilling fluid additives (for • Drilling reuse: remove • Drilling reuse: MF and UF • Large water consumption
water example, surfactants, oxidants, suspended solids and • Desalination: NF, RO • High total dissolved solids
strong acids), oil and grease, scalants (for low-salinity waters), • Fluctuating water quantity
radium and divalent cations • Disposal: remove oily forward osmosis (FO) and and quality
compounds and reduce membrane distillation • Regulations over disposal
salinity (MD)‡ • High membrane-fouling
potential
Coal-fired power Toxic metals (for example, • Remove dissolved toxic • MF, UF, NF, RO, FO and • Potential zero liquid
plant flue gas arsenic, selenium and mercury), metals and reduce salinity MD discharge requirements
desulfurization ammonia and organic acids
*Electrodialysis is an electrically driven desalination method that uses ion-exchange membranes. ‡Membrane distillation is a thermally driven desalination process in
which a hydrophobic porous membrane separates a hot feed stream from a cold permeate. A vapour pressure gradient drives water vapour flux.

2 | MAY 2016 | VOLUME 1 www.nature.com/natrevmats


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

of membrane materials, both polymeric and inorganic, evaporate from a liquid film that includes polymer and non-
have been used. Polymeric membranes are by far the solvent, and thermally induced phase separation, in
most widespread, largely owing to their high process- which a high-temperature solution of dissolved polymer
ability and low cost. State-of-the-art polymeric mem- is slowly cooled until phase separation occurs.
branes and their corresponding fabrication methods are During phase inversion, several key parameters
discussed in this section. determine the resulting membrane morphology: for
example, the nature of the solvent and non-solvent, poly­
Phase inversion membranes. Most commercially avail- mer type, polymer concentration and fabrication tech-
able porous membranes that are used for MF and UF, nique. As shown in FIG. 1b, a wide range of morphologies
as well as some dense membranes used for RO and FO, can result, including asymmetric membranes for UF that
are formed by phase inversion. This process involves the have a thin nanoporous selective layer (which is often
controlled precipitation of a dissolved polymer in a thin called the active layer) and a microporous underlying
film to produce a porous membrane structure17. The main structure, and relatively symmetric microporous mem-
phase inversion technique used is non-solvent-induced branes for MF. However, owing to the stochastic nature
phase separation (NIPS), in which a film of polymer of the phase inversion process, membranes formed by
dissolved in solvent is immersed in a non-solvent bath, traditional phase inversion have polydisperse pore size
typically a water bath, leading to solvent–non-solvent distributions, which adversely affect the selectivity of the
exchange and phase separation into polymer-rich and resulting active layer, as discussed later.
polymer-poor phases (FIG. 1a). These two phases ultimately In contrast to phase inversion, an alternative process
form the solid membrane matrix (from the polymer-rich called track etching can form UF and MF membranes
phase) and pores (from the polymer-poor phase). Other with pores of uniform size17. Track etching involves two
phase inversion techniques include controlled solvent steps: bombardment with charged particles to partially
evaporation, in which a volatile solvent is allowed to degrade a track of polymer, followed by chemical etching

Box 1 | Membrane processes for water purification and desalination


0.1 nm 1 nm 10 nm 100 nm 1 μm 10 μm

Hydrated ions Viruses


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

Reverse osmosis,
Ultrafiltration
forward osmosis

Nanofiltration Microfiltration

Solution–diffusion Size exclusion

Several common membrane processes for water purification and desalination separate largely Nature Reviews
on the basis |ofMaterials
solute
size. Processes that are driven by hydraulic pressure include microfiltration (MF), ultrafiltration (UF), nanofiltration
(NF) and reverse osmosis (RO). Additionally, the process of forward osmosis (FO) is driven by an osmotic pressure
difference between the feed source water and a specialized solution of high osmotic pressure, which is referred to as
a draw solution17,124. The processes are classified on the basis of the size of the solutes that are retained, as shown in
the figure (in which the sizes refer to the diameter). The main function of MF membranes is to remove suspended
particles and microbial pathogens. UF membranes are designed to retain macromolecules, such as natural organic
matter, as well as smaller pathogens, such as viruses, which are only partially removed by MF. The molecular weight
cut-off (that is, the solute size at which 90% of species are rejected) of UF membranes is between approximately 5
and 500 kDa. MF and UF membranes are porous, and molecular separation is based on a sieving or size exclusion
mechanism. NF membranes can effectively remove scale-forming ions, such as calcium and magnesium, and can
partially reduce salinity. Typical molecular weight cut-offs of NF membranes are between 100 and 300 Da. RO and FO
membranes are designed for desalination. Current RO and FO membranes are non-porous and can remove nearly all
ions in addition to uncharged solutes of molecular weight greater than around 100 Da. Molecular transport in RO and
FO is governed by a solution–diffusion mechanism in which molecules partition into the selective layer of the
membrane, diffuse through the polymer network and desorb in the permeate17,24. Separation of species stems from
differences in their solubility and/or diffusivity in the selective layer. Separation in NF is based on a combination of
sieving and solution–diffusion mechanisms.

NATURE REVIEWS | MATERIALS VOLUME 1 | MAY 2016 | 3


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

a b

Polymer, solvent
and additives
Thin film
Solid support
40 μm 20 μm

Non-solvent Solvent
Solid membrane
matrix
Pores

250 nm 2 μm

c m-Phenylene Trimesoyl chloride Crosslinked polyamide d


diamine
O O O O
H2N
Cl Cl N N N
H H H

NH2
O Cl O N N 500 nm
H H

Sulfur
Nitrogen

200 nm 200 nm

Figure 1 | State-of-the-art membranes for water treatment. a | A schematic diagram of phase inversion by non-solvent-
Nature Reviews | Materials
induced phase separation to form microfiltration (MF) and ultrafiltration (UF) membranes. A casting solution consisting of
polymer, solvent and desired additives is cast on a solid support as a thin film, which is followed by immersion in a non-solvent to
induce phase separation. b | Scanning electron microscopy (SEM) images of membranes formed from phase inversion. The upper
left panel is a cross-section of an asymmetric UF membrane with finger-like macrovoids cast from 9% polysulfone (PSf) in
dimethylformamide (DMF). The upper right panel is a cross-section of an asymmetric UF membrane with a sponge-like structure
cast from 12% PSf in DMF. The lower left panel is a top view of a hand-cast PSf UF membrane. The lower right panel is a top view of
a commercial 0.22-μm nominal pore size polyvinylidene fluoride MF membrane (EMD Millipore, Billerica, Massachusetts, USA).
c | A schematic representation of interfacial polymerization to form thin-film composite (TFC) reverse osmosis membranes.
Aqueous m‑phenylenediamine diffuses to the water/organic interface, reacting with trimesoyl chloride to form a dense
polyamide film with ridge-and-valley morphology. d | Microscopy images of a hand-cast TFC membrane. The upper panel
shows a top-view SEM image of rough polyamide film at a 45° viewing angle. The lower left panel is a transmission electron
microscopy (TEM) image of a thin cross-section showing the rough polyamide film formed on a PSf UF membrane. The
lower right panel shows a corresponding scanning TEM energy-dispersive X‑ray spectroscopy image; the presence of
sulfur and nitrogen delineates the PSf layer and polyamide film, respectively. The upper panels of part b are adapted with
permission from REF. 125, Elsevier. The upper panel of part d is adapted with permission from REF. 126, American Chemical
Society. The lower panels of part d are adapted with permission from REF. 20, American Chemical Society.

to form pores. The main limitation of track-etched poly­amide membranes can achieve water permeability
membranes is their low porosity (that is, low void vol- and salt rejection far exceeding those for asymmetric
ume within the membrane), which must be kept below cellulose-acetate-based membranes, which were the
5% to prevent pore overlap. The low porosity results in first-generation RO membranes developed more than
low water permeability, largely limiting track-etched 50 years ago17. This performance, combined with the
membranes to analytical separations. stability of TFC membranes over a wide pH range
(pH 2–11 for continuous operation), has led to their
Thin-film composite polyamide membranes. Thin- widespread use in desalination applications.
film composite (TFC) polyamide membranes are the The polyamide selective layer is formed through inter-
gold standard for NF, RO and FO applications15. These facial polymerization of a diamine — most commonly,
membranes comprise a non-porous, highly crosslinked m‑phenylenediamine (MPD) for RO, FO and NF, and
polyamide selective layer and an underlying porous piperazine (PIP) for NF — with a triacyl chloride, par-
support layer, typically made of polysulfone. TFC ticularly trimesoyl chloride (TMC). During interfacial

4 | MAY 2016 | VOLUME 1 www.nature.com/natrevmats


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

polymerization (FIG. 1c), the polysulfone support is first This simplified relationship best describes an idealized
brought into contact with the aqueous diamine solution membrane with uniformly sized cylindrical pores (that
to allow diamine penetration into the support. After is, an isoporous membrane). In reality, membranes that
removing the excess solution, the support is immersed in are formed by phase inversion have non-cylindrical and
an organic phase containing TMC, which is not soluble in tortuous pores that vary substantially in size.
water. As diamine monomers diffuse to the water/organic In addition to affecting water permeability, the
interface, they react with TMC to form the polyamide pore size distribution largely determines the rejection
film, which adheres to the polysulfone support through (R = 1 − cpermeate/cfeed) of solutes present in the feed solution.
physical interactions. Subsequent hydrolysis of non- For a solute of radius a, rejection by a pore of radius rp
reacted acyl chloride groups yields carboxyl groups, (with rp ≥ a) is a function of the ratio a/rp, as shown by
which confer a negative charge to the membrane sur- equation (3)23:
face. TFC RO membranes formed using aromatic MPD R = 1 – [2(1 – rap )2 – (1 – rap )4]exp[–0.7146( rap )2] (3)
and TMC monomers are fairly hydrophobic (with water
contact angles of 50–60°), whereas TFC NF membranes For complete solute rejection (cpermeate = 0), the entire
formed using PIP and TMC are more hydrophilic (with pore size distribution must be smaller than the solute of
water contact angles of approximately 30°)19. Both MPD- interest (rp ≤ a). Similarly, to completely reject a mixture of
based and PIP-based TFC membranes are sensitive to solutes, the largest pore must be smaller than the smallest
chemical degradation by strong oxidants, most notably solute that is to be retained.
chlorine17. This sensitivity eliminates the usage of chlo- For non-porous, dense membranes, water and solute
rine as a membrane-cleaning agent and necessitates transport is governed by the solution–diffusion model.
removal of chlorine before the membrane step if it is Water and solute molecules partition into the active layer
present in the feed stream. of the membrane, diffuse through the polymer matrix
The surface morphology of TFC membranes also down their chemical potential gradients and desorb into
depends on the monomers used. PIP-based membranes the permeate solution. The solubility and diffusivity
are smooth and have a root-mean-square roughness of combine to define the diffusive water permeability (Pw)
less than 10 nm. By contrast, MPD-based membranes and diffusive solute permeability (Ps), both of which are
have a characteristic ‘ridge-and-valley’ structure, typically intrinsic material properties — that is, they are independ-
having a root-mean-square roughness greater than 50 nm ent of thickness. The water permeability coefficient of the
and an overall end‑to‑end thickness of 100–500 nm. The active layer is related to the diffusive water permeability
roughness is due to protruding polyamide nodules20, as and layer thickness by equation (4)24:
can be seen in FIG. 1d. The nodules are generally hollow; PwVw
transmission electron microscopy images suggest that the A= (4)
δmRgT
nodule walls are roughly 10–30 nm thick20,21. Although
the rough nodular morphology increases fouling pro- where Vw is the molar volume of water, Rg is the gas con-
pensity (as discussed later), the increased surface area stant and T is the absolute temperature. Similarly, solute
stemming from this morphology may also increase water transport is modelled as Fickian diffusion (equation (5))24:
permeability 21,22. Ps
Js = Δcm = BΔcm (5)
δm
Inherent limitations and urgent requirements
The permeability–selectivity trade-off. Water flux where B is the solute permeability coefficient and Δcm
through porous (that is, MF and UF) membranes is the solute concentration difference across the active
and non-porous (that is, NF and RO) membranes is layer. Together, the coefficients A and B largely define the
described by equation (1)17: selective layer performance of non-porous membranes.
The ideal membrane has both high water permeabil-
Jw = A(ΔP –Δπ m) (1)
ity and high selectivity — a combination that is difficult
where Jw is the volumetric water flux, A is the water to attain. In recent studies, it has been proposed that
permeability coefficient, ΔP is the applied hydraulic porous and non-porous water separation membranes
transmembrane pressure and Δπm is the osmotic pres- exhibit a permeability–selectivity trade-off that is sim-
sure difference across the active layer. In typical water ilar to the Robeson plot that has found extended use in
treatment applications, Δπm is negligible for MF and UF, the field of polymeric gas separation membranes23,25,26.
whereas it is a crucial parameter for NF and RO. These trade-offs all have a common implication: an
For porous membranes, flow through the active layer increase in permeability for a highly permeating species
is modelled as laminar flow through an array of cylin- (for example, water) comes with an even greater increase
drical pores (FIG. 2a, inset). Water permeability is related in permeability for relatively highly retained species (for
by the Hagen–Poiseuille equation (equation (2)) 17 to the example, macromolecules for UF or salt for RO).
solution viscosity (μ) and morphological characteristics For porous membranes, selectivity is intrinsically
of the active layer — namely, the surface porosity (ε), linked with active layer pore size. This relationship is
pore radius (rp) and active layer thickness (δm): shown in FIG. 2a for UF membranes, in which selectivity
for water over bovine serum albumin, given as the sepa-
εrp2
A= (2) ration factor 1/(1 − R), decreases sharply with increasing
8μδm water permeability at fixed surface porosity and active

NATURE REVIEWS | MATERIALS VOLUME 1 | MAY 2016 | 5


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

layer thickness23. This trade-off occurs for an ideal iso- a loss in permselectivity for independently fabricated
porous membrane when the pore size exceeds the solute membranes. These improvements are possibly due to
size and is even more pronounced for real phase inversion optimization of the interfacial polymerization reaction
membranes, which have a heterogeneous pore size dis- conditions to yield slightly different film chemistries,
tribution owing to the stochastic nature of selective layer polymer network structures and film morphologies.
formation. For either case, decreasing the active layer An updated trade-off relationship is shown in FIG. 2b,
thickness or increasing surface porosity would increase reflecting currently achieved active layer performance.
water permeability without sacrificing selectivity. Although the pattern of incremental performance gains
For non-porous membranes, a permeability–selectivity in commercial TFC membranes is likely to continue,
trade-off was originally found by comparing the diffu- water–salt separation in these membranes is still fun-
sive water and salt permeabilities for several bulk poly­ damentally dependent on the separating abilities of
mer materials25. A cubic relationship was proposed, the aromatic polyamide material. Therefore, an upper
wherein Ps ∝ Pw 3. A similar trade-off relationship was performance limit is still expected to exist for TFC
subsequently found to hold true for coefficients A and membranes, motivating research on novel materials to
B (that is, B ∝ A3) for a commercial TFC RO membrane achieve further performance gains.
that was treated stepwise with chlorine and alkaline to
increase permeability27, as shown in FIG. 2b. Also shown Selectivity, not permeability, is key for desalination.
in FIG. 2b are the permselectivities (A/B) for current Research on novel active layers for desalination mem-
commercial membranes, values for which were calcu- branes seeks to improve the water permeability or water–
lated on the basis of the technical specifications pro- solute selectivity, with permeability often being the focus.
vided by the manufacturer 28. A comparison with the However, for both seawater RO (SWRO) and FO desal-
proposed TFC trade-off relationship shows that many ination, process modelling shows that increased water
of the membranes that are currently available outper- permeability above levels that are currently achievable
form the previously proposed trade-off. In other words, (that is, 0.5–0.8 μm s−1 bar−1 or 2–3 l m−2 h−1 bar−1; FIG. 2b)
in contrast to the results for a single chemically treated would negligibly decrease energy requirements and capi-
membrane, incremental gains in the water permeabil- tal costs29,30. For example, increasing the water permeabil-
ity coefficient have been attained without experiencing ity coefficient from 3 to 10 L m−2 h−1 bar−1 would decrease

a Water permeability coefficient, A (L m–2 h–1 bar–1) b Water permeability coefficient, A (L m–2 h–1 bar–1)
0 1000 2000 1 10
104

a 100
Water–salt permselectivity, A/B (bar–1)

103 rp
Separation factor, (1 – R)–1

102
10

101

100 1
0 1 2 3 4 5 6 7 8 1 10
Water permeability coefficient, A (10 m s bar )
–4 –1 –1
Water permeability coefficient, A (10–6 m s–1 bar–1)
Acrylic- and acrylonitrile-based polymers TFC RO membranes
Polysulfone-based polymers Chlorine- and alkaline-treated RO membrane
Cellulose-based polymers
Nature Reviews | Materials
Figure 2 | Permeability–selectivity trade-off relationships for ultrafiltration and reverse osmosis membranes.
a | Ultrafiltration (UF) membranes. Experimentally measured water permeabilities and rejection values of bovine serum
albumin are shown for membranes formed by phase inversion using polyacrylonitrile, polysulfone-based polymers and
cellulose-based polymers. The separation factor uses true rejection values (that is, rejection over just the active layer of the
membrane). The lines are modelled using equations (2) and (3) and a fixed surface porosity/active layer thickness ratio (ε/δm)
of 1 μm−1. The solid line assumes a log-normal distribution of pore sizes with a standard deviation of 0.2 relative to the mean
radius. The dashed line assumes a uniform pore size. The insets show how increasing the pore radius (rp) relative to the
solute radius (a) results in the observed trade-off. Data for panel a are from REF. 23. b | Reverse osmosis (RO) membranes.
The blue squares represent the permselectivity values for commercial thin-film composite (TFC) RO membranes. These
values are calculated from technical documentation of the manufacturer28. The green triangles represent permselectivity
values measured for a chlorine- and alkaline-treated commercial seawater RO membrane27. The dashed line is the
corresponding proposed trade-off relationship for TFC membranes. The solid line is an updated trade-off relationship that
better fits the performance of current membranes.

6 | MAY 2016 | VOLUME 1 www.nature.com/natrevmats


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

the SWRO energy requirements by less than 2%29. but their application is limited by a narrow operating
This limited difference largely stems from the single- pH range of 2–8 and an operating temperature of less
stage operation of RO, which necessitates the use of a than 30 °C (REF. 36). In addition to material proper-
hydraulic pressure greater than the osmotic pressure of ties, the polydisperse pore size distribution of MF and
the brine exiting the RO stage, irrespective of the mem- UF membranes that arises from phase inversion can
brane permeability 15. The use of hydraulic pressure is the increase fouling propensity. Large pores tend to foul
main determinant of the energy used by the RO stage. rapidly owing to high local fluxes and can suffer from
Instead of attaining higher water permeability, a irreversible fouling (that is, fouling by substances that
more effective goal for desalination membrane materi- are not removed with rinsing steps) because of foulant
als research would be to attain improved selectivity for penetration into the membrane interior 38.
water over all dissolved solutes. For example, despite a TFC polyamide membranes used in RO, FO and
high salt rejection (>99.5% at standard test conditions) NF are highly susceptible to all three types of fouling,
for SWRO membranes, multiple RO passes are still largely owing to a combination of surface morphology,
sometimes needed, such as in the production of high- hydrophobicity and charge. The high surface roughness
purity water for industrial use31. In addition, rejection of MPD-based TFC membranes enhances all types of
by current membranes of small neutral species is fairly fouling as a result of increased surface area, thus provid-
poor 32,33. Some of these species, such as boric acid in sea- ing greater foulant–membrane interaction and greater
water and some micropollutants in wastewater, adversely opportunity for attachment 39. MPD-based TFC mem-
affect human and/or crop health. Incomplete rejection of branes are also fairly hydrophobic, thereby increasing
these species necessitates extensive post-treatment steps, rates of organic and biological fouling 40. Surface charge
such as further RO passes, ion exchange and oxidative exacerbates fouling for both MPD- and PIP-based TFC
chemical degradation2,34. These additional steps require polyamide membranes because of the surface carboxyl
substantially higher chemical and energy usage, as well functional groups. Through calcium bridging, these car-
as incurring higher capital costs, which could be avoided boxyl groups increase the binding of organic foulants
if more selective RO membranes were available. such as natural organic matter (that is, partially aromatic
macromolecules with high levels of carboxylic and phe-
High fouling propensity. Membrane fouling — that is, nolic functional groups)41. The enrichment of calcium
the accumulation of substances on the membrane sur- at the surface can also induce calcium sulfate scaling
face or within the membrane pores — is a major obstacle through surface nucleation and growth42.
for the efficient operation of membrane systems. There
are three types of fouling: organic fouling by adsorption Designing highly selective membranes
of dissolved organic matter; scaling by deposition of Recent advances in chemical synthesis and directed
precipitated salts or by surface nucleation and growth assembly offer the possibility of molecular-level design
of sparingly soluble salts; and biological fouling (bio- of selective layers, potentially facilitating the fabrication
fouling) by deposition and growth of microorganisms of highly selective membranes for both water purifi-
to form strongly adherent biofilms17. Excessive fouling cation and desalination (that is, for UF, RO and FO).
deteriorates membrane performance (that is, water flux For desalination membranes, specially designed water
and selectivity), necessitates chemical cleaning, which is channels could lead to a shift in how water itself moves
unsustainable and shortens membrane life, and increases through the membrane: from partitioning into and dif-
energy consumption and operating costs17. The sever- fusing through polymeric thin films to single-file trans-
ity of fouling can vary substantially and is affected not port through active-layer-spanning water channels.
only by feed water quality and process conditions (for Such sub-nanoporous water channels, if well designed,
example, water flux and hydrodynamics), but also by the could offer a significant improvement in desalination
fouling propensity of the membrane itself. The develop- performance.
ment of membrane materials with low fouling propensity
remains highly important. Aquaporin as a model water channel. In the design of
Many UF and MF applications require a high per- sub-nanoporous water channels, materials research can
formance despite feed streams that have a high fouling gain insight from the biological water channel aqua-
potential: for example, the microorganism-filled sludge porin. These ubiquitous, integral membrane proteins
that accumulates in membrane bioreactors for munici- have homologues found in bacteria, archaea, plants
pal wastewater treatment 35. Phase inversion UF and MF and animals. They exist as tetramers of four identical
membranes commonly comprise synthetic polymers, monomers, and each monomer bears a water channel43.
including polyvinylidene fluoride (PVDF), poly­sulfone, Aquaporin is remarkable for its unique transport prop-
poly­ethersulfone (PES) and polyacrylonitrile (PAN)17,36. erties: some types of aquaporin are not only highly per-
These poly­mers are easily processable, chemically, meable to water, but also extremely selective, rejecting
thermally and mechanically robust, and inexpensive. ions, protons and even small neutral solutes, such as
However, they are also fairly hydrophobic, which can urea, that readily permeate through current desalination
exacerbate organic and biological fouling through membranes44–47.
increased adsorption of foulants to the membrane sur- The high permeability of the aquaporin water
face and interior pores37. Cellulose-acetate-based mem- channel stems from its hourglass structure (FIG. 3a),
branes are more hydrophilic and more fouling resistant, which features wide entrance vestibules and a narrow

NATURE REVIEWS | MATERIALS VOLUME 1 | MAY 2016 | 7


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

a Aquaporin Carbon Synthetically b


nanotubes designed nanochannels Block copolymer
(10–100 kDa)

5–50 nm

Polymerizable
surfactants
(~1 kDa) Graphene-based
frameworks

500 nm

Figure 3 | Selective membranes formed using molecular-level design. a | Sub-nanoporous materials Nature Reviews | Materials
with promise for
desalination. The top panel shows discrete channels spanning a thin, nearly impermeable film, including the membrane
protein aquaporin, single-walled carbon nanotubes with diameters ≤0.8 nm to exclude salt, self-assembled channel-forming
macrocycles and single-molecule synthetic water channels. The bottom panel shows contiguous media that fully form a
selective layer, including inverse hexagonal channels formed using wedge-shaped polymerizable surfactants, tortuous
channels formed in bicontinuous cubic phase using polymerizable surfactants and graphene-based frameworks, shown with
either graphene oxide or reduced graphene oxide as the grey sheets and covalent linkers in green. In all of the schematics,
water molecules are shown traversing the nanochannels. b | Controlled formation of the nearly isoporous ultrafiltration
membrane using block copolymers. Scanning electron microscopy image in panel b is from REF. 90, Nature Publishing Group.

hydrophobic centre that supports single-file transport of or block copolymer bilayer 49,50. Protein stability is an
water molecules46. Selectivity is size- and charge-based: additional concern. Although AqpZ reconstituted in
0.8 nm above the channel centre, the channel constricts vesicles remains active after storage at 4 °C for several
to just 0.28 nm in diameter (which is approximately the months43, it is likely that it will not be stable enough for
size of a water molecule) owing to protruding phenyl­ continuous use.
alanine, histidine and arginine side chains. The arginine
side chain provides the constriction site with a fixed Discrete synthetic nanochannels. Using aquaporin as a
positive charge. This tight, charged constriction site model, it is theoretically possible to systematically design
provides a high energy barrier that prevents permeation discrete water channels that are suitable for desalination
of charged and neutral solutes44–46,48. and stable enough for easy processing and use. Such
Owing to its exceptional transport properties, aqua- channels should ideally exhibit the following attributes:
porin has recently been assessed as the selective agent in molecular-sieve-like constriction to exclude salt and
desalination membranes49–51. Selectivity of a defect-free small neutral solutes; stable alignment within a poly-
aquaporin-based membrane should be very high and, meric thin film; intrinsic water permeability and achiev-
depending on the achievable packing density, water able packing density that are high enough to match the
permeability could match that of TFC membranes water permeability coefficient of TFC membranes;
(TABLE 2). Membrane fabrication has typically followed a chemical and mechanical stability; and ease of fabrication
vesicle-based approach. The fairly stable bacterial aqua- using consistent and scalable synthesis methods.
porin Z (AqpZ) is first incorporated into unilamellar Despite their structural simplicity, carbon nanotubes
lipid or block copolymer vesicles, with AqpZ spanning (CNTs) have the potential for remarkable performance
the vesicle bilayer. Block copolymers — that is, poly- as nanochannels for water purification and desali-
mers comprising two or more monomers divided into nation. Much of their promise for desalination was
chemically distinct sequences — with a hydrophobic demonstrated by molecular dynamics simulations for a
polydimethylsiloxane (PDMS) block have been most single-walled (6,6)CNT (REF. 53) (FIG. 3a). The simulations
commonly used50,51. Using physical or chemical interac- showed the formation of a single-file water wire with
tions, the vesicles are then ruptured on a porous support ultra-fast water conduction (TABLE 2), which is largely
to form a planar bilayer, which becomes the selective attributed to the atomically smooth, non-polar inte-
layer of the membrane. Membranes that are formed rior. Simulations have also indicated that (6,6)CNTs,
using this strategy have often exhibited poor salt rejec- which have a diameter of 0.81 nm between opposing
tion52, probably because of defects, such as unruptured carbon centres and an inner (pore) diameter of 0.47 nm,
vesicles, although salt rejection of up to 98% has been can fully reject salt 54. Salt passage increases rapidly
reported50. Membranes have also suffered from poor with diameter: just 58% salt rejection is expected for
stability, largely because of the non-crosslinked lipid (8,8)CNTs with a 1.1-nm centre-to-centre diameter.

8 | MAY 2016 | VOLUME 1 www.nature.com/natrevmats


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

In two early experimental works, aligned-CNT-based Synthetically designed nanochannels offer a custom-
membranes were fabricated using chemical vapour izable alternative to CNTs. Although these nanochannels
deposition (CVD) methods55,56, yielding CNTs with lack the atomically smooth interior present in CNTs,
average inner diameters of 1.6 and 7 nm. Experiments their properties can be individually tailored, allowing far
validated the proposed ultra-fast water permeation: greater control over physical dimensions, performance,
water permeability was found to be three to five orders processing and alignment. Synthetic nano­channels
of magnitude higher than that predicted by continuum can be divided into two classes: single-molecule and
flow models. However, the smallest particle rejected supramolecular nanochannels (FIG. 3a).
was colloidal gold with a 2-nm diameter, indicating A recent single-molecule example is the peptide-
that these membranes fit in the tight UF range. To appended pillar[5]arene channel63. Pillar[5]arenes are
enhance CNT selectivity and potentially to allow desal- macrocycles (that is, cyclic macromolecules) comprising
ination, tip functionalization has been explored using five chemically linked hydroquinone groups, in which the
ligands such as biotin57, aliphatic groups58, charged benzene rings are oriented orthogonally to the macro­
groups57–60 and zwitterions61. Although tip functionali- cycle plane64. In peptide-appended pillar[5]arene chan-
zation has proven somewhat effective, the approach is nels, the macrocycle, which somewhat resembles a CNT,
limited by a lack of control over ligand placement and defines the channel diameter, whereas the appended pep-
the permeability–selectivity properties of the chosen tide arms control the length of the channel. The channels
flexible ligand. have been experimentally observed to readily insert into
A more ideal approach for ultra-selective CNT and to self-align within lipid bilayers. They were also seen
membranes would be to directly use the rigid molec- to self-assemble into densely packed arrays. Although the
ular sieving properties of highly monodisperse, small channels were highly permeable to water (TABLE 2), they
single-walled CNTs that intrinsically reject salt, as seen did not reject ions and had a molecular weight cut-off
in molecular dynamics simulations54. Such a membrane of approximately 420 Da, making them unsuitable for
has yet to be experimentally demonstrated, mainly desalination. It may be possible to improve selectivity
because of the heterogeneous size (that is, diameter) through functionalization of the appended peptides but,
distribution that typically arises during CNT synthe- as discussed for CNTs, enhancing selectivity through the
sis. Recent advances in monodisperse CNT synthesis use of smaller-diameter central constriction would be the
may overcome this limitation. For example, using a flat ideal solution.
platinum catalyst and a tailored precursor for CVD, sin- Supramolecular synthetic nanochannels comprise
gle-walled (6,6)CNTs with lengths greater than 100 nm monomer units that self-assemble to form a water-
were selectively synthesized62. Scaling up this process, if conducting channel. Two promising examples with sev-
it is possible, would result in CNTs that are well suited eral similarities are modified cyclic peptide nanotubes65
for desalination. and m‑phenylene ethynylene nanotubes66. Both types of

Table 2 | Transport properties for desalination-focused discrete water channels


Channel type Average single- Pore density Areal pore Measurement Demonstrated
channel water to match high- density method rejection
permeability water-flux TFC to match
coefficient (channels cm−2) TFC
at 25 °C
(molecules s−1 Pa−1)
Bacterial 5–36‡ (0.7–5.2) × 1012 7–47% Vesicle size All solutes
aquaporin Z43,127 change and
(~4 nm in length)* voltage clamp
(6,6)CNTs54,56 110§ 2.4 × 1011 0.25% Molecular NaCl (by simulation)
(1.4 nm in length) dynamics 2-nm colloidal gold
(by experiment)||
Peptide-appended 2.4‡ 1.1 × 1013 32% Vesicle size >420 Da
pillar[5]arenes63 change
(~4 nm in length)*
Modified cyclic 11¶ 2.4 × 1012 8% Molecular Not tested
peptide nanotubes128 dynamics
(2 nm in length)
m‑Phenylene 0.34# 7.9 × 1013 630% Vesicle size KCl and NaCl
ethynylene change
nanotubes66
(~4 nm in length)*
CNT, carbon nanotube; TFC, thin-film composite. *Approximate thickness of a lipid bilayer. ‡Permeabilities adjusted for
temperature using measured Arrhenius activation energies. §Simulated at 27 °C. ||For an aligned double-walled carbon nanotube
membrane with a 1.6-nm average pore diameter and with rejection measured using dead-end filtration. ¶For methylated cyclic
peptides to correspond to previous experimental work65. #Measured at 12 °C; value is unadjusted as Arrhenius activation energy
was not determined.

NATURE REVIEWS | MATERIALS VOLUME 1 | MAY 2016 | 9


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

nanotubes are formed using macrocyclic rings that have range of physicochemical properties and functionalities.
chemically modifiable interiors to control selectivity. The For both small molecules and block copolymers, self-
rings self-assemble by inter-ring hydrogen bonding and assembly occurs because of partitioning of chemically
π–π stacking to form solvent-stable, high-aspect-ratio distinct segments of the molecules.
nanotubes. Through molecular-level design, the two Although small molecules and block copolymers
nanotube types were synthesized to have very small pore can self-assemble into various structures (principally as
widths (0.47 nm for modified cyclic peptide nanotubes a function of composition), hexagonally packed cylin-
and 0.64 nm for m-phenylene ethynylene nanotubes). drical and bicontinuous cubic mesophases (FIG. 3) are
Both types also have promising transport behaviour, as the structures of interest for membrane separations,
shown in TABLE 2. In particular, m‑phenylene ethynylene in which the cylindrical or bicontinuous cubic micro­
nanotubes have been shown experimentally to reject domains act as the pores for size-selective transport.
sodium, potassium and chloride ions while maintain- These mesophases form over limited ranges of composi-
ing high water permeability in open (hydrated) chan- tion that are thermodynamically defined. For both block
nels. However, channels were open only 5% of the time, copolymers and small molecules, the characteristic pore
resulting in low average water permeability compared size is approximately set by the product of the size of the
with TFC membranes (TABLE 2). molecule and the volume fraction of its pore-forming
For all discrete nanochannels, the current challenge segment. Currently accessible pore diameters range
is predominantly the synthesis of channels with the from 5 to 50 nm (which lies in the UF range) for block
desired water and solute transport properties. An equally copolymers and 0.5 to 2 nm (which lies in the RO, FO
important technical challenge is developing scalable solu- and NF range) for small molecules. In most cases, the
tion-phase methods for the assembly and alignment of pores are formed by the selective removal (for example,
these channels in robust selective layers. Structural fea- etching by base hydrolysis of polylactide73) or shrinkage
tures that aid alignment are likely to be important and of material from the self-assembled structures, but for
will be a major factor in nanochannel design. In addition, small molecules, the porosity may be an intrinsic feature
much potential lies in using the self-assembling proper- of the system. Maximum porosity is usually 35–40%, as
ties of amphiphilic surfactants and block copolymers. it is defined by the small molecule or block copolymer
Indeed, this approach has already been applied with some used and the composition range over which the desired
success for CNTs67 and cyclic peptides68. Self-assembled structure forms.
materials can also be used directly to form nanoporous The ideal self-assembled selective layer has straight
membranes, as described in the next subsection. pores that fully span the layer. For cylindrical meso-
phases, this requires the alignment of the cylindrical
Self-assembled materials. Nanostructured polymer mem- pores perpendicular to the film surface as well as physical
branes based on self-assembled materials have been a continuity of the pores throughout the film. By contrast,
topic of considerable interest for more than two decades. bicontinuous cubic mesophases intrinsically provide
The underlying motivation is the potential to derive highly three-dimensionally percolated pathways throughout
selective membranes because of the narrow pore size the system. However, these pathways are inherently tor-
distribution that results from equilibrium self-assembly. tuous, which increases hydraulic resistance. In addition,
This contrasts sharply with the broad size distributions in bicontinuous cubic phases are considerably less acces-
conventional membranes that result from kinetically sible than cylinders owing to the limited composition
dictated fabrication processes such as phase inversion. range in which they are stable — if they are present
Two broad classes of materials have been used. at all — in most small molecule and block copolymer
The first class is reactive small molecules with a molar systems. Efforts to develop high-performance mem-
mass generally less than 1 kDa that self-assemble into branes based on self-assembled mesophases have there-
lyotropic or thermotropic liquid crystalline meso- fore focused heavily on molecular design to ensure the
phases69,70. Lyotropic liquid crystalline mesophases are stable formation of bicontinuous cubic structures and
structurally ordered systems that form principally as a on the development of scalable directed self-assembly
function of small molecule concentration in a solvent methods to allow vertical orientation of cylindrical pores
(typically water), whereas thermotropic liquid crystal- in thin films.
line mesophases mainly form as a function of tempera- Recent years have witnessed substantial progress
ture in the absence of a solvent. Before they can be used in molecular design for desalination-focused small
as membranes, liquid crystalline mesophases must be molecule membranes. Lyotropic liquid crystalline bi­­
poly­merized or crosslinked to form sufficiently robust continuous cubic mesophases based on ionic gemini
polymer films. The second class is block copolymers with surfactants are particularly promising 74–77. For example, a
a molar mass generally greater than 10 kDa. Block copol- crosslinkable phosphonium system with pores of 0.75 nm
ymers self-assemble into mesophases that are structurally in diameter demonstrated 95% NaCl rejection and more
similar to those in small molecule systems71,72. Physical than 99% rejection of sucrose and MgCl2. Although water
entanglement of the polymer chains and/or the glassy permeability was low because of the 40‑μm thickness of
nature of the materials can provide mechanical integ- the selective layer, the thickness-normalized water per-
rity without an explicit need for crosslinking. The ver- meability was comparable to that of commercial TFC RO
satility of modern synthetic polymer chemistry enables membranes, indicating that this material could be prac-
block copolymer membranes to be designed with a wide tically effective if thin selective layers of approximately

10 | MAY 2016 | VOLUME 1 www.nature.com/natrevmats


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

30 nm in thickness can be realized76. Several lyotropic To produce nanoporous graphene, defects are intro-
and thermotropic liquid crystalline phases (for example, duced into pristine, single-layer graphene by plasma
hexagonal and bicontinuous cubic) that use gallic acid etching or bombardment by ions or electrons97,98. Because
derivatives have also been developed78–83. For these mate- of the ultra-low thickness of the resulting pores, nano­
rials, the pore size was readily tuneable over a range of porous graphene was projected by molecular dynamics
0.8–2 nm by varying the type of counterion present in salt simulations to have 102–103 times greater water permea-
forms of the molecule84 and by using sacrificial species bility coefficients than current TFC RO membranes (the
that can be rinsed away after initially co‑assembling at effect of thickness is expressed in equation (2)). Complete
the centre of the cylindrical structures82,85,86. Gallic acid salt rejection was projected to be possible for hydroxy-
derivatives have also been used to form thin films (that lated pores that were 0.45 nm in diameter 99. Such perfor-
is, selective layers) on porous supports. Spin-coating 83 mance has recently been achieved for a 5‑μm‑diameter
and roll-casting 81 processes yielded film thicknesses of sample of nanoporous graphene formed by oxygen
50–100 nm and around 600 nm, respectively. In most of plasma etching of graphene grown through CVD97.
the work carried out thus far on small molecule mem- However, scaling up such a membrane is extremely chal-
branes, separation properties of the resulting membranes lenging, because it requires the inexpensive formation
have fallen into the NF range. Further advances in molec- of a large area of defect-free, single-layer graphene, as
ular design are crucial for fine-tuning both pore size and well as the scalable formation of uniformly sized nano-
permeability to address desalination applications more pores100–102. In addition, even if achieved, the usefulness
effectively. of nanoporous graphene for desalination applications is
Progress has also been made in directing self- doubtful. As discussed earlier, increased water perme-
assembly over large areas to achieve vertically aligned ability has a minimal effect above the current levels of
cylinders. Magnetic fields provide a readily scalable TFC desalination membranes, and unless the nanopores
means of achieving such orientation control and have can be even more finely tuned to also reject small neutral
been successfully used in both small molecule87 and solutes, nano­porous graphene will be at a disadvantage
block copolymer88 systems. Recent work has shown an compared with TFC membranes.
alternative field-free approach based on so-called soft Graphene-based framework membranes differ in
confinement, in which symmetric physical confine- morphology and water transport mechanism from
ment of a small molecule mesophase during ordering nanop­orous graphene and comprise a multi-layered
results in the spontaneous display of vertically oriented stack of finite-sized graphene sheets95,96,103 (FIG. 3a). They
microdomains89. The resulting films exhibit physically are formed as laminates through vacuum filtration or
continuous, vertically aligned 1-nm pores. layer-by-layer deposition of graphene oxide, which is
For block copolymer membranes, fabrication meth- inexpensively produced by oxidation of graphite and is
ods that combine self-assembly with a kinetic component stable when suspended in water. Graphene oxide frame-
have proven highly effective. For example, self-assembly works allow ultra-fast permeation of water, which is pos-
with non-solvent-induced phase separation (SNIPS) is tulated to be due to the slip flow along the atomically
a single, readily scalable process that can produce verti- smooth, unoxidized graphene channels; this is analogous
cally oriented pores in a thin selective layer at the mem- to the water flow within carbon nanotubes96. However,
brane surface with an integrally formed microporous the spacing between contiguous sheets increases from
underlying support layer 90,91 (FIG. 3b). As SNIPS com- 0.9 ± 0.1 nm in humid air to 1.3 ± 0.1 nm when immersed
bines self-assembly with kinetically limited phase inver- in water, resulting in rapid permeation of monovalent
sion, the resulting pore size distribution is broader than and divalent cations and the corresponding counteri-
that produced in equilibrium self-assembly processes. ons95. Graphene oxide frameworks still show a sharp
Nevertheless, these membranes display sharper mole­ size-based cut-off, but the cut-off is too large for desali-
cular weight cut-off characteristics and thus improved nation. In addition, graphene oxide laminates are gener-
selectivity relative to NIPS-formed UF membranes in the ally unstable in water, unless they are stabilized by trace
10–40-nm regime considered91–93. The combination of multivalent cations104.
block copolymer microphase separation and kinetically To potentially facilitate desalination, interlayer spac-
limited structure formation also features in the formation ing can be decreased and stability can be increased
of membranes by polymerization-induced microphase through chemical or thermal reduction to remove oxy-
separation; this has recently been used to fabricate a gen species. Full reduction decreases the interlayer spac-
bicontinuous cubic phase with 4-nm pore size94. ing to 0.36 nm, which is too small to allow water flow
between sheets and thus restricts water permeation to
Graphene. Owing to its atomically smooth and atom- defect-driven flow105. Although a reduced graphene oxide
ically thin nature, graphene has become a widely membrane was recently shown to allow water permeation
researched material for membrane separations, espe- with high salt retention when tested in FO, the defect-
cially for desalination95–97. Graphene-based desalination driven permeation in this approach limits water perme-
membranes have been proposed in two forms: nano­ ability 106. Another approach for decreasing interlayer
porous graphene and graphene-based frameworks. spacing and increasing stability is covalent crosslink-
Both forms are envisioned to serve as selective layers ing 103,107,108. However, a covalent crosslinking method has
and to operate as molecular sieves with size-based yet to be shown that facilitates salt rejection. Depending
exclusion of undesired solutes. on the success of these efforts, water-treatment-focused

NATURE REVIEWS | MATERIALS VOLUME 1 | MAY 2016 | 11


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

a b Surface modification that renders the membrane


Hydrophilic polymers Polymers with low surface energy surface highly hydrophilic substantially reduces organic
Poly(ethylene oxide) Poly(dimethylsiloxane) fouling and the adhesion of bacteria that eventually
n leads to biofouling. The most common hydrophilic
O
O O n
n Si materials for the preparation of fouling-resistant sur-
O
O O F n faces are polyethylene oxide (PEO)-based and zwitter-
O
O– F ionic polymers110,111 (FIG. 4a). PEO, which is also known
N+ S F as polyethylene glycol (PEG), binds water molecules by
O F F F
hydrogen bonding, whereas zwitterionic polymers —
Poly(sulfobetaine) Poly(hexafluoro-
butylmethacrylate) for example, poly(sulfobetaine) and poly(carboxybe-
Hydration layer taine) — bind water molecules more strongly through
electrostatic interactions111. The hydration layer of the
Hydrophilic side chain hydrophilic PEO and zwitterionic polymers provides a
Low-surface-energy block steric repulsive barrier that prevents the adsorption of
organic molecules and bacteria. Brush layers of these
Hydrophobic hydrophilic materials have been successful in prevent-
backbone
ing protein and bacteria adhesion in a wide range of
biomedical and industrial applications111,112.
Fouling resistance Fouling release Because of the adverse effects that fouling has
Water flow Water flow on membrane processes, the literature on develop-
ing fouling-resistant membranes is vast, and readers
are referred to recent comprehensive reviews113,114. A
few promising and widely applicable approaches for
designing fouling-resistant TFC and phase inversion
membranes are discussed below.

Anti-fouling TFC membranes by surface grafting.


Surface grafting of hydrophilic polymers is one of the
most promising and versatile anti-fouling strategies
for TFC membranes. One approach involves binding
PEG-based or other anti-fouling materials to the free
Nature Reviews | Materials
Figure 4 | Polymeric materials and self-segregation approach for fouling surface carboxyl or primary amine groups of the poly­
minimization in phase inversion membranes. In this approach, the phase inversion amide layer 113,114. This method, however, often yields
casting solution contains a hydrophobic polymer and a comb copolymer with a limited fouling resistance owing to non-uniform and
hydrophobic backbone and side chains that are at least partially hydrophilic. During incomplete surface coverage of anti-fouling materials
non-solvent-induced phase separation in water, the hydrophilic side chains
on the polyamide active layer, which is a direct result
self-segregate at the membrane/water interface, whereas the backbone anchors itself in
the membrane matrix. a | Side chains are entirely hydrophilic, comprising poly(ethylene of the limited number of accessible pendant carboxyl or
oxide) or zwitterionic polymers. The resulting hydration layer inhibits irreversible amine functional groups on the membrane surface115,116.
adsorption of organic and biological foulants. b | Side chains contain an additional In addition, the method is limited even when using a
hydrophobic block with low surface energy. The low-surface-energy block promotes more ideal substrate, as steric repulsion of pre-formed
foulant release in the presence of surface turbulence (for example, high cross-flows). polymers limits the achievable brush density 117. A
more promising method for grafting dense and uni-
form anti-fouling polymer brushes on TFC membranes
graphene oxide framework membranes may be effective involves the use of surface-initiated controlled radical
only for NF-range separations and not suitable for the polymerization techniques. Of these techniques, atom
more urgently needed desalination applications. transfer radical polymerization (ATRP) is particularly
versatile and robust because it can precisely control
Designing fouling-resistant membranes the density, length and architecture of the anti-fouling
As discussed earlier, phase inversion and TFC mem- brush layer 117. A recent study has successfully demon-
branes are prone to organic and biological fouling strated the use of ATRP to graft a dense brush layer of
because of their inherent surface chemical and physi- poly(sulfobetaine methacrylate), a zwitterionic polymer,
cal properties. Proper tailoring of the surface chemical to the polyamide surface of TFC RO membranes118. The
properties of membranes can substantially enhance their resulting membrane exhibited a very low water contact
fouling resistance. Past work on the adsorption of pro- angle (<10°), negligible adsorption of biomolecules (that
teins to various surfaces provides useful insights into the is, proteins), substantial reduction of foulant adhesion
necessary surface properties that minimize adsorption of forces to the membrane surface and a negligible change
organic foulants and microorganisms to membrane sur- in the permselectivity of the membrane.
faces. Specifically, it was reported that surfaces with high Surface grafting often results in reduced water per-
resistance to the adsorption of proteins are hydrophilic, meability and thus should be optimized to balance the
contain hydrogen bond acceptors, are electroneutral and gain in fouling resistance and reduction in water perme-
do not contain hydrogen bond donors109. ability 114. ATRP is advantageous in this regard because

12 | MAY 2016 | VOLUME 1 www.nature.com/natrevmats


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

of the control it allows over the density and thickness of Conclusion and outlook
the brush layer. However, methods for uniform immo- In this Review, we identified the lack of control over
bilization of the initiator molecules on the rough poly- water and solute permeabilities as the key limitation
amide surface of TFC membranes need to be developed hampering membrane performance. This crucial con-
to ensure dense and uniform brush layers. Long-term straint is attributed to current membrane fabrication
anti-fouling performance and chemical stability of the methods that still remain more of an art than a science.
grafted brush layer, particularly under chemical clean- Several novel approaches for molecular-level design of
ing operations, are also important considerations for TFC isoporous films that can function as selective layers have
surface modification. Furthermore, although model sur- been discussed. With these next-generation membranes,
faces and small membrane coupons have been used in the solute rejection in both UF and desalination (that is, RO
laboratory, upscaling to membrane modules and reduc- and FO) membranes will be based on a sieving or size-
ing the cost of membrane fabrication remain challenging. exclusion mechanism. Salt retention by size exclu-
sion will represent a paradigm shift for desalination
Anti-fouling porous membranes by surface segrega- membranes, as solute and water transport in current
tion. When applied to UF membranes, surface grafting TFC membranes is governed by a solution–diffusion
leaves the interior pore walls unmodified, resulting in mechanism.
adsorption of organic foulants to the pore walls and Substantial progress has been made in the fabri­
irreversible fouling 2,119. A more robust surface modifi- cation of isoporous UF separation membranes by some
cation approach for phase inversion membranes is in situ of the molecular-level design approaches discussed in
surface segregation119,120. In this approach, an amphi- this Review. Fabrication of block copolymer UF mem-
philic copolymer comprising a hydrophobic backbone branes with uniform, high-density pores is now possi-
and hydrophilic side chains is blended with the hydro­ ble using scalable approaches, such as SNIPS. However,
phobic base polymer in the casting solution. During highly selective, isoporous UF membranes may offer
precipitation of the base polymer in the coagulation only marginal improvements over conventional UF
bath, the hydrophobic backbone of the amphiphilic co­­ membranes for water purification applications, as high
polymers ensures strong anchoring within the polymer selectivity is not an important objective in the current
bulk matrix, whereas the hydrophilic segments segregate applications of UF as pre-treatment before RO or as
to all polymer/water interfaces, creating a surface with a standalone water purification system. Isoporous UF
an exposed hydrophilic brush layer that imparts fouling membranes are likely to have an important role in
resistance to the entire membrane (FIG. 4a). This one-step future applications that involve recovery of substances
membrane fabrication method has been successfully from various wastewaters, as well as in other industrial
demonstrated with the use of PAN-graft-PEO — which is separations.
an amphiphilic comb copolymer with a water-insoluble Next-generation, highly selective desalination
PAN backbone and hydrophilic PEO side chains — as mem­branes will represent an important advancement.
an additive in the fabrication of a PAN UF membrane Enhanced membrane selectivity will have an important
that exhibits complete resistance to irreversible fouling role in improving water quality and eliminating the need
by three types of organic foulants2,120. for additional separation stages — for example, for low-
The in situ surface segregation method for fabricating ering boric acid concentration in seawater desalination
highly hydrophilic UF membranes was recently extended to acceptable levels for agriculture — thereby reducing
to the fabrication of membranes with surface proper- energy usage and the cost of desalination. The molecular-
ties that provide both resistance to irreversible foulant level design approaches for desalination membranes
adsorption and enhanced foulant release121. The method discussed in this Review represent an active area of
involves the use of a ternary amphiphilic block co­­ research. Despite these major efforts, however, it is still
polymer that comprises a hydrophobic anchoring block, not possible to consistently fabricate even small-area
a hydrophilic fouling-resistant block and a nonpolar, membranes with greater water permeability and selec-
low-surface-energy fouling-release block as an additive tivity than current state‑of‑the-art TFC membranes. A
during the phase inversion process. If the amphiphilic particular challenge for several of these materials will
copolymer is properly designed and the correct solvents be the formation of thin, defect-free selective layers
are chosen, the hydrophilic and fouling-release segments supported by microporous layers. After the concepts of
are forced to segregate to the polymer/water interface. these novel membranes are successfully demonstrated
This method has been successfully demonstrated for the at the laboratory scale, upscaling the small-area mem-
fabrication of cellulose acetate121, PVDF122, and PES123 branes to low-cost, industrial-scale modules will be an
UF membranes with surfaces containing hydrophilic additional challenge.
PEO-based blocks for fouling resistance and PDMS Fouling control in next-generation membranes, in
or fluorinated-methacrylate-based polymer blocks for addition to current membranes, will remain a major
fouling release (FIG. 4b). The membranes exhibit excellent challenge, because novel materials for high separation
performance in inhibiting the nonspecific adsorption of performance are often not compatible with fouling
organic foulants as well as in preventing the accumula- resistance. Surface grafting of fouling-resistant poly-
tion and coalescence of organic and oily foulants on the mer brushes — such as PEO and zwitterionic polymers
membrane surface — when surface turbulence is present — will probably remain the best approach for impart-
— owing to the fouling release property. ing fouling resistance for next-generation, isoporous

NATURE REVIEWS | MATERIALS VOLUME 1 | MAY 2016 | 13


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

membranes. For ‘tight’ UF membranes, surface grafting membranes are likely to overcome another limitation
will not impart fouling resistance to the interior pore of current TFC membranes: the high sensitivity of the
walls; hence, proper choice of new polymer chemistries polyamide selective layer to oxidants such as chlorine
for membrane fabrication with fouling and chemical and ozone. This will allow better control of biofouling,
resistance will be crucial. Next-generation desalination which is the Achilles heel of membrane desalination.

1. Elimelech, M. The global challenge for adequate and 24. Geise, G. M., Paul, D. R. & Freeman, B. D. 45. Hub, J. S. & de Groot, B. L. Mechanism of selectivity in
safe water. J. Water Supply Res. Technol. 55, 3–10 Fundamental water and salt transport properties of aquaporins and aquaglyceroporins. Proc. Natl Acad.
(2006). polymeric materials. Prog. Polym. Sci. 39, 1–42 Sci. USA 105, 1198–1203 (2008).
2. Shannon, M. A. et al. Science and technology for (2014). 46. Sui, H., Han, B. G., Lee, J. K., Walian, P. & Jap, B. K.
water purification in the coming decades. Nature 452, 25. Geise, G. M., Park, H. B., Sagle, A. C., Freeman, B. D. Structural basis of water-specific transport through the
301–310 (2008). & McGrath, J. E. Water permeability and water/salt AQP1 water channel. Nature 414, 872–878 (2001).
3. Schwarzenbach, R. P. et al. The challenge of selectivity tradeoff in polymers for desalination. In this X‑ray crystallography study, a high-resolution
micropollutants in aquatic systems. Science 313, J. Membr. Sci. 369, 130–138 (2011). structure of aquaporin 1 was attained, which
1072–1077 (2006). A water–salt selectivity trade-off for desalination elucidated the basis for its selectivity for water.
4. Kolpin, D. W. et al. Pharmaceuticals, hormones, and using polymeric membranes is proposed for the 47. Murata, K. et al. Structural determinants of water
other organic wastewater contaminants in U.S. first time in this study. permeation through aquaporin‑1. Nature 407,
streams, 1999–2000: a national reconnaissance. 26. Robeson, L. M. The upper bound revisited. J. Membr. 599–605 (2000).
Environ. Sci. Technol. 36, 1202–1211 (2002). Sci. 320, 390–400 (2008). 48. Savage, D. F., O’Connell, J. D., Miercke, L. J. W., Finer-
5. Tang, J. Y., Busetti, F., Charrois, J. W. & Escher, B. I. 27. Yip, N. Y. & Elimelech, M. Performance limiting effects Moore, J. & Stroud, R. M. Structural context shapes
Which chemicals drive biological effects in wastewater in power generation from salinity gradients by the aquaporin selectivity filter. Proc. Natl Acad. Sci.
and recycled water? Water Res. 60, 289–299 pressure retarded osmosis. Environ. Sci. Technol. 45, USA 107, 17164–17169 (2010).
(2014). 10273–10282 (2011). 49. Wang, M. et al. Layer‑by‑layer assembly of aquaporin
6. Vidic, R. D., Brantley, S. L., Vandenbossche, J. M., 28. Iwahashi, H. et al. Advanced RO system for high Z‑incorporated biomimetic membranes for water
Yoxtheimer, D. & Abad, J. D. Impact of shale gas temperature and high concentration seawater purification. Environ. Sci. Technol. 49, 3761–3768
development on regional water quality. Science 340, desalination at the Arabian Gulf. IDA World Congress (2015).
1235009 (2013). (San Diego, 2015). 50. Wang, H. et al. Highly permeable and selective pore-
7. Shaffer, D. L. et al. Desalination and reuse of high- 29. Cohen-Tanugi, D., McGovern, R. K., Dave, S. H., spanning biomimetic membrane embedded with
salinity shale gas produced water: drivers, Lienhard, J. H. & Grossman, J. C. Quantifying the aquaporin Z. Small 8, 1185–1190 (2012).
technologies, and future directions. Environ. Sci. potential of ultra-permeable membranes for water 51. Kumar, M., Grzelakowski, M., Zilles, J., Clark, M. &
Technol. 47, 9569–9583 (2013). desalination. Energy Environ. Sci. 7, 1134–1141 (2014). Meier, W. Highly permeable polymeric membranes
8. Gregory, K. B., Vidic, R. D. & Dzombak, D. A. Water 30. Deshmukh, A., Yip, N. Y., Lin, S. & Elimelech, M. based on the incorporation of the functional water
management challenges associated with the Desalination by forward osmosis: identifying channel protein aquaporin Z. Proc. Natl Acad. Sci.
production of shale gas by hydraulic fracturing. performance limiting parameters through module- USA 104, 20719–20724 (2007).
Elements 7, 181–186 (2011). scale modeling. J. Membr. Sci. 491, 159–167 52. Zhong, P. S., Chung, T.‑S., Jeyaseelan, K. & Armugam, A.
9. Bond, R. & Veerapaneni, S. Zero liquid discharge for (2015). Aquaporin-embedded biomimetic membranes for
inland desalination (Awwa Research Foundation, 31. Singh, R. Production of high-purity water by membrane nanofiltration. J. Membr. Sci. 407–408, 27–33 (2012).
2007). processes. Desalin. Water Treat. 3, 99–110 (2012). 53. Hummer, G., Rasaiah, J. C. & Noworyta, J. P. Water
10. Mickley, M. Survey of high-recovery and zero liquid 32. Miyashita, Y., Park, S.-H., Hyung, H., Huang, C.-H. & conduction through the hydrophobic channel of a
discharge technologies for water utilities (WateReuse Kim, J.-H. Removal of N‑nitrosamines and their carbon nanotube. Nature 414, 188–190 (2001).
Foundation, 2008). precursors by nanofiltration and reverse osmosis This study provides the first demonstration, using
11. van Loosdrecht, M. C. M. & Brdjanovic, D. membranes. J. Environ. Eng. 135, 788–795 (2009). molecular dynamics, of ultra-fast water permeation
Anticipating the next century of wastewater 33. Ozaki, H. & Li, H. Rejection of organic compounds by through the interior of carbon nanotubes.
treatment. Science 344, 1452–1453 (2014). ultra-low pressure reverse osmosis membrane. Water 54. Corry, B. Designing carbon nanotube membranes for
12. Grant, S. B. et al. Taking the “waste” out of Res. 36, 123–130 (2002). efficient water desalination. J. Phys. Chem. B 112,
“wastewater” for human water security and ecosystem 34. Fritzmann, C., Löwenberg, J., Wintgens, T. & Melin, T. 1427–1434 (2008).
sustainability. Science 337, 681–686 (2012). State‑of‑the-art of reverse osmosis desalination. 55. Majumder, M., Chopra, N., Andrews, R. & Hinds, B. J.
13. Sales, C. M. & Lee, P. K. Resource recovery from Desalination 216, 1–76 (2007). Nanoscale hydrodynamics: enhanced flow in carbon
wastewater: application of meta-omics to phosphorus 35. Le‑Clech, P., Chen, V. & Fane, T. A. G. Fouling in nanotubes. Nature 438, 44 (2005).
and carbon management. Curr. Opin. Biotechnol. 33, membrane bioreactors used in wastewater treatment. 56. Holt, J. K. et al. Fast mass transport through
260–267 (2015). J. Membr. Sci. 284, 17–53 (2006). sub‑2‑nanometer carbon nanotubes. Science 312,
14. Wang, X. et al. Probabilistic evaluation of integrating 36. Cheryan, M. Ultrafiltration and Microfiltration 1034–1037 (2006).
resource recovery into wastewater treatment to Handbook 31–70 (CRC Press, 1998). 57. Hinds, B. J. et al. Aligned multiwalled carbon
improve environmental sustainability. Proc. Natl Acad. 37. Matthiasson, E. The role of macromolecular nanotube membranes. Science 303, 62–65 (2004).
Sci. USA 112, 1630–1635 (2015). adsorption in fouling of ultrafiltration membranes. 58. Majumder, M., Chopra, N. & Hinds, B. J. Effect of tip
15. Elimelech, M. & Phillip, W. A. The future of seawater J. Membr. Sci. 16, 23–36 (1983). functionalization on transport through vertically
desalination: energy, technology, and the environment. 38. Belfort, G., Davis, R. H. & Zydney, A. L. The behavior of oriented carbon nanotube membranes. J. Am. Chem.
Science 333, 712–717 (2011). suspensions and macromolecular solutions in crossflow Soc. 127, 9062–9070 (2005).
16. Semiat, R. Energy issues in desalination processes. microfiltration. J. Membr. Sci. 96, 1–58 (1994). 59. Fornasiero, F. et al. Ion exclusion by sub‑2‑nm carbon
Environ. Sci. Technol. 42, 8193–8201 (2008). 39. Elimelech, M., Zhu, X., Childress, A. E. & Hong, S. Role nanotube pores. Proc. Natl Acad. Sci. USA 105,
17. Baker, R. W. Membrane Technology and Applications of membrane surface morphology in colloidal fouling 17250–17255 (2008).
(John Wiley & Sons, 2012). of cellulose acetate and composite aromatic 60. Corry, B. Water and ion transport through functionalised
18. Gin, D. L. & Noble, R. D. Designing the next polyamide reverse osmosis membranes. J. Membr. carbon nanotubes: implications for desalination
generation of chemical separation membranes. Sci. 127, 101–109 (1997). technology. Energy Environ. Sci. 4, 751–759 (2011).
Science 332, 674–676 (2011). The roughness of polyamide thin-film composite 61. Chan, W. F. et al. Zwitterion functionalized carbon
19. Tang, C. Y., Kwon, Y.‑N. & Leckie, J. O. Effect of membrane surfaces is shown in this study to nanotube/polyamide nanocomposite membranes for
membrane chemistry and coating layer on exacerbate fouling. water desalination. ACS Nano 7, 5308–5319 (2013).
physiochemical properties of thin film composite 40. Herzberg, M. & Elimelech, M. Biofouling of reverse 62. Sanchez-Valencia, J. R. et al. Controlled synthesis of
polyamide RO and NF membranes. Desalination 242, osmosis membranes: role of biofilm-enhanced osmotic single-chirality carbon nanotubes. Nature 512,
168–182 (2009). pressure. J. Membr. Sci. 295, 11–20 (2007). 61–64 (2014).
20. Lu, X. et al. Elements provide a clue: nanoscale 41. Li, Q. & Elimelech, M. Organic fouling and chemical 63. Shen, Y. X. et al. Highly permeable artificial water
characterization of thin-film composite polyamide cleaning of nanofiltration membranes: measurements channels that can self-assemble into two-dimensional
membranes. ACS Appl. Mater. Interfaces 7, and mechanisms. Environ. Sci. Technol. 38, arrays. Proc. Natl Acad. Sci. USA 112, 9810–9815
16917–16922 (2015). 4683–4693 (2004). (2015).
21. Karan, S., Jiang, Z. & Livingston, A. G. Sub‑10 nm 42. Mi, B. & Elimelech, M. Gypsum scaling and cleaning in 64. Hu, X. B., Chen, Z., Tang, G., Hou, J. L. & Li, Z. T. Single-
polyamide nanofilms with ultrafast solvent transport forward osmosis: measurements and mechanisms. molecular artificial transmembrane water channels.
for molecular separation. Science 348, 1347–1351 Environ. Sci. Technol. 44, 2022–2028 (2010). J. Am. Chem. Soc. 134, 8384–8387 (2012).
(2015). 43. Borgnia, M. J., Kozono, D., Calamita, G., 65. Hourani, R. et al. Processable cyclic peptide
22. Kwak, S. Y., Jung, S. G. & Kim, S. H. Structure-motion- Maloney, P. C. & Agre, P. Functional reconstitution and nanotubes with tunable interiors. J. Am. Chem. Soc.
performance relationship of flux-enhanced reverse characterization of AqpZ, the E. coli water channel 133, 15296–15299 (2011).
osmosis (RO) membranes composed of aromatic protein. J. Mol. Biol. 291, 1169–1179 (1999). 66. Zhou, X. et al. Self-assembling subnanometer pores
polyamide thin films. Environ. Sci. Technol. 35, 44. Beitz, E., Wu, B., Holm, L. M., Schultz, J. E. & with unusual mass-transport properties. Nat.
4334–4340 (2001). Zeuthen, T. Point mutations in the aromatic/arginine Commun. 3, 949 (2012).
23. Mehta, A. & Zydney, A. L. Permeability and selectivity region in aquaporin 1 allow passage of urea, glycerol, Highly selective artificial water channels that
analysis for ultrafiltration membranes. J. Membr. Sci. ammonia, and protons. Proc. Natl Acad. Sci. USA operate by rigid molecular sieving are reported in
249, 245–249 (2005). 103, 269–274 (2006). this study.

14 | MAY 2016 | VOLUME 1 www.nature.com/natrevmats


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

67. Mauter, M. S., Elimelech, M. & Osuji, C. O. magnetic field directed self-assembly. ACS Nano 8, 111. Jiang, S. & Cao, Z. Ultralow-fouling, functionalizable,
Nanocomposites of vertically aligned single-walled 11977–11986 (2014). and hydrolyzable zwitterionic materials and their
carbon nanotubes by magnetic alignment and 88. Gopinadhan, M. et al. Thermally switchable aligned derivatives for biological applications. Adv. Mater. 22,
polymerization of a lyotropic precursor. ACS Nano 4, nanopores by magnetic-field directed self-assembly of 920–932 (2010).
6651–6658 (2010). block copolymers. Adv. Mater. 26, 5148–5154 (2014). 112. Callow, J. A. & Callow, M. E. Trends in the development
68. Xu, T. et al. Subnanometer porous thin films by the 89. Feng, X. et al. Thin polymer films with continuous of environmentally friendly fouling-resistant marine
co‑assembly of nanotube subunits and block vertically aligned 1 nm pores fabricated by soft coatings. Nat. Commun. 2, 244 (2011).
copolymers. ACS Nano 5, 1376–1384 (2011). confinement. ACS Nano 10, 150–158 (2015). 113. Kang, G. D. & Cao, Y. M. Development of antifouling
69. Gin, D. L., Gu, W., Pindzola, B. A. & Zhou, W. J. 90. Peinemann, K. V., Abetz, V. & Simon, P. F. Asymmetric reverse osmosis membranes for water treatment: a
Polymerized lyotropic liquid crystal assemblies for superstructure formed in a block copolymer via phase review. Water Res. 46, 584–600 (2012).
materials applications. Acc. Chem. Res. 34, 973–980 separation. Nat. Mater. 6, 992–996 (2007). 114. Rana, D. & Matsuura, T. Surface modifications for
(2001). In this study, block copolymer self-assembly and antifouling membranes. Chem. Rev. 110, 2448–2471
70. Gin, D. L., Bara, J. E., Noble, R. D. & Elliott, B. J. phase inversion are combined for the first time to (2010).
Polymerized lyotropic liquid crystal assemblies for readily form porous membranes with a low pore 115. Shaffer, D. L., Jaramillo, H., Romero-Vargas Castrillón, S.,
membrane applications. Macromol. Rapid. Commun. size polydispersity. Lu, X. & Elimelech, M. Post-fabrication modification of
29, 367–389 (2008). 91. Phillip, W. A. et al. Tuning structure and properties of forward osmosis membranes with a poly(ethylene glycol)
71. Zhang, Y., Sargent, J. L., Boudouris, B. W. & graded triblock terpolymer-based mesoporous and block copolymer for improved organic fouling resistance.
Phillip, W. A. Nanoporous membranes generated from hybrid films. Nano Lett. 11, 2892–2900 (2011). J. Membr. Sci. 490, 209–219 (2015).
self-assembled block polymer precursors: quo vadis? 92. Gu, Y. & Wiesner, U. Tailoring pore size of graded 116. Van Wagner, E. M., Sagle, A. C., Sharma, M. M.,
J. Appl. Polym. Sci. 132, 41683 (2015). mesoporous block copolymer membranes: moving La, Y.‑H. & Freeman, B. D. Surface modification of
72. Jackson, E. A. & Hillmyer, M. A. Nanoporous from ultrafiltration toward nanofiltration. commercial polyamide desalination membranes using
membranes derived from block copolymers: from drug Macromolecules 48, 6153–6159 (2015). poly(ethylene glycol) diglycidyl ether to enhance
delivery to water filtration. ACS Nano 4, 3548–3553 93. Clodt, J. I. et al. Performance study of isoporous membrane fouling resistance. J. Membr. Sci. 367,
(2010). membranes with tailored pore sizes. J. Membr. Sci. 273–287 (2011).
73. Zalusky, A. S., Olayo-Valles, R., Wolf, J. H. & 495, 334–340 (2015). 117. Barbey, R. et al. Polymer brushes via surface-initiated
Hillmyer, M. A. Ordered nanoporous polymers from 94. Seo, M. & Hillmyer, M. A. Reticulated nanoporous controlled radical polymerization: synthesis,
polystyrene–polylactide block copolymers. J. Am. polymers by controlled polymerization-induced characterization, properties, and applications. Chem.
Chem. Soc. 124, 12761–12773 (2002). microphase separation. Science 336, 1422–1425 Rev. 109, 5437–5527 (2009).
74. Sorenson, G. P., Coppage, K. L. & Mahanthappa, M. K. (2012). 118. Ye, G., Lee, J., Perreault, F. & Elimelech, M. Controlled
Unusually stable aqueous lyotropic gyroid phases from 95. Joshi, R. K. et al. Precise and ultrafast molecular architecture of dual-functional block copolymer brushes
gemini dicarboxylate surfactants. J. Am. Chem. Soc. sieving through graphene oxide membranes. Science on thin-film composite membranes for integrated
133, 14928–14931 (2011). 343, 752–754 (2014). “defending” and “attacking” strategies against biofouling.
75. Hatakeyama, E. S., Wiesenauer, B. R., Gabriel, C. J., This is an experimental demonstration of molecular ACS Appl. Mater. Interfaces 7, 23069–23079 (2015).
Noble, R. D. & Gin, D. L. Nanoporous, bicontinuous sieving through graphene oxide laminates, albeit 119. Kang, S., Asatekin, A., Mayes, A. M. & Elimelech, M.
cubic lyotropic liquid crystal networks via with a size cut-off that is too large for desalination. Protein antifouling mechanisms of PAN UF
polymerizable gemini ammonium surfactants. Chem. 96. Nair, R. R., Wu, H. A., Jayaram, P. N., Grigorieva, I. V. membranes incorporating PAN‑g‑PEO additive.
Mater. 22, 4525–4527 (2010). & Geim, A. K. Unimpeded permeation of water J. Membr. Sci. 296, 42–50 (2007).
76. Zhou, M. et al. New type of membrane material for through helium-leak-tight graphene-based 120. Asatekin, A., Kang, S., Elimelech, M. & Mayes, A. M.
water desalination based on a cross-linked membranes. Science 335, 442–444 (2012). Anti-fouling ultrafiltration membranes containing
bicontinuous cubic lyotropic liquid crystal assembly. 97. Surwade, S. P. et al. Water desalination using polyacrylonitrile-graft-poly(ethylene oxide) comb
J. Am. Chem. Soc. 129, 9574–9575 (2007). nanoporous single-layer graphene. Nat. Nanotechnol. copolymer additives. J. Membr. Sci. 298, 136–146
77. Pindzola, B. A., Jin, J. & Gin, D. L. Cross-linked normal 10, 459–464 (2015). (2007).
hexagonal and bicontinuous cubic assemblies via 98. Russo, C. J. & Golovchenko, J. A. Atom‑by‑atom The in situ surface segregation approach is used in
polymerizable gemini amphiphiles. J. Am. Chem. Soc. nucleation and growth of graphene nanopores. Proc. this study to readily fabricate fouling-resistant
125, 2940–2949 (2003). Natl Acad. Sci. USA 109, 5953–5957 (2012). polyacrylonitrile UF membranes.
78. Soberats, B. et al. 3D Anhydrous proton-transporting 99. Cohen-Tanugi, D. & Grossman, J. C. Water 121. Chen, W. et al. Engineering a robust, versatile
nanochannels formed by self-assembly of liquid desalination across nanoporous graphene. Nano Lett. amphiphilic membrane surface through forced surface
crystals composed of a sulfobetaine and a sulfonic 12, 3602–3608 (2012). segregation for ultralow flux-decline. Adv. Funct.
acid. J. Am. Chem. Soc. 135, 15286–15289 (2013). 100. O’Hern, S. C. et al. Selective ionic transport through Mater. 21, 191–198 (2011).
79. Kerr, R. L., Miller, S. A., Shoemaker, R. K., Elliott, B. J. tunable subnanometer pores in single-layer graphene 122. Chen, W. et al. Efficient wastewater treatment by
& Gin, D. L. New type of Li ion conductor with 3D membranes. Nano Lett. 14, 1234–1241 (2014). membranes through constructing tunable antifouling
interconnected nanopores via polymerization of a 101. O’Hern, S. C. et al. Nanofiltration across defect-sealed membrane surfaces. Environ. Sci. Technol. 45,
liquid organic electrolyte-filled lyotropic liquid-crystal nanoporous monolayer graphene. Nano Lett. 15, 6545–6552 (2011).
assembly. J. Am. Chem. Soc. 131, 15972–15973 3254–3260 (2015). 123. Zhao, X. et al. Engineering amphiphilic membrane
(2009). 102. O’Hern, S. C. et al. Selective molecular transport surfaces based on PEO and PDMS segments for
80. Smith, R. C., Fischer, W. M. & Gin, D. L. Ordered through intrinsic defects in a single layer of CVD improved antifouling performances. J. Membr. Sci.
poly(p‑phenylenevinylene) matrix nanocomposites via graphene. ACS Nano 6, 10130–10138 (2012). 450, 111–123 (2014).
lyotropic liquid-crystalline monomers. J. Am. Chem. 103. Mi, B. Graphene oxide membranes for ionic and 124. Shaffer, D. L., Werber, J. R., Jaramillo, H., Lin, S. &
Soc. 119, 4092–4093 (1997). molecular sieving. Science 343, 740–742 (2014). Elimelech, M. Forward osmosis: where are we now?
This is an early demonstration of the formation of 104. Yeh, C. N., Raidongia, K., Shao, J., Yang, Q. H. & Desalination 356, 271–284 (2015).
a nanoporous polymeric structure using surfactants Huang, J. On the origin of the stability of graphene 125. Tiraferri, A., Yip, N. Y., Phillip, W. A., Schiffman, J. D.
based on polymerizable gallic acid. oxide membranes in water. Nat. Chem. 7, 166–170 & Elimelech, M. Relating performance of thin-film
81. Zhou, M., Kidd, T. J., Noble, R. D. & Gin, D. L. (2014). composite forward osmosis membranes to support
Supported lyotropic liquid-crystal polymer 105. Su, Y. et al. Impermeable barrier films and protective layer formation and structure. J. Membr. Sci. 367,
membranes: promising materials for molecular-size- coatings based on reduced graphene oxide. Nat. 340–352 (2011).
selective aqueous nanofiltration. Adv. Mater. 17, Commun. 5, 4843 (2014). 126. Lu, X., Arias Chavez, L. H., Romero-Vargas
1850–1853 (2005). 106. Liu, H., Wang, H. & Zhang, X. Facile fabrication of Castrillón, S., Ma, J. & Elimelech, M. Influence of
82. Broer, D. J., Bastiaansen, C. M., Debije, M. G. & freestanding ultrathin reduced graphene oxide active layer and support layer surface structures on
Schenning, A. P. Functional organic materials based membranes for water purification. Adv. Mater. 27, organic fouling propensity of thin-film composite
on polymerized liquid-crystal monomers: 249–254 (2015). forward osmosis membranes. Environ. Sci. Technol.
supramolecular hydrogen-bonded systems. Angew. 107. Hung, W.‑S. et al. Cross-linking with diamine 49, 1436–1444 (2015).
Chem. Int. Ed. Engl. 51, 7102–7109 (2012). monomers to prepare composite graphene oxide- 127. Pohl, P., Saparov, S. M., Borgnia, M. J. & Agre, P.
83. Henmi, M. et al. Self-organized liquid-crystalline framework membranes with varying d‑spacing. Chem. Highly selective water channel activity measured by
nanostructured membranes for water treatment: Mater. 26, 2983–2990 (2014). voltage clamp: analysis of planar lipid bilayers
selective permeation of ions. Adv. Mater. 24, 108. Zhang, Y., Zhang, S. & Chung, T. S. Nanometric reconstituted with purified AqpZ. Proc. Natl Acad.
2238–2241 (2012). graphene oxide framework membranes with enhanced Sci. USA 98, 9624–9629 (2001).
84. Deng, H., Gin, D. L. & Smith, R. C. Polymerizable heavy metal removal via nanofiltration. Environ. Sci. 128. Ruiz, L., Wu, Y. & Keten, S. Tailoring the water
lyotropic liquid crystals containing transition-metal and Technol. 49, 10235–10242 (2015). structure and transport in nanotubes with tunable
lanthanide ions: architectural control and introduction 109. Ostuni, E., Chapman, R. G., Holmlin, R. E., Takayama, S. interiors. Nanoscale 7, 121–132 (2015).
of new properties into nanostructured polymers. J. Am. & Whitesides, G. M. A survey of structure–property
Chem. Soc. 120, 3522–3523 (1998). relationships of surfaces that resist the adsorption of Acknowledgements
85. Lee, H.‑K. et al. Synthesis of a nanoporous polymer protein. Langmuir 17, 5605–5620 (2001). The authors acknowledge support received from the US
with hexagonal channels from supramolecular discotic Through adsorption measurements of proteins on National Science Foundation (NSF) under award numbers
liquid crystals. Angew. Chem. Int. Ed. Engl. 40, model surfaces with defined functional groups, this CBET-1437630 and CMMI-1246804 and through the NSF
2669–2671 (2001). study describes general characteristics of Nanosystems Engineering Research Center for
86. Ishida, Y. et al. Guest-responsive covalent frameworks non-fouling surfaces. Nanotechnology-Enabled Water Treatment (ERC‑1449500).
by the cross-linking of liquid-crystalline salts: tuning of 110. Banerjee, I., Pangule, R. C. & Kane, R. S. Antifouling The authors also acknowledge the NSF Graduate Research
lattice flexibility by the design of polymerizable units. coatings: recent developments in the design of Fellowship awarded to J.R.W.
Chemistry 17, 14752–14762 (2011). surfaces that prevent fouling by proteins, bacteria,
87. Feng, X. et al. Scalable fabrication of polymer and marine organisms. Adv. Mater. 23, 690–718 Competing interests statement
membranes with vertically aligned 1 nm pores by (2011). The authors declare no competing interests.

NATURE REVIEWS | MATERIALS VOLUME 1 | MAY 2016 | 15


©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
O N L I N E O N LY

Figure permissions
Figure 1
The upper panels of part b are adapted with permission from REF. 125,
Elsevier. The upper panel of part d is adapted with permission from
REF. 126, American Chemical Society. The lower panels of part d are
adapted with permission from REF. 20, American Chemical

Figure 3
Scanning electron microscopy image in panel b is from REF. 90, Nature
Publishing Group.

Subject Terms
Physical sciences / Materials science / Nanoscale materials [URI
/639/301/357]
Physical sciences / Engineering / Chemical engineering
[URI /639/166/898]
Physical sciences / Chemistry / Materials chemistry / Soft materials
/ Polymers
[URI /639/638/298/923/1028]
Physical sciences / Chemistry / Materials chemistry / Graphene
[URI /639/638/298/918]
Physical sciences / Materials science / Nanoscale materials / Carbon
nanotubes and fullerenes
[URI /639/301/357/73]
Physical sciences / Chemistry / Environmental chemistry / Pollution
remediation
[URI /639/638/169/896]
Earth and environmental sciences / Environmental sciences /
Environmental chemistry / Pollution remediation
[URI /704/172/169/896]
Physical sciences / Chemistry / Environmental chemistry
[URI /639/638/169]

Web summary

000 Materials for next-generation desalination


and water purification membranes
Jay R. Werber, Chinedum O. Osuji and Menachem
Elimelech
Membranes have an increasingly important role in
alleviating water scarcity and the pollution of aquatic
environments. Promising molecular-level design
approaches are reviewed for membrane materials,
focusing on how these materials address the urgent
requirements of water treatment applications.
©
2
0
1
6
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.

You might also like