ArticlePDF Available

Wetting, Scaling, and Fouling in Membrane Distillation: State-of-the-Art Insights on Fundamental Mechanisms and Mitigation Strategies

Authors:

Abstract and Figures

Membrane distillation (MD) has been garnering increasing attention in research and development, since it has been proposed as a promising technology for desalinating hypersaline brine from various industries using low-grade thermal energy. However, depending on the application context, MD faces several important technical challenges that would lead to compromised performance or even process failure. These challenges include pore wetting, mineral scaling, and membrane fouling. This review is devoted to providing a state-of-the-art understanding of fundamental mechanisms and mitigation strategies regarding these three challenges. Guided by the fundamental understanding of each membrane failure mechanism, we discuss both operational and material strategies that can potentially address the three technical challenges. In particular, the material strategies involve the development of MD membranes with tailored special wetting properties to impart resistance against different types of membrane failure. Lastly, we also discuss research needs and best practices in future studies to further enhance our ability to overcome technical challenges toward the practical, sustainable, and scalable applications of MD.
Content may be subject to copyright.
Wetting, Scaling, and Fouling in Membrane Distillation: State-of-
the-Art Insights on Fundamental Mechanisms and Mitigation
Strategies
Thomas Horseman, Yiming Yin, KoSS Christie, Zhangxin Wang, Tiezheng Tong,*and Shihong Lin*
Cite This: ACS EST Engg. 2021, 1, 117140
Read Online
ACCESS Metrics & More Article Recommendations
ABSTRACT: Membrane distillation (MD) has been garnering
increasing attention in research and development, since it has been
proposed as a promising technology for desalinating hypersaline
brine from various industries using low-grade thermal energy.
However, depending on the application context, MD faces several
important technical challenges that would lead to compromised
performance or even process failure. These challenges include pore
wetting, mineral scaling, and membrane fouling. This review is
devoted to providing a state-of-the-art understanding of
fundamental mechanisms and mitigation strategies regarding
these three challenges. Guided by the fundamental understanding
of each membrane failure mechanism, we discuss both operational
and material strategies that can potentially address the three technical challenges. In particular, the material strategies involve the
development of MD membranes with tailored special wetting properties to impart resistance against dierent types of membrane
failure. Lastly, we also discuss research needs and best practices in future studies to further enhance our ability to overcome technical
challenges toward the practical, sustainable, and scalable applications of MD.
KEYWORDS: membrane distillation, wetting, scaling, fouling, mitigation strategies
INTRODUCTION
Membrane distillation (MD) is a membrane-based thermal
desalination process that has received extensive and growing
research and development interests in the past few decades.
While MD has multiple congurations, each case involves the
use of a nonwetted (typically hydrophobic), microporous
membrane to serve as an airgap that separates the feed and
distillate solutions from mixing. The transport of water vapor
from the hot, salty feed solution to the cold distillate is driven
by a partial vapor pressure gradient. This partial pressure
gradient is typically induced by the temperature gradient and,
in certain cases, enhanced by a partial vacuum.
13
The interest in MD has grown substantially in recent years
due to the increasing demand for modular systems capable of
treating hypersaline water including oil- and gas-produced
water, brine from inland brackish water desalination, and brine
generated in zero liquid discharge (ZLD) processes.
47
Because of growing water scarcity and more stringent
regulations, these hypersaline brines are becoming both
unconventional sources of water and hazardous wastes of
increasing environmental concern. MD is the most promising
modular (down-scalable) technology capable of treating high-
salinity feedwater using low-grade thermal energy and, thus,
has several unique advantages for treating hypersaline brine, as
compared to the state-the-of-art desalination process, reverse
osmosis,
1,8,9
or conventional thermal distillation pro-
cesses.
1,1013
More recently, MD has also been explored as
an advanced technological platform for solar-thermal desalina-
tion due to its ability to implement latent heat recovery.
1416
Another major category of membrane processes that
envelopes MD is membrane contactor (MC), where a
nonwetted microporous membrane is also used as an airgap.
However, the intended species of mass transfer in MC is the
volatile solute instead of water vapor. For instance, MC has
been actively explored for recovering valuable gases such as
ammonia and methane, sequestrating carbon dioxide, oxygen-
ation/deoxygenation, and removal of volatile contami-
nants.
17,18
The mass transfer in an MC is also driven by a
partial vapor pressure gradient, which is often induced by a
concentration gradient and sometimes assisted by a thermal
Received: June 12, 2020
Revised: August 7, 2020
Accepted: August 11, 2020
Published: October 1, 2020
Reviewpubs.acs.org/estengg
© 2020 American Chemical Society 117
https://dx.doi.org/10.1021/acsestengg.0c00025
ACS EST Engg. 2021, 1, 117140
Made available for limited times for personal research and study only license.
Downloaded via 154.16.101.205 on July 19, 2021 at 12:03:03 (UTC).
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
gradient and/or partial vacuum (fundamentally, all mass-
transfer processes in MD and MC are driven by a chemical
potential gradient reected as a partial vapor pressure
gradient).
In both MD and MC, the microporous membranes serve as
not only a medium for vapor transport but also a barrier to
direct liquid permeation and, thus, must be maintained free
from pore blocking and wetting. However, the feedwater in
many promising applications of MD or MC often contains
constituents that promote fouling and wetting of conventional
hydrophobic membranes.
1924
For example, organic matter,
such as proteins from the food and beverage industry or oil
particles that exist in oil- and gas-produced water, are potent
foulants, especially for hydrophobic membranes.
4,2527
Low-
surface-tension, water-miscible liquids may exist in the
feedwater of MC, reducing the overall surface tension of the
feedwater and resulting in pore wetting.
28,29
In addition,
synthetic surfactants and natural amphiphiles may also reduce
the surface tension of the feedwater and induce wetting.
3032
For MD used in treating hypersaline brine, an additional
challenge is mineral scaling, that is, the formation and/or
accumulation of salt precipitates on the membrane surface that
results in signicant ux reduction and, in some cases, even
pore wetting.
3335
All these technical challenges, namely, wetting, scaling, and
fouling, constrain the practical adoption of MD and MC for
treating a wide spectrum of feedwater. In particular, these
limitations pose a paradox for MD as an eective technology
for desalinating and concentrating hypersaline brines: on the
one hand, MD is very promising for such applications due to
its (theoretical) capability of handling hypersaline brine using
low-grade thermal energy; on the other hand, concentrating
brine inevitably increases the concentrations of salts, foulants,
and surfactantswhatever constituents that originally exist in
the feedwaterand thus intensies the propensity of scaling,
fouling, and wetting and limits the (practical) applicability of
MD in various applications.
1,9,3638
It is therefore of
paramount importance for the community to gain fundamental
understanding of these challenges facing MD and MC and to
develop eective strategies of mitigating these technical
challenges to enable these membrane-based vapor-transport
processes to achieve their full potential for practical
applications. This review is precisely organized to assess the
recent advances in the fundamental understanding and
technological development we have made as a community
toward addressing these technical challenges.
In this review, we will discuss the fundamental mechanisms
and mitigation strategies for the three technical challenges in
membrane-based vapor-transfer processes: wetting, mineral
scaling, and fouling. We will focus our discussion on MD, as
most of the developments in addressing these challenges have
been made in the context of MD. However, many principles
and strategies to be discussed will also apply to MC. We
highlight the most recent research ndings that enrich or
improve our mechanistic knowledge on wetting, scaling, and
fouling in MD desalination. An important part of the
mitigation strategies involves the application of membranes
with special wetting properties, which is a relatively new eld
fueled by the recent advances in material science in
understanding and developing surfaces with special wettability.
However, we intend in this review to focus our discussion on
the fundamental mechanisms instead of presenting a
comprehensive survey of dierent membrane fabrication
methods.
PORE WETTING
The most essential function of a membrane in MD and MC
systems is to separate an aqueous stream from another liquid
or gas stream in order to facilitate volatile component transport
between the two streams. In MD and MC applications,
Figure 1. (A) Schematic diagram of membrane wetting, featuring three wetting detection techniques based on distillate conductivity, optical
transmittance, and transmembrane impedance. (B) Illustration of how dierent properties associated with dierent detection techniques vary in
dierent stages of the wetting process. We note that distillate conductivity only starts to increase when at least some pores are fully penetrated by
the feed solution. Thus, the distillate conductivity does not capture any information when some pores are only partially inltrated (but not a single
pore has been fully penetrated). Detection techniques based on transmembrane impedance and optical transmittance can provide information
regarding this stage of partial inltration.
ACS ES&T Engineering pubs.acs.org/estengg Review
https://dx.doi.org/10.1021/acsestengg.0c00025
ACS EST Engg. 2021, 1, 117140
118
membrane failure occurs when the liquid feed solution freely
permeates through the membrane, a phenomena known as
pore wetting.
27,28
In MD, wetting results in reduced salt
rejection and, thus, contamination of the distillate
stream;
2,3942
whereas, in MC applications, wetting results in
reduced mass-transfer and separation eciencies.
4345
In MD,
the simplest and most common method for detecting wetting
is measuring distillate electrical conductivity
46
(Figure 1A).
The major limitation of this technique, however, is that it only
detects wetting once the feed liquid has already fully
penetrated at least a fraction of the membrane pores and
contaminated the distillate.
Recently, in situ techniques based on single-frequency
impedance
47
and light transmittance
48
have been developed
to provide dynamic insights into the wetting process. With the
technique based on single-frequency impedance, the air-lled
pores separating the salty feed solution and distillate stream
can be modeled as an equivalent circuit
49,50
(Figure 1A). The
total impedance of this equivalent circuit decreases as the air-
gap becomes progressively thinner due to the propagation of
feed solution/air interface toward the distillate (Figure 1B).
With the optical technique based on light transmittance
(Figure 1A),theintensityoftransmittedlightisalso
dependent on the thickness of the air-lled pores (Figure
1B). The membrane transitions from being opaque when
unwetted to being translucent when wetted.
48
While the
wetting detection method is not the focus of this review, we
want to emphasize that techniques are being actively developed
to extract more information about the dynamic wetting
phenomenon that has not been accessible by previous
detection techniques that rely on the change of the distillate
properties. The ability to unveil more information about the
dynamic behavior of wetting is critical for enhancing our
fundamental understanding of pore wetting.
Mechanism of Pore Wetting. General Criterion of Pore
Wetting. Historically, pore wetting is explained based on force
balance at the triple phase boundary with the help of an
important concept called liquid entry pressure (LEP). LEP is
the minimum hydrostatic pressure required to push the liquid
into the membrane pores (for ideal cylindrical pores, entry is
equivalent to penetration) and can be estimated using eq 1
51,52
B
r
L
EP 2cos
L0
γθ
=−
(1)
where γLis the liquid surface tension, θ0is the intrinsic contact
angle between the liquid and solid membrane material, ris the
equivalent pore radius, and Bis a geometric factor accounting
for the noncylindrical nature of the membrane pore geometry
(B= 1 for perfectly cylindrical pores). In general, reducing the
surface tension of the feed solution reduces both γLand cos θ0
(as θ0is also a function of γL), thus reducing LEP and
facilitating pore wetting.
29,41,5153
The general criterion for
membrane pore wetting is that the transmembrane pressure,
ΔP, exceeds LEP.
51,54
PLE
PΔ
(2)
This criterion, as specied by the above inequality, applies well
in most cases, provided the LEP can be accurately determined.
If the membrane is composed of perfect cylindrical pores, the
system will transition from an unwetted state (Figure 2A) to a
wetted state (Figure 2C), as long as ΔPexceeds LEP, without
going through any stable transition state (Figure 2B)
observable by measuring the transmembrane impedance or
optical transmittance. This abrupt transition from unwetted
state to wetted state has been observed when (1) the hydraulic
pressure of the feed solution undergoes a signicant stepwise
increase and (2) the surface tension of the feed solution is
suddenly reduced by the addition of a large amount of water-
miscible liquids such as alcohol. The absence of observable
transition state (i.e., the nearly instantaneous wetting as shown
in Figure 2D) is consistent with the kinetics of an expanding
Poiseuille ow.
29,55,56
Assuming an ideal cylindrical pore, the
depth of pore inltration l(t) can be estimated using eq 3
29
Figure 2. Various states in a wetting process. (A) Unwetted state: all pores are neither partially inltrated nor fully penetrated. (B) Transition state:
some pores are partially inltrated, but no pore is fully penetrated. (C) Wetted state: some or all pores are fully penetrated. (D) Behavior (i.e., salt
rejection and transmembrane impedance) of wetting induced by having a hydrostatic pressure exceeding the LEP. (E) Behavior of wetting induced
by addition of surfactants.
ACS ES&T Engineering pubs.acs.org/estengg Review
https://dx.doi.org/10.1021/acsestengg.0c00025
ACS EST Engg. 2021, 1, 117140
119
lt rPt() 4(LEP)
2
μ
(3)
where ris the pore radius, μis the dynamic viscosity of the
feed liquid, tis time with t=0dened as the point that feed
solution begins intruding into the membrane pores (see Wang
et al.
29
for a more detailed derivation). Calculations based on
eq 3 suggest that liquid penetration through a membrane pore
typically takes only seconds, which explains the absence of an
observable transition state.
Surfactant-Induced Wetting. What happens when the feed
solution contains surfactants? Equation 2 suggests that wetting
would occur just as in the case when low-surface-tension
(LST) water-miscible liquid is added, except that we need a
much lower concentration of surfactants, as they are very
eective in reducing the liquid surface tension. However, both
the impedance and optical transmittance-based techniques
revealed that wetting induced by the addition of surfactants is
transient, not instantaneous; that is, the transition state as
illustrated in Figure 2B is observable for an extended period of
time.
47,48
The observable transition state suggests that
surfactant-induced wetting may have a very dierent
mechanism as compared to that of wetting induced by
increasing the ΔPor reducing the LEP via adding LST water-
miscible liquids. The key to explaining the dynamic behavior of
surfactant-induced wetting in MD is to understand that (1)
surfactants readily adsorb onto a hydrophobic surface
immersed in water and that (2) surfactants are very eective
in reducing liquid surface tension, and thus it only requires a
very low concentration of surfactants to reduce the LEP to be
below the ΔPvalue.
The dynamic wetting behavior of surfactant-induced wetting
has been recently elucidated and modeled,
29,31,32
and it is
illustrated in Figure 3A. On the one hand, the fast adsorption
of surfactants onto the hydrophobic membrane pore wall
constantly removes surfactant from the solution near the
water/air interface (i.e., the wetting frontier). On the other
hand, advective transport due to water vapor ux and diusive
transport due to concentration gradient replenish surfactants
from the feed solution (outside the pore) to the wetting
frontier. The combination of the adsorption-driven depletion
and transport-driven replenishment results in a surfactant
concentration at the wetting frontier that is lower than the bulk
concentration. The relevant LEP of the system should be
calculated, not based on the surfactant concentration of the
feed solution (in the bulk), but based on the surfactant
concentration and the corresponding surface tension of the
feed solution at the wetting frontier. In this context, eq 2 again
becomes applicable, as the criterion for pore wetting, except
that in this case the LEP of the wetting frontier, not of the bulk
feed solution, should be compared with ΔP.
The theory presented above explains why instantaneous
wetting does not occur even if LEP calculated from the bulk
surfactant concentration is lower than ΔP. However, pore
wetting still occurs gradually, because the pore wall surface has
a limited capability for surfactant adsorption. When the wetted
Figure 3. (A) Mechanism of surfactant-induced pore wetting featuring surfactants adsorption at the wetting frontier and the transport of surfactants
from the bulk feed solution to the wetting frontier via advection and diusion. (B) Bridge buildinganalogy of the dynamic wetting model: passing
the river (equivalent to wicking through the pore) requires extending the bridge (equivalent to saturating the pore surface surrounding the
wetting frontier). The impacts of (C) bulk surfactant concentration, (D) vapor ux, and (E) surfactant type, on the kinetics of wetting as
quantied by breakthrough time or its inverse. The corresponding variables in the bridge buildinganalogy are also given on the top of each panel.
Adapted with permission from (C, D) ref 31 and (E) ref 32; Copyright 2018 and 2019 Elsevier, respectively.
ACS ES&T Engineering pubs.acs.org/estengg Review
https://dx.doi.org/10.1021/acsestengg.0c00025
ACS EST Engg. 2021, 1, 117140
120
surface of the pore is saturated, adsorptive depletion of
surfactants no longer occurs, which leads to the increase in the
surfactant concentration at the wetting frontier and, corre-
spondingly, to the reduction of the surface tension and LEP.
Once the LEP becomes lower than ΔP, wetting occurs, and the
wetting frontier propagates toward the distillate. This forward
propagation exposes a new region of the pore to the feed
solution, which again enables rapid adsorption of surfactants to
increase the LEP beyond ΔPuntil that new region is again
saturated with surfactants. Such a process of adsorption,
saturation, and forward propagation of the wetting frontier
repeats itself and results in the experimentally observed
transient wetting. While this phenomenon is described here
as a stepwise process to facilitate understanding, it is
continuous in a real process.
It may be helpful to illustrate this new model for surfactant-
induced wetting with an interesting analogy of bridge
building(Figure 3B). To build a bridge (i.e., for the feed
solution to fully penetrate a hydrophobic pore), bricks (i.e.,
surfactant molecules) must be transported to the bridge
frontier and laid there to further extend the bridge. The brick-
laying process (i.e., the adsorption of surfactants to the pore
wall at frontier) is very fast. Therefore, the kinetics of the
bridge-building process is limited by how fast the bricks are
transported to the bridge frontier. This bridge building analogy
leads to several important implications. For example, if the
brick-transporting truck carries more bricks per truck, which is
equivalent to having a higher feed concentration of surfactants
(as the advection term is proportional to concentration), the
bridge building (pore wetting) process is faster (Figure 3C;
note that, if the bulk surfactant concentration is too low,
wetting does not occur, because ΔPis lower than the LEP of
the bulk solution). In addition, if the truck moves faster, which
is equivalent to having a higher ux (as the advection term is
proportional to ow velocity), the bridge building (pore
wetting) process is also faster (Figure 3D). These two
dependences of wetting kinetics on experimental conditions
have been shown experimentally and predicted accurately by
the dynamic wetting model.
31,32
The most interesting conclusion from the model regards
how the surfactant type aects the wetting kinetics. The bridge
building analogy suggests that, if the bricks are bigger
(primarily, longer), the bridge will be extended faster provided
(1) all trucks move at the same speed and (2) each truck
carries the same number of bricks. Equivalently, the dynamic
wetting model predicts that biggersurfactants will lead to
faster pore wetting (i.e., shorter breakthrough time), which has
also been experimentally conrmed
31,32
(Figure 3E). Here, the
size of the surfactants can be quantied using surface excess
concentration Γ, which can be evaluated using the Gibbs
adsorption isotherm.
5759
Specically, a very nice linear
correlation has been experimentally shown between the surface
excess concentration and the breakthrough time for a series of
surfactants of dierent charges (Figure 3E, except for sodium
dodecyl sulfate, for which diusion has an important
contribution to transport. See a detailed explanation in the
work of Wang et. al.
32
). This experimentally validated model
emphasizes the sizeof surfactants as a key property that
inuences how eciently dierent surfactants induce wetting
(assuming the same molar concentration), which is in contrast
to the previous understanding that focused on the interaction
between the surfactant and the pore surface.
30,60
According to
the dynamic wetting model, the surfactant-pore surface
interaction is irrelevant to the kinetics of pore wetting, because
adsorption, which is much faster than transport of surfactants
to the frontier, is not the rate-limiting step.
The dynamic model for surfactant-induced wetting is the
rst quantitative model for predicting the kinetics of surfactant-
induced wetting. The model, which provides a satisfactory
tting to the experimental data from multiple dimensions, is
based on the assumption that wetting occurs only when the ΔP
exceeds the LEP at the wetting frontier. Another plausible
mechanism for propagation of the wetting frontier is via the
autophilic eect, that is, the surfactants can adsorb onto the
unwetted part of the surface ahead of the triple phase boundary
and thereby reduce the surface energy of the unwetted part of
the surface near the wetting frontier, making it hydrophilic.
61,62
As a result, the propagation of the liquidair interface toward
the distillate is spontaneous and can occur without applied
pressure. The autophilicexplanation, however, has challenges
in explaining why there exists a minimum bulk surfactant
concentration below which wetting does not occur (Figure
3A). Moreover, the subtle impact of ΔPon the wetting
kinetics, which has been experimentally validated, also appears
to conrm the assumed wetting criterion based on comparing
the ΔPand the LEP. Whether an autophilic eect does play an
important role in surfactant-induced wetting requires further
investigation. However, even assuming an autophilic eect is
important, because the adsorption of surfactants onto a
hydrophobic surface immersed in water is highly energetically
favorable, we speculate that the autophilic eect can only exert
its contribution when a wetted pore surface is saturated with
surfactants. In other words, the bridge building analogy and the
kinetic model of pore wetting focusing on the critical role of
surfactant transport should apply regardless of the exact
mechanism of frontier propagation.
Wetting Mitigation Strategies. Pretreatments. There
are limited pretreatment techniques that may mitigate
membrane pore wetting. For example, physical pretreatments
such as microltration (MF), ultraltration (UF), or nano-
ltration (NF) can eectively remove amphiphilic proteins that
may eventually wet the pores of a hydrophobic membrane.
63,64
However, they cannot mitigate pore wetting induced by
surfactants or LST water-miscible liquids. Surfactants can be
removed from solution via ion exchange,
65
coagulation,
66,67
oatation or foam fractionation,
6870
or biodegradation.
71,72
While some of the aforementioned pretreatment methods have
not yet been tested for MD or MC and a knowledge gap exists
in matching the appropriate pretreatment method given a real
water application, some studies have shown promise in using
these pretreatment strategies for MD and MC. For example,
dispersing bubbles into a surfactant-containing feed solution
from the textile industry formed oating foams that removed
surfactants from the bulk solution. This foam fractionation
pretreatment in turn eliminated the surfactant-induced wetting
that had previously occurred in the downstream MD process.
70
In another study, MC used for ammonia stripping from a
model wastewater stream was plagued by eventual wetting due
to amphiphilic protein molecules. After the integration of the
MC process with a microltration or ultraltration pretreat-
ment, the tested poly(tetrauorethylene) (PTFE) and
polypropylene (PP) membranes realized a twofold and
fourfold increase in the ammonia mass-transfer coecients,
respectively, and membrane wetting was never observed
throughout the duration of the 30 h experiment.
64
For a
feed solution containing LST water-miscible liquids as often
ACS ES&T Engineering pubs.acs.org/estengg Review
https://dx.doi.org/10.1021/acsestengg.0c00025
ACS EST Engg. 2021, 1, 117140
121
encountered in MC, no viable pretreatment exists for wetting
mitigation, as the removal of the wetting agent is precisely the
technical goal of MC itself. In general, if the feedwater has a
composition that may promote pore wetting in MD or MC,
options to mitigate wetting via pretreatment are fairly limited.
Therefore, a dierent approach based on developing a wetting-
resistant membrane is needed for more versatile and robust
wetting mitigation.
Omniphobic Membranes. Because all wetting processes,
whether induced by surfactants or alcohol, follow the same
principle described by eq 2, we can develop membranes that
resist wetting by LST liquids as a strategy for wetting
mitigation. Membranes that are resistant to wetting by LST
liquids, such as oil, are called oleophobic membranes. In most
cases, oleophobic membranes are also hydrophobic and thus
are referred to as amphiphobic or omniphobic membranes.
73,74
Without specifying the medium, the denitions of wetting
property assume air as the medium. Strictly speaking,
omniphobicity literally refers to resistance to wetting by all
liquids, whereas amphiphobicity refers to wetting resistance to
both water and oil. The dierence between the two concepts is
more quantitative than qualitative. For instance, a membrane
may be referred to as amphiphobic if it is resistant to wetting
by mineral oil (γ30 mN m1) but not by ethanol (γ22
mN m1) and n-decane (γ24 mN m1). Therefore, we use
omniphobic membranes here to refer broadly to membranes
with resistance to wetting by LST liquids.
The development of an omniphobic membrane for MD
application was rst reported in 2014
28
and has since attracted
extensive research interest.
7382
For example, omniphobic
membranes have shown promising wetting resistance in
treating produced water from the oil and gas
73,74
and coking
industries.
76
However, the fundamental mechanism of
achieving omniphobicity had been elucidated by the material
science community several years earlier.
8386
Without going
through the theoretical derivation (interested readers are
referred to these publications
83,84
), here we summarize the
criteria for developing omniphobic membranes and the basic
principle behind such criteria.
The two major criteria for making an omniphobic
membrane or, more generally, an omniphobic surface, are
that (1) the material has a low surface energy and (2) the
surface has a reentrant texture. The rst criterion is shared by
both hydrophobic and omniphobic membranes and is often
met by using uoro-polymers and surface modiers, which are
known to be of low surface energy and chemically relatively
stable. In describing the wetting state of a general textured
substrate surface, if the liquid is in full contact with the
substrate, the system is in a Wenzel state; if the liquid is
supported by a composite surface comprising the protrusions
of the solid substrate and the air pockets between these
protrusions, the system is in a CassieBaxter state.
In the context of membrane wetting, the composite interface
at the wetting frontier that comprises the membrane pores and
the solid surface itself envelopes the aforementioned surface
texture. The CassieBaxter state corresponds to the unwetted
state depicted in Figure 2A, when the LEP at the wetting
frontier is high enough to deter any liquid inltration into the
membrane pores toward the direction of the distillate stream.
Likewise, the Wenzel state corresponds to the wetted state
depicted in Figure 2C, when the LEP at the wetting frontier
satises the condition stipulated by eq 2 and thus results in
pore wetting. To prevent wetting of any degree, the wetting
frontier needs to be maintained in a CassieBaxter state at the
membrane surface. The thermodynamics of wetting suggests
that the maintenance of a CassieBaxter state for high-surface-
tension liquids, such as water, can be achieved as long as the
porous membrane is made of low-surface-energy materials.
75,87
However, it also suggests that the CassieBaxter state is not
stable (i.e., it has a higher free energy than that of the Wenzel
state) if the surface tension of the liquid is suciently low.
Here comes the important role of the reentrant texture (the
second criterion) that sustains a metastable CassieBaxter
state.
Figure 4. (A) Local intrinsic contact angles at the local triple-phase boundary for a reentrant structure (top) and a nonreentrant structure
(bottom). (B) Illustration of how a liquid intrusion into a pore with reentrant geometry would change the contact area of the liquid/solid and
liquid/air interfaces. (C) Free energy prole as a function of the position of the liquid/air interface for pores with reentrant and nonreentrant
geometries.
ACS ES&T Engineering pubs.acs.org/estengg Review
https://dx.doi.org/10.1021/acsestengg.0c00025
ACS EST Engg. 2021, 1, 117140
122
A reentrant texture is a concave topography in which the
cross-sectional area of a pore increases with the depth of the
pore. A classic reentrant texture is an array of inverted cones.
Figure 4A shows the cross-section of space between two
inverted cones (top) versus that between two cylinders
(bottom). The requirement of a reentrant texture to maintain
a CassieBaxter state with even LST liquids can be explained
using both force balance and free energy analysis. Without a
reentrant texture, the maintenance of a CassieBaxter state
requires that local contact angle be higher than 90°(Figure 4A
bottom), which can be satised if the solid material
constituting the pores has a low surface energy and the liquid
has high surface tension (e.g., water). However, with LST
liquids, a local contact angle does not exceed 90°even with
solid material with low surface energy. In this case, the Cassie
Baxter state can only be maintained with a reentrant texture
with which the local contact angle can be lower than 90°
(Figure 4A top).
The necessity of reentrant texture for achieving omnipho-
bicity can also be elucidated from the perspective of free
energy. When liquid inltrates into a pore, part of the solid/air
interface is replaced by a solid/liquid interface (Figure 4B). If
the solid surface has a low surface energy and the liquid has a
high surface tension, the replacement of the solid/air interface
by a solid/liquid interface is energetically unfavorable, which
explains the stable CassieBaxter state for water contacting a
hydrophobic membrane. When an LST liquid is considered,
this replacement is energetically favorable even if the pore is
made of low-surface-energy materials. However, with and only
with reentrant texture, the liquid inltration expands the
liquid/air interface, which is energetically unfavorable. If the
unfavorable expansion of the liquid/air interface outweighs the
favorable expansion of the solid/liquid interface, a free energy
barrier would exist in the process of the transition from the
CassieBaxter state to the Wenzel state when the pore is fully
inltrated (Figure 4C). Therefore, a metastable CassieBaxter
state can exist with LST liquid in the presence of a reentrant
structure. On the contrary, the transition from the Cassie
Baxter state to the Wenzel state would monotonically reduce
the free energy if the system does not have a reentrant texture
that promotes expansion of the liquid/air interface upon pore
inltration (Figure 4C).
While inverted cones are convenient models for illustrating a
reentrant structure, they are not a practical structure to
engineer on a membrane. Instead, bers from electrospinning
or nanoparticles from spraying or solution-phase adsorption
are used to construct the reentrant texture for omniphobic
membranes.
7476,78
Recent studies also nd that omniphobic
membranes can be obtained by simple chemical modication
to reduce the surface energy of commercial hydrophobic
membranes, which suggests that the structure of some
commercial hydrophobic membranes is already reentrant.
88
It also suggests that whether a CassieBaxter state can be
maintained depends on the intricate interplay between liquid
surface tension, surface energy of the solid, and the pore
geometry. If the membrane is made of material of lower surface
energy and/or has a more robust reentrant structure (the
readers can refer to ref 84 for the theory about the robustness
of oleophobicity), the membrane may become more
omniphobicand can sustain stable MD performance with a
feed solution of a lower surface tension. In an extreme case
with a doubly reentrant structure, even a surface of high surface
energy (e.g., silica) can be omniphobic.
89
As mentioned above,
this article is not focusing on membrane fabrication methods.
Therefore, readers who are interested in fabricating omni-
phobic membranes can refer to these review papers.
38,82,90
MINERAL SCALING
Mineral scaling is extremely relevant to MD, as its most
opportunistic applications are for treating hypersaline indus-
Figure 5. (A) CNT considering contributions from both reduction in bulk free energy and increase in interfacial free energy. (B) Reducing free
energy barrier for nucleation via (1) heteronucleation and (2) oversaturation. (C) Local spatial distribution of temperature and salt concentration
in the direction normal to the membrane, featuring temperature and concentration polarizations. Specically, T, IAP, and SI represent temperature,
ion activity product, and saturation index, respectively; the rst subscripts, F and D, denote the properties of the feed solution and distillate,
respectively; the second subscripts, B and M, denote the properties in the bulk and at the membrane surface, respectively. (D) Spatial distribution
of temperature (represented by colorred symbolizing high temperature and blue symbolizing low temperature) and salt concentration
(represented by density of white dots) in the MD module.
ACS ES&T Engineering pubs.acs.org/estengg Review
https://dx.doi.org/10.1021/acsestengg.0c00025
ACS EST Engg. 2021, 1, 117140
123
trial wastewater where reverse osmosis (RO) is not feasible.
1
Depending on the pretreatment methods used, hypersaline
wastewaters often contain ions that form sparingly soluble
minerals (i.e., sulfates, carbonates, and silicates).
91
During MD,
these compounds are concentrated as water evaporates, which
eventually leads to the development of a precipitated layer of
mineral crystals on the membrane surface, also known as
mineral scaling.
34
Mineral scaling can occur via the deposition
of precipitates formed in the bulk solution or direct nucleation
and growth of precipitates on the membrane surface. In either
case, scaling reduces process eciency or leads to process
failure, and it must be properly addressed to maximize the
potential for applying MD toward addressing broader water
treatment challenges.
92
Mineral scaling is mostly detected by monitoring the water
vapor ux over time.
93,94
Typically, an appreciable decrease in
water vapor ux is indicative of scaling, as the scaling layer
blocks the membrane pores. Although this method involves a
straightforward indicator that is directly relevant to membrane
performance, it only allows scaling to be detected after much of
the membrane surface has been covered and provides little
information regarding the growth kinetics, spatial distribution,
morphology, and composition of the scale layer. Ex situ
membrane autopsy has helped elucidate scaling mechanisms
and how they relate to operation parameters and membrane
characteristics. A suite of standard material characterization
techniques has been used for membrane autopsy, including
scanning electron microscopy (SEM) with energy-dispersive
X-ray spectroscopy (EDS), Fourier-transformed infrared
(FTIR) spectroscopy, atomic force microscopy (AFM), X-ray
diraction (XRD), and X-ray photoelectron spectroscopy
(XPS).
9597
There has also been some progress on the development of in
situ scaling observation techniques. For example, optical
coherence tomography (OCT) has revealed that hardly any
scaling particles were visible before the vapor ux decline.
98,99
Also, a laser-based optical technique has been used to measure
the concentration proles of metal chloride salts in the MD
feed solution near the membrane surface.
100
However, more
powerful and easy-to-implement in situ scaling detection
methods are still in need to better elucidate the dynamics of
scaling, especially during the early stages of crystal growth.
Mechanism of Scaling. General Scaling Criterion.
Mineral precipitates form when the speciesconcentration
product exceeds its solubility product. Like other mineraliza-
tion processes occurring in nature, the formation of mineral
scales in membrane desalination follows the theories
established under the framework of classic nucleation theory
(CNT).
101
The CNT describes the change of free energy
(ΔG) associated with the formation of a nascent cluster of a
new solid phase (from ions), which is expressed as the sum of
two terms (eq 4,Figure 5A)
102
Gn A
l
n
μγ
Δ
=− Δ +
(4)
where nis the number of ions in a cluster, Δμrefers to the
chemical potential dierence of the ions between their solution
phase and solid phase, Ais the surface area of the cluster, and
γln refers to the interfacial energy between the liquid and the
nucleus. The rst term on the right side of eq 4 represents the
energy change associated with the formation of bulk mineral
from ngrowth units (i.e., bulk free energy), which is always
negative under supersaturated conditions; the second term
regards the energy gain (penalty) associated with increasing
the interfacial area (i.e., interfacial free energy). Assuming a
spherical nucleus of radius r, the change in the free energy
during homogeneous nucleation is given by eqs 5 and (6)
102
Gr N
vrr() 4
34
hom A
m
3
ln
2
πμ πγ
Δ
=− Δ+
(5)
where NAand vmare the Avogadro number and molar volume,
respectively. The chemical potential dierence Δμis depend-
ent on the degree of oversaturation
kT K
ln IAP
B
sp
i
k
j
j
j
j
j
j
y
{
z
z
z
z
z
z
μ
Δ
=
(6)
where kBis the Boltzmann constant, Tis the absolute
temperature, IAP is the ion activity product, and Ksp is the
equilibrium constant (or solubility product). As shown in
Figure 5A, ΔGhom(r) reaches a maximum value (ΔG*) when
the mineral nucleus grows to a critical size r*. Only the nuclei
with sizes larger than r*can evolve into thermodynamically
stable minerals. Therefore, ΔG*represents the height of the
energy barrier of mineral nucleation.
Equations 5 and (6) bridge the degree of saturation of the
feed solutions to the thermodynamic energy barrier of scale
formation. Saturation index (SI) is often used in the literature
to quantify the scaling potential of the feedwater in membrane
desalination.
95,103106
While there are multiple denitions of
SI in the literature (it has been dened as ln(IAP/Ksp),
log10(IAP/Ksp), or (IAP/Ksp)), the SI based on all these
denitions consistently increases as the solution becomes more
concentrated. With an increase of SI, Δμincreases (according
to eq 6) and ΔGdecreases (according eq 4 or (5)), resulting in
a lower nucleation energy barrier (ΔG*in Figure 5B) and
easier mineral formation.
For heterogeneous nucleation, which could take place on
membrane surfaces or foreign particles in the bulk solution, the
interfacial energy at the liquid-substrate (γls), substrate-nucleus
(γsm), and liquid-nucleus (γls) interfaces needs to be
considered.
102
The presence of a substrate surface decreases
the interfacial free energy (Figure 5C), which also leads to a
reduction of the nucleation energy barrier. The energy barrier
of heterogeneous nucleation (ΔGhet
*) relates to that of
homogeneous nucleation (ΔGhom
*) by a correction factor
called the wetting function f(θn/w)
107
Gf G
G
() 1
4(1 cos )
(2 cos )
het n/w hom n/w 2
n/w hom
θθ
θ
Δ*
*=−
×+ Δ
*(7)
where θn/w is the contact angle of a nucleus (assuming a
spherical cap geometry) on the substrate in water as the
medium (we note that θnshould not be misinterpreted as the
in-air water contact angle of the membrane substrate, which
will be denoted as θw). Because the morphology of the nucleus
is determined by the crystal structure and not necessarily a
spherical cap, θn/w is more of a hypothetic metric to quantify
the interaction between the mineral nucleus and the substrate
based on the Young equation.
cos n/w ls sn
ln
θγγ
γ
=
(8)
Most membrane substrates in MD are nonpolar polymers that
do not have polar interaction (also called acidbase
ACS ES&T Engineering pubs.acs.org/estengg Review
https://dx.doi.org/10.1021/acsestengg.0c00025
ACS EST Engg. 2021, 1, 117140
124
interaction according to Van Oss
108
). Therefore, the γls and γsm
terms in eq 8 can be written as
109
2
ls l s l
D
s
D
γ
γγ γγ=+− (9)
and
2
sn s n s
D
n
D
γ
γγ γγ=+ − (10)
where γl,γs, and γnare the surface energy of the liquid,
substrate, and nucleus, respectively, and γl
D,γs
D, and γn
Dare the
contributions of dispersion force (also called Lifshitz-van der
Waals interaction by Van Oss) to the surface energy of the
liquid, substrate, and nucleus, respectively. Considering eqs 9,
(10), and the equality of γsγs
Dfor a nonpolar substrate, eq 8
can be rewritten as
cos 2( )
n/w
ln sl
D
n
D
ln
θ
γγ γγ γ
γ
=
−− −
(11)
Depending on the temperature, γl
Dis slightly higher than 20
mN m1for water. Without known exception, γn
Dis always
larger than γl
D(e.g., γn
D37 mN m1for silica
110
and γn
D47
mN m1for gypsum
111
). Therefore, l
D
γ
n
D
γis negative,
and thus cos θn/w positively correlates with γs. In other words,
as the substrate becomes more hydrophobic (as a result of
reduced γs), the contact angle θn/w in eq 7 also increases (as
cos θn/w decreases).
The value of f(θn/w)ineq 7 ranges between 0 and 1 when
θn/w changes from 0 to π, consistent with the above discussion
that ΔGhet
*is lower than ΔGhom
*. The occurrence of
heterogeneous nucleation is more favorable than homogeneous
nucleation, and thus the presence of a surface (e.g., membrane)
promotes scale formation. While ΔGhom
*is a constant at a
specic temperature and SI, a dierence in membrane
materials may result in dierent energy barriers for
heterogeneous nucleation. This partially lays the foundation
of fabricating scaling-resistant membranes by tuning mem-
brane surface properties to inhibit heterogeneous nuclea-
tion.
95,96,103,112,113
The CNT theory also provides insights into the change of
scaling potential due to the concentration and temperature
polarization in an MD process. Concentration polarization is a
phenomenon in which solutes accumulate within a boundary
layer near the membrane surface as a result of the advective
transport of solute toward a nonpermeable boundary and the
back-diusion due to concentration gradient (Figure 5C).
Concentration polarization leads to a higher IAP at the
membrane surface than in the bulk, which consequently aects
the SI at the membrane surface. Besides, temperature
polarization refers to a decrease of feedwater temperature
from the bulk to the membrane surface due to the limited heat
transfer rate across the boundary layer (Figure 5C).
114
Temperature polarization reduces the temperature at the
membrane surface as compared to that of the bulk and, also,
thereby alters SI via changing the temperature-dependent Ksp.
For example, the solubility of NaCl and silica is enhanced with
increased temperature,
115,116
whereas an inverse correlation
between temperature and solubility was reported for calcite
(CaCO3) and Na2SO4.
116118
There are also minerals for
which the correlation between solubility and temperature is
not monotonic. For example, the solubility of gypsum (CaSO4·
2H2O) peaks at 40 °C.
119
The combination of the eects of
concentration and temperature polarization may result in an SI
at the membrane surface that diers substantially from that in
the bulk, which adds to the impact of heterogeneous
nucleation.
In practical MD membrane modules, the mass and heat
transfer across the membrane also causes spatiotemporal
variations in temperature and solute concentration along the
channels. As more water is recovered, the solute concentration
in the feed stream gradually increases. However, such an
increase in solute concentration occurs more temporally than
spatially, as the maximum water recovery per pass is limited to
a few percent.
10,11
Therefore, the scaling potential always
increases over time, as water recovery increases. In contrast,
the spatial distribution of temperature along the module is
substantial due to the large amount of latent heat transfer
associated with vapor transfer (Figure 5D). Consequently, the
dependence of scaling potential on the module position (at a
given moment) is strongly dependent on the type of
correlation between solubility and temperature.
Understanding Mineral Scaling Beyond CNT. Scaling in
MD is a rather complex process that involves multiple
fundamental phenomena in parallel and/or in sequence.
These phenomena include, but are not limited to, ion
transport, nucleation, precipitate growth, particle deposition,
and adhesion. While understanding CNT and concentration/
temperature polarization are important to understand scaling,
knowledge gaps still exist to precisely predict long-term
membrane performance in MD subject to mineral scaling.
Typically, membrane performance refers to vapor ux and salt
rejection. These two metrics can be easily measured and are
most relevant in practical operation. A widely investigated
parameter that is particularly relevant to understanding the
kinetics of mineral scaling is the induction time. The relevant
induction time is the period from process initiation to the
onset of irreversible ux decline due to mineral scaling.
Notably, the induction time of ux decline caused by mineral
scaling in membrane desalination is fundamentally distinct
from the induction time of scale formation, which refers to the
time between the generation of a supersaturated solution and
the formation of the rst nuclei.
120
The induction of scale
formation, a physicochemical phenomenon, is believed to
occur much earlier than the onset of ux decline, an
operational phenomenon. The relationship between those
two inductionsis still unknown and requires further
investigation.
The roles of heterogeneous and homogeneous nucleation in
regulating scaling behavior in MD have not been fully
understood. Although heterogeneous nucleation occurs more
favorably than homogeneous nucleation due to the reduction
of the energy barrier, both may take place in an MD process,
especially considering the high salinity of the feedwater. The
contributions of the heterogeneous and homogeneous
nucleations to mineral scaling are dicult to quantify
separately, but a better understanding of their respective
contributions is essential for predicting scaling kinetics and
developing scaling mitigation strategies. For example, homoge-
neous silica nucleation has been shown to play a dominant role
in initiating the decline of water vapor ux, while
heterogeneous silica nucleation facilitates the membrane pore
blockage.
121
In another study, ltering a supersaturated
calcium sulfate feed solution to remove homogeneously
precipitated crystals proved to limit the rate of ux decline
due to scaling.
122
If that is proven generally true for all
ACS ES&T Engineering pubs.acs.org/estengg Review
https://dx.doi.org/10.1021/acsestengg.0c00025
ACS EST Engg. 2021, 1, 117140
125
common scaling species, the inhibition of precipitation in the
bulk solution and the abatement of precipitate deposition onto
the membrane surface will be more eective in mitigating
membrane scaling than hindering heterogeneous nucleation by
altering membrane surface properties.
In addition, to develop an improved general framework for
understanding mineral scaling in MD, we also need to acquire
a better understanding of the species-dependent behavior of
MD scaling. Dierent types of scaling could have distinct
mechanisms of precipitate formation at the molecular level and
dierent consequences in their macroscopic scaling behavior
(Figure 6).
35
Some scales, including gypsum, calcite, and NaCl,
are formed via a crystallization process (Figure 6A). In the case
of gypsum scaling, it has been shown that crystal growth causes
scaling intrusion (into the pores) and pore deformation
(Figure 6B). The pore deformation eventually leads to
membrane wetting, as LEP is reduced with larger deformed
pores.
35,123,124
In contrast, silica scaling results from the
polymerization of silicic acid followed by gelation (Figure
6C).
121
Consequently, it does not result in the same intrusion
and pore deformation phenomena observed in gypsum scaling
(Figure 6D). In consequence, MD with silica scaling is only
plagued with uxdeclinebutnotmembranewetting.
Therefore, mineral scaling formed by crystallization should
receive special attention in MD desalination of wastewater with
mixed salts, due to its potential for causing membrane wetting.
It would be valuable to evaluate whether wetting-resistant
membranes, such as those described above, are also eective to
mitigate membrane wetting caused by inorganic scaling.
Scaling Mitigation Strategies. Pretreatments. The use
of antiscalants is the most straightforward approach of
controlling mineral scaling in membrane desalination including
MD.
125134
For example, organic phosphonate derivatives
were found to mitigate the precipitation of calcium scales in
the MD treatment of seawater RO brine,
133
while the
application of a proprietary polymeric compound was reported
to retard calcite scaling in MD with coal seam gas RO brine as
the feedwater.
135
Additionally, carboxylic-based polymeric
molecules also inhibit the nucleation of calcite and gypsum
in MD.
128
Antiscalants typically serve as nucleation inhibitors
that hinder homogeneous nucleation in feed solutions. The
two major mechanisms that are likely responsible for scaling
inhibition are (1) the formation of soluble complexes in the
bulk solution, which decreases the SI by reducing availability of
the freeions for precipitation, and (2) the direct adsorption
onto the nuclei surface, which retards the mineral growth.
9,37
Antiscalants that are widely used in industry are weak acids
such as phosphonate derivatives and polymeric molecules
anchored with carboxylic groups. Under near-neutral con-
ditions of desalination, such antiscalants are partially or fully
deprotonated, exposing negative active sites that are able to
form complexes with multivalent cations in the solution.
136139
For example, poly(acrylate acid) (PAA), a carboxylic derivative
polymer with pKaof 4.4,
140
is highly deprotonated at
approximately the pH of 7 and may strongly chelate with
Ca2+ to reduce its activity for precipitating with CO32or
SO42to form calcite or gypsum.
141,142
However, this complex
formation mechanism is challenged by the fact that the
antiscalants can be highly eective at a very low concen-
trationmuch lower than the concentration of Ca2+ stoichio-
metrically.
143,144
Another possible mechanism for the role of antiscalants in
scaling inhibition is the adsorption of antiscalant molecules
onto the surface of crystal nuclei via either electrostatic
interaction or ligand exchange.
142147
Adsorption of anti-
scalants may contribute to scaling mitigation in two dierent
ways. First, the adsorption of antiscalants on the nuclei surface
increases the interfacial free energy at the liquid-nucleus
interface,
148
thereby increasing the energy barrier of nucleation
according to the CNT as discussed in the previous section.
This consequently reduces the nucleation rate and extends the
induction time of scaling.
128,145,148
Second, the attachment of
antiscalants reduces the active nuclei surface area for crystal
growth.
149,150
Figure 6. (A) Mechanism of gypsum scaling-induced wetting via pore deformation under crystallization pressure. (B) SEM-EDS images showing
gypsum intrusion into the membrane pores. (C) Silica scaling via gelation of silica nanoparticles. (D) SEM-EDS images showing the absence of
silica intrusion into membrane pores. Adapted with permission from ref 35.
ACS ES&T Engineering pubs.acs.org/estengg Review
https://dx.doi.org/10.1021/acsestengg.0c00025
ACS EST Engg. 2021, 1, 117140
126
Besides the use of antiscalants, other strategies have been
applied for the mitigation of mineral scaling in MD.
Coagulation and softening are common approaches to remove
scalants from the feedwater prior to the MD process.
151
For
example, precipitative softening has been shown to reduce the
concentrations of several scale-forming species (e.g., Ca2+,Sr
2+,
and silica) in produced water from shale oil and gas
production, decreasing the corresponding SI of sparingly
soluble salts such as calcite, silica, and strontianite (SrCO3).
152
This resulted in the eective mitigation of mineral scaling and
improved water recovery in the subsequent MD process. In
addition to these chemical-based pretreatments, nanoltration
has been explored for softening the feedwater before MD to
mitigate scaling.
33
Superhydrophobic Membranes. The use of a super-
hydrophobic membrane can complement the pretreatment
strategies and contribute to the overall reduction in membrane
scaling potential.
95,96,112,113,153,154
Superhydrophobic mem-
branes are characterized by their extreme nonwetting property.
In general, a surface is typically considered superhydrophobic if
(1) the static water contact angle is higher than 150°and (2)
the sliding angle is below 10°(alternatively, contact angle
hysteresis, which is the dierence between the advancing and
receding contact angles, can be used instead of a sliding angle).
In the MD literature, superhydrophobic membranes are
sometimes also referred to as slippery membranes.
1,112,113
Membrane superhydrophobicity is achieved by minimizing the
liquidsolid contact area via introducing rough, and preferably
hierarchical, surface texture. If we denote ϕas the areal fraction
of a solidliquid interface over the entire liquid interface (with
both solid and air), the apparent water contact angle θw,A (i.e.,
the water contact angle actually measured with the membrane)
relates to the intrinsic contact angle θw,0 (i.e., the water contact
angle measured with a molecularly smooth, nonporous surface
made of the same material as the membrane) via the following
equation.
155
cos (cos 1) 1
w,A w,0
θϕθ=+(12)
The simplest and most common fabrication technique to
create a superhydrophobic membrane is to coat the membrane
surface with nanoparticles grafted with peruorinated or other
ultralow surface energy,
95,96
although many other techniques
can also be used.
96,112
In all cases, the major requirement is
that the membrane surface has a high degree of roughness and
low surface energy.
Several studies have shown that scaling can be mitigated
with superhydrophobic membranes to dierent extents
depending on the scaling type and other factors.
95,96,112,113,156
With the use of superhydrophobic membranes, gypsum scaling
was substantially delayed, and NaCl scaling was not even
observed.
96,112,113
Also, the superhydrophobic membrane also
showed promising eectiveness in reducing mineral scaling in
MD treatment of industrial wastewater such as the blowdown
water from the cooling tower of a power plant.
95
Although the
exact mechanisms of scaling mitigation by superhydrophobic
membranes are unclear, there are several possible explanations.
First, the small liquidsolid contact area with a super-
hydrophobic membrane reduces the area available for crystal
deposition or nucleation on the membrane surface (Figure
7A). For example, consider a hydrophobic and super-
hydrophobic membrane made from the same material with
θw,0 = 105°. For the rough, superhydrophobic membrane with
θw,A = 165°, the areal fraction of liquidsolid contact ϕis
estimated to be only 5% based on eq 7. In comparison, for
the hydrophobic membrane with θw,A = 120°,ϕis estimated to
be 67%. The larger liquidsolid contact oers more area for
heterogeneous nucleation and crystal deposition and/or
adhesion (Figure 7B).
Figure 7. Possible mechanisms for the scaling mitigation with superhydrophobic membranes (top) relative to conventional hydrophobic
membranes (bottom). (A, B) Reducing solid/liquid contact area. (C, D) Reducing nucleation propensity by reducing surface energy. (E, F)
Enhancing boundary cross-ow and eliminating stagnant zones in a partially wetted pore. Adapted with permission from ref 96.
ACS ES&T Engineering pubs.acs.org/estengg Review
https://dx.doi.org/10.1021/acsestengg.0c00025
ACS EST Engg. 2021, 1, 117140
127
The second possible explanation of the scaling mitigation by
a superhydrophobic membrane is the reduced propensity for
heterogeneous nucleation, which can be explained by the CNT
and eq 7 introduced in the last section. Equation 7 suggests
that the nucleation on an interface with higher θn/w is more
dicult, because f(θn/w) is a monotonically increasing function
for θn/w [0,π]. As discussed earlier, θn/w positively correlates
with the intrinsic (in-air) water contact angle θw,A. Therefore,
we should expect that nucleation on a surface with lower
surface energy γsalso has a higher energy barrier than
nucleation on a surface with higher γs. Consequently,
nucleation propensity is substantially lower with a super-
hydrophobic membrane for two reasons. First and most
importantly, with a superhydrophobic membrane, there is a
larger fraction of the liquidair interface (Figure 7C), which
has the lowest nucleation propensity (as γs0 for air).
Second, because the fabrication of a superhydrophobic
membrane typically uses materials (e.g., peruorinated
compounds) with a lower γsthan that used for fabricating
hydrophobic membranes (e.g., poly(vinyl diuoride) or
polypropylene), the nucleation propensity on the liquid
solid interface is also lower with a superhydrophobic
membrane (Figure 7C) than with a hydrophobic membrane
(Figure 7D).
The third explanation of scaling mitigation by a super-
hydrophobic membrane is the slip boundary condition for
liquid ow along the membrane surface (Figure 7E). The slip
boundary condition may have two major impacts on scale
formation.
113
The rst impact is the reduced residence time for
crystal growth and interaction with a superhydrophobic
membrane surface as compared to that for a ow along a
hydrophobic membrane with a no-slip condition (Figure 7F).
The second impact is the higher ow velocity and turbulence
intensication near the surface of a superhydrophobic
membrane. This second eect may reduce the oversaturation
level near the membrane surface by promoting better mixing to
minimize the concentration polarization. It may also aid in
dislodging precipitated crystals or even preventing their
deposition.
117
The rst and second explanations of scaling mitigation
concern the free energy of a system, which is essentially a
thermodynamic argument. In comparison, the third contribu-
tion is a hydrodynamic argument. While all these explanations
are mechanistically reasonable, to what extent each explanation
contributes to the overall eect of scaling mitigation by
superhydrophobic membrane is practically dicult to quantify.
Operation Strategies. Besides pretreatment and the use of
superhydrophobic membranes, various operation strategies can
also reduce mineral scaling in MD. These strategies include
water ushing,
133
microbubble aeration,
157
electrophoretic
mixing,
158
ow/temperature reversal,
159
and MD integrated
with a crystallizer,
116,160167
which hinder scaling by either
removing scalants from the feedwater or disrupting the
nucleation process. For example, periodic ushing using
deionized water has been shown to maintain a stable water
vapor ux of MD in gypsum-saturated feed solutions, probably
due to the removal of gypsum nuclei from the membrane
surface.
133
However, the use of deionized water (product of
MD) for cleaning is not desirable, and the eectiveness of this
approach diminishes as more water is recovered. In another
study, ow and temperature reversal between feed and
distillate streams was proved eective in mitigating mineral
scaling when using MD to treat hypersaline water from the
Great Salt Lake.
159
Furthermore, microbubbles have been also
used to eectively reduce vapor ux decline caused by salt
precipitation. The authors proposed that the negatively
charged microbubbles not only reduced the eect of
concentration polarization but also attracted divalent ions
(e.g., Ca2+ and Mg2+) at the waterair interface to reduce their
availability in the bulk solution for scale formation.
157
Scaling
mitigation can also be achieved by applying an alternating
current to the surface of electrically conducting membranes
fabricated via depositing a network of carbon nanotubes on a
hydrophobic membrane.
158
Such a strategy has been shown to
be eective for mitigating both gypsum and silica scaling, likely
due to the electrophoretic mixing that disrupts the
concentration polarization layer. Also, the generation of an
electrical double layer creates a concentration imbalance
Figure 8. (A) Top-down SEM image and (B) cross-section elemental mapping based on SEM with EDS for a hydrophobic membrane scaled with
gypsum. (C) Top-down SEM image and (D) cross-section elemental mapping based on SEM-EDS for a superhydrophobic membrane. In both
cases, the membranes were subject to intermittent back-purging of the same frequency, duration, and intensity. (E, F) Illustration of how (E) partial
pore intrusion into hydrophobic membrane or (F) the lack of it with superhydrophobic membrane inuences the eectiveness of scaling mitigation
by purging or pulse ow. Adapted with permission from ref 174.
ACS ES&T Engineering pubs.acs.org/estengg Review
https://dx.doi.org/10.1021/acsestengg.0c00025
ACS EST Engg. 2021, 1, 117140
128
between anions and cations, which also contributes to slowing
the scale formation.
158
Another integrated process that has the benet of pure water
recovery while simultaneously recovering valuable precipitates
is membrane distillation-crystallization (MDC). In MDC, MD
is used to concentrate the aqueous solution to the desired
supersaturation, while a crystallizer is used to precipitate and
remove mineral crystals from the solution. While the
integration of MD with a brine crystallizer may allow for an
enhanced water recovery ratio and delay scaling induction
time,
168,169
the MD membrane is still subjected to high
saturation solutions and, thus, eventually suers from mineral
scaling.
170,171
Backwashing is a common strategy for membrane cleaning
in microltration.
172,173
Because an MD membrane needs to
be maintained nonwetted, back-purging (with gas) can be used
instead of backwashing (with water). It has been found that
intermittently back-purging a hydrophobic membrane cannot
mitigate gypsum scaling due to the robust adhesion of the scale
layer to the hydrophobic membrane surface (Figure 8A).
However, when intermittent back-purging is combined with
superhydrophobic membranes, gypsum scaling can be virtually
eliminated (Figure 8 B), even though the use of a
superhydrophobic membrane alone without back-purging can
only delay gypsum scaling.
174
Back-purging a hydrophobic
membrane was ineective, because the partial intrusion of the
feedwater into the pores promotes very robust adhesion of the
formed crystal to the membrane (Figure 8C,E). In comparison,
the minimum solidliquid contact area and lack of pore
intrusion in the case of a superhydrophobic membrane makes
back-purging eective (Figure 8D,F).
In a more recent study, a similar approach integrating
superhydrophobic membrane and pulse ow (i.e., ow with a
variable hydraulic pressure) was shown to be also eective in
eliminating gypsum scaling.
175
Again, neither pulse ow with
hydrophobic membrane nor superhydrophobic membrane
alone without pulse ow was eective in mitigating gypsum
scaling. While pulse ow is likely more practical than back-
purging in a real operation, both studies reveal that the same
importance of synergy between the material and operation
strategies in mitigating gypsum scaling. Regardless of the
specic operation approach, recharging the air layer on the
surface of a superhydrophobic membrane seems to be critical
to sustain long-term MD performance against gypsum scaling.
Other approaches toward this goal of air-layer recharging,
including the direct injection of nanobubbles or the in situ
generation of bubbles via electrolysis, may be explored. Finally,
it is not entirely clear why a superhydrophobic membrane
alone, without back-purging or pulse ow, can eectively
eliminate scaling by NaCl but not gypsum. We can only
postulate at this point that the crystal morphology plays an
important role. Elucidating the impact of scaling type on the
eectiveness of dierent scaling mitigation approaches is
important for developing strategies that are universally eective
for real feedwater with complex compositions.
FOULING
Fouling is a common problem to all membrane processes and
is an umbrella term that can include organic fouling, inorganic
fouling, and biological fouling. In MD processes, mineral
scaling has been referred to as inorganic fouling in some cases.
However, inorganic fouling in other membrane processes often
involves only inorganic particles that originally exist in the
feedwater, excluding mineral scaling. This type of inorganic
fouling (sometimes referred to as colloidal fouling) is not a
major concern in MD or MC, as inorganic particles can be
readily removed using simple pretreatments. Therefore, we will
focus on organic and biological fouling in this discussion.
Generally speaking, fouling is a membrane failure common
in both MD and MD applications, where species in the feed
solution accumulate on the membrane surface and block its
pores, reducing the ux of water vapor (in MD) and volatile
component (in MC), and it may eventually lead to pore
wetting.
124,176
From an operational perspective, the low
operating pressure in MC and MD reduces the propensity to
form a dense, compact, irreversible foulant layer as compared
to pressure-driven membrane processes such as RO or
NF.
177,178
From a material perspective, however, MD and
MC membranes are inherently prone to organic fouling due to
the long-range hydrophobic interaction between the hydro-
phobic membrane and many common hydrophobic organic
foulants.
179,180
The most common method for detecting membrane fouling
is by simply monitoring the ux of water vapor (in MD) or
volatile species (in MC), which declines as the membrane
becomes fouled. Several studies have also used electro-
impedance spectroscopy to detect organic fouling in
situ.
181184
In brief, the working mechanism is that the foulant
layer adds electrical resistance to the membrane surface. This
new detection technique allows for the elucidation of the
mechanisms and time scales of organic fouling, which in turn
enables early fouling detection, more ecient operation, and
more eective fouling mitigation in practice.
Organic foulants relevant to MD and MC include proteins,
humic acids, and, in some cases, emulsied oil droplets. Oil
fouling is relevant to MD, because MD has been actively
explored for treating oil- and gas-produced water.
4,73,152
Biological fouling (or just biofouling) is caused by the growth
of bacteria, fungi, and algae, especially by the formation of
biolm. Technically, organic foulants are also found in
biofouling, as substances from microbes, including proteins
and extracellular polymeric substances, are organics. In this
section we will review the mechanisms in which these foulants
interact with the membrane surfaces, making the distinction
between the saline conditions typical to MD and lack thereof
in MC. In particular, we will discuss the eects of solution
composition on fouling mechanisms and how these mecha-
nisms apply to each of the individual foulant categories. Finally,
we will review the common, and most eective, mitigation
strategies that have shown promise for practical MC and MD
applications.
Mechanism of Organic Fouling. In general, organic
fouling is a phenomenon where organic colloids interact with a
membrane surface in an attractive manner and accumulate to
cover the membrane surface. Colloidal interaction is often
modeled by the DerjaguinLandauVerweyOverbeek
(DLVO) theory, which considers the relative contributions
of the electric double layer (EDL) and van der Waals (vdW)
interactions. The EDL interaction, which is repulsive for two
similarly charged interacting surfaces, is strongly dependent on
the ionic strength. In the context of MD, which is most
promising for desalinating hypersaline brine, the ionic strength
is so high that no interaction energy barrier should exist,
because the vdW contribution always outcompetes the EDL
contribution. Therefore, it may be argued that the DLVO
theory is less relevant to MD due to the lack of interaction
ACS ES&T Engineering pubs.acs.org/estengg Review
https://dx.doi.org/10.1021/acsestengg.0c00025
ACS EST Engg. 2021, 1, 117140
129
energy barrier in any practical condition. The DLVO theory
may be relevant to understanding fouling propensity in the
context of MC, where salinity (and thus ionic strength) can be
quite low. Even so, the DLVO theory should only be used to
provide qualitative reasoning for the observed phenomena but
should not be used as a predictive tool to quantify the fouling
propensity. Specically, a microporous hydrophobic membrane
is dierent in many aspects from the idealized impermeable
surface in a particle deposition to which the DLVO theory has
been applied.
Another important interaction that is strongly relevant to the
organic fouling of hydrophobic membranes is the long-ranged,
hydrophobic interaction.
179,180,185
The hydrophobic interac-
tion is associated with the congurational rearrangement of
water molecules and tends to be an order of magnitude
stronger and longer-ranged than the vDW interactions,
decaying exponentially with distance.
179
The exact mechanism
of the widely observed hydrophobic interaction has long been
debated. One probable explanation is the capillary bridging
between two surfaces by an air bubble that tend to develop on
a hydrophobic surface.
186,187
With this mechanism, a hydro-
phobic interaction is all the more relevant to MD and MC, not
only because hydrophobic membranes are used but also
because an air layer is intrinsically present in a nonwetted
membrane in a CassieBaxter state. Considering the
substantially high strength of the hydrophobic interaction
and the heavily shielded EDL interaction in high ionic
strength, we argue that the long-ranged hydrophobic
interaction dominates the membrane fouling process, making
hydrophobic organic foulants very troublesome for sustaining
long-term MD or MC performance.
188,189
This is particularly
true for oil foulants that are not heavily emulsied by excessive
surfactants.
Mechanism of Biofouling. Biofouling pertains to both
MC and MD when feedwater containing a high content of
biological microorganisms, such as those in the wastewaters of
the pharmaceutical, food and beverage industries, and
municipal wastewater treatment facilities, is in direct contact
with the membrane surface.
27,190
The high-salinity and high-
temperature conditions of the feed solution in MD may lead
one to expect a low potential for biofouling.
191
However, there
are bacteria and other microorganisms that are resistant to high
salinity and temperature,
192
and moderate temperature can
even promote the growth of thermophilic bacteria.
176,193,194
While the mechanism of biofouling is complicated due to the
presence of multiple foulant species with dierent physico-
chemical properties, it is important to discuss the general
process and time scales of biofouling.
Biofouling is a particularly unwelcome fouling phenomenon,
because it would lead to the formation of a large gel-like
structure on the membrane surface. Such a structure, namely
biolm,is dicult to remove physically or via chemical
cleaning.
195
The growth of biolm is typically characterized by
several unique stages that develop consecutively
195
(Figure 9).
In the early stages of biofouling, a conditioning lm develops,
where organic foulants such as proteins, humic substances, and
other organic matter deposit on the membrane surface in a lm
with thickness on the order of nanometers.
195,196
The
conditioning lm from organic fouling serves as a precursor
for further foulants attachment, as it alters the surface
properties exposed to the feed solution.
197200
Next, several
attractive interactions, including vdW, hydrophobic, and
hydrogen-bonding interactions, result in reversible bacteria
cell attachment, while larger aggregate particles and protein
aggregates (namely, protobiolms) irreversibly attach to the
conditioning lm.
197
These two processes occur on the time
scale of seconds to minutes. After a few hours, the bacteria cells
bind irreversibly to the surface. This irreversible attachment is
due to the reinforcement of the extracellular polymeric
substances (EPS), which are produced within the microbe
colonies of the biolm and mostly consist of protein,
polysaccharides, some humic substances, and
DNA.
176,193,195,201
The kinetic growth and mechanical properties of the EPS are
heavily inuenced by the salinity and specic ionic species.
202
For example, the presence of divalent cations such as calcium
and magnesium is known to reinforce the biolm by forming
complexes with the polysaccharides in the EPS matrix, forming
a more dense, interconnected network.
195,201,203
Furthermore,
divalent cations in the feedwater aid the initial adhesion of
organic foulants, such as humic substances, by forming
complexes in the bulk that settle more easily and deposit on
the membrane surface.
134,204206
After several hours up to
days, the biolm is fully reinforced by the constant secretion of
Figure 9. Dierent stages of biolm formation and the respective time scales. Dissolved and colloidal organics deposit onto the surface and form a
conditioning lm to which bacteria adhere. Protobiolm and bacteria then directly deposit onto the conditioning lm. These two processes occur
on the time scale of minutes to seconds, but eventually the EPS excreted from the bacteria enhances the cell attachment and facilitates the
formation of a robust and continuous biolm. The biolm eventually becomes a source from which bacteria and bacterial clusters are dispersed.
Adapted with permission from ref 196.
ACS ES&T Engineering pubs.acs.org/estengg Review
https://dx.doi.org/10.1021/acsestengg.0c00025
ACS EST Engg. 2021, 1, 117140
130
EPS and can even begin to spread and grow with the
colonization of new areas on the membrane by living bacteria
from within the biolm.
195
Fouling Mitigation Strategies. Pretreatments and
Membrane Cleaning. Well-established physical and chemical
pretreatments can be used to reduce the concentration of
foulants in the feed solution. Physical pretreatments are often
very eective in mitigating membrane fouling. For example,
pretreatment with nanoltration or ultraltration very
eectively limits ux decline due to fouling.
33,36,42
Coagu-
lation/occulation is a highly cost-eective approach to
remove foulants from the feed solution by forming ocs that
enmesh the foulants. These ocs can be removed by
sedimentation, ltration through porous media, or micro-
ltration.
5,151,207
For example, coagulation pretreatment of
recirculating cooling water with poly(aluminum chloride) has
shown to improve elimination of total organic carbon and
reduce the water vapor ux decline due to organic fouling in
MD.
151
Other methods that may also be eective include
otation (especially for oil foulants) and activated carbon
adsorption. A recent study using MD to treat real produced
water from the shale oil and gas industry showed that
precipitative softening/coagulation using aluminum sulfate
followed by walnut shell ltration can very eectively reduce
fouling in MD.
152
We expect that pretreatment methods or
treatment trains that have been proven eective for fouling
mitigation in other membrane processes are also eective for
MD and that the choice of pretreatments for fouling mitigation
depends largely on the feedwater characteristics rather than the
membrane processes.
In addition to pretreatment, there are additional chemical
and physical strategies that are eective in removing foulants
that have already attached to membrane surfaces. For example,
the use of micro/nanobubbles via direct injection into the feed
Figure 10. Comparison between in-air (top) and underwater (bottom) wetting properties for (A) hydrophobic membrane, (B) omniphobic
membrane, and (C) composite membrane with a hydrophobic substrate and a hydrophilic coating.
Figure 11. Illustration of the eectiveness of dierent types of membranes with special wetting properties in mitigating wetting, fouling, and scaling.
A Janus (h) membrane is a membrane with a hydrophobic (thus h) substrate and an in-air hydrophilic and underwater oleophobic coating
(denoted by the red line in the gure). A Janus (o) membrane is a membrane with an omniphobic (thus o) substrate and an in-air hydrophilic
and underwater oleophobic coating. An omniphobic-slippery membrane is both omniphobic and superhydrophobic (i.e., slippery). The superscript
*for F(representing fouling) indicates the uncertainty or limited information for the eectiveness evaluation, primarily because of the multiple
types of fouling that have not been comprehensively assessed for the respective membranes. These eectiveness ratings are based on the following
studies: superhydrophobic membranes,
95,96,112,113,153,154,174
omniphobic membranes,
28,7382
Janus (h) membranes,
38,188,189,219,220,235237
Janus (o)
membranes,
219,238
and omniphobic-slippery membranes.
239
ACS ES&T Engineering pubs.acs.org/estengg Review
https://dx.doi.org/10.1021/acsestengg.0c00025
ACS EST Engg. 2021, 1, 117140
131
solution
157,208211
or using ultrasound or electrolysis to induce
bubble nucleation near the membrane surface
212215
has been
shown to eectively limit and remove organic foulants.
Generally speaking, the bubbles both physically dislodge
foulants from the membrane surface and remove foulants
from the feedwater by capturing them at the airliquid
interface of the bubbles (similar to otation).
215217
In a study
where membrane biofouling occurred in the MD desalination
of seawater from the South China Sea, ultrasonic cleaning
eectively removed the organic foulants, and transmembrane
water vapor ux was restored to its initial rate.
213
Other
common membrane cleaning methods, including those using
cleaning agents based on alkaline, acidic, surfactant, and
chelating agents, have also been shown to be eective to
dierent extents for removing the foulant layer.
23,218
While pretreatment methods that are eective for organic
fouling mitigation can also reduce biofouling to a certain
degree via reducing the microbe concentration of the feed
solution, methods for cleaning organic fouling may not be
eective for alleviating biofouling, especially after biolms have
formed on the membrane surface. The most eective approach
for preventing biofouling or cleaning a biolm once it forms is
to apply disinfectants (e.g., chlorination).
178
Because most
disinfectants are also strong oxidants, disinfection is not
applicable for RO or NF membranes that are primarily based
on a polyamide that would suer a performance loss after
prolonged exposure to oxidants.
178
For MD, however, applying
disinfectants can be a viable approach, because hydrophobic
membranes, especially those made from uoropolymers, are
typically chemically stable.
Fouling-Resistant Membranes. Many recent studies on
MD fouling are focused on oil fouling, because it presents a
critical challenge when MD is used to desalinate oil- and gas-
produced water. Hydrophobic MD membranes are inherently
prone to oil fouling due to the strong hydrophobic interaction.
Oil fouling is fatal to a hydrophobic membrane, because it is
dicult to be reversed due to pore wicking by oil (Figure
10A). It may be expected that an omniphobic membrane is
resistant to oil fouling, because an omniphobic surface is, by
denition, both hydrophobic and oleophobic. However, it is
important to emphasize that all wetting properties, without a
prex descriptor, are dened in air. An omniphobic surface
(and membrane) is in-air oleophobic but not underwater
oleophobic. In fact, an omniphobic membrane is underwater
oleophilic,
38
because the low surface energy of the material
required for fabricating an omniphobic membrane leads to a
strong hydrophobic interaction just as that between oil
droplets and a hydrophobic membrane. However, an
omniphobic membrane diers from a hydrophobic membrane
by having a reentrant structure that prevents the oil droplets
from wicking into the pores (Figure 11B). In other words,
while oil droplets can still foul an omniphobic membrane by
covering the membrane surface and blocking pores for vapor
transfer, the fouling can be reversed by cleaning due to the
absence of pore wicking. Interestingly, if the oil-in-water
emulsion is stabilized by a high concentration of surfactants, an
omniphobic membrane would not even be fouled, because
now both the oil droplets and the membrane surface are
rendered hydrophilic due to the adsorbed surfactants, thereby
eliminating the attractive hydrophobic interaction.
38
While the
same principle should apply to hydrophobic membranes,
hydrophobic membranes are susceptible to wetting by
surfactants.
On the basis of underwater wettability, a robust material
strategy to develop an MD or MC membrane resistant to oil
fouling is to apply a (in-air) hydrophilic coating layer on a
hydrophobic membrane surface
38,188,189,219,220
(Figure 10C).
An in-air hydrophilic surface is typically underwater
oleophobic due to the hydration layer that results from the
strong attractive interaction between water and the high
surface energy moieties of the hydrophilic coating. This
hydration layer prevents oil fouling, as the spreading of oil on
the membrane surface requires dehydration of the coating
layer, which is energetically unfavorable.
189
The mitigation of
fouling using such a hydration layer is more robust, as the
nonfoulingcondition is a thermodynamically stable state.
Finally, we note that in-air hydrophilicity is only a necessary,
but not a sucient, condition for underwater oleophobicity.
To develop a robust underwater oleophobic surface, the eects
of electrostatic interaction between oil and surface must also be
considered. For example, while a negatively charged (in-air)
hydrophilic surface is robustly resistant to fouling by negatively
charged oil droplets,
221,222
a positively charged (in-air)
hydrophilic surface has shown to be underwater oleophobic
only upon its initial contact with a solution containing
negatively charged oil droplets. Over the time scale of tens
of minutes, the positively charged (in-air) hydrophilic becomes
increasingly oleophilic and wetted by the oil droplets of
opposite charge.
221
Expectedly, the MD membrane with a
negatively charged hydrophilic coating is substantially more
robust in sustaining a stable MD performance than that with a
positively charged hydrophilic coating, because the former is
robustly underwater oleophobic, while the latter is not.
Ultimately, it is the underwater wetting property that matters
for oil-fouling resistance.
In general, this rationale also applies to the mitigation of
organic foulants other than oil droplets, such as NOM and
proteins. Hydrophilic surfaces with identical charge to the
organic foulants in question are more robust in mitigating
fouling due to the presence of a hydration layer and
electrostatic repulsion.
223231
This is especially true for organic
foulants with hydrophobic moieties such as proteins. Some-
what paradoxical to the observation that an in-air hydrophilic,
underwater oleophobic surface imparts robust fouling resist-
ance, some studies have also reported that superhydrophobic
membranes are also eective for mitigating fouling by organics
such as humic substances.
232234
Similar to the mechanisms
for scaling mitigation, mitigating organic fouling with a
superhydrophobic membrane is more attributable to decreased
liquidsolid contact between the feed solution and membrane
surface. Because there is still some small fraction of liquid
solid contact, long-term biofouling experiments show a small
ux decline, as the humic substances eventually attach to the
surface where there is that liquidsolid contact.
232,233
OUTLOOK
Despite the several unique advantages of MD in desalinating
hypersaline brine, MD also faces multiple unique challenges as
compared to pressure-driven membrane processes that have
been scaled up for practical applications. For example, virtually
all sweet spotapplications for MD involve concentrating
high-salinity streams, which inherently has a higher propensity
for mineral scaling than other membrane processes treating
feedwater with low to moderate salinity.
1,6
Furthermore, pore
wetting is a technical challenge that is relevant to MD (or MC)
but not to any other membrane processes. Lastly, while fouling
ACS ES&T Engineering pubs.acs.org/estengg Review
https://dx.doi.org/10.1021/acsestengg.0c00025
ACS EST Engg. 2021, 1, 117140
132
is common in all membrane processes, the use of a
hydrophobic membrane and the presence of an air layer in
MD processes enhances the hydrophobic interaction and,
thereby, increases the propensity of fouling by hydrophobic
organics.
189
Nonetheless, the absence of an applied hydraulic
pressure in MD prevents the formation of a very dense foulant
layer, which may alleviate irreversible fouling.
177,178
Any of
these three problems, scaling, wetting, and fouling, if not
properly addressed, can substantially compromise the perform-
ance or even lead to a complete process failure of MD (or
MC).
Fortunately, the rich collective experience in other
membrane processes can help us address the problems of
fouling and scaling using the proper measures of pretreatment,
feedwater conditioning, (e.g., dosing antiscalants or disinfec-
tants), and membrane cleaning. These approaches, though not
scientically novel, are often economical and eective over a
broad range of foulants or scalants. Other process innovations,
such as the use of micro/nanobubbles, intermittent back-
purging, and pulse ow, have also been proven eective in
enhancing the robustness of the MD performance.
157,174,175
However, these approaches have their limitations. For example,
the challenge of pore wetting is dicult to overcome by
removing wetting agents from the feedwater due to the very
limited capability of achieving such a separation cost-
eectively. In this case, developing novel MD membranes
with special wettability provides an alternative path for
addressing the challenge of pore wetting. For fouling and
scaling, developing better membranes is also an important
supplement to, or even a substitute of, pretreatments and
operational measures. In certain cases, the material and
operational approaches are synergistic and do not function
eectively without each other.
Thanks to the recent development of novel membranes with
special wetting properties, we now have eective solutions to
each of the three major challenges in MD: wetting, scaling, and
fouling. The general rules of thumb are that (1) omniphobic
membranes can prevent wetting, (2) superhydrophobic
membranes are eective in mitigating mineral scaling, and
(3) Janus membranes (i.e., composite membranes) with an in-
air hydrophilic and underwater oleophobic coating are capable
of reducing fouling, in particular, oil fouling. However, none of
these membranes have been proven to be a robust and
universal solution to all three challenges. To engineer
resistance against multiple failure mechanisms, it is possible
to integrate two types of wetting properties into one composite
membrane. For example, a Janus (o) membrane, which
comprises an omniphobic substrate and a fouling-resistant,
hydrophilic surface coating, has been shown to be eective in
mitigating both wetting and fouling.
219
In another example, a
membrane that is both omniphobic and superhydrophobic
(also known as slippery) has been shown to be simultaneously
wetting and scaling resistant.
239
These general principles are
summarized in Figure 11.
The eectiveness of dierent membranes in mitigating
organic and biological fouling is less clear-cut, as dierent types
of foulants behave very dierently, and not all types of fouling
(e.g., oil, natural organic matter, and biological foulants) have
been tested with membranes with dierent types of wetting
properties. Membranes that are resistant to fouling by natural
organic matter may not be resistant to oil fouling, and vice
versa. Moreover, even the oil-fouling resistance of the same
membrane also depends on whether and to what extent the oil
droplets are stabilized by surfactants. When oil droplets are
stabilized by excessive surfactants, the membrane may fail via
the mechanism of wetting instead of fouling.
38
In some cases, all types of failure mechanisms in MD have
been referred to as fouling. For example, a membrane might
be referred to as anti-fouling, while it was actually tested
against a variety of agents that would induce wetting, scaling,
and fouling (of the narrower denition).
9,36,240
While putting
all these mechanisms under the umbrella of fouling may sound
pragmatically convenient from an operation point of view (it
essentially becomes a proxy term for membrane failure),
doing so obscures the fundamental dierence behind these
mechanisms and is unconstructive for the systematic under-
standing of membrane failure and the development of
mitigation strategies. In other cases, however, the membrane
fails via multiple mechanisms concurrently, and distinction
between them is inherently dicult, because one type of failure
can induce the other. For example, it has been reported that
both organic fouling and mineral scaling can result in pore
wetting.
37
In the case of gypsum scaling, wetting is caused by
pore deformation instead of the reduced surface tension of the
feed solution due to the presence of surfactants or LST
liquids.
35
In addition to more clearly distinguishing between dierent
failure mechanisms, we also need to be more mindful and
precise in categorizing a membrane with special wettability.
For instance, a superhydrophobic membrane can be both
oleophobic and oleophilic. When a study shows that a
superhydrophobic membrane is wetting resistant without
actually measuring the oil-wetting property of the membrane,
the conclusion can be confusingthe wetting resistance may
likely be attributable to the oleophobicity imparted by the re-
entrant structure instead of the superhydrophobicity imparted
by the high degree of roughness.
Finally, although a limited number of MD studies was
performed using real feedwater such as RO brines or industrial
wastewater, a majority of reported studies on addressing the
challenges of wetting, scaling, and fouling in MD, especially
those related to novel membrane development, used simple
feed solutions. Most studies either focused on a single type of
membrane failure or investigated multiple mechanisms
separately. While these studies are important to enhancing
our fundamental understanding, practical MD processes often
involve more complex feed solutions and the synergy of
dierent failure mechanisms. Examples include, but are not
limited to, scaling by dierent types of minerals, simultaneous
mineral scaling and organic fouling, and simultaneous fouling
and wetting. Therefore, future research eorts on pretreatment,
operation, or membrane development should be more directed
toward understanding the combined eects of multiple failure
mechanisms using feed solutions with more complex
compositions. Such eorts will potentially bridge the gap
between fundamental understanding and engineering practice,
thereby enabling MD to become a reliable process for treating
various types of hypersaline brine.
AUTHOR INFORMATION
Corresponding Authors
Tiezheng Tong Department of Civil and Environmental
Engineering, Colorado State University, Fort Collins, Colorado
80523, United States; orcid.org/0000-0002-9289-3330;
Email: tiezheng.tong@colostate.edu
ACS ES&T Engineering pubs.acs.org/estengg Review
https://dx.doi.org/10.1021/acsestengg.0c00025
ACS EST Engg. 2021, 1, 117140
133
Shihong Lin Department of Chemical and Biomolecular
Engineering and Department of Civil and Environmental
Engineering, Vanderbilt University, Nashville, Tennessee 37235-
1831, United States; orcid.org/0000-0001-9832-9127;
Email: shihong.lin@vanderbilt.edu
Authors
Thomas Horseman Department of Chemical and
Biomolecular Engineering, Vanderbilt University, Nashville,
Tennessee 37235-1831, United States; orcid.org/0000-
0002-4660-1448
Yiming Yin Department of Civil and Environmental
Engineering, Colorado State University, Fort Collins, Colorado
80523, United States
KoSS Christie Department of Civil and Environmental
Engineering, Vanderbilt University, Nashville, Tennessee 37235-
1831, United States; orcid.org/0000-0002-7039-7889
Zhangxin Wang Department of Chemical and Environmental
Engineering, Yale University, New Haven, Connecticut 06511,
United States
Complete contact information is available at:
https://pubs.acs.org/10.1021/acsestengg.0c00025
Notes
The authors declare no competing nancial interest.
ACKNOWLEDGMENTS
The authors are grateful for the generous support by the
National Science Foundation via Grant Nos. 1739884 (S.L.),
1705048 (T.H. and Z.W.), and DGE-1145194 (K.S.S.C.) and
by the Bureau of Reclamation (USBR) under the Department
of Interior via DWPR Agreement R18AC00108 (T.T.).
REFERENCES
(1) Deshmukh, A.; Boo, C.; Karanikola, V.; Lin, S.; Straub, A. P.;
Tong, T.; Warsinger, D. M.; Elimelech, M. Membrane Distillation at
the Water-Energy Nexus: Limits, Opportunities, and Challenges.
Energy Environ. Sci. 2018,11 (5), 11771196.
(2) Lawson, K. W.; Lloyd, D. R. Membrane Distillation. J. Membr.
Sci. 1997,124 (1), 125.
(3) Alkhudhiri, A.; Darwish, N.; Hilal, N. Membrane Distillation: A
Comprehensive Review. Desalination 2012,287,218.
(4) Shaffer, D. L.; Arias Chavez, L. H.; Ben-Sasson, M.; Romero-
Vargas Castrillón, S.; Yip, N. Y.; Elimelech, M. Desalination and
Reuse of High-Salinity Shale Gas Produced Water: Drivers,
Technologies, and Future Directions. Environ. Sci. Technol. 2013,47
(17), 95699583.
(5) Sardari, K.; Fyfe, P.; Lincicome, D.; Ranil Wickramasinghe, S.
Combined Electrocoagulation and Membrane Distillation for
Treating High Salinity Produced Waters. J. Membr. Sci. 2018,564,
8296.
(6) Tong, T.; Elimelech, M. The Global Rise of Zero Liquid
Discharge for Wastewater Management: Drivers, Technologies, and
Future Directions. Environ. Sci. Technol. 2016,50 (13), 68466855.
(7) Choi, Y.; Naidu, G.; Nghiem, L. D.; Lee, S.; Vigneswaran, S.
Membrane Distillation Crystallization for Brine Mining and Zero
Liquid Discharge: Opportunities, Challenges, and Recent Progress.
Environ. Sci. Water Res. Technol. 2019,5(7), 12021221.
(8) Lin, S. Energy Efficiency of Desalination: Fundamental Insights
from Intuitive Interpretation. Environ. Sci. Technol. 2019,54 (1), 76
84.
(9) Tijing, L. D.; Woo, Y. C.; Choi, J. S.; Lee, S.; Kim, S. H.; Shon,
H. K. Fouling and Its Control in Membrane Distillation-A Review. J.
Membr. Sci. 2015,475, 215244.
(10) Christie, K. S. S.; Horseman, T.; Lin, S. Energy Efficiency of
Membrane Distillation: Simplified Analysis, Heat Recovery, and the
Use of Waste-Heat. Environ. Int. 2020,138, 105588.
(11) Lin, S.; Yip, N. Y.; Elimelech, M. Direct Contact Membrane
Distillation with Heat Recovery: Thermodynamic Insights from
Module Scale Modeling. J. Membr. Sci. 2014,453, 498515.
(12) Camacho, L. M.; Dumée, L.; Zhang, J.; Li, J. De; Duke, M.;
Gomez, J.; Gray, S. Advances in Membrane Distillation for Water
Desalination and Purification Applications. Water 2013,5(1), 94
196.
(13) Gingerich, D. B.; Mauter, M. S. Quantity, Quality, and
Availability of Waste Heat from United States Thermal Power
Generation. Environ. Sci. Technol. 2015,49 (14), 82978306.
(14) Burgess, G.; Lovegrove, K. Solar Thermal Powered
Desalination: Membrane Versus Distillation TechnologiesSolar 2005.
(15) Qiblawey, H. M.; Banat, F. Solar Thermal Desalination
Technologies. Desalination 2008,220, 633644.
(16) Wang, Z.; Horseman, T.; Straub, A. P.; Yip, N. Y.; Li, D.;
Elimelech, M.; Lin, S. Pathways and Challenges for Ecient Solar-
Thermal Desalination. Sci. Adv. 2019,5(7). .
(17) Curcio, E.; Drioli, E. Membrane Distillation and Related
Operations - A Review. Sep. Purif. Rev. 2005,34 (1), 3586.
(18) Rongwong, W.; Lee, J.; Goh, K.; Karahan, H. E.; Bae, T.-H.
Membrane-Based Technologies for Post-Treatment of Anaerobic
Euentsnpj Clean Water201811DOI: 10.1038/s41545-018-0021-y
(19) Yu, X.; An, L.; Yang, J.; Tu, S. T.; Yan, J. CO2 Capture Using a
Superhydrophobic Ceramic Membrane Contactor. J. Membr. Sci.
2015,496,112.
(20) Lv, Y.; Yu, X.; Jia, J.; Tu, S. T.; Yan, J.; Dahlquist, E. Fabrication
and Characterization of Superhydrophobic Polypropylene Hollow
Fiber Membranes for Carbon Dioxide Absorption. Appl. Energy 2012,
90 (1), 167174.
(21) Wang, R.; Zhang, H. Y.; Feron, P. H. M.; Liang, D. T. Influence
of Membrane Wetting on CO2 Capture in Microporous Hollow Fiber
Membrane Contactors. Sep. Purif. Technol. 2005,46 (12), 3340.
(22) Shao, J.; Liu, H.; He, Y. Boiler Feed Water Deoxygenation
Using Hollow Fiber Membrane Contactor. Desalination 2008,234
(13), 370377.
(23) Henares, M.; Ferrero, P.; San-Valero, P.; Martínez-Soria, V.;
Izquierdo, M. Performance of a Polypropylene Membrane Contactor
for the Recovery of Dissolved Methane from Anaerobic Effluents:
Mass Transfer Evaluation, Long-Term Operation and Cleaning
Strategies. J. Membr. Sci. 2018,563, 926937.
(24) Rongwong, W.; Goh, K.; Sethunga, G. S. M. D. P.; Bae, T. H.
Fouling Formation in Membrane Contactors for Methane Recovery
from Anaerobic Effluents. J. Membr. Sci. 2019,573, 534543.
(25) Darestani, M.; Haigh, V.; Couperthwaite, S. J.; Millar, G. J.;
Nghiem, L. D. Hollow Fibre Membrane Contactors for Ammonia
Recovery: Current Status and Future Developments. J. Environ. Chem.
Eng. 2017,5(2), 13491359.
(26) Gryta, M.; Tomaszewska, M.; Grzechulska, J.; Morawski, a. W.
Membrane Distillation of NaCl Solution Containing Natural Organic
Matter. J. Membr. Sci. 2001,181 (2), 279287.
(27) Hausmann, A.; Sanciolo, P.; Vasiljevic, T.; Weeks, M.; Schroën,
K.; Gray, S.; Duke, M. Fouling of Dairy Components on Hydrophobic
Polytetrafluoroethylene (PTFE) Membranes for Membrane Distil-
lation. J. Membr. Sci. 2013,442, 149159.
(28) Lin, S.; Nejati, S.; Boo, C.; Hu, Y.; Osuji, C. O.; Elimelech, M.
Omniphobic Membrane for Robust Membrane Distillation. Environ.
Sci. Technol. Lett. 2014,1(11), 443447.
(29) Wang, Z.; Chen, Y.; Sun, X.; Duddu, R.; Lin, S. Mechanism of
Pore Wetting in Membrane Distillation with Alcohol vs. Surfactant. J.
Membr. Sci. 2018,559, 183195.
(30) Chew, N. G. P.; Zhao, S.; Loh, C. H.; Permogorov, N.; Wang,
R. Surfactant Effects on Water Recovery from Produced Water via
Direct-Contact Membrane Distillation. J. Membr. Sci. 2017,528,
126134.
ACS ES&T Engineering pubs.acs.org/estengg Review
https://dx.doi.org/10.1021/acsestengg.0c00025
ACS EST Engg. 2021, 1, 117140
134
(31) Wang, Z.; Chen, Y.; Lin, S. Kinetic Model for Surfactant-
Induced Pore Wetting in Membrane Distillation. J. Membr. Sci. 2018,
564, 275288.
(32) Wang, Z.; Chen, Y.; Zhang, F.; Lin, S. Significance of Surface
Excess Concentration in the Kinetics of Surfactant-Induced Pore
Wetting in Membrane Distillation. Desalination 2019,450,4653.
(33) Karakulski, K.; Gryta, M. Water Demineralisation by NF/MD
Integrated Processes. Desalination 2005,177 (13), 109119.
(34) He, F.; Gilron, J.; Lee, H.; Song, L.; Sirkar, K. K. Potential for
Scaling by Sparingly Soluble Salts in Crossflow DCMD. J. Membr. Sci.
2008,311 (12), 6880.
(35) Christie, K. S. S.; Yin, Y.; Lin, S.; Tong, T. Distinct Behaviors
between Gypsum and Silica Scaling in Membrane Distillation.
Environ. Sci. Technol. 2019,54 (1), 568576.
(36) Gryta, M. Fouling in Direct Contact Membrane Distillation
Process. J. Membr. Sci. 2008,325 (1), 383394.
(37) Warsinger, D. M.; Swaminathan, J.; Guillen-Burrieza, E.; Arafat,
H. A.; Lienhard V, J. H. Scaling and Fouling in Membrane Distillation
for Desalination Applications: A Review. Desalination 2015,356,
294313.
(38) Wang, Z.; Lin, S. Membrane Fouling and Wetting in Membrane
Distillation and Their Mitigation by Novel Membranes with Special
Wettability. Water Res. 2017,112,3847.
(39) Guillen-Burrieza, E.; Mavukkandy, M. O.; Bilad, M. R.; Arafat,
H. A. Understanding Wetting Phenomena in Membrane Distillation
and How Operational Parameters Can Affect It. J. Membr. Sci. 2016,
515, 163174.
(40) Guillen-Burrieza, E.; Ruiz-Aguirre, A.; Zaragoza, G.; Arafat, H.
A. Membrane Fouling and Cleaning in Long Term Plant-Scale
Membrane Distillation Operations. J. Membr. Sci. 2014,468, 360
372.
(41) Rezaei, M.; Warsinger, D. M.; Lienhard V, J. H.; Duke, M. C.;
Matsuura, T.; Samhaber, W. M. Wetting Phenomena in Membrane
Distillation: Mechanisms, Reversal, and Prevention. Water Res. 2018,
139, 329352.
(42) Gryta, M. Long-Term Performance of Membrane Distillation
Process. J. Membr. Sci. 2005,265 (12), 153159.
(43) Lu, J. G.; Zheng, Y. F.; Cheng, M. D. Wetting Mechanism in
Mass Transfer Process of Hydrophobic Membrane Gas Absorption. J.
Membr. Sci. 2008,308 (12), 180190.
(44) Gabelman, A.; Hwang, S. T. Hollow Fiber Membrane
Contactors. J. Membr. Sci. 1999,159 (12), 61106.
(45) Kreulen, H.; Smolders, C. A.; Versteeg, G. F.; Van Swaaij, W. P.
M. Determination of Mass Transfer Rates in Wetted and Non-Wetted
Microporous Membranes. Chem. Eng. Sci. 1993,48 (11), 20932102.
(46) Gryta, M.; Barancewicz, M. Influence of Morphology of PVDF
Capillary Membranes on the Performance of Direct Contact
Membrane Distillation. J. Membr. Sci. 2010,358 (12), 158167.
(47) Chen, Y.; Wang, Z.; Jennings, G. K.; Lin, S. Probing Pore
Wetting in Membrane Distillation Using Impedance: Early Detection
and Mechanism of Surfactant-Induced Wetting. Environ. Sci. Technol.
Lett. 2017,4(11), 505510.
(48) Jacob, P.; Dejean, B.; Laborie, S.; Cabassud, C. An Optical In-
Situ Tool for Visualizing and Understanding Wetting Dynamics in
Membrane Distillation. J. Membr. Sci. 2020,595, 117587.
(49) Tuberquia, J. C.; Song, W. S.; Jennings, G. K. Investigating the
Superhydrophobic Behavior for Underwater Surfaces Using Impe-
dance-Based Methods. Anal. Chem. 2011,83 (16), 61846190.
(50) Tuberquia, J. C.; Nizamidin, N.; Jennings, G. K. Effect of
Superhydrophobicity on the Barrier Properties of Polymethylene
Films. Langmuir 2010,26 (17), 1403914046.
(51) Franken, A. C. M.; Nolten, J. A. M.; Mulder, M. H. V;
Bargeman, D.; Smolders, C. A. Wetting Criteria for the Applicability
of Membrane Distillation. J. Membr. Sci. 1987,33 (3), 315328.
(52) García-Payo, M. C.; Izquierdo-Gil, M. A.; Fernández-Pineda, C.
Wetting Study of Hydrophobic Membranes via Liquid Entry Pressure
Measurements with Aqueous Alcohol Solutions. J. Colloid Interface Sci.
2000,230, 420431.
(53) Jacob, P.; Laborie, S.; Cabassud, C. Visualizing and Evaluating
Wetting in Membrane Distillation: New Methodology and Indicators
Based on Detection of Dissolved Tracer Intrusion (DDTI).
Desalination 2018,443, 307322.
(54) Zmievskii, Y. G. Determination of Critical Pressure in
Membrane Distillation Process. Pet. Chem. 2015,55 (4), 308314.
(55) Churaev, N. V.; Martynov, G. A.; Starov, V. M.; Zorin, Z. M.
Some Features of Capillary Imbibition of Surfactant Solutions. Colloid
Polym. Sci. 1981,259, 747752.
(56) Starov, V. M.; Zhdanov, S. A.; Velarde, M. G. Capillary
Imbibition of Surfactant Solutions in Porous Media and Thin
Capillaries: Partial Wetting Case. J. Colloid Interface Sci. 2004,273
(2), 589595.
(57) Rosen, M. J.; Cohen, A. W.; Dahanayake, M.; Hua, X. Y.
Relationship of Structure to Properties in Surfactants. 10. Surface and
Thermodynamic Properties of 2-Dodecyloxypoly(Ethenoxyethanol)s,
C12H25(OC2H4)XOH, in Aqueous Solution. J. Phys. Chem. 1982,
86 (4), 541545.
(58) Dong, B.; Li, N.; Zheng, L.; Yu, L.; Inoue, T. Surface
Adsorption and Micelle Formation of Surface Active Ionic Liquids in
Aqueous Solution. Langmuir 2007,23 (8), 41784182.
(59) Adamson, A. W.; Gast, A. P. Physical Chemistry of Surfaces. Z.
Phys. Chem. 1967,150, 134.
(60) Tan, Y. Z.; Velioglu, S.; Han, L.; Joseph, B. D.; Unnithan, L. G.;
Chew, J. W. Effect of Surfactant Hydrophobicity and Charge Type on
Membrane Distillation Performance. J. Membr. Sci. 2019,587,
117168.
(61) Milne, A. J. B.; Amirfazli, A. Autophilic Effect: Wetting of
Hydrophobic Surfaces by Surfactant Solutions. Langmuir 2010,26
(7), 46684674.
(62) Kumar, N.; Varanasi, K.; Tilton, R. D.; Garoff, S. Surfactant
Self-Assembly Ahead of the Contact Line on a Hydrophobic Surface
and Its Implications for Wetting. Langmuir 2003,19 (13), 5366
5373.
(63) Bailey, A. F. G.; Barbe, A. M.; Hogan, P. A.; Johnson, R. A.;
Sheng, J. The Effect of Ultrafiltration on the Subsequent
Concentration of Grape Juice by Osmotic Distillation. J. Membr.
Sci. 2000,164 (12), 195204.
(64) Zarebska, A.; Amor, Á. C.; Ciurkot, K.; Karring, H.; Thygesen,
O.; Andersen, T. P.; Hägg, M. B.; Christensen, K. V.; Norddahl, B.
Fouling Mitigation in Membrane Distillation Processes during
Ammonia Stripping from Pig Manure. J. Membr. Sci. 2015,484,
119132.
(65) Kowalska, I. Nanofiltration -Ion Exchange System for Effective
Surfactant Removal from Water Solutions. Braz. J. Chem. Eng. 2014,
31 (4), 887894.
(66) Beltrán-Heredia, J.; Sánchez-Martín, J.; Barrado-Moreno, M.
Long-Chain Anionic Surfactants in Aqueous Solution. Removal by
Moringa Oleifera Coagulant. Chem. Eng. J. 2012,180, 128136.
(67) Aboulhassan, M. A.; Souabi, S.; Yaacoubi, A.; Baudu, M.
Removal of Surfactant from Industrial Wastewaters by Coagulation
Flocculation Process. Int. J. Environ. Sci. Technol. 2006,3(4), 327
332.
(68) Chen, S.; Timmons, M. B.; Bisogni, J. J.; Aneshansley, D. J.
Modeling Surfactant Removal in Foam Fractionation: I - Theoretical
Development. Aquac. Eng. 1994,13 (3), 163181.
(69) Lee, J.; Maa, J. R. Separation of a Surface Active Solute by
Foam Fractionation. Int. Commun. Heat Mass Transfer 1986,13, 465
473.
(70) Dow, N.; Villalobos García, J.; Niadoo, L.; Milne, N.; Zhang, J.;
Gray, S.; Duke, M. Demonstration of Membrane Distillation on
Textile Waste Water Assessment of Long Term Performance,
Membrane Cleaning and Waste Heat Integration. Environ. Sci.
Water Res. Technol. 2017,3(3), 433449.
(71) Camacho-Muñoz, D.; Martín, J.; Santos, J. L.; Aparicio, I.;
Alonso, E. Occurrence of Surfactants in Wastewater: Hourly and
Seasonal Variations in Urban and Industrial Wastewaters from Seville
(Southern Spain). Sci. Total Environ. 2014,468469, 977984.
ACS ES&T Engineering pubs.acs.org/estengg Review
https://dx.doi.org/10.1021/acsestengg.0c00025
ACS EST Engg. 2021, 1, 117140
135
(72) Chen, H. J.; Tseng, D. H.; Huang, S. L. Biodegradation of
Octylphenol Polyethoxylate Surfactant Triton X-100 by Selected
Microorganisms. Bioresour. Technol. 2005,96 (13), 14831491.
(73) Boo, C.; Lee, J.; Elimelech, M. Omniphobic Polyvinylidene
Fluoride (PVDF) Membrane for Desalination of Shale Gas Produced
Water by Membrane Distillation. Environ. Sci. Technol. 2016,50 (22),
1227512282.
(74) Chul Woo, Y.; Chen, Y.; Tijing, L. D.; Phuntsho, S.; He, T.;
Choi, J. S.; Kim, S. H.; Kyong Shon, H. CF4 Plasma-Modified
Omniphobic Electrospun Nanofiber Membrane for Produced Water
Brine Treatment by Membrane Distillation. J. Membr. Sci. 2017,529,
234242.
(75) Lee, J.; Boo, C.; Ryu, W. H.; Taylor, A. D.; Elimelech, M.
Development of Omniphobic Desalination Membranes Using a
Charged Electrospun Nanofiber Scaffold. ACS Appl. Mater. Interfaces
2016,8(17), 1115411161.
(76) Li, J.; Guo, S.; Xu, Z.; Li, J.; Pan, Z.; Du, Z.; Cheng, F.
Preparation of Omniphobic PVDF Membranes with Silica Nano-
particles for Treating Coking Wastewater Using Direct Contact
Membrane Distillation: Electrostatic Adsorption vs. Chemical
Bonding. J. Membr. Sci. 2019,574, 349357.
(77) Chen, L. H.; Huang, A.; Chen, Y. R.; Chen, C. H.; Hsu, C. C.;
Tsai, F. Y.; Tung, K. L. Omniphobic Membranes for Direct Contact
Membrane Distillation: Effective Deposition of Zinc Oxide Nano-
particles. Desalination 2018,428, 255263.
(78) Lu, K. J.; Zuo, J.; Chang, J.; Kuan, H. N.; Chung, T. S.
Omniphobic Hollow-Fiber Membranes for Vacuum Membrane
Distillation. Environ. Sci. Technol. 2018,52 (7), 44724480.
(79) Lu, C.; Su, C.; Cao, H.; Ma, X.; Duan, F.; Chang, J.; Li, Y. F-
POSS Based Omniphobic Membrane for Robust Membrane
Distillation. Mater. Lett. 2018,228,8588.
(80) Zheng, R.; Chen, Y.; Wang, J.; Song, J.; Li, X. M.; He, T.
Preparation of Omniphobic PVDF Membrane with Hierarchical
Structure for Treating Saline Oily Wastewater Using Direct Contact
Membrane Distillation. J. Membr. Sci. 2018,555, 197205.
(81) Deng, L.; Ye, H.; Li, X.; Li, P.; Zhang, J.; Wang, X.; Zhu, M.;
Hsiao, B. S. Self-Roughened Omniphobic Coatings on Nanofibrous
Membrane for Membrane Distillation. Sep. Purif. Technol. 2018,206,
1425.
(82) Lu, K. J.; Chen, Y.; Chung, T. S. Design of Omniphobic
Interfaces for Membrane Distillation A Review. Water Res. 2019,
162,6477.
(83) Tuteja, A.; Choi, W.; Ma, M.; Mabry, J. M.; Mazzella, S. a;
Rutledge, G. C.; McKinley, G. H.; Cohen, R. E. Designing
Superoleophobic Surfaces. Science 2007,318 (5856), 16181622.
(84) Tuteja, A.; Choi, W.; Mabry, J. M.; McKinley, G. H.; Cohen, R.
E. Robust Omniphobic Surfaces. Proc. Natl. Acad. Sci. U. S. A. 2008,
105 (47), 1820018205.
(85) Dufour, R.; Harnois, M.; Thomy, V.; Boukherroub, R.; Senez,
V. Contact Angle Hysteresis Origins: Investigation on Super-
Omniphobic Surfaces. Soft Matter 2011,7(19), 93809387.
(86) Chhatre, S. S.; Choi, W.; Tuteja, A.; Park, K. C.; Mabry, J. M.;
McKinley, G. H.; Cohen, R. E. Scale Dependence of Omniphobic
Mesh Surfaces. Langmuir 2010,26 (6), 40274035.
(87) Boo, C.; Lee, J.; Elimelech, M. Engineering Surface Energy and
Nanostructure of Microporous Films for Expanded Membrane
Distillation Applications. Environ. Sci. Technol. 2016,50 (15),
81128119.
(88) Wang, W.; Du, X.; Vahabi, H.; Zhao, S.; Yin, Y.; Kota, A. K.;
Tong, T. Trade-off in Membrane Distillation with Monolithic
Omniphobic Membranes. Nat. Commun. 2019,10 (1), 19.
(89) Liu, T.; Kim, C.-J. Turning a Surface Superrepellent Even to
Completely Wetting Liquids. Science 2014,346 (6213), 10961100.
(90) Yao, M.; Tijing, L. D.; Naidu, G.; Kim, S. H.; Matsuyama, H.;
Fane, A. G.; Shon, H. K. A Review of Membrane Wettability for the
Treatment of Saline Water Deploying Membrane Distillation.
Desalination 2020,479, 114312.
(91) Burton, F.; Tsuchihasi, R.; Tchobanoglous, G.; Stensel, H. D.
Wastewater Engineering: Treatment and Resource Recovery, 5th ed.;
McGraw-Hill Education, 2013.
(92) Gryta, M. Studies of Membrane Scaling during Water
Desalination by Membrane Distillation. Chem. Pap. 2019,73 (3),
591600.
(93) McGaughey, A. L.; Gustafson, R. D.; Childress, A. E. Effect of
Long-Term Operation on Membrane Surface Characteristics and
Performance in Membrane Distillation. J. Membr. Sci. 2017,543,
143150.
(94) Ramezanianpour, M.; Sivakumar, M. An Analytical Flux
Decline Model for Membrane Distillation. Desalination 2014,345,
112.
(95) Karanikola, V.; Boo, C.; Rolf, J.; Elimelech, M. Engineered
Slippery Surface to Mitigate Gypsum Scaling in Membrane
Distillation for Treatment of Hypersaline Industrial Wastewaters.
Environ. Sci. Technol. 2018,52 (24), 1436214370.
(96) Su, C.; Horseman, T.; Cao, H.; Christie, K.; Li, Y.; Lin, S.
Robust Superhydrophobic Membrane for Membrane Distillation with
Excellent Scaling Resistance. Environ. Sci. Technol. 2019,53 (20),
1180111809.
(97) Xie, M.; Gray, S. R. Gypsum Scaling in Forward Osmosis: Role
of Membrane Surface Chemistry. J. Membr. Sci. 2016,513, 250259.
(98) Fortunato, L.; Jang, Y.; Lee, J. G.; Jeong, S.; Lee, S.; Leiknes, T.
O.; Ghaffour, N. Fouling Development in Direct Contact Membrane
Distillation: Non-Invasive Monitoring and Destructive Analysis.
Water Res. 2018,132,3441.
(99) Lee, J.-G.; Jang, Y.; Fortunato, L.; Jeong, S.; Lee, S.; Leiknes, T.;
Ghaffour, N. An Advanced Online Monitoring Approach to Study the
Scaling Behavior in Direct Contact Membrane Distillation. J. Membr.
Sci. 2018,546,5060.
(100) Lokare, O. R.; Ji, P.; Wadekar, S.; Dutt, G.; Vidic, R. D.
Concentration Polarization in Membrane Distillation: I. Development
of a Laser-Based Spectrophotometric Method for in-Situ Character-
ization. J. Membr. Sci. 2019,581, 462471.
(101)Gibbs,J.W.OntheEquilibriumofHeterogeneous
Substances. Am. J. Sci. 1879,2, 300320.
(102) Tong, T.; Wallace, A. F.; Zhao, S.; Wang, Z. Mineral Scaling in
Membrane Desalination: Mechanisms, Mitigation Strategies, and
Feasibility of Scaling-Resistant Membranes. J. Membr. Sci. 2019,579,
5269.
(103) Tong, T.; Zhao, S.; Boo, C.; Hashmi, S. M.; Elimelech, M.
Relating Silica Scaling in Reverse Osmosis to Membrane Surface
Properties. Environ. Sci. Technol. 2017,51 (8), 43964406.
(104) Giuffre, A. J.; Hamm, L. M.; Han, N.; De Yoreo, J. J.; Dove, P.
M. Polysaccharide Chemistry Regulates Kinetics of Calcite
Nucleation through Competition of Interfacial Energies. Proc. Natl.
Acad. Sci. U. S. A. 2013,110 (23), 92619266.
(105) Hamm, L. M.; Giuffre, A. J.; Han, N.; Tao, J.; Wang, D.; De
Yoreo, J. J.; Dove, P. M. Reconciling Disparate Views of Template-
Directed Nucleation through Measurement of Calcite Nucleation
Kinetics and Binding Energies. Proc. Natl. Acad. Sci. U. S. A. 2014,111
(4), 13041309.
(106) Rahardianto, A.; Mccool, B. C.; Cohen, Y. Reverse Osmosis
Desalting of Inland Brackish Water of High Gypsum Scaling
Propensity: Kinetics and Mitigation of Membrane Mineral Scaling.
Environ. Sci. Technol. 2008,42 (12), 42924297.
(107) Wu, W.; Nancollas, G. H. Interfacial Free Energies and
Crystallization in Aqueous Media. J. Colloid Interface Sci. 1996,182
(2), 365373.
(108) Van Oss, C. J.; Chaudhury, M. K.; Good, R. J. Interfacial
Lifshitz-van Der Waals and Polar Interactions in Macroscopic
Systems. Chem. Rev. 1988,88 (6), 927941.
(109) Israelachvili, J. Intermolecular and Surface Forces, 3rd ed.;
Science Direct, 2010. DOI: 10.1016/B978-0-12-375182-9.10025-9.
(110) Chibowski, E.; Holysz, L. Use of the Washburn Equation for
Surface Free Energy Determination. Langmuir 1992,8(2), 710716.
ACS ES&T Engineering pubs.acs.org/estengg Review
https://dx.doi.org/10.1021/acsestengg.0c00025
ACS EST Engg. 2021, 1, 117140
136
(111) Teng, F.; Zeng, H.; Liu, Q. Understanding the Deposition and
Surface Interactions of Gypsum. J. Phys. Chem. C 2011,115 (35),
1748517494.
(112) Zheng, R.; Yin, H.; Liu, Y.; He, H.; Zhang, Y.; Li, X.-M.; Ji, Y.;
Xiao, Z.; Yuan, X.; He, T.; Li, D. Slippery for Scaling Resistance in
Membrane Distillation: A Novel Porous Micropillared Super-
hydrophobic Surface. Water Res. 2019,155, 152161.
(113) Xiao, Z.; Li, Z.; Guo, H.; Liu, Y.; Wang, Y.; Yin, H.; Li, X.;
Song, J.; Nghiem, L. D.; He, T. Scaling Mitigation in Membrane
Distillation: From Superhydrophobic to Slippery. Desalination 2019,
466,3643.
(114) Martínez-Díez, L.; Vázquez-González, M. Temperature and
Concentration Polarization in Membrane Distillation of Aqueous Salt
Solutions. J. Membr. Sci. 1999,156 (2), 265273.
(115) Gilron, J.; Ladizansky, Y.; Korin, E. Silica Fouling in Direct
Contact Membrane Distillation. Ind. Eng. Chem. Res. 2013,52 (31),
1052110529.
(116) Tun, C. M.; Fane, A. G.; Matheickal, J. T.; Sheikholeslami, R.
Membrane Distillation Crystallization of Concentrated Salts Flux
and Crystal Formation. J. Membr. Sci. 2005,257, 144155.
(117) Warsinger, D. M.; Tow, E. W.; Swaminathan, J.; Lienhard V, J.
H. Theoretical Framework for Predicting Inorganic Fouling in
Membrane Distillation and Experimental Validation with Calcium
Sulfate. J. Membr. Sci. 2017,528, 381390.
(118) Fein, J. B.; Walther, J. V. Calcite Solubility in Supercritical
CO2H2O Fluids. Geochim. Cosmochim. Acta 1987,51 (6), 1665
1673.
(119) Azimi, G. Evaluating the Potential of Scaling Due to Calcium
Compounds in Hydrometallurgical Processes, Ph.D. Thesis; University of
Toronto, 2010.
(120) He, S.; Oddo, J. E.; Tomson, M. B. The Nucleation Kinetics of
Calcium Sulfate Dihydrate in NaCl Solutions up to 6 m and 90°C. J.
Colloid Interface Sci. 1994,162 (2), 297303.
(121) Yin, Y.; Wang, W.; Kota, A. K.; Zhao, S.; Tong, T. Elucidating
Mechanisms of Silica Scaling in Membrane Distillation: Effects of
Membrane Surface Wettability. Environ. Sci. Water Res. Technol. 2019,
5(11), 20042014.
(122) Warsinger, C.; Martin, D.; Swaminathan, J.; Chung, H. W.;
Jeong, S.; Lienhard, J. H.; Martin Warsinger, D.; Warsinger, D. M.
Effect of Filtration and Particulate Fouling in Membrane Distillation.
International Desalination Association 2015,014.
(123) Barbe, A. M.; Hogan, P. A.; Johnson, R. A. Surface
Morphology Changes during Initial Usage of Hydrophobic, Micro-
porous Polypropylene Membranes. J. Membr. Sci. 2000,172 (12),
149156.
(124) Gryta, M. Influence of Polypropylene Membrane Surface
Porosity on the Performance of Membrane Distillation Process. J.
Membr. Sci. 2007,287 (1), 6778.
(125) Andrés-Mañas, J. A.; Ruiz-Aguirre, A.; Acién, F. G.; Zaragoza,
G. Assessment of a Pilot System for Seawater Desalination Based on
Vacuum Multi-Effect Membrane Distillation with Enhanced Heat
Recovery. Desalination 2018,443, 110121.
(126) Yu, X.; Yang, H.; Lei, H.; Shapiro, A. Experimental Evaluation
on Concentrating Cooling Tower Blowdown Water by Direct
Contact Membrane Distillation. Desalination 2013,323, 134141.
(127) Peng, Y.; Ge, J.; Li, Z.; Wang, S. Effects of Anti-Scaling and
Cleaning Chemicals on Membrane Scale in Direct Contact
Membrane Distillation Process for RO Brine Concentrate. Sep.
Purif. Technol. 2015,154,2226.
(128) He, F.; Sirkar, K. K.; Gilron, J. Effects of Antiscalants to
Mitigate Membrane Scaling by Direct Contact Membrane Distillation.
J. Membr. Sci. 2009,345,5358.
(129) Qu, F.; Yan, Z.; Yu, H.; Fan, G.; Pang, H.; Rong, H.; He, J.
Effect of Residual Commercial Antiscalants on Gypsum Scaling and
Membrane Wetting during Direct Contact Membrane Distillation.
Desalination 2020,486, 114493.
(130) Minier-Matar, J.; Hussain, A.; Janson, A.; Benyahia, F.;
Adham, S. Field Evaluation of Membrane Distillation Technologies
for Desalination of Highly Saline Brines. Desalination 2014,351,
101108.
(131) Hou, D.; Dai, G.; Wang, J.; Fan, H.; Luan, Z.; Fu, C. Boron
Removal and Desalination from Seawater by PVDF Flat-Sheet
Membrane through Direct Contact Membrane Distillation. Desalina-
tion 2013,326, 115124.
(132) Gryta, M. Polyphosphates Used for Membrane Scaling
Inhibition during Water Desalination by Membrane Distillation.
Desalination 2012,285, 170176.
(133) Zhang, P.; Knötig, P.; Gray, S.; Duke, M. Scale Reduction and
Cleaning Techniques during Direct Contact Membrane Distillation of
Seawater Reverse Osmosis Brine. Desalination 2015,374,2030.
(134) Gryta, M.; Tomaszewska, M.; Karakulski, K. Wastewater
Treatment by Membrane Distillation. Desalination 2006,198 (13),
6773.
(135) Zhang, P.; Pabstmann, A.; Gray, S.; Duke, M. Silica Fouling
during Direct Contact Membrane Distillation of Coal Seam Gas Brine
with High Sodium Bicarbonate and Low Hardness. Desalination 2018,
444, 107117.
(136) Sun, X.; Zhang, J.; Yin, C.; Zhang, J.; Han, J. Poly(Aspartic
Acid)-Tryptophan Grafted Copolymer and Its Scale-Inhibition
Performance. J. Appl. Polym. Sci. 2015,132 (45), 29.
(137) Duan, W.; Oota, H.; Sawada, K. Stability and Structure of
Ethylenedinitrilopoly(Methylphosphonate) Complexes of the Alka-
line-Earth Metal Ions in Aqueous Solution. J. Chem. Soc., Dalton
Trans. 1999,493 (17), 30753080.
(138) Deluchat, V.; Bollinger, J. C.; Serpaud, B.; Caullet, C. Divalent
Cations Speciation with Three Phosphonate Ligands in the PH-Range
of Natural Waters. Talanta 1997,44 (5), 897907.
(139) Lioliou, M. G.; Paraskeva, C. A.; Koutsoukos, P. G.; Payatakes,
A. C. Calcium Sulfate Precipitation in the Presence of Water-Soluble
Polymers. J. Colloid Interface Sci. 2006,303 (1), 164170.
(140) Battaglia, G.; Crea, F.; Crea, P.; Sammartano, S. The
Protonation of Polyacrylate in Seawater. Analysis of Concentration
Effects. Ann. Chim. 2005,95 (910), 643656.
(141) Liu, Z.; Wang, S.; Zhang, L.; Liu, Z. Dynamic Synergistic Scale
Inhibition Performance of IA/SAS/SHP Copolymer with Magnetic
Field and Electrostatic Field. Desalination 2015,362,2633.
(142) Ahmed, S. B.; Tlili, M. M.; Amor, M. B. Influence of a
Polyacrylate Antiscalant on Gypsum Nucleation and Growth. Cryst.
Res. Technol. 2008,43 (9), 935942.
(143) Shih, W. Y.; Albrecht, K.; Glater, J.; Cohen, Y. A Dual-Probe
Approach for Evaluation of Gypsum Crystallization in Response to
Antiscalant Treatment. Desalination 2004,169 (3), 213221.
(144) Le Gouellec, Y. A.; Elimelech, M. Control of Calcium Sulfate
(Gypsum) Scale in Nanofiltration of Saline Agricultural Drainage
Water. Environ. Eng. Sci. 2002,19 (6), 387397.
(145) Jain, T.; Sanchez, E.; Owens-Bennett, E.; Trussell, R.; Walker,
S.; Liu, H. Impacts of Antiscalants on the Formation of Calcium
Solids: Implication on Scaling Potential of Desalination Concentrate.
Environ. Sci. Water Res. Technol. 2019,5(7), 12851294.
(146) Shih, W. Y.; Gao, J.; Rahardianto, A.; Glater, J.; Cohen, Y.;
Gabelich, C. J. Ranking of Antiscalant Performance for Gypsum Scale
Suppression in the Presence of Residual Aluminum. Desalination
2006,196 (13), 280292.
(147) Rabizadeh, T.; Morgan, D. J.; Peacock, C. L.; Benning, L. G.
Effectiveness of Green Additives vs Poly(Acrylic Acid) in Inhibiting
Calcium Sulfate Dihydrate Crystallization. Ind. Eng. Chem. Res. 2019,
58 (4), 15611569.
(148) Oh, H. J.; Choung, Y. K.; Lee, S.; Choi, J. S.; Hwang, T. M.;
Kim, J. H. Scale Formation in Reverse Osmosis Desalination: Model
Development. Desalination 2009,238 (13), 333346.
(149) Weijnen, M. P. C.; Van Rosmalen, G. M. Adsorption of
Phosphonates on Gypsum Crystals. J. Cryst. Growth 1986,79 (13),
157168.
(150) Rosenberg, Y. O.; Reznik, I. J.; Zmora-Nahum, S.; Ganor, J.
The Effect of PH on the Formation of a Gypsum Scale in the
Presence of a Phosphonate Antiscalant. Desalination 2012,284, 207
220.
ACS ES&T Engineering pubs.acs.org/estengg Review
https://dx.doi.org/10.1021/acsestengg.0c00025
ACS EST Engg. 2021, 1, 117140
137
(151) Wang, J.; Qu, D.; Tie, M.; Ren, H.; Peng, X.; Luan, Z. Effect of
Coagulation Pretreatment on Membrane Distillation Process for
Desalination of Recirculating Cooling Water. Sep. Purif. Technol.
2008,64 (1), 108115.
(152) Zhang, Z.; Du, X.; Carlson, K. H.; Robbins, C. A.; Tong, T.
Effective Treatment of Shale Oil and Gas Produced Water by
Membrane Distillation Coupled with Precipitative Softening and
Walnut Shell Filtration. Desalination 2019,454,8290.
(153) Yang, C.; Li, X. M.; Gilron, J.; Kong, D.-F.; Yin, Y.; Oren, Y.;
Linder, C.; He, T. CF4 Plasma-Modified Superhydrophobic PVDF
Membranes for Direct Contact Membrane Distillation. J. Membr. Sci.
2014,456, 155161.
(154) Zhang, W.; Li, Y.; Liu, J.; Li, B.; Wang, S. Fabrication of
Hierarchical Poly (Vinylidene Fluoride) Micro/Nano-Composite
Membrane with Anti-Fouling Property for Membrane Distillation. J.
Membr. Sci. 2017,535, 258267.
(155) Milne, A. J. B.; Amirfazli, A. The Cassie Equation: How It Is
Meant to Be Used. Adv. Colloid Interface Sci. 2012,170 (12), 48
55.
(156) Meng, S.; Ye, Y.; Mansouri, J.; Chen, V. Crystallization
Behavior of Salts during Membrane Distillation with Hydrophobic
and Superhydrophobic Capillary Membranes. J. Membr. Sci. 2015,
473, 165176.
(157) Ye, Y.; Yu, S.; Hou, L.; Liu, B.; Xia, Q.; Liu, G.; Li, P.
Microbubble Aeration Enhances Performance of Vacuum Membrane
Distillation Desalination by Alleviating Membrane Scaling. Water Res.
2019,149, 588595.
(158) Rao, U.; Iddya, A.; Jung, B.; Khor, C. M.; Hendren, Z.; Turchi,
C.; Cath, T.; Hoek, E. M. V.; Ramon, G. Z.; Jassby, D. Mineral Scale
Prevention on Electrically Conducting Membrane Distillation
Membranes Using Induced Electrophoretic Mixing. Environ. Sci.
Technol. 2020,54 (6), 36783690.
(159) Hickenbottom, K. L.; Cath, T. Y. Sustainable Operation of
Membrane Distillation for Enhancement of Mineral Recovery from
Hypersaline Solutions. J. Membr. Sci. 2014,454, 426435.
(160) Zou, T.; Kang, G.; Zhou, M.; Li, M.; Cao, Y. Submerged
Vacuum Membrane Distillation Crystallization (S-VMDC) with
Turbulent Intensification for the Concentration of NaCl Solution.
Sep. Purif. Technol. 2019,211, 151161.
(161) Edwie, F.; Chung, T. S. Development of Simultaneous
Membrane Distillation-Crystallization (SMDC) Technology for
Treatment of Saturated Brine. Chem. Eng. Sci. 2013,98, 160172.
(162) Curcio, E.; Drioli, E. Membrane Distillation and Related
Operations - A Review. Sep. Purif. Rev. 2005,34 (1), 3586.
(163) Choi, Y.; Naidu, G.; Nghiem, L. D.; Lee, S.; Vigneswaran, S.
Membrane Distillation Crystallization for Brine Mining and Zero
Liquid Discharge: Opportunities, Challenges, and Recent Progress.
Environ. Sci. Water Res. Technol. 2019,5(7), 12021221.
(164) Anisi, F.; Thomas, K. M.; Kramer, H. J. M. Membrane-
Assisted Crystallization: Membrane Characterization, Modelling and
Experiments. Chem. Eng. Sci. 2017,158, 277286.
(165) Ruiz Salmón, I.; Luis, P. Membrane Crystallization via
Membrane Distillation. Chem. Eng. Process. 2018,123, 258271.
(166) Gryta, M. Chemical Pretreatment of Feed Water for
Membrane Distillation. Chem. Pap. 2008,62 (1), 100105.
(167) Luo, L.; Zhao, J.; Chung, T.-S. Integration of Membrane
Distillation (MD) and Solid Hollow Fiber Cooling Crystallization
(SHFCC) Systems for Simultaneous Production of Water and Salt
Crystals. J. Membr. Sci. 2018,564, 905915.
(168) Macedonio, F.; Curcio, E.; Drioli, E. Integrated Membrane
Systems for Seawater Desalination: Energetic and Exergetic Analysis,
Economic Evaluation, Experimental Study. Desalination 2007,203
(13), 260276.
(169) Quist-Jensen, C. A.; Macedonio, F.; Horbez, D.; Drioli, E.
Reclamation of Sodium Sulfate from Industrial Wastewater by Using
Membrane Distillation and Membrane Crystallization. Desalination
2017,401, 112119.
(170) Kim, J.; Kim, J.; Hong, S. Recovery of Water and Minerals
from Shale Gas Produced Water by Membrane Distillation
Crystallization. Water Res. 2018,129, 447459.
(171) Shin, Y.; Sohn, J. Mechanisms for Scale Formation in
Simultaneous Membrane Distillation Crystallization: Effect of Flow
Rate. J. Ind. Eng. Chem. 2016,35, 318324.
(172) Chellam, S.; Jacangelo, J. G.; Bonacquisti, T. P. Modeling and
Experimental Verification of Pilot-Scale Hollow Fiber, Direct Flow
Microfiltration with Periodic Backwashing. Environ. Sci. Technol.
1998,32 (1), 7581.
(173) Su, S. K.; Liu, J. C.; Wiley, R. C. Cross-Flow Microfiltration
with Gas Backwash of Apple Juice. J. Food Sci. 1993,58 (3), 638641.
(174) Horseman, T.; Su, C.; Christie, K. S. S.; Lin, S. Highly
Effective Scaling Mitigation in Membrane Distillation Using a
Superhydrophobic Membrane with Gas Purging. Environ. Sci. Technol.
Lett. 2019,6(7), 423429.
(175) Liu, Y.; Li, Z.; Xiao, Z.; Yin, H.; Li, X.; He, T. Synergy of
Slippery Surface and Pulse Flow: An Anti-Scaling Solution for Direct
Contact Membrane Distillation. J. Membr. Sci. 2020,603, 118035.
(176) Bogler, A.; Bar-Zeev, E. Membrane Distillation Biofouling:
Impact of Feedwater Temperature on Biofilm Characteristics and
Membrane Performance. Environ. Sci. Technol. 2018,52 (17), 10019
10029.
(177) Kwan, S. E.; Bar-Zeev, E.; Elimelech, M. Biofouling in
Forward Osmosis and Reverse Osmosis: Measurements and
Mechanisms. J. Membr. Sci. 2015,493, 703708.
(178) Yoon, H.; Baek, Y.; Yu, J.; Yoon, J. Biofouling Occurrence
Process and Its Control in the Forward Osmosis. Desalination 2013,
325,3036.
(179) Israelachvili, J.; Pashley, R. The Hydrophobic Interaction Is
Long Range, Decaying Exponentially with Distance. Nature 1982,300
(5890), 341342.
(180) Meyer, E. E.; Rosenberg, K. J.; Israelachvili, J. Recent Progress
in Understanding Hydrophobic Interactions. Proc. Natl. Acad. Sci. U.
S. A. 2006,103 (43), 1573915746.
(181) Ho, J. S.; Sim, L. N.; Webster, R. D.; Viswanath, B.; Coster, H.
G. L.; Fane, A. G. Monitoring Fouling Behavior of Reverse Osmosis
Membranes Using Electrical Impedance Spectroscopy: A Field Trial
Study. Desalination 2017,407,7584.
(182) Kavanagh, J. M.; Hussain, S.; Chilcott, T. C.; Coster, H. G. L.
Fouling of Reverse Osmosis Membranes Using Electrical Impedance
Spectroscopy: Measurements and Simulations. Desalination 2009,236
(13), 187193.
(183) Ahmed, F. E.; Hilal, N.; Hashaikeh, R. Electrically Conductive
Membranes for in Situ Fouling Detection in Membrane Distillation
Using Impedance Spectroscopy. J. Membr. Sci. 2018,556,6672.
(184) Zhang, N.; Halali, M. A.; de Lannoy, C.-F. Detection of
Fouling on Electrically Conductive Membranes by Electrical
Impedance Spectroscopy. Sep. Purif. Technol. 2020,242, 116823.
(185) Tsao, Y.; Evans, D.; Wennerstrom, H. Long-Range Attractive
Force between Hydrophobic Surfaces Observed by Atomic Force
Microscopy. Science 1993,262 (5133), 547550.
(186) Wallqvist, V.; Claesson, P. M.; Swerin, A.; O
̈stlund, C.;
Schoelkopf, J.; Gane, P. A. C. Influence of Surface Topography on
Adhesive and Long-Range Capillary Forces between Hydrophobic
Surfaces in Water. Langmuir 2009,25 (16), 91979207.
(187) Hampton, M. A.; Nguyen, A. V. Nanobubbles and the
Nanobubble Bridging Capillary Force. Adv. Colloid Interface Sci. 2010,
154 (12), 3055.
(188) Hou, D.; Wang, Z.; Wang, K.; Wang, J.; Lin, S. Composite
Membrane with Electrospun Multiscale-Textured Surface for Robust
Oil-Fouling Resistance in Membrane Distillation. J. Membr. Sci. 2018,
546, 179187.
(189) Wang, Z.; Hou, D.; Lin, S. Composite Membrane with
Underwater-Oleophobic Surface for Anti-Oil-Fouling Membrane
Distillation. Environ. Sci. Technol. 2016,50 (7), 38663874.
(190) Qiu, H.; Peng, Y.; Ge, L.; Villacorta Hernandez, B.; Zhu, Z.
Pore Channel Surface Modification for Enhancing Anti-Fouling
Membrane Distillation. Appl. Surf. Sci. 2018,443, 217226.
ACS ES&T Engineering pubs.acs.org/estengg Review
https://dx.doi.org/10.1021/acsestengg.0c00025
ACS EST Engg. 2021, 1, 117140
138
(191) Vrouwenvelder, H. S.; van Paassen, J. A. M.; Folmer, H. C.;
Hofman, J. A. M. H.; Nederlof, M. M.; van der Kooij, D. Biofouling of
Membranes for Drinking Water Production. Desalination 1998,118
(13), 157166.
(192) Gryta, M. The Assessment of Microorganism Growth in the
Membrane Distillation System. Desalination 2002,142 (1), 7988.
(193) Krivorot, M.; Kushmaro, A.; Oren, Y.; Gilron, J. Factors
Affecting Biofilm Formation and Biofouling in Membrane Distillation
of Seawater. J. Membr. Sci. 2011,376 (12), 1524.
(194) Frock, A. D.; Kelly, R. M. Extreme Thermophiles: Moving
beyond Single-Enzyme Biocatalysis. Curr. Opin. Chem. Eng. 2012,1
(4), 363372.
(195) Bogler, A.; Lin, S.; Bar-Zeev, E. Biofouling of Membrane
Distillation, Forward Osmosis and Pressure Retarded Osmosis:
Principles, Impacts and Future Directions. J. Membr. Sci. 2017,542,
378398.
(196) Bar-Zeev, E.; Passow, U.; Romero-Vargas Castrillón, S.;
Elimelech, M. Transparent Exopolymer Particles: From Aquatic
Environments and Engineered Systems to Membrane Biofouling.
Environ. Sci. Technol. 2015,49 (2), 691707.
(197) Bar-Zeev, E.; Berman-Frank, I.; Girshevitz, O.; Berman, T.
Revised Paradigm of Aquatic Biofilm Formation Facilitated by
Microgel Transparent Exopolymer Particles. Proc. Natl. Acad. Sci. U.
S. A. 2012,109 (23), 91199124.
(198) Schneider, R. P.; Chadwick, B. R.; Jankowski, J.; Acworth, I.
Determination of Physicochemical Parameters of Solids Covered with
Conditioning Films from Groundwaters Using Contact Angles.
Comparative Analysis of Different Thermodynamic Approaches
Utilizing a Range of Diagnostic Liquids. Colloids Surf., A 1997,126
(1), 123.
(199) Phattaranawik, J.; Fane, A. G.; Pasquier, A. C. S.; Bing, W.;
Wong, F. S. Experimental Study and Design of a Submerged
Membrane Distillation Bioreactor. Chem. Eng. Technol. 2009,32 (1),
3844.
(200) De Kerchove, A. J.; Elimelech, M. Impact of Alginate
Conditioning Film on Deposition Kinetics of Motile and Nonmotile
Pseudomonas Aeruginosa Strains. Appl. Environ. Microbiol. 2007,73
(16), 52275234.
(201) Ferrando, D.; Toubiana, D.; Kandiyote, N. S.; Nguyen, T. H.;
Nejidat, A.; Herzberg, M. Ambivalent Role of Calcium in the
Viscoelastic Properties of Extracellular Polymeric Substances and the
Consequent Fouling of Reverse Osmosis Membranes. Desalination
2018,429,1219.
(202) Qiu, G.; Ting, Y. P. Short-Term Fouling Propensity and Flux
Behavior in an Osmotic Membrane Bioreactor for Wastewater
Treatment. Desalination 2014,332 (1), 9199.
(203) Fang, Y.; Al-Assaf, S.; Phillips, G. O.; Nishinari, K.; Funami,
T.; Williams, P. A.; Li, A. Multiple Steps and Critical Behaviors of the
Binding of Calcium to Alginate. J. Phys. Chem. B 2007,111 (10),
24562462.
(204) Qin, W.; Zhang, J.; Xie, Z.; Ng, D.; Ye, Y.; Gray, S. R.; Xie, M.
Synergistic Effect of Combined Colloidal and Organic Fouling in
Membrane Distillation: Measurements and Mechanisms. Environ. Sci.
Water Res. Technol. 2017,3(1), 119127.
(205) Shao, S.; Liang, H.; Qu, F.; Li, K.; Chang, H.; Yu, H.; Li, G.
Combined Influence by Humic Acid (HA) and Powdered Activated
Carbon (PAC) Particles on Ultrafiltration Membrane Fouling. J.
Membr. Sci. 2016,500,99105.
(206) Srisurichan, S.; Jiraratananon, R.; Fane, A. G. Humic Acid
Fouling in the Membrane Distillation Process. Desalination 2005,174,
6372.
(207) Zhao, K.; Sun, J.; Hu, C.; Qu, J. Membrane Fouling Reduction
through Electrochemically Regulating Flocs Aggregation in an
Electro-Coagulation Membrane Reactor. J. Environ. Sci. 2019,83,
144151.
(208) Etchepare, R.; Oliveira, H.; Azevedo, A.; Rubio, J. Separation
of Emulsified Crude Oil in Saline Water by Dissolved Air Flotation
with Micro and Nanobubbles. Sep. Purif. Technol. 2017,186, 326
332.
(209) Khuntia, S.; Majumder, S. K.; Ghosh, P. Microbubble-Aided
Water and Wastewater Purification: A Review. Rev. Chem. Eng. 2012,
28 (46), 191221.
(210) Agarwal, A.; Ng, W. J.; Liu, Y. Principle and Applications of
Microbubble and Nanobubble Technology for Water Treatment.
Chemosphere 2011,84 (9), 11751180.
(211) Agarwal, A.; Xu, H.; Ng, W. J.; Liu, Y. Biofilm Detachment by
Self-Collapsing Air Microbubbles: A Potential Chemical-Free
Cleaning Technology for Membrane Biofouling. J. Mater. Chem.
2012,22 (5), 22032207.
(212) Qasim, M.; Darwish, N. N.; Mhiyo, S.; Darwish, N. A.; Hilal,
N. The Use of Ultrasound to Mitigate Membrane Fouling in
Desalination and Water Treatment. Desalination 2018,443, 143164.
(213) Hsu, S. T.; Cheng, K. T.; Chiou, J. S. Seawater Desalination by
Direct Contact Membrane Distillation. Desalination 2002,143 (3),
279287.
(214) Hou, D.; Lin, D.; Zhao, C.; Wang, J.; Fu, C. Control of
Protein (BSA) Fouling by Ultrasonic Irradiation during Membrane
Distillation Process. Sep. Purif. Technol. 2017,175, 287297.
(215) Zhu, J.; An, H.; Alheshibri, M.; Liu, L.; Terpstra, P. M. J.; Liu,
G.; Craig, V. S. J. Cleaning with Bulk Nanobubbles. Langmuir 2016,
32 (43), 1120311211.
(216) Wu, Z. H.; Chen, H. B.; Dong, Y. M.; Mao, H. L.; Sun, J. L.;
Chen, S. F.; Craig, V. S. J.; Hu, J. Cleaning Using Nanobubbles:
Defouling by Electrochemical Generation of Bubbles. J. Colloid
Interface Sci. 2008,328 (1), 1014.
(217) Liu, G.; Wu, Z.; Craig, V. S. J. Cleaning of Protein-Coated
Surfaces Using Nanobubbles: An Investigation Using a Quartz Crystal
Microbalance. J. Phys. Chem. C 2008,112 (43), 1674816753.
(218) Ebrahim, S. Cleaning and Regeneration of Membranes in
Desalination and Wastewater Applications: State-of-the-Art. Desalina-
tion 1994,96 (13), 225238.
(219) Huang, Y. X.; Wang, Z.; Jin, J.; Lin, S. Novel Janus Membrane
for Membrane Distillation with Simultaneous Fouling and Wetting
Resistance. Environ. Sci. Technol. 2017,51 (22), 1330413310.
(220) Wang, Z.; Lin, S. The Impact of Low-Surface-Energy
Functional Groups on Oil Fouling Resistance in Membrane
Distillation. J. Membr. Sci. 2017,527,6877.
(221) Wang, Z.; Jin, J.; Hou, D.; Lin, S. Tailoring Surface Charge
and Wetting Property for Robust Oil-Fouling Mitigation in
Membrane Distillation. J. Membr. Sci. 2016,516, 113122.
(222) Ulbricht, M. Advanced Functional Polymer Membranes.
Polymer 2006,47 (7), 22172262.
(223) Boo, C.; Hong, S.; Elimelech, M. Relating Organic Fouling in
Membrane Distillation to Intermolecular Adhesion Forces and
Interfacial Surface Energies. Environ. Sci. Technol. 2018,52 (24),
1419814207.
(224) Wang, Y. N.; Tang, C. Y. Protein Fouling of Nanofiltration,
Reverse Osmosis, and Ultrafiltration Membranes-The Role of
Hydrodynamic Conditions, Solution Chemistry, and Membrane
Properties. J. Membr. Sci. 2011,376 (12), 275282.
(225) Van Wagner, E. M.; Sagle, A. C.; Sharma, M. M.; La, Y. H.;
Freeman, B. D. Surface Modification of Commercial Polyamide
Desalination Membranes Using Poly(Ethylene Glycol) Diglycidyl
Ether to Enhance Membrane Fouling Resistance. J. Membr. Sci. 2011,
367 (12), 273287.
(226) Choudhury, R. R.; Gohil, J. M.; Mohanty, S.; Nayak, S. K.
Antifouling, Fouling Release and Antimicrobial Materials for Surface
Modification of Reverse Osmosis and Nanofiltration Membranes. J.
Mater. Chem. A 2018,6(2), 313333.
(227) Miller, D. J.; Dreyer, D. R.; Bielawski, C. W.; Paul, D. R.;
Freeman,B.D.SurfaceModificationofWaterPurification
Membranes. Angew. Chem., Int. Ed. 2017,56 (17), 46624711.
(228) Liu, C.; Chen, L.; Zhu, L. Fouling Mechanism of
Hydrophobic Polytetrafluoroethylene (PTFE) Membrane by Differ-
ently Charged Organics during Direct Contact Membrane Distillation
(DCMD) Process: An Especial Interest in the Feed Properties. J.
Membr. Sci. 2018,548, 125135.
ACS ES&T Engineering pubs.acs.org/estengg Review
https://dx.doi.org/10.1021/acsestengg.0c00025
ACS EST Engg. 2021, 1, 117140
139
(229) Combe, C.; Molis, E.; Lucas, P.; Riley, R.; Clark, M. M. The
Effect of CA Membrane Properties on Adsorptive Fouling by Humic
Acid. J. Membr. Sci. 1999,154 (1), 7387.
(230) Kochkodan, V.; Hilal, N. A Comprehensive Review on Surface
Modified Polymer Membranes for Biofouling Mitigation. Desalination
2015,356, 187207.
(231) Li, F.; Meng, J.; Ye, J.; Yang, B.; Tian, Q.; Deng, C. Surface
Modification of PES Ultrafiltration Membrane by Polydopamine
Coating and Poly(Ethylene Glycol) Grafting: Morphology, Stability,
and Anti-Fouling. Desalination 2014,344, 422430.
(232) Razmjou, A.; Arifin, E.; Dong, G.; Mansouri, J.; Chen, V.
Superhydrophobic Modification of TiO2 Nanocomposite PVDF
Membranes for Applications in Membrane Distillation. J. Membr.
Sci. 2012,415416, 850863.
(233) Hou, D.; Christie, K. S. S.; Wang, K.; Tang, M.; Wang, D.;
Wang, J. Biomimetic Superhydrophobic Membrane for Membrane
Distillation with Robust Wetting and Fouling Resistance. J. Membr.
Sci. 2020,599, 117708.
(234) Shan, H.; Liu, J.; Li, X.; Li, Y.; Tezel, F. H.; Li, B.; Wang, S.
Nanocoated Amphiphobic Membrane for Flux Enhancement and
Comprehensive Anti-Fouling Performance in Direct Contact
Membrane Distillation. J. Membr. Sci. 2018,567, 166180.
(235) Zhu, Z.; Liu, Z.; Zhong, L.; Song, C.; Shi, W.; Cui, F.; Wang,
W. Breathable and Asymmetrically Superwettable Janus Membrane
with Robust Oil-Fouling Resistance for Durable Membrane
Distillation. J. Membr. Sci. 2018,563, 602609.
(236) Chew, N. G. P.; Zhao, S.; Malde, C.; Wang, R. Super-
oleophobic Surface Modification for Robust Membrane Distillation
Performance. J. Membr. Sci. 2017,541, 162173.
(237) Chew, N. G. P.; Zhang, Y.; Goh, K.; Ho, J. S.; Xu, R.; Wang,
R. Hierarchically Structured Janus Membrane Surfaces for Enhanced
Membrane Distillation Performance. ACS Appl. Mater. Interfaces
2019,11 (28), 2552425534.
(238) Li, C.; Li, X.; Du, X.; Tong, T.; Cath, T. Y.; Lee, J. Antiwetting
and Antifouling Janus Membrane for Desalination of Saline Oily
Wastewater by Membrane Distillation. ACS Appl. Mater. Interfaces
2019,11 (20), 1845618465.
(239) Chen, Y.; Lu, K. J.; Chung, T.-S. An Omniphobic Slippery
Membrane with Simultaneous Anti-Wetting and Anti-Scaling Proper-
ties for Robust Membrane Distillation. J. Membr. Sci. 2020,595,
117572.
(240) Lu, X.; Peng, Y.; Ge, L.; Lin, R.; Zhu, Z.; Liu, S. Amphiphobic
PVDF Composite Membranes for Anti-Fouling Direct Contact
Membrane Distillation. J. Membr. Sci. 2016,505,6169.
ACS ES&T Engineering pubs.acs.org/estengg Review
https://dx.doi.org/10.1021/acsestengg.0c00025
ACS EST Engg. 2021, 1, 117140
140
Article
Full-text available
Membrane wetting caused by substances in the feed solution with low surface tension has posed serious issues for membrane distillation (MD). Omniphobic membranes could be helpful when dealing with this problem because they strongly repel liquids with a wide range of surface tensions. We fabricated electrospun omniphobic membranes containing PVDF‐TEOS‐PFDTMS, PVDF‐HFP‐TMOS‐PFDTMS, and PVDF‐HFP‐TEOS‐PFDTMS. All electrospun membranes have shown contact angles >90°, but the electrospun membrane containing PVDF‐HFP‐TEOS‐PFDTMS has shown super repellence with a contact angle >150° and maintains stable salt rejection and water flux in direct contact MD processes in the presence of sodium dodecyl sulfate surfactant (up to 0.3 mM). Furthermore, the morphology and crystallinity of the electrospun membranes were studied by scanning electron microscopy and Fourier‐transform infrared spectroscopy, and the percentage composition of elements was analyzed by x‐ray photoelectron spectroscopy. liquid entry pressure, mechanical strength, and the porosity of the electrospun omniphobic membranes were also studied.
Article
The apparent tortuosity due to adsorption of diffusing tracers in a porous material is determined by a scaling approach and is used to analyze recent data on LiCl and alkane diffusion in mesoporous silica. The slope of the adsorption isotherm at small loadings is written as β = qA/qG, where qA is the adsorption–desorption ratio and qG = ϵ/(as) – 1 is a geometrical factor depending on the range a of the tracer-wall interaction, the porosity ϵ, and the specific surface area s. The adsorption leads to a decrease of effective diffusion coefficient, which is quantified by multiplying the geometrical tortuosity factor τgeom by an apparent tortuosity factor τapp. In wide pores or when the adsorption barrier is high, τapp = β + 1, as obtained in previous works, but in narrow pores there is an additional contribution from frequent adsorption–desorption transitions. These results are obtained in media with parallel pores of constant cross sections, where the ratio between the effective pore width ϵ/s and the actual width is ≈0.25. Applications to mesoporous silica samples are justified by the small deviations from this ideal ratio. In the analysis of alkane self-diffusion data, the fractions of adsorbed molecules predicted in a recent theoretical work are used to estimate τgeom of the silica samples, which is ≫1 only in the sample with the narrowest pores (nominal 3 nm). The application of the model to Li⁺ ion diffusion leads to similar values of τgeom and to a difference of energy barriers of desorption and adsorption for those ions of ∼0.06 eV. Comparatively, alkane self-diffusion provides the correct order of magnitude of τgeom, with adsorption playing a less important role, whereas adsorption effects on Li⁺ diffusion are much more important.
Article
Full-text available
Membrane distillation (MD) is a thermal desalination process that is advantageous due to its ability to harness low-grade waste heat to separate highly saline feedstock. However, like any thermal desalination process, the energy efficiency depends on the ability to recover latent heat from condensation in the distillate. In direct contact MD (DCMD), this can be achieved by integrating a heat exchanger (HX) to recover latent heat stored in the distillate stream to preheat the incoming feed stream. Based on the principle of equal heat capacity flows, we derive a simple and intuitive expression for the optimal flow rate ratio between the feed and distillate streams to best recover this latent. Following the principle of energy balance, we derive simple expressions for the specific thermal energy consumption (SECth) and gained output ratio (GOR) of DCMD with and without a coupled HX for latent heat recovery, revealing an intuitive critical condition that indicates whether DCMD should or should not be coupled with HX. As MD is attractive for its ability to use low-grade waste heat as a heat source, we also evaluate the energy efficiency of DCMD powered by a waste heat stream. A waste heat stream differs fundamentally from a conventional constant-temperature heat source in that the temperature of the waste heat stream decreases as heat is extracted from it. We discuss the implication of this fundamental difference on energy efficiency and how we should analyze the energy efficiency of DCMD powered by waste heat streams. A new metric, namely specific yield, is proposed to quantify the performance of DCMD powered by waste heat stream. Our analysis suggests that, for a single-stage DCMD powered by a waste heat stream, whether implementing latent heat recovery or not only affects conventional metrics for energy efficiency (e.g. SECth and GOR) but not the specific yield. Overall, this analysis presents an intuitive and important framework for evaluating and optimizing energy efficiency in DCMD.
Chapter
This anthology is the first collection of primary science articles written by scientists working in America during the nineteenth century.
Article
Membrane distillation (MD) is considered as a promising technology for reverse osmosis (RO) brine treatment, but application of MD is significantly hampered by gypsum scaling and subsequent membrane wetting. As antiscalants are widely used in RO desalination, it is of interest to understand effects of residual antiscalants on membrane scaling during RO brine treatment using MD. In this work, MDC220, PTP-0100 and SHMP were adopted to study effects of residual antiscalants on water vapor flux, distillate conductivity and scale growth using a direct contact membrane distillation (DCMD) device. Feed temperature and antiscalant dose were investigated in ranges of 40–60 °C and 0–10 mg L⁻¹, respectively. Moreover, morphological characterization of scale layers in the presence of antiscalants were performed using a field emission scanning electron microscopy. The results show that all the antiscalants neither decreased feed surface tension nor worsened distillate quality. The presence of antiscalants effectively retarded flux decline and distillate conductivity increase; SHMP and MDC220 outperformed PTP-0100 in restricting gypsum scaling. The performance of antiscalants was temperature dependent, and increased feed temperature adversely affected scaling inhibition performance. The dose of antiscalants played a significant role in gypsum scaling restriction and the inhibition performance was superb at a MDC220 dose not less than 5 mg L⁻¹.
Article
Detecting the onset of membrane fouling is critical for effectively removing membrane foulants during microfiltration (MF) separation. This work investigates the use of electrical impedance spectroscopy (EIS) on the surface of electrically conductive membranes (ECMs) to measure early development of membrane surface fouling. An electrochemical cell was developed in which an ECM acted as a working electrode and a graphite electrode acted as the counter electrode. Conductive membranes were fabricated by coating single-walled/double-walled carbon nanotubes (f-SW/DWCNT) on microfiltration polyethersulfone (PES) supporting membranes. Membrane fouling was simulated by pressure depositing different amounts of latex beads onto the surface of the membrane in a dead-end filtration cell. Changes in membrane water permeability were correlated to the degree of membrane fouling. Clean membranes had water permeability of 392±28 LMH/bar. Reduction of membrane water permeability of 13.8±3.3 %, 15.8±4.7 %, 17.8±0.5 % and 27.1±4.6 % were observed for membranes covered with 0.028mg/m², 0.28 mg/m², 1.40 mg/m² and 2.80 mg/m² on the membranes, respectively. These small differences in fouling degree were statistically resolvable in measured Nyquist plots. It was observed that the diameter of the higher frequency charge transfer region (10⁴ ∼ 10⁶ Hz) of the Nyquist plot semicircles increased with greater fouling. These observations were hypothesized to correspond to decreasing surface conductivities of the membranes by the incorporation of insulating materials (latex beads) within the porous conductive coating. This proposed hypothesis was supported by measured EIS results modeled with a theoretical equivalent circuit. Fouled membrane surface conductivity, surface hydrophilicity, and pore size were measured by SEM, four-point probe conductivity, contact angle, and MWCO experiments, respectively, to compare conventional characterization techniques with non-destructive EIS measurements.
Article
The growth of mineral crystals on surfaces is a challenge across multiple industrial processes. Membrane-based desalination processes, in particular, are plagued by crystal growth (known as scaling), which restricts the flow of water through the membrane, can cause membrane wetting in membrane distillation, and can lead to physical destruction of the membrane material. Scaling occurs when super-saturated conditions develop along the membrane surface due to the passage of water through the membrane, a process known as concentration polarization. To reduce scaling, concentration polarization is minimized by encouraging turbulent conditions and by reducing the amount of water recovered from the saline feed. In addition, anti-scaling chemicals can be used to reduce the availability of cations. Here, we report on an energy efficient electrophoretic mixing method capable of nearly eliminating CaSO4 and silicate scaling on electrically conducting membrane distillation (ECMD) membranes. The ECMD membrane material is composed of a percolating layer of carbon nanotubes deposited on a porous polypropylene support and cross-linked by poly(vinyl alcohol). The application of low alternating potentials (2 Vpp, 1 Hz) had a dramatic impact on scale formation, with the impact highly dependent on the frequency of the applied signal, and in the case of silicate, on the pH of the solution.
Article
Mineral scaling constrains membrane distillation (MD) and limits its application in treating hypersaline wastewater. Addressing this challenge requires enhanced fundamental understanding of the scaling phenomenon. However, MD scaling with different types of scalants may have distinctive mechanisms and consequences which have not been systematically investigated in the literature. In this work, we compared gypsum and silica scaling in MD and demonstrated that gypsum scaling caused earlier water flux decline and induced membrane wetting that was not observed in silica scaling. Microscopic imaging and elemental mapping revealed contrasting scale morphology and distribution for gypsum and silica, respectively. Notably, while gypsum crystals grew both on the membrane surface and deep in the membrane matrix, silica only formed on the membrane surface in the form of a relatively thin film composed of connected sub-micron silica particles. We attribute the intrusion of gypsum into membrane pores to the crystallization pressure as a result of rapid, oriented crystal growth, which leads to pore deformation and the subsequent membrane wetting. In contrast, the silica scale layer was formed via polymerization of silicic acid and gelation of silica particles, which were less intrusive and had a milder effect on membrane pore structure. This hypothesis was supported by the result of tensile strength testing, which showed that the MD membrane was significantly weakened by gypsum scaling. The fact that different scaling mechanisms could yield different consequences on membrane performance provides valuable insights for the future development of cost-effective strategies for scaling control.
Article
Desalination has become an essential toolset to combat the worsening water stress resulting from population and industrial growth and exacerbated by climate change. Various technologies have been developed to desalinate feedwater with a wide spectrum of salinity. While energy consumption is an important consideration in many desalination studies, it is challenging to make (intuitive) sense of energy efficiency due to the different mechanisms of various desalination processes and the very different separations achieved. This perspective aims to provide an intuitive, thermodynamics-based interpretation of energy efficiency by illustrating how energy consumption breaks down into minimum energy of separation and the irreversible energy dissipation. The energy efficiencies of different desalination processes are summarized and rationalized based on their working mechanisms. Notably, an important new concept called the minimum mean voltage is proposed as a convenient tool to evaluate the energy efficiency of electrochemical desalination processes. Lastly, the intrinsic tradeoff between energy efficiency and desalination rate and the relevance of energy efficiency in different desalination applications are discussed.
Article
Membrane distillation (MD) is a promising membrane-based thermal process capable of desalinating highly saline water. However, its application is limited by fouling and wetting of commercial hydrophobic membranes. Inspired by the lotus leaf, we developed a biomimetic superhydrophobic polyvinylidene fluoride (PVDF) membrane for robust MD via self-assembly method. The hierarchical micro-nanoscale texture on the membrane surface was constructed by grafting the spherical polyvinylsilsesquioxane (PVSQ) nanoparticles onto micron-sized silica particles (SiPs). The membrane surface energy can be simultaneously lowered due to the hydrophobic groups including vinyls and methoxyls created from the condensation reaction of the vinyltrimethoxysilane (VTMOS). The resulting membrane showed a very high water contact angle (~160°) and a low water sliding angle (<15°), demonstrating the strong hydrophobicity that is expected for a lotus-leaf-like surface. Compared to the commercial PVDF membrane, the fabricated superhydrophobic membrane exhibited superior resistance against wetting by a surfactant (sodium dodecyl sulfate) and fouling by humic acid during MD experiments. Our results suggest that biomimetic superhydrophobic membranes can enhance the operational robustness of MD and facilitate the efficient use of MD for desalinating saline wastewaters in favor of beneficial water reuse and resource recovery.