You are on page 1of 23

Desalination 495 (2020) 114659

Contents lists available at ScienceDirect

Desalination
journal homepage: www.elsevier.com/locate/desal

Hybrid technologies: The future of energy efficient desalination – A review T


Farah Ejaz Ahmed, Raed Hashaikeh , Nidal Hilal
⁎ ⁎

NYUAD Water Research Center, New York University Abu Dhabi, Abu Dhabi, United Arab Emirates

GRAPHICAL ABSTRACT

ARTICLE INFO ABSTRACT

Keywords: Further reduction in energy consumption is an attractive prospect for both well-established technologies such as
Desalination reverse osmosis and electrodialysis, and for emerging desalination technologies struggling to reach commer­
Hybrid cialization. One way of reducing energy consumption and meeting target water demands is by playing on the
Reverse osmosis strengths of two or more processes through hybridization. Other key objectives of hybridization include flexible
Salinity gradient
operation, increased plant capacity and/or meeting specific water quality requirements. Here, we provide a
Forward osmosis
critical review of hybrid desalination systems, and methods used to optimize such systems with respect to these
Membrane distillation
objectives. Providing a brief overview of current status and energy consumption in both mature and developing
desalination processes, we review the advantages and challenges of hybridization of these processes to overcome
limitations of standalone systems. For instance, coupling of electrodialysis with reverse osmosis helps overcome
the low recovery of reverse osmosis systems. On the other hand, reverse osmosis can be coupled with membrane
distillation for treatment of hypersaline feed solutions or for zero liquid discharge brine treatment. Forward
osmosis relies on hybridization with a low-cost separation process to recover the draw solution. Some promising
candidates for this have been nanofiltration and membrane distillation. Membrane distillation hybrids are sui­
table when thermal energy can be supplied at very low costs. We also review the applicability of salinity gradient
power technologies with desalination systems and identify gaps that need to be addressed for effective upscaling
and implementation of such hybrid systems.

1. Introduction adequate food and livelihood. Despite making up most of the Earth's
surface, clean water remains a precious resource to which billions of
Water is essential to life and to fundamental human rights such as people have little or no access. The UN predicts that up to 5.7 billion


Corresponding authors.
E-mail addresses: raed.hashaikeh@nyu.edu (R. Hashaikeh), nidal.hilal@nyu.edu (N. Hilal).

https://doi.org/10.1016/j.desal.2020.114659
Received 14 July 2020; Received in revised form 30 July 2020; Accepted 30 July 2020
0011-9164/ © 2020 Elsevier B.V. All rights reserved.
F.E. Ahmed, et al. Desalination 495 (2020) 114659

people could be affected by water scarcity by 2050 as more and more water quality and the individual processes that make up the hybrid
people continue to live under water stress [1]. Some of the factors system.
contributing to growing water scarcity are population growth, urbani­ Hybrid desalination systems, first proposed towards the end of the
zation, climate change as well as depletion of limited freshwater re­ 20th century, have traditionally constituted thermal/RO systems,
sources. It is vital that we explore develop technologies and benefit touting the combined benefits of high separation efficiency of multi-
from alternative water sources such as seawater to meet our increasing stage flash (MSF) and low energy consumption of RO [25–30]. This
water needs [2]. However, many widely available water sources are MSF-RO hybrid arrangement has been adopted by existing and new
complex in composition and can contain large amounts of dissolved plants in Saudi Arabia and the UAE to expand water production capa­
salts, that first need to be removed via desalination. Desalination can city, as briefly reviewed by Hamed [24]. However, in 2007, Ismat
treat brackish water and remove salts and minerals from seawater [3]. Kamal [31] concluded that there is no justification for new hybrid
Today, desalination meets nearly 1% of the global water demand and power/thermal/SWRO plants. She argued that such hybridization
the global water desalination market is predicted to grow at 8% be­ makes economic and technological sense only when refurbishing ex­
tween 2018 and 2025 [2]. Water-intensive industries such as oil and isting thermal desalination plants by adding SWRO, which could in fact
gas, agriculture and chemical manufacturing are expected to support lead to a several-fold increase in water output and improved fuel effi­
this growth over the next few years. ciency, whereas new hybrid MSF-RO plants are not techno-economic­
Traditionally, desalination processes either rely on thermal dis­ ally feasible. Today, hybrid desalination technologies encompass a
tillation such multi-stage flash (MSF) and multi-effect distillation broad spectrum that includes integration of RO with other membrane
(MED) or on mechanically driven membrane processes such as reverse processes such as electrodialysis (ED) and membrane distillation (MD),
osmosis (RO). The growing acceptance, and often preference, of as well as the hybridization of RO or MSF with other emerging desa­
membrane technologies is based on ease of operation, compactness, lination technologies. Although, the use of more than one power source
limited use of chemicals, low energy consumption and the development integrated with a single desalination process is also referred to as hy­
of enhanced membrane materials [4]. brid, such systems are not included in the scope of this review. Hybrid
Emerging desalination technologies such as freeze desalination (FD) systems were initially employed for small-scale desalination systems,
and osmotically driven processes like forward osmosis (FO), have also but have recently been extended to large-scale plants, to reduce energy
recently garnered interest although they still face many challenges on cost [23].
the road to commercialization. In the past, researchers have proposed hybrid desalination tech­
Constituting roughly 20–30%, energy is a major factor in the overall nologies as a potential solution towards energy efficient desalination
cost of product water from desalination plants [5]. It does not come as a [9,32]. The development of new membrane processes such as FO and
surprise that there is a vast amount on literature available on energy pressure retarded osmosis (PRO) has caused a revival of research in
consumption of desalination processes. Some reviews target specific hybrid desalination systems. In 2015, Ang et al. [33] reviewed the
processes, while others take a broader approach towards energy con­ performance and applications of integrated/hybrid membrane systems
sumption in the desalination industry. Shrivastava et al. broke down the for desalination and water treatment. They encouraged researchers to
energy efficiency of reverse osmosis, and found an asymptoting energy focus on efficient designs of integrated membrane systems and high-
efficiency, indicating that RO systems currently operate very close to energy efficiency processes; however, they left out osmotically driven
their thermodynamic limit [6]. Similarly, Zarzo et al. [7] consider membrane processes. Soon after, Chekli et al. assessed recent devel­
major energy contributions and methods to optimize energy costs for opments and applications of FO systems integrated with thermal,
RO desalination. Ghalavand et al. [8] compare the energy consumption membrane and/or chemical processes for desalination and water
of leading desalination technologies: mainly seawater reverse osmosis treatment, citing high recovery rates, ability to treat challenging feeds
(SWRO), brackish water reverse osmosis (BWRO), MSF and MED. Re­ and reduced environmental impact as advantages of hybrid FO systems
cently, Nassrullah et al. consider both thermal and membrane desali­ [34]. Hybrid desalination systems are characterized not only be low
nation processes, providing a systematic review of the evolution of specific energy consumption (SEC), but also high plant capacity, flex­
energy consumption and breakdown of desalination processes [9]. They ibility in operation and meeting specific water quality demands. The
emphasize the need for continued attention on renewable energy. The current work presents a critical review of the past, present and future of
number of studies, and in turn reviews, focusing solely on renewable hybrid desalination technologies with respect to these aspects.
energy for desalination [10–22] are manifest to the continuous efforts
in integrating renewable energy sources such as solar, wind and geo­
2. Overview of desalination processes
thermal power with desalination, which also includes implementation
of large-scale solar-powered desalination plants.
The acceptance and growth of desalination have been primarily
Flexibility in operation while meeting specific product water quality
driven by reduction in costs [35], of which energy consumption is a key
demands and brine disposal regulations with high water production
component [36]. Membrane-based desalination, which made up
rates comes at the price of high energy consumption. Specific demands
roughly 10% of the world's seawater desalination capacity in 1999, has
in water quality often require the use of additional passes and/or in­
now surpassed traditional thermal technologies primarily due to lower
creased stress on the energy costs of a stand-alone system. There is
energy costs [37].
growing interest in combining the benefits of two or more systems, to
meet specific water quality goals and/or reduce energy consumption.
Hybridization of desalination technologies is often aimed at meeting 2.1. Current status and energy consumption in desalination systems
one or more objectives, such as eliminating the need for a second pass,
increasing water recovery rate or reducing brine salinity. Hybrid sys­ 2.1.1. Thermal desalination processes
tems have been considered as economically superior alternatives to In the Middle East and North Africa region, thermal desalination
standalone systems due to their ability to reduce energy consumption such as multi-stage flash (MSF) and multi-effect distillation (MED) are
and therefore cost of desalinated water through improved recovery rate the predominant desalination technologies making up to 85% of total
and/or water quality [23,24]. Additionally, hybridization of salinity production capacity [38]. Thu et al. attribute this to the following
gradient power generation technologies with desalination serves as reasons: greater seawater salinity the Gulf region, the recurrent algal
integrated energy recovery and may potentially lower the cost of de­ blooms that would cause severe fouling of RO membranes and the
salination. Effective hybridization requires the optimization of opera­ number of large-scale power generation plants that enable cost-effec­
tional parameters and configuration depending on feed type, target tive colocation of power and desalination plants.

2
F.E. Ahmed, et al. Desalination 495 (2020) 114659

2.1.1.1. Multi-stage flash (MSF) system also depends on whether the energy is supplied with waste heat
2.1.1.1.1. Overview of MSF. As its name suggests, MSF relies on from a power generation plant as is the case in many GCC countries. In
flash evaporation of part of the feed as it moves through multiple such cogeneration plants, high availability steam produced by a steam
stages. Flashing occurs when the pressure in each stage is lower than turbine is partly used to supply thermal energy to desalination systems
the vapor pressure of the heated liquid. In MSF, each stage is at a higher including MSF and MED [44,47].
temperature than the previous stage and the pressure in each stage
corresponds to the boiling point of water at that temperature. Cold feed 2.1.1.2. Multi-effect distillation (MED)
is heated as it passes through each stage and is further heated in the 2.1.1.2.1. Overview of MED. Multi-effect distillation (MED) is the
brine heater. As the brine flows back through the stages, it has a oldest industrial desalination process in use today [48,49]. The MED
temperature higher than the boiling point at the given pressure and evaporator consists of cells, called effects, at decreasing pressure and
thus a fraction of the brine boils to steam. The vapor is then condensed temperature from first to last, with temperature typically between 65
on the external surface of heat exchanger tubes [39]. The two most and 90 °C [50]. The lower pressure in each effect enables water to
common MSF configurations are once-through MSF (MSF-OT) and brine evaporate at low temperatures as the boiling point of water decreases
mixing MSF (BM-MSF). In OT-MSF, the brine blowdown, or the with pressure [51]. Each effect consists of evaporator tube bundles on
unevaporated brine leaving the last stage, is returned to sea. In BM- which seawater is sprayed. Heating steam or hot water through the
MSF, part of the brine is mixed with incoming feed. Evaporator surfaces tubes is supplied in the first effect and it transfers energy to the
are a key component of both MSF and MED [40]. Evaporator surfaces seawater in each effect, causing partial evaporation [50]. Clean
can consist of submerged tubes or falling films and plates, among distillate water is produced when steam is condensed in the last effect
others. using a heat exchanger. This product water is pumped into a storage
Today, MSF accounts for roughly 23% of total installed desalination tank while the brine is pumped back into the sea.
capacity [41], as is seen in Fig. 1. In the Middle East, desalination is 2.1.1.2.2. Energy consumption in MED. MED plants producing water
dominated by MSF despite its higher capital investment and operating in the range of 5000 to 50,000 m3/day typically require between 145
costs as compared to other existing technologies. This is attributed to and 230 MJ/m3 of thermal energy which translates into 12.2 to 19.1
the accumulated experience, simple construction, use of low-cost con­ kWh/m3 of electrical energy. In addition, pumps account for 2 to 2.5
struction materials and proven reliability. As MSF desalination makes kWh/m3 of additional electrical energy [9]. The energy consumption in
up a significant portion of the global installed desalination capacity and MED is linked to the configuration of feed and vapor flow. The three
remains the most widely used desalination technology in the Middle main configurations are forward feed, backward feed and parallel feed.
East due to availability of low-grade heat and cogeneration of power As shown in Fig. 1, MED accounts for roughly 8% of the world's
and water, it is important to implement strategies to further minimize installed desalination capacity.
the cost of running these systems. In the following section, we will The top brine temperature (TBT) in MED plants is limited by cal­
consider the evolution of energy consumption in MSF systems. cium sulfate scaling whereas the temperature at the bottom of the
2.1.1.1.2. Energy consumption in MSF. Although many factors such condenser is limited by the normal temperature of seawater that is used
as the maximum temperature of the heat source, salt concentration in as cooling water. Although many plants operate at a TBT of 70 °C to
the flashing brine, configuration and number of stages as well as the minimize potential of scaling on tube surfaces, a lower TBT translates
design of the heat-exchange devices affect the energy consumption in into greater required heat transfer area is required for evaporation [37].
an MSF system, Al-Karaghouli and Kazmerski emphasize that reduction Another factor affecting the performance of MED is the number of ef­
in energy consumption in MSF can be driven by increasing the gain fects. Studies show that the performance ratio in MED increases with
output ratio (GOR), number of stages and heat transfer area [42]. GOR the number of effects [37]. This is because the number of effects in­
is used to measure the efficiency of a thermal desalination system and is creases specific heat transfer area, reduces the temperature drop across
defined as the mass of kg of water produced per kg of steam used each effect and allows for enhanced internal energy recovery. This re­
[8,43]. Practical values of 8 to 12 kgdistillate / kgsteam are typically duces SEC and thus enhances the performance ratio of the system [9].
reported for commercial MSF systems [43]. Scale formation, fouling However, the number of effects is limited by the difference between the
and corrosion of pipes also reduce the energy efficiency of MSF systems. condensing temperature at the first hot plant effect and that in the final
MSF plants consume heat energy equivalent to ~20 kWh/m3 in condenser [49] as well as the minimum temperature difference between
addition to mechanical equivalent energy of 4 kWh/m3 [44,45]. El- consecutive effects [37]. MED requires less thermal energy than MSF as
Naser reported that the energy consumption by MSF plants falls 12–24 it typically operates at lower temperatures [9,52] and is thermo­
kWh of power for every cubic meter of water produced [46]. It must be dynamically the most efficient of thermal distillation processes [53].
noted that the cost of the thermal energy for a thermal desalination
2.1.2. Reverse osmosis
2.1.2.1. Overview. Today, RO is a leading membrane process and is
considered an established desalination technology. Prior to widespread
commercialization of RO, desalination involved high-energy thermal
processes that relied on phase change. RO relies on reversing the flow of
osmosis by applying a hydraulic pressure greater than the osmotic
pressure of a solution, thus forcing water to move from a high-solute
region to a low-solute region through a semi-permeable non-porous
membrane. As RO is not a phase change process, the theoretical energy
consumption to produce fresh water from RO is significantly lower than
that needed to evaporate water from a feed solution. This has made RO
an attractive technology over the years with large-scale plants installed
worldwide [54]. In many parts of the world, RO desalination has
surpassed thermal desalination in installed capacity. The low recovery
of RO plants especially for seawater desalination combined with the
vast knowledge available on optimization of RO systems makes RO.
Fig. 1. Breakdown of worldwide desalination capacity by technology [41].

3
F.E. Ahmed, et al. Desalination 495 (2020) 114659

2.1.2.2. Energy consumption in reverse osmosis systems. According to the


laws of thermodynamics, the minimum theoretical energy to produce
fresh water from seawater at 50% recovery is 1.14 kWh/m3 [55].
However, pump inefficiencies, other losses as well as the energy
consumed in downstream and upstream processes result in seawater
RO plants typically operating at 10 kWh/m3 without energy recovery
and at 2–5 kWh/m3 with energy recovery. The energy consumption of
smaller plants can be > 15 kWh/m3. Brackish water RO requires
around 1–3 kWh/m3 [56].
Kim et al. [57] have critically analyzed trends in energy consump­
tion for SWRO systems by using data from > 70 large-scale SWRO
plants (with capacity ≥10,000 m3/day). The increasing need for de­
salination, reflected by growing plant capacity (Fig. 1), complements a
gradual decrease in energy consumption.
Data collected by Nikolas Voutchkov from 20 desalination plants of
production capacity ≥40,000 m3/day between 2005 and 2010 shows
that the average energy consumption of SWRO membrane systems is
3.1 kWh/m3 [58]. A comparable figure of 3.7 kWh/m3 has been found
by Raphael Semiat [59]. Voutchkov emphasizes that this energy use
represents the desalination component, which contributes roughly 65%
- 80% of the total energy demand for a desalination plant, while pre­ Fig. 3. Specific energy consumption (SEC) in kWh/m3 for desalination of saline
treatment of seawater and, occasionally, product water delivery are the water from the Arabian Sea, typical seawater, Tampa Bay and brackish
next largest users of energy. groundwater as a function of membrane permeability [62]. Published by the
The overall energy of RO systems depends on several factors such as Royal Society of Chemistry.
feed characteristics, recovery rate, pump efficiency, membrane per­
meability, RO configuration (determined by target permeate quantity Combining these two yields a total energy demand curve with a
and quality), and the type of energy recover device (ERD) used, if any. shallow minimum at around 50% recovery, indicating that even
Many of these factors and the evolution of pumps, ERDs, membranes major improvements in RO membranes may not lead to significant
and configurations over decades have been discussed in detail else­ reduction in RO energy consumption today [59].
where [9,57,58,60]; thus only a brief overview is provided here in order Recently, Patel et al. reached a similar conclusion for desalination
to guide our discussion on energy and hybrid desalination technologies. technologies through their systematic analysis of materials on energy
2.1.2.2.1. Feed characteristics. Feed temperature is said to not be consumption: ultrapermeable materials will do little to further improve
strongly related to SEC for RO systems [57]. As Koutsou et al. point out energy efficiency of desalination systems (Fig. 3) [62]. On the contrary,
[61], an increase in feed-water temperature generally increases issues related to concentration polarization and membrane fouling may
membrane permeability, lowers the pressure required and thus leads amplify at the high fluxes achieved by such membranes and efforts
to a reduction of SEC, especially for low-salinity feeds. For desalination should instead emphasize aspects of membrane-solute selectivity for
of high salinity feeds such as seawater, an increase in temperature specific applications such as removal of boron or chlorides [62].
increases the osmotic pressure and counters the positive effects and an 2.1.2.2.3. High-pressure pumps and ERDs. Pressurizing the feed
optimum temperature is reached at ~30 °C. With respect to salinity, through high-pressure pumps is the largest contributor to energy
higher feed total dissolved solids (TDS) or greater osmotic pressure consumption and operating cost in an RO system. Karabelas et al.
requires a higher applied pressure. Hence, the greater the feed salinity, agree that, in addition to membrane permeability, pump and ERD
the greater the energy consumption. inefficiencies have the greatest effect on SEC [63]. In RO systems, a
2.1.2.2.2. Recovery ratio and energy consumption. Semiat calculated large part of the energy leaves in the form of pressurized brine. This
the energy requirement of RO as a function of recovery ratio, as shown holds especially true for seawater RO systems which typically operate at
in Fig. 2. The energy demand for pumping and pretreatment decreases low recovery rates of 35–45% [64] and also require greater applied
with increasing recovery ratio, while the energy demand for the RO pressure. ERDs were first applied to RO systems in the 80s with the aim
process increases as recovery ratio due to pressure changes [59]. of recovering this energy and using it to pressurize incoming feed, thus
reducing the energy consumption of the system [65,66].
The first generation of ERDs used for RO was of the centrifugal type.
Widely used devices of this type include Francis turbines, Pelton wheels
and turbochargers. These were limited in capacity and had efficiencies
of < 80%. The advent of isobaric ERDs in the form of piston-type work
exchangers and rotary pressure exchangers (PX) translated into un­
limited capacity and efficiencies reaching as high as 98% [67–69],
which led to RO energy savings of up to 40% [70].
The Pelton turbine (PT) remains a widely used commercial ERD
technology despite it being a centrifugal device. Among isobaric de­
vices, PX are widely being implemented in RO plants, although other
technologies such as axial piston motors (APM) and pressure in­
tensifiers (PI) are also commonly available ERD technologies. PX ERDs
operate by bringing the pressurized brine into contact with low-pres­
sure feed to transfer energy at high efficiency. Although the choice of
ERD also depends on the minimum brine flow rate [22,71], PX devices
offer low payback time, are scalable and require little maintenance.
[67,70,72–74]. Improvement in the energy efficiency of ERDs as they
Fig. 2. Energy demand in RO as a function of water recovery [59].

4
F.E. Ahmed, et al. Desalination 495 (2020) 114659

Fig. 4. (a) Production capacity of SWRO desalination plants over commissioning years; (b) Specific energy consumption (kWh/m3) for SWRO desalination plants
using ERDs of different types [57]. DWEER: dual work exchanger energy recovery; FT: Francis turbine; PT: Pelton turbine; PX: pressure exchanger.

transitioned from turbine to isobaric has caused significant reduction in With a shift from fossil fuels to renewable energy (RE), the coupling
SEC. Fig. 4 shows the reduction in SEC of SWRO plants and the different of commercially viable RE technologies such as photovoltaics (PV) with
types of ERDs employed over the years. desalination technologies is set to achieve new energy trends. Al Khafji,
ERDs are not only beneficial for SWRO systems, but have also shown the world's largest PV-powered reverse osmosis (PV-RO) desalination
to reduce the energy consumption for brackish water RO (BWRO) [75]. plant is set to become the world's largest PV-powered reverse osmosis
For a medium-scale BWRO desalination plant, an ERD with efficiency of (PV-RO) desalination plant with a production capacity of 60,000 m3/
90% could reduce energy consumption by almost 54%. day [22]. Semiat argues that solar powered desalination is only suitable
Table 1 summarizes the relative weight of each factor on the SEC of when batteries are used to ensure continuous operation, which adds to
small-scale RO systems [76]. It can be seen that SEC depends strongly energy costs [22,59].
on feed and product water quality, applied pressure and the use of
energy recovery devices [76]. 2.1.3. Nanofiltration
Fath et al. analyze trends in production rates and energy con­ 2.1.3.1. Overview. With membrane properties that lie between
sumption for different desalination technologies among nations of the ultrafiltration (UF) and RO, nanofiltration has emerged as a unique
Gulf Cooperation Council between 2005 and 2025 [77]. In Saudi and versatile pressure-driven separation technology which many
Arabia, a country that leads desalination and is also home to the world's applications. With lower required pressure and therefore lower
largest desalination plant in Al-Jubail, a 2.5-fold increase in installed energy consumption when compared to RO [78], NF is capable of
desalination capacity is accompanied by a similar increase in energy separating multivalent inorganic salts and small organic molecules from
consumption from 48,000 GWh to 119,000 GWh. A similar trend was feed streams at high rejection rates. NF membranes exhibit rejection
observed for the UAE. Given rapid global advances in both membrane rates between 75 and 99% for divalent salts and 30–50% for
productivity and selectivity, this increase in energy consumption could monovalent salts [79]. As the rejection of some small-size organic
be surprising. One of the factors has been the slower transition from contaminants is lower in NF, it is often used in combination with other
high-energy consuming thermal technologies to membrane technolo­ separation processes such as adsorption to achieve similar removal of
gies in the region as compared to the rest of the world. The study micropollutants [80].
suggested energy consumption strategies in the form of novel pre­
treatment methods and configurations for thermal desalination, and
low-pressure membranes for the growth of RO [77]. 2.1.3.2. Nanofiltration as pretreatment to desalination. Adequate and
reliable pretreatment can minimize fouling and scaling issues in the
desalination system and lower operating costs [81–83]. In addition to
Table 1 removal of large impurities as is the case with conventional
Relative importance of various factors in determining the SEC of small-scale RO pretreatment and low-pressure membrane processes such as
systems, as analyzed by Stillwell and Webber [76]. microfiltration (MF) to ultrafiltration (UF), softening of the feed and
efficient removal of divalent ions at the pretreatment stage would not
Factor Variable Relative weight
only reduce the osmotic pressure of the feed and lower the energy
3
Raw water flow (m /day) qrw 0.067a required for RO, but would control membrane scaling and increase the
Product water flow (m3/day) qpw 0.072a lifespan of RO membranes. Pretreatment to desalination processes, both
Inverse recovery 1 0.071a thermal and RO, have thus been a major application of NF [84]. By
1 R
Raw water TDS (mg/L) crw 0.15a removing scaling species such as calcium sulfate and calcium
Product water TDS (mg/L) cpw 0.19a carbonate, NF pretreatment enables the use of higher top brine
Pressure (bar) P 0.12a
temperatures (TBT) in thermal desalination, and increases GOR.
Energy recovery equipment ER 0.15a
Temperature (∘C) T 0.034 Apart from desalination, NF membranes are also widely used in food
and dairy, textile, paper pharmaceutical, and petrochemical industries
a
Relative weight is statistically significant. [85].

5
F.E. Ahmed, et al. Desalination 495 (2020) 114659

Although ionic transport in NF membranes is not fully understood, negative electrodes respectively. Combined with the selective separa­
as is shown by recent efforts in modelling NF mass transport [86]. tion of cations and anions through the stack of CEMs and AEMs, this
However, there have been sufficient studies to show that charge, size ionic transport results in channels with low-concentration (diluate) and
exclusion and dielectric effects are all partially responsible for ionic high-concentration (concentrate) solutions [100]. Separation via ED
transport through NF membranes [87,88]. Charge polarity between the takes place when cations moving through the CEM are prevented from
membrane and the solute causes separation by attraction/repulsion the following AEM and the similar takes place for anions [99]. This
forces. Steric effects are related to transport due to the relative size of results in removal of salt and subsequent accumulation in concentrate
the ions and membrane pores, which for NF membranes are on the channels. A single cell pair refers to the assembly of an AEM and a CEM
order of 1 nm [89,90]. Transport due to dielectric effects results from and adjacent channels [100]. Ion exchange membranes used for ED are
differences in dielectric constant between bulk and membrane pores characterized by high selectivity and low electrical resistance. Like
[87]. other membranes, these membranes are prone to fouling in the form
accumulation of impurities on the surface. Reversing the polarity of the
2.1.3.3. Energy consumption in nanofiltration. Many of the factors that ED system mitigates fouling by removing foulants from the surface of
determine energy consumption for RO are shared by NF since it is also a the membranes [101–104]. This modified process, known as electro­
pressure-drive process. As discussed earlier, RO systems have energy dialysis reversal (EDR) was first developed by Ionics in 1974 and
demands of 1.5–2.5 kWh/m3, with the highest contributor to cost being eliminated the need to use acid or anti-scalant chemicals into ED sys­
high-pressure pumps. Since NF membranes operate at lower pressures, tems [105]. Fig. 6 shows key milestones in the development of ED
the energy demand is also lower at 0.6–1.2 kWh/m3 [80]. Some technology.
challenges associated with nanofiltration include membrane fouling,
low separation of micropollutants and the need to develop a deeper 2.1.4.2. Energy consumption in electrodialysis. An industrial scale ED
understanding of mass transport [91,92]. Advances in membrane system can consist of 300–500 cell pairs [107,108]. The potential
materials [93–95], especially anti-fouling membranes [96,97], have difference applied and hence the energy consumption depends strongly
partially driven nanofiltration towards lower energy costs. NF as on the salt concentration of the feed solution, as well as the membrane
pretreatment to RO or MSF has not been considered a hybrid used. The rate of salt removal is proportional to the electric current
desalination technology in this review. [109,110]. Values between 0.75 and 1.5 V/cell are typical, requiring a
high potential difference for a large-scale ED system. Higher
2.1.4. Electrodialysis concentration of ions would require a higher potential difference for
2.1.4.1. Overview and current status. Electrodialysis (ED) is an electro- efficient separation. Given the high concentration of ions in seawater,
membrane process in which ionic and non-ionic components are the electrical energy required for seawater desalination by ED would be
removed under the effect of an electric field. Although it was first unfeasible, rendering it suitable for low-concentration solutions such as
patented in 1890, ED became a commercially established process for brackish water with < 5000 mg/L of total dissolved solids (TDS) [111].
desalinating brackish water > 50 years ago [98]. Anions migrate Energy is also required for pumping and for the reactions that take
towards the positive electrode while cations migrate towards the place at the electrodes, the latter of which accounts for 1–3% of the
negative electrode. An ED system consists of alternately arranged total energy consumption [112]. The efficiency of the ED process
cation exchange membranes (CEM) and anion exchange membranes depends on the properties of the ion exchange membranes used, feed
(AEM), as shown in Fig. 5. water composition and amount of salt removal required. The electrical
When an electric field is applied across the electrodes, an ionic conductivity of the membrane determines the potential drop across the
current forces anions and cations to migrate towards the positive and stack of membranes and thus the energy efficiency of the overall system

Fig. 5. Schematic of electrodialysis desalination [99].

6
F.E. Ahmed, et al. Desalination 495 (2020) 114659

Fig. 6. Milestones in the development of electrodialysis and related technologies. Adopted from [106,107].

[106]. Theoretically, the energy required for ED to produce water of 2.1.5. Emerging desalination technologies
800 mg/L is 3.3 kWh/m3 for brackish water and 26 kWh/m3 for Many argue that the cost of desalination today, particularly sea­
seawater desalination [113]. On average, 0.5–1.1 kWh/m3 for water desalination, remains high despite increasing industrial installa­
pumping, 0.7 kWh for each 1000 mg/L TDS removed and roughly 5% tion capacity. Key issues associated with RO for example include high
accounts for energy losses in a brackish water ED desalination system energy consumption, high capital cost and membrane fouling. In ad­
[112]. According to a review by Sajtar and Bagley, energy consumption dition, other growing challenges in the form of brine treatment and
for removal between 0 and 2000 mg/L TDS from a feed stream ranges disposal require new approaches. This has led scientists and engineers
from 0.1 to 1 kWh/m3 [110]. Although ED is typically applied as a to delve into novel desalination processes. Some promising emerging
room temperature process, introducing a temperature gradient or desalination technologies that have received significant interest in re­
increasing the temperature of the system can cause energy reductions cent years but are not yet widely applied at a commercial scale are
[114]. Increasing the temperature increases ion mobility, reduces discussed in this section.
electrical resistance of the solution and decreases solution viscosity.
Benneker et al. found that the energy required for ED can be reduced by 2.1.5.1. Membrane distillation
9% if the temperature of one of the feed streams is increased by 20 °C 2.1.5.1.1. Overview and current status. Membrane distillation is a
[115]. However, loss of rigidity or degradation of polymer membranes separation process in which a porous hydrophobic membrane is in
could occur at high temperatures limiting the upper temperature for ED contact with aqueous heated feed solution on one side. A vapor pressure
design and operation. A higher current density lowers the required difference across the membrane arises from the temperature difference
membrane area which leads to a reduction of capital cost. However, the between the hot solution and the colder permeate flowing on the other
voltage applied also increases with current density, thus increasing side. This vapor pressure difference causes more volatile compounds to
energy costs. Depending on the water salinity, 0.5–10 kWh/m3 of evaporate and transport through the membrane pores. MD is a versatile
electrical energy can be consumed by an ED system [56]. For example, separation technique with applications ranging from the separation of
to lower TDS from 1500 ppm to 500 ppm, an ED unit would consume dairy compounds, juices, pharmaceutical compounds to desalination
~1.5 kWh/m3. Recently, multi-stage electrodialysis systems have been and treatment of oily wastewater [121]. MD has found niche
investigated for reduced energy consumption. Chehayeb et al. found applications within desalination in recent years where it may
that energy consumption for brackish water desalination can be compete with conventional processes such as RO. These include brine
reduced by 29% by using a two-stage system provided that fixed concentration [122–125] or simple, small-scale desalination systems in
costs are lowered [116]. In ED, the rate of salt removal is proportional remote areas powered using waste heat or renewable energy
to the electric current applied. Limiting current density (LCD) is the [126–128].
phenomenon of the electric current being limited by the rate of 2.1.5.1.2. Energy consumption in membrane distillation. The thermal
diffusion of ions to the membrane surface and occurs when the ion energy used for heating the feed is the largest contributor of energy in
concentration at the membrane surface reaches zero. Thus, LCD i.e. MD. Thermal energy efficiency corresponds to heat energy such as
current per unit area of membrane (mA/cm2) represents the maximum latent heat and conductive heat. MD energy costs can be low if waste
transport rate of ions through the membrane [117]. Doornbusch et al. heat is used to supply thermal energy to the system. When the energy
used the limiting current density of each stage to determine the being supplied comes at a cost however, the goal should be to maximize
maximum desalination rate per each stage and evaluate the potential thermal efficiency. As discussed above, a niche application for MD is
for multistage ED with 4 stages for seawater desalination. They were brine concentration, for which RO would not be energetically feasible
able to lower the salinity of seawater from 500 mM to 11.4 mM TDS at due to the high osmotic pressure of brine streams. MD has not still not
an energy consumption of 3.6 kWh/m3, which is not far off from state- achieved widespread commercial implementation in desalination due
of-the-art RO systems [118]. ED mass transfer continues to grow in the to high energy consumption and in turn the high cost of product water
overlimiting current region. Membrane modification plays a strong role [129].
in ED desalination performance at overlimiting currents, enabling a There are four basic MD configurations namely, direct contact
shift of optimum voltage range corresponding to minimum costs [119]. membrane distillation (DCMD), air-gap membrane distillation (AGMD),
The application of ED remains limited by the high cost of electrodes and vacuum membrane distillation (VMD) and sweeping gas membrane
ion exchange membranes, and the electrically-driven degradation of distillation (SGMD). All of them involve direct contact of the membrane
polymeric membranes [120]. with a heated feed solution. They differ in how the vapor pressure
gradient is induced and how transported vapor is collected on the

7
F.E. Ahmed, et al. Desalination 495 (2020) 114659

Table 2
Pros and cons of the four major MD configurations [121].
Configuration Pros Cons

DCMD The easiest and simplest configuration to realize practically, flux is more stable Flux obtained is relatively lower than vacuum configurations under the identical
than VMD for the feeds with fouling tendency, high gained output ratio, it operating conditions, thermal polarization is highest among all the
might be the most appropriate configuration for removal of volatile configurations, flux is relatively more sensitive to feed concentration, the
components. permeate quality is sensitive to membrane wetting, suitable mainly for aqueous
solutions.
VMD High flux, can be used for recovery of aroma compounds and related Higher probability of pore wetting, higher fouling, minimum selectivity of
substances, the permeate quality is stable despite of some wetting, no volatile components, require vacuum pump and external condenser.
possibility of wetting from distillate side, thermal polarization is very low.
AGMD Relatively high flux, low thermal losses, no wetting on permeate side, less Air gap provides an additional resistance to vapors, difficult module designing,
fouling tendency. difficult to model due to the involvement of too many variables, lowest gained
output ratio.
SGMD Thermal polarization is lower, no wetting from permeate side, permeate
quality independent of membrane wetting

permeate side [121]. The energy consumed in MD is largely dependent molecular weight, high solubility in water and a high degree of dis­
on the configuration used. However, non-standardized testing condi­ sociation [164,165]. Low molecular weight and viscosity in water also
tions make it difficult to compare configurations across studies. indicate a greater diffusion coefficient, reducing the effects of con­
Due to simplicity, most lab-scale MD studies adopt the DCMD con­ centration polarization during FO [166].
figuration [130]. However, it has been reported that both VMD and As such, multivalent ionic solutes are preferred candidates. Second,
AGMD have greater thermal energy efficiency [131,132], which makes the draw solution must be readily regenerated or reconcentrated using
them popular choices for companies seeking to commercialize MD simple low-cost methods so as to keep overall operation costs and en­
processes [121]. Table 2, adopted from Drioli et al. [121], shows the ergy consumptions low [163]. The dependence of osmotic pressure on
benefits and drawbacks of each of the four conventional MD config­ solution concentration for various draw solutions was calculated and
urations. presented by Cath et al. [167], as shown in Fig. 7.
In the last decade alone, much progress has been made in novel MD Although forward osmosis was first explored in the late 60s, the first
membrane materials [133–143]. Despite developments in membrane commercial application did not emerge until 2002. The world's first FO
fabrication and process design, SEC of MD systems is still vague with a desalination plant was built in 2010, followed by large-scale pilot de­
wide range of reported values. This dispersion has been studied by monstrations in other locations. Since then, there has been an increase
Khayet et al. [144], who also suggested that standard methods should in FO system suppliers in the market. Fig. 8 shows key milestones in the
be implemented to calculate water production costs for MD. Operating development of FO as a desalination technology.
conditions such as recovery rate have not been optimized in many MD 2.1.5.2.2. Energy consumption in FO. FO is widely promoted as a
studies. When comparing relative flow arrangements between feed and low-energy desalination technique. Although drawing water from the
permeate, Swaminathan et al. [145] found that countercurrent flow feed solution to the concentrated DS is a low-energy process driven by
yielded highest energy efficiency, followed by crossflow and parallel. osmotic flow, it is the DS recovery step that dictates the energy
Table 3 illustrates reported SEC values reported for selected MD sys­ consumption of the entire process. During the osmosis step, a low-
tems [146]. pressure pump is needed to overcome the pressure drop in the feed
channel, and the energy consumed is ~0.10–0.11 kWh/m3 at 50%
water recovery [169,170]. Values of 0.2–0.55 kWh/m3 have also been
2.1.5.2. Forward osmosis
reported for the osmosis step [168]. Moon and Lee suggest that FO
2.1.5.2.1. Overview. Osmosis is the spontaneous and selective
desalination ranges from 3 to 8 kWh/m3, depending on the energy
passage of solvent from a dilute solution to a concentrated solution
required for draw solute regeneration [171]. Future work in reducing
through a semipermeable membrane separating the two [159,160]. The
the energy consumption in FO systems is geared towards process
pressure needed to stop this spontaneous movement across the
configuration as well as large-scale process implementation [168].
membrane and maintain equilibrium is known as osmotic pressure.
For a more detailed review on the technological status of and factors
When hydraulic pressure is applied to counter and exceed the osmotic
affecting energy consumption in FO desalination systems, the reader is
pressure, osmosis is reversed i.e. ‘reverse osmosis’ and water is
referred [168] and [9].
transported from high solute to low solute region. On the other hand,
transport in osmotically driven processes such as forward osmosis (FO)
is guided by an osmotic pressure gradient across the semipermeable 2.1.5.3. Freeze desalination
membrane. As a result, separation in FO requires little or no hydraulic 2.1.5.3.1. Overview and current status. Although the concept of
pressure as a concentrated draw solution (DS) with a greater osmotic water desalination by freezing was first proposed in the late 1600s
pressure draws in water molecules from the feed solution through a [172], it was not considered to be of practical use before the invention
membrane [161]. The then diluted DS is then further processed to of refrigeration systems. When the Office of Saline water was created in
separate draw solutes from purified water. FO does not require high 1955, freeze desalination (FD), along with other desalination
pressure and, for the same reason, is not as prone to fouling and scaling techniques received renewed interest [173,174]. Since then, as
as RO membranes. FO is also a strong candidate for hybrid systems membrane technologies moved towards commercialization, progress
where other separation techniques are applied to recover the draw in FD as a viable desalination technology remained slow, and only
solute after the direct osmosis step. picked up pace in the last 10 years. According to the Web of Science
Apart from feed characteristics and membrane properties, the databases, out of 82 publications with ‘freeze desalination’ in their title,
choice of appropriate DS strongly affects the performance of an FO 52 were published between 2010 and 2019. In freeze desalination, ice
system [162]. An appropriate draw solution must fulfill several re­ crystals are formed from the chilled feed and later melting, thus
quirements. First, it must be able to generate a high osmotic pressure, to allowing fresh water to be extracted [175]. Freeze desalination (FD)
ensure a high osmotic pressure difference between feed and draw so­ relies on salts and impurities being rejected by ice crystals during the
lutions [163]. This can be achieved by using ionic compounds with low freezing process which results in a high separation efficiency [176]. The

8
F.E. Ahmed, et al. Desalination 495 (2020) 114659

[147]
[148]
[149]
[150]
[151]
[152]
[153]
[154]
[155]
[156]

[157]

[158]
Refs.
Plant capacity (m3/h)

0.0034–0.0094
931 (overall)

2.67–6.94
3.46–19
0.2–20
0.05

3.85

~697 to 10,457b
8100.8–9089.5
SEC (kWh/m3)

~65 to ~ 127

~130 to 1700
3550–4580
600–1600
140–200

200–300
1.6–27.5

1500
30.8

130
Tap water, synthetic seawater

Simulated reverse osmosis


Radioactive solution

Underground water
Brackish water

Distilled water

Distilled water

Wastewater
Feed type

Seawater

Seawater
Seawater

Seawater
brine

Fig. 7. Osmotic pressure as a function of solution concentration at 25C for


various potential draw solutions as calculated by Cath et al. [167].
13.4–14.4
Operating conditions

Tp (°C)

18–21

observation of brine pockets trapped in frozen sea ice indicates that


5–30

30

20

frozen sea ice is not pure fresh water and the ice should be separated

from the brine before melting to obtain high quality product water
313–343

39.8–59
15–22a

[172]. Thus, the three main steps involved in FD are ice crystallization,
Tf (°C)

35–80
60–85

50–70

separation of ice from brine and melting. Fig. 9 shows a flow diagram of
80
80

60

85

components in any freeze desalination process [172]. The crystallizer is


the vessel in which ice crystals are formed. These are then separated
PTFE with PP support, mean pore size 0.5 ± 0.08, porosity 91 ± 0.5, active layer thickness

from brine in the separator/washer unit which often consists of a


centrifuge, filter or wash column [177,178]. Product water is obtained
by melting the ice crystals in the last stage of the process.
In addition to exhibiting high separation efficiency, FD is insensitive
Commercial membranes from Membrana with Pore size 0.2 μ and thickness 91 μ

to fouling and does not require any chemicals. Commercialization of FD


has been limited by difficulties related to separation of the ice crystals
Flat sheet PP, thickness 400 μm, Pore size 0.1 μ, porosity 70%, Am 5 m2

from the concentrated brine solution, as well as lack of optimization of


process parameters [179].
LDPE, thickness 76 μ, Pore size 0.3 mμ, porosity 85%, Am 7.4 m2

2.1.5.3.2. Energy consumption in FD. Compared to other phase


PP models from Microdyn Nadir, Pore size 0.2 μ, porosity 73%
Spiral wound PTFE (SEP GmbH), Pore size 0.2 μ, porosity 80%

Several commercial membranes with different characteristics

change processes that involve distillation, FD has a significantly lower


energy requirement as the latent heat of vaporization of water is
2256.7 kJ/kg, which is roughly 7 times the latent heat of fusion of ice
(333.5 kJ/kg) [178,180]. Cold energy can also be provided from
regasification of liquefied natural gas (LNG) [181–188]. Most cold
Specific energy consumption (SEC) of selected MD systems [146].

energy during LNG regasification is wasted, but if used in FD systems,


could keep desalination costs low. For 80% water recovery, FD needs
PVDF hollow fiber, thickness 240 μm
PTFE, Pore size 0.2 μ, porosity 80%

between 12.31 and 13.78 kW/m3 of energy [189]. Although this is


PP, thickness 35 μ, Pore size 0.1 μ

higher than energy consumed by RO systems, it is significantly lower


than other phase change processes. If this energy is provided mainly by
46 ± 1 μm, Am 0.67 m2
Membrane characteristics

waste cold energy from LNG regasification, FD has the potential to be a


PTFE, Pore size 0.2 μ

low-cost desalination alternative.

2.1.6. Salinity gradient energy harvesting


Techniques for harvesting salinity energy are closely related to de­
salination techniques. These techniques convert chemical energy in
salinity gradients to useful energy. Often, these are hybridized with
DCMD in hybrid systems

desalination technologies to simultaneously generate clean water and


energy. Two of these techniques, namely reverse electrodialysis (RED)
and pressure retarded osmosis (PRO) are introduced here. Although
Configuration

there are other techniques in literature for conversion of salinity gra­


dient to useful energy, PRO and RED are the most commonly studied
Table 3

AGMD
AGMD
AGMD

AGMD
DCMD

DCMD

DCMD

DCMD

DCMD
VMD

VMD

[190]. They differ in operating principles, membrane properties and


performance.

9
F.E. Ahmed, et al. Desalination 495 (2020) 114659

Fig. 8. Milestones in the development of FO desalination [168].

Fig. 9. Flow diagram of freeze desalination [172].

Fig. 10. Schematic of power generation via reverse electrodialysis (RED) stack with two cells [191].

2.1.6.1. Reverse electrodialysis. RED systems, like ED systems, consist of were achieved when IEMs were tailored for RED along with improve­
alternating cation and anion ion exchange membranes (IEMs) forming a ments in stack design [192]. As is the case in ED, membrane properties
stack. Concentrated and diluted salt solutions flow through alternating play a significant role in determining the performance of the process.
compartments to create a salinity gradient [190]. Ion transport through Recently, high power density membranes have been developed at the
the membranes due to chemical potential difference between solutions lab-scale. These include nanoporous membranes that demonstrate
is guided by IEMs and a voltage is generated across the membranes power densities 1–3 orders of magnitude greater than conventional
[191]. This voltage is known as the membrane potential. The electric IEMs [193]. However, IEMs remain 2–3 times more costly than RO
potential difference generated across the stack is an accumulation of the membranes [194]. Other factors that play a role in enhancing RED
potential differences across each membrane [190,191]. Fig. 10 shows a power density include spacer and electrode system design.
schematic of an RED system. The two solutions (concentrated and Although there are no full-scale RED plants to date, pilot facilities
diluted) are continuously supplied to maintain current flow. have been commissioned in the Netherlands and in Italy [192,195].
In the late 20th century, the power density of RED systems was too Existing pilot plants are limited to capacities of < 330 W and exhibit
low for practical purposes (~ 0.4 W/m2). Power densities of 1.27 W/m2 low power density of < 1 W/m2 [192]. Low-cost, high power density

10
F.E. Ahmed, et al. Desalination 495 (2020) 114659

RED stacks at the industrial scale are crucial for the commercialization
of RED systems [196]. Scaling up of RED systems depends on the via­
bility of RED pilot plants and improvement of hydrodynamics and
spacer designs for greater power density [197].

2.1.6.2. Pressure retarded osmosis. The first PRO system was designed
by Loeb and Norman in what they referred to as osmotic power plants
[198]. Both FO and PRO rely on the osmotic pressure difference
between two solutions separated by a semipermeable membrane that
allows the solvent to pass through but not the solute. The feed solution
is at a lower osmotic pressure than the draw solution, causing the draw
solution to ‘draw’ water from the feed to the draw via a positive osmotic
pressure gradient [199,200]. If the volume expansion of the draw
solution is restricted, the hydraulic pressure of the draw reservoir will
Fig. 11. FO-NF hybrid schematic as proposed by Zhao et al.
increase and the pressurized flow of water can be used to drive a
turbine and generate electricity [201,202].
A pilot facility based on PRO was opened by Statkraft in Tofte, but for brackish water desalination [210]. Fig. 11 shows a schematic of
Norway in 2009. For PRO plants to be commercially viable, the energy the hybrid process [210]. This schematic is applicable to most FO-NF
generated should be adequately higher than the energy consumed in processes where the diluted draw solution is treated with NF to recover
pretreatment, feed pumps, chemicals, etc. [203]. PRO membranes the draw solution and to produce clean water. Compared to stand-alone
would be required to generate 5 W/m2 and be comparable in price to RO, the FO-NF system requires less pressure, is less prone to fouling and
RO membranes [204,205]. The unavailability of low-cost high-per­ also exhibits higher flux recovery upon cleaning. All of these lower the
forming membranes eventually caused the plant to shut down in 2013. energy consumption of the desalination system although they did not
Since then, research efforts in the development low-cost membranes present SEC values.
with high power densities ≥5 W/m2 and using hypersaline waters have Considerable research effort has also been directed towards ferti­
intensified [206,207]. lizer-drawn forward osmosis (FDFO), in which fertilizers are used as
draw solution and the desalinated water can be used directly for irri­
3. Hybrid desalination technologies gation [211–218]. However, it has been noted that the concentration of
fertilizer in the diluted draw solution may still be too high for it to be
A hybrid desalination system is the integration of two or more de­ directly used as fertilizer [219]. Hence, hybridizing FDFO with se­
salination systems in order to enhance performance and/or reduce costs paration processes such as NF has been studied in the lab and pilot
compared to individual components. Hybrid processes seek to over­ scales. Kim et al. investigated an integrated FDFO-NF pilot in the Mil­
come the limitations of individual processes, improve overall system dura region of Australia [220,221]. They studied the effect of NF feed
performance and/or meet specific requirements using a synergistic flow rate, feed concentration and applied pressure on water flux and
approach. Since hybrid systems require substantial investment, the ammonium sulphate rejection, and found that feed concentration was
optimization of hybrid configurations is an important part of the pro­ the most dominant factor affecting these parameters [221]. The pilot-
cess. This section aims to critically review hybrid desalination tech­ scale hybrid system consisted of two spiral wound FO modules and one
nologies at the lab-scale and pilot scale as appropriate. spiral wound NF module [220]. To prevent the use of a two-pass NF,
which would increase costs, they suggested that the fertilizer should be
3.1. Forward osmosis-nanofiltration hybrid systems diluted to 70% of the initial concentration during FDFO which would
allow a much lower nutrient concentration in the product water and
For the recovery of draw solutes in FO, thermal and reverse osmosis make it suitable for direct irrigation. When applied to saline ground­
have both been suggested. The advantage of applying thermal desali­ water from coal mines, the group found that the pilot-scale FDFO-NF
nation is that separation is independent of feed concentration. On the hybrid system operating at NF recovery of 84% only required 1.08 kW/
other hand, RO theoretically and practically consumes less energy and m3 of energy, which is 14% and 21% less than that needed by MF-RO
requires low capital when compared to thermal systems. However, the and UF-RO systems with similar initial feed conditions [217,220]. Since
cost of operating an RO system is still relatively high. As a result, to MF and UF pretreatment lower the SEC of RO systems by reducing
keep FO sustainable, efforts have been put into hybridizing FO with membrane cleaning and replacement requirements [222], it can be
lower-energy alternatives for draw solute recovery. One such process is inferred that the FDFO-NF hybrid system operates at a lower SEC than
nanofiltration, in which 99% removal of divalent salts [208] can be standalone RO. FO-NF hybrid systems have also been applied for was­
achieved at a lower hydraulic pressure than RO. tewater treatment and reuse [223,224].
Tan and Ng proposed a hybrid forward osmosis – nanofiltration (FO-
NF) process for seawater desalination in which NF is applied to re­ 3.2. Electrodialysis – reverse osmosis hybrid systems
generate the draw solution and produce clean water [209]. Empha­
sizing the importance of choosing an appropriate draw solute, they RO systems, especially those treating challenging feeds, operate at
compared seven potential draw solutes: NaCl, KCl, MgCl2, CaCl2, limited recovery rates due to issues related to scaling and retentate
MgSO4, Na2SO4 and C6H12O6. A commercial cellulose triacetate (CTA) salinity [225,226]. Increasing recovery in RO systems requires multiple
membrane was used for FO while a thin-film composite membrane from stages and thus significantly increased capital and operation costs
GE Osmotics was applied to NF. They found that solute rejection in NF [227]. On the other hand, ED systems can be operated at higher re­
depended on the concentration of dilute draw solution. Only Na2SO4 as covery rates, although this is also eventually limited by scale formation.
a draw solute was able to maintain solute rejection of > 90% even as Electrical desalination systems such as ED also cannot achieve the high
the concentration of dilute draw solution increased. However, they salt rejection of RO membranes [228], and the magnitude of applied
tested each process separately and left optimizing the hybrid system, current and hence the energy consumed increases with the amount of
including draw solution concentration, and energy consumption ana­ salt rejected [228]. The high recovery rates possible in ED combined
lysis to future work. with high salt removal rates in RO provide an opportunity for hy­
Following this work, Zhao et al. designed a similar FO-NF process bridization to produce high-quality permeate at high recovery and low

11
F.E. Ahmed, et al. Desalination 495 (2020) 114659

Fig. 12. Standalone ED and hybrid ED-RO configurations [230].


Fig. 13. Flow diagram of ED-RO hybrid where ED is applied to high salinity
region and RO is applied to low salinity region. TS: treated stream, CS: con­
SEC [229].
centrated stream; EW: electrode wash; ED: electrodialysis; RO: reverse osmosis
The concept of a hybrid system of an electrodialysis-reverse osmosis
[229].
(ED-RO) hybrid system was first explored in 1981 by Schmoldt et al.
[118]. They proposed the use of ED as a second stage to control
permeate quality. However, the lack of high-flux, high-selectivity of the described ED-RO hybrid is shown in Fig. 13. The ED system is
membranes back then translated into a high energy consumption of operated as pretreatment in the high salinity region where high solution
7.94 kWh/m3 for RO of seawater of 45,000 ppm [118]. At the time, conductivity translates into greater efficiency. The RO unit operates in
they established that for a plant capacity of 1000 m3/day with feed the low salinity region where the osmotic pressure and thus the energy
TDS < 4000 mg/L, the investment cost for ED is lower than that for required for RO is low. Water was produced with TDS between 50 and
RO. Today, energy required for RO constitutes a small fraction of this 120 ppm at recoveries > 50% for all feed TDS between 2000 and
value, as seen in section 2.2. They noted that the development of high 4000 ppm. The lab-scale ED system consisted of 30 cell pairs and an
flux, high salt-rejection membranes would not only reduce the cost of effective membrane area of 200 cm2. RO alone requires 7.8 kWh/m3 to
the RO system, but also lower the TDS of incoming feed and hence the produce water at 10–20% recovery. The hybrid ED-RO system consumed
electrical energy consumption of the ED unit. 8–10 kWh/m3, but at a much higher recovery rate of 50–60% [229]. The
McGovern et al. modelled ED-RO systems for both brackish water energy consumption of the hybrid ED-RO system is comparable or
desalination [230] and for desalination of highly saline brines [231]. slightly higher than RO alone, but with 3–5 times the recovery of stan­
For brackish water desalination, the cost of ED can be reduced by hy­ dalone RO. Such high recovery rates are not feasible in standalone RO at
bridization with RO using the system shown in Fig. 12. The single-stage comparable SEC. A similar scheme that applies ED as pretreatment was
RO unit treats feed streams of 3000 ppm TDS at 50% recovery, and the proposed by Galama et al. for seawater desalination [234].
concentrate stream from RO becomes the feed to the ED at 6000 ppm Turek et al. [235] compared four different configurations for sea­
TDS [230]. They attributed two advantages of a hybrid system: re­ water desalination plants: single-stage standalone RO, hybrid ED-RO
duction in total membrane area as higher current densities (and system, a nanofiltration-RO (NF-SWRO) system and an NF-SWRO-ED
therefore salt removal rates) are possible at higher salinities, and since system. They compared SEC and recovery rates for the four systems. As
the final product is a blend of RO and ED permeates, the required seen in Table 4, a high recovery of 69% was achieved for the hybrid NF-
salinity of the ED permeate can be higher than the final product water SWRO-ED system at 6.90 kWh/m3. Although SEC of the standalone RO
salinity [230]. Percentage reduction in cost is greatest when both feed system was much less (2.76 kWh/m3), this single-stage RO system
and product salinity are low, and highest when salt removal occurs at operates at a recovery rate of only 43%. Furthermore, they highlight
high salinity. They suggest that if drinking water is required as product that the saturated brine produced by the system can be used in the
water is required, i.e. 500 ppm TDS, the relative cost from the single- production of salt or in the chlor-alkali industry.
stage RO would have to be 65% of that from stand-alone ED to justify So far, we have considered ED-RO systems in which ED was used as
hybridization. If product water requirements are more relaxed, then the the second stage in RO to either purify RO permeate or to desalt RO
savings in applying a hybrid system are even less. brine and blend it with RO permeate. In a contrasting approach,
Since ED is a salt removal process as opposed to water removal, Pellegrino et al. [236] present a theoretical hybrid system in which
treatment of RO concentrate with ED results in a highly concentrated hollow-fiber RO membranes are packed in a density similar to ED
brine that can be further used for salt production [232]. When applied spacers and act as spacer channels between AEMs and CEMs, as shown
to treatment brine with TDS 120,000 ppm, a similar arrangement of in Fig. 14. ED helps remove some salt from the RO concentrate stream
counterflow ED systems with RO proposed by the same group could not and also lowers concentration polarization via electrophoretic move­
perform at high recovery rates due to concentration differences be­ ment of ions [236]. Depending on the transport properties of both RO
tween streams, greater osmotic pressure and lower efficiency. The membranes and IEMs for ED, they identified feed water TDS and pro­
current density corresponding to lowest water cost was much higher duct recovery rates that would result in energy savings compared to RO
than that needed to maximize efficiency [231]. alone. However, such a prototype for this system has not yet been
The concept of incorporating ED on RO concentrate had been ap­ tested.
plied previously by Zhang et al. [233] who modified a UF-RO system
treating secondary effluent from a wastewater treatment plant. At a
Table 4
higher overall recovery rate of 95%, the quality of the ED effluent was Specific energy consumption and water recovery for SWRO, SWRO-ED, NF-
similar to the system without RO, but with higher total organic carbon SWRO and NF-SWRO-ED systems [235].
(TOC). This was resolved by adding an ozonation process to reduce the
System Energy consumption [kWh m−3] Water recovery [%]
accumulation of organic compounds [233]. However, this study did not
provide data on energy consumption. SWRO 2.76 42.6
Thampy et al. [229] demonstrated an ED-RO hybrid process for SWRO-ED 7.77 81.1
brackish water desalination (feed TDS: 2000–4000 ppm) at high recovery NF-SWRO 3.93 41.2
NF-SWRO-ED 6.90 69.0
such that the permeate from ED becomes the feed to RO. A flow diagram

12
F.E. Ahmed, et al. Desalination 495 (2020) 114659

Fig. 14. (a) Schematic cell pair in hybrid ED-RO by Pellegrino et al., (b) Cross-section view of stack configuration for hybrid ED-RO. M+ and X- are cations and
anions respectively. The rectangular elements are CEMs and AEMs. The circles are RO hollow fiber membranes. [120,236].

Although various studies have been implemented in which either 3.4.1. Draw solution for FO-MD hybrid processes
ED or RO are used as the second stage to the other technology, there are Zhao et al. exploited temperature-controlled osmotic pressure using
few experimental studies on a hybrid single-stage process. Additionally, a thermoresponsive copolymer as the draw solute in an FO-MD hybrid
the industrial use of ED-RO systems is not yet widespread as further process for seawater desalination [242]. They used poly(sodium
optimization of these processes is needed. A comparison of selected ED- styrene-4-sulfonate-co-n-isopropylacrylamide) (PSSS-PNIPAM) as the
RO studies has been presented in Table 5. draw solute. When dissolved in water, the polyelectrolyte PSSS gen­
erates high osmotic pressure, facilitating extraction of water from sea­
water at a flux of 4 LMH. When the diluted draw solution is heated to
3.3. Reverse osmosis – membrane distillation hybrid systems above the low critical solution temperature (LCST) of PNIPAM, ag­
glomeration of polymer chains causes a reduction in osmotic pressure
Low capital cost, high separation efficiency and ability to operate at and an increase in vapor pressure of the solution (Fig. 15). They note
high recovery make MD a promising candidate for hybrid separation that although MD requires 29 kWh/m3 to heat water to 50 °C making it
technologies [237,238]. There have been a few studies investigating the more energy-intensive than RO, capital cost is low and using waste heat
hybridization of MD and RO, where MD is used to treat the concentrate would make the process cost-effective [242]. The concept of a poly­
stream from the RO process. Choi et al. assessed the economic feasi­ electrolyte-based draw solution for FO-MD hybrid systems was pro­
bility of a hybrid membrane distillation-reverse osmosis (RO-MD) posed by Ge et al. [243]. Their group applied poly(acrylic acid) sodium
system for seawater desalination. They found that and RO-MD hybrid (PAAeNa) salt as the draw solute followed by MD for reconcentration
system or an MD stand-alone system could compete with RO only when for wastewater treatment, and found that efficient treatment required a
the recovery and flux are greater than that for RO and the thermal stabilized PAA-Na concentration of 0.48 g/mL at a temperature of
energy supplied for MD comes at a relatively low cost [239]. Naidu 66 °C. The same group developed a multi-charged hydroacid complex to
et al. [240] applied DCMD for wastewater RO concentrate treatment enable FO-MD hybrid systems for seawater/brackish water desalination
with granular activated carbon (GAC) pretreatment to mitigate the ef­ [244], as well as the removal of arsenic As(III) from water [245].
fects of organic fouling. With wastewater RO concentrate, MD is able to This allows improved separation via MD for recovery of the draw
achieve a high water recovery rate of 85%. Energy consumption for RO- solution. They attributed improved regeneration via MD to the ther­
MD systems for desalination is still unclear and should be further in­ moresponsive behavior of the draw solute to improved regeneration via
vestigated. MD. More studies on this of materials that respond to temperature in
the temperature range desirable for MD would assist in further opti­
mization of the FO-MD process. Guo et al. [246] applied Na+-functio­
3.4. Forward osmosis – membrane distillation hybrid systems nalized carbon quantum dots (Na_CQDs) as draw solute for FO as the
feed solution and found that at a concentration of 0.4 g/mL, Na_CQDs
Although FO is considered a low-energy technology relying on the resulted in higher flux than 2.0 M NaCl, a model FO draw solution.
spontaneous process of osmosis to transport water from the feed solu­ After each FO run, the diluted draw solution was reconcentrated using
tion to the concentrated draw solution, the cost of recovering the draw MD at 45 °C. FO. MD flux when DI water is applied as FO feed is shown
solution is often ignored. Practical implementation of FO desalination in Fig. 16b. Osmolality, or osmotic concentration, gives an indication of
requires that a simple, low-cost separation process be applied to pro­ the osmotic pressure of the solution. As mentioned earlier, a draw so­
duce clean water from the diluted draw solution. Continuous re­ lute should have a high osmotic pressure to induce an osmotic pressure
concentration of the draw solution would also keep costs low by pre­ gradient and draw water from the feed solution. Fig. 16a shows the
venting the need for frequent replace of the draw solution. MD was first effect of draw solute concentration on osmolality. When tested with
proposed for this purpose by Cath et al. in 2007 [241]. Since then, the seawater near the coast of Singapore, FO flux was lowered to ~10 LMH
simultaneous growth of FO and MD with novel membranes, new draw due to lower osmotic pressure gradient between seawater and the draw
solutes for FO, and an attempt to reduce energy consumption has been solution. The system could be further optimized by testing at different
complemented by research in hybrid FO-MD. MD feed temperatures. In addition, the study did not consider energy

13
F.E. Ahmed, et al. Desalination 495 (2020) 114659

Reference

[229]

[233]

[230]

[231]
alone: 7.8
ED-RO: 8–10
SEC (kWh/m3)

• RO



Performance at high recoveries is limited by

Fig. 15. Schematic of FO-MD hybrid system for seawater desalination using a
thermoresponsive copolymer PSSS-PNIPAM as draw solute [242].

consumption, which plays a critical role in the practical application.


concentration differences

Nguyen et al. reported the use of a high charge of phosphate as the


alone: 10–20%

draw solute in an FO-MD system. Three types of polytetra­


50–60%
alone: 75%

fluoroethylene (PTFE) MD membranes were investigated for the re­


• ED-RO: 95%

generation of the draw solute, and 100% recovery of the draw solute
Recovery rate

• ED-RO:

was achieved [247]. They found that the pH of the draw solution
• RO
• RO

strongly affected FO flux. The draw solute was used in the concentra­
50%

tion of high-nutrient sludge but could also be extended to desalination.


It is evident that more studies are needed to optimize draw solute
Hybrid preferred over ED alone only when product

concentration as well as MD operating conditions, especially tempera­


ture, to render FO-MD systems more efficient. Additionally, novel draw
solutes [248,249] should be further extended to desalination and stu­
died specifically for seawater and brackish water desalination applica­
tions. Data on energy consumption also remains vague.
TDS requirement is strict

3.4.2. Integrated FO-MD module design


Most studies on FO-MD hybrid systems explore limited integration
Product TDS (mg/L)

where the product of one process becomes the feed of another. Few
studies focus on a single hybrid process that combines elements of two
or more processes. Maintaining their position as pioneers in FO-MD
50–120

hybrid systems, Cath et al. were the first to report a more integrated
300

design with a single module in which both processes were sealed to­

gether, as shown in Fig. 17 [250].


Feed TDS (mg/L)

Kim et al. comment that the FO draw solution and MD feed solution
flowing in the same channel simultaneously could limit MD perfor­
2000–4000

2550–3550

120,000

mance due to heat transfer from MD feed to FO feed and MD permeate


3000

[251]. To tackle this issue, Ghaffour et al. [252] recently proposed an


osmotically and thermally isolated FO-MD integrated module. In this
ED of RO concentrate; ED product water blended with RO

design (Fig. 18) the FO DS and MD feed streams are separated using an
isolation barrier. This barrier does not prevent MD product (/brine) to
flow into the FO DS inlet. They suggest that this design maintains os­
motic and thermal isolation, allowing both processes to operate at high
ED as pretreatment to lower RO feed salinity
Key parameters from selected ED-RO hybridization studies.

efficiency [251].
They also established that the flow rate and concentration of the
initial DS determined long-term operation of the integrated system,
with greater flow rate and lower concentration pointing to greater flux
of the module [251]. Although such integrated designs of FO-MD hy­
permeate to produce water

Counterflow ED with RO

brids provide obvious benefits in the form of compactness and low


ED of RO concentrate

footprint, more work is needed in design improvement and in evalu­


ating the energy consumption to consumption of such modules.
Hybridization

Koo et al. found that the efficiency of FO-MD hybrid systems de­
pends on the draw solute transport through the MD membrane.
Selectivity of the draw solute can be increased by reducing the tem­
perature difference between MD feed and distillate, but this also lowers
MD flux [253].
Hypersaline brine

Many FO-MD systems have been applied to treatment of municipal


Brackish water

Brackish water
Wastewater

and industrial wastewater [254–261], and the efficiency of these sys­


Feed type

tems needs to be investigated with respect to desalination of different


Table 5

feed types.

14
F.E. Ahmed, et al. Desalination 495 (2020) 114659

Fig. 16. (a) Osmolality, or osmotic concentration, as a function of NaCQDs concentration; (b) FO flux with DI water as feed and 0.4 g/mL Na_CQDs as draw solution;
MD flux with diluted Na_CQDs as the feed solution [246]

Table 6
Comparison of SEC for selected FO hybrid systems [263].
System Specific energy Reference
consumption [kWh/m3]

Hybrid FO-Crystallization-RO (sodium 2.35 [264]


periodate as a draw solute)
Seawater RO 4.00–7.00 [265,266]
FO-RO hybrid for 25% total recovery 2.40 [267]
Hybrid FO-ED-RO 2.48 [263]

elevated energy costs. Bitaw et al. compared SEC for selected hybrid FO
systems, as shown in Table 6. They also developed a hybrid FO-ED-RO
system in which ammonium chloride was selected as the optimum draw
solute. The draw solution was recovered first by ED and the outlet from
ED became the RO feed. Their work showed that the relative energy
consumed by ED and RO depends on the intermediate concentration at
the ED outlet/RO inlet with higher concentrations increasing the con­
Fig. 17. Schematic of FO-MD module patented by Cath et al. [250,251].
tribution of RO to energy consumption.

3.6. Freeze desalination – membrane distillation hybrid systems

The advantage of MD is that the separation efficiency does not de­


pend on the salinity of the feed and a 100% separation efficiency is
possible even with hypersaline feed. Other processes such as FD and RO
can overcome their limitations in treating concentrated feed by hy­
bridization with MD [268]. Such hybridization of MD with other de­
salination processes to further concentrate brine allows the realization
of zero liquid discharge [269–271]. Additionally, waste cold energy can
be used for both freezing in FD and for cooling the distillate in MD.
These reasons have led to the development of hybrid FD-MD systems in
which the FD concentrate is treated by MD to produce water and/or the
same source of waste cold energy is utilized by both processes.
Fig. 18. Schematic of FO-MD integrated module proposed by Kim et al. [251] In 2012, Peng Wang and Tai-Shung Chung demonstrated the design
of an FD-MD hybrid in which seawater is first treated via FD, generating
3.5. Forward osmosis – electrodialysis purified water from melted ice [181]. The brine byproduct is treated by
MD, resulting in further concentration of brine and production of high
Electrodialysis is not commonly applied for the recovery of draw quality product water through condensation of water vapor on the
solutes in FO desalination. Zou et al. demonstrated a hybrid FO-ED distillate side of the MD membrane. They applied an LNG vaporizer to
system for wastewater treatment, with energy consumption of 0.72 provide cold energy to both the FD crystallizer and the MD distillate.
kWh/m3 [262]. However, extending to seawater desalination, a more Chang et al. evaluated the feasibility of such an FD-MD hybrid
concentrated draw solution will be required to overcome the high os­ system employing VMD for seawater desalination with a high water
motic pressure of seawater. Recovery of such a solution will lead to recovery of 74% [272]. They used regasification to provide cold energy
for FD whereas both the regasification process and solar panels were

15
F.E. Ahmed, et al. Desalination 495 (2020) 114659

Fig. 19. Schematic of hybrid freeze desalination, membrane distillation and crystallization system [273].

considered as sources of electricity to heat the VMD feed. Heat in­ Kwon et al. show that, compared to using seawater as the high
tegration between the feed, clean ice and vaporized steam can lower the concentration solution for RED, RO and FO brines can increase power
cooling energy requirement by 27%. generation by 1.5 and 2 times respectively. For their system, a max­
Lu et al. [273] developed a mathematical model for a hybrid system imum power of 1.48 W/m2 was obtained when RO brine was used while
consisting of FD and membrane distillation-crystallization (MD-C). MD- 1.86 W/m2 was obtained when FO brine was used, as seen in Fig. 21.
C is itself an integrated process combining MD and crystallization to Applying RED for energy recovery in each of these processes could
recover both clean water and valuable resources in crystal form [274]. reduce energy cost by 7.8% for RO and 13.5% for FO [276]. They used
In crystallization, specific techniques such as reducing temperature and concentrations of 0.6 mol/L, 1.2 mol/L and 2.4 mol/L for seawater, RO
removing solvent are applied to control saturation and extract resources brine and FO brine respectively.
from brine. The effect of various parameters such as MD feed tem­ Wang et al. developed a hybrid RED/ED system to recover salinity
perature, feed TDS, distillate temperature and FD recovery ratio on gradient power (SGP) and phenol in the treatment of phenol-containing
energy consumption was evaluated for the hybrid FD-MD-C system. wastewater treatment [277]. The advantage of this hybridization is
Optimal parameters for minimizing SEC were MD feed temperature of lower energy consumption than standalone ED and greater water pro­
80 °C, distillate temperature of 5 °C, feed TDS of 26,500 ppm and an FD duction rate than standalone RED. The energy consumption of the hy­
recovery of 30% [273]. However, at 810 kWh/m3 of hot energy and brid system is 17.65 kWh/m3 which is significantly lower than 25.32
830 kWh/m3 of cold energy with heat integration, the hybrid FD-MD-C kWh/m3 for a standalone ED system [277]. Interestingly, this study
system remains energy-intensive. A schematic of the hybrid system is showed that the type of wastewater stream can be optimized to achieve
shown in Fig. 19. high desalination rate and SGP.
Tufa et al. proposed a hybrid membrane distillation-reverse elec­
4. Future perspectives trodialysis (MD-RED) system to produce fresh water and energy from
seawater [278]. Feeding the RED system with DCMD brine at different
4.1. Hybridization of salinity gradient power and water desalination temperatures and flow velocities could generate a power density of
technologies 0.9–2.4 W/m2 per cell. Mei and Tang explored the implications of co-
locating RED with RO desalination [279]. Optimizing the RED system,
Integrated application of salinity gradient energy harvesting tech­ they suggested that maximum power was obtained when the low sali­
nologies with desalination technologies serves two noteworthy pur­ nity stream was of moderate salinity, i.e. 0.01–0.02 M NaCl whereas an
poses. First, high salinity brine from desalination processes can be used increase in salinity and feed temperature of the high salinity stream led
as the concentrated solution for power production. In RO, this equates to improved power generation. As post-treatment for RO, RED could
to further recovering brine energy in addition to that recovered with reduce almost 50% of TDS from the RO brine.
ERDs. This also provides a method of diluting the brine before disposal. Li et al. modelled various configurations for the RED-RO hybrid
Second, combining energy harvesting with desalination allows the po­ process in terms of energy consumption [280]. Compared to conven­
tential development of self-sufficient desalination systems, which en­ tional SWRO, they found that hybridization with RED results in much
ables production of freshwater at considerably lower costs. Research in lower energy consumption for desalination and the possibility of gen­
this area are still in its infancy. This section explores such systems in erating energy. For practical applications involving integration of RED
with RED or PRO are integrated with traditional and emerging desali­ with desalination processes, combinations with multiple desalination
nation processes. technologies may need to be explored. Additionally, the limitations of
RED systems have to be overcome through the improvement of elec­
4.1.1. Reverse electrodialysis hybrid systems trodes, membranes and optimization of feed solutions [281].
Highly saline brine streams from desalination systems such as RO
and ED can be fed into the high concentration compartments of RED 4.1.2. Pressure retarded osmosis hybrid systems
systems. A conceptual illustration developed by Tufa et al. for the in­ PRO can be used as an energy recovery system in desalination
tegration of RED systems with other desalination technologies can be processes providing benefits similar to those outlined for RED above.
seen in Fig. 20. They cite several advantages for the integration of RED This section reviews hybrid systems in which PRO is integrated with
with desalination technologies, namely: use of high salinity brine as selected desalination processes to harness energy and produce water at
high concentration stream of RED which reduces ohmic losses, appli­ a lower cost compared to standalone desalination systems.
cation of power generated by RED to reduce energy consumption of Prante et al. developed a reverse osmosis-pressure retarded osmosis
desalination, no extra pumps to recirculate seawater and brine streams (RO-PRO) system in which the energy generated in PRO offsets the
through RED, no additional pretreatment and dilution of brine from energy required for RO. They found that the minimum theoretical SEC
desalination in RED which reduces environmental impact [275]. for the modelled system was 1.2 kWh/m3 operating at 50% RO

16
F.E. Ahmed, et al. Desalination 495 (2020) 114659

Fig. 20. Conceptual illustration of integration of RED with desalination technologies [275].

Fig. 22. Specific energy consumption of RO-PRO as a function of PRO dilution


for several RO recoveries for ideal (thermodynamically reversible) and model
RO-PRO systems [282].

mathematical model with seawater and urban wastewater as feed so­


lutions [284]. Out of six potential RO-PRO systems, he found that the
one in which the diluted PRO draw outlet is mixed with RO feed brings
down SEC by 49% when compared to a standalone RO system, without
the need for additional ERD units. Wan and Chung also modelled PRO
as an energy recovery system together with a pressure exchanger (PX)
ERD in a seawater RO system [285]. Using both PX and PRO for energy
recovery brings the SEC of RO to 1.08 kWh/m3 and 1.15 kWh/m3 at 25
and 50% recovery, respectively. PRO as an energy recovery system was
Fig. 21. Experimental and theoretical maximum power density for various previously proposed by Sharqawy et al. [286]
concentrated solutions applied as the high concentration solution in RED [276]. Although researchers have investigated the design of RO-PRO sys­
tems [287] many studies are still at the modelling stage with little
recovery [282], which is a 40% energy reduction from an RO SEC of 2.0 practical implementation to date. Commercialization of PRO-RO re­
kWh/m3. Fig. 22 shows the variation of SEC with PRO dilution for the mains a possibility but requires focus on experimental lab-scale and
RO-PRO system at different RO recovery rates, for both the thermo­ pilot-scale studies. Additionally, the development of low-cost high
dynamically reversible, or ideal, case and the modelled case. power density membranes for PRO will also be crucial in making RO-
Kim et al. also modelled an RO-PRO system and evaluated the water PRO systems economically viable solutions for reducing the energy
and energy return rate for such a system [283]. They found that a de­ consumption of RO desalination.
crease in RO plant size leads to reduction in water and energy return Lin et al. integrated PRO with MD by coupling the mass and energy
rate, whereas the PRO plant size does not play a significant role. flows between MD and PRO stages at different heat source temperatures
Senthil optimized the design of a hybrid RO-PRO system using a between 40 and 80 °C [288]. They suggested that there is an optimal
relative flow rate between the MD permeate and feed streams that

17
F.E. Ahmed, et al. Desalination 495 (2020) 114659

reduces overall energy efficiency of the PRO-MD system. However, they with respect to methods used to optimize hybrid systems. We first
cautioned that low thermal efficiencies should be taken into con­ provide an overview of established and emerging desalination tech­
sideration when choosing the heat source for such a system. nologies, focusing on their current status and future potential in terms
The prospect of multihybrid desalination systems with PRO for of energy use. We then review state-of-the-art hybridization of various
energy recovery has not been widely explored. Kim et al. evaluated the desalination processes, identifying specific applications, gaps in litera­
performance of an RO-MD-PRO system numerically [289] by com­ ture and future direction of the hybridization of desalination technol­
bining previously validated models of each of the three processes. In the ogies, with each other and with salinity gradient power technologies as
proposed hybrid process, RO brine is partially used as MD feed and the well. The low recovery of seawater reverse osmosis can be overcome by
MD concentrate is mixed with the remaining RO brine and employed as coupling with electrodialysis systems. On the other hand, the im­
the PRO draw solution. They considered the effect of RO brine division practicality of treating hypersaline solutions with reverse osmosis
ratio, plant size and cost of MD thermal energy as factors affecting the means that processes whose separation efficiency does not depend on
performance of the hybrid plant. They determined that, for a plant size feed concentration, such as membrane distillation, may be used in
of 2000 m3/day, the hybrid system could generate clean water at an conjunction for the treatment of high salinity feeds and/or RO brine.
SEC of 1.61–1.78 kWh/m3 depending on brine division ratio, compared The appropriate choice of separation process employed in osmotically
to 1.9 kWh/m3 required for stand-alone RO, provided that free or low- driven separation processes such as forward osmosis is key in de­
cost thermal energy is supplied for the MD component. termining overall energy consumption and thus commercial viability.
When selecting a salinity-gradient power technology, it is important This separation process is applied to separate water from the diluted
to assess the strengths of both PRO and RED. PRO is attractive for draw solution and recover the draw solution. As RO applies higher
power generation using concentrated saline brines due to high power hydraulic pressures than nanofiltration, the latter has emerged as a
density and energy recovery [190,290]. On the other hand, RED is potential process for integration with FO. In cases where waste or low-
suitable for energy generation with moderately concentrated solutions grade heat is available, MD has also been investigated as a viable se­
such as seawater and river water. However, there is immense room for paration process for FO. There has been considerable effort in opti­
improvement for both technologies. For example, PRO will benefit from mizing draw solutions for specific hybrid processes involving FO such
the development of membranes with greater permeability. Although as forward osmosis-membrane distillation (FO-MD). This includes the
improving the characteristics of IEMs is also crucial for RED develop­ development of thermally responsive draw solutes. Apart from factors
ment, other elements such as spacer and system design will allow RED that control the energy consumption of each of the processes involved
to achieve better performance [190]. in the hybrid system, the design of the integrated module plays an
important role in the overall energy consumption. We also review re­
4.2. Renewable energy for hybrid desalination systems cent advances in power generation from salinity gradients using pres­
sure retarded osmosis and reverse electrodialysis and recent advances
The decreasing cost of renewable energy especially photovoltaic and challenges in applying these technologies to recover energy and
(PV) technology has led to several advances in integrated renewable further lower the cost of desalination and/or simultaneously generate
energy with desalination technologies. Such systems seek to lower the energy and power. Further research focused on the following key areas,
carbon footprint, providing a sustainable and efficient means of energy as identified in this review, is likely to drive progress in the area of
supply. To date, integration of renewable energy systems with hybrid hybrid desalination technologies towards more energy efficient sys­
and multihybrid desalination technologies remains an unexplored do­ tems.
main. Today, PV powered RO is a technologically mature and com­
mercially available system, the energy cost of which depends greatly on • In many studies, hybridization is carried out by one process followed
the energy consumption of each of its components [22]. As we have by another. Very few designs for integrated system coupling have
discussed earlier, the energy consumption of RO can further be reduced been suggested and optimized in literature, with even fewer proto­
through hybridization with other processes. This would help further types. In some cases, each process is optimized separately but the
lower the cost of PV-powered RO. effect on the integrated system has not been investigated. For ex­
Similarly, solar thermal energy as a renewable low-cost source of ample, in FO-NF hybrids, optimization of the draw solute type and
heating could be beneficial for MD hybrid systems [291], where the solution concentration with respect to the performance of the hybrid
cost of heat supply greatly affects the energy efficiency of the entire system is still needed.
hybrid system. • For ED-RO hybrid systems, more work is needed targeting the use of
Although there are limitations such as seasonal supply and addi­ seawater as the feed solution.
tional cost of batteries for storage, the integration of renewable energy • Energy consumption of hybrid systems remains unclear as many
sources with hybrid technologies could potentially be a viable pathway studies with different operating conditions cannot be compared. The
to energy efficient desalination and is an area that needs further in­ benefits and drawbacks of hybridization in terms of energy con­
vestigation. SEC of hybrid desalination systems and of the renewable sumption remains unclear especially for RO-MD, FO-MD and FD-MD
energy sources must be optimized through both modelling and ex­ processes. Research efforts should be directed towards design im­
perimental studies before the commercial applicability of these systems provement and evaluation of energy consumption.
is explored. • In the case of MD hybrid systems, often the effect of MD feed tem­
perature on the energy efficiency of hybrid processes is neglected.
5. Conclusion • Combinations of multiple desalination technologies in multi-stage
multi-hybrid systems should be considered to reduce energy con­
The energy consumption of both thermal and membrane processes sumption and meet product water targets. This would pave the way
is widely acknowledged as a drawback of established desalination for upscaling of hybrid desalination systems.
technologies and a hindrance to the growth of emerging technologies. • Overcoming limitations related to salinity gradient power technol­
Novel configurations, improvement in process design and the devel­ ogies and pushing them towards commercialization would render
opment of high-performance materials have led to significant reduction hybrid desalination more economically viable as salinity gradient
in energy costs over the years. However, further reduction in energy power could be used as energy recovery systems on their own or
can be sought through hybridization of desalination processes to with other ERDs in desalination systems. In addition to the devel­
overcome limitations of each process and benefit from their strengths. opment of low-cost high power density membranes and systems for
This paper provides a critical review of hybrid desalination processes reverse electrodialysis and pressure retarded osmosis, the

18
F.E. Ahmed, et al. Desalination 495 (2020) 114659

implementation and testing of pilot plants would speed up their [2] V. Barawkar, Growth expected for global water desalination market from 2018 to
transition and make them more commercially viable for industrial 2025, Water Technology, 2019.
[3] A.K. Plappally, J.H. Lienhard V, Energy requirements for water production,
scale operation with other desalination processes.

treatment, end use, reclamation, and disposal, Renew. Sust. Energ. Rev. 16 (7)
More studies are needed to evaluate the impact of applying re­ (2012) 4818–4848.
newable energy systems with hybrid desalination technologies. [4] K. Wang, et al., Mechanical properties of water desalination and wastewater
treatment membranes, Desalination 401 (2017) 190–205.
[5] M. Busch, W.E. Mickols, Reducing energy consumption in seawater desalination,
Abbreviations Desalination 165 (2004) 299–312.
[6] A. Shrivastava, S. Rosenberg, M. Peery, Energy efficiency breakdown of reverse
osmosis and its implications on future innovation roadmap for desalination,
MSF multi-stage flash Desalination 368 (2015) 181–192.
MED multi-effect distillation [7] D. Zarzo, D. Prats, Desalination and energy consumption. What can we expect in
RO reverse osmosis the near future? Desalination 427 (2018) 1–9.
[8] Y. Ghalavand, M.S. Hatamipour, A. Rahimi, A review on energy consumption of
FD freeze desalination
desalination processes, Desalin. Water Treat. 54 (6) (2015) 1526–1541.
FO freeze osmosis [9] H. Nassrullah, et al., Energy for desalination: a state-of-the-art review,
SWRO seawater reverse osmosis Desalination 491 (2020) 114569.
BWRO brackish water reverse osmosis [10] C. Charcosset, A review of membrane processes and renewable energies for de­
salination, Desalination 245 (1–3) (2009) 214–231.
ED electrodialysis [11] E. Mathioulakis, V. Belessiotis, E. Delyannis, Desalination by using alternative
MD membrane distillation energy: review and state-of-the-art, Desalination 203 (1–3) (2007) 346–365.
PRO pressure retarded osmosis [12] A. Subramani, et al., Energy minimization strategies and renewable energy utili­
zation for desalination: a review, Water Res. 45 (5) (2011) 1907–1920.
MSF-OT once-through MSF [13] M.A. Eltawil, Z. Zhengming, L. Yuan, A review of renewable energy technologies
BM-MSF brine mixing MSF integrated with desalination systems, Renew. Sust. Energ. Rev. 13 (9) (2009)
TBT top brine temperature 2245–2262.
[14] N. Ghaffour, et al., Renewable energy-driven desalination technologies: a com­
ERD energy recovery device prehensive review on challenges and potential applications of integrated systems,
TDS total dissolved solids Desalination 356 (2015) 94–114.
PX pressure exchanger [15] H. Sharon, K. Reddy, A review of solar energy driven desalination technologies,
Renew. Sust. Energ. Rev. 41 (2015) 1080–1118.
PT Pelton turbine [16] A. Alkaisi, R. Mossad, A. Sharifian-Barforoush, A review of the water desalination
DWEER dual work exchanger energy recovery systems integrated with renewable energy, Energy Procedia 110 (2017) 268–274.
APM axial piston motors [17] C. Li, Y. Goswami, E. Stefanakos, Solar assisted sea water desalination: a review,
Renew. Sust. Energ. Rev. 19 (2013) 136–163.
RE renewable energy
[18] M.T. Ali, H.E. Fath, P.R. Armstrong, A comprehensive techno-economical review
PV photovoltaic of indirect solar desalination, Renew. Sust. Energ. Rev. 15 (8) (2011) 4187–4199.
UF ultrafiltration [19] Q. Ma, H. Lu, Wind energy technologies integrated with desalination systems:
MF microfiltration review and state-of-the-art, Desalination 277 (1–3) (2011) 274–280.
[20] M. Chandrashekara, A. Yadav, Water desalination system using solar heat: a re­
NF nanofiltration view, Renew. Sust. Energ. Rev. 67 (2017) 1308–1330.
CDI capacitive deionization [21] L. Garcia-Rodriguez, Seawater desalination driven by renewable energies: a re­
CEM cation exchange membrane view, Desalination 143 (2) (2002) 103–113.
[22] F.E. Ahmed, R. Hashaikeh, N. Hilal, Solar powered desalination – technology,
AEM anion exchange membranes energy and future outlook, Desalination 453 (2019) 54–76.
LCD limiting current density [23] A. Al Bloushi, et al., Chapter 3 - environmental impact and technoeconomic
DCMD direct contact membrane distillation analysis of hybrid MSF/RO desalination: the case study of Al Taweelah A2 plant,
in: V.G. Gude (Ed.), Sustainable Desalination Handbook, Butterworth-Heinemann,
AGMD air-gap membrane distillation 2018, pp. 55–97.
VMD vacuum membrane distillation [24] O.A. Hamed, Overview of hybrid desalination systems — current status and future
SGMD sweeping gas membrane distillation prospects, Desalination 186 (1) (2005) 207–214.
[25] M.A. Darwish, Critical comparison between energy consumption in large capacity
DS draw solution
reverse osmosis (RO) and multistage flash (MSF) seawater desalting plants,
LNG liquefied natural gas Desalination 63 (1987) 143–161.
RED reverse electrodialysis [26] A.M.R. Al-Marafie, Prospects of hybrid RO-MSF desalting plants in Kuwait,
Desalination 72 (3) (1989) 395–404.
IEM ion exchange membrane
[27] E. El-Sayed, et al., Pilot study of MSFRO hybrid systems, Desalination 120 (1)
FDFO fertilizer-drawn forward osmosis (1998) 121–128.
GAC granular activated carbon [28] S.P. Agashichev, M.E. El-Dahshan, Reverse osmosis incorporated into existing
PSSS-PNIPAM poly(sodium styrene-4-sulfonate-co-n-isopropylacrylamide) cogenerating systems as a sustainable technological alternative for United Arab
Emirates, Desalination 157 (1) (2003) 33–49.
LCST low critical solution temperature [29] Z.K. Al-Bahri, W.T. Hanbury, T. Hodgkiess, Optimum feed temperatures for sea­
PAA-Na poly(acrylic acid) sodium water reverse osmosis plant operation in an MSF/SWRO hybrid plant, Desalination
Na_CQD Na+-functionalized carbon quantum dots 138 (1) (2001) 335–339.
[30] A.M. Helal, et al., Optimal design of hybrid RO/MSF desalination plants part II:
PTFE polytetrafluoroethylene (PTFE) results and discussion, Desalination 160 (1) (2004) 13–27.
[31] I. Kamal, Myth and reality of the hybrid desalination process, Desalination 230 (1)
Declaration of competing interest (2008) 269–280.
[32] N. Ghaffour, T.M. Missimer, G.L. Amy, Technical review and evaluation of the
economics of water desalination: current and future challenges for better water
The authors of this manuscript declare that there is no conflict of supply sustainability, Desalination 309 (2013) 197–207.
[33] W.L. Ang, et al., A review on the applicability of integrated/hybrid membrane
interest. processes in water treatment and desalination plants, Desalination 363 (2015)
2–18.
Acknowledgement [34] L. Chekli, et al., A comprehensive review of hybrid forward osmosis systems:
performance, applications and future prospects, J. Membr. Sci. 497 (2016)
430–449.
The authors would like to acknowledge the support of NYU Abu [35] J. Bundschuh, J. Hoinkis, Renewable Energy Applications for Freshwater
Dhabi for funding this work. Production, CRC Press, 2012.
[36] S. Avlonitis, K. Kouroumbas, N. Vlachakis, Energy consumption and membrane
replacement cost for seawater RO desalination plants, Desalination 157 (1) (2003)
References 151–158.
[37] A.D. Khawaji, I.K. Kutubkhanah, J.-M. Wie, Advances in seawater desalination
[1] WWAP, The United Nations World Water Development Report 2018: Nature-based technologies, Desalination 221 (1) (2008) 47–69.
Solutions, UNESCO, Paris, 2018. [38] K. Thu, et al., A hybrid multi-effect distillation and adsorption cycle, Appl. Energy

19
F.E. Ahmed, et al. Desalination 495 (2020) 114659

104 (2013) 810–821. [76] A.S. Stillwell, M.E. Webber, Predicting the specific energy consumption of reverse
[39] H. Ettouney, Conventional thermal processes, in: G. Micale, L. Rizzuti, A. Cipollina osmosis desalination, Water 8 (12) (2016) 601.
(Eds.), Seawater Desalination: Conventional and Renewable Energy Processes, [77] H. Fath, A. Sadik, T. Mezher, Present and future trend in the production and en­
Springer Berlin Heidelberg, Berlin, Heidelberg, 2009, pp. 17–40. ergy consumption of desalinated water in GCC countries, Int. J. Therm. Environ.
[40] J.B. Tonner, S. Hinge, C. Legorreta, Plates—the next breakthrough in thermal Eng. 5 (2) (2013) 155–165.
desalination, Desalination 134 (1–3) (2001) 205–211. [78] M.K. Wafi, et al., Nanofiltration as a cost-saving desalination process, SN Appl. Sci.
[41] V.G. Gude, Desalination and sustainability – an appraisal and current perspective, 1 (7) (2019) 751.
Water Res. 89 (2016) 87–106. [79] L.K. Wang, et al., Membrane and Desalination Technologies, Humana Press, 2010.
[42] A. Al-Karaghouli, L.L. Kazmerski, Energy consumption and water production cost [80] C. Kazner, T. Wintgens, P. Dillon, Water Reclamation Technologies for Safe
of conventional and renewable-energy-powered desalination processes, Renew. Managed Aquifer Recharge, IWA Publishing, 2012.
Sust. Energ. Rev. 24 (2013) 343–356. [81] L. Henthorne, B. Boysen, State-of-the-art of reverse osmosis desalination pre­
[43] I.S. Al-Mutaz, A.M. Al-Namlah, Characteristics of dual purpose MSF desalination treatment, Desalination 356 (2015) 129–139.
plants, Desalination 166 (2004) 287–294. [82] P.H. Wolf, S. Siverns, S. Monti, UF membranes for RO desalination pretreatment,
[44] M.A. Darwish, N.M. Al-Najem, Energy consumption by multi-stage flash and re­ Desalination 182 (1) (2005) 293–300.
verse osmosis desalters, Appl. Therm. Eng. 20 (5) (2000) 399–416. [83] I. Sutzkover-Gutman, D. Hasson, Feed water pretreatment for desalination plants,
[45] M.A. Darwish, F. Al Asfour, N. Al-Najem, Energy consumption in equivalent work Desalination 264 (3) (2010) 289–296.
by different desalting methods: case study for Kuwait, Desalination 152 (1) (2003) [84] B.A. Abdelkader, M.A. Antar, Z. Khan, Nanofiltration as a pretreatment step in
83–92. seawater desalination: a review, Arab. J. Sci. Eng. 43 (9) (2018) 4413–4432.
[46] H. El-Naser, Management of Scarce Water Resources: A Middle Eastern [85] L.W. Jye, A.F. Ismail, Nanofiltration Membranes : Synthesis, Characterization, and
Experience, WIT Press, 2009. Applications, CRC Press LLC, Boca Raton, UNITED KINGDOM, 2016.
[47] D. Wogan, S. Pradhan, S. Albardi, GCC Energy System Overview - 2017, King [86] F.E. Ahmed, et al., Mathematical and optimization modelling in desalination:
Abdullah Petroleum Studies and Research Center (KAPSARC), 2017. state-of-the-art and future direction, Desalination 469 (2019) 114092.
[48] B. Rahimi, H.T. Chua, Chapter 1 - introduction to desalination, in: B. Rahimi, [87] W.R. Bowen, A.W. Mohammad, Characterization and prediction of nanofiltration
H.T. Chua (Eds.), Low Grade Heat Driven Multi-effect Distillation and membrane performance—a general assessment, Chem. Eng. Res. Des. 76 (8)
Desalination, Elsevier, Amsterdam, 2017, pp. 1–17. (1998) 885–893.
[49] M. Al-Shammiri, M. Safar, Multi-effect distillation plants: state of the art, [88] O. Agboola, et al., Theoretical performance of nanofiltration membranes for
Desalination 126 (1) (1999) 45–59. wastewater treatment, Environ. Chem. Lett. 13 (1) (2015) 37–47.
[50] X. Wang, et al., Low grade heat driven multi-effect distillation technology, Int. J. [89] Pore Size, Pore size distribution, and roughness at the membrane surface, in:
Heat Mass Transf. 54 (25) (2011) 5497–5503. K.C. Khulbe, C.Y. Feng, T. Matsuura (Eds.), Synthetic Polymeric Membranes:
[51] J.A. Fleck, The influence of pressure on boiling water reactor dynamic behavior at Characterization by Atomic Force Microscopy, Springer Berlin Heidelberg, Berlin,
atmospheric pressure, Nucl. Sci. Eng. 9 (2) (1961) 271–280. Heidelberg, 2008, pp. 101–139.
[52] M.A. Darwish, A. Alsairafi, Technical comparison between TVC/MEB and MSF, [90] A. Street, et al., Nanotechnology Applications for Clean Water: Solutions for
Desalination 170 (3) (2004) 223–239. Improving Water Quality, Elsevier Science, 2018.
[53] A. Ophir, F. Lokiec, Advanced MED process for most economical sea water desa­ [91] B. Van der Bruggen, M. Mänttäri, M. Nyström, Drawbacks of applying nanofil­
lination, Desalination 182 (1) (2005) 187–198. tration and how to avoid them: a review, Sep. Purif. Technol. 63 (2) (2008)
[54] N. Hilal, A.F. Ismail, C. Wright, Membrane Fabrication, Taylor & Francis Group, 251–263.
London, UNITED KINGDOM, 2015. [92] A.W. Mohammad, et al., Nanofiltration membranes review: recent advances and
[55] M. Elimelech, W.A. Phillip, The future of seawater desalination: energy, tech­ future prospects, Desalination 356 (2015) 226–254.
nology, and the environment, Science 333 (6043) (2011) 712–717. [93] N. Song, et al., A review of graphene-based separation membrane: materials,
[56] L. Rizzuti, H.M. Ettouney, A. Cipollina, Solar Desalination for the 21st Century: A characteristics, preparation and applications, Desalination 437 (2018) 59–72.
Review of Modern Technologies and Researches on Desalination Coupled to [94] S. Tul Muntha, A. Kausar, M. Siddiq, Advances in polymeric nanofiltration
Renewable Energies, Springer, 2007. membrane: a review, Polym.-Plast. Technol. Eng. 56 (8) (2017) 841–856.
[57] J. Kim, et al., A comprehensive review of energy consumption of seawater reverse [95] A. Wahab Mohammad, N. Hilal, M. Nizam Abu Seman, A study on producing
osmosis desalination plants, Appl. Energy 254 (2019) 113652. composite nanofiltration membranes with optimized properties, Desalination 158
[58] N. Voutchkov, Energy use for membrane seawater desalination – current status (1) (2003) 73–78.
and trends, Desalination 431 (2018) 2–14. [96] M.A. Seman, M. Khayet, N. Hilal, Development of antifouling properties and
[59] R. Semiat, Energy issues in desalination processes, Environ. Sci. Technol. 42 (22) performance of nanofiltration membranes modified by interfacial polymerisation,
(2008) 8193–8201. Desalination 273 (1) (2011) 36–47.
[60] S. Mirza, Reduction of energy consumption in process plants using nanofiltration [97] N. Hilal, M. Khayet, C.J. Wright, Membrane Modification : Technology and
and reverse osmosis, Desalination 224 (1) (2008) 132–142. Applications, Taylor & Francis Group, Baton Rouge, UNITED STATES, 2012.
[61] C.P. Koutsou, et al., Analysis of temperature effects on the specific energy con­ [98] H. Strathmann, Electrodialysis, a mature technology with a multitude of new
sumption in reverse osmosis desalination processes, Desalination 476 (2020) applications, Desalination 264 (3) (2010) 268–288.
114213. [99] N.A.A. Qasem, B.A. Qureshi, S.M. Zubair, Improvement in design of electrodialysis
[62] S.K. Patel, et al., The relative insignificance of advanced materials in enhancing desalination plants by considering the Donnan potential, Desalination 441 (2018)
the energy efficiency of desalination technologies, Energy Environ. Sci. 20 (2020) 62–76.
1694–1710, https://doi.org/10.1039/D0EE00341G. [100] A. Campione, et al., A hierarchical model for novel schemes of electrodialysis
[63] A.J. Karabelas, et al., Analysis of specific energy consumption in reverse osmosis desalination, Desalination 465 (2019) 79–93.
desalination processes, Desalination 431 (2018) 15–21. [101] A. Merkel, A.M. Ashrafi, An investigation on the application of pulsed electro­
[64] X. Wang, et al., Comparative study on stand-alone and parallel operating schemes dialysis reversal in whey desalination, Int. J. Mol. Sci. 20 (8) (2019) 1918.
of energy recovery device for SWRO system, Desalination 254 (1–3) (2010) [102] J.J. Schoeman, Evaluation of electrodialysis for the treatment of a hazardous
170–174. leachate, Desalination 224 (1) (2008) 178–182.
[65] B. Peñate, L. García-Rodríguez, Energy optimisation of existing SWRO (seawater [103] Y. Oren, et al., Modified heterogeneous anion-exchange membranes for desalina­
reverse osmosis) plants with ERT (energy recovery turbines): technical and ther­ tion of brackish and recycled water, Environ. Eng. Sci. 19 (6) (2002) 512–529.
moeconomic assessment, Energy 36 (1) (2011) 613–626. [104] W.S. Walker, Y. Kim, D.F. Lawler, Treatment of model inland brackish ground­
[66] R.L. Stover, Retrofits to improve desalination plants, Desalin. Water Treat. 13 water reverse osmosis concentrate with electrodialysis—part I: sensitivity to su­
(1–3) (2010) 33–41. perficial velocity, Desalination 344 (2014) 152–162.
[67] R.L. Stover, Seawater reverse osmosis with isobaric energy recovery devices, [105] Y. Tanaka, Ion Exchange Membranes: Fundamentals and Applications, Elsevier
Desalination 203 (1) (2007) 168–175. Science, 2015.
[68] W. School, ‘New Water’ Offers an Ocean of Hope, Knowledge@Wharton, 2013. [106] A. Campione, et al., Electrodialysis for water desalination: a critical assessment of
[69] J.P. MacHarg, B. Sessions, Modular pumps bring efficiencies, IDA J. Desalin. Water recent developments on process fundamentals, models and applications,
Reuse 5 (2) (2013) 106–113. Desalination 434 (2018) 121–160.
[70] Arenas Urrea, S., et al., Technical review, evaluation and efficiency of energy [107] S. Al-Amshawee, et al., Electrodialysis desalination for water and wastewater: a
recovery devices installed in the Canary Islands desalination plants. Desalination, review, Chem. Eng. J. 380 (2020) 122231.
2018. [108] V. Montiel, et al., Membrane processes, Electrodialysis, in: G. Kreysa, K.-i. Ota,
[71] J. Rheinländer, D. Geyer, Photovoltaic reverse osmosis and Electrodialysis, in: R.F. Savinell (Eds.), Encyclopedia of Applied Electrochemistry, Springer New
G. Micale, L. Rizzuti, A. Cipollina (Eds.), Seawater Desalination: Conventional and York, New York, NY, 2014, pp. 1224–1229.
Renewable Energy Processes, Springer Berlin Heidelberg, Berlin, Heidelberg, [109] Interior, U.S.D.o.t. and U.S.O.o.S. Water, Saline Water Conversion Report for,
2009, pp. 189–211. United States Department of the Interior, 1965.
[72] ERIEnergy, How the PX Pressure Exchanger® Works, (2013). [110] E.T. Sajtar, D.M. Bagley, Electrodialysis reversal: process and cost approximations
[73] G. Migliorini, E. Luzzo, Seawater Reverse Osmosis Plant Using the Pressure for treating coal-bed methane waters, Desalin. Water Treat. 2 (1–3) (2009)
Exchanger for Energy Recovery: A Calculation Model, 165 (2004), pp. 289–298. 284–294.
[74] C. Harris, Energy recovery for membrane desalination, Desalination 125 (1–3) [111] Chapter 1 - Overview of Ion-Exchange Membrane Processes, in: H. Strathmann
(1999) 173–180. (Ed.), Membrane Science and Technology, Elsevier, 2004, pp. 1–22.
[75] A.A. Alsarayreh, et al., Evaluation and minimisation of energy consumption in a [112] A.M. Bernardes, M.A.S. Rodrigues, J.Z. Ferreira, Electrodialysis and Water Reuse:
medium-scale reverse osmosis brackish water desalination plant, J. Clean. Prod. Novel Approaches, Springer, Berlin Heidelberg, 2013.
248 (2020) 119220. [113] E. Korngold, Electrodialysis unit: optimization and calculation of energy

20
F.E. Ahmed, et al. Desalination 495 (2020) 114659

requirement, Desalination 40 (1–2) (1982) 171–179. [150] F. Banat, et al., Performance evaluation of the “large SMADES” autonomous de­
[114] F.B. Leitz, M.A. Accomazzo, W.A. McRae, High temperature electrodialysis, salination solar-driven membrane distillation plant in Aqaba, Jordan, Desalination
Desalination 14 (1) (1974) 33–41. 217 (1) (2007) 17–28.
[115] A.M. Benneker, et al., Effect of temperature gradients in (reverse) electrodialysis in [151] F. Macedonio, E. Curcio, E. Drioli, Integrated membrane systems for seawater
the Ohmic regime, J. Membr. Sci. 548 (2018) 421–428. desalination: energetic and exergetic analysis, economic evaluation, experimental
[116] K.M. Chehayeb, K.G. Nayar, J.H. Lienhard, On the merits of using multi-stage and study, Desalination 203 (1) (2007) 260–276.
counterflow electrodialysis for reduced energy consumption, Desalination 439 [152] A. Criscuoli, M.C. Carnevale, E. Drioli, Evaluation of energy requirements in
(2018) 1–16. membrane distillation, Chem. Eng. Process. Process Intensif. 47 (7) (2008)
[117] R.W. Baker, Membrane Technology and Applications, Wiley, 2004. 1098–1105.
[118] G.J. Doornbusch, et al., Experimental investigation of multistage electrodialysis [153] X. Wang, et al., Feasibility research of potable water production via solar-heated
for seawater desalination, Desalination 464 (2019) 105–114. hollow fiber membrane distillation system, Desalination 247 (1–3) (2009)
[119] V.V. Nikonenko, et al., Desalination at overlimiting currents: state-of-the-art and 403–411.
perspectives, Desalination 342 (2014) 85–106. [154] H.C. Duong, et al., Evaluating energy consumption of air gap membrane distilla­
[120] T. Xu, C. Huang, Electrodialysis-based separation technologies: a critical review, tion for seawater desalination at pilot scale level, Sep. Purif. Technol. 166 (2016)
AICHE J. 54 (12) (2008) 3147–3159. 55–62.
[121] E. Drioli, A. Ali, F. Macedonio, Membrane distillation: recent developments and [155] A. Criscuoli, M.C. Carnevale, E. Drioli, Modeling the performance of flat and ca­
perspectives, Desalination 356 (2015) 56–84. pillary membrane modules in vacuum membrane distillation, J. Membr. Sci. 447
[122] G. Amy, et al., Membrane-based seawater desalination: present and future pro­ (2013) 369–375.
spects, Desalination 401 (2017) 16–21. [156] G. Guan, et al., Evaluation of hollow fiber-based direct contact and vacuum
[123] S. Adham, et al., Application of membrane distillation for desalting brines from membrane distillation systems using aspen process simulation, J. Membr. Sci. 464
thermal desalination plants, Desalination 314 (2013) 101–108. (2014) 127–139.
[124] J. Minier-Matar, et al., Field evaluation of membrane distillation technologies for [157] N. Dow, et al., Pilot trial of membrane distillation driven by low grade waste heat:
desalination of highly saline brines, Desalination 351 (2014) 101–108. membrane fouling and energy assessment, Desalination 391 (2016) 30–42.
[125] U.K. Kesieme, et al., Economic analysis of desalination technologies in the context [158] M.I. Ali, et al., Effects of membrane properties on water production cost in small
of carbon pricing, and opportunities for membrane distillation, Desalination 323 scale membrane distillation systems, Desalination 306 (2012) 60–71.
(2013) 66–74. [159] J. Feher, 2.7 - Osmosis and osmotic pressure, in: J. Feher (Ed.), Quantitative
[126] F. Banat, R. Jumah, M. Garaibeh, Exploitation of solar energy collected by solar Human Physiology, Second edition, Academic Press, Boston, 2017, pp. 182–198.
stills for desalination by membrane distillation, Renew. Energy 25 (2) (2002) [160] B.H. Ozer, P.C. Brazy, 2 - Body fluid compartments and their regulation, in:
293–305. A.V. Moorthy (Ed.), Pathophysiology of Kidney Disease and Hypertension, W.B.
[127] K. Schneider, et al., Membranes and modules for transmembrane distillation, J. Saunders, 2009, pp. 17–27.
Membr. Sci. 39 (1) (1988) 25–42. [161] M. Eyvaz, et al., Forward osmosis membranes–a review: part I, Osmotically Driven
[128] E. Curcio, E. Drioli, Membrane distillation and related operations—a review, Sep. Membrane Processes-Approach, Development and Current Status, IntechOpen,
Purif. Rev. 34 (1) (2005) 35–86. 2018.
[129] G. Zaragoza, J.A. Andrés-Mañas, A. Ruiz-Aguirre, Commercial scale membrane [162] M. Amjad, et al., Novel draw solution for forward osmosis based solar desalina­
distillation for solar desalination, npj Clean Water 1 (1) (2018) 20. tion, Appl. Energy 230 (2018) 220–231.
[130] E. Drioli, A. Criscuoli, E. Curcio, Membrane Contactors: Fundamentals, [163] J.R. McCutcheon, R.L. McGinnis, M. Elimelech, Desalination by ammonia–carbon
Applications and Potentialities, Elsevier, 2011. dioxide forward osmosis: influence of draw and feed solution concentrations on
[131] F. Benyahia, Membrane-Distillation in Desalination, CRC Press, 2019. process performance, J. Membr. Sci. 278 (1) (2006) 114–123.
[132] E.K. Summers, H.A. Arafat, J.H. Lienhard, Energy efficiency comparison of single- [164] Q. Ge, M. Ling, T.-S. Chung, Draw solutions for forward osmosis processes: de­
stage membrane distillation (MD) desalination cycles in different configurations, velopments, challenges, and prospects for the future, J. Membr. Sci. 442 (2013)
Desalination 290 (2012) 54–66. 225–237.
[133] B.S. Lalia, et al., Fabrication and characterization of polyvinylidenefluoride-co- [165] N. Akther, et al., Recent advancements in forward osmosis desalination: a review,
hexafluoropropylene (PVDF-HFP) electrospun membranes for direct contact Chem. Eng. J. 281 (2015) 502–522.
membrane distillation, J. Membr. Sci. 428 (2013) 104–115. [166] D.J. Johnson, et al., Osmotic’s potential: an overview of draw solutes for forward
[134] F. Guo, et al., Desalination by membrane distillation using electrospun polyamide osmosis, Desalination 434 (2018) 100–120.
fiber membranes with surface fluorination by chemical vapor deposition, ACS [167] T.Y. Cath, A.E. Childress, M. Elimelech, Forward osmosis: principles, applications,
Appl. Mater. Interfaces 7 (15) (2015) 8225–8232. and recent developments, J. Membr. Sci. 281 (1) (2006) 70–87.
[135] Y. Liao, et al., Electrospun superhydrophobic membranes with unique structures [168] A.M. Awad, et al., The status of forward osmosis technology implementation,
for membrane distillation, ACS Appl. Mater. Interfaces 6 (18) (2014) Desalination 461 (2019) 10–21.
16035–16048. [169] N.M. Mazlan, D. Peshev, A.G. Livingston, Energy consumption for desalination —
[136] L.D. Tijing, et al., Recent progress of membrane distillation using electrospun a comparison of forward osmosis with reverse osmosis, and the potential for
nanofibrous membrane, J. Membr. Sci. 453 (2014) 435–462. perfect membranes, Desalination 377 (2016) 138–151.
[137] C. Su, et al., Novel three-dimensional superhydrophobic and strength-enhanced [170] R.K. McGovern, J.H. Lienhard V, On the potential of forward osmosis to en­
electrospun membranes for long-term membrane distillation, Sep. Purif. Technol. ergetically outperform reverse osmosis desalination, J. Membr. Sci. 469 (2014)
178 (2017) 279–287. 245–250.
[138] K.-J. Lu, J. Zuo, T.-S. Chung, Novel PVDF membranes comprising n-butylamine [171] A.S. Moon, M. Lee, Energy consumption in forward osmosis-desalination com­
functionalized graphene oxide for direct contact membrane distillation, J. Membr. pared to other desalination techniques, World Acad. Sci. Eng. Technol. 65 (2012)
Sci. 539 (2017) 34–42. 537–539.
[139] X. Li, et al., A novel profiled core–shell nanofibrous membrane for wastewater [172] P.M. Williams, et al., Technology for freeze concentration in the desalination in­
treatment by direct contact membrane distillation, J. Mater. Chem. A 4 (37) dustry, Desalination 356 (2015) 314–327.
(2016) 14453–14463. [173] H.M. Hendrickson, R.W. Moulton, Research and Development of Processes for
[140] H.C. Duong, et al., A novel electrospun, hydrophobic, and elastomeric styrene- Desalting Water by Freezing, US Department of the Interior, 1956.
butadiene-styrene membrane for membrane distillation applications, J. Membr. [174] H.F. Wiegandt, Desalting of Seawater by Freezing, (1968).
Sci. 549 (2018) 420–427. [175] R. Fujioka, et al., Application of progressive freeze-concentration for desalination,
[141] G. Zuo, R. Wang, Novel membrane surface modification to enhance anti-oil fouling Desalination 319 (2013) 33–37.
property for membrane distillation application, J. Membr. Sci. 447 (2013) 26–35. [176] M. Shafiur Rahman, M. Ahmed, X.D. Chen, Freezingmelting process and desali­
[142] Y.-X. Huang, et al., Novel Janus membrane for membrane distillation with si­ nation: review of present status and future prospects, Int. J. Nucl. Desalin. 2 (3)
multaneous fouling and wetting resistance, Environ. Sci. Technol. 51 (22) (2017) (2007) 253–264.
13304–13310. [177] F.G.F. Qin, et al., Experimental study of wash columns used for separating ice from
[143] M.E. Leitch, et al., Bacterial nanocellulose aerogel membranes: novel high-porosity ice-slurry, Desalination 218 (1) (2008) 223–228.
materials for membrane distillation, Environ. Sci. Technol. Lett. 3 (3) (2016) [178] M.S. Rahman, M. Ahmed, X.D. Chen, Freezing-melting process and desalination: I.
85–91. Review of the state-of-the-art, Sep. Purif. Rev. 35 (2) (2006) 59–96.
[144] M. Khayet, Solar desalination by membrane distillation: dispersion in energy [179] P. Englezos, The freeze concentration process and its applications, Dev. Chem.
consumption analysis and water production costs (a review), Desalination 308 Eng. Miner. Process. 2 (1) (1994) 3–15.
(2013) 89–101. [180] G.F.C. Rogers, Y.R. Mayhew, Thermodynamic and Transport Properties of Fluids,
[145] J. Swaminathan, et al., Energy efficiency of permeate gap and novel conductive Blackwell, 1995.
gap membrane distillation, J. Membr. Sci. 502 (2016) 171–178. [181] P. Wang, T.-S. Chung, A conceptual demonstration of freeze desalination–mem­
[146] W. Jantaporn, A. Ali, P. Aimar, Specific energy requirement of direct contact brane distillation (FD–MD) hybrid desalination process utilizing liquefied natural
membrane distillation, Chem. Eng. Res. Des. 128 (2017) 15–26. gas (LNG) cold energy, Water Res. 46 (13) (2012) 4037–4052.
[147] G. Zakrzewska-Trznadel, M. Harasimowicz, A.G. Chmielewski, Concentration of [182] W. Cao, C. Beggs, I.M. Mujtaba, Theoretical approach of freeze seawater desali­
radioactive components in liquid low-level radioactive waste by membrane dis­ nation on flake ice maker utilizing LNG cold energy, Desalination 355 (2015)
tillation, J. Membr. Sci. 163 (2) (1999) 257–264. 22–32.
[148] J. Koschikowski, M. Wieghaus, M. Rommel, Solar thermal-driven desalination [183] J. Chang, et al., Freeze desalination of seawater using LNG cold energy, Water Res.
plants based on membrane distillation, Desalination 156 (1) (2003) 295–304. 102 (2016) 282–293.
[149] S. Bouguecha, B. Hamrouni, M. Dhahbi, Small scale desalination pilots powered by [184] W. Lin, M. Huang, A. Gu, A seawater freeze desalination prototype system utilizing
renewable energy sources: case studies, Desalination 183 (1) (2005) 151–165. LNG cold energy, Int. J. Hydrog. Energy 42 (29) (2017) 18691–18698.

21
F.E. Ahmed, et al. Desalination 495 (2020) 114659

[185] C. Xie, et al., A direct contact type ice generator for seawater freezing desalination solutions, J. Membr. Sci. 375 (1) (2011) 172–181.
using LNG cold energy, Desalination 435 (2018) 293–300. [220] J.E. Kim, et al., Environmental and economic impacts of fertilizer drawn forward
[186] A. Messineo, D. Panno, Potential applications using LNG cold energy in Sicily, Int. osmosis and nanofiltration hybrid system, Desalination 416 (2017) 76–85.
J. Energy Res. 32 (11) (2008) 1058–1064. [221] J.E. Kim, S. Phuntsho, H.K. Shon, Pilot-scale nanofiltration system as post-treat­
[187] T. He, et al., LNG cold energy utilization: prospects and challenges, Energy 170 ment for fertilizer-drawn forward osmosis desalination for direct fertigation,
(2019) 557–568. Desalin. Water Treat. 51 (31−33) (2013) 6265–6273.
[188] C. Xie, et al., High efficient seawater freezing desalination technology by utilizing [222] S.F. Anis, R. Hashaikeh, N. Hilal, Reverse osmosis pretreatment technologies and
cold energy of LNG, IDA J. Desalin. Water Reuse 6 (1) (2014) 5–9. future trends: a comprehensive review, Desalination 452 (2019) 159–195.
[189] A.A. Madani, Zero-discharge direct-contact freezing/solar evaporator desalination [223] M. Giagnorio, et al., Hybrid forward osmosis–nanofiltration for wastewater reuse:
complex, Desalination 85 (2) (1992) 179–195. system design, Membranes 9 (5) (2019) 61.
[190] J.W. Post, et al., Salinity-gradient power: evaluation of pressure-retarded osmosis [224] B. Corzo, et al., Long-term evaluation of a forward osmosis-nanofiltration de­
and reverse electrodialysis, J. Membr. Sci. 288 (1) (2007) 218–230. monstration plant for wastewater reuse in agriculture, Chem. Eng. J. 338 (2018)
[191] B. Zhang, et al., Chapter 6 - Pressure retarded osmosis and reverse electrodialysis 383–391.
as power generation membrane systems, in: A. Basile, A. Cassano, A. Figoli (Eds.), [225] E.M.V. Hoek, et al., Modeling the effects of fouling on full-scale reverse osmosis
Current Trends and Future Developments on (Bio-) Membranes, Elsevier, 2019, processes, J. Membr. Sci. 314 (1) (2008) 33–49.
pp. 133–152. [226] M. Turek, et al., Prospects for high water recovery membrane desalination,
[192] Y. Mei, C.Y. Tang, Recent developments and future perspectives of reverse elec­ Desalination 401 (2017) 180–189.
trodialysis technology: a review, Desalination 425 (2018) 156–174. [227] R.L. Stover, Industrial and brackish water treatment with closed circuit reverse
[193] W. Guo, et al., Energy harvesting with single-ion-selective nanopores: a con­ osmosis, Desalin. Water Treat. 51 (4–6) (2013) 1124–1130.
centration-gradient-driven nanofluidic power source, Adv. Funct. Mater. 20 (8) [228] B. Kim, et al., Purification of high salinity brine by multi-stage ion concentration
(2010) 1339–1344. polarization desalination, Sci. Rep. 6 (1) (2016) 31850.
[194] J.G. Hong, et al., Potential ion exchange membranes and system performance in [229] S. Thampy, et al., Development of hybrid electrodialysis-reverse osmosis domestic
reverse electrodialysis for power generation: a review, J. Membr. Sci. 486 (2015) desalination unit for high recovery of product water, Desalination 282 (2011)
71–88. 104–108.
[195] F. Giacalone, et al., Exergy analysis of reverse electrodialysis, Energy Convers. [230] R.K. McGovern, S.M. Zubair, J.H. Lienhard V, The benefits of hybridising elec­
Manag. 164 (2018) 588–602. trodialysis with reverse osmosis, J. Membr. Sci. 469 (2014) 326–335.
[196] J. Moreno, et al., Upscaling reverse electrodialysis, Environ. Sci. Technol. 52 (18) [231] R.K. McGovern, S.M. Zubair, J.H. Lienhard V, Hybrid electrodialysis reverse os­
(2018) 10856–10863. mosis system design and its optimization for treatment of highly saline brines, IDA
[197] P. Długołȩcki, et al., Practical potential of reverse electrodialysis as process for J. Desalin. Water Reuse 6 (1) (2014) 15–23.
sustainable energy generation, Environ. Sci. Technol. 43 (17) (2009) 6888–6894. [232] C. Jiang, et al., Electrodialysis of concentrated brine from RO plant to produce
[198] S. Loeb, et al., The osmotic power plant, iece, 1976. coarse salt and freshwater, J. Membr. Sci. 450 (2014) 323–330.
[199] K.L. Lee, R.W. Baker, H.K. Lonsdale, Membranes for power generation by pressure- [233] Y. Zhang, et al., Electrodialysis on RO concentrate to improve water recovery in
retarded osmosis, J. Membr. Sci. 8 (2) (1981) 141–171. wastewater reclamation, J. Membr. Sci. 378 (1) (2011) 101–110.
[200] F. Helfer, C. Lemckert, Y.G. Anissimov, Osmotic power with pressure retarded [234] A.H. Galama, et al., Seawater predesalination with electrodialysis, Desalination
osmosis: theory, performance and trends – a review, J. Membr. Sci. 453 (2014) 342 (2014) 61–69.
337–358. [235] M. Turek, et al., Energy consumption and gypsum scaling assessment in a hybrid
[201] A.P. Straub, A. Deshmukh, M. Elimelech, Pressure-retarded osmosis for power Nanofiltration-reverse osmosis-electrodialysis system, Chem. Eng. Technol. 41 (2)
generation from salinity gradients: is it viable? Energy Environ. Sci. 9 (1) (2016) (2018) 392–400.
31–48. [236] J. Pellegrino, C. Gorman, L. Richards, A speculative hybrid reverse osmosis/
[202] A. Cipollina, G. Micale, Sustainable Energy from Salinity Gradients, Elsevier electrodialysis unit operation, Desalination 214 (1) (2007) 11–30.
Science, 2016. [237] P. Wang, T.-S. Chung, Recent advances in membrane distillation processes:
[203] W. Gai, D.L. Zhao, T.-S. Chung, Novel thin film composite hollow fiber membranes membrane development, configuration design and application exploring, J.
incorporated with carbon quantum dots for osmotic power generation, J. Membr. Membr. Sci. 474 (2015) 39–56.
Sci. 551 (2018) 94–102. [238] D. Eumine Suk, T. Matsuura, Membrane-based hybrid processes: a review, Sep.
[204] Statkraft discontinues investments in pressure retarded osmosis, Sci. Technol. 41 (4) (2006) 595–626.
ForwardOsmosisTech, 2014. [239] Y.-J. Choi, et al., Evaluation of economic feasibility of reverse osmosis and
[205] Z.L. Cheng, X. Li, T.-S. Chung, The forward osmosis-pressure retarded osmosis membrane distillation hybrid system for desalination, Desalin. Water Treat. 57
(FO-PRO) hybrid system: a new process to mitigate membrane fouling for sus­ (51) (2016) 24662–24673.
tainable osmotic power generation, J. Membr. Sci. 559 (2018) 63–74. [240] G. Naidu, et al., Membrane distillation for wastewater reverse osmosis concentrate
[206] J.M. Salamanca, O. Álvarez-Silva, F. Tadeo, Potential and analysis of an osmotic treatment with water reuse potential, J. Membr. Sci. 524 (2017) 565–575.
power plant in the Magdalena River using experimental field-data, Energy 180 [241] Cath, T.Y., A.E. Childress, and C.R. Martinetti, Water extraction system for pur­
(2019) 548–555. ifying liquid, e.g. seawater, comprises membrane-distillation desalination unit
[207] N. Bajraktari, C. Hélix-Nielsen, H.T. Madsen, Pressure retarded osmosis from hy­ having input and output and used to remove salt solutes from intermediate draw
persaline sources — a review, Desalination 413 (2017) 65–85. solution, and forward-osmosis system. UNIV NEVADA SYSTEM HIGHER
[208] M. Park, et al., Application of nanofiltration pretreatment to remove divalent ions EDUCATION (UNEV-C) CATH T Y (CATH-Individual) CHILDRESS A E (CHIL-
for economical seawater reverse osmosis desalination, Desalin. Water Treat. 57 Individual) MARTINETTI C R (MART-Individual).
(44) (2016) 20661–20670. [242] D. Zhao, et al., Thermoresponsive copolymer-based draw solution for seawater
[209] C. Tan, H. Ng, A novel hybrid forward osmosis-nanofiltration (FO-NF) process for desalination in a combined process of forward osmosis and membrane distillation,
seawater desalination: draw solution selection and system configuration, Desalin. Desalination 348 (2014) 26–32.
Water Treat. 13 (1–3) (2010) 356–361. [243] Q. Ge, et al., Polyelectrolyte-promoted forward osmosis–membrane distillation
[210] S. Zhao, L. Zou, D. Mulcahy, Brackish water desalination by a hybrid forward (FO–MD) hybrid process for dye wastewater treatment, Environ. Sci. Technol. 46
osmosis–nanofiltration system using divalent draw solute, Desalination 284 (11) (2012) 6236–6243.
(2012) 175–181. [244] P. Wang, et al., Evaluation of hydroacid complex in the forward osmosis–mem­
[211] S. Phuntsho, et al., Blended fertilizers as draw solutions for fertilizer-drawn for­ brane distillation (FO–MD) system for desalination, J. Membr. Sci. 494
ward osmosis desalination, Environ. Sci. Technol. 46 (8) (2012) 4567–4575. (2015) 1–7.
[212] Y. Kim, et al., Selection of suitable fertilizer draw solute for a novel fertilizer- [245] Q. Ge, G. Han, T.-S. Chung, Effective As(III) removal by a multi-charged hydroacid
drawn forward osmosis–anaerobic membrane bioreactor hybrid system, Bioresour. complex draw solute facilitated forward osmosis-membrane distillation (FO-MD)
Technol. 210 (2016) 26–34. processes, Environ. Sci. Technol. 50 (5) (2016) 2363–2370.
[213] L. Chekli, et al., Evaluation of fertilizer-drawn forward osmosis for sustainable [246] C.X. Guo, et al., Na+−functionalized carbon quantum dots: a new draw solute in
agriculture and water reuse in arid regions, J. Environ. Manag. 187 (2017) forward osmosis for seawater desalination, Chem. Commun. 50 (55) (2014)
137–145. 7318–7321.
[214] P. Nasr, H. Sewilam, Investigating the performance of ammonium sulphate draw [247] N.C. Nguyen, et al., Exploring high charge of phosphate as new draw solute in a
solution in fertilizer drawn forward osmosis process, Clean Techn. Environ. Policy forward osmosis–membrane distillation hybrid system for concentrating high-
18 (3) (2016) 717–727. nutrient sludge, Sci. Total Environ. 557-558 (2016) 44–50.
[215] S. Li, et al., Impact of reverse nutrient diffusion on membrane biofouling in fer­ [248] Q. Long, et al., Recent advance on draw solutes development in forward osmosis,
tilizer-drawn forward osmosis, J. Membr. Sci. 539 (2017) 108–115. Processes 6 (9) (2018) 165.
[216] P. Nasr, H. Sewilam, Investigating fertilizer drawn forward osmosis process for [249] L. Zohrabian, N.P. Hankins, R.W. Field, Hybrid forward osmosis-membrane dis­
groundwater desalination for irrigation in Egypt, Desalin. Water Treat. 57 (56) tillation system: demonstration of technical feasibility, J. Water Process Eng. 33
(2016) 26932–26942. (2020) 101042.
[217] S. Phuntsho, et al., Fertiliser drawn forward osmosis process: pilot-scale desali­ [250] Cath, T.Y., A.E. Childress, and C.R. Martinetti, Combined membrane-distillation-
nation of mine impaired water for fertigation, J. Membr. Sci. 508 (2016) 22–31. forward-osmosis systems and methods of use. 2011, Google Patents.
[218] W. Suwaileh, D. Johnson, N. Hilal, Brackish water desalination for agriculture: [251] Y. Kim, et al., Osmotically and thermally isolated forward osmosis–membrane
assessing the performance of inorganic fertilizer draw solutions, Desalination 456 distillation (FO–MD) integrated module, Environ. Sci. Technol. 53 (7) (2019)
(2019) 53–63. 3488–3498.
[219] S. Phuntsho, et al., A novel low energy fertilizer driven forward osmosis desali­ [252] Ghaffour, N., et al., Osmotically and thermally isolated forward osmosis-mem­
nation for direct fertigation: evaluating the performance of fertilizer draw brane distillation (FO-MD) integrated module for water treatment applications.

22
F.E. Ahmed, et al. Desalination 495 (2020) 114659

2018, Google Patents. desalination-vacuum membrane distillation hybrid systems powered by LNG re­
[253] J.-W. Koo, et al., Integration of forward osmosis with membrane distillation: effect gasification and solar energy, Desalination 449 (2019) 16–25.
of operating conditions, Desalin. Water Treat. 51 (25–27) (2013) 5355–5361. [273] K.J. Lu, et al., Design of zero liquid discharge desalination (ZLDD) systems con­
[254] M. Xie, et al., Toward resource recovery from wastewater: extraction of phos­ sisting of freeze desalination, membrane distillation, and crystallization powered
phorus from digested sludge using a hybrid forward osmosis–membrane distilla­ by green energies, Desalination 458 (2019) 66–75.
tion process, Environ. Sci. Technol. Lett. 1 (2) (2014) 191–195. [274] Y. Choi, et al., Membrane distillation crystallization for brine mining and zero
[255] H. Song, J. Liu, Forward osmosis membrane bioreactor using Bacillus and mem­ liquid discharge: opportunities, challenges, and recent progress, Environ. Sci.:
brane distillation hybrid system for treating dairy wastewater, Environ. Technol. Water Res. Technol. 5 (7) (2019) 1202–1221.
(2019) 1–12. [275] R.A. Tufa, et al., Progress and prospects in reverse electrodialysis for salinity
[256] S. Zhang, et al., Sustainable water recovery from oily wastewater via forward gradient energy conversion and storage, Appl. Energy 225 (2018) 290–331.
osmosis-membrane distillation (FO-MD), Water Res. 52 (2014) 112–121. [276] K. Kwon, et al., Brine recovery using reverse electrodialysis in membrane-based
[257] Q. Liu, et al., Integrated forward osmosis-membrane distillation process for human desalination processes, Desalination 362 (2015) 1–10.
urine treatment, Water Res. 91 (2016) 45–54. [277] Q. Wang, et al., Hybrid RED/ED system: simultaneous osmotic energy recovery
[258] G.S. Arcanjo, et al., Draw solution solute selection for a hybrid forward osmosis- and desalination of high-salinity wastewater, Desalination 405 (2017) 59–67.
membrane distillation module: effects on trace organic compound rejection, water [278] Ramato A. Tufa, et al., Membrane distillation and reverse electrodialysis for near-
flux and polarization, Chem. Eng. J. 400 (2020) 125857. zero liquid discharge and low energy seawater desalination, J. Membr. Sci. 496
[259] F. Volpin, et al., Optimisation of a forward osmosis and membrane distillation (2015) 325–333.
hybrid system for the treatment of source-separated urine, Sep. Purif. Technol. 212 [279] Y. Mei, C.Y. Tang, Co-locating reverse electrodialysis with reverse osmosis desa­
(2019) 368–375. lination: synergies and implications, J. Membr. Sci. 539 (2017) 305–312.
[260] X.-M. Li, et al., Water reclamation from shale gas drilling flow-back fluid using a [280] W. Li, et al., A novel hybrid process of reverse electrodialysis and reverse osmosis
novel forward osmosis–vacuum membrane distillation hybrid system, Water Sci. for low energy seawater desalination and brine management, Appl. Energy 104
Technol. 69 (5) (2014) 1036–1044. (2013) 592–602.
[261] W. Suwaileh, et al., An integrated fertilizer driven forward osmosis- renewables [281] H. Tian, et al., Unique applications and improvements of reverse electrodialysis: a
powered membrane distillation system for brackish water desalination: a com­ review and outlook, Appl. Energy 262 (2020) 114482.
bined experimental and theoretical approach, Desalination 471 (2019) 114126. [282] J.L. Prante, et al., RO-PRO desalination: an integrated low-energy approach to
[262] S. Zou, Z. He, Electrodialysis recovery of reverse-fluxed fertilizer draw solute seawater desalination, Appl. Energy 120 (2014) 104–114.
during forward osmosis water treatment, Chem. Eng. J. 330 (2017) 550–558. [283] J. Kim, et al., Reverse osmosis (RO) and pressure retarded osmosis (PRO) hybrid
[263] T.N. Bitaw, K. Park, D.R. Yang, Optimization on a new hybrid forward osmosis- processes: model-based scenario study, Desalination 322 (2013) 121–130.
electrodialysis-reverse osmosis seawater desalination process, Desalination 398 [284] S. S, S. S, Reverse osmosis–pressure retarded osmosis hybrid system: modelling,
(2016) 265–281. simulation and optimization, Desalination 389 (2016) 78–97.
[264] B. Gu, J.H. Kim, D.R. Yang, Theoretical analysis of a seawater desalination process [285] C.F. Wan, T.-S. Chung, Energy recovery by pressure retarded osmosis (PRO) in
integrating forward osmosis, crystallization, and reverse osmosis, J. Membr. Sci. SWRO–PRO integrated processes, Appl. Energy 162 (2016) 687–698.
444 (2013) 440–448. [286] M.H. Sharqawy, S.M. Zubair, J.H. Lienhard, Second law analysis of reverse os­
[265] L.F. Greenlee, et al., Reverse osmosis desalination: water sources, technology, and mosis desalination plants: an alternative design using pressure retarded osmosis,
today’s challenges, Water Res. 43 (9) (2009) 2317–2348. Energy 36 (11) (2011) 6617–6626.
[266] S. Miller, H. Shemer, R. Semiat, Energy and environmental issues in desalination, [287] D.I. Kim, et al., Pressure retarded osmosis (PRO) for integrating seawater desali­
Desalination 366 (2015) 2–8. nation and wastewater reclamation: energy consumption and fouling, J. Membr.
[267] D.L. Shaffer, et al., Seawater desalination for agriculture by integrated forward Sci. 483 (2015) 34–41.
and reverse osmosis: improved product water quality for potentially less energy, J. [288] S. Lin, et al., Hybrid pressure retarded osmosis–membrane distillation system for
Membr. Sci. 415 (2012) 1–8. power generation from low-grade heat: thermodynamic analysis and energy effi­
[268] L. Martinez-Diez, F. Florido-Diaz, Desalination of brines by membrane distillation, ciency, Environ. Sci. Technol. 48 (9) (2014) 5306–5313.
Desalination 137 (1–3) (2001) 267–273. [289] J. Kim, et al., Performance analysis of reverse osmosis, membrane distillation, and
[269] D. González, J. Amigo, F. Suárez, Membrane distillation: perspectives for sus­ pressure-retarded osmosis hybrid processes, Desalination 380 (2016) 85–92.
tainable and improved desalination, Renew. Sust. Energ. Rev. 80 (2017) 238–259. [290] N.Y. Yip, M. Elimelech, Comparison of energy efficiency and power density in
[270] K. Nakoa, et al., Sustainable zero liquid discharge desalination (SZLDD), Sol. pressure retarded osmosis and reverse electrodialysis, Environ. Sci. Technol. 48
Energy 135 (2016) 337–347. (18) (2014) 11002–11012.
[271] J.-H. Tsai, et al., Membrane-based zero liquid discharge: myth or reality? J. [291] B. Ricci, et al., Critical performance assessment of a submerged hybrid forward
Taiwan Inst. Chem. Eng. 80 (2017) 192–202. osmosis-membrane distillation system, Desalination 468 (2019) 114082.
[272] J. Chang, et al., Membrane development and energy analysis of freeze

23

You might also like