You are on page 1of 14

w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 1 9 0 7 e1 9 2 0

Available at www.sciencedirect.com

journal homepage: www.elsevier.com/locate/watres

Review

Energy minimization strategies and renewable energy


utilization for desalination: A review
Arun Subramani a,*, Mohammad Badruzzaman a, Joan Oppenheimer a,
Joseph G. Jacangelo a,b
a
b

MWH Americas Inc., 618 Michillinda Avenue, Arcadia, CA 91007, USA


The Johns Hopkins University, Baltimore, MD 21205, USA

article info

abstract

Article history:

Energy is a significant cost in the economics of desalinating waters, but water scarcity is

Received 25 October 2010

driving the rapid expansion in global installed capacity of desalination facilities. Conven-

Received in revised form

tional fossil fuels have been utilized as their main energy source, but recent concerns over

27 December 2010

greenhouse gas (GHG) emissions have promoted global development and implementation of

Accepted 31 December 2010

energy minimization strategies and cleaner energy supplies. In this paper, a comprehensive

Available online 9 January 2011

review of energy minimization strategies for membrane-based desalination processes and


utilization of lower GHG emission renewable energy resources is presented. The review

Keywords:

covers the utilization of energy efficient design, high efficiency pumping, energy recovery

Reverse osmosis

devices, advanced membrane materials (nanocomposite, nanotube, and biomimetic),

Nanotechnology

innovative technologies (forward osmosis, ion concentration polarization, and capacitive

Renewable energy

deionization), and renewable energy resources (solar, wind, and geothermal). Utilization of

Energy recovery

energy efficient design combined with high efficiency pumping and energy recovery devices

Water sources

have proven effective in full-scale applications. Integration of advanced membrane


materials and innovative technologies for desalination show promise but lack long-term
operational data. Implementation of renewable energy resources depends upon geographyspecific abundance, a feasible means of handling renewable energy power intermittency,
and solving technological and economic scale-up and permitting issues.
2011 Elsevier Ltd. All rights reserved.

Contents
1.
2.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Minimization of energy usage for desalination processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1. Enhanced system design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2. High efficiency pumping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3. Energy recovery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.4. Advanced membrane materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.4.1. Nanocomposite membranes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

* Corresponding author. Tel.: 1 626 568 6002; fax: 1 626 568 6015.
E-mail address: Arun.Subramani@us.mwhglobal.com (A. Subramani).
0043-1354/$ e see front matter 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.watres.2010.12.032

1908
1908
1908
1910
1910
1910
1911

1908

3.

4.
5.

1.

w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 1 9 0 7 e1 9 2 0

2.4.2. Nanotube membranes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .


2.4.3. Biomimetic membranes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.5. Innovative technologies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.5.1. Forward osmosis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.5.2. Ion concentration polarization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.5.3. Capacitive deionization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Renewable energy utilization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1. Solar energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1.1. Solar thermal processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1.2. Solar electromechanical process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2. Wind energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3. Geothermal energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.4. Hybrid systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.5. Design and implementation of renewable energy systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Research needs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Introduction

In response to increasing water demand, many municipalities


and water suppliers are considering more energy intensive
seawater desalination to supplement inadequate freshwater
sources (GWI, 2010). Desalination has been successfully
implemented to provide additional water to communities
experiencing shortages by applying processes developed over
the last 40 years (Gleick, 2006). It is estimated that the capital
expenditures for new desalination plants will exceed $17 billion
by the year 2016 out of which more than $13 billion is expected
to be targeted for seawater reverse osmosis (RO) (GWI, 2010).
Desalination processes are broadly categorized as thermal
or membrane-based technologies based on the separation
process adopted (Greenlee et al., 2009). Although thermal
desalination has remained the primary technology of choice
in the Middle East, membrane processes have rapidly developed since the 1960s (Loeb and Sourirajan, 1963) and
currently surpass thermal processes in new plant installations (Greenlee et al., 2009). Membrane-based desalination
technologies are favored over thermal-based desalination in
regions where the cost of energy for steam production is high
(NREL, 2006). The energy requirements for seawater desalination using thermal-based technologies are on the order of
7e14 kWh/m3 when compared to 2e6 kWh/m3 for membranebased technologies (Veerapaneni et al., 2007; Semiat, 2008;
Anderson et al., 2010). The energy requirements are lower
for RO, but energy consumption still remains the major
operational cost component due to the high pressure pumps
required to feed water to the RO process. These pumps are
responsible for more than 40% of the total energy costs
(Service, 2006; Souari and Hassairi, 2007). Reducing energy
consumption is, therefore, critical for lowering the cost of
desalination and addressing environmental concerns about
GHG emissions from the continued use of conventional fossil
fuels as the primary energy source for seawater desalination
plants. A large number of energy minimization approaches

1911
1912
1912
1912
1913
1913
1913
1913
1913
1915
1915
1916
1916
1916
1917
1917
1917
1917

and renewable energy alternatives are rapidly being developed, investigated and implemented around the globe
(Charcosset, 2009). Thus, providing an updated, comprehensive review of sustainable design and operational strategies to
reduce energy usage and GHG emissions is warranted. This
paper critically reviews in a holistic manner the latest developments and technologies for reducing energy consumption
by reverse osmosis desalination processes and addresses
strategies for integrating renewable energy as a source of
alternative clean energy supply. The paper is organized by
a discussion about energy minimization strategies for desalination followed by a discussion on the utilization of renewable energy resources to reduce GHG emissions.

2.
Minimization of energy usage for
desalination processes
Factors influential in minimizing energy usage in desalination
processes using RO membranes can be classified according to
enhanced system design, high efficiency pumping, energy
recovery, advanced membrane materials, and innovative
technologies. Each of these factors is described in more detail
below.

2.1.

Enhanced system design

The design and configuration of membrane units have


a significant effect on the performance and economics of an RO
plant (Wilf and Bartels, 2005). In the past, membrane units for
seawater were usually configured as two stages with six
elements per pressure vessel. The two-stage system resulted in
a high feed and concentrate flow, which reduced concentration
polarization at the expense of a greater feed pressure needed to
compensate for the increased pressure drop across the RO train.
Design efforts to reduce power consumption resulted in the use
of single-stage configurations for high salinity feed water

w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 1 9 0 7 e1 9 2 0

applications, and in some cases, the use of seven (or) eight


elements per pressure vessel (Wilf and Bartels, 2005; Petry et al.,
2007). The pressure drop reduction in using a single-stage rather
than a two-stage system was reported to result in a 2.5% lower
power requirement (Wilf and Bartels, 2005).
More recently, further reduction in RO desalination cost
has been shown to occur from optimal process configuration
and control schemes. Theoretical cost minimization framework have been developed and experimentally implemented
using a controller to quantify the effect of energy cost with
respect to membrane cost, brine management cost, energy
recovery, and feed salinity fluctuation (Zhu et al., 2009b, 2010).
A control system utilizing real-time sensor data and userdefined permeate flow requirements has been implemented
to compute in real-time the energy-optimal set-points for
controlling concentrate valve position and feed flow rate
(Bartman et al., 2009, 2010). Implementation of the control
system demonstrated the ability to achieve energy-optimal
operation of the RO system close to the theoretically predicted
energy consumption curves.
When stringent water quality requirements mandate the
use of multi-pass RO (Sauvet-Goichon, 2007), the overall power
consumption of the RO system can be lowered if a portion of
the first pass permeate is pumped to the second pass (Zhu
et al., 2009). Since the permeate produced from the front-end
elements is lower in salinity than the permeate produced at the
back-end elements, lower feed pressure is required for the
second pass when the front-end permeate is utilized as feed to
the second pass. In a multi-pass system, the lowest energy
consumption is obtained when membranes with the highest
salt rejection is used in the first-pass (Zhu et al., 2009a). In
another study, various mixing operations between feed,
concentrate, and permeate streams were evaluated to assess
their potential on energy usage (Zhu et al., 2010a). It was
determined that various mixing approaches may provide
certain operational or system design advantages but they do
not provide an advantage from an energy usage perspective.
A novel design modification to reduce pressure drop across
membrane elements is the use of a pressure vessel with
a center port design (van Paassen et al., 2005). In this innovative configuration, feed water enters the pressure vessel
through two feed ports on each end of the pressure vessel in
the first stage. The concentrate is collected through a middle
port and flows to a similar port on the pressure vessels in the
second stage. Thus, the flow path is reduced by half and
although the membrane unit has eight elements per pressure
vessel, the flow path length is reduced to four elements per
stage, creating a lower pressure drop that lowers the feed
pressure. A 15% reduction in the feed pressure has been
reported using the center port design when compared to
a conventional side port design (Wilf and Hudkins, 2010). The
disadvantage of the center port design is the potential for
scaling due to excessive concentration polarization. Thus,
pilot testing and long-term operational data are recommended before considering implementation of the center port
design in order to determine the influence of water quality
variations on feed water recovery.
Reduction in energy consumption for RO systems treating
high salinity feed water has also been achieved by using a twostage hybrid system with concentrate staging (Veerapaneni

1909

et al., 2005). The first stage consists of high rejection brackish


water membrane elements (or) high permeability seawater
membrane elements. The second stage consists of standard
seawater elements. Using a two-stage system with brackish
(or) low-pressure seawater membranes in the first stage lowers
feed pressure requirements due to lower membrane resistance
(Veerapaneni et al., 2007) As most of the permeate is produced
in the first stage with the high permeability membranes, the
pressure of only a small fraction of the remaining flow is
boosted, resulting in significant energy savings. A two-pass
nanofiltration (NF) membrane system also substantially
reduces the energy consumption (Long, 2008). The power
consumption of a two-pass NF process was estimated to be
2.06 kWh/m3 compared to 2.32 kWh/m3 for a two-stage hybrid
brackish and seawater membrane system (Long, 2008). A 5%
and 12% reduction in energy consumption was obtained when
using a hybrid brackish/seawater or two-pass NF element
system, respectively (Long, 2008).
Energy consumption is also reduced by minimizing the
pressure drop across membrane elements (Macedonio and
Drioli, 2010). An approach by which to reduce the axial pressure drop in membrane elements involves the use of a novel
feed spacer design that reduces the hydraulic pressure drop
in the RO elements (Subramani et al., 2006; Guillen and Hoek,
2009). The feed spacer pattern used in most spiral wound
membrane elements causes a variation in the flow path of the
feed water resulting in a higher axial pressure drop than flow in
an open channel (Guillen and Hoek, 2009). Although feed spacer
geometry was found to have a marginal impact on mass transfer, thinner spacer filaments spread apart substantially reduced
hydraulic pressure losses. In addition, certain non-circular
spacer filament shapes produced lower hydraulic losses when
compared to conventional circular spacer filament shapes
(Guillen and Hoek, 2009). Although various feed spacer geometries have been shown to reduce hydraulic pressure loss in RO
elements, actual data from pilot-scale and full-scale operation
are still minimal since spiral wound elements with novel feed
spacer configurations are not readily available. Commercialization of feed spacers that reduce the axial pressure drop across
membrane elements could potentially reduce the feed pressure
requirements during RO seawater desalination.
A plant design approach for improving the economics of
desalination and at the same time reduce the impact on
environment due to brine discharge is the co-location of
membrane desalination plants with existing coastal power
generation stations (Voutchkov, 2004). In this approach,
overall desalination power demand and associated costs of
water production are reduced as a result of the use of warmer
source water. The cooling water discharged from the
condensers in a power plant is 5e15  C warmer that the source
ocean water. When this water is used by the RO plant, 5e8%
lower feed pressure is required to desalinate the water when
compared to desalination of colder source ocean water. This
approach also has the advantage of sharing a common intake
facility. In the Middle East, RO and thermal-based technologies are combined to provide a hybrid design (Cardona and
Piacentino, 2004). Such hybrid designs not only result in
capital savings by sharing a common intake and outfall
facility but also have a 40e50% increase in water production
related to pre-heating of feed water to the RO plant.

1910

2.2.

w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 1 9 0 7 e1 9 2 0

High efficiency pumping

With respect to pumping, energy is predominantly consumed


from operation of primary feed pumps, second pass feed
pumps (as required), pretreatment pumps, product water
transfer pumps, chemical feed pumps, and water distribution
pumps. The distribution of power usage in a two-stage
seawater RO system is shown in Fig. 1. More than 80% of the
power is required for the operation of the primary feed pumps
(Wilf and Bartels, 2005). Although the flow and head of
a pumping system are determined by the design specifications
of the RO system, the selection and operation of pumps and
other elements of a pumping system play an important role in
reducing overall energy usage in the plant.
To achieve the highest possible pumping efficiency, several
procedures are performed including: (1) verifying energy efficient operation of the pumping system, (2) utilizing a premium
efficiency motor, and (3) utilizing a variable frequency drive
(Manth et al., 2003). To achieve an energy efficient operation,
a pumps speed must fall within a specified range for optimal
efficiency or the best efficiency point (Veerapaneni et al., 2007).
The use of high speed and high flow pumps at lower total
dynamic head provides the optimal speed needed for highest
efficiency. To accommodate the variability of feed pressure
with time (due to salinity and temperature fluctuations)
without the necessity to throttle high pressure pumps or energy
recovery devices, a variable frequency drive is often incorporated into the electric motor unit that drives the high pressure
pump (Torre, 2008). All of the above mentioned pumping
methods have been demonstrated to significantly improve
efficiency and reduce energy requirements at full scale.

2.3.

Energy recovery

Energy consumption for RO desalination processes is reduced


by using energy recovery devices (ERD) that have been shown
to recover energy from the RO concentrate (Andrews and
Laker, 2001; Wang et al., 2004). Before the concentrate stream
is sent for disposal, pressure from the stream is recovered by
passing it through an ERD. The fraction of power recovered
depends on the type and efficiency of the equipment used. Two
broad classes exist for ERDs (Wang et al., 2004). Class I devices
use hydraulic power to cause a positive displacement within
the recovery device, and the hydraulic energy is directly
transferred in one step (Greenlee et al., 2009). Class II devices
use the hydraulic energy of the RO concentrate in a two-step

Fig. 1 e Distribution of power usage in a two-stage


seawater RO system (Wilf and Bartels, 2005).

process that first converts the energy to centrifugal mechanical energy and then back to hydraulic energy. Most of the
seawater desalination plants in operation today use a Class I
type of ERD (Greenlee et al., 2009). When an ERD is used,
a fraction of the feed must bypass the primary high-pressure
pump and a booster pump is used to account for pressure
losses in the RO membrane modules, piping, and ERD
(Greenlee et al., 2009). The pressure or work exchanger (PWE) is
a Class I type of ERD. The pelton wheel, reverse running turbine
pump, and turbo charger are examples of a Class II type of ERD.
Efficiency greater than 95% can be achieved using a Class
I type of ERD (Greenlee et al., 2009). The PWE transfers the
hydraulic energy of the pressurized RO concentrate stream to
the RO feed water stream (Avlonitis et al., 2003; Stover, 2007).
PWE systems can be categorized as two types: those that
provide a physical barrier (piston) between the RO concentrate
stream and feed side of the system, such as a Dual Work
Exchanger Energy Recovery (DWEER), and those without
a physical barrier such as a Pressure Exchanger (PX) (Cameron
and Clemente, 2008; Mirza, 2008). In the case of a DWEER, the
system is based on moving pistons in cylinders which is well
suited for a wide range of water viscosities and densities, but
results in a large foot print (Mirza, 2008). A PX device has
higher efficiency since no transformational losses occur in the
device, but individual PXs have limited flow rates and
a higher capacity must be achieved by arranging several
devices in series. PX devices are also associated with very high
noise levels requiring a sound abatement enclosure (Mirza,
2008). Another disadvantage of a PX device is the degree of
mixing that occurs between the feed water and concentrate
stream. A feed salinity increase of 1.5%e3.0% caused by such
mixing will increase the required feed pressure for the RO
system (Wang et al., 2004, 2005).
Class 2 Centrifugal ERDs (such as the pelton wheel and turbo
charger) are limited in capacity and are usually optimized for
narrow flow and pressure operating conditions (Stover, 2004,
2007). The turbo charger is typically used in smaller capacity
RO installations (Oklejas et al., 2005). The reverse running
turbine pump is not suitable for a low flow range due to poor
efficiency (Mirza, 2008). The efficiency of commercial pelton
wheels can reach 90% (Stover, 2007). The overall efficiency of
the mechanically coupled reverse running turbine pump is in
the 75%e85% range. For the submersible generator type, the
overall efficiency is in the 62%e75% range (Mirza, 2008). The
efficiency of the turbo charger ranges from 55% to 60%.

2.4.

Advanced membrane materials

Significant improvements in the salt rejection capacity and


permeability of RO membranes for treating high salinity feed
waters have been achieved in recent years. In 1980, seawater
RO systems consumed more than 26 kWh/m3. Today, seawater
RO systems consume on average only 3.4 kWh/m3 (Chang et al.,
2008). The minimum theoretical energy use (50% recovery) is
about 1.08 kWh/m3 for seawater desalination (Voutchkov,
2010). Thus, there are further avenues for improving the
permeability of RO membranes using novel membrane materials such that the energy consumption is minimized. But, the
new generation membranes must provide at least double the
permeability of current generation RO membranes. This is

1911

w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 1 9 0 7 e1 9 2 0

based on a recent approach to determine the minimization of


energy costs by improving membrane permeability (Zhu et al.,
2009c). A dimensionless factor was used to reflect the impact of
feed water osmotic pressure, salt rejection requirement,
membrane permeability, and purchase price of electrical
energy and membrane module. It was estimated that unless
the permeability of the RO membrane is doubled and the
capital cost of pressure vessels directly impacted by a lower
membrane area requirement, further improvements in
seawater RO membrane permeability is less likely to significantly reduce the cost of desalination. New generation RO
membrane which show promise in providing more than
double the permeability of currently available RO membranes
are discussed below. New generation RO membranes offer
reduced feed pressure requirements while maintaining rejection. Todays high productivity membrane elements are
designed with two features that include more fresh water per
membrane element and higher surface area and denser
membrane packing (Voutchkov, 2007). New generation RO
membranes can be broadly classified as nanocomposite,
nanotube, and biomimetic membranes. A comparison of these
advanced membranes is described in Table 1.

2.4.1.

Nanocomposite membranes

Thin film nanocomposite RO membranes are made by


combining zeolite nanoparticles dispersed within a traditional
polymide thin film (Jeong et al., 2007; Hoek and Ghosh, 2009).
The zeolite nanoparticles are dispersed in one or more of the
monomer solutions used to create the membrane by an interfacial polymerization process. Incorporation of zeolite nanoparticles in the polymer matrix of seawater RO membranes has
enhanced flux to more than double that of a commercial product with 99.7% salt rejection. Utilization of nanocompositebased RO membranes can result in 20% lower energy
consumption (NanoH2O, 2010). Although RO membranes using
zeolite nanoparticles have been reported to show substantial reduction in the feed pressure requirement, long-term

operational data are still unavailable. Chemical stability of the


incorporated zeolite nanoparticles within the polyamide
matrix also needs to be demonstrated. If the nanocomposite
membranes are not chemically compatible in a wide pH range,
their applicability would be limited. Also, rejection of the
nanocomposite membrane has been reported only for sodium
chloride, magnesium sulfate, and polyethylene glycol (Jeong
et al., 2007). Rejection of specific constituents, such as boron,
in seawater by RO membranes has become a concern recently
due to stringent discharge limits (Greenlee et al., 2009). The use
of nanocomposite membranes for seawater desalination
would require that more rejection data for specific constituents
in seawater be available.

2.4.2.

Nanotube membranes

The use of carbon nanotubes have also been shown to consume


lower energy than conventional seawater RO desalination
systems (Truskett, 2003; Holt and Park, 2006; Sholl and Johnson,
2006; Corry, 2008; Jia et al., 2010). The transport of water and ions
occurs through membranes formed using carbon nanotubes
. Membranes incorporating
ranging in diameter from 6 to 11 A
carbon nanotubes are promising candidates for water desalination using RO since the size and uniformity of the tubes
should achieve the desired salt rejection. A ten-fold permeability increase is expected using a carbon nanotube RO
membrane which should result in a 30e50% savings in energy
usage. Simulations have shown that boron nitride nanotubes
have superior water flow properties to carbon nanotubes and
they can achieve 100% salt rejection (Hilder et al., 2009). The use
can also be used to functionalize
of a nanotube radius of 4.14 A
the membrane to become cation-selective. When a nanotube
is used, the membrane can be functionalized to
radius of 5.52 A
become anion-selective (Hilder et al., 2009). Similar to the
nanocomposite membrane, long-term operational data, scaleup information, chemical compatibility specifications and
tensile strength of the nanotube-based RO membrane is lacking. Information available on the rejection property and water

Table 1 e Comparison of advanced material based reverse osmosis membranes.


Membrane type

Principle

Energy consumption

Advantages

Drawbacks

Nanocomposite

Zeolite nanoparticles
incorporated in
polyamide matrix
creating enhanced
transport of water
molecules.

20% lower energy


consumption than
conventional seawater
RO (NanoH2O, 2010).

More than double


the flux of currently
available seawater
RO membranes
(NanoH2O, 2010).

Nanotube

Transport of water
molecules through
structured carbon
and boron nitride
nanotubes.
Aquaporins used to
regulate transport of
water molecules.

30e50% lower energy


consumption than
conventional seawater
RO (Hilder et al., 2009).

Ten e fold higher


flux than currently
available seawater
RO membranes
(Hilder et al., 2009).
Hundred times
permeable than
currently available
seawater RO
membranes
(AquaZ, 2010).

Chemical compatibility
and structural
stability is not known.
Rejection of specific
contaminants is not known.
Long-term operational
data not available.
Only modeling results available.
Rejection of specific contaminants
is not known.

Biomimetic

Energy consumption
is not known.

Inability to withstand high


operating pressures.
Rejection of specific contaminants
is not known.
Long-term operational
data not available.

1912

w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 1 9 0 7 e1 9 2 0

flux of the nanotube membrane is solely based on modeling


results and a significant amount of research still needs to be
conducted to determine if these nanotubes can be polymerized
and casted to achieve the results demonstrated through
modeling.

2.4.3.

Biomimetic membranes

New developments have also occurred with the use of biomimetic membranes for desalination (Bowen, 2006). Biomimetic
membranes are designed to mimic the highly selective transport of water across cell membranes. Natural proteins known as
aquaporins are used to regulate the flow of water providing
increased permeability and high solute rejection. Aquaporins
act as water channels which selectively allow water molecules
to pass through while the transport of ions is restricted by an
electrostatic tuning mechanism of the channel interior. This
results in only water molecules being transported through the
aquaporin channels and charged ions being rejected (Sui et al.,
2001; Gong et al., 2007). Aquaporin membranes are considered
to be a hundred times more permeable than commercial RO
membranes with anticipated specific power consumption
savings of 70% of specific power consumption (AquaZ, 2010).
Highly permeable and selective membranes based on the
incorporation of the functional water channel protein Aquaporin Z into a novel triblock copolymer has been shown to have
significantly higher water transport than existing RO
membranes (Kumar et al., 2007). A particular difficulty with
biomimetic membranes that needs to be overcome is their
inability to withstand high operating pressures.

2.5.

Innovative technologies

Innovative technologies utilizing the principles of separation


technology with membranes, osmosis and electric fields have
been introduced in recent years. These technologies have
the potential to offer a substantial reduction in energy consumption for desalination. A comparison of these innovative

technologies is described in Table 2. The technologies are discussed below.

2.5.1.

Forward osmosis

In the forward osmosis process, instead of using hydraulic


pressure similar to a conventional RO desalination process,
a concentrated draw solution is used to generate high osmotic pressure, which pulls water across a semipermeable
membrane from the feed solution (McCutcheon et al., 2005;
Chou et al., 2010). The draw solutes are then separated from
the diluted draw solution to recycle the solutes and to produce
clean product water. A mixture of ammonia and carbon
dioxide gas has been used as the predominant draw solution
(McCutcheon et al., 2006). Other draw solutions utilized are salt
solutions and magnetic nanoparticles (Yang et al., 2009).
Forward osmosis is a combination of membrane and thermal
processes. Specific energy consumption of less than 0.25 kWh/
m3 has been reported for the membrane portion of the process
(Cath et al., 2006; McGinnis and Elimelech, 2007). In a recent
study, a combination of forward osmosis and reverse osmosis
was found to produce a higher flux than the forward osmosis
process alone, thus reducing the specific energy consumption
(Choi et al., 2009). But, the results were valid only for certain
operating conditions and effect of membrane material on
energy efficiency of the process is unknown. The regeneration
of the draw solution requires significant amount of energy and
unless waste heat is available for regeneration of the draw
solution, the forward osmosis process is not considered more
efficient than an RO process (McGinnis et al., 2007; Semiat,
2008). The process also has the advantage of a lower fouling
propensity than RO as a result of the absence of hydraulic
pressure and application of novel thin film composite
membranes (Mi and Elimelech, 2010). A particular drawback of
the forward osmosis process is the utilization of an appropriate
membrane to reduce internal concentration polarization and
increase efficiency (McGinnis and Elimelech, 2007). Although
the forward osmosis process shows promise in terms of
better performance with respect to fouling and scaling on the

Table 2 e Comparison of innovative desalination technologies with reverse osmosis.


Technology

Principle

Energy
consumption

Forward osmosis

Utilizes natural osmosis to


dilute seawater feed
stream using a draw
solution with higher
osmotic pressure than
the seawater feed.

0.25e0.84 kWh/m3
(Cath et al., 2006; McGinnis
and Elimelech, 2007)

Ion concentration
polarization

Nanofluidics in
combination with
ion concentration
polarization utilized
to desalinate seawater.
Ions electrosorbed
by polarization of
electrode (carbon aerogels)
by a direct current
power source.

3.5 kWh/m3 (Kim et


al., 2010).

Capacitive
deionization

1.37e1.67 kWh/m3
(brackish water)
(Welgemoed, 2005).
Energy consumption
of seawater is not known.

Advantages
Lower energy consumption
than RO (McGinnis and
Elimelech, 2007). Lower
fouling potential than
RO due to absence of
transmembrane pressure
(Mi and Elimelech, 2010).
Lower energy consumption
than RO. Absence of
membranes and applied
pressure (Kim et al., 2010).
Lower energy consumption
than RO for brackish
water treatment (Oren, 2010).
Absence of membrane
and applied pressure.

Drawbacks
More applicable than
RO only when waste
heat source is
available (McGinnis
and Elimelech, 2007).
Full-scale operation
data is not available.
Process suited for
small and medium e
scale systems. Fullscale operational
data is not available.
Low feed water
recovery (Oren, 2010).
Full-scale operational
data not available.

w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 1 9 0 7 e1 9 2 0

membrane surface, the rejection capability of the membranes


for specific contaminants such as boron, which is a design
limiting factor for seawater desalination, is unknown. Studies
have been conducted to improve the membrane property used
for forward osmosis (Yang et al., 2009) but long-term operational data for the forward osmosis process is unavailable and
only bench-scale results have been published.

achievable is very low (Oren, 2008). Capacitive deionization has


primarily been used for brackish water desalination but with
improvements in electrode material and better process control
strategies, the technology holds promise for seawater desalination (Oren, 2008; Anderson et al., 2010).

3.
2.5.2.

Renewable energy utilization

Ion concentration polarization

Ion concentration polarization has been utilized to desalinate


seawater using an energy efficient process (Kim et al., 2010). In
this process, micro and nanofluidics in combination with ion
concentration polarization are used to desalinate seawater.
Ion concentration polarization is a fundamental transport
mechanism that occurs when an ionic current is passed
through an ion-selective membrane. But, in the newly developed process, no membranes are utilized. An electrical
potential is used to create a repulsion zone that acts as
a membrane in separating charged ions, bacteria, viruses, and
microbes from seawater flowing through a 500 mm  100 mm
microchannel. Water flows through the microchannel
tangential to a nanochannel where the voltage is applied. The
resulting force creates a repulsion zone and the stream splits
into two smaller channels at a nanojunction. The two streams
created are the treated water and concentrate. More than 99%
salt rejection and 50% recovery has been reported using this
process (Kim et al., 2010). The ion concentration polarization
process has been reported to consume approximately 3.5 kWh/
m3, which is comparable to seawater desalination using RO
membranes (Kim et al., 2010). The advantage of ion concentration is that the process is fouling free since membranes are
not used. The process is best suited for small to medium scale
systems with the possibility of battery-powered operation. The
process is not efficient in the removal of neutrally charged
organics and needs to be combined with other processes to
achieve treatment goals. Results from the ion concentration
polarization are not available for large-scale systems and only
modeling and bench-scale results have been reported.

2.5.3.

1913

Capacitive deionization

Capacitive deionization technology is not a recent development, but several challenges with the identification of an
optimum material for the manufacturing of the associated
electrodes have delayed commercialization (Farmar et al., 1997;
Dermentzis and Ouzounis, 2008; Lee et al., 2009). This technology was developed as a non-polluting, energy efficient and
cost effective alternative to desalination technologies such as
RO (Welgemoed, 2005). In this technology, a saline solution
flows through an unrestricted capacitor type module consisting
of numerous pairs of high-surface area electrodes. The electrode material, typically a carbon aerogel, contains a high
specific surface area (400e1100 m2 per g) and a very low electrical resistivity. Anions and cations in solution are electrosorbed by the electric field upon polarization of each electrode
pair by direct current power sources. For desalination of
brackish water, energy consumption of 1.37e1.67 kWh/m3 has
been reported using this technology (Welgemoed, 2005). Energy
consumption for high salinity waters (such as seawater) is not
readily available in the literature. The main drawback with
capacitive deionization is that the feed water recovery

Renewable energy resources are the best energy supply option


for stand-alone desalination systems in remote regions where
energy supply from an electricity grid is not readily accessible
(Schafer et al., 2007; Gude et al., 2010). In urban regions,
renewable energy may provide treatment cost reductions due
to the implementation of a diversified portfolio of energy
sources and reduces GHG emissions. For an RO system used to
desalinate seawater with traditional fossil fuel based energy
source, CO2 emissions of 1.78 kg/m3 of desalted water and NOx
emissions of 4.05 g/m3 of desalted water have been reported
(Raluy et al., 2005). When RO was operated with electricity
generated from wind or solar energy, GHG emissions were
substantially lower. For RO integrated with wind energy
resource, CO2 emissions were 0.1 kg/m3 and NOx emissions
were 0.4 g/m3 (Raluy et al., 2005). For RO integrated with solar
photovoltaic energy resource, CO2 emissions were 0.6e0.9 kg/
m3 and NOx emissions were 1.8e2.1 g/m3 (Raluy et al., 2005).
Thus, an important avenue for reducing GHG emissions is the
utilization of renewable energy sources in place of fossil fuels
(NAS, 2010). A comparison of renewable energy resources is
shown in Table 3 and is discussed in detail below.

3.1.

Solar energy

Solar energy is one of the most promising sources of renewable


energy. The quantity of solar energy received by earth is
a function of the season, with the highest quantity of incoming
solar energy received during the summer months (Kiehl and
Trenberth, 1997). Desalination using solar energy can be categorized as thermal and electromechanical processes. Thermal
processes use solar thermal energy whereas electromechanical
processes use photovoltaic cells.

3.1.1.

Solar thermal processes

Solar thermal desalination processes are characterized as


either direct or indirect processes (Qiblawey and Banat, 2008).
Direct processes consist of all parts integrated into one system
whereas indirect processes refer to heat coming from a separate solar collecting device such as solar collectors or solar
ponds. The predominant solar thermal processes that are
integrated with or used as desalination systems are solar stills
and solar ponds. Utilization of solar energy for desalination is
described in detail below.

3.1.1.1. Solar stills. A solar still is a simple device that is used to


convert saline water into drinking water (Qiblawey and Banat,
2008). Solar stills are used for direct solar desalination which is
mainly suited for small production systems and regions where
the freshwater demand is less than 200 m3 per day (Rodriguez,
2002). The solar still consists of a transparent cover (glass or
plastic) which encloses saline water. The principle of operation

1914

w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 1 9 0 7 e1 9 2 0

is evaporation and condensation. The solar still cover traps solar


energy within the enclosure. The trapped solar energy heats the
saline water causing evaporation and condensation on the inner
surface of the sloped transparent cover. As the saline water is
heated, its vapor pressure increases. The resultant water vapor
is condensed on the underside of the roof cover and runs down
into troughs, which collect the distillate. The distillate obtained
is of high quality with most salts, organic and inorganic
components removed. The solar still requires frequent flushing
to prevent salt precipitation. Design problems encountered with
solar stills are brine depth, vapor tightness of the enclosure,
distillate leakage, methods of thermal insulation, and cover
slope, shape, and material (Eibling and Talbert, 1971). In practice,
heat losses will occur through a still. Currently available stateof-the-art single-effect solar stills have an efficiency of approximately 30e40% (Rodriguez, 2002). The solar still of the basin
type is the oldest method and improvements in its design have
been made to increase its efficiency (Naim et al., 2003). Modifications using passive methods include basin stills, wick stills,
diffusion stills, stills integrated with greenhouses, and other
configurations (Ahsan et al., 2010; Tabrizi et al., 2010; Murugavel
and Srithar, 2011).
A major drawback with the solar still is the energy loss in the
form of latent heat of condensation. In order to solve this
problem, a humidification-dehumidification (HD) principle has
been developed (Mathioulakis et al., 2007). Solar desalination
based on the HD principle results in an increase in the overall
efficiency of the desalination plant and is therefore considered
a promising technique for small capacity, solar-driven desalination plants (Mathioulakis et al., 2007). The basic principle of
the HD process is the evaporation of high salinity water and
condensation of water vapor from the humid air at ambient
pressure. The HD process is based on the fact that air can be
mixed with significant quantities of vapor. The vapor carrying

capability of air increases with temperature. Fresh water is


produced by condensing out the water vapor, which results in
dehumidification of the air. A significant advantage of the HD
technology is that it provides a means for low pressure, low
temperature desalination that operates off of waste heat and is
potentially very cost competitive (Parekh et al., 2004). Since the
heat transfer coefficient of the condensing vapor from air
is much lower than for pure water, the heat transfer area
needed is enormously high and a disadvantage for the process
(Semiat, 2008).

3.1.1.2. Solar ponds. Solar ponds utilize a saline gradient to


reduce the heat loss that normally occurs when the less dense
water heated by the sun rises to the surface of a pond and
loses energy to the atmosphere by convection and radiation
(Kalagirou, 2005). The objective of the solar pond is to utilize
a saline gradient to combat the thermal gradient and create
a stagnant and insulating zone in the upper part of the pond that
traps the hot fluid in the lower section of the pond. New technologies combining solar thermal energy and desalination have
been shown to utilize up to 80% less energy than conventional
desalination technologies (Saltworks, 2010). In this thermoionic desalination technology, the energy consumption is
reduced by harnessing low temperature heat to overcome the
energy barrier for desalination. Salt water is evaporated to
produce a concentrated salt solution. The concentrated gradient
energy from the concentrated salt solution is then used to
operate a desalination system. Using desalination brine for solar
ponds not only provides a preferable alternative to environmental disposal, but also a convenient and inexpensive source
of solar pond salinity.
3.1.1.3. Concentrated solar power. A new study has estimated
that 25% of the worlds electricity could come from concentrated

Table 3 e Comparison of renewable energy resources for desalination.


Renewable
energy
resource
Solar

Wind

Geothermal

Application

Advantages

Disadvantages

Solar still: Direct conversion of


saline to potable water.
Solar pond: Utilization of
salinity gradient to store
heat and produce steam
for electricity generation.
Concentrated solar power:
Hot fluid used in turbine
generator for producing
electricity.
Photovoltaic cell: Conversion
of sunlight directly into
electricity to power
RO desalination.
Wind turbine: Wind energy
used to generate electricity
to power RO desalination.
Geothermal steam to
generate electricity to
power RO desalination.

Simple process. Inexpensive material of


construction can be utilized (Qiblawey
and Banat, 2008).
Beneficial use of desalination brine
(Qiblawey, 2008).
Same equipment used in conventional
power plants can be used for
concentrated solar power
plants (DOE, 2010).
Hybrid designs with other (wind)
renewable energy sources are
easily achievable. Well suited for
desalination plants requiring
electrical power (Eltawil et al., 2009).
Well suited for desalination plants
requiring electrical power
(Eltawil et al., 2009).
Continuous power output, predictable
resource, thermal storage is not
necessary (Barbier, 2002; EGEC, 2010).

Energy loss in the form of


latent heat of condensation
(Mathioulakis et al., 2007).
Large land area requirement
(Kalagirou, 2005).
Capital cost intensive. Output is
intermittent (Trieb et al., 2009).
Large land area requirement.
Capital cost intensive. Output
is intermittent (Kalagirou, 2005).

Output is intermittent. Resource


is location dependant and
unpredictable (Kalagirou, 2005).
Resource is limited to certain
locations (Kalagirou, 2005).

w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 1 9 0 7 e1 9 2 0

solar power (CSP) by the year 2050 (Trieb et al., 2009). Most
commercial CSP facilities use a system of curved mirrors to
collect the suns energy to heat a fluid flowing through tubes.
The hot fluid is then used to boil water in a conventional steamturbine generator to produce electricity. Concentrating solar
power typically uses a Dish/Sterling system. Other methods to
concentrate solar energy utilize a parabolic trough, solar tower,
or linear Fresnel (Trieb et al., 2009). Large mirror fields concentrate the sunlight to produce high temperature steam for power
generation that can be used for seawater desalination. Part of
the harvested solar thermal energy is used during the day and
conventional electricity is used during the night for continuous
operation.

3.1.2.

1915

Kalagirou, 2005). When electrical loads require an alternating voltage, an inverter is used to transform direct current
into alternating current. The batteries allow operation at
constant flow and pressure. They are sized to stabilize the
power supply to the RO unit on a daily basis, as well as to
account for fluctuations in solar energy and water demand.
Battery less PV-powered RO systems have been tested before
(Thomson and Infield, 2003) but certain disadvantages such as
longer operation in stand by mode needs to be overcome. It is
also common practice to connect PV systems to the local
electricity grid. During the day, the energy generated from the
PV systems is used directly from the grid and power is
utilized from the electricity grid at night. Thus, the grid acts as
an energy storage system.

Solar electromechanical process

Desalination using an electromechanical process involves the


application of photovoltaic (PV) cells. The PV process converts
sunlight directly into electricity. A PV cell consists of two or
more thin layers of semiconducting material (mostly silicon).
When the semiconducting material is exposed to sunlight,
electrical charges are generated and are conducted away by
metal contacts as direct current (DC). The PV sector has been
growing at 20% per annum or more for several years and is
now a multi-billion dollar business in Europe (Infield, 2009).
A total of around 5.95 GW of capacity has been installed
worldwide.
Photovoltaic cells are either monocrystalline silicon cells,
muticrystalline silicon cells, or amorphous cells. Monocrystalline cells are made of very pure monocrystalline silicon
whereas multicrystalline cells are produced using numerous
grains of monocrystalline cells (Kalagirou, 2005). The PV panel
is the principle building block of a PV system and any number
of panels can be connected together to give the desired electrical output. The choice of photovoltaic active material has
important effects on system design and performance. Both
the composition and atomic structure of the cell is an
important consideration for assessing performance. Materials
other than silicon which are used for photovoltaic cells are
silver, cadmium telluride, and copper indium selenide
(Kalagirou, 2005; Jacobson and Delucchi, 2009). The advantage
of using cadmium-based and copper-based material for PV
cells is that they are manufactured by relatively inexpensive
industrial processes when compared to crystalline silicon
technologies (Kalagirou, 2005). Limited supplies of tellurium
and indium could reduce the prospects for some types of thinfilm solar cells and large-scale production could be restricted
by the availability of silver (Jacobson and Delucchi, 2009).
The most popular combination of PV cells is with RO. PV is
considered a proper solution for small applications in sunny
areas (less than 50 m3/d) (Mathioulakis et al., 2007). The
feasibility of PV-powered RO for desalination at remote sites is
a proven combination. Since both technologies are fairly
mature, PV-powered RO requires minimum maintenance
(Bayod-Rujula and Martinez-Gracia, 2009). Stand-alone PV
systems are used in areas that are not easily accessible to
electricity. A stand-alone system is independent of the electricity grid, with the energy produced being stored in batteries
(Bayod-Rujula and Martinez-Gracia, 2009). Typically, a standalone PV-powered RO system will consist of a module,
batteries, and charge controller (Bouguecha et al., 2005;

3.2.

Wind energy

Wind has re-emerged as one of the most important and fastest growing sustainable energy resources since wind turbines
were first commercialized in the 1970s (Garca-Rodrguez,
2002; Ackermann and Soder, 2002). Wind turbines are
mature technologies and are commercially available. Windpowered desalination represents one of the most promising
renewable energy options for desalination, especially for
coastal areas with high availability of wind energy resources
(Zejli et al., 2004; Forstmeier et al., 2007).
After solar energy, wind energy is the most widely used
renewable energy source for small capacity desalination
plants (Kalagirou, 2005). The two common approaches for
wind-powered desalination systems include connecting both
the wind turbines and the desalination system to a grid, or
direct coupling of the wind turbines with the desalination
system (Ackermann and Soder, 2002). The latter option is
likely to be a stand-alone system at remote locations which
have no electricity grid. In this case, the desalination system
may be affected by power variations and interruptions caused
by the power source (wind). Hence, the stand-alone wind
desalination systems are often hybrid systems, combined
with another type of renewable energy source (for instance
solar), or a back-up system such as batteries or diesel generators (Mathioulakis et al., 2007). For stand-alone wind energydriven desalination units, the reported cost of fresh water
produced ranged from $1.35 per m3 to $6.7 per m3 when
compared to RO cost of about $1.0 per m3 (Karagiannis and
Soldatos, 2008; Mezher et al., 2011). The primary concern
with the use of wind energy for desalination is that wind
speed is highly variable. Wind speed varies both geographically and temporally and varies over a multitude of temporal
and spatial time scales (NREL, 2006). In terms of using a wind
turbine to generate power for desalination, this variation is
amplified by the fact that the available energy in the wind
varies as the cube of the wind speed (NREL, 2006). Thus, the
choice of location of the wind farm is critical for the exploitation of wind resources for power generation to ensure
superior economic performance.
The theoretical maximum aerodynamic conversion efficiency from wind to mechanical power for wind turbines is
59% (Kalagirou, 2005). The need to economize on blade costs
tends to lead to the construction of slender bladed, fast
running wind turbines with peak efficiencies close to 45%.

1916

w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 1 9 0 7 e1 9 2 0

Strategies for improving wind turbine performance include


the utilization of lighter turbine blades and improving power
storage systems (Spang, 2006; Thumthae and Chitsomboon,
2009). Wind turbines with blades that are approximately 40%
lighter than standard turbine blades could reduce capital costs
by up to 20e25% (Fairley, 2002). Lighter wind turbine design
with the blades placed on hinges allows flexibility for reduced
drag during high winds (Spang, 2006; Ameku et al., 2008).
Utilization of two blades per turbine rather than three blades
leads to lighter turbine design, but, the three-blade wind
turbines have a lower noise level than two-blade wind
turbines and the rotor moment of inertia is easier to handle.
Wind generators with vertical axis turbines convert wind
energy into electrical energy at a greater efficiency than
horizontal axis turbines and result in approximately 5%
increase in energy production and significantly reduce the
investment cost per kWh (Spang, 2006). With respect to the
materials available for constructing wind turbines, enough
concrete and steel exist and both these commodities are fully
recyclable (Jacobson and Delucchi, 2009). Increased production of wind turbines in the future could be slowed due to the
availability of rare-earth metals such as neodymium, but low
cost sources available in China could be utilized to compensate for a shortage of essential metals for manufacturing wind
turbines (Jacobson and Delucchi, 2009).

3.3.

Geothermal energy

Geothermal energy sources are classified in terms of the


measured temperature as low (<100  C), medium (100e150  C)
and high temperature (>150  C). Geothermal energy is usually
extracted with ground heat exchangers (Kalagirou, 2005). The
primary advantage of geothermal energy compared to solar and
wind, is that it is both continuous and predictable; as such,
thermal storage is unnecessary (Barbier, 2002). Geothermal
energy is being evaluated for desalination in Queensland,
Australia (Queensland Geothermal Energy Center of Excellence,
2010). This energy center has estimated that a geothermal plant
in the 1000e100,000 m3/d capacity can easily provide the entire
fresh water needs for an outback city at the cost of around
$0.73e1.46/m3. The first installation of geothermal energypowered desalination plants was reported by the Bureau of
Reclamation of the United States Department of the Interior in
the 1970s (Awerbuch et al., 1976). Geothermal energy has also
been investigated for desalination recently on the Baja California peninsula (Hiriart, 2008).
Geothermal energy presents a mature technology which
can be used to provide energy for desalination systems. High
temperature geothermal fluids generate electricity to power
RO plants and are used directly as shaft power for mechanically driven desalination. The main advantage of using
geothermal energy for desalination is that it is a stable and
reliable heat supply 24 h a day, 365 days a year. Also, desalination using geothermal energy source is environmentally
friendly with no emission of greenhouse gases (EGEC, 2010).

3.4.

Hybrid systems

Combinations of wind and solar energy have been used for


driving desalination systems (Koutroulis and Kolokotsa, 2010;

Karellas et al., 2011). The purpose of using hybrid wind-solar


systems for desalination is based on the fact that in certain
locations, wind and solar time profiles do not coincide
(Mathioulakis et al., 2007). The complementary features of wind
and solar resources make use of hybrid wind-solar systems to
drive desalinations systems (Charcosset, 2009). Hybrid windsolar PV systems have been implemented in the Sultanate of
Oman, Israel, Mexico, Germany, and Italy (Petersen et al., 1981;
Al Malki et al., 1998; Weiner et al., 2001; Pretner and Iannelli,
2002). Two RO desalination plants supplied by a 6 kW wind
energy converter and a 2.5 kW solar generator have been
designed for remote areas (Petersen et al., 1981). Stand-alone
systems for seawater desalination using hybrid wind-PV
system have also been designed (Mohamed and Papadakis,
2004). Using wind and solar conditions in Eritrea, East Africa,
the hourly water production was determined to be 35 m3/d with
a specific energy consumption of about 2.33 kWh/m3 (Gilau and
Small, 2008). Although several studies have been performed
using hybrid renewable energy desalination systems, none of
them represent large-scale applications.

3.5.
Design and implementation of renewable energy
systems
Renewable energy desalination systems need to be designed
using an iterative approach (Voivontas et al., 2001). The first
step of the approach involves the definition of a list of alternative technologies that satisfy the water demand. A second
step focuses on a detailed design analysis of each candidate
option made to determine the plant capacity, the structure of
the power unit and the operational characteristics. The final
step involves a financial analysis of the investment associated
with the selected renewable energy-desalination combination.
The most challenging issue associated with the implementation of renewable energy-desalination technology is the
optimum matching of the intermittent renewable energy
power output with the steady energy demand of the desalination process. Power supply management and demand-side
management are considered as the two options available to
address this problem (Voivontas et al., 2001). In the first case, an
appropriately controlled hybrid renewable energy resource unit
that is capable of providing a steady energy output is used. This
unit is sized at the nominal power demand of the desalination
process. In the demand-side management option, the desalination process only operates when the energy output of the
renewable energy resource unit is able to cover the energy
demand.
Other options available to address the issue of intermittent
renewable energy power output are different types of energy
storage such as electro-mechanical, virtual (through process
modification), and grid energy (Kalogirou, 1997). Compressed
air energy storage plants have also been used when energy
produced from a wind turbine exceeds grid load capacity
(BINE, 2010). For limited periods, the compressed air stores
cover the short-term reserve requirement, which are needed
due to the unpredictable forecasts of wind power feeding the
grid. In this case, wind turbines do not have to deactivate in
the event of a grid overload, and if there is excess supply of
electrical energy, the storage technology refines base-load
electricity, converting it to peak-load electricity (BINE, 2010).

w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 1 9 0 7 e1 9 2 0

The selection of the appropriate renewable energy


resource depends on several factors including plant size, feed
water salinity, remoteness, availability of grid electricity,
technical infrastructure, and the type and potential of the
local renewable energy resource and storage options. In
addition, socio-economic factors and policy need to be
considered as a driver for renewable energy resource implementation. The applicability of renewable energy resources
for desalination strongly depends on the local availability of
renewable energy and the quality of water required after
treatment. In addition, some combinations of resources are
better suited for large size plants, whereas some others are
better suited for small scale applications. Other important
factors that need to be considered are the capital cost of the
equipment and the land area required for the equipment
installation. When considering resource availability, solar
thermal energy and photovoltaics are considered to be a better
choice over wind and geothermal energy which are locationdependant. When considering the continuity and predictability of power output, geothermal energy is the most reliable
resource as the output is intermittent and less predictable for
solar thermal, photovoltaic, and wind energy.
Renewable energy resources provide various advantages
but their application for desalination has been limited. The
reasons for their limited application are technology, cost, and
availability. Although desalination technologies are mature,
technologies for the storage of renewable energy are not
completely mature and avenues for design improvements still
exist. Prices for renewable energy technologies are decreasing
but the capital costs still prohibit their commercialization at
a large-scale. Renewable energy is also not sufficiently available in certain locales and for regions with adequate supplies,
a lack of adequate storage strategies has sometimes impeded
development due to supply intermittency.

4.

1917

With respect to integrating renewable energy resources with


desalination systems in order to reduce GHG emissions,
research is still needed to bring down the cost of these alternative energy supplies, establish appropriate storage systems
to smooth out resource intermittency, and incorporate renewable sources into the electric grid for utilization by desalination
facilities in regions with poor renewable energy resources.

5.

Conclusions

Minimization of energy utilization by RO desalination can be


achieved in several ways. One promising avenue for reducing
desalination energy is through the use of advanced membrane
materials and application of innovative technologies. Innovative technologies such as forward osmosis and ion concentration polarization show promise but long-term operational
data are lacking. Integration of renewable energy resources is
needed to reduce GHG emissions and eliminate desalinations
dependency on fossil fuels and ultimately reduce the cost of
energy. The selection of appropriate renewable energy
resources depends on factors such as plant size, feed water
salinity, plant location, availability of grid electricity, technical
infrastructure, and the availability of local renewable energy
resources, and storage options. Technology advances are
important, but economic and political factors are also critical
to large-scale deployment of renewable energy.

Acknowledgements
The authors would like to thank the WateReuse Research
Foundation (WRRF) and the California Energy Commission
(CEC) for project funding (Project # WRRF-08 13).

Research needs

Minimization of the energy required for seawater RO desalination through utilization of efficient system design, high efficiency pumping, and energy recovery devices has been studied
extensively and near optimal performance characteristics have
already been tested and achieved. Further design improvements in these categories will only provide marginal reduction
in energy consumption further. Research avenues that show
the most promise for reducing energy usage lie in the development and testing of advanced membrane materials which
can enhance the performance of the membrane, in terms of flux
and rejection, and reduce feed pressure requirements. The
development of nanocomposite, nanotube, and biomimetic
membranes show promise but much more data are necessary
in order to validate the application of these membranes under
normal operation and chemical cleaning conditions. Innovative
technologies, such as forward osmosis, require a more efficient
recovery of the draw solution and methods to reduce internal
concentration polarization, inorganic scaling, and fouling of the
membrane. Similarly, the application of ion concentration
polarization and capacitive deionization technologies requires
further study in order to enhance feed water recovery to the
point where this technology might be economically feasible.

references

Ackermann, T., Soder, L., 2002. An overview of wind energystatus. Renewable and Sustainable Energy Reviews 6, 67e128.
Ahsan, A., Islam, K.M., Fukuhara, T., Ghazali, A.H., 2010.
Experimental study on evaporation, condensation and
production of a new tubular solar still. Desalination 260 (1e3),
172e179.
Al Malki, A., Al Amri, M., Al Jabri, H., 1998. Experimental study of
using renewable energy in the rural areas of Oman. Renewable
Energy 14, 319e324.
Ameku, K., Nagai, B.M., Roy, J.N., 2008. Design of a 3 kW wind
turbine generator with think airfoil blades. Experimental
Thermal and Fluid Science 32 (8), 1723e1730.
Anderson, M.A., Cudero, A.L., Palma, J., 2010. Capacitive
deionization as an electrochemical means of saving energy
and delivering clean water. Comparison to present
desalination practices: will it compete? Electrochemica Acta
55, 3845e3856.
Andrews, W.T., Laker, D.S., 2001. A twelve-year history of large
scale application of work exchanger energy recovery
technology. Desalination 138 (1e3), 201e206.
AquaZ, 2010. http://www.danfoss-aquaz.com/files/billeder/
Aquaporin_membrane.JPG (accessed 09.30.10).

1918

w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 1 9 0 7 e1 9 2 0

Avlonitis, S.A., Kouroumbas, K., Vlachakis, N., 2003. Energy


consumption and membrane replacement cost for seawater
RO desalination plants. Desalination 157, 151e158.
Awerbuch, L., Lindemuth, T.E., May, S.C., Rogers, A.N., 1976.
Geothermal energy recovery process. Desalination 19, 325e336.
BINE, 2010. Compressed air energy storage power plants.
Available from: http://www.bine.info/fileadmin/content/
Publikationen/Englische_Infos/projekt_0507_engl_internetx.
pdf (accessed 09.30.10).
Barbier, E., 2002. Geothermal energy technology and current
status: an overview. Renewable and Sustainable Energy
Reviews 6, 3e65.
Bartman, A.R., Christofides, P.D., Cohen, Y., 2009. Non-linear
model-based control of an experimental reverse-osmosis
water desalination system. Industrial and Engineering
Chemistry Research 48, 6126e6136.
Bartman, A.R., Zhu, A., Christofides, P.D., Cohen, Y., 2010.
Minimizing energy consumption in reverse osmosis
membrane desalination using optimization-based control.
Journal of Process Control 20, 1261e1269.
Bayod-Rujula, A.A., Martinez-Gracia, A., 2009. Photovoltaic system
for brackish water desalination by electrodialysis and electricity
generation. Desalination and Water Treatment 7, 142e151.
Bouguecha, S., Hamrouni, B., Dhahbi, M., 2005. Small scale
desalination pilots powered by renewable energy sources:
case studies. Desalination 183, 151e165.
Bowen, R., 2006. Biomimetic separations e learning from the early
development of biological membranes. Desalination 199,
225e227.
Cameron, I.B., Clemente, R.B., 2008. SWRO with ERIs PX pressure
exchanger device-a global survey. Desalination 221, 136e142.
Cardona, E., Piacentino, A., 2004. Optimal design of cogeneration
plants for seawater desalination. Desalination 166, 411e426.
Cath, Z., Childress, A., Elimelech, M., 2006. Forward osmosis:
principles, applications, and recent developments. Journal of
Membrane Science 281 (1e2), 70e87.
Chang, Y., Reardon, D.J., Kwan, P., Boyd, G., Brant, J., Rakness, K.,
Furukawa, D., 2008. Evaluation of dynamic energy
consumption of advanced water and wastewater treatment
technologies. AWWARF Final Report, ISBN 978-1-60573-033-2.
Charcosset, C., 2009. A review of membrane processes and
renewable energies for desalination. Desalination 245,
214e231.
Choi, Y., Choi, J., Oh, H., Lee, S., Yang, D., Kim, J., 2009. Toward
a combined system of forward osmosis and reverse osmosis
for seawater desalination. Desalination 247 (1e3), 239e246.
Chou, S., Shi, L., Wang, R., Tang, C.Y., Qiu, C., Fane, A.G., 2010.
Characteristics and potential applications of a novel forward
osmosis hollow fiber membrane. Desalination 261 (3),
365e372.
Corry, B., 2008. Designing carbon nanotube membranes for
efficient water desalination. The Journal of Physical Chemistry
B 112 (5), 1427e1434.
Dermentzis, K., Ouzounis, K., 2008. Continuous capacitive
deionization e electrodialysis reversal through electrostatic
shielding for desalination and deionization of water.
Electrochimica Acta 53 (24), 7123e7130.
Eibling, J.A., Talbert, S.G., 1971. Solar stills for community use e
digest of technology. Solar Energy 13, 263e276.
Eltawil, M.A., Zhengming, Z., Yuan, L., 2009. A review of
renewable energy technologies integrated with desalination
systems. Renewable and Sustainable Energy Reviews 13,
2245e2262.
European Geothermal Energy Council (EGEC), 2010. Geothermal
desalination. Available from: http://www.egec.org/target/
Brochure%20DESALINATION.pdf (accessed 07.20.10).
Fairley, P., 2002. Wind Power for Pennies. MIT Technology Review,
Cambridge, MA.

Farmar, J.C., Tran, T.D., Richardson, J.H., Fix, D.V., May, S.C.,
Thomson, S.L., 1997. The application of carbon aerogel
electrodes to desalinate and waste treatment. Report No.
231717. Lawrence Livermore National Laboratory.
Forstmeier, M., Mannerheim, F., DAmato, F., Shah, M., Liu, Y.,
Baldea, M., Stella, A., 2007. Feasibility study on wind-powered
desalination. Desalination 203, 463e470.
Garca-Rodrguez, L., 2002. Seawater desalination driven by
renewable energies: a review. Desalination 143, 103e113.
Gilau, A.M., Small, M.J., 2008. Designing cost-effective seawater
reverse osmosis system under optimal energy options.
Renewable Energy 33, 617e630.
Gleick, P.H., 2006. The Worlds Water 2006e2007, The Biennial
Report on Freshwater Resources. Island Press, Chicago.
Global Water Intelligence (GWI), 2010. Desalination Markets, 2010.
Gong, X., Li, J., Lu, H., Wan, R., Li, J., Hu, J., Fang, H., 2007. A chargedriven molecular water pump. Nature Nanotechnology 2,
709e712.
Greenlee, L.F., Lawler, D.F., Freeman, B.D., Marrot, B., Moulin, P.,
2009. Reverse osmosis desalination: water sources, technology,
and todays challenges. Water Research 43, 2317e2348.
Gude, V.G., Nirmalakhandan, N., Deng, S., 2010. Renewable and
sustainable approaches for desalination. Renewable and
Sustainable Energy Reviews 14 (9), 2641e2654.
Guillen, G., Hoek, E.M.V., 2009. Modeling the impacts of feed
spacer geometry on reverse osmosis and nanofiltration
processes. Chemical Engineering Journal 149 (1e3), 221e231.
Hilder, T.A., Gordon, D., Chung, S., 2009. Salt rejection and water
transport through boron nitride nanotubes. Small 5 (19),
2183e2190.
Hiriart, G., 2008. Geothermal energy for desalination seawater.
International Geological Congress, Oslo.
Hoek, E.M.V., Ghosh, A., 2009. Nanotechnology-based membranes
for water purification. Nanotechnology Applications for Clean
Water, 47e58.
Holt, J., Park, H.G., 2006. Fast mass transport through sub 2nanometer carbon nanotubes. Science 312 (5766), 1034e1037.
Infield, D.G., 2009. An overview of renewable energy technologies
with a view to stand alone power generation and water
provision. Desalination 248, 494e499.
Jacobson, M.Z., Delucchi, M.A., 2009. A Path to Sustainable
Energy. Scientific American, pp. 58e65.
Jeong,B.H.,Hoek,E.M.V.,Yan,Y.,Huang,X.,Subramani,A.,Hurwitz,G.,
Ghosh, A.K., Jawor, A., 2007. Interfacial polymerization of thin film
nanocomposites: a new concept for reverse osmosis membranes.
Journal of Membrane Science 294, 1e7.
Jia, Y., Li, H., Wang, M., Wu, L., Hu, Y., 2010. Carbon nanotube:
possible candidate for forward osmosis. Separation and
Purification Technology 75 (1), 55e60.
Kalagirou, S.A., 2005. Seawater desalination using renewable
energy sources. Progress in Energy and Combustion Science
31, 242e281.
Kalogirou, S.A., 1997. Survey of solar desalination systems and
system selection. Energy e The International Journal 22, 69e81.
Karagiannis, I.C., Soldatos, P.G., 2008. Water desalination cost
literature: review and assessment. Desalination 223, 448e456.
Karellas, S., Terzis, K., Manolakos, D., 2011. Investigation of an
autonomous hybrid solar thermal ORC e PV RO desalination
system. The Chalki island case. Renewable Energy 36 (2), 583e590.
Kiehl, J.T., Trenberth, K.E., 1997. Earths annual global mean
energy budget. Bulletin of the American Meteorological
Association 78, 197e208.
Kim, S.J., Ko, S.H., Kang, K.H., Han, J., 2010. Direct seawater
desalination by ion concentration polarization. Nature
Nanotechnology 5, 297e301.
Koutroulis, E., Kolokotsa, D., 2010. Design optimization of
desalination systems power-supplied by PV and W/G energy
sources. Desalination 258 (1e3), 171e181.

w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 1 9 0 7 e1 9 2 0

Kumar, M., Grzelakowski, M., Zilles, J., Clark, M., Meier, W., 2007.
Highly permeable polymeric membranes based on the
incorporation of the functional water channel protein
Aquaporin Z. Proceedings of the National Academy of
Sciences 104 (52), 20719e20724.
Lee, J., Park, K., Yoon, S., Park, P., Park, K., Lee, C., 2009.
Desalination performance of a carbon-based composite
electrode. Desalination 237 (1e3), 155e161.
Loeb, S., Sourirajan, S., 1963. Seawater demineralization by
means of an osmotic membrane. Advances in Chemistry
Series 38, 117e132.
Long, B., 2008. Optimization of desalination for low energy.
Presentation at the Singapore International Water Week,
Singapore.
Macedonio, F., Drioli, E., 2010. An exergetic analysis of a membrane
desalination system. Desalination 261 (3), 293e299.
Manth, T., Gabor, M., Oklejas, E., 2003. Minimizing RO energy
consumption under variable conditions of operation.
Desalination 157 (1e3), 9e21.
Mathioulakis, E., Belessiotis, V., Delyannis, E., 2007. Desalination
by using alternative energy: review and state-of-the-art.
Desalination 203, 346e365.
McCutcheon, J., McGinnis, R.L., Elimelech, M., 2005. A novel
ammoniaecarbon dioxide forward (direct) osmosis
desalination process. Desalination 174 (1), 1e11.
McCutcheon, J., McGinnis, R.L., Elimelech, M., 2006. Desalination
by ammoniaecarbon dioxide forward osmosis: influence of
draw and feed solution concentrations on performance on
process performance. Journal of Membrane Science 278 (1e2),
114e123.
McGinnis, R.L., Elimelech, M., 2007. Energy requirements of
ammoniaecarbon dioxide forward osmosis desalination.
Desalination 207, 370e382.
McGinnis, R.L., McCutcheon, J., Elimelech, M., 2007. Forward
Osmosis Energy Use: Comparisons to RO, MSF, MED.
Presentation at North American membrane Society (NAMS)
annual Conference, April 14e17, Orlando, Florida.
Mezher, T., Fath, H., Abbas, Z., Khaled, A., 2011. Techno-economic
assessment and environmental impacts of desalination
technologies. Desalination 266 (1e3), 263e273.
Mi, B., Elimelech, M., 2010. Organic fouling of forward osmosis
membranes: fouling reversibility and cleaning without
chemical reagents. Journal of Membrane Science 348 (1e2),
337e345.
Mirza, S., 2008. Reduction of energy consumption in process
plants using nanofiltration and reverse osmosis. Desalination
224, 132e142.
Mohamed, E., Papadakis, G., 2004. Design, simulation and
economic analysis of a stand-alone reverse osmosis
desalination unit powered by wind turbines and
photovoltaics. Desalination 164, 87e97.
Murugavel, K.K., Srithar, K., 2011. Performance study on basin
type double solar still with different wick materials and
minimum mass of water. Renewable Energy 36 (2),
612e620.
Naim, M., Mervat, A., El-Kawi, A., 2003. Non-conventional solar
stills. Part I: non-conventional solar stills with charcoal
particles as absorber medium. Desalination 153, 55e64.
NanoH2O, 2010. Nanotechnology advances reverse osmosis
membrane performance. Available from: http://www.
nanoh2o.com/Technology.php5?categoryEconomics
(accessed 09.30.10).
National Academy of Sciences (NAS), 2010. Electricity from
renewable resources: status, prospects, and impediments. The
National Academy Press, Washington D.C.
National Renewable Energy Laboratory (NREL), 2006. Integrated
wind energy/desalination system. Final Report SR500e39485.

1919

Oklejas, M., Stidham, K., Weidmann, M., 2005. Improve energy


recovery in gas processing plant using an HPT. Hydrocarbon
Process, 43e46.
Oren, Y., 2008. Capacitive deionization (CDI) for desalination and
water treatment e past, present and future (a review).
Desalination 228 (1e3), 10e29.
Parekh, S., Farid, M., Selman, J., Al-Hallaj, S., 2004. Solar
desalination with a humidificationedehumidification
technique e a comprehensive technical review. Desalination
160, 167e186.
Petersen, G., Fries, S., Mohn, J., Muller, A., 1981. Wind and solar
powered reverse osmosis desalination units e design, start up,
operating experiences. Desalination 39, 125e135.
Petry, M., Sanz, M.A., Langlais, C., Bonnelye, V., Durand, J.,
Guevara, D., Nardes, W.M., Saemi, C.H., 2007. The El Coloso
(Chile) reverse osmosis plant. Desalination 203, 141e152.
Pretner, A., Iannelli, M., 2002. Feasibility study and assessment of
the technical, administrative and financial viability of the
Voltano desalination plant (Agrigento, Sicily). Desalination
153, 313e320.
Qiblawey, H.M., Banat, F., 2008. Solar thermal desalination
technologies. Desalination 220, 633e644.
Queensland Geothermal Energy Center of Excellence, 2010.
Available from: http://www.uq.edu.au/news/?article20223
(accessed 09.30.10).
Raluy, R.G., Serra, L., Uche, J., 2005. Life cycle assessment of
desalination technologies integrated with renewable energies.
Desalination 183, 81e93.
Rodriguez, L.G., 2002. Seawater desalination driven by renewable
energies: a review. Desalination 143, 103e113.
Saltworks, 2010. Thermo-ionic energy conversion. Available
from: http://www.saltworkstech.com/technology.php
(accessed 09.01.10).
Sauvet-Goichon, B., 2007. Ashkelon desalination plant e
a successful challenge. Desalination 203, 75e81.
Schafer, A.I., Broeckmann, A., Richards, B.S., 2007. Renewable
energy powered membrane technology. 1. Development and
characterization of a photovoltaic hybrid membrane system.
Environmental Science and Technology 41 (3), 998e1003.
Semiat, R., 2008. Energy issues in desalination processes.
Environmental Science and Technology 42 (22), 8193e8201.
Service, R.F., 2006. Desalination freshens up. Science 313,
1088e1090.
Sholl, D.S., Johnson, J.K., 2006. Making high-flux membranes with
carbon nanotubes. Science 312 (5776), 1003e1004.
Souari, L., Hassairi, M., 2007. Seawater desalination by reverse
osmosis: the true needs for energy. Desalination 206 (1e3),
465e473.
Spang, E., 2006. The potential for wind-powered desalination in
water-scarce countries. Masters Thesis, The Fletcher School,
Tufts University, Massachusetts, United States.
Stover, R., 2004. Development of a fourth generation energy
recovery device. A CTOs notebook. Desalination 165, 313e321.
Stover, R., 2007. Seawater reverse osmosis with isobaric energy
recovery devices. Desalination 203, 168e175.
Subramani, A., Kim, S., Hoek, E.M.V., 2006. Pressure, flow, and
concentration profiles in open and spacer-filled membrane
channels. Journal of Membrane Science 277 (1e2), 7e17.
Sui, H., Han, B.G., Lee, J.K., Walian, P., Jap, B.K., 2001. Structural
basis of water specific transport through the AQP1 water
channel. Nature 414, 872e878.
Tabrizi, F.F., Dashtban, M., Moghaddam, H., Razzaghi, K., 2010.
Effect of water flow rate on internal heat and mass transfer
and daily productivity of a weir-type cascade solar still.
Desalination 260 (1e3), 239e247.
Thomson, M., Infield, D., 2003. A photovoltaic e powered
seawater reverse e osmosis system without batteries.
Desalination 153 (1e3), 1e8.

1920

w a t e r r e s e a r c h 4 5 ( 2 0 1 1 ) 1 9 0 7 e1 9 2 0

Thumthae, C., Chitsomboon, T., 2009. Optimal angle of attack for


untwisted blade wind turbine. Renewable Energy 34 (5),
1279e1284.
Torre, A., 2008. Efficiency optimization in SWRO plant: high
efficiency and low maintenance pumps. Desalination 221
(1e3), 151e157.
Trieb, F., Muller-Steinhagen, H., Kern, J., Scharfe, J., Kabariti, M.,
Al Taher, A., 2009. Technologies for large scale seawater
desalination using concentrated solar radiation. Desalination
235 (1e3), 33e43.
Truskett, T.M., 2003. Subtleties of water in small spaces.
Proceedings of the National Academy of Science 100 (18),
10139e10140.
van Paassen, J., van der Meer, W., Post, J., 2005. Optiflux: from
innovation to realization. Desalination 178, 325e331.
Veerapaneni, S., Jordan, B., Leitner, G., Freeman, S., Madhavan, J.,
2005. Optimization of RO desalination process energy
consumption. International Desalination Association World
Congress, Singapore.
Veerapaneni, S., Long, B., Freeman, S., Bond, R., 2007. Reducing
energy consumption for desalination. Journal of the American
Water Works Association 99 (6), 95e106.
Voivontas, D., Misirlis, K., Manoli, E., Arampatzis, G.,
Assimacopoulos, D., Zervos, A., 2001. A tool for the design of
desalination plants powered by renewable energies.
Desalination 133, 175e198.
Voutchkov, N., 2004. Seawater desalination costs cut through
power plant co-location. Filtration and Separation 41 (7), 24e26.
Voutchkov, N., 2007. Advances in seawater desalination
technology. Water Conditioning and Purification Available
from: http://www.wcponline.com/pdf/0709Voutchkov.pdf
(accessed 09.07.10).
Voutchkov, N., 2010. Membrane Seawater Desalination Overview
and Recent Trends. Desalination: An energy solution.
Presentation at International Desalination Association
Conference, November 2e3, Huntington Beach, California.
Wang, Y., Wang, S.C., Xu, S.C., 2004. Experimental studies on
dynamic process of energy recovery device for RO
desalination plants. Desalination 160 (2), 187e193.
Wang, Y., Wang, S., Xu, S., 2005. Investigations on characteristics
and efficiency of a positive displacement energy recovery unit.
Desalination 177 (1e3), 179e185.

Weiner, D., Fisher, D., Moses, E.J., Katz, B., Meron, G., 2001.
Operation experience of a solar- and wind-powered
desalination demonstration plant. Desalination 137, 7e13.
Welgemoed, T.J., 2005. Capacitive Deionization Technology:
development and evaluation of an industrial prototype
system. Ph.D. Dissertation, University of Pretoria.
Wilf, M., Bartels, C., 2005. Optimization of seawater RO systems
design. Desalination 173, 1e12.
Wilf, M., Hudkins, J., 2010. Energy Efficient Configuration of RO
Desalination Units. Proceedings of Water Environment
Federation Membrane Applications Conference, Anaheim,
California.
Yang, Q., Wang, K.Y., Chung, T.S., 2009. Dual-layer hollow fibers
with enhanced flux as novel forward osmosis membranes for
water production. Environmental Science and Technology 43
(8), 2800e2805.
Zejli, D., Benchrifa, R., Bennouna, A., Zazi, K., 2004. Economic
analysis of wind-powered desalination in the south of
Morocco. Desalination 165, 219e230.
Zhu, A., Christofides, P.D., Cohen, Y., 2009. Effect of
thermodynamic restriction on energy cost optimization of RO
membrane water desalination. Industrial and Engineering
Chemistry Research 48, 6010e6021.
Zhu, A., Christofides, P.D., Cohen, Y., 2009a. Minimization of
energy consumption for a two-pass membrane desalination:
effect of energy recovery, membrane rejection and retentate
recycling. Journal of Membrane Science 339, 126e137.
Zhu, A., Christofides, P.D., Cohen, Y., 2009b. Energy consumption
optimization of reverse osmosis membrane water
desalination subject to feed salinity fluctuation. Industrial and
Engineering Chemistry Research 48, 9581e9589.
Zhu, A., Christofides, P.D., Cohen, Y., 2009c. On RO membrane
and energy costs and associated incentives for future
enhancements of membrane permeability. Journal of
Membrane Science 344 (1e2), 1e5.
Zhu, A., Rahardianto, A., Christofides, P.D., Cohen, Y., 2010.
Reverse osmosis desalination with high permeability
membranes e cost optimization and research needs.
Desalination and Water Treatment 15, 256e266.
Zhu, A., Christofides, P.D., Cohen, Y., 2010a. Effect of stream
mixing on RO energy cost minimization. Desalination 261 (3),
232e239.

You might also like