Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Desalination 342 (2014) 107–117

Contents lists available at ScienceDirect

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

Concentration of NaCl from seawater reverse osmosis brines for the


chlor-alkali industry by electrodialysis
Mònica Reig a, Sandra Casas a,d, Carlos Aladjem b, César Valderrama a,⁎, Oriol Gibert a,d, Fernando Valero c,
Carlos Miguel Centeno c, Enric Larrotcha d, José Luis Cortina a,d
a
Chemical Engineering Dept. UPC-Barcelona TECH, Av. Diagonal 647, 08028 Barcelona, Spain
b
SOLVAY Ibérica SL.C/Marie Curie 1-3-5, 08760 Martorell, Spain
c
Aigües Ter LLobregat (ATLL) Sant Martí de l'Erm, 30. 08970 Sant Joan Despí, Barcelona, Spain
d
CETAQUA Carretera d'Esplugues, 75, 08940 Cornellà de Llobregat, Spain

H I G H L I G H T S

• IXM-ED pilot plant was used to evaluate the efficiency in concentrating a SWRO brine.
• NaCl concentration was evaluated covering the temperature range in a semiarid climate.
• Results show that IXM-ED can concentrate SWRO brines up to 245 g/L NaCl.
• Multivalent ions removal was achieved due to IXM selectivity and ion complexation.
• Energy consumption was 0.12 kWh/kg NaCl for 185 g NaCl/l at 27 °C and 0.35kA/m2.

a r t i c l e i n f o a b s t r a c t

Article history: Currently, numerous studies are focused on the valorisation of seawater desalination reverse osmosis brines.
Received 3 August 2013 Electrodialysis can be used to concentrate one of the primary components (NaCl) and obtain a suitable raw ma-
Received in revised form 17 December 2013 terial for industrial applications, such as the chlor-alkali industry. An electrodialysis pilot plant was used to eval-
Accepted 18 December 2013
uate the efficiency of concentrating a seawater reverse osmosis (SWRO) brine under representative full-scale
Available online 11 January 2014
operational conditions covering the temperature range of a semiarid climate. The results indicate that electrodi-
Keywords:
alysis is a technology that can concentrate SWRO brines from approximately 70 to 245 g/L NaCl, achieving an ad-
Ion exchange membranes ditional intrinsic purification of major multivalent ions (Ca2+, Mg2+, SO2− 4 ) due to the selectivity patterns of ion

Energy consumption exchange membranes and the ion-complexation reactions in the concentrated brines. However, minor compo-
Brine reuse nents, such as Ni and Cu, are concentrated due to the formation of Cu and Ni complexes with chloride ions to
Chlor-alkali industry form monocharged species (e.g., NiCl+ and CuCl+). Energy consumption values of 0.12 kWh/kg NaCl for 185 g
NaCl/l at 27 °C and 0.35 kA/m2 or 0.19 kWh/kg NaCl for 203 g NaCl/l at 27 °C and 0.50 kA/m2 were reached.
These results were compared with the data obtained from the literature for salt production by electrodialysers.
© 2013 Elsevier B.V. All rights reserved.

1. Introduction the high concentrations of salts and increases in the concentration of


transition and heavy metals [4–8].
The desalination of seawater and brackish water sources is a com- Traditional management of RO concentrates from desalination
mon method for providing fresh drinking water around the world. The plants is mainly conditioned by the location of the plant. In the case of
use of reverse osmosis (RO) membranes is becoming an increasingly in-land plants, well injection into deep aquifers is one of the preferred
popular option for desalination, due to significant improvements in en- options. In the last decade, new demonstration projects have been ad-
ergy recovery systems and pre-treatment processes over the past cou- dressed to achieve an effluent volume reduction by either solar evapo-
ple of decades [1–3]. However, the disposal of the brines generated by ration ponds or thermal evaporation [9,10]. Brine volume reduction by
the desalination process poses significant environmental issues, due to evaporation techniques results in a solid product that can more easily
be disposed of compared to the original concentrate, whereas the low
salinity effluent can be reused to increase the water production ratio
or it can be directly discharged into surface or ground water bodies
⁎ Corresponding author at: Departament d’Enginyeria Química, Universitat Politècnica
de Catalunya, Av. Diagonal 647, 08028 Barcelona, Spain. Tel.: +34 93 4011818; fax:
[11–13].
+34 93 401 58 14. The normal management option in coastal areas is the direct dis-
E-mail address: cesar.alberto.valderrama@upc.edu (C. Valderrama). charge into the sea or, after appropriate dilution, with a secondary

0011-9164/$ – see front matter © 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.desal.2013.12.021
108 M. Reig et al. / Desalination 342 (2014) 107–117

discharge (e.g., cooling streams of power plants) [14,15], although strin- using RO brines saturated with CaSO4 and silica, brines were concentrat-
gent regulations limit the development of desalination plants due to the ed from 1.5% to approximately 10% with similar specific energy require-
restrictions and requirements on brine discharge. Environmentally ments (stack cell) and electrical efficiency as the previous study.
sound disposal of seawater and industrial desalination brines will be a Assuming that this process or similar techniques can be adopted, the
growing need in the next decades [16,17]. production of highly concentrated brines that can be thermally treated
In the case of direct discharge, due to the adverse effects of brine dis- to a dry product for the recovery of selected chemical resources can be
charge, together with its associated costs, research efforts in the last de- performed. The use of SWRO brines as raw material for salt production
cades have been centred on reducing the impact of brines by reducing was evaluated in the seventies by the Government Chemical Industrial
the volume and/or by diminishing the most critical pollution load Research Institute of Japan [35]. Seawater was desalinated by a flash
[18,19]. However, the exponential increase of desalination potential, es- evaporator (e.g., RO was not fully developed at this stage), and the
pecially in the south of Europe, the Middle East and Australia, has pro- discharged brine was concentrated by ED for further use in the produc-
moted the concept of defining valorisation routes of brines or brines tion of chlorine and sodium chloride. Tanaka et al. [36] recently sug-
by-products. gested that for areas with fresh water availability problems (e.g.
Conventionally, four components are extracted from seawater by Southeast Asia, the Middle and Near East), where desalination plants
evaporation: table salt (sodium chloride) and the by-products potassi- are installed, RO brine valorisation will be a need. Tanaka et al. [36] iden-
um chloride, magnesium salts and bromide salts [20–22]. Extraction tified the possibility to obtain salt, which is not produced for industrial
processes for other components may be feasible because they are suffi- use in a manufacturing plant by means of ED technology. Moreover, in-
ciently valuable or rare on land, as proposed by Le Dirach et al. [23] who dustry in these countries is in development and the consumption of
identified eight elements (sodium, magnesium, rubidium, potassium, chlorine and sodium hydroxide is expected. In this study [36] was deter-
phosphor, indium, caesium and germanium) that are potentially mined that the energy consumption in a salt manufacturing process
economically and technically viable. More recently, Jepessen et al. [24] (200,000 t NaCl/y as production capacity) using the brine discharged
examined the potential for economic extraction of rubidium and phos- from an RO seawater desalination plant was 80% of the energy con-
phorus and analysed the potential cost of potable water production for sumption in the process using seawater. It was also found that the opti-
varying levels of extraction of sodium chloride, taking as reference for mum current density that minimised energy conditions was 0.3 kA/m2
the calculations a typical, large RO plant (100,000 m3/d). These types for ED fed with either brine discharged from the RO desalination plant
of solutions are aligned with recent developments in zero liquid dis- or seawater.
charge desalination systems [25]. These entail further processing of Most of the research work has been based on the technical feasibility
the concentrate until dry salts are obtained, which can be disposed to of isolating by-products, mainly as salts, where morphology and purity
landfills or used in other applications. These systems have multiple ben- requirements need to be fulfilled. Analysis of the total amounts of
efits: avoiding discharge to surface or ground waters, flexibility in site potential raw materials present in the brines (NaCl, MgCl2, MgSO4,
selection, and efficient reuse of water [26]. From an environmental per- CaCO3, KCl, KBr, and CO2(g)) has opened the possibility of linking the
spective, such systems are desirable; however, further concentration of seawater desalination plants to chemical production plants (e.g.,
the brine can be achieved only by thermal processes, which adds signif- chlor-alkali and fertiliser industries) [37–40] or even energy production
icantly to the overall cost of desalination [27]. (liquid fuels), as recently proposed by Eisaman et al. [41], with a process
These problems can be overcome by developing 1) technologies that based on CO2(g) extraction from seawater using bipolar-electrodialysis.
reduce brine volumes to increase salts concentrations and 2) separation The present work reports the evaluation of ED as a technological solu-
technologies for the isolation of desired salts and purification, such as tion to recover a by-product (NaCl) for the chlor-alkali industry from
crystallisation processes [28]. The different options have been thor- SWRO desalination brines.
oughly reviewed [21,29]. Conventional treatments, such as concentra-
tion in evaporation ponds, have several disadvantages, such as 2. Materials and Methods
extensive land use and low productivity [30,31]. Thus, investigation on
new approaches to improve the valorisation options is a current de- 2.1. Seawater reverse osmosis (SWRO) desalination brines
mand. Membrane distillation (e.g., vacuum membrane distillation) has
been studied as an alternative for the processing of highly concentrated The ED unit was fed with SWRO brine pumped directly from the
aqueous solutions over conventional distillation processes because it brine deposits of the El Prat Seawater Desalination Plant (Aigues Ter
operates in the range of 60–80 °C and provides a high contact area per Llobregat), Barcelona (Spain). The average composition is shown in
unit of equipment volume, allowing very compact installations and a re- Table 1. The brine was concentrated under different operation condi-
duced footprint. However, development is still at the research level tions to obtain the maximum NaCl concentration and the minimum en-
without successful industrial implementation, and the efforts are ergy consumption. Several experiments were conducted in a range of
centred on the recovery of valuable by-products to reduce the high cap- temperatures from 10 to 28 °C and currents densities from 0.30 to
ital and operational expenditures and make it more economically at- 0.60 kA/m2. The experimental programme included 24 months of oper-
tractive [29–31]. ation to cover the lower seawater temperature of the Mediterranean
The major technical barrier to achieving zero liquid discharge sys-
tems has historically been the problem of scale formation (precipitation
of alkaline earth metals such as CaCO3, Mg(OH)2 and CaSO4) in highly Table 1
Composition of seawater reverse osmosis desalination brine of El Prat Desalination Plant
concentrated brines [32]. Recently, Kornold et al. [33] proposed electro- (Barcelona, Spain).
dialysis (ED) to increase the brine concentration and thus reducing both
the volume of the brine effluent and the cost of its disposal. The main Major components Concentration Minor components Concentration
(g/L) (mg/L)
drawback of this technique (membrane precipitation of CaSO4) is the
continuous removal of gypsum from the brine using a separate precipi- Cl− 38.8 ± 0.4 SiO2 b1
Na+ 20.8 ± 0.3 Al3+ b0.5
tator, in which gypsum seeds precipitated the excess CaSO4 in the
SO2−
4 5.41 ± 0.2 Fe3+, Fe3+ b0.2
oversaturated solution. By applying ED to synthetic brine effluents, sim- Mg2+ 2.64 ± 0.2 Ba2+ b0.2
ulating effluents from the desalination of brackish and industrial waters, Ca2+
0.83 ± 0.04 Ni2+ 0.07 ± 0.02
their salt concentration was increased from 0.2–2% to 12–20% with an K+ 0.75 ± 0.05 Cu2+ 0.03 ± 0.01
energy consumption of 1.0–7.0 kW h/m3 in contrast to approximately Br− 0.13 ± 0.06 Mn2+ 0.01 ± 0.01
Sr2+ 0.016 ± 0.003 Cr3+ 0.007 ± 0.003
25 kW h/m3 needed by thermal evaporation [34]. In a second study
M. Reig et al. / Desalination 342 (2014) 107–117 109

Sea in winter time (10 °C) and the higher seawater temperatures of 2.4. Limiting current density (LCD)
summer time (28 °C).
As the brine was concentrated along the ED process, the potential The maximum operating current intensity using SWRO brine was
precipitation of the brine components was estimated. The saturation established at the ED pilot plant using a procedure described previously
index of minerals whose precipitation could limit the concentration [43]. In the voltage range studied with the SWRO brine, thanks to the
step was calculated with the PHREEQC code [42] using the PITZER equi- single-pass configuration and the high concentration of the solution,
librium data base. The saturation index of inlet brine and its evolution as no increase of resistance was observed, and the LCD could not be deter-
a function of the concentration phenomena occurring on the concen- mined. Although the maximum nominal voltage that could be applied
trate circuit were calculated to determine the potential risks of to the stack was 65 V, which corresponds to a maximum current density
precipitation. of 0.65 kA/m2, the pilot plant operated at current densities not higher
than 0.60 kA/m2 to avoid the formation of H2 bubbles inside the stack.
To reduce this problem, the cell was placed vertically and a second gas-
2.2. Analytical methodology and chemical analysis ket in the cathode circuit was added to help with gas removal.

Samples were taken from the feed tank, the inlet brine, and the 2.5. Evaluation of the brine concentration data analysis: concentration
diluate and concentrate flows leaving the stack every 2 hours. First, the factors (CF), membrane selectivity (SAB ) and energy consumption (ENaCl)
pH was measured in every sample using a glass electrode. The NaCl,
SO2−
4 , Ca
2+
and Mg2+ levels in the samples were analysed using differ- The concentration factors (CF) were calculated according to Eq. (1):
ent analytical methods. Cl−was determined potentiometrically through
precipitation with AgNO3 and an AgCl electrode in a METHROM 721 in- C A ðt Þ
C FA ¼ ð1Þ
strument. Ca2+ and Mg2+ were determined by atomic absorption spec- C A;0
trophotometry using a Perkin Elmer Analyst 300. SO2− 4 was determined
by ionic chromatography using a Methrom 761 Compact IC equipped where CA,0 and CA(t) are initial concentration and the concentration at a
with an Anion Dual 2–6.1006.100 column. Finally, NaCl was determined given time t of component A in the concentrated loop, respectively.
by electrically balancing the major ions of the solution. The conductivity The membrane selectivity (SAB) was calculated as proposed by Zhang
and temperature of the diluate and concentrate streams were also mon- et al. [44] according to Eq. (2):
itored during operation to ensure the concentration process was success-
A C F A −C F B
ful. Minor and trace elements of the brines (Sr, Cu, Ni, Cd, Hg, and Pb) SB ¼ ð2Þ
ð1−C F A Þ þ ð1−C F B Þ
were analysed by ICP-OES, Variant 725 at the end of each experiment.
Along the experimental programme, the membrane stack was open
where CFA and CFB are the concentration factors of components A and B,
for membrane maintenance when needed. Membrane surfaces, espe-
respectively.
cially those close to both electrodes, were examined by scanning elec-
The energy consumption (E) for the NaCl concentration was calcu-
tron microscopy (SEM) and energy dispersive X-ray spectroscopy
lated in two different ways:
(EDS) using a JEOL 3400® scanning electron microscope with an energy
dispersive system. a) In a batch process, EB(NaCl) ((kWh/kg NaCl)), as the energy necessary
to increase the NaCl concentration in the tank, was calculated ac-
cording to Eq. (3):
2.3. ED pilot plant description and operation
V cell  I·t
EBðNaClÞ ¼ ð3Þ
The ED pilot plant has been described elsewhere [43]. The ion- v tank  C NaCl
exchange membranes stack was an EURODIA AQUALIZER SV-10 with
where Vcell is the mean membrane stack potential (V), I is the ap-
50 cell pairs made up of Neosepta cation-exchange membranes
plied current intensity (A), t is the operation t (h) and vtank is the vol-
(CIMS) and anion-exchange membranes (ACS) (1000 cm2 active sur-
ume of the concentrate tank (L).
face area per membrane). The stack dimensions were 620x450x313
b) In a continuous process, EC(NaCl) (kWh/kg NaCl), as the energy neces-
mm. The intermembrane distance was 0.43 mm, whereas linear flow
sary to generate the product overflow in the concentrated brine
velocity at the inlet of desalting and concentrating cells was around
tank, calculated according to Eq. (4) [43]:
10.8 cm/s.
The brine flow rate through the ion-exchange membranes stack was V cell  I
0.50 m3/h in both the concentrate and the diluate compartments, and EC ðNaClÞ ¼ ð4Þ
P NaCl overflow
0.15 m3/h in the electrolyte chambers. The diluate and electrolyte cir-
cuits had a single-pass design to operate at higher current densities
where PNaCl_ overflow is the concentrate brine production overflow (g
and minimise the problems of the increase of temperature in the cell
NaCl/h), and Vcell is defined as [45]:
and the precipitation of CaSO4 in the diluate compartments. The con-
centrate flow was recirculated until the maximum concentration was V cell ¼ ðr mem þ r dc þ r cc ÞI þ V mem ð5Þ
reached. Generated H2 accumulation in the catholyte chamber was
prevented by ventilation, whereas Cl2 formed in the anolyte chamber where rmem, rdc, and rcc describe the ion exchange membrane, the
was neutralised with sodium bisulphite. The volume of the feed diluate circuit and the concentrate circuit resistances, and Vmem is
(SWRO brine) and concentrate containers were 1 m3 and 0.25 m3, the average membrane potential [36]. Vcell was experimentally
respectively. measured.
Hydrochloric acid was added to keep the pH of the cathodic cir-
cuit below 3, below 7 and 5.5 in the diluate and the concentrate 3. Results and discussion
stream, respectively. The current densities were varied between
0.30 and 0.60 kA/m2. The inlet and outlet temperatures were moni- 3.1. Evaluation of the NaCl concentration.
tored during all the experiments. The brine concentration process was
monitored by in-line measurements of conductivity, temperature, pH, The evolution of the NaCl concentration results for different current
flow-rate, pressure, current intensity and voltage. densities (from 0.30kA/m2 to 0.6 kA/m2) at three different temperature
110 M. Reig et al. / Desalination 342 (2014) 107–117

ranges: a) T N 25 °C, b) 25 °C ≥ T ≥ 20 °C and c) T b 20 °C, are plotted on the intensity and temperature conditions. A gradient of NaCl concen-
in Fig. 1. A summary of the concentration factors achieved for the more tration appears and two mass transport phenomena occur: diffusion of
relevant experiments is given in Table 2. NaCl from the concentrate compartment to the diluate compartment
For the three temperature ranges, the concentration of NaCl in the (NaCl back diffusion) and osmotic flux. In this case, water is transported
concentrate tank increased gradually with time (from an initial 65 g from the diluate compartment to the concentrate one. These two mass
NaCl/L) until a plateau was reached after several hours at concentra- transfer phenomena diminish the desired objective of increasing the
tions between 210 mg/L and 265 mg/L, depending on the current den- NaCl concentration on the concentrate tank and increase the energy re-
sity applied: the higher the intensity applied, the higher the final quirements of the process. As seen in Fig. 2, for current intensities of
concentration of NaCl reached (Fig. 1). In the first step, an increase of 0.3kA/m2 (Fig. 2A) and 0.4kA/m2 (Fig. 2B) at the low temperature
the electrical potential is applied, favouring two main mass transport range (b20 °C), NaCl concentrations of 220 and 250 g/L, respectively,
phenomena: a) the transfer of charged species in the solution to the were achieved after 25 hours of operation, whereas only 170 and
cathode or anode through the ion-exchange membranes (ion migration 200 g/L were achieved at the high temperature range (N25 °C). Length-
flux) and b) the flux of water (as hydrated water to ions) due to ion mi- ening the experiment time did not provide any substantial increase in
gration, through the ion-exchange membranes (electro-osmosis flux). the NaCl concentration (increase differences less than approximately
However, the concentration gradient between the diluate and the con- 2 g/L) after 15 hours of additional concentration. Although the produc-
centrate circuits increased from 0 up to 90 to 160 g/L NaCl, depending tion of concentrate brines close to NaCl saturation could be desirable

Fig. 1. NaCl concentration vs. time at various current densities and different temperatures applied A) T N 25 °C, B) 25 °C ≥ T ≥ 20 °C, and C) T b 20 °C.
M. Reig et al. / Desalination 342 (2014) 107–117 111

Table 2
Composition of concentrated brines, concentration factors (CF) and time necessary to reach the maximum NaCl concentration for experiments at 30A (0.3 kA/m2) and 40A (0.4 kA/m2) at
14 °C and 10 °C, and at 35A (0.35 kA/m2), 50A (0.5 kA/m2) and 60A (0.6 kA/m2) at 28° inlet SWRO brine.

0.30 kA/m2 0.40 kA/m2 0.35 kA/m2 0.50 kA/m2 0.60 kA/m2
(14 °C) (10 °C) (28 °C) (28 °C) (28 °C)

g/L C0 Cf CF C0 Cf CF C0 Cf CF C0 Cf CF C0 Cf CF
NaCl 71.5 239 3.3 57.7 264 4.5 65 210 3.2 65 244 3.8 65.1 261 4
K+ 0.89 3.33 3.7 0.68 3.55 5.2 0.70 3.10 4.4 0.77 3.2 4.2 0.74 3.50 4.7
Mg2+ 2.5 1.3 0.5 2.5 1.2 0.5 2.2 1.1 0.5 2.3 1.3 0.6 2.3 1.2 0.5
Ca2+ 0.8 0.7 0.9 0.8 0.7 0.9 0.6 0.4 0.6 0.7 0.3 0.4 0.7 0.3 0.4
SO2−
4 4.7 1.4 0.3 4.9 1.0 0.2 5.4 1.7 0.3

mg/L
Al3+ 0.05 DL – 0.03 DL – 0.28 0.17 0.6 0.11 0.10 1 0.08 DL –
Ni2+ 0.06 0.14 2.3 0.05 0.12 2.4 0.06 0.11 1.8 0.06 0.20 3.3 0.06 0.12 2.0
Sr2+ 15.0 12.7 0.9 14.7 10.7 0.7 13.2 13.2 1.0 14.0 14.0 1.0 13.6 13.0 0.9
Cu2+ 0.02 0.03 1.5 0.03 0.04 1.3 0.02 0.2 10 0.03 0.2 6.6 0.02 0.13 6.5
Time (h) 35 38 15 14 13

DL: Detection limit.

(290 g/L) under the studied conditions (current densities and tempera- of these results is valuable for the identification of energy requirements
ture), those values could not be reached. An extra evaporation or the ad- ranges as a function of the brine source, although, there is a limitation
dition of a solid salt might be required, depending on the application, due to the scarce published data available to compare.
e.g., raw material for the chlor-alkali industry. The temperature influence on the concentration process in ED is
The concentration factors (CF) achieved were higher than those for more complex than that exerted by the current intensity because the
monovalent ions (Na+, K+ and Cl−). In Table 3, CF and the selectivity different mass transfer processes are temperature dependent. Most of
of the ACS and CIMS membranes operated at 0.3 kA/m2 with SWRO the published data are at laboratory scale and typically at room temper-
brines are compared with published data on brine concentrations ature (20–25 °C). Temperature influences the main transport processes
under similar concentration ranges. The most similar studies are the affecting the ionic species and water and also affects the properties of
use of CLS-25 T and AVS-4 T Neosepta membranes for salt production the brines (e.g., density and electrical conductivity), the properties of
at 0.2 kA/m2 [46], Selemion ASA and CMA membranes at 0.3 kA/m2 the solvent (diffusivity) and the membrane properties (electrical con-
[47], Selemion CMV and AMV [48] and Selemion CMT and AMT Ashahi ductivity and membrane resistance). Only Tanaka et al. [49–51] report-
Glass [33]. CFs were higher in the reviewed studies because the initial ed on the concentration of seawater with ED with similar membranes
concentrations were lower. CFs obtained by Turek et al. [48] were espe- and a similar tendency of the salt concentrations was identified. The ap-
cially high because the brine was treated using batch ED. The comparison proach developed by Tanaka for the concentration of seawater to

Fig. 2. NaCl concentration vs. time at various temperatures and different electric current applied A) 0.3kA/m2 and B) 0.4kA/m2.
112 M. Reig et al. / Desalination 342 (2014) 107–117

Table 3 produce salt water was derived of experimental equations from a high
Concentration factors (CF) and selectivity obtained at 0.3 kA/m2 and 14 °C with SWRO number of experimental assays of ion transport through the ion ex-
brines and CIMS/AMS membranes compared with the data from different previous expe-
riences reported in the literature.
change membranes in addition to the correction of temperature for
physical parameters, such as density, electrical conductivity and mem-
CIMS//ACS CLS-25 ASA/CMA CMV/AMV CMT/AMT brane resistance. A model describing the concentration process has
T/AVS-4 T [47] [48] [33,34]
been published recently [52,53]. Membrane pair characteristics of
[46]
three different IX membranes were measured as a function of the cur-
Feed inlet SWD-RO brine SW SW BWRO brine BWRO brine
rent density and seawater temperature allowed for the determination
Na+ 3.2 5.7 9.7 11.1 –
Ca2+ 0.8 7.6 6.3 10.4 1.1 of the hydraulic permeabilities and the cell electric resistances by
Mg2+ 0.5 6.0 2.7 10.8 1.8 using the same empirical functions of temperature. Temperature influ-
Cl− 3.2 5.7 11.5 11.1 1.4 ence on the performance of the electrodialyser to increase the NaCl con-
SO2−
4 0.3 0.6 0.2 10.3 1.3 centration in the brine and the energy consumption implies an increase
S(Na/Ca) −1.2 0.2 −0.2 0.0 –
of consumption with a decrease of temperature. A decrease of the tem-
S(Na/Mg) −1.6 0.0 −0.7 0.0 –
S(Cl/SO4) −1.9 −1.2 −1.2 0.0 −0.1 perature from 60 °C to 25 °C is accompanied by an increase of the

Fig. 3. Divalent ion concentration vs. time at different electric current applied. A) Ca2+ high temperature, B) Mg2+ high temperature, and C)SO2−
4 low temperature.
M. Reig et al. / Desalination 342 (2014) 107–117 113

energy consumption up to 20%; however, the increase of temperature minor components (Ca2 +, Mg2 + and SO24 −), the concentrations de-
limits the concentration factor, which could be reduced by up to 30%. creased in the concentration tank, reaching a CF close to or below 1 as
The direct current electric resistance of a membrane pair is calculat- a consequence of the concentration reduction from the initial values
ed, and it is predominant over that of a desalting cell and a concentrat- of the SWRO brines. The sulphate, calcium and magnesium concentra-
ing cell. It is necessary to decrease the electric resistance of an ion- tions in the diluate were similar to the inlet brine composition, given
exchange membrane for reducing energy consumption in a salt that the circuit had a single-pass design. This confirmed the fact that
manufacturing process. multi-charged ionic species (Ca2+, Mg2+ and SO2− 4 ) in the concentrate
tank were mainly diluted by water transport (osmotic water ion flux)
inside the cell and that the migration of these ions (ion flux) was not sig-
3.2. Evaluation of the behaviour of minor and trace species
nificant because the membranes are selective to monovalent species.
For this reason, multivalent ions concentration plunged and as the
The ion exchange membranes used in this study (ACS and CIMS)
time increases, a reduction of the slope of this function is observed;
proved to be highly selective to mono-charged ions (e.g., Na+, K+ and
obtaining a value close to zero, and the diffusion process, back diffusion
Cl−), as the selectivity factor was the highest in all cases reviewed.
and osmotic flux are the dominant phenomena.
The influence of the brine composition on the speciation of the main
The measured values indicated that CF was lower for sulphate (0.2–
components of the SWRO brine (Cl−, Na+ and K+) as a function of pH
0.3) than for Ca (0.4–0.9) and Mg (0.5–0.6) ions. In addition to the influ-
was determined with the Hydra and Medusa codes [54]. The results in-
ence of the dilution effect due to the electro-osmotic flow, other factors
dicated that the main species present in the brine solution were Cl− (I)
are affecting ions selectivity with the ion exchange membranes or the
as Cl− (N 98%); Na(I) as Na+ (90%), NaClaq (8%) and NaSO− 4 (2%); and
potential effect of the ions speciation. Literature review shows scarce
K(I) as K+ (87%), KCl(9%) and KSO−4 (4%).
data on the ion exchange selectivity coefficients for chloride/sulphate
The evolution of the concentration of Ca2 +, Mg2 + and SO24 − as a
with ACS and Na/Ca and Na/Mg with CIMS. Only CFs could be found on
function of time for different current intensities is shown in Fig. 3. For
the application of ED for the concentration of seawater or on the

Cu(OH) 2
1.0
1.0 A Mg(OH) 2 (c)
A
0.8
Mg 2+ 0.8

Cu 2+
Fraction

0.6
Fraction

0.6

0.4 0.4

CuCl +
0.2 0.2
Cu(OH) 4 2-
MgSO CuSO 4
MgCl +
4
CuCl 2 Cu(OH) 3-

0.0 0.0
2 4 6 8 10 12 2 4 6 8 10 12
pH
pH
Ni(OH) 2 (c)
1.0
1.0
B B
Ca 2+
0.8 0.8
Fraction

0.6
Fraction

0.6
Ni 2+

0.4 0.4
NiCl 2

0.2 0.2 NiCl +


CaSO4 NiSO 4 Ni(OH) 3-
CaCl +
0.0 0.0
2 4 6 8 10 12 2 4 6 8 10 12
pH pH

1.0 1.0
C SO 4 2- C
Sr 2+
0.8 0.8
Fraction

0.6
Fraction

0.6

0.4 0.4

MgSO 4
0.2 0.2
SrSO 4
HSO4-
SrCl +
CaSO 4
0.0 0.0
2 4 6 8 10 12 2 4 6 8 10 12
pH pH

Fig. 4. Speciation of Mg2+, Ca2+ and SO2−


4 (fraction percentage) in the SWRO brine as a Fig. 5. Speciation of Cu2+, Ni2+ and Sr2+ (fraction percentage) in the SWRO brine as a
function of the pH. Calculations performed with the Hydra and Medusa codes [54]. function of the pH. Calculations performed with the Hydra and Medusa codes [54].
114 M. Reig et al. / Desalination 342 (2014) 107–117

treatment of seawater RO desalination brines, where pre-concentrations higher water flux in the stack, reaching approximately 50% of the trans-
factors were from 0.3 to 10 for SO2−4 , from 0.8 to 10 for Ca
2+
and from 0.5 port. The current efficiency of the chloride ion was improved after
2+
to 11 for Mg [34,49–51]. A second effect to be taken into account to ex- cleanings of the membranes, reaching more than 92% of the current
plain the differences in behaviour between the three ions is an indication transport for the lower inlet temperatures.
that the selectivity of the cation and the anion membranes for Ca2+, In Table 4, the brine concentration obtained at the end of the exper-
Mg2+ and SO2− 4 could be influenced by the properties of the involved iments is compared to the membrane electrolysis requirements in the
species in solution. The speciation diagrams as a function of the pH of chlorine production industry. The brine needs to have higher NaCl con-
Ca2+, Mg2+ and SO2− 4 in the treated brines are plotted in Fig. 4. In the centration and impurities content must be lower than the achieved
pH working range of 5 to 7, the high concentration values of chloride values to avoid precipitation on the membrane surface or to increase
(1.1 M) is promoting the formation of Ca and Mg chloride complexes the efficiency of the process. These requirements are specified in the
(CaCl+ and MgCl+), and the strong complexing properties of sulphate last column of Table 4. The elements that do not meet the electrolysis re-
ions with Ca2+ and Mg2+ promote the formation of the non-charged quirements are Ca, Mg, Sr, Cu and Ni. In all cases, the brine should be fur-
complexes (MgSO4(aq) and CaSO4(aq)). In the treated brines, the three ther concentrated to reach the saturation limit. This could be achieved
species are present in solution as mixtures of bi-charged, mono- by adding a small amount of solid salt (between 90 and 40 g/L NaCl).
charged and non-charged species with the following distribution (%): Moreover, the brine would also need a purification step to remove cal-
Mg: Mg2+ (80%), MgSO4aq (12%) and MgCl+ (8%); Ca: Ca2+ (82%), cium, magnesium, copper, nickel and strontium in all cases studied.
CaSO4aq (10%) and CaCl+ (8%); and sulphate: SO24 − (72%), MgSO4aq After ten months of operation, the stack was opened to analyse the
(22%) and CaSO4aq (6%). The lower concentration factors for sulphate membrane integrity. The membranes were examined by SEM-EDS,
demonstrate the higher selectivity of the anion exchange membrane and no precipitates were observed on their surfaces. The absence of pre-
for chloride than sulphate, with the transport potentially enhanced due cipitates of magnesium, calcium and carbonates supports the deduction
to the presence of two non-charged species that suffer from electrical ex- provided in the analysis of the brine saturation indexes. The initial inlet
clusion on the anion exchange membranes. In the case of the Ca and Mg brines were oversaturated with calcium carbonate, magnesium carbon-
species, the selectivity of the cation exchange membranes for Na over Ca ate and calcium magnesium carbonate. Additionally, the evolution of
and Mg is lower than the selectivity value for chloride/sulphate, and an the saturation index for NaCl, KCl, MgCl2•6H2O, CaSO4•2H2O, SrSO4,
additional 8% of the Ca and Mg species are in the form of monovalent CaCO3, MgCO3 and CaMg(CO3)2 as a function of the concentration pro-
species and could more easily be transported through the cation ex- cess indicates that the brine could be concentrated up to 4 times (ap-
change membranes. The cation concentration order of Ca(II) N Mg(II) proximately 280 g/L NaCl) without risk of NaCl precipitation. Because
is in agreement with the fact that the hydrated ionic radii of Ca2+ is the brines contained the antiscalant from the RO process, precipitation
lower than for Mg2+ (stokes radius of 0.349 over 0.429 nm), and a of carbonate and sulphate salts did not occur easily, despite their over-
slightly higher CF for calcium was obtained compared to magnesium. saturation, thus benefiting the ED process. The antiscalants used in the
However, other minor components present in the brines, (Ni(II) and RO process (1–2.5 mg/L of PC-1020) were mainly super-threshold
Cu(II)), did not follow the general trend; their concentrations increased agents that were able to stabilise supersaturated salt solutions to pre-
in all the experiments. Fig. 5 shows that in the pH range studied (5–7), vent precipitation of carbonate and sulphate salts. Moreover, HCl was
the formation of mono-charged species of nickel and copper are NiCl+ added at different points of the concentrate to maintain a constant pH
and CuCl+, and their transport through the membrane is enhanced. of 5.5 and to ensure that most of the inorganic carbon was present as bi-
The concentration factors of copper are 2 to 3 times higher than the con- carbonate, which does not cause scaling. At the end of the two years of
centration factors of nickel, depending on the temperature and the cur- the research evaluation programme, the stack was open again and a
rent density. The speciation diagrams of Fig. 5 show that the fraction of membrane autopsy was completed. In addition to the visual inspection
CuCl+ in the brine is up to 2 times greater than the fraction of NiCl+. Con- of the anion and cation exchange membranes, those identified in this
sequently, it can obtain higher concentration factors because more cop- step as potentially more affected were analysed by SEM-EDAX, FTIR
per is available to be transported. Contrary to the case of Sr(II), CF and XPS. The SEM-EDAX analysis showed the scattered presence of
between 0.7 and 1 are described for Ca(II), along with similar complexa- some precipitates, in most cases, crystalline forms containing typical
tion behaviour with the formation of SrCl+. For the trace metals present (O and Si), and also the presence of non-crystalline precipitates rich in
in the brines, the concentration order was Cu (II) N Ni (II) N Sr (II). O, Fe, Al and in minor cases, Mn. Samples were also analysed by XRD,
Diluate production from the cell was constant in all experiments at and no crystalline compounds were detected, or their size was so
approximately 50–52 g NaCl/L, which means that only 6–8 g/L NaCl small that there was no diffraction. The membrane surfaces were
was removed in each pass, depending on the current density applied. analysed, and the main elements detected in the cationic membrane
The current efficiency showed that chloride transported more than were C, O, S, Na, Cl, and K, and the main elements detected in the anionic
80% of the current in the stack for the experiments at low temperature. membrane were C,O, N and Cl. The FTIR analysis did not detect the pres-
When the inlet temperatures were high, this value was reduced due to a ence of mineral forms of carbon, such as Ca and Mg carbonates, and the

Table 4
Comparison of concentrated brines obtained with ED and membrane cell requirements.

0.30 kA/m2 0.40 kA/m2 0.35 kA/m2 0.50 kA/m2 0.60 kA/m2 Membrane cell
(14 °C) (10 °C) (28 °C) (28 °C) (28 °C) requirements

g/L
NaCl 239 264 210 244 261 300
K+ 3.33 3.55 3.10 3.20 3.50 –
Ca2+ 0.6 0.7 0.4 0.3 0.3 20 μg/L
Mg2+ 1.3 1.2 1.1 1.3 1.2
SO2−
4 1.4 1.0 1.7 8

mg/L
Al3+ DL DL 0.17 0.1 DL 0.1
Ni2+ 0.14 0.12 0.11 0.2 0.12 0.01
Sr2+ 12.7 10.7 13.2 14.0 13.0 0.4
Cu2+ 0.03 0.04 0.2 0.2 0.1 0.01
M. Reig et al. / Desalination 342 (2014) 107–117 115

Fig. 6. Energy consumption working under batch configuration as a function of NaCl concentration in the tank and current density applied for different inlet temperatures.

peak assignments were in agreement with the main functional groups the electrode potentials are not considered in industrial cells because
of the anionic and cationic ion exchange resins. Finally, the results of the voltage produced is low compared to that produced by the total
the XPS analysis of both cationic and anionic membranes detected C, number of cells, for pilot plants, the electrode potential significantly in-
O, S, Na and Cl for the cationic membranes and C, O, N, Na and Cl in creases the energy consumption and must be taken into account.
the anionic membranes. Only Si was detected as a foreign element in The energy consumption sharply increased from 0.5 to 0.6 kA/m2
both membranes, and Ca and Mg were not detected. (28 °C), thus, 0.5 kA/m2 was considered the maximum operable point
at higher inlet temperatures. The temperature affects the product flow
3.3. Energy consumption on the concentration of NaCl (ENaCl) from seawater rate, and lower flow rates were obtained at higher temperatures.
reverse osmosis desalination brines A comparison of the performance of an industrial electrodialyser to
obtain solid salt from seawater and the results obtained in the pilot
The energy consumption increases needed to increase the NaCl con- plant in Barcelona in continuous operation and batch mode can be
centration in the tank (batch operation (EB(NaCl))) over the concentra- found in Table 5. The values obtained in the pilot plant were within the
tion reached at different current densities are shown in Fig. 6. EB(NaCl) range of the industrial process in all cases studied. Higher concentrations
rises gradually in all the experiments until the concentration in the were reached in the pilot plant with slightly higher current densities and
tank reaches approximately 210 g/L NaCl. The energy consumption in- energy consumptions. For laboratory scale studies [55], the energy con-
creased dramatically because osmosis and ion diffusion fluxes inside sumption with NaCl solutions up to approximately 100 g/L ranged be-
the stack reach a maximum due to the concentration gradient between tween 0.18 and 0.33 kWh/kg with 80% solute recovery and current
the diluate and the concentrate compartments. The NaCl concentration densities of 0.20 to 0.30 kA/m2 were reported. Furthermore, Turek et al.
in the tank increased very slowly from this point. Then, the energy [56] worked at laboratory scale obtaining more than 300 g/L NaCl in
applied was used to slowly increase the NaCl concentration of the the concentrate stream with a current density of 0.6kA/m2, nevertheless,
tank, and the specific energy consumption increased. If the process the energy consumption reported was 0.28 kWh/kg NaCl transported.
needs to be operated in a batch mode, the optimum energy consump- Technical targets when operating electrodialyzers to produce solid salt
tion concentration would be obtained approximately 200 g/L NaCl at from seawater in Japan are obtaining a concentration higher than 200 g/L
0.2–0.3 kWh/kg NaCl in the range of 0.3–0.4 kA/m2, depending on the NaCl with energy consumption lower than 0.12 kWh/kg NaCl [57–59].
current density applied and the feed inlet temperature. This value was nearly reached 0.12 kWh/kgNaCl when operating the
A summary of the energy consumption for continuous operation (E- plant in a continuous mode at 27 °C at 0.35 kA/m2. Also, for 203 g/L
C(NaCl)) for different operation conditions is shown in Table 5. Although NaCl at 27 °C with and at 0.50 kA/m2 an ENaCl of 0.19 kWh/kg was

Table 5
Comparison of the performance of an industrial electrodialyzer using seawater in Japan and the pilot plant using SWRO brine in Barcelona.

Current density Temperature Energy consumption NaCl concentration


(kA/m2) (°C) (kWh/kg NaCl) (g NaCl/L)

Industrial cell [45] 0.27 23.5 0.160 174


Batch mode 0.30 14.0 0.20 199
0.40 10.0 0.30 209
0.35 28.0 0.29 208
0.50 28.0 0.35 186
0.60 28.0 0.38 208
Continuous mode 0.30 16.0 0.16 176
0.30 18.0 0.15 176
0.35 27.0 0.12 185
0.40 17.0 0.20 178
0.45 18.0 0.26 219
0.50 20.0 0.26 246
0.50 27.0 0.19 203
0.60 20.0 0.30 245
0.60 27.0 0.24 244
116 M. Reig et al. / Desalination 342 (2014) 107–117

Fig. 7. Relationship between the energy consumption working under continuous configuration and the current density for experiments at low and high temperature range for the IXM-ED.

obtained. However, in order to enhance the performance of the ED, it is temperature and current densities used. At higher inlet temperatures
expected to increase the NaCl in the brine and decrease the energy con- lower concentrations were obtained, but higher production flows and
sumption necessary to produce one tonne of NaCl (ENaCl (kWh/KgNaCl)). lower energy consumptions could be reached. At higher current densi-
The actual technical target for strengthening the competitive position of ties applied, higher production flows and concentrations could be ob-
an electrodialysis system in the salt market was discussed by Tanaka tained but higher energy consumptions were recorded. ED pilot plant
[60] to be CNaCl N 200 g NaCl/L and ENaCl b 120 (kWh/t NaCl). Fig. 7 was designed in one single pass for diluate, which allowed to work
shows the relationship between CNaCl and EC(NaCl) obtained in the with higher current densities and to get higher concentrations. ED in-
electrodyalisis experiments performed under the different conditions of trinsically purified polyvalent ions, which were diluted due to osmosis
intensity and temperature, and Fig. 8 compares the EC(NaCl) data to achieve and electro-osmosis phenomena. Nickel and copper were an exception
a desired CNaCl showing that as the target indicates a reduction in energy and were concentrated in the brine produced. In the operating pH, these
consumption is desirable. EC(NaCl) depends on the membrane cell poten- ions were in the form of univalent compounds and so, could pass
tial (Vcell) which includes the ohmic term (rmem, rdc and rcc) and mem- through the membranes. The concentration order of divalent and triva-
brane potential term (see Eq. (5)). As for the electric resistances in the lent ions was highly dependent on electric charge and ionic radii. Poly-
ohmic term (rmem, rdc and rcc) the membrane resistance (rmem) is pre- valent ions with lower electric charge got more concentrated in all the
dominant over rdc and rcc [52] so reduction of energy consumption is as- experiments. Migration of these ions was not a primary phenomenon
sociated to the reduction of the membrane resistance. However, this and their final composition was mainly determined by the osmosis
situation is accompanied by a decrease in CNaCl, because of the increase and electro-osmosis fluxes.
in density. Then rationale proposed [52] for the reduction of the mem- The optimal ED operation point was determined when working in a
brane electric resistance by diminishing the membrane thickness is the continuous mode at 0.35 kA/m2 of 185 g/L NaCl with 0.12 kWh/kg NaCl
development of a double-layered membrane consisting of a fine porous energy consumption were obtained. This energy consumption values
thinner functional layer and porous reinforced layer. were in the range of the industrial cells used in Japan to produce solid
salt from seawater although higher concentrations can be obtained
4. Conclusions with SWRO brine. Further optimization will be needed to reach the
technical target to get brines with NaCl contents higher than 200 g/L
Electrodialysis has proved to be a feasible technology for SWRO with less than 0.12 kWh/kg. As can be seen, the energy consumption
brine concentration. ED performance was highly dependent on inlet values obtained were in the range of those reported in the literature,

Fig. 8. Relationship between the NaCl concentration and the continuous energy consumption necessary to produce one tonne of NaCl.
M. Reig et al. / Desalination 342 (2014) 107–117 117

but more research is needed to obtain data on energy consumption [29] P. Chelme-Ayala, D.W. Smith, M.G. El-Din, Membrane concentrate management op-
tions: a comprehensive critical review, Can. J. Civ. Eng. 36 (2009) 1107–1119.
under different operating conditions so that the energy consumption [30] M. Ahmed, W.H. Shayya, D. Hoey, A. Mahendran, R. Morris, J. Al-Handaly, Use of
of the pilot plant can be reduced and optimal NaCl concentrations can evaporation ponds for brine disposal in desalination plants, Desalination 130
be reached. (2000) 155–168.
[31] J. Gilron, Y. Folkman, R. Savliev, M. Waisman, O. Kedem, WAIV — wind aided inten-
sified evaporation for reduction of desalination brine volume, Desalination 158
References (2003) 205–214.
[32] Y. Oren, E. Korngold, N. Daltrophe, R. Messalem, Y. Volkman, L. Aronov, M.
[1] L. Malaeb, G.M. Ayoub, Reverse osmosis technology for water treatment: State of the Weismann, J. Gilron, Pilot studies on high recovery BWRO-EDR for near zero liquid
art review, Desalination 267 (2011) 1–8. discharge approach, Desalination 261 (2010) 321–330.
[2] L.F. Greenlee, D.F. Lawler, B.D. Freeman, B. Marrot, P. Moulin, Reverse osmosis desa- [33] E. Korngold, L. Aronov, N. Daltrophe, Electrodialysis of brine solutions discharged
lination: Water sources, technology, and today's challenges, Water Res. 43 (2009) from an RO plant, Desalination 242 (2009) 215–227.
2317–2348. [34] E. Korngold, L. Aronov, N. Belayev, K. Kock, Electrodialysis with brine solutions
[3] C. Fritzmann, T. Löwenberg, T. Wintgens, T. Mellin, State-of-the-art of reverse osmo- oversaturated with calcium sulfate, Desalination 172 (2005) 63–75.
sis desalination, Desalination 216 (2007) 1–76. [35] S. Ishizaka, Seawater desalination and by-product utilization, Soda Chlorine 20
[4] G. Mauguin, P. Corsin, Concentrate and other waste disposals from SWRO plants: (1969) 1.
characterization and reduction of their environmental impact, Desalination 182 [36] Y. Tanaka, Ion-exchange membrane electrodialysis program and its application
(2005) 355–364. to multi-stage continuous saline water desalination, Desalination 301 (2012)
[5] D.A. Roberts, E.L. Johnston, N.A. Knott, Impacts of desalination plant discharges on 10–25.
the marine environment: A critical review of published studies, Water Res. 44 [37] M. Badruzzaman, J. Oppenheimer, S. Adham, M. Kumar, Innovative beneficial reuse
(2010) 5117–5128. of reverse osmosis concentrate using bipolar membrane electrodialysis and
[6] R. Einav, F. Lokiec, Environmental aspects of a desalination plant in Ashkelon, Desa- electrochlorination processes, J. Membr. Sci. 326 (2009) 392–399.
lination 156 (1–3) (2003) 79–85. [38] N. Melián-Martel, J.J. Sadhwani, S. Ovidio Pérez Báez, Saline waste disposal reuse for
[7] A. Loya-Fernández, L.M. Ferrero-Vicente, C. Marco-Méndez, E. Martínez-García, J. desalination plants for the chlor-alkali industry. The particular case of pozo
Zubcoff, J.L. Sánchez-Lizaso, Comparing four mixing zone models with brine dis- izquierdo SWRO desalination plant, Desalination 281 (2011) 35–41.
charge measurements from a reverse osmosis desalination plant in Spain, Desalina- [39] M. El-Naas, O. Al-Marzouqi, Chaalal, A combined approach for the management of
tion 286 (2012) 217–224. desalination reject brine and capture of CO2, Desalination 251 (2010) 70–74.
[8] M. Ahmad, P. Williams, Assessment of desalination technologies for high saline [40] Y. Tanaka, R. Ehara, S. Itoi, T. Goto, Ion-exchange membrane electrodialytic salt pro-
brine applications e discussion paper, Desalin. Water Treat. 30 (2011) 22–36. duction using brine discharged from a reverse osmosis seawater desalination plant,
[9] P. Bond, S. Veerapaneni, Zeroing in on ZLD technologies for inland desalination, J. J. Membr. Sci. 222 (2003) 71–86.
Am. Water Works Assoc. 100 (2008) 76–89. [41] M.D. Eisaman, K. Parajuly, A.V. Tuganov, C. Eldershaw, N. Chang, K.A. Littau, CO2 ex-
[10] C.J. Gabelich, P. Xu, Y. Cohen, Chapter 10 Concentrate treatment for inland desalting, traction from seawater using bipolar membrane electrodialysis, Energy Environ. Sci.
Sustain. Sci. Eng. 2 (2010) 295–326. 5 (2012) 7346–7352.
[11] M.C. Mickley, Membrane Concentrate Disposal: Practices and Regulation. Report [42] D.L. Parkhurst, PHREEQC: a computer program for speciation, reaction-path, advec-
No.123. Bureau of Reclamation, Denver, CO, Water Treatment Engineering and Re- tive transport, and inverse geochemical calculations, U.S. Geological Survey Water
search Group; Mickley and Associates, Boulder, CO (2006). Resources, Investigation Report, 1995, pp. 95–422.
[12] M. Ahmed, A. Arakel, D. Hoey, M.R. Thumarukudy, M.F.A. Goosen, M. Al-Haddabi, A. [43] S. Casas, C. Aladjem, J.L. Cortina, E. Larrotcha, L.V. Cremades, Seawater reverse osmo-
Al-Belushi, Feasibility of salt production from inland RO desalination plant reject sis brines as a new salt source for the chlor-alkali industry: Integration of NaCl con-
brine: a case study, Desalination 158 (2003) 109–117. centration by electrodialysis, Solvent Extr. Ion Exch. 30 (2012) 322–332.
[13] F. Mohammadesmaeili, M.K. Badr, M. Abbaszadegan, P. Fox, Mineral recovery from [44] Y. Zhang, K. Ghyselbrecht, B. Meesschaert, L. Pinoy, B. Van der Bruggen, Electrodial-
inland reverse osmosis concentrate using isothermal evaporation, Water Res. 44 ysis on RO concentrate to improve water recovery in wastewater reclamation, J.
(2010) 6021–6030. Membr. Sci. 378 (2011) 101–110.
[14] N. Voutchkov, Overview of seawater concentrate disposal alternatives, Desalination [45] Y. Tanaka, Development of a computer simulation program of batch ion-exchange
273 (2011) 205–219. membrane electrodialysis for saline water desalination, Desalination 320 (2013)
[15] I. Muñoz, A.R. Fernández-Alba, Reducing the environmental impacts of reverse os- 118–133.
mosis desalination by using brackish groundwater resources, Water Res. 42 [46] R. Yamane, M. Ichikawa, Y. Mituzani, Y. Onoue, Concentrated brine production from
(2008) 801–811. seawater by electrodialysis using ion exchange membranes, Industrial and Engi-
[16] J.L. Sánchez-Lizaso, J. Romero, J. Ruiz, E. Gacia, J.L. Buceta, O. Invers, Y. Fernández neering Chemistry Process Design andDevelopment 8 (1969) 21–26.
Torquemada, A. Mas, A. Ruiz-Mateo, M. Manzanera, Salinity tolerance of the Medi- [47] W. Blancke, Unita di Recerca Electtrolisi Rosignano, Production of NaCl concentrated
terranean seagrass Posidonia oceánica: recommendations to minimize the impact brine from seawater for chlor-alkali membrane electrolysis: exploitation of saturat-
of brine discharges from desalination plants, Desalination 221 (2008) 602–607. ed super purified seawater in a chlor-alkali micro pilot cell, Solvay Internal Report,
[17] S. Lattemann, T. Hoepner, Environmental impact and impact assessment of seawater 2007.
desalination, Desalination 220 (2008) 1–15. [48] M. Turek, J. Was, P. Dydo, Brackish water desalination in RO–single pass EDR system,
[18] G.L. Meerganz von Medeazza, Direct and socially-induced environmental impacts of Desalin. Water Treat. 7 (2009) 263–266.
desalination, Desalination 185 (2005) 57–70. [49] Y. Tanaka, A computer simulation of batch ion exchange membrane electrodialysis
[19] F. Hajbi, H. Hammi, A. M'Nif, Reuse of RO desalination plant reject brine, J. Phase for desalination of saline water, Desalination 249 (2009) 1039–1047.
Equilib. Diffus. 31 (2010) 341–347. [50] Y. Tanaka, Simulation of an ion exchange membrane electrodialysis process for con-
[20] A. Pérez-González, A.M. Urtiaga, R. Ibáñez, I. Ortiz, State of the art and review on the tinuous saline water desalination, Desalin. Water Treat. 22 (2010) 1–15.
treatment technologies of water reverse osmosis concentrates, Water Res. 46 [51] Y. Tanaka, A computer simulation of ion exchange membrane electrodialysis for
(2012) 267–283. concentration of seawater, Membr. Water Treat. 1 (2010) 13–37.
[21] D.H. Kim, A review of desalting process techniques and economic analysis of the re- [52] Y. Tanaka, Ion-exchange membrane electrodialysis for saline water desalination and
covery of salts from retentates, Desalination 270 (2011) 1–8. its application to seawater concentration, Ind. Eng. Chem. Res. 50 (2011)
[22] O. Gibert, C. Valderrama, M. Peterkóva, J.L. Cortina, Evaluation of selective sorbents 7494–7503.
for the extraction of valuable metal ions (Cs, Rb, Li, U) from reverse osmosis rejected [53] Y. Tanaka, Ion-exchange membrane electrodialysis of saline water and its numerical
brine, Solvent Extr. Ion Exch. 28 (2010) 543–562. analysis, Ind. Eng. Chem. Res. 50 (2011) 10765–10777.
[23] J. Le Dirach, S. Nisan, C. Poletiko, Extraction of strategic materials from the concen- [54] I. Puigdomènech, Chemical equilibrium software Hydra and Medusa, Inorganic
trated brine rejected by integrated nuclear desalination systems, Desalination 182 725 Chemistry Department, Royal Institute of Technology, Stockholm, Sweden,
(2005) 449–460. 2001.
[24] T. Jeppessen, L. Shu, G. Keir, V. Jegatheesan, Metal recovery from reverse osmosis [55] M. Fidaleo, M. Moresi, Optimal strategy to model the electrodialytic recovery of a
concentrate, J. Clean. Prod. 17 (2009) 703–707. strong electrolyte, J. Membr. Sci. 260 (2005) 90–111.
[25] A. Neilly, V. Jegatheesan, L. Shu, Evaluating the potential for zero discharge from re- [56] M. Turek, K. Mitko, M. Chorążewska, P. Dydo, Use of the desalination brines in the
verse osmosis desalination using integrated processes — a review, Desalin. Water saturation of membrane electrolysis feed. Desalination and water treatment, Desali-
Treat. 11 (2009) 58–65. nation 51 (2013) 2749–2754.
[26] C.J. Gabelich, A. Rahardianto, C.R. Northrup, T.I. Yun, Y. Cohen, Process evaluation of [57] T. Fujita, Current challenges of salt production technology, Bull. Soc. Sea Water Sci.
intermediate chemical demineralization for water recovery enhancement in Jpn. 63 (2009) 15–20.
production-scale brackish water desalting, Desalination 272 (2011) 36–45. [58] Y. Tanaka, A computer simulation of feed and bleed ion exchange membrane elec-
[27] M. Turek, Seawater desalination and salt production in a hybrid membrane thermal trodialysis for desalination of saline water, Desalination 254 (2010) 99–107.
process, Desalination 153 (2002) 173–177. [59] Y. Tanaka, A computer simulation of continuous ion exchange membrane, electrodi-
[28] X. Ji, E. Curcio, S. Al Obaidani, G. Di Profio, E. Fontananova, E. Drioli, Membrane alysis for desalination of saline water, Desalination 249 (2009) 809–821.
distillation-crystallization of seawater reverse osmosis brines, Sep. Purif. Technol. [60] Y. Tanaka, Ion-exchange membrane electrodialysis program and its application to
71 (2010) 76–82. multi-stage continuous saline water desalination, Desalination 301 (2012) 1–25.

You might also like