You are on page 1of 7

Desalination 362 (2015) 5258

Contents lists available at ScienceDirect

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

Two-stage sulfate removal from reject brine in inland desalination with


zero-liquid discharge
Dema Almasri a,, Khaled A. Mahmoud a, Ahmed Abdel-Wahab b,
a
b

Qatar Environment and Energy Research Institute, Qatar Foundation, Doha 5825, Qatar
Chemical Engineering Program, Texas A&M University at Qatar, Education City, Doha 23874, Qatar

H I G H L I G H T S

Two-stage process for SO4 removal from NF-reject brine


Successful removal at reasonable lime and aluminum ratios
Effective removal at a high pH range
The experimental model developed accurately predicted experimental results.

a r t i c l e

i n f o

Article history:
Received 15 September 2014
Received in revised form 15 January 2015
Accepted 2 February 2015
Available online 9 February 2015
Keywords:
Sulfate
Ettringite
Nanoltration
Reject brine
Zero liquid discharge
Ultra high lime with aluminum

a b s t r a c t
This research proposes a two-stage process for the removal of sulfate from a nano-ltration (NF) reject stream. In
Stage 1, calcium chloride is used to partially remove the majority of sulfate as calcium sulfate precipitate while
the remaining sulfate is removed in Stage 2 as calcium sulfoaluminate precipitate. Kinetics of the removal process
in both stages was investigated and the results indicated that the removal kinetics is not a limiting factor for the
practicability of the process. Equilibrium characteristics of sulfate precipitation were evaluated and optimum
conditions for maximum sulfate removal were obtained. Sulfate removal in Stage 1 was independent of the pH
while efcient sulfate removal in Stage 2 was found to be above pH 11. Efcient sulfate removal was observed
at practical ranges of lime and sodium aluminate doses. An equilibrium model was developed to describe the
chemical behavior in Stage 2. A valid explanation for the mechanism of sulfate removal in Stage 2 was the formation of a solid mixture consisting of ettringite and monosulfate as conrmed by X-ray diffraction (XRD) analysis.
This approach is expected to reduce the volumes of the overall precipitated solids and in turn improve the overall
efciency of the ZLD treatment system.
2015 Elsevier B.V. All rights reserved.

1. Introduction
The deterioration of groundwater quality due to seawater intrusion
and other human activities in the Gulf region and many arid countries
has led to the use of inland desalination techniques for the production
of good quality water. Clearly, any desalination process produces two
streams; a clean water product stream and a reject concentrate stream
(called brine) that must be disposed of. Reject brine disposal into the
sea is not an option for inland desalination. Improper disposal of reject
brine from inland plants results in the following problems [1]: (1) Pollution of groundwater resources that are used as feed water for desalination plants due to high salt content and the presence of other harmful
chemicals in the concentrate; (2) production decline in agricultural
lands caused by the deposition of airborne salts from dried concentrate;
Corresponding authors.
E-mail addresses: delmasri@qf.org.qa (D. Almasri),
ahmed.abdel-wahab@qatar.tamu.edu (A. Abdel-Wahab).

http://dx.doi.org/10.1016/j.desal.2015.02.008
0011-9164/ 2015 Elsevier B.V. All rights reserved.

(3) formation of eyesores caused by the improper disposal of concentrates on nearby land; and (4) rendering treated municipal sewage efuent unsuitable for agriculture due to high total dissolved solids
(TDS). Environmental issues caused by the improper disposal of reject
brine pose major challenges for inland desalination and urge the need
for zero liquid discharge (ZLD) [2,3].
Most ZLD systems in operation today use thermal processes, evaporation ponds, or a combination of these methods [4]. While thermal
evaporation is a proven process, it is energy-intensive and suffers
from high capital and operating costs. On the other hand, evaporation
ponds typically require large areas of land, and their construction
costs are high. Furthermore, water evaporated from a pond is a lost
resource.
Given the need for ZLD and the drawbacks of existing ZLD methods,
it is imperative to nd alternative ZLD treatment technologies that provide more affordable concentrate management. This can be achieved
through the development of inexpensive treatment techniques that
maximize water recovery while minimizing the volume of concentrated

D. Almasri et al. / Desalination 362 (2015) 5258

53

Fig. 1. Schematic diagram of the two-stage RO with intermediate brine treatment.

brine that needs to be vaporized. When brackish groundwater is desalinated using reverse osmosis (RO), the constituents of the water are concentrated in the reject brine by a factor of 310 times that of raw water,
causing membrane fouling depending on the desalinated water recovery [5,6]. Sulfate salts, such as barium sulfate or calcium sulfate, form a
scale on the membranes during water recovery. Although the calcium
sulfate scale is more soluble than other mineral scales such as CaCO3,
BaSO4, and SrSO4, once formed, it is difcult to eradicate mechanically
and is insoluble in mineral acids and other conventional solvents
[79]. Sulfate is present in high concentrations (~ 700 mg/L as SO24 )
in brackish groundwater in Qatar, resulting in limited water recovery
in desalination systems.
Several methods were proposed for sulfate removal from water including chemical treatment by precipitation [1012], membrane separation using nanoltration [1315], ion-exchange and/or adsorption
[1618] and biological treatment [11,18] or a combination of these processes. However, these methods suffer from limited removal efciency
and/or high cost. In addition, ion exchange and membrane separation
involve the generation of a liquid waste stream that would require
proper management.
Sulfate can also be removed by precipitation with calcium and aluminum as calcium sulfoaluminate which has a low solubility product.

Important evidence to consider is the performance of sulfate in the


chemistry of Portland cement. The stability of calcium sulfoaluminate,
commonly known as ettringite (Ca6Al2(SO4)3(OH)12), at equilibrium relies on adequate activities of calcium, aluminum, and sulfate [19]. It also
depends on the pH, temperature, and sulfate availability. It has been reported that the most favorable pH conditions for the formation and
stability of ettringite are in the domain of 11 and 12.5 [19,20]. Increasing
the pH above the stability range will cause the ettringite to convert to
calcium hydroxide leaving behind a solution of sulfate and aluminate
ions. Damidot and Glasser reported from thermodynamic calculations that ettringite was stable at all times relative to monosulfate
(Ca4Al2(SO4) (OH)12) at 25 C [19,21]. In order for ettringite to develop
over monosulfate, the aluminum-to-sulfate ratio has to be lower than
1.0 [19]. The availability of more aluminum than sulfate will result in
sulfate developing into monosulfate and excess aluminum existing in
the hydroxyl-substituted AFm (alumina, ferric oxide, monosulfate),
hydroxy-AFm, phase. If there is a slight excess of sulfate, a mixture of
monosulfate and ettringite will develop in the system. As the available
sulfate level increases, more ettringite and less monosulfate will
develop [22,23].
Previous research has shown that sulfate was efciently removed
from recycled cooling water by precipitation with calcium and

Fig. 2. Schematic of the two-stage precipitation process of NF reject brine.

54

D. Almasri et al. / Desalination 362 (2015) 5258

aluminum at high pH using a process called ultra-high lime with aluminum (UHLA) [2427]. The UHLA process is an advanced alteration of the
lime softening treatment method and involves the addition of aluminum to enhance the removal of sulfate from the system as ettringite. Research conducted by Abdel-Wahab & Batchelor revealed sulfate
removal kinetics that was suitable enough for practical applications
[2527]. However, little information was available to support its application for the removal of sulfate from reject brine in ZLD systems.
Also, high sulfate concentrations in the reject brine from the rst stage
of a two-stage RO system depicted in Fig. 1 requires high doses of aluminum which results in higher treatment costs [24,28].
An attractive alternative for removing sulfate from reject brine is to
apply nanoltration (NF) before the second stage RO. However, in
order to achieve ZLD for the entire treatment system, the reject stream
from NF should be treated and recycled in the treatment system or combined with the feed of the second stage RO. The purpose of this paper is
to implement a two-stage treatment process for the removal of sulfate
from NF reject brine. In Stage 1, calcium is added to remove sulfate by
calcium sulfate precipitation while in Stage 2 both lime and aluminum
are added to remove the remaining sulfate from Stage 1 efuent by calcium sulfoaluminate precipitation. This approach would result in reducing the doses of aluminum required to remove sulfate if all sulfate from
the rst stage RO reject brine is to be removed using the UHLA process
alone. Also, this approach leads to reducing volumes of the overall precipitated solids in the treatment system. An equilibrium model was also
developed to simulate sulfate removal from the NF reject brine.
2. Experimental
2.1. Materials
The chemicals used in this research were sodium sulfate anhydrous
(Fisher), calcium chloride (96%, Sigma Aldrich, anhydrous), calcium hydroxide (ACS), sodium aluminate (EMD), sodium hydroxide (VWR), sodium chloride (Fisher Scientic) and hydrochloric acid (37%, VWR). All
solutions were prepared using deionized water (DI water).
2.2. Kinetic and equilibrium experiments
A bench scale two-stage process was developed to remove sulfate
from the NF reject stream as shown in Fig. 2. In Stage 1, sulfate is removed by precipitation of calcium sulfate while the remaining amount
was removed in Stage 2 by precipitation of calcium sulfoaluminate.
Kinetic and equilibrium experiments were performed in completely
mixed batch reactors using 250 mL high-density polyethylene (HDPE)
sealed plastic bottles that served as the reactor system. All experiments
were conducted in duplicates for quality assurance. The reactors were
tightly sealed after the addition of reagents and were rapidly mixed at
room temperature (2224 C). First, kinetic experiments were carried
out to obtain the optimum equilibrium time for sulfate removal in
each stage. The experiments were conducted at an initial sulfate concentration of 97 mM which is equivalent to the average concentration
of sulfate in brine rejected from the nano-ltration unit. Calcium chloride (CaCl2) was used as a calcium source in Stage 1. Lime (Ca(OH)2)
did not result in signicant removal of sulfate in this stage. At this
high sulfate concentration, a high lime dosage was required which
resulted in a high pH above the lime solubility limit, leaving insoluble lime in the reactor. Lime and sodium aluminate were used in
Stage 2 to provide aluminum and the additional calcium required
to remove the remaining sulfate in Stage 2 by calcium sulfoaluminate
precipitation.
The reactors were mixed by placing them on a shaker running at a
rate of 200 rpm at room temperature. For the kinetic experiments, samples were taken at 0.5, 1, 2, 4 and 8 h using a plastic syringe and ltered
using 0.45 m Whatman membrane lters (VWR). The ltered samples were then acidied to a pH 2 and refrigerated until the time of

Fig. 3. Kinetics of sulfate and calcium removal in Stage 1.

analysis. Equilibrium experiments were conducted at different chemical


reagent doses to initial sulfate concentration ratios to obtain the optimum ratios for maximum sulfate removal. Sodium sulfate was used as
the source of sulfate in all kinetic and equilibrium experiments.
2.3. Analysis
Experiments followed the analytical standard methods acquired
from the Standard Methods for the Examination of Water and Waste
water [29]. Ca2+, SO2
4 , and Cl were analyzed using a computerized
Dionex ICS-5000 Reagent Free system. Aluminum was analyzed using a
computerized ICP-OES that utilized inductively coupled plasma emitting
electromagnetic radiation at an analytical wavelength of 396.152 nm
with a plasma radial view. The pH was measured using a pH meter
(SympHony SP70P) with a SympHony electrode standardized with
pH 4, pH 7, pH 10, and pH 12 (VWR) buffers.
XRD analysis (Rigaku Ultima V) was performed on the dry precipitated solids from Stage 2 experiments, providing the most efcient sulfate removal, to identify the solid phases formed during the treatment
process. The solids were identied by comparing the peaks and the corresponding 2 values with the standard data from the Joint Committee
on Powder Diffraction Standards (JCPDS) Cards [30].
2.4. Equilibrium model development
An equilibrium model was developed to describe the chemical behavior in the treatment system. The model was used to predict the
nal sulfate concentration in the treated brine with the given chemical
reagents doses and initial sulfate concentrations. The solubilities of the
solids in the system were assumed to be controlled by precipitation.

Fig. 4. Effect of pH on sulfate removal in Stage 1. Initial CaCl2 concentration was 150% of the
initial sulfate amount.

D. Almasri et al. / Desalination 362 (2015) 5258

55

approximately 12.1 mM, consistent with the solubility and solubility


product of calcium sulfate precipitation [3436].
The effect of pH on sulfate removal was evaluated at a pH range of 7
12 and showed a negligible effect on the removal efciency of sulfate
(Fig. 4). This hypothesis is in agreement with previous reports stating
that the hydroxides available in the solution do not inuence the precipitation of calcium sulfate at a high pH range [37].
In this stage, the removal of sulfate at initial calcium to sulfate ratios
of 0.5, 1, 1.5, and 2 was also studied. Maximum sulfate removal was observed at a stoichiometric ratio of 2 as shown in Fig. 5. This is consistent
with Le Chatelier's principle for the following reaction:
2

Ca
Fig. 5. Effect of calcium doses on sulfate removal in Stage 1.

The geochemical modeling software, PHREEQC [31] was used to calculate nal concentrations knowing initial concentrations and reaction
equilibrium constants. Initial concentrations and chemical doses for
every set of experiments were dened in the PHREEQC input le. Individual element amounts were modied to attain charge balance and
equilibrium within a pure phase. Speciation computations were carried
out on each solution which were then utilized for succeeding batchreaction calculations. The initial concentrations of individual components in every solid solution were dened in the input data le. All computations concerning solid solutions assumed that the solid solutions
dissolved completely and re-precipitated in equilibrium with the solution [32]. The PHREEQC database le was updated to include the solubility product of the solids predicted to be produced in the process. Other
aqueous species and solids utilized by the model were dened in the database le.
3. Results and discussion
3.1. Kinetics and equilibrium characteristics of sulfate removal in Stage 1
To evaluate the kinetics of sulfate removal in Stage 1, calcium chloride concentration was added at 200% of the initial sulfate solution.
The calcium dose was added in excess, at twice the theoretical ratio of
CaSO4 precipitate, in order to enhance the sulfate removal efciency
[33]. The kinetic results indicated rapid removal of sulfate and calcium
(Fig. 3), being almost complete within 30 min of the reaction time. In
addition, a metastable equilibrium was observed at a reaction time of
2 h, which is the time that will be used for the equilibrium experiments.
The nal concentration of sulfate in the solution was still high, at

SO4 CaSO4 anhydrite; Ksp 2:0110

34:

Increasing calcium or sulfate concentrations will result in more


solids formed to maintain a thermodynamic equilibrium. The theoretical values of the nal sulfate and calcium concentrations based on the
solubility of calcium sulfate were obtained using the equilibrium modeling software, PHREEQC [31]. The model-predicted results versus experimental values of nal sulfate concentrations are shown in Fig. 5 and
indicate that calcium sulfate precipitation followed the equilibrium reaction and solubility product according to Eq. (1).
3.2. Equilibrium characteristics of sulfate removal in Stage 2
In order to simulate the real two-stage treatment system, the brine
source in Stage 2 was the ltrate from Stage 1 which was found to
have sulfate and calcium concentrations of approximately 23 3 mM
and 24 3 mM, respectively, conducted at an initial Ca:SO4 ratio of
1:1 for maximum efciency. Since CaCl2 is more costly than lime
(Table S1), the sources of calcium in Stage 2 were lime and the excess
calcium coming from the Stage 1 ltrate. Combinations of lime doses
(0%, 100%, 200%, and 300% of initial sulfate concentration) and sodium
aluminate doses (67%, 100%, 150%, and 200% of the initial sulfate concentration) were investigated at a reaction time of 2 h in Stage 2
[2527]. The 67% sodium aluminate dose corresponds to the aluminum
to sulfate ratio found in Ca6Al2(SO4)3(OH)12.
Efcient sulfate removal was observed in this stage at practical
ranges of lime and sodium aluminate doses (Fig. 6). Sulfate removal increased with increasing doses of lime and sodium aluminate, signifying
sulfate removal via precipitation with calcium and aluminum to form
CaAlSO4OH precipitates. Data shown in Fig. 6(a) indicate that stoichiometric ratios of lime and sodium aluminate above 100% and 67%
of the initial sulfate concentration, respectively, resulted in minor additional sulfate removal compared to additional doses added. Also, nal
calcium and aluminum concentrations corresponding to these doses

Fig. 6. Effect of lime and aluminum dose on (a) nal sulfate concentration, (b) nal calcium concentration and (c) nal aluminum concentration in Stage 2.

56

D. Almasri et al. / Desalination 362 (2015) 5258

Fig. 8. Measured and model predicted values of the ratio of Ca:SO4 removed in Stage 2 assuming solid mixture formation of calcium sulfoaluminate and monosulfate.

Fig. 7. Effect of pH on sulfate removal in Stage 2.

(Fig. 6(b) and (c)) are almost at their minimum values. Therefore, lime
and sodium aluminate ratios of 100% and 67% to initial sulfate concentration in Stage 2 could be adequate for practical applications and
these ratios were used for the subsequent experiments in this section.
The reagent costs and amounts are also of great importance to the treatment of NF brine in this process. A brief cost analysis is provided in
Table S1 to provide an insight on the chemical ratios to be added in
order to achieve an economically feasible and efcient process.
Although a NF unit is employed to remove sulfate, it also partially
removes chloride and as a result, the NF reject brine contains chloride.
Experiments were carried out to determine the inuence of initial chloride concentration on the removal efciency of sulfate in Stage 2. Results
revealed very slight decrease of sulfate removal at a very high concentration of chloride (Fig. S1).
The effect of pH on sulfate removal was evaluated at pH values ranging between 9.8 and 12.0. It was observed that sulfate removal in this
stage was highly dependent on the pH, whereby sulfate removal increased with increasing pH (Fig. 7). This agrees with previous studies
which report the most favorable pH conditions for the formation and
stability of the calcium sulfoaluminate solid to be in the domain of 11
and 12.5 [19,20].
A summary of the pH corresponding to the nal solutions and nal
sulfate concentrations in Stage 2 equilibrium experiments with varying
initial aluminum and lime doses is provided in Table S2.

3.3. Equilibrium modeling for sulfate removal in Stage 2


It was hypothesized that sulfate removal in Stage 2 was primarily controlled by the formation of calcium sulfoaluminate solid
(Ca6Al2(SO4)3(OH)12) assuming the following reaction:
6Ca

3SO4

2AlOH4 4OH Ca6 Al2 SO4 3 OH12 s:

develop is monosulfate (Ca4Al2(SO4)(OH)12) and forms through the


following reaction:
2

4Ca

SO4 2AlOH4 4OH Ca4 Al2 SO4 OH12 s:

Other possible solids that could precipitate include gypsum, lime,


gibbsite, and calcium chloroaluminate. An equilibrium model was developed in order to understand the precipitation mechanisms in Stage
2 and develop a tool that can predict nal concentrations knowing initial concentrations and chemical doses in an actual treatment system.
The solubility products of the solid phases that could form in the system
were added to the PHREEQC database and are shown in the following
Table 1.
Initial sulfate and calcium concentrations as well as chemical doses
of lime and sodium aluminate were inputted into PHREEQC to calculate
different scenarios assuming pure solid phases and a solid mixture of
different solid phases. The hypothesis made assuming the formation of
a solid mixture of calcium sulfoaluminate and monosulfate agreed
well with experimental results. The ratios of calcium to sulfate removed
at different lime and aluminum doses were calculated using the equilibrium model and were compared with the measured values (Fig. 8). If
calcium sulfoaluminate (Ca6Al2(SO4)3(OH)12) was the only solid
formed, the ratio of Ca:SO4 removed should be equivalent to 2, however,
Fig. 8 shows that the ratio falls in a range between 1.6 and 4.7, indicating
that monosulfate (Ca4Al2(SO4)(OH)12) could have also precipitated.
Fig. 9 depicts the nal sulfate concentrations at different lime and
aluminum doses calculated using the equilibrium model. Initial calcium
concentrations in the gure correspond to the calcium derived from
Stage 1 and calcium from the lime added. At high aluminum doses,
the measured values agreed with the model predictions. However, results for the experiments conducted at the low aluminate dose,

However, the ratios of sulfate removed to calcium and aluminum removed deviated from the theoretical ratio of the calcium sulfoaluminate
solid. This indicated that another solid phase(s) or a solid mixture containing more than one solid could have formed. One solid assumed to
Table 1
Solubility product of solid phases that could form in the UHLA process.
Solid name

Chemical formula

Log (Ksp)

Reference

Ettringite
Monosulfate
Calcium chloroaluminate
Gibbsite
Gypsum
Anhydrite
Lime

Ca6Al2(SO4)3(OH)12
Ca4Al2(SO4)(OH)12
Ca4Al2Cl2(OH)12
Al(OH)3
CaSO42H2O
CaSO4
Ca(OH)2

43.13
30.0
27.10
33.5
3.62
3.70
22.81

[34]
[21]
[38]
[39]
[35]
[34]
[39]

Fig. 9. Comparison between measured and model predicted values for nal sulfate concentrations in Stage 2. The points represent measured concentrations and the lines represent model predicted concentrations.

D. Almasri et al. / Desalination 362 (2015) 5258

18.4 mM, showed some deviation from the model predictions which
could be attributed to some lime not dissolving in the solution, leading
to less sulfate removal in the experiments performed at this dose.
The effect of lime and aluminum doses on the development of individual solids in the mixture was also evaluated using PHREEQC (Fig. 10).
An increase in aluminum dose was found to increase the precipitation of
monosulfate and decrease the precipitation of calcium sulfoaluminate
which is consistent with the sulfate to aluminum stoichiometric ratios
in these solids. The monosulfate solid includes a higher aluminum to sulfate ratio (2:1) than calcium sulfoaluminate (2:3). Therefore, increasing
the aluminum dose above the stoichiometric ratio of aluminum to sulfate corresponding to the stoichiometric ratio in calcium sulfoaluminate
solid is neither desirable nor cost effective.
The presence of calcium sulfoaluminate (ettringite) in the solids was
conrmed by XRD analysis. As shown in Fig. 11, the peaks at 8.72, 15.42
and 17.9 2 correspond to ettringite while monosulfate peaks were
found at 20.4, 22.4 and 31.2 2. These results agree with the hypothesis
that a solid mixture of ettringite and monosulfate was formed. XRD patterns displayed peaks which showed the presence of lime (at 17.3, 33.5,
and 46.1 2), calcium chloroaluminate (at 22.6 and 31.12) and gibbsite
(at 17.3, 20.08, and 44.8 2). The results obtained from the XRD analysis display the presence of the same solids that were assumed to exist by
the equilibrium model, which conrms the credibility of the model in
portraying the means of sulfate removal by the UHLA process.
The above approach is expected to decrease the volumes of the overall precipitated solids coming from the NF brine, by efciently precipitating the sulfate as CaSO4 in Stage 1 and calcium sulfoaluminate in
Stage 2. Signicant sulfate removal was found at practical doses of
lime and sodium aluminate (at 100% and 67% of the initial Stage 2 sulfate amount, respectively), leaving a nal sulfate amount of approximately 4 mM at a pH of about 12 (Table S2). This allows for the Stage
2 efuent to be recycled back into the treatment system or combined
with the second stage RO feed for feasibility.
4. Conclusions
The bench scale two-stage process showed effective removal of sulfate. Kinetics experiments for Stage 1 and Stage 2 of the process showed
fast removal of sulfate, being almost complete within 2 h. Results
showed that high sulfate removal in Stage 1 using CaCl2 as a reagent
was limited by the solubility of calcium sulfate, resulting in the need
of a Stage 2 for efcient sulfate removal via the precipitation of sulfate
as calcium sulfoaluminate. Lime and aluminum doses at ratios of 100%
and 67% of the initial sulfate concentration, respectively, provided optimum conditions for efcient sulfate removal. The removal of sulfate in
Stage 2 greatly depended on the pH, and effective sulfate removal was
observed above pH 11. An equilibrium model was developed using
PHREEQC to depict the chemical behavior in the two-stage process
and to predict the nal sulfate concentrations and solids formed at the

Fig. 10. Fractions of solids in the mixture in Stage 2.

57

Fig. 11. XRD patterns for the solids in Stage 2.

provided initial conditions. The model accurately predicted the experimental results and revealed that lime doses and aluminum doses inuenced the ratio of calcium to sulfate removed, forming either ettringite
or monosulfate precipitates, depending on the added amounts.

Acknowledgments
The researchers are grateful to Texas A&M University at Qatar and
Qatar Foundation for providing nancial support to perform this research work.

Appendix A. Supplementary data


Supplementary data to this article can be found online at http://dx.
doi.org/10.1016/j.desal.2015.02.008.

References
[1] H. Khordagui, Environmental Aspects of Brine Reject from Desalination Industry in
the ESCWA Region, ESCWA, Beirut, 1997.
[2] M.H. El-Naas, Reject brine management, in: M. Schorr (Ed.)Desalination, Trends and
Technologies, 2011, pp. 237252.
[3] A. Mohamed, M. Maraqa, J. Al Handhaly, Impact of land disposal of reject brine from
desalination plants on soil and groundwater, Desalination 182 (2005) 411433.
[4] J. Morillo, J. Usero, D. Rosado, H. El Bakouri, A. Riaza, F.-J. Bernaola, Comparative
study of brine management technologies for desalination plants, Desalination 336
(2014) 3249.
[5] R.Y. Ning, T.L. Troyer, Tandom reverse osmosis process for zero-liquid discharge, Desalination 237 (2009) 238242.
[6] R. Silva, L. Cadorin, J. Rubio, Sulphate ions removal from an aqueous solution. I: coprecipitation with hydrolysed aluminum-bearing salts, Miner. Eng. 23 (2010)
12201226.
[7] CSM, CSM Technical Manual: Reverse Osmosis Membrane, Saehan Industries, Korea,
2006, pp. 3237.
[8] Y. Wang, Composite Fouling of Calcium Sulfate and Calcium Carbonate in a Dynamic
Seawater Reverse Osmosis Unit, The University of New South Wales, 2005.
[9] A. Karabelas, A. Karanasiou, S. Mitrouli, Incipient membrane scaling by calcium sulfate during desalination in narrow spacer-lled channels, Desalination 345 (2014)
146157.
[10] A. Geldenhuys, J. Maree, M. De Beer, P. Hlabela, An integrated limestone/lime process for partial sulphate removal, J. South. Afr. Inst. Min. Metall. 103 (2003)
345354.
[11] INAP, Treatment of Sulphate in Mine Efuents, In: International Network for Acid
Prevention, 2003.
[12] R. Bowell, R. Connelly, J. Ellis, J. Cowan, A. Wood, J. Barta, P. Edwards, A review of sulfate removal options for mine waters, Proceedings of the International Mine Water
Association symposium, 2004, pp. 7591.
[13] R. Bowell, Sulphate and salt minerals: the problem of treating mine waste, Min. Environ. Manag. 8 (2000) 1113.
[14] M. Bader, Innovative processes to desulfate the Paradox Valley brine, Desalination
229 (2008) 5267.

58

D. Almasri et al. / Desalination 362 (2015) 5258

[15] N. Eckert, T. Drackett, I. Bailey, A. Merz, F. Mok, S. Lashkari, G. Ramasubbo,


Nanoltration process for enhanced brine recovery and sulfate removal, in, Patent
Application No. WO2013131183 A1, 2013.
[16] J. Schoeman, A. Steyn, Investigation into alternative water treatment technologies
for the treatment of underground mine water discharged by Grootvlei Proprietary
Mines Ltd into the Blesbokspruit in South Africa, Desalination 133 (2001) 1330.
[17] C. Namasivayam, D. Sangeetha, Application of coconut coir pith for the removal of
sulfate and other anions from water, Desalination 219 (2008) 113.
[18] W. Cao, Z. Dang, X. Zhou, X. Yi, P. Wu, N. Zhu, G. Lu, Removal of sulphate from aqueous solution using modied rice straw: preparation, characterization and adsorption
performance, Carbohydr. Polym. 85 (2011) 571577.
[19] M. Chrysochoou, D. Dermatas, Evaluation of ettringite and hydrocalumite formation
for heavy metal immobilization: literature review and experimental study, J. Hazard. Mater. 136 (2006) 2033.
[20] G.J. McCarthy, D.J. Hassett, J.A. Bender, Synthesis, crystal chemistry and stability of
ettringite, a material with potential applications in hazardous waste immobilization,
MRS Proceedings, Cambridge Univ Press, 1991, p. 129.
[21] D. Damidot, F.P. Glasser, Thermodynamic investigation of the CaOAl2O3CaSO4
H2O system at 25 C and the inuence of Na2O, Cem. Concr. Res. 23 (1993) 221238.
[22] B. Clark, P. Brown, The formation of calcium sulfoaluminate hydrate compounds:
part II, Cem. Concr. Res. 30 (2000) 233240.
[23] N.B. Winter, Understanding Cement: An Introduction to Cement Production, Cement Hydration and Deleterious Processes in Concrete, WHD Microanalysis Consultants, 2009.
[24] A. Abdel-Wahab, B. Batchelor, Chloride removal from recycled cooling water using
ultra-high lime with aluminum process, Water Environ. Res. 256263 (2002).
[25] A. Abdel-Wahab, B. Batchelor, Effects of pH, temperature, and water quality on chloride removal with ultra-high lime with aluminum process, Water Environ. Res. 78
(2006) 930937.
[26] A. Abdel-Wahab, B. Batchelor, Interactions between chloride and sulfate or silica removals using an advanced lime-aluminum softening process, Water Environ. Res.
78 (2006) 24742479.

[27] A. Abdel-Wahab, B. Batchelor, Interactions between chloride and sulfate or silica removals from wastewater using an advanced lime-aluminum softening process:
equilibrium modeling, Water Environ. Res. 79 (2007) 528535.
[28] S. Tait, W.P. Clarke, J. Keller, D.J. Batstone, Removal of sulfate from high-strength
wastewater by crystallisation, Water Res. 43 (2009) 762772.
[29] A.D. Eaton, M.A.H. Franson, Standard Methods for the Examination of Water &
Wastewater, 2005.
[30] P.D. File, Joint Committee on Powder Diffraction Standards, ASTM, Philadelphia, Pa,
1967. 9185.
[31] D.L. Parkhurst, C. Appelo, User's Guide to PHREEQC (Version 2): A Computer Program for Speciation, Batch-reaction, One-dimensional Transport, and Inverse Geochemical Calculations, 1999.
[32] D.L. Parkhurst, C. Appelo, Description of Input and Examples for PHREEQC Version
3A Computer Program for Speciation, Batch-reaction, One-dimensional Transport,
and Inverse Geochemical Calculations, 2013.
[33] F. Alimi, H. Ell, A. Gadri, Kinetics of the precipitation of calcium sulfate dihydrate in
a desalination unit, Desalination 158 (2003) 916.
[34] R. Damons, F. Petersen, An aspen model for the treatment of acid mine water, Eur. J.
Miner. Process. Environ. Prot. 2 (2002) 6981.
[35] G. Azimi, V. Papangelakis, J. Dutrizac, Modelling of calcium sulphate solubility in
concentrated multi-component sulphate solutions, Fluid Phase Equilib. 260 (2007)
300315.
[36] J. Dutrizac, Calcium sulphate solubilities in simulated zinc processing solutions, Hydrometallurgy 65 (2002) 109135.
[37] C.T. Benatti, C.R.G. Tavares, E. Lenzi, Sulfate removal from waste chemicals by precipitation, J. Environ. Manag. 90 (2009) 504511.
[38] U.A. Birnin-Yauri, F.P. Glasser, 28, 1713, Friedel's salt, Ca2Al(OH)6(Cl, OH)2H2O: its
solid solutions and their role in chloride binding, Cem. Concr. Res. 28 (1998) 1713.
[39] W. Stumm, J.J. Morgan, Aquatic Chemistry: Chemical Equilibrium and Rates in Natural Waters, Wiley, New York, 1996.

You might also like