Você está na página 1de 13

Desalination 273 (2011) 2335

Contents lists available at ScienceDirect

Desalination
j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / d e s a l

Boron removal from saline water: A comprehensive review


N. Hilal a,, G.J. Kim b, C. Somereld b
a
b

The Centre for Clean Water Technologies, Multidisciplinary Nanotechnology Centre, School of Engineering, Swansea University, Singleton Park, Swansea SA2 8PP, UK
Faculty of Engineering, The University of Nottingham, Nottingham NG7 2RD, UK

a r t i c l e

i n f o

Article history:
Received 25 January 2010
Received in revised form 5 May 2010
Accepted 6 May 2010
Available online 23 June 2010
Keywords:
Boron removal
Reverse osmosis
Boron selective resins
AMF
Saline water
Desalination

a b s t r a c t
Boron is an essential micronutrient for plants and animals as well as a useful component for numerous
industries. It is necessary to produce low boron containing water from RO desalination plants for both
human consumption and for agriculture. For plants, a small amount of boron is necessary for their growth
and development, but boron becomes toxic if the amount is slightly greater than required. Desalinated
seawater from RO plants often contains high boron content and, when used for irrigation, has been proven to
be damaging to crops including blackberry, lemon, and grapefruit. Apart from the toxic effects of boron on
plants, boron should be removed from RO desalination plants to comply with the current guideline value,
0.5 mg/L, for potable water issued by the World Health Organisation (WHO). Currently there is no simple
method to remove boron from saline water. The use of multi-pass reverse osmosis membrane (RO) with pH
modication and the use of ion exchange using boron selective resins (BSRs) have both been considered as
effective methods for the removal of boron. A hybrid process, Adsorption Membrane Filtration (AMF), has
received attention as an emerging technology for boron removal with a high efciency and low operating
costs. The purpose of this review is to give an overview on boron in general and to discuss its toxicity. The
problems of boron in the MENA (Middle East and North Africa) region are discussed as well as technologies,
current and future, for the removal of boron from seawater. The focus is placed on current RO and ion
exchange methodologies using BSRs as well as the future for the AMF method. The fundamentals of each
process, the effects of experimental parameters, and ndings are discussed.
2010 Elsevier B.V. All rights reserved.

Contents
1.

2.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . .
1.1.
Boron distribution in nature . . . . . . . . . . . . . .
1.2.
Importance of boron . . . . . . . . . . . . . . . . .
1.3.
Toxicity of boron . . . . . . . . . . . . . . . . . . .
1.4.
Boron problems and drinking water standards . . . . .
1.5.
Boron in seawater . . . . . . . . . . . . . . . . . . .
The overview of boron removal technologies . . . . . . . . .
2.1.
Reverse osmosis . . . . . . . . . . . . . . . . . . . .
2.1.1.
Principles and congurations . . . . . . . . .
2.1.2.
Effects of each parameter on boron rejection . .
2.2.
Ion exchange methods . . . . . . . . . . . . . . . . .
2.2.1.
Mechanism of boron removal . . . . . . . . .
2.2.2.
Research on synthesis of BSRs . . . . . . . . .
2.2.3.
Batch sorption studies . . . . . . . . . . . .
2.2.4.
Effects of each parameter on boron removal . .
2.2.5.
Elution and regeneration studies . . . . . . . .
2.3.
Adsorption membrane ltration (AMF) . . . . . . . . .
2.3.1.
Basic principles . . . . . . . . . . . . . . . .
2.3.2.
Adsorption membrane ltration for the removal
2.3.3.
Effects of each parameter on boron removal . .

Corresponding author. Tel.: +44 1792 295587.


E-mail address: n.hilal@swansea.ac.uk (N. Hilal).
0011-9164/$ see front matter 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.desal.2010.05.012

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
of
.

. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
boron
. . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

24
24
24
24
24
25
26
26
26
26
29
29
29
29
30
31
32
32
32
33

24

N. Hilal et al. / Desalination 273 (2011) 2335

3.
Future research . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
4.
Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

1. Introduction
1.1. Boron distribution in nature
Boron is widely distributed in both the hydrosphere and lithosphere
of the earth [1]. In nature, boron is never found elemental but always as
various forms of complex compounds combining with oxygen and
other elements. In the lithosphere, it is found in rocks and soils. The
concentration of boron in the earth's crust varies from 1 to 500 mg/kg,
depending on the nature of the rock [2]. According to Krauskopf [3], an
average of boron in the earth's crust is around 10 mg/kg, representing
only 0.001% of the elemental composition of the earth. The amount of
boron in soils ranges from 2 to 100 mg/kg with an average of 30 mg/kg
[4]. The total amount of boron stored in the lithosphere is estimated as:
the continental and oceanic crusts (1018 kg B), coal deposits (1010 kg B),
commercial borate deposits (1010 kg B), and biomass (1010 kg B) [5]. In
the hydrosphere, boron exists in seawater at around 4.5 mg/L and in
ground water with concentrations raging from 0.3 to 100 mg/L [6]. The
distribution of boron contents in the hydrosphere was identied by
Argust [5] as the oceans (1015 kg B), groundwater (1011 kg B), ice
(1011 kg B), and surface waters (108 kg B).
1.2. Importance of boron
Boron is an essential micronutrient affecting plant growth and
development. The concentration of boron in irrigation water or in soils
plays an important role in both crop yields and the quality of produce
[7]. The most widely accepted role of boron is as a structural component of primary cell walls [8]. As well as the role of boron in cell wall
formation, boron is also involved in the structural integrity of the cell
wall providing stability to the cell wall matrix. Also boron is involved
in lignications, membrane transports, enzyme interactions, nucleic
acid synthesis, phenol and carbohydrate metabolism, and the transport of sugar, polyol and hydroxy acid [7,911].
There have been a number of studies proving that boron is an
essential or at least a benecial element for human beings and animals
[12,13]. Fort et al. [14] reported that boron is essential for the
development of an embryo for at least vertebrates and found that
disrupted embryonic development is caused by the deciency of boron.
From the results of nutritional research, it was suggested boron gives
benecial effects related to bone metabolism [15,16]. Moreover, it was
reported that dietary boron involves in immune function of organisms
[12,17].
Boron is also a useful component for numerous industries. It has
been widely used as ingredients for a range of products. The glass
industry is the biggest single user, which consumes more than half of
the total world production of boron compounds. Besides the glass
industry, boron is used in cleaning products, semiconductors, cancer
treatment, cosmetics, etc. [18]. In the nuclear industry, it is crucial to
utilise isotope boron-10 in order to control the reaction rate of nuclear
reaction and avoid a nuclear explosion [19].
1.3. Toxicity of boron
As discussed in the previous chapter, boron is essential for the
growth of plants in many ways. However, it becomes toxic to plants
when the amount of boron is slightly greater than required. There is
little doubt that the range of boron concentration between deciency
and excess is narrow, for instance, in sunower 0.5 ppm boron affords

good growth while 1 ppm is denitely toxic [20]. Table 1 shows the
maximum permissible concentration of boron in soil water without
reducing yield for a variety of crops and also the relative tolerance of
these crops.
In regard to the adverse effects of boron on plants, it has been
extensively reviewed [8]. The physiological adverse effect of boron
involves the reduction of root cell division, retarded shoot and root
growth, inhibition of photosynthesis, deposition of lignin and suberin,
decrease in leaf chlorophyll, etc [8]. In most crops, the symptoms of
boron toxicity are shown as burned edges on the older leaves, yellowing
of the leaf tips, accelerated decay, and ultimately plant expiration. These
symptoms are highly dependent on the amount of boron excess and the
tolerance of the plant [21].
The toxic effects of boron in animals and humans remain largely
unknown. Boron toxicity depends on the length, frequency, and level
of exposure and therefore it is difcult to quantify. A vast amount of
research has been undertaken to discover adverse effects of boron in
animals. A chronic exposure of boron may cause cutaneous disorders
[23], retarded growth [24] and have adverse impact on the male
reproductive system in rats and mice [2528].
1.4. Boron problems and drinking water standards
In the MENA region, there has been, and continues to be, signicant
fresh water shortages due mainly to low annual rainfall combined with
geological characteristics. As it can be seen from Table 2, the rate of
population growth in Africa, South and South-West Asia containing the
MENA region is higher than the others. This growth is a crucial contributing factor for water shortages in these areas. In order to obtain
fresh water for drinking and irrigation, desalination technology has been
widely used in these regions since the rst industrial desalination
plant was built in 1957 [29].
In the MENA region, to compensate for low rainfall, a large amount
of irrigation water is used for the cultivation of plants. However,
RO desalinated water used for this purpose often contains a high
concentration of boron [31]. Normally, under low rainfall condition,
boron is not sufciently leached by rain and it persists in soils [31].
Thus, irrigation water containing high boron levels can be a factor
accelerating the deposition of boron in both soils and plants. Consequently it is important to produce low boron containing water
from desalination plants to prevent the effects of boron toxicity on
plants.
Boron in seawater is effectively removed to nearly zero concentration by thermal desalination technology. This effectiveness is not
mirrored by reverse osmosis desalination technology where boron
removal is shown to be insufcient. This is because a large proportion
of boron in seawater generally exists as uncharged boric acid which
can diffuse through the RO membrane in the way similar to water.
This in turn decreases the percentage rejection of boron from seawater [32,33]. Boron removal from seawater is a challenging problem
for RO desalination plants. RO desalination has become more
dominant (Fig. 1) as thermal desalination loses favour on energy
grounds. RO desalination plants built in the 1980s and 1990s have a
limited ability to remove boron in comparison to newly built RO
desalination plants. These older desalination plants are still operating
and retrot modications are necessary to ensure the concentration
of boron product is as low as required.
For many years boron was not considered as a toxic element.
In 1958, 1963 and 1971 there was no mention of boron in the WHO

N. Hilal et al. / Desalination 273 (2011) 2335

25

Table 1
Boron Tolerance Limits for Agricultural Crops [22] Notes. (a) Maximum permissible concentration in soil water without yield reduction. Boron tolerances may vary, depending upon
climate, soil conditions, and crop varieties. (b) Tolerance based on reduction in vegetative growth.
Tolerance

Concentration of boron in soil water (mg/L)

Agricultural crops

Extremely sensitive
Very sensitive

b 0.5
0.50.75

Sensitive

0.751.0

Moderately sensitive
Moderately tolerant

1.02.0
2.04.0

Tolerant
Very tolerant
Extremely tolerant

4.06.0
6.010.0
10.010.5

Blackberry(b), Lemon(b)
Avocado(b), grapefruit(b), orange(b), apricot(b), peach(b), cherry(b),
plum(b), persimmon(b), Kadota Fig(b), grape(b), walnut(b), pecan(b), onion
Garlic, sweet potato, wheat, sunower, mung bean(b), sesame(b), lupine(b), strawberry(b),
Jerusalem artichoke(b), kidney bean(b), snap bean, lima bean(b), peanut
Broccoli, red pepper, pea(b), carrot, radish, potato, cucumber, lettuce(b)
Cabbage(b), turnip, Kentucky bluegrass(b), barley, cowpea, oats, corn, artichoke(b), tobacco(b),
mustard(b), sweet clover(b), squash, muskmelon(b), cauliower
Alfalfa(b), purple vetch(b), parsley (b), red beet, sugar beet, tomato
Sorghum, cotton, celery(b)
Asparagus(b)

International Standards for Drinking-water. The rst mention of


boron in drinking water was encountered in the document, Drinkingwater Quality, published in 1984 where the conclusion made indicated
that no action was required. In other words, boron was not considered
harmful at this time. Provisional guidelines for boron concentration in
drinking water by WHO were rst introduced in 1993. However, since
boron has been shown to induce several harmful effects on animals
in laboratory studies, the recommended guideline value of 0.3 mg/L,
based on NOAEL (No Observed Adverse Effect Level) was set in the
1993 guidelines. This guideline value was increased to 0.5 mg/L in
the published guidelines in 1998 as it was difcult to comply with the
guideline due to the poor performance of boron removal processes
encountered in water treatment. This 0.5 mg/L value is still valid in
2010. However, according to the draft of the background document for
development of WHO Guidelines [35], boron is included in the plan of
work of the rolling revision. Extensive data from the UK and the USA on
dietary intakes of boron showed that the intake from air and food is
lower than expected. This could lead the increase in the allocation to
drinking water from 10% [6] to 40% [35], without approaching the
Tolerable Daily Intake (TDI). In this case, the guideline value of boron
would rise to 2.4 mg/L.
Table 3 shows the boron standards of drinking water in several
regions. As it can be seen from the table, most countries do not
follow the WHO recommendation. Saudi Arabia is the only country
complying with the guideline in the list. Meanwhile, the USA has no
federal regulations of boron and Canada and Australia have set the
maximum boron concentration much higher than the guideline. There
are two major reasons for this. Firstly, there is insufcient information proving harmful effects on human beings. Secondly, it is difcult
and/or expensive to remove boron from water to achieve the guideline
value.
From Table 1, it is seen that to enable the growth of agricultural crops
without toxic effects, some boron sensitive crops require a lower boron

concentration than encountered in drinking water. Consequently, in


these cases, such as the MENA region where a large amount of irrigation water is supported by desalination plants, boron concentration
in desalinated water should be low enough to grow plants regardless
of the potable water guideline.
1.5. Boron in seawater
In aqueous environments, dissolved boron is present as several
species, depending on the concentration of boron. At low boron
concentration (216 mg/L), dissolved boron is mainly found as the
mononuclear boron species, B(OH)3 and B(OH)
4 . At higher concentrations with the increase in pH, polynuclear boron species such as
B2O(OH)2
or those incorporating B3O3 rings such as B3O3(OH)
6
4 ,
B4O5(OH)2
,
and B5O6(OH)
4
4 are formed [49]. According to Su and
Suarez [50], those polynuclear ions are negligible at the concentrations lower than about 290 mg/L. As the concentration of boron in seawater is around 4.8 mg/L [51], it is acceptable that only mononuclear
species B(OH)3 and B(OH)
4 are present in seawater.
Boric acid, B(OH)3, behaves as a very weak Lewis acid according to
the dissociation reaction in Eq. (1).

BOH3 H2 OBOH4 H :

The distribution of two components, boric acid and borate ion,


depends on the dissociation constant of boric acid (pKa). The constant
is a function of temperature, pressure, pH, and ionic strength. The
equilibrium between boric acid and borate ion is important due to the

Table 2
Demographic trends of the world population by region [30].
Region

East and North-East Asia


South-East Asia
South and South-West Asia
Africa
Europe
Latin America & Carib.
North America
World

Population growth rate


19901995

19952000

20002005

2007

1.0
1.8
2.2
2.6
0.2
1.7
1.1
1.6

0.9
1.6
1.9
2.5
0.1
1.6
1.0
1.4

0.6
1.4
1.7
2.3
0.2
1.3
1.0
1.2

0.5
1.3
1.6
2.3
0.1
1.3
1.0
1.2

Fig. 1. The cumulative contracted desalination plants by technologies until 2008 [34].

26

N. Hilal et al. / Desalination 273 (2011) 2335

Table 3
Boron in drinking water standards by regions.
Region

Maximum boron
References and comments
concentration (mg/L)

Saudi Arabia

0.5

United States of America


(USA)
State of Minnesota
State of New Hampshire
State of Florida
State of Maine
State of Wisconsin
State of California
European Union (EU)
including UK
South Korea

0.6
0.63
0.63
0.63
0.9
1
1

Japan
New Zealand
Israel

1
1.4
b 1.5

Australia
Canada

4
5

WHO recommendation

0.5

SASO [36,37] Bottled and


unbottled drinking water
USEPA [38] No federal
regulations of boron
USEPA [39]
USEPA [39]
USEPA [39]
USEPA [39]
USEPA [39]
USEPA [39]
EEA [40]

Ministry of Environment [41]


Changed from 0.3 mg/L
NIPH [42]
Ministry of Health [43]
ISRAEL MEP [44] This is for
industrial sewage, it should be
lower for drinking water
NHMRC [45,46]
CDW [47] It has not changed
since 1990
WHO [6] changed from 0.3 mg/L
[48]

contribution to the buffering properties and the physical properties of


seawater such as a sound adsorption property [51]. The equilibrium
is also important for the calculation of alkalinity shown in Eq. (2).
The equation shows it is necessary to take into account the borate
ion concentration in quantitative calculations of alkalinity when its
concentrations are signicant.

Alkalinity HCO3  2CO3  BOH4  OH  2PO4 


2

HPO4  SiOOH3 H HSO4 :

There have been several studies to discover the correlations


between the dissociation constant of boric acid and each governing
factor such as temperature, pressure, pH, and salinity. Dickson [52]
investigated the dissociation of boric acid in seawater over ranges of
salinity (500045,000 ppm) and temperature (045 C) and devised a
formula to estimate the boric acid dissociation constant. Roy et al. [53]
have subsequently found that the measurement of the dissociation
constant undertaken by Dickson is reliable. Regarding the effect of
pressure on the dissociation constant, Culberson and Pytkowicz [54]
determined the pressure coefcients of boric acid for pressures up to
1000 atm at 2 C, 11.6 C, and 22 C.
From the equations given by Dickson [52], and Culberson and
Pytkowicz [54], the distribution of boric acid and borate in seawater can
be calculated. Fig. 2 shows the variation of distribution by the changes in
each parameter. As can be seen from Fig. 2, pH is the most inuential
parameter while the effect of pressure is negligible at normal pressure
ranges. In regard to temperature and salinity, boric acid is easily dissociated as a temperature and salinity increase.
2. The overview of boron removal technologies
2.1. Reverse osmosis
2.1.1. Principles and congurations
Reverse osmosis is a widely utilised membrane technology for
seawater desalination. The basic concept of Sea Water Reverse Osmosis
(SWRO) is to utilise a semi-permeable membrane where most of the

dissolved species are rejected while water permeates it. Normally, water
spontaneously diffuses through a semi-permeable membrane from a
dilute solution to a concentrated solution. However, the reverse process
is realised by applying external pressure enough to overcome the
osmotic pressure on seawater. This phenomenon is the basis of seawater desalination and water treatment by RO.
Boron in the form of negatively charged borate ion, B(OH)
4 , is easily
separated by RO membranes like other dissolved ionic substances. However, it is difcult to remove non-dissociated neutrally charged boric
acid in seawater by RO. Reported rejection rates vary and are membrane
dependant. Prats et al. [32], reported the rejection of boric acid in the
region 4060% at pH ranges of 5.59.5 while borate ion removal under the
same conditions is more than 95%. Therefore the overall boron removal
by RO is dependent on the boric acid/borate ion ratio and using a RO
membrane for feed water with a high proportion of boric acid will lead to
unsatisfactory levels of boron in the permeate water.
According to Glueckstern and Priel [55], currently available RO
membranes for seawater have the ability to remove boron (boric acid
and borate ions) from 85 to 90% at standard test conditions by manufacturers. However, this corresponds to around 7880% boron rejection
at normal operation conditions of SWRO systems. Koseoglu et al. [56]
investigated the removal of boron from sample seawater using two
commercially available high boron rejection membranes. From the data
obtained, around 90% of removal was achieved at normal operation
conditions (pH 8.2, 55.2 bar, 25 C) when using the high boron rejection
membranes.
In order to reduce the concentration of boron in RO permeate and
comply with stringent requirements, several designs have been developed. Redondo et al. [57] discussed four proposed concepts from an
economic and technical point of view. The schematic diagrams of each
design concept and results are shown in Table 4. From the study, it was
found that desalination cost varies with required boron concentration in
products. Typical water production costs for boron contents of 0.61 mg/L
was 0.380.50 $/m3 while it was 0.470.60 $/m3 for boron contents of
0.30.5 mg/L in product water. Additional costs for removal of boron
to a level of 0.4 mg/L were estimated in a range of 0.050.07 $/m3 in a
large system and 50% or more in small systems.
Glueckstern and Priel [55] optimised 8 different design concepts to
minimise boron removal costs. It was reported that cost reduction can
be achieved by hybridisation of the second pass RO with BSR. However,
in that situation, an optimisation of the process should be carried out
because the salinity of product is higher than a conventional 2 pass
system. Research done by Taniguchi et al. [58] supported the fact that a
combination with SWRO, BWRO, and BSR is an effective way to reduce
production costs associated with high boron removal.
Several design methods were investigated in a full scale plant operation by Faigon and Hefer [59]. From the comparison between two pass
design with recycle and cascade design shown in Fig. 3, it was found that
the cascade design has several advantages over the two pass system.
For example, by using the cascade design, it is possible to adjust operational conditions according to plant temporal needs and to reduce the
cost of water production.
The cascade design was used to retrot the SWRO plant in Eilat.
The SWRO plant was, rstly, retrotted to increase the production to
10,000 m3/day from 8000 m3/day in 1999. However, due to the ageing
of the membranes since 1997, not only was there a decrease in
production capacity to 9500 m3/day, but also an increase in the boron
and chloride concentrations in the product. A second retrot in 2004,
increased production to 10,000 m3/day and reduced the boron
concentration to below 0.4 ppm. Nadav et al. [60] reported in 2005
that the retrot was still successful with boron in the permeate
reported as 0.39 ppm.
2.1.2. Effects of each parameter on boron rejection
The removal of boron by RO is affected by many factors, such as,
temperature, pressure, pH, feed ow rate, feed salinity or ionic

N. Hilal et al. / Desalination 273 (2011) 2335

27

Fig. 2. The distribution of boric acid and borate in seawater by the changes of each parameter. Notes. (a) pH = 8, temperature = 25 C, pressure = 1 atm (b) pH = 8,
salinity = 35,000 ppm, pressure = 1 atm (c) temperature = 25 C, salinity = 35,000 ppm, pressure = 1 atm (d) pH = 8, temperature = 10 C, salinity = 34,800 ppm.

Table 4
Boron removal processes for seawater desalination (Feed concentration 45 mg/L and temperature 1826 C) (summarised from Redondo et al. [57]).
Design concept

Achievable concentration and cost

1 pass SWRO (with/without pH changes)

0.81.0 mg/L (0.380.52 $/m3)

2 pass with increased pH

0.40.5 mg/L (0.450.55 $/m3)

2 pass with BSR

Less than 0.4 mg/L (0.500.55 $/m3)

Cascade design (IDE process)

Less than 0.4 mg/L (0.470.52 $/m3)

Examples of simplied schematic diagram

28

N. Hilal et al. / Desalination 273 (2011) 2335

Fig. 3. Schematic drawings of (a) the two pass design with recycle and (b) the cascade design (Modied from [59]).

strength, an initial boron concentration, and recovery ratio. Thus, it is


critical to nd out a relationship between % removal of boron and
affecting parameters and to develop reliable designs in order to
optimise the boron removal process. Parameters affecting on the
performance of boron removal have received much research
attention. The results in Table 5 indicate that each parameter shows
the similar effect on boron removal regardless of membrane used.
The impact of pH on boron removal by RO membranes is dominant
while initial boron concentration and feed ow rate are negligible.
Generally, boron rejection increases from around 5075% at pH 78 to
over 95% at pH 10.5 and is mainly due to the increased proportion of
borate ions as the pH rises. There is also a smaller effect on rejection
which is governed by the type of membrane used [32,61]. In regard to
feed pressure, the boron rejection tends to increase as the feed
pressure becomes high [32,56,62]. However, as an exception, the
removal of boron decreased by the increase in feed pressure from
700 psi to 800 psi when the Filmtec SW30HR membrane was used
[56]. Complexation reactions can also be used to increase boron

rejection and this is achieved by an addition of mannitol or Fe which


cause the formation of large complexes [63,64].
Increased feed salinity and/or temperature shows increased
amounts of boron in the permeate. As it can be seen Fig. 2,
theoretically higher salinity and temperature cause lower pKa value,
which increases the fraction of borate ion at certain pH. Thus, it seems
possible to remove boron at lower pH by increasing salinity or
temperature. However, the results clearly showed that boron removal
declines as the salinity increases. According to Oo and Song [65], it
might be explained as the effect of charge neutralisation or hindering
of membrane surface potential at high salinity. For example,
electrostatic repulsion between membrane surface and borate ion
becomes less dominant or membrane charge density is changed and
caused the diffusion of ions faster through the membrane when the
membrane surface is less negatively charged at high salinity. Similar
to increased salinity, increased temperature also results in an increase
in boron in the permeate, despite the proportion of borate ion
increasing at a higher temperature. This can be explained by the fact

Table 5
The effect of each parameter on boron rejection by RO.
Parameters

Boron rejection References Used membranes

pH increases

Increases

Feed pressure increases

Increases

[32]
[56]
[61]
[32]

Feed ow rate increases


Initial boron concentration
increases
Feed salinity increases

No effect
No effect

Recovery ratio increases

Decreases

[56]
[62]
[66]
[61]
[56]
[67]
[56]
[61]
[65]
[67]

Mannitol or Fe increases

Increases

[63]

Feed temperature increases Decreases

Decreases

[64]

Hydranautics CPA2 Toray SU-710, SUL-G10


Toray UTC-80-AB Filmtec SW30HR
Woongjin Chemical RE8040-SR, RE8040-SHN
Hydranautics CPA2 Toray SU-710, SUL-G10

Comments

pH 5.510.5
pH 8.2 and 10.5
pH 6.510.5
The operation pressure was chosen to obtain recoveries
of 10, 20, 30 and 40 (Brackish water)
Toray UTC-80-AB Filmtec SW30HR
600800 psi No effects for SW30HR (Slightly decrease)
Filmtec SWHR, BW-30 GE membrane AG
1535 bar
Hydranautics SWC4+, SW30 HR LE Toray TM820, TM820A 15 C, 25 C, 35 C at two pH conditions (6.2 and 9.5)
Woongjin Chemical RE8040-SR, RE8040-SHN
15 35 C and pH conditions (8.18.3)
Toray UTC-80-AB Filmtec SW30HR
Cross-ow velocity 0.51.0 m/s
Nitto Denko NTR-70SWC, ES10-D4
No remarkable effects lower than 35 mg/L
Toray UTC-80-AB Filmtec SW30HR
No signicant effects up to 6.6 mg/L
Woongjin Chemical RE8040-SR, RE8040-SHN
20,00045,000 ppm
Hydranautics ESPA1, LFC1, CPA2
50015,000 ppm (Brackish water)
Nitto Denko NTR-70SWC, ES10-D4
50%, 80%, 90% Recovery Impacts of high recovery is
enhanced as pH goes up.
Hydranautics ESPA1
Addition of Fe improves boron rejection. 95% rejection
was obtained when ratio of Fe/B was 12
Filmtec NF200, BW30, SW30
Boron rejection is enhanced by addition of mannitol due
to complexation reaction

N. Hilal et al. / Desalination 273 (2011) 2335

29

Fig. 4. Schematic drawings of the neutral cis-diol monoborate ester (left), the
monoborate complex (middle) and the bis(diol) borate complex (right) [49].

that total ux through RO membranes including salt ux increases at


higher temperature.

Fig. 5. Structural formula of N-methyl-D-glucamine (left) and monoborate complex


(right).

2.2. Ion exchange methods


2.2.1. Mechanism of boron removal
Boron removal by a boron specic ion exchange resin was developed as early as in the 1970s, in order to remove borate from magnesium brine in the ceramic industry [68]. The principle of boron
selective resins is the use of complexation reactions. Boric acid reacts
with compounds possessing multi hydroxyl groups (polyols) to form
a variety of borate esters, in accordance with Eq. (3) where R is a
hydrocarbon group. The esters formed rapidly dissociate releasing
protons. Thus, the amount of acidication produced upon the addition
of the polyols is proportional to the extent of ester formation, which
is used for monitoring the reaction [49].
B3 3RHR3 32 O:

The stability of the borate complex formed strongly depends on


the type of diol used. According to Power and Woods [49], a strong
complex is formed when the diol used involves the hydroxyl groups
oriented in such a way that they accurately match the structural
parameters required by tetrahedrally coordinated boron. For example,
a stable complex is formed by the reaction of boron with compounds
possessing cis-diol system such as D-mannitol, D-sorbitol, and D-ribose
[64]. In this case, stable borate complex, cis-diol monoborate ester or
bis-diol borate complex is formed, which are shown in Fig. 4.
Regarding commercially available BSRs, these resins have a macroporous polystyrene matrix, on which N-methyl-D-glucamine functional
groups are attached. Commercially available BSRs are tabulated in
Table 6. The functional group has a tertiary amine end and a polyol end,
which is shown in Fig. 5. The role of the tertiary amine in the functional
group is to neutralise the proton brought by the formation of tetra
borate complex. Amine protonation shown in Eq. (4) is critical to
prevent the decrease of pH by proton released from the dissociation of
borate esters. As far as hydroxyl groups are concerned, there are 5
hydroxyl groups in N-methyl-D-glucamine. This allows the formation
of a strong complex with boron and improves the possibility of complexation by offering several sites for boron.

CH2 NCH3 CH2 H CH2 N HCH3 CH2

2.2.2. Research on synthesis of BSRs


There have been efforts to improve the ability of BSRs since
polystyrene based resins with N-methyl-D-glucamine function emerged

as boron specic resins in the mid 1960s [69]. As the performance of the
resins depends on its polymeric support and functional groups, it has
been of interest to develop new polymeric support or functional groups.
Bicak et al. [70] prepared terpolymers of glycidyl methacrylate
(GMA)methyl methacrylate (MMA)divinyl benzene (DVB) and
used the terpolymers as a support for boron specic resins possessing
N-methyl-D-glucamine as functional groups. From sorption and elution
tests, it was found that the resin showed good stability in terms of
particle disintegration for long term uses and better performance of
regeneration in comparison to common polymeric boron sorbents.
Senkal and Bicak [71] demonstrated the boron chelating ability
of iminodipropylene glycol and then extended this boron chelate
sorption onto polymer supported iminopropylene glycols. In their experiments, GMAMMADVB beads were prepared by using the same
method used in Bick et al. [70] and used as a support polymer. From
batch sorption and regeneration tests, it was found that the resin used
shows an excellent boron loading capacity of 32 mg/g and has a reasonable regenerability.
Wang et al. [72] developed a chelating polymeric sorbent having
N-methylglucamine groups as functional group and poly GMA-co-TRIM
as support. The resin showed low swelling degree and high capacity
of boron sorption. Although the capacity was the same as that of a commercially available N-methylglucamine type polystyrene resin, the rate
of sorption was more rapid. It was found that fast sorption kinetics is
caused by its low swelling characteristic, permanent macroporous structures and more hydrophilic characteristics.
Parschova et al. [73] compared the performance of each sorbent
possessing N-methyl-D-glucamine functional groups with different
polymeric supports, polypropylenestyrene, polypropyleneGMA, viscoseGMA and commercially used polystyreneDVB (Purolite D-4123),
respectively. The breakthrough capacity of the resins synthesised was
much lower than the commercial BSR under the conditions investigated.
However, viscoseGMA based sorbent showed much faster sorption
kinetics and was easy to regenerate even with diluted (0.1 mol/L)
hydrochloric acid.

2.2.3. Batch sorption studies


The batch sorption test is usually carried out to nd the optimum
amount of resin. The kinetic characteristics and parameters such as pH
and temperature are necessary for optimising the performance of

Table 6
Commercially available boron selective resins.
Manufacturer

Product name

Polymer structure

Functional groups

Mean diameter (m)

Total capacity (eq/L)

Amberlite

PWA10
IRA743
CRB01
CRB02
BSR-1
S108
S110

Macroporous
Macroporous
Macroporous
Macroporous
Macroporous
Macroporous
Macroporous

N-Methylglucamine
N-Methylglucamine
N-Methylglucamine
N-Methylglucamine
N-Methylglucamine
N-Methylglucamine
N-Methylglucamine

3001200
500700
3001200
3001200
550 50
650 70
600 100

0.7
0.7
1.2
0.9
0.7
0.6
0.8

Diaion
Dowex
Purolite

polystyrene
polystyrene
polystyrene-DVB
polystyrene-DVB
polystyrene-DVB
polystyrene-DVB
polystyrene-DVB

30

N. Hilal et al. / Desalination 273 (2011) 2335

the resin. The efciency of boron removal from aqueous solution is


encountered with increase in the resin/solution ratio. Because the
amount of absorbates is limited in treated solution, optimisation of
resin amount is used to prevent the use of unnecessary amounts of BSR
[74,75].
The optimum amount of resin required to remove boron to the target
point depends on the boron uptake capacity of resin at equilibrium.
The capacity varies with the change of initial boron concentration in
solution and the resin used. From the isotherm study carried out by
Simonnot et al. [76] using Amberlite IRA743, the capacity of boron
uptake of resin improves to 1.87 g B/bed with the increase in the initial
boron concentration in solution to 20 mg B/L. Above 20 mg B/L boron
uptake increases to reach about 2.3 g B/L and then remains constant. It
was mentioned that the isotherm data should not be interpreted in
terms of the Langmuir isotherm, because boron uptake by complexation
is not a reversible adsorption process. Ozturk and Kose [77] reported
that the amount of boron absorbed (mg B/g resin) increased as the
concentration of boron in solution rose. The amount of boron absorbed
at equilibrium increased from 5 to 15.5 mg B/g resin with the increment
of initial boron concentration from 100 to 1000 mg B/L. It was noted that
the empirical Langmuir isotherm can adequately describe the equilibrium relation between resin and liquid phases of the ion exchange
process.
In regard to sorption kinetic studies, several different kinetic models
can be used, which are as follows: Lagergren pseudo-rst and pseudosecond-order, Elovich equation and parabolic diffusion model, etc.
Kabay et al. [78,79] investigated the kinetics of two boron selective
resins, Diaion CRB02 and Dowex-XUS 43594.00. The obtained data from
batch sorption tests were tted to sorption kinetics by using pseudorst-order equation and pseudo-second-order equation. From the
results, it was found that sorption kinetics for Diaion CRB02 and
Dowex-XUX 43594.00 t well to the Ho pseudo-second-order kinetics
model. Boncukcuoglu et al. [74] carried out kinetic studies of boron
removal by Amberlite IRA 743. They predicted the process kinetics by
adopting heterogeneous uidsolid reaction models and obtained an
empirical kinetic model derived from a second-order pseudo-homogeneous reaction model.
The mechanism of ion exchange in heterogeneous systems, consists
of several steps, namely, solution diffusion, lm diffusion, and particle
diffusion, which are shown in Fig. 6 [80]. The rate of ion exchange
depends on the rate-determining step, which is dened by rate of the
slowest step. As it provides useful information on rate parameters for
design purposes, it is important to determine the rate-determining step
[77]. Generally, for ion exchange processes, lm diffusion or particle
diffusion is considered as the rate-determining step because the rate of

solution diffusion is able to be assisted by hydrodynamic turbulences


(e.g. by stirring) [80]. According to Levenspiel [81], the lm diffusion
controlled process is directly dependent on the concentration of the
solution, while the particle diffusion controlled process is independent.
In order to determine the rate-determining step of boron removal
processes, there has been a number of studies. Boncukcuoglu et al. [74]
investigated the ion exchange kinetics of the removal of the boron by
using various reaction parameters and established an empirical model
to describe the boron removal process by using Amberlite IRA 743. They
concluded that the rate-determining step of the process is lm diffusion
as the direct effect of concentration to the process is found from the
empirical model developed. Yan et al. [82] carried out the batch sorption
tests of boron by using XSC-800. In order to determine the ratedetermining step of the process, the experimental data obtained by
batch sorption tests were tted to the moving boundary model. It was
found that the moving boundary model could not satisfactorily describe
the process. From the modied model by introducing a time correction
factor to express the process more accurately, it was found that the
process is controlled by particle diffusion.
Ozturk and Kose [77] examined the boron removal process from
aqueous solutions using Dowex 2 8 anion exchange resin. The generalised ion exchange model was applied to t kinetic data obtained at
three different temperatures. From the results, it was conrmed that
the sorption process is controlled by a combination of lm and particle
diffusion. Meanwhile, it was reported that the rate-determining step
of Dowex-XUS 43594.00 and Diaion CRB02 is conrmed as lm
diffusion and particle diffusion respectively, according to the innite
solution volume model (ISV) [78].
From the studies for the determination of the sorption controlling
step, it seems that the rate-determining step can be varied by the type of
resins used and the experimental conditions. This is because the mass
transfer in the exchanger phase depends on the physico-chemical
properties of the resins used [80]. Furthermore, the mass transfer
through the lm can be changed with the alteration of agitation, particle
size, and the degree of cross-linking functional groups.
2.2.4. Effects of each parameter on boron removal
The removal of boron by BSRs depends on several process
parameters as well as the type of resins used. Boncukcuoglu et al. [74]
investigated the effects of several experimental parameters on the
performance of boron removal by batch sorption operation using
Amberlite IRA 743. The parameters selected were the ratio of resin/
solution, initial boron concentration, pH, stirring speed and temperature. In regard to the effect of the ratio of resin/solution, it was found that
the rate and performance of boron removal improved with the increase

Fig. 6. General mechanism of the ion exchange process [80]. (1) dissociation of the dissolved complexes containing rst ion; (2) diffusion of the rst ion from solution towards the
inter-phase lm; (3) diffusion of the rst ion through the inter-phase lm; (4) diffusion of the rst ion inside the material phase; (5) association between the rst ion and functional
group; (6) dissociation of the associates between the second ion and functional group; (7) diffusion of the second ion inside the material phase towards the surface; (8) diffusion of
the second ion through the inter-phase lm; (9) diffusion and random distribution of the second ion in the solution; (10) formation of the second ion complexes in the solution.

N. Hilal et al. / Desalination 273 (2011) 2335

in the proportion of resin. Similarly, the increase in temperature showed


the effect of stimulating reaction rate and slightly increasing % boron
removal. However, the opposite result was obtained by Ozturk and Kose
[77]. They investigated boron removal from aqueous solution of 600 mg
B/L concentration by using Dowex 2 8. Their work, at temperatures
from 25 to 45 C concluded that boron rejection decreased as the temperature increased.
As far as the effect of pH is concerned, it was reported that the
removal of boron is optimised at around pH 99.5 [74,77]. The size of
resin is also an important process parameter. As resin size decreases
the rate of boron removal will increase. Kabay et al. [79] reported that
a high percent removal of boron was achieved when the particle size
decreased due to the improvement of diffusion rate with increased
surface area. In regard to initial concentration of boron and stirring
speed, the removal of boron decreased as the initial boron
concentration increased and there was no signicant effect found
when stirring speed was changed from 350 to 750 rpm [74]. The
independence of boron removal on the speed of stirring was
conrmed by Yan et al. [82].
The studies to discover the effect of experimental parameters on
boron removal have also been performed in column mode operation.
In column mode operation, breakthrough capacity is considered as the
key factor indicating the performance of a boron removal process as it
is directly related to the boron uptake capacity of resins. The results of
several researches are tabulated in Table 7 and show an agreement in
the effect of ow rate and the concentration of boron in solution.
However there is disagreement in the results obtained for the effects
of pH and ionic strength and this is discussed below.
In regard to the effect of ow rate, it was found that the
breakthrough capacity reduces as the ow rate increases due to the
decrease in the contact time between resin and solution containing
boron [79,82]. Similarly, the improvement of the breakthrough
capacity can be achieved by increasing the Height/Diameter (H/D)
ratio due to the increase in contact time and the enhancement of
liquid distribution in the column [82]. The breakthrough of boron in
the eluent from feed solutions with a high boron concentration is due
to limitations associated with the boron uptake capacity of the resin
and the amount of resin used under experimental conditions.
The effect of increasing ionic strength was reported by Yan et al.
[82] to decrease boron rejection while Simonnot et al. [76] found that
there was no signicant effect on the removal of boron. Yan et al. [82]
carried out column tests by adding 1.252.5 g NaCl/mg B at pH 10
while Simonnot et al. [76] used 00.292 g NaCl/mg B to increase ionic
strength at pH ranges of 5 to 5.5. Thus, the disagreement may be
explained as a different experimental condition in pH or the amount

31

of NaCl added. The difference of the boron selectivity between


Amberlite IRA 743 and XSC-800 could also be considered as the cause.

2.2.5. Elution and regeneration studies


Huge amounts of chemicals are used to regenerate the saturated
BSRs used in the boron removal process. This is problematic in terms
of not only the volumes of reagents used, but also the cost incurred. To
illustrate this, it may be necessary to use 4 Bed Volumes (BV) of
0.25 M HCl and 4 BV of 0.25 M NaOH. For a system having a 10 m3 bed,
suitable for treating 3800 m3 of water at 5 mg B/L, this means that
40 m3 of HCl and 40 m3 of NaOH are required. Thus, a key point for the
implementation of the boron removal by BSRs is to minimise the
regenerant volumes required [76].
Generally, regeneration process consists of two main steps,
deboronation using acid regenerants, such as HCl and H2SO4, and
neutralisation using basic regenerant, typically, NaOH. Nadav [21]
compared the breakthrough capacity of the boron removal process
after regeneration by three different methods. It was concluded that
regeneration with H2SO4 followed by NaOH gives an increased mg B/g
resin absorption capacity in comparison to regeneration with H2SO4
alone. This can be explained as the gross uniformity of the chemical
potential of the resin is improved by NaOH.
Simonnot et al. [76] performed tests for the column regeneration by
using different types of acidic and basic regenerants. It was found that
deboronation was achieved by a weak acid, CH3COOH, but it needed 50
BV. A regenerant acid can be reused partially or totally to reduce the
consumption of acid. The elution performance of reused acid decreased
from 94% to 78% as the molar fraction of boric acid in the regenerant
solution increased from 0.1 to 0.5. It was concluded that the choice of
base was less important as its function is for neutralisation only.
Kabay et al. [83] carried out recycle tests with a boron selective resin,
Diaion CRB02. The resin was tested for 10 sorptionwashingelution
washingregenerationwashing cycles in geothermal eld. In the
research 5% H2SO4 and 4% NaOH were used for elution and regeneration
respectively. It was observed that the breakthrough capacity of resin
decreased slightly between cycles 2 and 7, and reached a plateau between cycles 7 and 10. The capacity reduction could be considered due
to degradation caused by the inuence of high temperature, around
70 C, applied during the whole process. The similar recycle tests were
done by Kose and Ozturk [84] using a strong base anion exchange resin,
Dowex 2 8. The boron on the resin was eluted with 0.5 M HCl and
regenerated with 2 N NaOH. From the 3 cycle tests, it was observed that
the boron capacity was increased after the rst cycle while there was no
noticeable change after the 2nd and 3rd cycles.

Table 7
The effects of each parameter on breakthrough capacity of BSRs.
Parameters

Boron rejection References Used resins

pH increases

No effects

[76]

Increases

[82]

Flow rate increases

Decreases

Feed temperature increases

Increases

[79]
[82]
[77]
[82]

Initial boron concentration increases Decreases


Ionic strength increases

No effects
Decreases

[76]
[82]
[76]
[82]

The ratio of H/D increases

Increases

[82]

Amberlite IRA743

Comments

pH 5.58, concentration of boron 20 mg B/L, ionic strength


00.1 mol NaCl/L
XSC-800
pH 412, concentration of boron 40 mg B/L, ow rate
16 BV/h, H/D ratio of 15
Diaion CRB02 Dowex XUS 43594.00 15 and 20 BV/h, concentration of boron 1.51.6 mg B/L
XSC-800
5, 16, and 30 BV/h, concentration of boron 40 mg/L, 293 K, pH 10
Dowex 28
39 and 45 mL/h, concentration of boron 600 mg/L, pH 5.8
XSC-800
293323 K, concentration of boron 40 mg/L, ow rate 16 BV/h,
pH 10, H/D ratio 15
Amberlite IRA743
15 mg B/L and 5500 mg B/L
XSC-800
40, 50, and 100 mg B/L, ow rate 16 BV/h, 293 K, pH 10, H/D ratio 15
Amberlite IRA743
00.1 mol NaCl/L, concentration of boron 20 mg B/L, pH 5.58
XSC-800
50, 75, and 100 g NaCl/L, concentration of boron 40 mg B/L, Flow rate
16 BV/h, pH 10, H/D ratio 15
XSC-800
10, 12, and 15 H/D, concentration of boron 40 mg B/L, Flow rate
16 BV/h, pH 10

32

N. Hilal et al. / Desalination 273 (2011) 2335

Fig. 7. Flow sheet of (a) a conventional xed bed column and (b) the integrated system [85].

2.3. Adsorption membrane ltration (AMF)


2.3.1. Basic principles
Adsorption membrane ltration (AMF) process has been considered
as an alternative separation process. The concept of the hybrid process is
to combine a sorption process with membrane separation. In this hybrid
process, solutes are absorbed by sorbents followed by a membrane
separation of the saturated sorbents [85]. Fig. 7 shows the ow sheet of
the integrated process in comparison to a conventional sorption process.
The major advantage of the hybrid process over a conventional
sorption process is the possibility of using ne sorbents. As discussed in
the previous chapter, the kinetic performance of sorbents and the boron
uptake are improved as the size of sorbents decreases. Consequently it
is possible to reduce the amount of sorbent required by using ne
sorbents, which leads to a reduction of sorbent cost. For membranesorption integrated processes, the pressure drop which results in high
energy costs is independent on particle size but is a function of the
membrane resistance [85]. That is, it is possible to use a ne particle size
and improve the process performance and to decrease the amount of
sorbents required without a pressure drop.
On the other hand, for a conventional sorption process with a xed
bed column, it is impracticable to use ne particles as sorbents due to
the resulting signicant pressure drop. According to KozenyCarman
equation Eq. 5 which is used to calculate the pressure drop of the uid
through a packed bed of solids, the pressure drop in a xed bed is
inversely proportional to the square of the particle size.
P =

150V0 L12
:
2s D2p 3

Fig. 8. Flow sheet of the AMF system.

Where P is the pressure drop, L is the total height of the bed, V0 is


the supercial velocity, is the viscosity of the uid, is the porosity of
the bed, s is the sphericity of the particles in the packed bed, and Dp
is the sorbent particle diameter having a spherical shape.
Koltuniewicz et al. [85] compared the total operation costs of the
hybrid system in comparison to two conventional packed bed systems
using sorbents of 1 mm and 1 m, respectively. It was found that the
operating cost of the hybrid system is much cheaper than conventional
systems due to the decreased quantity of sorbents required and the
smaller pressure drop encountered.
2.3.2. Adsorption membrane ltration for the removal of boron
Kabay et al. [86] adopted the hybrid separation concept for the
removal of boron from aqueous solutions and examined the suitability
and performance of the hybrid process. As it is shown in Fig. 8, the
whole process suggested in the study is divided into ve stages, which
are as follows.

Boron sorption by BSRs


Membrane separation of saturated BSR from aqueous solution
Boron desorption using acid
Membrane separation of fresh BSR from BSR suspension
Regeneration of BSR using base.

For the experiment, in the rst stage, Diaion CRB02 and Dowex
XUS-43594.00 was employed as binding materials while in the second
stage a at-type hydrophobic Teon-Fluoropore Membrane with a
pore diameter of 0.2 m and thickness of 200 m was used as a
ltration material. It was found that boron sorption of both ground
BSRs (50 m) is rapid enough to use for the hybrid process. CRB02
needed 5 min to decrease boron concentration to half the initial
concentration while XUS-43594.00 took 30 min. As the results of
recycle tests, it was reported that the performance of both BSRs was
improved after ve cycles due to the regeneration step with NaOH
solution that gives a better access to functional groups.
Similar experiments were done by Yilmaz et al. [87]. In this study,
submerged hollow bre polypropylene membranes with 0.4 m pore
were used as ltration materials. As a submerged membrane system
requires only a low suction pressure, a lower energy is needed for its
operation in comparison to that for a cross ow microltration system
[88]. From the study, it was found that CRB02 is more suitable for AMF
system than XUS-43594.00 as it shows much faster boron sorption
kinetics as well as less time to achieve steady state conditions.
Blahusiak et al. [89] examined the suitability of using submerged
membranes for the suspension of high BSR (Dowex XUS-43594.00)
concentration. In terms of the rheological properties of the XUS-43594

N. Hilal et al. / Desalination 273 (2011) 2335

suspensions, it was suitable at 11% w/v of dry resins as the suspension


had low viscosity. For the experiments, hydrophilic capillary and hydrophobic polypropylene hollow bre membranes were prepared. With
the increase in trans-membrane pressure difference, the permeate ux
was increased with a slight loss at lower suspension concentrations,
below 7% w/v for a hydrophilic membrane and 4% w/v for a hydrophobic
membrane. However, a signicant drop of permeate ux was observed
at higher concentration. From the concentration dependence of the ux
data obtained, it was concluded that hydrophilic membranes are more
favourable for the AMF process.
After the desorption stage shown in Fig. 8, fresh BSRs are separated
from boron containing water and recycled ready for reuse in the rst
stage. The volumetric ow rate after desorption is relatively small, but
the concentration of suspension is high. Furthermore, pH is lowered
by the addition of acid. Thus, it is required to nd an appropriate
membrane with high acid resistance and to evaluate the performance
at high suspension concentration conditions. Onderkova et al. [90]
investigated a tubular ceramic microltration membrane Membralox
with an inner diameter of 7 mm and a length of 0.25 m. The decrease
in the permeate ux was moderate in the ranges of 416% w/v.
However, it was found that the permeate ux is nearly independent at
the point higher than 20% w/v. It was considered that the ceramic
membrane is suitable for the AMF application as it showed acceptable
permeate ux even at 20% w/v.
2.3.3. Effects of each parameter on boron removal
There were researches to nd out the effect of each experimental
parameter on the performance of hybrid process. Bryjak et al. [91]
developed mathematical modelling and investigated the effects of
several parameters such as, feed ow rate, BSR suspension concentration, and rate of suspension replacement on the process efciency.
The mathematical model was derived from the boron balance of the
whole system. The simplied model obtained is shown in Eq. (6).
J
Cp B = Cf B s X outqout
Jf

where Cp(B) is the boron concentration in permeate, Cf(B) it the initial


boron concentration, Js is the ux of suspension replacement, Jf is the
feed ux, X(out) is the concentration of solid in the suspension
replaced, and q(out) is the mass of boron in the suspension replaced.
It was found that Eq. (6) explains well the effect of process parameters. That is, high boron removal is achieved when feed ux ( Jf) is
low, and both the ux of suspension replacement ( Js) and the
concentration of BSR (X(out)) are high. Similar results were obtained
during 3 h and 48 h operation [92]. To achieve high boron removal, it
was found that the use of a more concentrated resin suspension is more
effective than increasing the rate of suspension replacement. Furthermore, it was mentioned that the use of more concentrated suspension
creates some kind of buffer preventing the AMF system malfunctioning.
Kabay et al. [93] investigated the effects of each parameter to
optimise the AMF system for boron removal from geothermal water
containing a high level of boron (89 mg B/L). Diaion CRB02 and
Dowex XUS-43594.00 were used for a short period of operation (2 h)
and a longer period of operation (24 h) respectively. In terms of the
effect of feed ow rate, resin concentration and resin particle size,
similar results were found. This means the effect of these parameters
is virtually the same regardless of the BSR used with boron concentration in the range of 29 ppm.
Blahusiak and Schlosser [94] performed a simulation and optimisation of the AMF process for boron removal from RO permeate based
on experimental data. It was found that dialtration, using permeate
of submerged microltration, is a critical factor affecting adsorbent
ow rate in the suspension recirculation loop and in the consumption
of chemicals during the desorption step. In the range of conditions
investigated, the adsorbent ow rate in the suspension recirculation

33

loop, specic consumption of acid, and membrane surface area of


submerged membrane were signicantly decreased as the dialtration ow rate increases up to 0.5 m3/h.
3. Future research
As the simplest method for reducing boron concentration from
seawater associated with the reduction of salinity, RO methods are
expected to have a great potential. A lot of efforts have been made on
optimising processes by changing process parameters or/and congurations. However, the main drawback of this method is the need of multipass RO systems to reduce boron level to the current recommended value
for boron, which causes the increase in costs. In order to render this
method more attractive, 1 pass RO system should be developed with
acceptable reduction efciency. Consequently, it is required to synthesise
a RO membrane having an ability of high boron rejection with lower
manufacturing costs.
Boron removal by ion exchange with BSRs shows the highest performance with conventional column mode operation. However, a large
amount of acid and base as a part of the BSRs regeneration process is
required. It is necessary to nd an effective regeneration process with
low consumption of chemicals, as well as an increasing sorption
capability, in order to make BSR processes a more favourable option
for the removal of boron. In regard to process parameters, an agreement
has been made on the effect of each process parameter from various
research. However, there is no comparison study yet between BSRs
under the same experimental conditions. For optimising a specic
process at certain process conditions, it is critical to have comparison
data on each BSR.
Recently, the AMF process has been demonstrated as an economical
treatment of RO permeate to remove boron with high performance.
Although it has a great potential for the future use, it has not been fully
developed, and therefore requires intensive further works. A special
attention should be paid to the membrane separation part in tubular and
submerged modules adopted. Characterisation of fouling property and
development of adequate cleaning protocols of each membrane are
important to optimise the process. Furthermore, for prolonged operation of the system without membrane fouling and degradation of
BSRs, it is required to characterise physical properties of suspensions
and material stability for long term operation.
4. Conclusions
Boron is an essential micronutrient for plants and animals, and also a
useful component for numerous commercial activities, including glass
industry and semiconductor manufacture. However, excess boron can
cause toxic effects in both plants and animals. Irrigation water from RO
desalination plants often contains higher boron content, which is
problematic for cultivation in the MENA region. As the deposition of
boron in both soils and plants is accelerated by irrigation water
possessing a large amount of boron, it is necessary to produce low boron
containing water from RO desalination plants to prevent toxic effects of
boron in plants.
In this paper, a comprehensive technical overview of boron removal
has been given on (1) reverse osmosis, (2) ion exchange with BSRs, and
(3) adsorption membrane ltration. The processes discussed all showed
degrees of suitability for the removal of boron from saline water to
comply with the current WHO recommendation.
Researches on the removal of boron by RO have focused on 1) system
congurations and 2) experimental parameters to nd the optimum
operating conditions. From the several studies concerning system
congurations, it can be concluded that multi pass RO systems are
required to reduce a boron level to the current recommended value for
boron. For the experimental parameters, increasing pH is shown to be
the governing parameter on the performance of boron removal by RO

34

N. Hilal et al. / Desalination 273 (2011) 2335

regardless of the RO membrane used. The performance of boron removal


is also improved at conditions of low temperature and low salinity.
Studies on boron removal by the use of BSRs have mainly focused on
nding optimised experimental parameters, and kinetic characteristics.
From the studies on experimental parameters and kinetics, it can be
concluded that contact time between BSRs and solution is the crucial
factor governing the performance of processes, while the surface area of
BSRs governs sorption kinetics. That is, due to the increase in contact
time, the removal of boron improves with the increase in H/D ratio and
the decrease in ow rate. Similarly, due to the increase in the surface
area of smaller size BSRs, the sorption rate is faster.
The AMF process has been identied as a promising boron removal
method with high performance and low operation cost. The main
advantage of AMF is the possibility of the use of ne BSR particles. As
mentioned previously, this improves the ability of boron uptake of BSRs,
which in turn leads to the reduction of the required amount of resin and
consequently a reduction in cost. Recently a laboratory scale system was
tested using Dowex XUS-43594.00 resin and the performance of boron
removal showed encouraging results.

References
[1] V. Morgan, Boron chemistry, in: W. Mellor (Ed.), Supplement to Mellor's
Comprehensive Treatise on Inorganic and Theoretical Chemistry, Longman,
New York, 1980.
[2] H. Aubert, M. Pinta, Trace elements in soils, Elsevier Scientic, Amsterdam, 1977.
[3] K.B. Krauskopf, Geochemistry of micronutrients, in: J.J. Mortvedt, P.M. Giordano,
W.L. Lindsay (Eds.), Micronutrients in Agriculture, Soil Science Society of America,
Madison, 1972.
[4] D.J. Swaine, The trace-element content of soils, 1955, , 1955.
[5] P. Argust, Distribution of boron in the environment, Biol. Trace Elem. Res. 66
(1998) 131143.
[6] WHO, Boron in drinking water, , 2003.
[7] P.H. Brown, N. Bellaloui, M.A. Wimmer, E.S. Bassil, J. Ruiz, H. Hu, H. Pfeffer, F.
Dannel, V. Romheld, Boron in plant biology, Plant Biol. 4 (2002) 205223.
[8] R. Reid, Update on boron toxicity and tolerance in plants, in: F. Xu, H.E. Goldbach, P.H.
Brown, R.W. Bell, T. Fujiwara, C.D. Hunt, S. Goldberg, L. Shi (Eds.), Advances in plant
and animal boron nutrition: proceedings of the 3 rd International symposium on all
aspects of plant and animal boron nutrition, Springer, Dordrecht, 2007.
[9] D.G. Blevins, K.M. Lukaszewski, Boron in plant structure and function, Annu. Rev.
Plant Physiol. Plant Mol. Biol. 49 (1998) 481500.
[10] L. Bolanos, K. Lukaszewski, I. Bonilla, D. Blevins, Why boron? Plant Physiol.
Biochem. 42 (2004) 907912.
[11] H.E. Goldbach, M.A. Wimmer, Boron in plants and animals: is there a role beyond
cell-wall structure? J. Plant Nutr. Soil Sci.-Z. Panzenernahrung Bodenkunde 170
(2007) 3948.
[12] C.D. Hunt, Dietary boron: an overview of the evidence for its role in immune
function, J. Trace Elem. Exp. Med. 16 (2003) 291306.
[13] F.H. Nielsen, B.J. Stoecker, J.G. Penland, Boron as a dietary factor for bone
microarchitecture and central nervous system function, in: F. Xu, H.E. Goldbach, P.
H. Brown, R.W. Bell, T. Fujiwara, C.D. Hunt, S. Goldberg, L. Shi (Eds.), Advances in
plant and animal boron nutrition: proceedings of the 3 rd International symposium
on all aspects of plant and animal boron nutrition, Springer, Dordrecht, 2007.
[14] D.J. Fort, E.L. Stover, P.L. Strong, F.J. Murray, C.L. Keen, Chronic feeding of a low
boron diet adversely affects reproduction and development in Xenopus laevis,
J. Nutr. 129 (1999) 20552060.
[15] C.D. Hunt, F.H. Nielsen, Interaction between boron and cholecalciferol in the chick,
in: J.M. Gawthorne, J. McHowell, C.L. White (Eds.), Trace element metabolism in
man and animals: proceedings of TEMA-4, Springer, Berlin, 1982.
[16] Y.S. Bai, C.D. Hunt, Dietary boron enhances efcacy of cholecalciferol in broiler
chicks, J. Trace Elem. Exp. Med. 9 (1996) 117132.
[17] J.W. Spears, T.A. Armstrong, Dietary boron: evidence for a role in immune
function, in: F. Xu, H.E. Goldbach, P.H. Brown, R.W. Bell, T. Fujiwara, C.D. Hunt, S.
Goldberg, L. Shi (Eds.), Advances in plant and animal boron nutrition: proceedings
of the 3 rd International symposium on all aspects of plant and animal boron
nutrition, Springer, Dordrecht, 2007.
[18] R. Adair, Boron, Rosen Publishing Group, New York, 2007.
[19] J.J. Duderstadt, L.J. Hamilton, Nuclear reactor analysis, Wiley, New York, 1976.
[20] S.V. Eaton, Effects of boron deciency and excess on plants, Plant Physiol. 15
(1940) 95107.
[21] N. Nadav, Boron removal from seawater reverse osmosis permeate utilizing
selective ion exchange resin, Desalination 124 (1999) 131135.
[22] E.V. Maas, Crop salt tolerance, in: K.K. Tanji (Ed.), Salinity Assessment and
Management, Amer Society of Civil Engineers, New York, 1990.
[23] R.J. Weir, R.S. Fisher, Toxicologic studies on borax and boric-acid, Toxicol. Appl.
Pharmacol. 23 (1972) 351364.
[24] D.A. Roe, D.B. McCormic, R.T. Lin, Effects of riboavin on boric-acid toxicity,
J. Pharm. Sci. 61 (1972) 10811085.

[25] N.Y. Tarasenko, A.A. Kasparov, O.M. Strongina, Effect of boric-acid on the
generative function in males, Gig. Tr. Prof. Zabol. 16 (1972) 1316.
[26] I.P. Lee, R.J. Sherins, R.L. Dixon, Evidence for induction of germinal aplasia in male
rats by environmental exposure to boron, Toxicol. Appl. Pharmacol. 45 (1978)
577590.
[27] National Toxicology Program, Toxicology and carcinogenesis studies of boric acid
in B6C3F1 mice, in, 1987.
[28] W.W. Ku, R.E. Chapin, R.F. Moseman, R.E. Brink, K.D. Pierce, K.Y. Adams, Tissue
disposition of boron in male Fischer rats, Toxicol. Appl. Pharmacol. 111 (1991) 145151.
[29] H.T. El-Dessouky, H.M. Ettouney, Fundamentals of Salt Water Desalination,
Elsevier Science, Amsterdam, 2002.
[30] ESCAP UN, Statistical yearbook for Asia and the Pacic 2008, in, 2008.
[31] R. Reid, Can we really increase yields by making crop plants tolerant to boron
toxicity? Plant Sci. 178 (2010) 911.
[32] D. Prats, M.F. Chillon-Arias, M. Rodriguez-Pastor, Analysis of the inuence of pH
and pressure on the elimination of boron in reverse osmosis, Desalination 128
(2000) 269273.
[33] A. Sagiv, R. Semiat, Analysis of parameters affecting boron permeation through
reverse osmosis membranes, J. Membr. Sci. 243 (2004) 7987.
[34] IDA, DesalData, Desalination in 2008 global market snapshot, in, 2008.
[35] WHO, Boron in drinking water background document for development of
WHO guidelines for drinking water quality, in, 2009.
[36] SASO, Bottled drinking water, in, Saudi Arabian Standards Organization, 2000.
[37] SASO, Unbottled drinking water, in, Saudi Arabian Standards Organization, 2000.
[38] USEPA, 2006 Edition of the drinking water standards and health advisories, in,
Washington, DC, 2006.
[39] USEPA, Drinking water health advisory for boron, in, 2008.
[40] EEA, The quality of water intended for human consumption, in, Council Directive,
98/83/EC, 1998.
[41] Ministry of Environment (Republic of Korea, Management of Drinking Water
Quality, 2009.
[42] NIPH, Seawater Desalination Facility on Okinawa in, 2006.
[43] Ministry of Health, Drinking-water Standards for New Zealand 2005, Ministry of
Health, Wellington, 2005.
[44] MEP, Business licensing regulations salt concentrations in industrial sewage, in,
2003.
[45] NHMRC, Draft Australian drinking water guidelines, in, National Health and
Medical Research Council, 2009.
[46] NHMRC, Australian drinking water guidelines 6, in, National Health and Medical
Research Council, Canberra, 2004.
[47] CDW, Guidelines for canadian drinking water quality summary table in, 2008.
[48] WHO, Guidelines for Drinking-Water Quality, 2nd editionWorld Health Organization, Geneva, 1993.
[49] P.P. Power, W.G. Woods, The chemistry of boron and its speciation in plants, Plant
Soil 193 (1997) 113.
[50] C.M. Su, D.L. Suarez, Coordination of adsorbed boron a FTIR spectrosopic study,
Environ. Sci. Technol. 29 (1995) 302311.
[51] R.E. Zeebe, A. Sanyal, J.D. Ortiz, D.A. Wolf-Gladrow, A theoretical study of the kinetics
of the boric acidborate equilibrium in seawater, Mar. Chem. 73 (2001) 113124.
[52] A.G. Dickson, Themodynamics of the dissociation of boric-acid in synthetic seawater
from 273.15K to 218.15K, Deep-Sea Res. A-Oceanogr. Res. Pap. 37 (1990) 755766.
[53] R.N. Roy, L.N. Roy, M. Lawson, K.M. Vogel, C.P. Moore, W. Davis, F.J. Millero,
Thermodynamics of the dissociation of boric-acid in seawater at S = 35 from 0 C
to 55 C, Mar. Chem. 44 (1993) 243248.
[54] C. Culberson, R.m. Pytkowicz, Effect of pressure on carbonic acid, boric acid and pH
in seawater, Limnol. Oceanogr. 13 (1968) 403417.
[55] P. Glueckstern, M. Priel, Optimization of boron removal in old and new SWRO
systems, Desalination 156 (2003) 219228.
[56] H. Koseoglu, N. Kabay, M. Yuksel, S. Sarp, O. Arar, M. Kitis, Boron removal from
seawater using high rejection SWRO membranes impact of pH, feed concentration,
pressure, and cross-ow velocity, Desalination 227 (2008) 253263.
[57] J. Redondo, M. Busch, J.P. De Witte, Boron removal from seawater using FILMTEC
(TM) high rejection SWRO membranes, Desalination 156 (2003) 229238.
[58] M. Taniguchi, Y. Fusaoka, T. Nishikawa, M. Kurihara, Boron removal in RO seawater
desalination, Desalination 167 (2004) 419426.
[59] M. Faigon, D. Hefer, Boron rejection in SWRO at high pH conditions versus cascade
design, Desalination 223 (2008) 1016.
[60] N. Nadav, M. Priel, P. Glueckstern, Boron removal from the permeate of a large
SWRO plant in Eilat, Desalination 185 (2005) 121129.
[61] P.V.X. Hung, S.H. Cho, S.H. Moon, Prediction of boron transport through seawater
reverse osmosis membranes using solution-diffusion model, Desalination 247
(2009) 3344.
[62] Y. Cengeloglu, G. Arslan, A. Tor, I. Kocak, N. Dursun, Removal of boron from water
by using reverse osmosis, Sep. Purif. Technol. 64 (2008) 141146.
[63] J.J. Qin, M.H. Oo, M.N. Wai, Y.M. Cao, Enhancement of boron removal in treatment of
spent rinse from a plating operation using RO, Desalination 172 (2005) 151156.
[64] N. Geffen, R. Semiat, M.S. Eisen, Y. Balazs, I. Katz, C.G. Dosoretz, Boron removal from
water by complexation to polyol compounds, J. Membr. Sci. 286 (2006) 4551.
[65] M.H. Oo, L.F. Song, Effect of pH and ionic strength on boron removal by RO
membranes, Desalination 246 (2009) 605612.
[66] H. Hyung, J.H. Kim, A mechanistic study on boron rejection by sea water reverse
osmosis membranes, J. Membr. Sci. 286 (2006) 269278.
[67] Y. Magara, A. Tabata, M. Kohki, M. Kawasaki, M. Hirose, Development of boron
reduction system for sea water desalination, Desalination 118 (1998) 2533.
[68] C. Jacob, Seawater desalination: boron removal by ion exchange technology,
Desalination 205 (2007) 4752.

N. Hilal et al. / Desalination 273 (2011) 2335


[69] N. Bicak, M. Gazi, B.F. Senkal, Polymer supported amino bis-(cis-propan 2, 3 diol)
functions for removal of trace boron from water, React. & Funct. Polym. 65 (2005)
143148.
[70] N. Bicak, N. Bulutcu, B.F. Senkal, M. Gazi, Modication of crosslinked glycidyl
methacrylate-based polymers for boron-specic column extraction, React. &
Funct. Polym. 47 (2001) 175184.
[71] B.F. Senkal, N. Bicak, Polymer supported iminodipropylene glycol functions for
removal of boron, React. & Funct. Polym. 55 (2003) 2733.
[72] L. Wang, T. Qi, Z. Gao, Y. Zhang, J. Chu, Synthesis of N-methylglucamine modied
macroporous poly(GMA-co-TRIM) and its performance as a boron sorbent, React.
& Funct. Polym. 67 (2007) 202209.
[73] H. Parschova, E. Mistova, Z. Matejka, L. Jelinek, N. Kabay, P. Kauppinen, Comparison of
several polymeric sorbents for selective boron removal from reverse osmosis
permeate, React. & Funct. Polym. 67 (2007) 16221627.
[74] R. Boncukcuoglu, A.E. Yilmaz, M.M. Kocakerim, M. Copur, An empirical model for
kinetics of boron removal from boron-containing wastewaters by ion exchange in
a batch reactor, Desalination 160 (2004) 159166.
[75] N. Kabay, I. Yilmaz, S. Yamac, S. Samatya, M. Yuksel, U. Yuksel, M. Arda, M. Saglam,
T. Iwanaga, K. Hirowatari, Removal and recovery of boron from geothermal
wastewater by selective ion exchange resins. I. Laboratory tests, React. & Funct.
Polym. 60 (2004) 163170.
[76] M.O. Simonnot, C. Castel, M. Nicolai, C. Rosin, M. Sardin, H. Jauffret, Boron removal
from drinking water with a boron selective resin: is the treatment really selective?
Water Res. 34 (2000) 109116.
[77] N. Ozturk, T.E. Kose, Boron removal from aqueous solutions by ion-exchange
resin: batch studies, Desalination 227 (2008) 233240.
[78] N. Kabay, S. Sarp, M. Yuksel, O. Arar, M. Bryjak, Removal of boron from seawater
by selective ion exchange resins, React. & Funct. Polym. 67 (2007) 16431650.
[79] N. Kabay, S. Sarp, M. Yuksel, M. Kitis, H. Koseoglu, O. Arar, M. Bryjak, R. Semiat,
Removal of boron from SWRO permeate by boron selective ion exchange resins
containing N-methyl glucamine groups, Desalination 223 (2008) 4956.
[80] A.A. Zagorodni, Ion exchange materials: properties and applications, Elsevier,
Amsterdam, London, 2007.
[81] O. Levenspiel, Chemical Reaction Engineering, 2nd ed, Wiley, New York, London,
1972.

35

[82] C.Y. Yan, W.T. Yi, P.H. Ma, X.C. Deng, F.Q. Li, Removal of boron from rened brine
by using selective ion exchange resins, J. Hazard. Mater. 154 (2008) 564571.
[83] N. Kabay, I. Yilmaz, S. Yamac, M. Yuksel, U. Yuksel, N. Yildirim, O. Aydogdu, T.
Iwanaga, K. Hirowatari, Removal and recovery of boron from geothermal wastewater
by selective ion-exchange resins II, Field Tests Desalination 167 (2004) 427438.
[84] T.E. Kose, N. Ozturk, Boron removal from aqueous solutions by ion-exchange
resin: column sorptionelution studies, J. Hazard. Mater. 152 (2008) 744749.
[85] A.B. Koltuniewicz, A. Witek, K. Bezak, Efciency of membrane-sorption integrated
processes, J. Membr. Sci. 239 (2004) 129141.
[86] N. Kabay, I. Yilmaz, M. Bryjak, M. Yuksel, Removal of boron from aqueous solutions
by a hybrid ion exchange-membrane process, Desalination 198 (2006) 158165.
[87] I. Yilmaz, N. Kabay, M. Brjyak, M. Yuksel, J. Wolska, A. Koltuniewicz, A submerged
membrane-ion-exchange hybrid process for boron removal, Desalination 198
(2006) 310315.
[88] W.S. Guo, W.G. Shim, S. Vigneswaran, H.H. Ngo, Effect of operating parameters in a
submerged membrane adsorption hybrid system: experiments and mathematical
modeling, J. Membr. Sci. 247 (2005) 6574.
[89] M. Blahusiak, B. Onderkova, S. Schlosser, J. Annus, Microltration of microparticulate boron adsorbent suspensions in submerged hollow bre and capillary
modules, Desalination 241 (2009) 138147.
[90] B. Onderkova, S. Schlosser, M. Blahusiak, M. Bugel, Microltration of suspensions
of microparticulate boron adsorbent through a ceramic membrane, Desalination
241 (2009) 148155.
[91] M. Bryjak, J. Wolska, N. Kabay, Removal of boron from seawater by adsorptionmembrane hybrid process: implementation and challenges, Desalination 223
(2008) 5762.
[92] M. Bryjak, J. Wolska, I. Soroko, N. Kabay, Adsorption-membrane ltration process
in boron removal from rst stage seawater RO permeate, Desalination 241 (2009)
127132.
[93] N. Kabay, I. Yilmaz-Ipek, I. Soroko, M. Makowski, O. Kirmizisakal, S. Yag, M. Bryjak,
M. Yuksel, Removal of boron from Balcova geothermal water by ion exchangemicroltration hybrid process, Desalination 241 (2009) 167173.
[94] M. Blahusiak, S. Schlosser, Simulation of the adsorption-microltration process for
boron removal from RO permeate, Desalination 241 (2009) 156166.

Você também pode gostar