Tải bản đầy đủ (.pdf) (6 trang)

DSpace at VNU: Sugarcane bagasse treated with hydrous ferric oxide as a potential adsorbent for the removal of As(V) from aqueous solutions

Bạn đang xem bản rút gọn của tài liệu. Xem và tải ngay bản đầy đủ của tài liệu tại đây (460.01 KB, 6 trang )

Food Chemistry 138 (2013) 133–138

Contents lists available at SciVerse ScienceDirect

Food Chemistry
journal homepage: www.elsevier.com/locate/foodchem

Sugarcane bagasse treated with hydrous ferric oxide as a potential adsorbent for
the removal of As(V) from aqueous solutions
E. Pehlivan a,⇑,1, H.T. Tran b, W.K.I. Ouédraogo c, C. Schmidt d, D. Zachmann d, M. Bahadir d
a

Department of Chemical Engineering, Selcuk University, Campus, 42079 Konya, Turkey
Hanoi University of Science, Hanoi, Vietnam
c
Laboratoire de Chimie Organique: Structure et Réactivité, UFR-SEA, Université de Ouagadougou, 03 BP 7021 Ouagadougou 03, Burkina Faso
d
Institute of Environmental and Sustainable Chemistry, Technische Universitaet Braunschweig, Germany
b

a r t i c l e

i n f o

Article history:
Received 16 January 2012
Received in revised form 1 September 2012
Accepted 24 September 2012
Available online 8 November 2012
Keywords:
Sugarcane bagasse


Adsorption
Arsenic
Iron oxide-based adsorbent

a b s t r a c t
The mechanism of As(V) removal from aqueous solutions by means of hydrated ferric oxide (HFO)-treated sugarcane bagasse (SCB-HFO) (Saccharum officinarum L.) was investigated. Effects of different parameters, such as pH value, initial arsenic concentration, adsorbent dosage, contact time and ionic strength,
on the As(V) adsorption were studied. The adsorption capacity of SCB-HFO for As(V) was found to be
22.1 mg/g under optimum conditions of pH 4, contact time 3 h and temperature 22 °C. Initial As(V) concentration influenced the removal efficiency of SCB-HFO. The desorption of As(V) from the adsorbent was
17% when using 30% HCl and 85% with 1 M NaOH solution. FTIR analyses evidenced two potential binding
sites associated with carboxyl and hydroxyl groups which are responsible for As(V) removal. Adsorption,
surface precipitation, ion exchange and complexation can be suggested as mechanisms for the As(V)
removal from the solution phase onto the surface of SCB-HFO.
Ó 2012 Elsevier Ltd. All rights reserved.

1. Introduction
Among various elemental species, arsenic is near the top of the
toxic list. Arsenic enters water bodies through both natural erosion
processes and anthropogenic activities. High levels of arsenic in
drinking water are a crucial problem in many countries, e.g. Mexico, Bangladesh, Vietnam, and Argentina (Farias et al., 2003; Fazal &
Kawaci, 2001).
Arsenic is released into the environment while using pesticides,
treating wood, producing glass and electronic devices, manufacturing copper and other metals, as well as producing fertilisers. Arsenic ions occur in surface and ground waters in both organic
and inorganic species, the inorganic forms being the predominant
ones, e.g. arsenite (H2AsO3À and arsenate (H2AsO4À) (Fazal & Kawaci, 2001).
While arsenic is toxic to plants and animals, inorganic arsenic
species are strong carcinogens to humans (Ng, 2005). Usually, arsenic is taken up and accumulated in the human body through
drinking water, the food chain, and inhalation of polluted air. The
human toxicity of arsenic ranges from skin lesions to cancer of
the brain, liver, kidney, and stomach. Arsenic intake causes disturbance of nervous system functions and can lead to death (Boddu,
⇑ Corresponding author. Tel.: +90 332 2232127; fax: +90 332 2410635.

E-mail address: (E. Pehlivan).
Present address: Department of Chemical Engineering, Selcuk University, Konya
42079, Turkey.
1

0308-8146/$ - see front matter Ó 2012 Elsevier Ltd. All rights reserved.
/>
Abburib, Talbottc, Smitha, & Haasch, 2008). Because of these effects, the World Health Organisation (WHO) and the United State
Environmental Protection Agency (USEPA) reduced the arsenic
standard concentration in drinking water from 50 to 10 mg/l (Environmental Protect Agency, 1999). If this concentration is exceeded
in surface and ground waters in many countries, it is essential to
develop effective methods for the removal of arsenic ions from
the water supply.
Arsenic removal, using low-cost adsorbents, such as lignocellulosic materials and agricultural by-products has been under investigation since the last decade. Agricultural by-products are of
particular interest since these materials are produced in great
amounts and are easily available worldwide. In Vietnam, sugarcane
industries produce large amounts of SCB that could be applied for
arsenic removal from water streams. It was reported that the main
components of SCB are cellulose (46.0%), hemicellulose (24.5%), lignin (19.95%), fat and waxes (3.5%), ash (2.4%), silica (2.0%), and others (1.7%) (Sene, Converti, Felipe, & Zilli, 2002). The polysaccharides
found in sugarcane bagasse are biopolymers with many hydroxyl
and/or phenolic groups that can be chemically modified to form
new compounds with various properties (Navarro, Sump, Fujii, &
Matsumura, 1996). The arsenic removal from water streams could
be achieved by using various techniques (Fe-electro-coagulation/
co-precipitation, coagulation–microfiltration, oxidation/precipitation, coagulation/precipitation, reverse osmosis, filtration, nanofiltration, ion-exchange) and different adsorbents (cellulose beads
loaded with iron oxyhydroxide, iron-oxide coated sand, granular


134


E. Pehlivan et al. / Food Chemistry 138 (2013) 133–138

ferric hydroxide, activated carbon, fly ash, zeolites, Calix[4]arenegrafted magnetite nanoparticles) have been used for the removal
of arsenic (Guo & Chen, 2005; Gupta, Basu, & De, 2007; Leupin &
Hug, 2005; Singh & Pant, 2004; Badruzzaman, Westerhoff, & Knappe, 2004; Mondal, Balomajumder, & Mohanty, 2007; Sayin et al.,
2010). Some of these new techniques are rather expensive for limited size water treatment systems situated in rural residential districts and they are in the developmental stages; consequently,
innovative cost-effective treatment processes are urgently needed.
Adsorption is considered as an economical and effective technique for arsenic removal because of its lower cost, and availability
of suitable adsorbents and their regeneration. Although the
adsorption capacity of agricultural by-products is usually less than
those of synthetic adsorbents, these materials could be an inexpensive alternative for water treatment plants. In order to enhance
their adsorption capacity, these materials are modified with various organic compounds having different functional groups.
The SCB contain biopolymers, mainly of polysaccharides with
hydroxyl, carboxyl and/or phenolic groups that can be chemically
modified to form new compounds with different properties. With
this investigation, we aimed at combining the beneficial effects
of both the polysaccharides and iron oxyhydroxides, such as SCBHFO, to form a new adsorbent for the removal of As(V) ion from
aqueous solutions and compare their performance with other
adsorbents for the same purpose. The influences of physical–chemical key parameters, e.g. pH, the initial concentration of arsenic, the
amount of adsorbent, contact time, the point of zero charge
(pHzpc) and ionic strength, were investigated in this study.
2. Materials and methods
2.1. The preparation of sugarcane bagasse SCB
SCB, obtained from a suburb of Hanoi, Vietnam, was powdered
in a ball mill (BLB Braunschweig) and sieved in a sieving machine
(Retsch, West of Germany). The sample having the sieve fraction of
125–200 lm was washed thoroughly with deionised water, and
dried in an aerated oven at 60 °C for 24 h. The air-dried and powdered SCB (50 g) was hydrolysed by using 1.15 M H2SO4 (w/w of
dry material, at 80 °C for 30 min) for removing starch, proteins
and sugars. Thereafter, the low molecular weight lignin compounds were removed by stirring the solid for 24 h at room temperature in 0.1 M NaOH solution (ratio SCB/sodium hydroxide is

5). After thorough washing, the adsorbent was dried in an oven
at 50 °C and stored in a desiccator prior to further experiments.
2.2. The modification of SCB with ferric nitrate Fe(NO3)3 to SCB-HFO
Ten grams of pretreated SCB were mixed for 48 h with 300 ml of
0.05 M ferric nitrate Fe(NO3)3 in a 1 l beaker. Aliquots of 1 M NaOH
were added dropwise into the beaker under continuous stirring,
keeping the pH between 2.8 and 3.5. After 48 h of the covering process, the suspension was filtered and washed with de-ionised
water several times until neutral pH was obtained. Coated adsorbent SCB-HFO was dried in an oven at 50 °C. Furthermore, it was
stored at room temperature until used.
2.3. Preparation of standards and reagents
Reagents used were purchased from Merck (Darmstadt, Germany) and were of analytical grade. MgSO4Á7H2O (Fluka, Seelze,
Germany), Na3PO4Á12H2O (Sigma–Aldrich, Seelze, Germany) and
NaNO3 (Merck) were used for studying the effects of ionic strength.
As(V) stock solution was purchased from Merck. Iron Nitrate
(Fe(NO3)3) was used for modification of the adsorbent. All

glassware was cleaned with diluted nitric acid and rinsed with
deionised water. Standard acid (0.1 M HCl) and base (0.1 M NaOH)
solutions were used for the pH adjustment of the solution.
The AAS standard solution of 1000 mg/l As(V) was prepared by
transferring the contents of a Titrisol ampule with As2O5 in H2O
(Merck, Germany) into a 1 l volumetric flask, which was filled up
to the mark at 20 °C according the instructions by Merck. The
working solutions of different concentrations were prepared by
diluting the stock solution immediately before starting the batch
studies. The As analyses were performed with a Hitachi Atomic
Absorption Spectrophotometer (HG-AAS, Series Z-2000; Hitachi
Corporation, Japan) which was connected to a hydride formation
system (model HFS-3; Hitachi). For hydride generation, the following solutions were used: (i) 1.2 M HCl (p.a., Merck); (ii) NaBH4–
NaOH-solution: solute 10 g NaBH4 (p.a., Fluka) in 1 l of H2O (Seralpure) by adding 4 g of NaOH (p.a., Merck); the solution was prepared immediately before use; (iii) KI-solution as a reduction

agent; 20% (w/v; reduction to As(III)). All standards, reference solutions, and sample solutions were adjusted to 0.24 M HCl and 2% KI.
The reduction agent was added at least 30 min before analysis. In
general, a 5-point calibration was run before starting the analyses
(0–20 lg/l). For monitoring of a possible signal drift, reference
solutions of 5 and 10 lg/l were used. As levels were measured
every 5–7 samples. For generation of hydride, HCl (1.2 M) and
NaBH4-solution were pumped into the reaction chamber in the hydride formation system; sample and standard solutions were
added. The flow rates of HCl and NaBH4 were 8–10 ml/min for
sample and standard 12 ml/min. A 12 cm quartz cuvette was
mounted above the standard burner flame zone that ran with air
(0.5 MPa) and C2H2 (1.2 l/min). Argon was used as carrier gas with
a flow rate of 0.3 l/min for constant transfer of As-hydride from the
reaction cell to the cuvette. The 193.7 nm emission line of the Ashollow-cathode lamp was used. Due to the long transport distances for reaction solutions and hydride gas, the absorption signals should be followed thoroughly; with pre-integration times
of at least 120 s for rinsing memory effects and to yield a constant
common analytical signal. An integration time of 5 s is standard.
Under these conditions, the instrumental detection limit was
0.2 lg/l. As a reducing agent, 2 ml of 30% HCl and 2 ml of 20%
(w/v) KI were added to 20 ml of the standard or sample As(V) solution and left for about 15 min for conversion of As(V) to As(III) ions.
2.4. Batch adsorption experiments
The defined amounts of SCB-HFO were added in 50 ml of aqueous As(V) solutions of different concentrations and shaken, using a
rotary shaker (Retsch, Germany), at 120 rpm for certain time intervals (15 min–24 h). Supernatant was filtered through a cellulose
acetate filter (pore size 0.2 lm) and analysed for As(V), using a
HG-AAS. The mass balance of As(V) adsorbed per mass unit of
the SCB-HFO (mg/g) was calculated by the following (Eq. (1)) (Altun & Pehlivan, 2012):

Q e ¼ ðC i À C e ÞV=W

ð1Þ

where Ci and Ce are the initial and equilibrium As(V) concentrations

in mg/l, respectively. V is volume of the As(V) solution in ml, and W
is the weight of adsorbent in mg.
The effect of initial pH (2–10) on the As(V) uptake by SCB-HFO
was studied by using 50 ml of 5 mg/l As(V) solution and 4 g/l of
adsorbent dosage at 23 °C. To study the effect of initial As(V) concentration, 10, 20, 30, 50, 75, 100, 200 and 300 mg As/l, 4 g adsorbent/l, pH 4, and a temperature 23 °C were applied. The effects of
contact time (15 min–24 h) and adsorbent amount (0.1–0.25 g)
were studied with an initial As(V) concentration of 2 mg/l, pH 4
and temperature 23 °C.


E. Pehlivan et al. / Food Chemistry 138 (2013) 133–138

135

3. Results and discussion

3.2. The influence of pH on As(V) sorption

3.1. FT-IR analysis

Solution pH normally has a large impact on adsorption performance (Arief, Trilestari, Sunarso, Indraswati, & Ismadji, 2008). The
effect of pH on As(V) adsorption was investigated using different
kinds of adsorbents and it produced similar results (Rahaman,
Basu, & Islam, 2008). Fig. 2b shows the relationships between
pH value and sorption yield of As(V). It was indicated that most
of the As(V) ions were bound to the adsorbent at an initial pH
range of 2–4. Therefore, there is more adsorption under acidic
conditions as well as in the near neutral region, i.e. at pH 2–6;
even more than 30% sorption is still observed up to pH 10. The
sorption yield reached a maximum value of 98% at pH 4 for

As(V) ions.
Many investigations were conducted on the adsorption of soluble arsenic species from adsorbent surfaces. The electrostatic force
is one important factor in the adsorption mechanisms. In the aqueous solution, the As(V) species predominate as a single negatively
charged anion (H2AsO4À) at pH 3–6 and a double negatively
charged form (HAsO42À) at pH up to 11, which can be adsorbed
on the SCB-HFO by substituting hydroxyl. In natural waters, the
electrostatic force between the negatively charged As(V) species
and the usually positively charged iron oxyhydroxide surface is
much stronger, resulting in a better adsorption of As(V). If the
arsenic concentration decreases, the electrostatic force for the
sorption is not strong enough to remove arsenic for meeting the
acceptable limit values in drinking water (Vaclavikova, Matik,
Jakabsky, & Hredzak, 2005).

SCB is a cellulose matrix, which has different binding sites,
including carboxyl (–COOH) and hydroxyl (–OH) groups. All infrared spectra of raw and treated materials were recorded using
an ATR technique at 4000–520 cmÀ1 and 32 scans with a BRUKER
FT-IR Tensor 27. The broad and strong band at 3343 cmÀ1 is due
to the hydroxyl group (–OH) and the absorption band at
2897 cmÀ1 due to the alkyl groups of the biomass. Absorption at
1729 cmÀ1 was attributed to stretching vibration of the carboxyl
group. The bands observed at 1034 cmÀ1 were assigned to C–O
stretching of alcohols and carboxylic acids (Fig. 1a).
Fig. 1b shows the SEM analysis of samples SCB and SCB-HFO.
The surface charge of sorbent was characterised by measuring
the zeta potential for the treated adsorbent that indicates the surface charge of a particle at a certain distance from the surface of
shear plane. For materials that undergo acid base reactions, the
surface charge depends on pH. The sorption of the As(V) is expected to be favoured at a pH less than pHzpc (zpc: zero point
charge) of the adsorbents (Kamala et al., 2005). SCB-HFO has ironoxyhydroxide groups, which are positively charged at pH smaller
than 6 (Fig. 2a). For that reason, the adsorption experiments were

performed at pH 4. The reactivity of HFO coated on SCB is similar to
iron oxide surface sites and the active form of this adsorbent is the
hydrolysed surface species „FeOH, which behaves like an
amphoteric site with a point of zero charge 5.8 (Fig. 2a).

Fig. 1a. FT-IR spectrum of raw SCB (black) and SCB-HFO (red). (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of
this article.)

Fig. 1b. SEM of SCB (left) and SCB-HFO (right).


136

E. Pehlivan et al. / Food Chemistry 138 (2013) 133–138

Different situations in the adsorption of As(III) and As(V) can be
expressed by their respective speciations in aqueous medium.
As(III) is present as an anion above pH 9 (and the experimental
data between pH 3 and 8); it is thus reasonable to accept that
the neutral form, H3AsO3, interacts with the adsorbent (Dupont,
Jolly, & Aplincourt, 2007). When pH increases from 4 to 10, the
As(V) biosorption decreased because of the decrease of electrostatic interactions between the positively charged surface groups
(„FeOH2+) and the anionic As(V) species which prevents the formation of surface complexes (Dupont et al., 2007). Physical forces,
e.g. van der Waals and London forces, might slightly overlap the
adsorption processes.
3.3. The determination of Fe(III) amount loaded onto SCB
The amount of Fe(III) (in hydrated ferric oxide) loaded onto SCB
was determined after acid digestion. In order to avoid uncontrolled
reactions, 0.1 g of SCB-HFO was kept for 24 h at room temperature
in aquatic conditions. The digestion was completed by raising the

temperature to 90 °C until the sample became dry; 10 ml of 1 M
HCl were added to the solution, and stirred several times. After
12 h, 2 ml of digested solution were drawn off and diluted to
10 ml. The concentration of Fe(III) in the solution was analysed
by GF-AAS.

Fig. 2a. The pH point zero of charge (pHpzc) of SCB-HFO.

120

As(V)

100

% Sorption

80

3.4. Adsorption isotherm models
60
40
20
0
0

2

4

6


8

10

12

Ce Ce
1
¼ þ
qe As As K b

pH
Fig. 2b. Sorption of As(V) on SCB-HFO as a function of pH (50 ppm As(V) in 50 ml of
solution at different pH values; temperature 22 ± 2 °C).

Another possible reaction mechanism assumes that hydroxyl
groups coordinate with As(V) ions (Boddu et al., 2008). At elevated
pH, As(V) sorption was decreased due to the competition of the
OH-groups bound to the matrix surface with the free hydroxyl
ions. Under alkaline conditions, the surface of SCB-HFO will become negatively charged, causing a repulsive force versus the anionic species of As(V), resulting in a decreasing sorption efficiency
(Rahaman et al., 2008). The formation of surface complexes between positively charged surface groups „FeOH2+ and negatively
charged As(V) ions can be described according to the following
equilibrium (2 and 3) (Sherman & Randall, 2003):

2

FeOHþ2

þ


H2 AsOÀ4

À

¡ ½ð FeOÞ2 À AsOðOHފ þ H3 O þ H2 O



þ

H2 AsOÀ4

À

þ

¡ ½ Fe À ðOÞ2 À AsOðOHފ þ H3 O

ð3Þ

In this situation, arsenate adsorption on the adsorbent can occur through non-specific coulombic interactions (outer-sphere
adsorption) and formation of monodentate (2) and bidentate (3)
surface complexes. In another case, ligand exchange on the surface
(inner-sphere adsorption) might happen during the following
equilibrium (4 and 5) (Jeon, Baek, Park, Oh, & Lee, 2009):

 FeOHþ2 þ H2 AsOÀ4 ¡  Fe À O À AsO3 H2 þ H2 O

ð4Þ


À
2  FeOH þ H2 AsO2À
4 ¡  Fe2 À O2 AsO2 H þ 2OH

ð5Þ

ð4Þ

where As (mol gÀ1) and Kb (l molÀ1) are the coefficients, qe is the
weight of adsorbate which was adsorbed per unit mass of adsorbent, and Ce is the analyte concentration in aqueous phase at equilibrium.Freundlich equation:

x
1=n
ð Þ ¼ kC e
m

ð5Þ

where 1/n is the adsorption intensity, k is the adsorption yield, x/m
is the mass of adsorbate adsorbed per unit mass of adsorbent and Ce
is the element concentration at equilibrium in the aqueous phase.
The modified formula of this equation was obtained as follows
(Eq. (6)):

log

þ

ð2Þ

FeOHþ2

Two well established equilibrium models, after Langmuir and
Freundlich, were applied for the adsorption study. The Freundlich
model assumes a heterogeneous sorbent surface and different
binding energies for the active sites (Altun & Pehlivan, 2007). The
Langmuir isotherm is frequently used for the adsorption of metal
ions from aqueous solutions (Langmuir, 1918).
The general form of the Langmuir model is given below
(4):Langmuir equation

x
m

¼ log k þ

1
log C e
n

ð6Þ

Table 1 shows that the values of Kf and n were 1.55 Â 10À6 and
0.18 for As(V) adsorption. However, R2 was calculated for Freundlich as 0.79, that is much lower than 0.99 for the Langmuir isotherm. It was shown that the Langmuir model described the
adsorption of the As(V) onto SCB-HFO better than did the Freundlich model. It could be concluded that the Langmuir isotherm model better fits the equilibrium data.
Fig. 3a shows that the non-linear relationship between the
amount of As(V) ion adsorbed on SCB-HFO depends on the arsenic
concentration. The maximum sorption capacity (As) of adsorbent
was found to be 22.1 mg/g for As(V). The Kb value was found to
be 0.45 for As(V) sorption.



137

E. Pehlivan et al. / Food Chemistry 138 (2013) 133–138
Table 1
Langmuir and Freundlich isotherm constants.

14

Freundlich isotherm parameters

2

2

12

As (mg/g)

Kb

R

Kf (mg/g)

n

R


22.1

0.45

0.99

1.55 Â 10À6

0.18

0.79

As(V) uptake (mg/g)

Langmuir isotherm parameters

25

10
8
6
4
2

As(V) uptake (mg/g)

20

0
0


0.05

15

0.1

0.15

0.2

0.25

0.3

Adsorbent amount (g)
Fig. 3b. Sorption of As(V) on SCB-HFO as a function of adsorbent amount (50 ppm
As(V) in 50 ml of solution at pH 4; adsorbent amount 0.1–0.25 g; temperature
22 ± 2 °C.

10

5

0

Table 3
The relationship of desorption and pH values.

0


50

100

150

200

250

300

350

As(V) concentration (ppm)

Leaching agent

pH

Desorption (%)

HCl (30%)
NaOH (1 M)

1.5
12
14


17
54
85

Fig. 3a. Sorption isotherm of As(V) on SC-HFO as a function of initial As (V)
concentration (10–300 ppm As(V); 50 ml; 0.2 g adsorbent; pH 4; 22 ± 2 °C; contact
time 3 h).

Iron compounds are reported to be effective for the removal of
As(V) ions. Several Fe(III) oxides/oxyhydroxides, e.g. amorphous
hydrous ferric oxide HFO (FeOOH), poorly crystalline hydrous ferric oxide–ferrihydride (Wilkie & Hering, 1996), goethite (a-FeOOH)
and akaganeite (b-FeOOH), were investigated for removing As(V)
from aqueous solutions. Other sorbents, based on iron oxides/oxyhydroxides, e.g. iron oxide-coated polymeric minerals (Katsoyiannis & Zouboulis, 2002), iron-hydroxide-coated alumina (Hlavay &
Polyak, 2005), and natural iron ores (Zhang, Singh, Paling, & Delides, 2004) were also investigated. A comparison of the removal
capacities, for As(V), of different sorbents materials is given in Table 2 showing that the SCB-HFO presented in this study had a medium sorption capacity compared with others.
3.5. The effect of adsorbent dose on sorption of As(V) by SCB-HFO
The effect of adsorbent amount on As(V) sorption was studied
with the initial As(V) ion concentration of 50 ppm at 22 ± 2 °C
and pH 4. The amount of SCB-HFO changed from 0.1–0.25 g. It
was indicated that the equilibrium concentration in the dissolved

phase decreased when increasing the amount of adsorbent
(Fig. 3b). The optimum amount of SCB-HFO was found to be
0.25 g/50 ml of As(V) solution. When the dosage was increased,
the number of surface sites in the structure of the adsorbent lattice
increased. This shows that the main factors governing the adsorption of arsenic species are the electrostatic interaction between
ironoxyhydroxide sites of the adsorbent and the anionic arsenic
species. Hence, facilitating the binding of arsenate resulted from
both electrostatic interactions and hydrogen bonding.
3.6. Desorption efficiency

The desorption of the adsorbed As(V) from SCB-HFO was studied by eluting with 30% HCl and 1 M NaOH (each 20 ml). The results were given in Table 3. The desorption of As(V), using 30%
HCl, was 17%, whereas the highest recovery of 85% was reached
with 1 M NaOH. pH of the solution phase was adjusted by adding
HCl and NaOH solutions. The results showed that the adsorbent
can be successfully reused upon treatment with 0.1 M NaOH solution, which may be referred to the displacement of As(V) bound to
the adsorbent with OHÀ ions.

Table 2
Arsenic removal capacities of different sorbents.
Adsorbent

Qmax (mmol gÀ1)

pH

References

Akaganeite
Akaganeite
Goethite

1.79
0.93
0.330

7.5
3.5
5.0

Deliyanni, Bakoyannakis, Zouboulis, and Matis (2003)

Vaclavikova et al. (2005)

Hydrous ferric oxide HFO
Fe(III) loaded resin

1.340
0.800

4.0
1.7

Fe-hydroxide coated alumina
Coconut-shell carbon
Peat-based carbon
Magnetite

0.210
0.430
0.070
0.350

6.6–7.2
5.0
5.0
6.5

(Rau, Gonzalo, & Valiente, 2003)
Hlavay and Polyak (2005)
Lorenzen, van Deventer, & Landi (1995)
Lorenzen, van Deventer, & Landi (1995)

Javier, Maria, de Joan, Miquel, and Lara (2007)

SCB-HFO

0.300

4.0

Present study

(Matis, Lehmann, & Zouboulis, 1999)
Wilkie and Hering (1996)


138

E. Pehlivan et al. / Food Chemistry 138 (2013) 133–138

3.7. The effects of the ionic strength (competing anions) on As(V)
removal
The ionic strength of the solution might compete with the As(V)
removal, in particular in the case of other co-occurring multiple
charged anions. Therefore, the removal of 50 ppm As(V) in 50 ml,
through 0.2 g of SCB-HFO, was investigated in parallel in the presence and absence of PO43À (50 ppm), NO3À (50 ppm), and SO42À
(250 ppm). The pH of the solutions was adjusted to 4.0 and the
samples were agitated for 3 h at 200 rpm. The results confirmed
that As(V) removal was suppressed by PO43À ions. The adsorption
capacity for As(V) was decreased by 6.5%, but the other anions
did not affect the adsorption process.
4. Conclusion

The occurrence of arsenic in water is of major concern in many
countries. The threshold values of arsenic in drinking water have
been set by public authorities, worldwide, at 10 lg lÀ1. In this
study, a novel adsorbent from sugarcane bagasse, as a low-cost
agro-waste, was developed, through treatment with iron(III)oxyhydroxide (SCB-HFO), that seems to have promising properties
for the removal of As(V) from aqueous solutions.
The main factors determining the adsorption of As(V) on this
sorbent are electrostatic interactions, ligand exchange, and chelation between positively charged surface groups „FeOH2+ and negatively charged As(V) ions. The adsorption capacity of SCB-HFO
was found to be 22.1 mg/g for As(V) under optimum conditions
of 3 h agitation at pH 4, and 22 °C. As(V) ions could be desorbed
successfully from SCB-HFO by using 1 M NaOH and the absorbent
was thus regenerated. Among typical anions in surface waters, only
phosphate (50 ppm) suppressed As(V) removal by the adsorbent
whereas nitrate and sulfate did not affect the sorption process.
The presented findings suggest that SCB-HFO is an inexpensive
adsorbent for As(V) removal from aqueous solutions.
Acknowledgements
This investigation was performed at the Guest Chair within the
project ‘‘Exceed-Excellence Center for Development CooperationSustainable Water Management in Developing Countries’’ at the
Technische Universitaet Braunschweig, Prof. Pehlivan being the
visiting professor, and Ms. Tran and Mr. Ouédraogo are the international exchange staff members. The Exceed Project is granted by
the German Federal Ministry for Economic Cooperation and Development (BMZ) and German Academic Exchange Service (DAAD);
their financial support is gratefully acknowledged.
References
Altun, T., & Pehlivan, E. (2007). Removal of copper(II) ions from aqueous solutions
by walnut- hazelnut- and almond-shells.. Clean, 35(6), 601–606.
Altun, T., & Pehlivan, E. (2012). Removal of Cr(VI) from aqueous solutions by
modified walnut shells. Food Chemistry, 132, 693–700.
Arief, V. O., Trilestari, K., Sunarso, J., Indraswati, N., & Ismadji, S. (2008). Recent
progress on biosorption of heavy metals from liquids using low cost

biosorbents: Characterization, biosorption parameters and mechanism
studies. Clean, 36, 937–962.
Badruzzaman, M., Westerhoff, P., & Knappe, D. R. U. (2004). Intraparticle diffusion
and adsorption of arsenate onto granular ferric hydroxide. Water Research, 38,
4002–4012.
Boddu, V. M., Abburib, K., Talbottc, J. L., Smitha, E. D., & Haasch, R. (2008). Removal
of As(III) and As(V) from aqueous medium using chitosan-coated biosorbent.
Water Research, 42, 633–642.

Deliyanni, E. A., Bakoyannakis, D. N., Zouboulis, A. I., & Matis, K. A. (2003). Sorption
of As(V) ions by akaganeite-type nanocrystals. Chemosphere, 50, 155–163.
Dupont, L., Jolly, G., & Aplincourt, M. (2007). Arsenic adsorption on
lignocellulosic substrate loaded with ferric ion. Environmental Chemistry
Letters, 5, 125–129.
Environmental Protection Agency (1999). Special Report, on Ingested inorganic
arsenic skin cancer: Nutritional essentiality; Report EPA/625 3-87-13,
Washington, DC, U.S.
Farias, S. S., Casa, V. A., Vàzquez, C., Ferpozzi, L., Pucci, G. N., & Cohen, I. M. (2003).
Natural contamination with arsenic and other trace elements in ground waters
of Argentine Pampean plain. Science of the Total Environment, 309, 187–199.
Fazal, M. A., & Kawaci, T. (2001). Extent and severity of ground water arsenic
contamination in Bangladesh. Water International, 26(3), 370–379.
Guo, X., & Chen, F. (2005). Removal of arsenic by bead cellulose loaded with iron
oxyhydroxide from groundwater. Environmental Science & Technology, 39,
6808–6818.
Gupta, A. S. D., Basu, J. K., & De, S. (2007). Sorption of arsenite using natural laterite
as adsorbent. Separation and Purification Technology, 55, 350–359.
Hlavay, J., & Polyak, K. (2005). Determination of surface properties of
ironhydroxide-coated alumina adsorbent prepared for removal of arsenic
from drinking water. Journal of Colloid and Interface Science, 284, 71–77.

Javier, G., Maria, M., de Joan, P., Miquel, R., & Lara, D. (2007). As(V) sorption onto
natural hematite, magnetite and goethite. Journal of Hazardous Materials, 141,
575–580.
Jeon, C., Baek, K., Park, J., Oh, Y., & Lee, S. (2009). Adsorption characteristics of As(V)
on iron-coated zeolite. Journal of Hazardous Materials, 163, 804–808.
Kamala, C. T., Chu, K. H., Chary, N. S., Pandey, P. K., Ramesh, S. L., Sastry, A. R. K., et al.
(2005). Removal of arsenic(III) from aqueous solutions using fresh and
immobilized plant biomass. Water Research, 39, 2815–2826.
Katsoyiannis, I. A., & Zouboulis, A. I. (2002). Removal of arsenic from contaminated
water sources by sorption onto iron-oxide-coated polymeric materials. Water
Research, 36, 5141–5155.
Langmuir, I. (1918). The sorption of gases on plane surfaces of glass, mica and
platinum. Journal of the American Chemical Society, 40, 1361–1403.
Leupin, O. X., & Hug, S. J. (2005). Oxidation and removal of arsenic(III) from aerated
groundwater by filtration through sand and zero-valent iron. Water Research,
39(9), 1729–1740.
Lorenzen, L., van Deventer, J. S. J., & Landi, W. M. (1995). Factors affecting the
mechanism of the sorption of As(V) species on activated carbon. Minerals
Engineering, 8, 557–569.
Matis, K. A, Lehmann, M., & Zouboulis, A. I (1999). Modeling sorption of metals from
aqueous solution onto mineral particles: The case of arsenic ions and goethite
ore. In P. Misaelides, F. Macašek, T. J. Pinnavaia, & C. Colella (Eds.), Natural
microporous materials in environmental technology (pp. 463–472). The
Netherlands: Kluwer.
Mondal, P., Balomajumder, C. B., & Mohanty, B. (2007). Quantitative separation of
As(III) and As(V) from a synthetic water solution using ion exchange columns in
the presence of Fe and Mn ions. Clean, 35(3), 255–260.
Navarro, R. R., Sump, K., Fujii, N., & Matsumura, M. (1996). Mercury removal from
wastewater using porous cellulose carrier modified with polyethyleneimine.
Water Research, 30(10), 2488–2494.

Ng, J. C. (2005). Environmental contamination of arsenic and its toxicological impact
on humans. Environmental Chemistry, 2, 146–160.
Rahaman, M. S., Basu, A., & Islam, M. R. (2008). The removal of As(III) and As(V) from
aqueous solutions by waste materials. Bioresource Technology, 99, 2815–2823.
Rau, I., Gonzalo, A., & Valiente, M. (2003). Arsenic (V) adsorption by immobilized
iron mediation. Modeling of the adsorption process and influence of interfering
anions. Reaction & Functional Polymer, 54, 85–94.
Sayin, S., Ozcan, F., Yilmaz, M., Tor, A., Memon, S., & Cengeloglu, Y. (2010). Synthesis
of Calix[4]arene-grafted magnetite nanoparticles and evaluation of their
arsenate as well as dichromate removal efficiency. Clean: Soil, Air, Water,
38(7), 639–648.
Sene, L., Converti, A., Felipe, M. G. A., & Zilli, M. (2002). Sugarcane bagasse as
alternative packing material for biofiltration of benzene polluted gaseous
streams. A preliminary study. Bioresource Technology, 83(2), 153–157.
Sherman, D. M., & Randall, S. R. (2003). Surface complexation of arsenic (V) to iron
(III) hydroxides: Structural mechanism from ab initio molecular geometries and
EXAFS spectroscopy. Geochimica et Cosmochimica Acta, 67, 4223–4230.
Singh, T. S., & Pant, K. K. (2004). Equilibrium, kinetics and thermodynamic studies
for adsorption of As(III) on activated alumina. Separation and Purification
Technology, 36, 139–147.
Vaclavikova, M., Matik, M., Jakabsky, S., & Hredzak, S. (2005). Preparation and
sorption properties of Fe-nanomaterials for removal of arsenic from waters, In
Book of abstract of NATO CCMS on clean products and processes, Norway 13.
Wilkie, J. A., & Hering, J. G. (1996). Adsorption of arsenic onto hydrous ferric oxide:
Effect on adsorbate/adsorbent ratios and co-occurring solutes. Colloids and
Surfaces A: Physicochemical and Engineering Aspects, 107, 97–110.
Zhang, W., Singh, P., Paling, E., & Delides, S. (2004). Arsenic removal from
contaminated water by natural iron ores. Minerals Engineering, 17, 517–524.




×