You are on page 1of 6

Adsorptive removal of Cu(II) from aqueous solutions using non-crosslinked and

crosslinked chitosan-coated bentonite beads


Maria Lourdes P. Dalida
a
, Ana Francia V. Mariano
a
, Cybelle M. Futalan
a
, Chi-Chuan Kan
b
,
Wan-Chi Tsai
c,
, Meng-Wei Wan
b,

a
Department of Chemical Engineering, Environmental Engineering Unit, University of the Philippines-Diliman, Quezon City 1011, Philippines
b
Department of Environmental Engineering and Science, Chia Nan University of Pharmacy and Science, Tainan 71710, Taiwan
c
Department of Medical Laboratory Science and Biotechnology, Kaohsiung Medical University, Kaohsiung, 80708, Taiwan
a b s t r a c t a r t i c l e i n f o
Article history:
Received 30 October 2010
Received in revised form 21 February 2011
Accepted 22 February 2011
Available online 21 March 2011
Keywords:
Adsorption isotherm
Bentonite
Chitosan
Crosslinking
Desorption
Batch experiments were executed to investigate the removal of Cu(II) from aqueous solutions using non-
crosslinked (CCB) and crosslinked chitosan-coated (CCB-ECH) bentonite beads. CCB and CCB-ECH beads were
characterized by BET surface area and pore diameter analysis and X-ray diffraction (XRD). The percentage
removal and adsorption capacity of Cu(II) ions were examined as a function of initial concentration and pH.
The equilibrium data of CCB agreed well with the Langmuir model while CCB-ECH beads showed a better t
with the Freundlich model. Based on the isothermstudy, CCB is a homogenous adsorbent, whereas CCB-ECHis
a heterogeneous adsorbent. The adsorption capacities of CCB and CCB-ECH at pH 4 are 12.21 and 9.43 mg/g,
respectively. The kinetic data correlated well with the pseudo-second order equation, which implies that
chemisorption is the rate-limiting step. The desorption study was performed using eluent solutions, tap water
(pH 7) and HCl solution (pH 1 and pH 3). The best Cu(II) recovery was obtained using HCl solution (pH 1).
2011 Elsevier B.V. All rights reserved.
1. Introduction
Efuents that contain signicant amount of heavy metals are
generated by industrial sources such as electroplating industry,
mining activities, metal fabrication, and battery manufacturing plants.
Generally, Hg(II), Pb(II), Cr(III), Ni(II), Cu(II), Cd(II) and Zn(II) are
considered toxic owing to their non-biodegradability, tendency to
accumulate in living organisms and the ability to undergo transfor-
mation when released into the environment [1]. Copper is one of the
essential nutrients that are needed by the body in trace quantities.
However, intake of Cu(II) at high dosages can cause health problems
such as lesions in the central nervous system, Wilson's disease, and
gastrointestinal disturbance that includes vomiting and nausea [2,3].
In addition, Cu(II) becomes toxic to aquatic organisms like sh at low
pH [4]. It is, therefore, important to remove any excess amount of
copper present in wastewater in order to protect the public health and
to prevent pollution of the surface water and groundwater.
Conventional technologies that treat metal-bearing efuents are
membrane separation, electrodeposition, ion exchange, chemical
precipitation, and solvent extraction. Though effective in removing
heavy metals in wastewater with high solute loadings, there are
several disadvantages such as high operating cost and ineffectivity in
removing trace quantities of heavy metals from waste streams.
Another disadvantage is the production of sludge or mud, which
requires proper disposal and connement [5]. On the contrary,
adsorption is an effective and economical method among the
physicochemical treatments. Activated carbon has been widely used
as a commercialized adsorbent for the removal of heavy metals from
wastewater. However, it is an expensive material. Studies on low cost
materials as potential adsorbents have included zeolites, chitin,
chitosan, and agricultural wastes such as maize leaf, banana pith,
peanut husk, pumpkin waste, rice hull and saw dust [6,7].
Chitosan is a product of the partial N-deacetylation of chitin, which
has several desirable properties like biodegradability, hydrophilicity,
anti-bacterial property, and non-toxicity [8]. It contains amino
(NH
2
) and hydroxyl (OH) groups that can serve as coordination
sites to bind heavy metals. However, chitosan has weak mechanical
and chemical properties where it easily dissolves in dilute organic
acids such as formic acid and acetic acid, and agglomerates to form a
gel in aqueous solution [5,9]. In order to overcome these limitations,
physical and chemical modication needs to be carried out on
chitosan.
Physical modication reduces the crystalline state of chitosan and
results in an expansion of the polymer network that allows an easy
access to its binding sites [9]. In addition, it decreases the amount of
chitosan needed to synthesize the composite material and helps in
overcoming the mass transfer limitations of chitosan [10,11]. On the
other hand, chemical modication enhances chitosan's mechanical
strength, improves its chemical stability in acidic and alkaline media,
and increases its resistance to microbiological and biochemical
Desalination 275 (2011) 154159
Corresponding authors. Tel.: +886 6 2660615; fax: +886 6 213 1291.
E-mail address: peterwan@mail.chna.edu.tw (M.-W. Wan).
0011-9164/$ see front matter 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.desal.2011.02.051
Contents lists available at ScienceDirect
Desalination
j our nal homepage: www. el sevi er. com/ l ocat e/ desal
degradation [4,9,12]. Previous studies on chitosan have utilized
different types of crosslinking agents like epichlorohydrin [9,1214],
glutaraldehyde [4,14], and ethylene glycol diglycidyl ether [10,15].
Clay materials are made of hydrous aluminosilicates, which have
high cation exchange capacity and large specic surface areas [16].
Moreover, clay minerals are abundant, chemically and mechanically
stable, and low-cost, making it an attractive immobilization material
for chitosan [17]. Bentonite, which is mainly composed of montmo-
rillonite, is a layered mineral with a crystalline structure and net
negative surface charge [8,16].
In this study, a combination of physical and chemical modication
was applied on chitosan, where it was immobilized on bentonite and
crosslinked with epichlorohydrin (ECH). Removal of Cu(II) from
aqueous solutions was investigated under static conditions using non-
crosslinked and crosslinked chitosan-coated bentonite. Effects of pH
and initial Cu(II) concentration on the adsorption capacity were
examined. The equilibrium data were analyzed using Langmuir and
Freundlich isotherm models. The kinetic constants of the experimen-
tal data were determined using the pseudo-rst order and pseudo-
second order equations. Regeneration studies were also carried out to
determine the reusability of the composite adsorbent.
2. Materials and method
2.1. Chemicals and equipment
Low-molecular weight chitosan with 75% to 85% degree of
deacetylation was purchased from Sigma Aldrich. Epichlorohydrin
(ECH), kaolinite and bentonite were also procured from Sigma
Aldrich. Copper sulfate anhydrous (CuSO
4
, 99%), sodium hydroxide
(NaOH, 99%) and hydrochloric acid (HCl, 37%) were purchased from
Merck KGaA. Deionized (DI) water was used to prepare all solutions.
The adsorbent was dried using a Channel Precision Oven model
DV452 220 V. Batch studies were carried out using a BT-350 YIH
shaker bath. An inductively coupled plasma optical emission
spectrometry (ICP-OES) Perkin Elmer DV2000 Series was utilized in
the analysis of residual Cu(II). The surface area analysis of bentonite,
chitosan, non-crosslinked and crosslinked chitosan-coated bentonite
beads was performed using a BET-N
2
adsorption GEMINI 2360
Micrometrics. The X-ray diffraction (XRD) patterns were obtained
using a Bruker D8 Series with a slow scan at 0.3/s in 2 range of 2
15.
2.2. Preparation of non-crosslinked chitosan-coated bentonite beads
Chitosan powder (5 g) was dissolved in 300 mL of 5% (v/v) HCl
solution under constant stirring for 2 h at 25 C. About 100 g of
bentonite was added and then mixed for 3 h. The resulting chitosan
clay solution was neutralized with 1 N NaOH that was added drop by
drop, until precipitation of chitosan onto the bentonite surface
occurred. The chitosan-coated bentonite beads were allowed to settle
and then ltered. The beads were washed using DI water to remove
any impurities and dried in the oven at 65 C for 24 h. After drying, the
beads were ground and sieved using ASTM sieve sizes #45 and #35.
The beads are then called non-crosslinked chitosan-coated bentonite
(CCB). The CCB beads with particle size ranging from 0.35 to 0.70 mm
were collected and utilized for the batch experiments.
2.3. Preparation of crosslinked chitosan-coated bentonite beads
The crosslinked chitosan-coated bentonite beads were prepared
using the method similar to Wan Ngah et al. [18] and Wan et al. [19]
with some modications.
To obtain a 1:1 molar ratio with chitosan, about 2.873 g of ECHwas
added into the chitosanclay solution. The solution was stirred using a
magnetic plate at 45 C for about 2 h. The solution was neutralized
using 1 N NaOH. After precipitation, the crosslinked chitosan-coated
bentonite beads (CCB-ECH) were ltered and rinsed using DI water in
order to remove any NaOH or HCl residue. The CCB-ECH beads were
dried for 24 h in the oven at 65 C, ground and sieved. The particle size
range of 0.35 to 0.70 mm of CCB-ECH beads were utilized in the batch
experiments.
2.4. Batch experiments
Experiments were carried out using 2.5 g of adsorbent and 30 mL
of Cu(II) solution in a 125 mL Erlenmeyer ask. The asks were
covered with paralm and were agitated at 50 rpm using a reciprocal
shaker bath at 25 C. The adsorbent was then ltered using Whatman
#40 in preparation for the analysis of residual Cu(II) in the
supernatant. The concentration of Cu(II) in the ltrate was analyzed
at a wavelength of 324.75 nm using ICP-OES. The effects of pH (pH 1
4) and initial concentration (100 to 2000 mg/L) on the removal
efciency and adsorption capacity were determined. The initial pH of
the solution was adjusted using 1 N NaOH or 1 N HCl solution.
The adsorption capacity of the composite adsorbent was calculated
using Eq. (1):
qe =
C0Ce V
W
1
where q
e
is the adsorption capacity (mg/g), V is the volume of the Cu
(II) solution (mL), W is the weight of CCB and CCB-ECH beads, and C
0
and C
e
are the initial and nal concentration (mg/L) of Cu(II) ions,
respectively.
2.5. Batch kinetics
The kinetic studies were performed using 2.5 g of adsorbent in
30 mL of Cu(II) solution at initial concentration of 100 to 2000 mg/L.
The asks were agitated at 50 rpm for pre-determined time intervals
(30 min to 4 h).
2.6. Equilibrium studies
Batch equilibrium studies were conducted using 30 mL aliquots of
Cu(II), where a xed amount (2.5 g) of adsorbent was added. The
initial Cu(II) concentration was varied in the range of 100 to 2000 mg/
L. For both CCB and CCB-ECHbeads, a contact time of 4 h and agitation
speed of 50 rpm were applied.
2.7. Regeneration studies
For the desorption study, 2.5 g of CCB-ECH and 30 mL aliquot of Cu
(II) solution were agitated at 50 rpm for 4 h contact time. The Cu(II)-
loaded CCB-ECH beads were ltered and washed with deionized
water to remove any unadsorbed Cu(II) ions. The ltrate was analyzed
for residual Cu(II).
The Cu(II)-loaded CCB-ECH beads were agitated at 50 rpm for 2 h
using 30 mL of tap water (pH 7), and HCl solution (pH 1 and 3). The
solution was ltered and analyzed using ICP-OES for Cu(II) in the
ltrate. The regenerated CCB-ECH beads were reused in the second
adsorption cycle, which applied the same procedure used in the batch
experiments (Section 2.4) with a contact time of 4 h.
3. Results and discussion
3.1. Surface area analysis
Table 1 lists the surface area and other physical properties of
chitosan, bentonite, CCB, and CCB-ECH. From the IUPAC recommen-
dation, the total porosity is divided into three categories based on the
155 M.L.P. Dalida et al. / Desalination 275 (2011) 154159
average pore diameter (d): macropores (dN50 nm), mesopores
(2 nmbdb50 nm), and micropores (db2 nm) [4]. According to
Table 1, chitosan, bentonite, CCB, and CCB-ECH are mesopores.
Chitosan is observed to have the least specic surface area (3.40 m
2
/
g) while bentonite has the highest surface area (94.28 m
2
/g). The
composite material of chitosan and bentonite (CCB) resulted in a
larger surface area than that of chitosan but less than that of
bentonite. Further reduction in the surface area was observed for
CCB when crosslinked to form CCB-ECH, where the value decreased
from 48.81 to 38.96 m
2
/g. The specic surface area, micropore area,
and total pore volume are in the order of bentonite N CCB N CCB-ECHN
chitosan. The CCB and CCB-ECH beads have inferior properties in
comparison to bentonite owing to formation of a occulated structure,
where an intermolecular hydrogen bond was formed between the
hydroxylated edgeedge silicate layers and the amino (NH
2
) or
hydroxyl (OH) functional group of chitosan [8]. Moreover, the
smaller surface areas of CCB and CCB-ECH are caused by the pore
blocking of chitosan and ECH molecules on the bentonite surface [20].
3.2. XRD analysis
Fig. 1 shows the XRD patterns of bentonite, chitosan, CCB and CCB-
ECH, respectively. XRD analysis can tell whether the chitosan and ECH
molecules entered into the interlayer spaces of bentonite. The XRD
pattern for bentonite has two pronounced peaks at 2=6.12 and
2=9.20. Upon addition of chitosan and ECH, no obvious shifts of the
peaks to a lower diffraction angle were observed. This indicates that
there is no change in the basal spacing of bentonite, implying that the
chitosan and ECHmolecules did not enter the silicate layers of the clay
mineral, hence no intercalation occurred. At 2=6.12 for CCB and
CCB-ECH beads, the peaks become relatively broader and less intense.
The interaction of bentonite with chitosan and ECH caused a slight
distortion of the intrinsic lattice arrangement of the silicate layers,
causing a decrease in the crystallinity, hence broader and less intense
peaks were observed.
3.3. Effect of pH and initial concentration
The effect of pH on the adsorption of Cu(II) ions by CCB and CCB-
ECH was studied in the pH range between 1 and 4. As shown in
Tables 2 and 3, the percentage removal and adsorption capacity of CCB
and CCB-ECH increase as the solution pH becomes less acidic from pH
1 to pH 4. The low uptake of Cu(II) ions at lower pH can be attributed
to the high concentration of H
+
ions, which compete against Cu(II) for
the binding sites on the CCB and CCB-ECH surface. In addition, most
amino groups of CCB and CCB-ECH become protonated (NH
3
+
) at
lower pH, which reduces the available binding sites for Cu(II) ions
[4,5,11]. As Cu(II) ions are transported from the solution to the
adsorbent, the protonated amino groups inhibit the approach of Cu(II)
due to the electrostatic repulsion force exerted by NH
3
+
on the
adsorbent surface [24]. When the pH is increased from pH 1 to 4, the
amino groups become deprotonated, hence it is free to interact and
bind with Cu(II) ions. Therefore, the maximum adsorption of Cu(II)
ions is observed at pH 4 on these two adsorbents.
The effect of initial Cu(II) concentration (100 to 2000 mg/L) on the
adsorption of Cu(II) ions onto CCB and CCB-ECH was also examined.
Tables 2 and 3 list the percentage removal and adsorption capacity at
different initial concentrations. As the initial concentration increased
from 100 to 2000 mg/L, the percentage Cu(II) removal was observed
to decrease. At higher concentrations, the binding sites on the
adsorbent were not sufcient to accommodate all Cu(II) ions, hence
a lower % removal was observed. On the other hand, higher initial
concentration caused the adsorption capacity of Cu(II) to increase. A
high initial concentration of 2000 mg/L provides a larger concentra-
tion gradient that would help overcome the mass transfer resistance
of the Cu(II) ions from the aqueous solution to the CCB and CCB-ECH
surfaces [11,13]. The adsorption capacity, q
e
was observed to be of
similar values at initial Cu(II) concentrations of 100 and 500 mg/L for
the two adsorbents. This is an indication that the initial concentration
was too low where the binding sites of the adsorbent was not yet
saturated by Cu(II) ions [2,3].
In the present study, CCB was crosslinked using ECH to increase
the chemical stability of the adsorbent in acidic media. ECH is a
crosslinking agent that belongs to the epoxide group [21]. In Table 3, it
is shown that non-crosslinked CCB beads have higher adsorption
capacities over CCB-ECH beads at 1000 and 2000 mg/L. Some of the
binding sites of chitosan were used in the crosslinking process, where
ECH interacted with the hydroxyl group of the chitosan molecule,
hence a decrease in q
e
values of CCB-ECH. In addition, the ECH
molecule is a short and rigid polymeric chain that contributes to a
reduction in the accessibility of the coordination sites on the
adsorbent, therefore there is less afnity of the Cu(II) ions to the
binding sites on CCB-ECH [21].
3.4. Adsorption kinetics
The kinetics of Cu(II) adsorption on CCB and CCB-ECH were
studied under pre-determined time intervals at several initial
concentrations from 100 to 2000 mg/L and at an optimum pH 4.
Kinetic equations were applied on the experimental data to
investigate the potential rate-determining step of the adsorption
process. In this study, two kinetic models were utilized, namely
pseudo-rst order and pseudo-second order equations.
The linear formof the pseudo-rst order equation of the Lagergren
model is given by Eq. (2):
log qeqt = log qe
k1t
2:303
2
where k
1
is the pseudo-rst order rate constant (min
1
), q
e
and q
t
are
the amount of Cu(II) ions adsorbed at equilibrium and time t,
Table 1
Physical properties of chitosan, bentonite, CCB and CCB-ECH adsorbents.
Chitosan Bentonite CCB CCB-ECH
Total pore volume (cm
3
/g) b0.0001 0.0119 0.0083 0.0049
Average pore diameter (nm) 6.75 5.67 5.12 6.34
Micropore area (m
2
/g) 0.18 28.19 13.98 10.45
BET surface area (m
2
/g) 3.40 94.28 48.81 38.96
Langmuir surface area (m
2
/g) 5.06 96.12 49.43 40.10
4 6 8 10 12 14
2
a
b
c
d
Fig. 1. XRD pattern of (a) bentonite, (b) chitosan, (c) CCB, and (d) CCB-ECH.
156 M.L.P. Dalida et al. / Desalination 275 (2011) 154159
respectively [22]. This equation describes a reversible equilibrium
between the solution and adsorbent [23].
The pseudo-second order equation can be expressed as:
t
qt
=
1
k2qe
2
+
t
qe
3
where k
2
(g/mgmin) is the kinetic rate constant [22].
The kinetic rate constants of the pseudo-rst order and pseudo-
second order equations for CCB and CCB-ECH are listed in Table 4. For
both adsorbents, the adsorption data have very low correlation
coefcient values for the pseudo-rst order equation. The pseudo-
second order equation correlated well with the experimental kinetic
data, as given by the high R
2
values (R
2
N0.99). This implies that
chemisorption is the rate-limiting step of Cu(II) adsorption for CCB
and CCB-ECH beads. It is likely that the adsorption process involves
the formation of covalent bond through sharing of electrons between
Cu(II) ions and the binding sites of the adsorbent [17]. The values of
the pseudo-second order kinetic rate constant (k
2
) values were
observed to increase with decreasing initial Cu(II) concentration. A
high concentration means that there were more Cu(II) ions competing
against each other to be adsorbed, hence a delay in the attainment of
equilibrium and lower k
2
values.
3.5. Adsorption equilibrium
The isotherm study was carried out by varying the initial
concentration (100 to 2000 mg/L) at 25 C and an optimum pH 4.
Isotherm models, namely Langmuir model and Freundlich model,
were used to describe the interaction between the solute and the
adsorbent.
The Langmuir isotherm describes an adsorption occurring on the
surface that has a nite number of sites with similar energy levels. It is
a model based on the following assumptions: homogenous adsorption
occurring on a monolayer surface coverage and without net
interaction between the adsorbed species [17]. It follows Henry's
law at low concentrations. The Langmuir equation has the form of:
1
qe
=
1
qmL
+
1
bqmLCe
4
where C
e
is the equilibrium concentration of Cu(II) (mg/L), q
e
is the
adsorption capacity at equilibrium (mg/g), q
mL
is the maximum
adsorption capacity at monolayer coverage (mg/g), and b is a
Langmuir constant that describes the afnity of sorbate to the binding
sites of the adsorbent (mL/mg) [7].
The Freundlich isotherm is represented by the equation:
log qe = log KF +
1
n
log Ce 5
where n and K
F
are the Freundlich constants, which are the intensity
and relative adsorption capacity, respectively [6]. It assumes adsorp-
tion by means of monolayer coverage by the solute on an energetically
heterogeneous surface [4].
Based on Table 5, the Cu(II) adsorption using the CCB beads shows
a good t to the Langmuir isothermdue to its high R
2
values (R
2
N0.98)
at pH 2 to pH 4. On the other hand, the adsorption using CCB-ECH
beads has a better t to the Freundlich isotherm model (R
2
N0.98). It
indicates that CCB is a homogenous adsorbent whereas CCB-ECH is a
heterogeneous type of adsorbent. The presence of the short ECH chain
on the CCB beads contributed to an increase in the existence of
heterogeneities in the surface of CCB-ECH beads. In a previous study,
similar results were obtained, where crosslinked chitosancellulose
showed more heterogeneity than chitosancellulose [15].
The q
mL
values were observed to increase as the pH became less
acidic from pH 1 to pH 4 for both adsorbents. In addition, CCB beads
have higher q
mL
capacity than CCB-ECH beads, except at pH 1. From
the Langmuir model, the maximum q
mL
values of CCB and CCB-ECH
beads are 12.21 mg/g and 9.43 mg/g, respectively. The Langmuir
constant, b can be used to compute for a dimensionless separation
factor, R
L
, as:
RL =
1
1+bC0
6
where b is the Langmuir equilibrium constant (L/mg) and C
0
is the
initial Cu(II) concentration (mg/L) [23]. The R
L
values indicate if the
adsorption is unfavorable (R
L
N1), irreversible (R
L
=0), or favorable
(0bR
L
b1). At pH 4, the R
L
values for CCB (0.011480.00058) and CCB-
ECH (0.001230.00006) indicate that Cu(II) uptake is favorable using
these two adsorbents.
The values of Freundlich parameter, n, for CCB and CCB-ECH are
greater than unity. It implies that the adsorption intensity is favorable
at high concentrations [4]. The calculated K
F
values follow the same
trend as shown by the q
mL
values, where CCB beads have higher K
F
values than those of CCB-ECH beads.
3.6. Regeneration study
Table 6 lists the adsorption uptake capacity of Cu(II) on the CCB-
ECH beads in the adsorptiondesorptionadsorption cycle. The
Table 2
Effect of initial concentration and pH on the percent (%) removal of Cu(II) ions.
Initial Cu(II) concentration
(mg/L)
CCB CCB-ECH
pH 1 pH 2 pH 3 pH 4 pH 1 pH 2 pH 3 pH 4
100 42.07 99.98 99.96 99.87 33.70 99.99 99.99 99.98
500 7.08 99.70 99.75 99.70 22.14 95.23 97.88 99.14
1000 14.78 87.56 93.52 94.46 14.82 81.00 86.10 91.76
2000 12.38 68.52 72.40 72.94 9.64 53.10 63.08 66.44
Table 3
Effect of initial concentration and pH on the adsorption of Cu(II) ions.
Initial Cu(II) concentration
(mg/L)
CCB CCB-ECH
pH 1 pH 2 pH 3 pH 4 pH 1 pH 2 pH 3 pH 4
100 0.50 1.20 1.20 1.20 0.40 1.20 1.20 1.20
500 0.42 5.98 5.98 5.98 1.33 5.71 5.87 5.95
1000 1.77 10.51 11.22 11.33 1.78 9.72 10.33 11.01
2000 2.97 16.44 17.37 17.50 2.31 12.74 15.14 15.94
157 M.L.P. Dalida et al. / Desalination 275 (2011) 154159
desorption agents utilized were tap water (pH 7) and HCl solution at
pH1 and pH3. It was observed that desorption using tap water (pH7)
and HCl solution (pH 3) yields very low Cu(II) recovery values of
0.02% to 3.37% and 0.07% to 13.45%, respectively. This indicates the
stability of the CCB-ECH beads in capturing Cu(II) ions under slightly
acidic and neutral conditions. It implies the potential of the CCB-ECH
beads to be used as a material in building permeable reactive barrier
(PRBs). Stability and efciency of capturing contaminants from
polluted groundwater are desirable characteristics of a material to
be used for PRBs.
Among the three desorbing agents, HCl solution (pH 1) has the
highest amount of Cu(II) recovered from CCB-ECH beads. After one
desorption cycle, the CCB-ECH beads were subjected to a second
adsorption cycle. The adsorption capacity at 100 and 500 mg/L using
pH 1 HCl solution have similar q
e
and % removal values as the rst
adsorption cycle. On the other hand, the q
e
at 1000 and 2000 mg/L
slightly dropped from 90.68% to 77.16% and 67.37% to 42.65%,
respectively. This indicates good regeneration capability of CCB-ECH
beads using HCl solution (pH 1) as the desorbing agent.
4. Conclusion
In this study, the removal capacities of CCB and CCB-ECH beads in
adsorbing Cu(II) from aqueous solution were investigated. Effects of
pH and initial concentration, as well as the kinetics and isotherm of
the Cu(II) adsorption, were determined. Removal of Cu(II) using CCB
can be best described by the Langmuir model while adsorption using
CCB-ECH correlates well with the Freundlich model. CCB beads have
better Cu(II) adsorption capacity over CCB-ECH beads for the entire
pH range (14) studied, but crosslinked CCB beads have better
regeneration properties even in acidic media. The pseudo-second
order equation best describes the Cu(II) adsorption onto CCB and CCB-
ECH. CCB-ECH can be utilized to construct lters for the remediation
of contaminated groundwater or in wastewater treatment for removal
of Cu(II).
Acknowledgements
The authors would like to acknowledge the Taiwan National
Science Council (NSC 99-2221-E-041-017) and the Philippine ERDT
scholarship for their nancial support.
References
[1] W.S. Wan Ngah, A. Kamari, Y.J. Koay, Equilibrium and kinetics studies of
adsorption of copper (II) on chitosan and chitosan/PVA beads, Int. J. Biol.
Macromol. 34 (2004) 155161.
[2] C.M. Futalan, C.C. Kan, M.L. Dalida, C. Pascua, M.W. Wan, Fixed-bed column studies
on the removal of copper using chitosan immobilized on bentonite, Carbohydr.
Polym. (2010), doi:10.1016/j.carbpol.2010.08.043.
[3] C.M. Futalan, C.C. Kan, M.L. Dalida, K.J. Hsien, C. Pascua, M.W. Wan, Comparative
and competitive adsorption of copper, lead, and nickel using chitosan immobi-
lized on bentonite, Carbohydr. Polym. (2010), doi:10.1016/j.carbpol.2010.08.013.
[4] W.S. Wan Ngah, S. Fatinathan, Adsorption of Cu(II) ions in aqueous solution using
chitosan beads, chitosan-GLA beads and chitosan-alginate beads, Chem. Eng. J.
143 (2008) 6272.
[5] S.R. Popuri, Y. Vijaya, V.M. Boddu, K. Abburi, Adsorptive removal of copper and
nickel ions from water using chitosan-coated PVC beads, Bioresour. Technol. 100
(2009) 194199.
[6] A. Kamari, W.S. Ngah, Isotherm, kinetic, and thermodynamic studies of lead and
copper uptake of H
2
SO
4
modied chitosan, Colloids Surf. 73 (2009) 257266.
[7] W.S. Wan Ngah, S. Ab Ghani, A. Kamari, Adsorption behavior of Fe(II) and Fe(III)
ions in aqueous solution on chitosan and crosslinked chitosan beads, Bioresour.
Technol. 96 (2005) 443450.
[8] W. Tan, Y. Zhang, Y.S. Szeto, L. Liao, A novel method to prepare chitosan/
montmorillonite nanocomposites in the presence of hydroxy-aluminum oligo-
meric cations, Compos. Sci. Technol. 68 (2008) 29172921.
[9] A.H. Chen, S.C. Liu, C.Y. Chen, C.Y. Chen, Comparative adsorption of Cu(II), Zn(II),
and Pb(II) ions in aqueous solution on the crosslinked chitosan with epichloro-
hydrin, J. Hazard. Mater. 154 (2008) 184191.
[10] A. Kamari, W.S. Wan Ngah, L.K. Liew, Chitosan and chemically modied chitosan
beads for acid dyes sorption, J. Environ. Sci. 21 (2009) 296302.
Table 6
Desorption study of Cu(II) from CCB-ECH beads.
pH Cu(II) concentration
(mg/L)
1st cycle adsorption 1st cycle desorption 2nd cycle adsorption
q
e
(mg/g) % removal C
e
(mg/L)
% desorbed q
e
(mg/g)
% removal
pH 7 100 1.1990 99.94% 0.02 0.02% 1.1912 99.28%
500 5.9526 99.23% 0.04 0.01% 5.1431 85.73%
1000 10.8843 90.73% 2.74 0.30% 7.2773 60.66%
2000 16.0570 66.91% 45.04 3.37% 5.9947 24.98%
pH 3 100 1.1990 99.94% 0.07 0.07% 1.1972 99.79%
500 5.9638 99.42% 1.20 0.24% 5.7138 95.25%
1000 10.9738 91.47% 12.60 1.38% 7.6649 63.72%
2000 16.0829 67.02% 180.27 13.45% 7.9799 33.25%
pH 1 100 1.1995 99.97% 5.79 5.80% 1.1988 99.91%
500 5.9637 99.43% 165.30 33.25% 5.8705 97.87%
1000 10.8781 90.68% 470.67 51.92% 9.2562 77.16%
2000 16.1649 67.37% 847.33 62.89% 10.2338 42.65%
Table 4
Kinetic parameters for pseudo-rst order and pseudo-second order equation for CCB
and CCB-ECH beads at pH 4.
Adsorbent Initial concentration
(mg/L)
Pseudo-rst order Pseudo-second order
k
1
(min
1
)
R
2
k
2
(g/mgmin)
R
2
CCB 100 0.00002 0.12600 0.83370 0.99910
500 0.00020 0.10900 0.16650 0.99980
1000 0.00030 0.13130 0.08330 0.99970
2000 0.00030 0.13000 0.05280 0.99960
CCB-ECH 100 0.00002 0.10050 0.83360 0.99990
500 0.00020 0.13410 0.16610 0.99980
1000 0.00030 0.14500 0.08530 0.99890
2000 0.00030 0.12410 0.05850 0.99970
Table 5
Langmuir and Freundlich isotherm constants for Cu(II) adsorption onto CCB beads and
CCB-ECH beads.
Adsorbent pH Langmuir Freundlich
b
(mL/mg)
q
mL
(mg/g)
R
2
n K
F
(mg/g)
R
2
CCB pH 4 0.8612 12.2100 0.9972 3.3400 3.7144 0.8981
pH 3 2.9202 10.5042 0.9912 3.7078 3.6669 0.9316
pH 2 6.1726 9.6432 0.9884 4.1391 3.6813 0.9462
pH 1 0.0150 1.0335 0.2632 2.0056 2.3564 0.5924
CCB-ECH pH 4 8.0992 9.4251 0.9860 4.0225 3.5367 0.9885
pH 3 15.0000 9.2593 0.9858 4.4248 3.4072 0.9914
pH 2 23.7400 8.4246 0.9871 4.9702 3.2144 0.9980
pH 1 0.0028 2.6116 0.9996 1.8646 2.6222 0.9748
158 M.L.P. Dalida et al. / Desalination 275 (2011) 154159
[11] M. Wan, C.C. Kan, B.D. Rogel, M.L.P. Dalida, Adsorption of copper (II) and lead (II)
ions from aqueous solution on chitosan-coated sand, Carbohydr. Polym. 80
(2010) 891899.
[12] W.S. Wan Ngah, M.A.K.M. Hanaah, S.S. Yong, Adsorption of humic acid from
aqueous solution on crosslinked chitosan-epichlorohydrin beads: kinetics and
isotherm studies, Colloids Surf. B 65 (2008) 1824.
[13] A.H. Chen, Y.Y. Huang, Adsorption of Remazol Black 5 from aqueous solution by
templated crosslinked-chitosans, J. Hazard. Mater. 177 (2010) 668675.
[14] R.S. Vieira, M.M. Beppu, Interaction of natural and crosslinked chitosan
membranes with Hg(II) ions, Colloids Surf. A 279 (2006) 196207.
[15] N. Li, R. Bai, Copper adsorption on chitosancellulose hydrogel beads: behaviors
and mechanisms, Sep. Purif. Technol. 42 (2005) 237247.
[16] K.G. Bhattacharya, S.S. Gupta, Adsorption of heavy metals on natural and modied
kaolinite and montmorillonite: a review, Adv. Colloid Interface Sci. 140 (2008)
114131.
[17] P. Wu, W. Wu, S. Li, N. Xing, N. Zhu, P. Li, J. Wu, Removal of Cd
+2
from aqueous
solution by adsorption using Fe-montmorillonite, J. Hazard. Mater. 169 (2009)
824830.
[18] W.S. Wan Ngah, C.S. Endud, R. Mayanar, Removal of copper (II) ions fromaqueous
solution onto chitosan and cross-linked chitosan beads, React. Funct. Polym. 50
(2002) 181190.
[19] M.W. Wan, I.G. Petrisor, H.T. Lai, D. Kim, T.F. Yen, Copper adsorption through
chitosan immobilized on sand to demonstrate the feasibility for in situ
decontamination, Carbohydr. Polym. 55 (2004) 249254.
[20] P. Monvisade, O. Siriphannon, Chitosan intercalated montmorillonite: preparation,
characterization and cationic dye adsorption, Appl. Clay Sci. 42 (2009) 427431.
[21] M.O. Machado, E. Lopes, K.S. Sousa, C. Airoldi, The effectiveness of the protected
amino group on crosslinked chitosans for copper removal and the thermodynamics
of interaction at the solid/liquid interface, Carbohydr. Polym. 77 (2009) 760766.
[22] B.H. Hameed, I.A.W. Tan, A.L. Ahmad, Adsorption isotherm, kinetic modeling and
mechanism of 2,4,6-trichlorophenol on coconut-husk based activated carbon,
Chem. Eng. J. 144 (2008) 235244.
[23] Y. Vijaya, S. Popuri, V. Boddu, A. Krishnaiah, Modied chitosan and calcium
alginate biopolymer sorbents for removal of nickel (II) through adsorption,
Carbohydr. Polym. 72 (2008) 261271.
[24] L. Jin, R. Bai, Mechanisms of lead adsorption on chitosan/PVA hydrogel beads,
Langmuir 18 (2002) 97659770.
159 M.L.P. Dalida et al. / Desalination 275 (2011) 154159

You might also like