Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Journal of Environmental Management 91 (2010) 798–806

Contents lists available at ScienceDirect

Journal of Environmental Management


journal homepage: www.elsevier.com/locate/jenvman

Review

A review and experimental verification of using chitosan and its derivatives as


adsorbents for selected heavy metals
Feng-Chin Wu a, Ru-Ling Tseng b, Ruey-Shin Juang c, d, *
a
Department of Chemical Engineering, National United University, Miao-Li 360, Taiwan
b
Department of Safety, Health and Environmental Engineering, National United University, Miao-Li 360, Taiwan
c
Department of Chemical Engineering and Materials Science, Yuan Ze University, 135 Yuan-Tung Road, Chung-Li 32003, Taiwan
d
Fuel Cell Center, Yuan Ze University, Chung-Li 32003, Taiwan

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

Article history: A literature survey on liquid-phase adsorption of selected heavy metals including Cu(II), Zn(II), Ni(II),
Received 30 April 2009 Cd(II), Pb(II), Hg(II), and Cr(VI) on chitosan (CTS) and its derivatives was made from the viewpoint of
Received in revised form adsorption capacity. This parameter was obtained from the Langmuir fit of isotherm data. The magnitude
20 October 2009
of adsorption capacity of heavy metals on pristine CTS was also used to discuss the mechanism of
Accepted 28 October 2009
Available online 14 November 2009
adsorption; that is, how many amino groups in CTS chains would coordinate with one heavy metal ion.
Furthermore, a newly defined parameter, the approaching equilibrium factor RL, was proposed to
quantitatively indicate the favorability of the related adsorption process and to judge the correctness of
Keywords:
Chitosan adsorption capacity determined by the Langmuir equation.
Chemical modification Ó 2009 Elsevier Ltd. All rights reserved.
Heavy metals
Adsorption capacity
Langmuir isotherm

1. Introduction mainly results from deacetylation of chitin (Guibal et al., 1994;


Onsoyen and Skaugrud, 1990; Yang and Zall, 1984). Of these alter-
Liquid-phase adsorption is an efficient method for the removal natives, CTS appears to be more attractive for this purpose because
of color, odor, and organic matter from process or waste effluents. chitin is the second abundant polymer in nature next to cellulose.
Powdered or granular activated carbons are the most widely used Also, CTS has many useful features such as hydrophilicity, biode-
adsorbents because they have an excellent adsorption ability for gradability, and anti-bacterial property. This biopolymer is a known
common organic matter, but their use is usually limited due to high sorbent, effective in the uptake of transition metals because the
cost (Bhattacharya and Venkobachar, 1984; Wang et al., 2003). This amino groups on CTS chains serve as the coordination sites (Guibal
has led many workers to search for cheaper substitutes such as fly et al., 1994; Mitani et al., 1991; Onsoyen and Skaugrud, 1990). In
ash, silica gel, zeolite, lignin, seaweed, wool wastes, agricultural contrast to chitin, CTS has superior adsorption ability for heavy
wastes (bagasse pith, maize cob, coconut shell, rice husk, etc.), metals due to its higher content of amino groups (Yang and Zall,
chitin, chitosan (CTS), and clay materials (diatomaceous earth, 1984). Chemical modifications of CTS such as carboxyalkyl-substi-
montmorillonite, bentonite, kaolinite, etc.). Compared to activated tution, aldehyde-crosslinking, ligand-crosslinking, and poly-
carbons, they have been studied with varying success for the amination are accessible to prevent it from dissolution in acidic
removal of pollutants from aqueous streams, especially for heavy media (pH < 2) or to enhance adsorption ability, or both (Guibal
metals and reactive dyes (Bailey et al., 1999; Wang et al., 2003). et al., 1994; Juang and Ju, 1997; Wu et al., 1999).
CTS is a partially acetylated glucosamine biopolymer that exists The chemical structure of CTS chain (C6H11NO4)n (n ¼ 1 and 3) is
in the cell wall of some fungi such as Mucorales strains; however, it shown in Fig. 1. CTS has a mass of 161 g per unit monomer, and has
a concentration of amino group (–NH2) of 6.21 mmol/g if it is
completely deacetylated. The adsorption of transition metals on
CTS is known to be mainly effected via coordination with the
* Corresponding author. Department of Chemical Engineering and Materials
Science, Yuan Ze University, 135 Yuan-Tung Road, Chung-Li 32003, Taiwan.
unprotonated amino group (Monteiro and Airoldi, 1999). The
Tel.: þ886 3 4638800x2555; fax: þ886 3 4559373. coordination of divalent heavy metals, e.g. Cu2þ, with the amino
E-mail address: rsjuang@saturn.yzu.edu.tw (R.-S. Juang). groups of CTS can be achieved by a molar ratio of 1:1 (Juang et al.,

0301-4797/$ – see front matter Ó 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jenvman.2009.10.018
F.-C. Wu et al. / Journal of Environmental Management 91 (2010) 798–806 799

Nomenclature HA-KA-CTS hydroxylamine hydrochloride modified KA-CTS


HQS-GLA-CTS 8-hydroyquinoline-5-sulfonic acid modified GLA-
BHFMBME-CTS N-N0 -[bis(2-hydroxy-3-formyl-5-methylbenzyl- CTS
dimethyl)]ethylenediamine crosslinked CTS KA-CTS a-ketoglutaric acid grafted CTS
BPMAMFP-CTS 2[bis(pyridylmethyl)aminomethyl]-4-methyl- KL Langmuir constant (L/mmol)
6-formylphenol functionalized CTS LY-CLCTS L-lysine modified CLCTS
C0 initial concentration of solute in the aqueous phase PEI-CTS poly(ethylenimine) modified CTS
(mmol/L) 2PM-CTS N-(2-pyridylmethyl) substituted CTS
Ce equilibrium concentration of solute in the aqueous 4PM-CTS N-(4-pyridylmethyl) substituted CTS
phase (mmol/L) qe equilibrium amount of adsorption (mmol/g)
CLCTS commercially available crosslinked CTS qmon amount of adsorption corresponding to monolayer
CTS chitosan coverage (mmol/g)
CM-CTS N,O-carboxymethyl substituted CTS r2 correlation coefficient of linearized Langmuir fit (–)
DTPA-CTS diethylenetriaminepentaacetic acid grafted CTS RB-CTS reactive blue 2 bound CTS
ECH-CTS epichlorohydrin crosslinked CTS RL approaching equilibrium factor based on Langmuir
EDTA-CTS ethylenediaminetetraacetic acid grafted CTS equation defined in Eq. (6) (–)
EGDE-CTS ethylene glycol diglycidylether crosslinked CTS RSF dimensionless separation factor of Langmuir equation
GL-CLCTS glycine modified CLCTS defined in Eq. (5) (–)
GLA-CTS glutaraldehyde crosslinked CTS V volume of the solution (L)
GO-CTS glyoxal crosslinked CTS W amount of the dry CTS used (g)

1999; Li and Bai, 2005), 1:2 (Paulino et al., 2007; Steenkamp et al., 2. Isotherm model
2002; Vijaya et al., 2008), and 1:4 (Zhao et al., 2007), as shown in
Fig. 2. It is also possible that multiple types of metal coordination 2.1. Linearized form of Langmuir equation
exist in one CTS chain, which includes intramolecular chelation
with three different configurations and intermolecular chelation Adsorption isotherm is important to describe how solutes
with four different forms (Debbaudt et al., 2004). Generally interact with adsorbent and so is critical in optimizing the use of
speaking, the type of coordination plays a crucial role in adsorption adsorbent. Correlation of isotherm data by empirical or theoretical
performance (capacity, kinetics, etc.). The theoretical (maximum) equations is thus essential to practical operation. The widely used
adsorption capacity of divalent metal is 6.2, 3.1, and 1.6 mmol/g Langmuir equation is given as:
when divalent metal and –NH2 group are bound by a molar ratio of
1:1, 1:2, and 1:4, respectively. qmon KL Ce
qe ¼ (1)
In this work, the adsorption of heavy metals including Cu2þ, 1 þ KL Ce
Zn , Ni2þ, Cd2þ, and Pb2þ from sulfate and nitrate solutions on

where qe (mmol/g) and Ce (mmol/L) are the equilibrium concen-
pristine CTS prepared from cuttlefish wastes was studied.
trations of solute in solid and liquid phases, respectively. Also, KL is
Furthermore, literature survey on adsorption capacity of selected
the Langmuir constant (L/mmol) and qmon is the amount of
heavy metals on pristine and chemically modified CTS was made in
adsorption corresponding to complete monolayer coverage (mmol/
terms of the Langmuir parameters and a newly proposed
g). The value of qmon is recognized as the adsorption capacity, which
approaching equilibrium factor RL. The mechanism of adsorption
is commonly a measure of adsorption ability of an adsorbent.
(coordination) was also discussed. It is expected that this paper can
The Langmuir parameters of KL and qmon can be graphically
give a rather overall insight in application potential of metal
obtained by the following linearized from:
adsorption using CTS and its derivatives.
     
Ce 1 1
¼ þ Ce (2)
qe KL qmon qmon
CH2OH
O O 2.2. Dimensionless form of Langmuir equation

OH Let Cref be a reference concentration that is assigned as the


highest Ce during an adsorption process. In this case, the corre-
H2N
sponding qe is set to be qref. From Eq. (2), we thus have
a n=1 !    
Cref 1 1
¼ þ Cref (3)
qref KL qmon qmon
CH2OH NH2 CH2OH
O O O O Dividing Eq. (3) by Eq. (2), we obtain
OH
! !" #
OH OH qe Ce 1 þ KL Cref
O O ¼   (4)
H 2N H2 N qref Cref 1 þ KL Cref Ce =Cref
CH2OH

b n=3 Eq. (4) is called the dimensionless form of Langmuir equation, in


which KLCref is the unique parameter when (qe/qref) is plotted
Fig. 1. Chemical structures of chitosan chains (C6H11NO4)n (a) n ¼ 1 and (b) n ¼ 3. against (Ce/Cref).
800 F.-C. Wu et al. / Journal of Environmental Management 91 (2010) 798–806

1.0

0.8
N O
Cu
CH2OH 0.6

qe/qref (-)
O O
O N
RL
0.4
OH 1
0.6
+
2+
Cu(H2N) + 2H 0.3
0.2 0.1
0.04
a b 0.01

0.0
0.0 0.2 0.4 0.6 0.8 1.0
CH2OH CH2OH
O O Ce/Cref (-)

O O Fig. 3. Characteristic equilibrium curves based on the Langmuir equation.


OH OH

H 2N H 2N
3. Materials and methods
Cu2+
3.1. Preparation of pristine CTS
NH2 NH2

HO
Dried cuttlefish wastes were immersed in 5 wt.% NaOH for 18 h
O O HO
to remove proteins and then in 5 wt.% HCl for 18 h to remove CaCO3
O O (weight ratio of the wastes to the solution, 0.1). The resulting
CH2OH CH2OH insoluble solid (40 g), chitin, was deacetylated in 50 wt.% NaOH
(800 g) at 90  C for 3 h. The final CTS was washed three times with
c deionized water (Millipore Milli-Q RO) and dried at 50  C in

a vacuum. Prior to further processing, the flakes were ground and
Fig. 2. Coordination of Cu with the –NH2 groups of chitosan by a molar ratio of (a)
sieved into a particle size range of 0.6–2.0 mm (10–30 mesh).
1:1, (b) 1:2, and (c) 1:4.
Physical properties of pristine CTS such as the degree of
deacetylation, molar mass, and BET surface area were measured.
2.3. Separation factor The former was obtained to be 96 mol% following the method of
Guibal et al. (1994). The molar mass was determined to be
The basic characteristics of Langmuir equation has been often 1.85  106 using the Mark–Houwink equation and the viscosity data
represented in terms of a dimensionless separation factor, RSF, of solutions containing different amounts of CTS in 0.1 mol/L acetic
defined as (Hall et al., 1966; McKay et al., 1987) acid and 0.2 mol/L NaCl (Roberts and Domszy, 1982). The BET
surface area was determined to be 12.5 m2/g from N2 adsorption
1
RSF ¼ (5) isotherms using a sorptiometer (Porous Materials Inc., BET-202A).
1 þ KL C0
where C0 is the highest, initial solute concentration in liquid phase. 3.2. Adsorption experiments
The RSF value implies that adsorption is either unfavorable (RSF > 1),
linear (RSF ¼ 1), favorable (0 < RSF < 1), or irreversible (RSF ¼ 0). Analytical grade reagents CuSO4, Cu(NO3)2, ZnSO4, Zn(NO3)2,
Based on similar concept, another dimensionless separation factor NiSO4, Ni(NO3)2, Cd(NO3)2, and Pb(NO3)2 were purchased from
RL (also, named as the approaching equilibrium factor) is defined as Merck Co. and used as received. The aqueous solution was prepared
follows: by dissolving metal salt in deionized water to the required
concentration (0.5–5 mmol/L) without pH adjustment. The initial
1 concentration of metal was kept low enough such that the change
RL ¼ (6)
1 þ KL Cref in solution pH before and after experiments was negligible.
In equilibrium experiments, an amount of CTS (0.1 g) and 0.1 L of
Combination of Eqs. (4) and (6) yields
an aqueous phase were placed in a 0.25-L glass-stoppered flask and
! !" # stirred for 5 days using a bath controlled at 30  C (Firstek Model
qe Ce 1 B603, Taiwan). Preliminary experiments had shown that the
¼   (7)
qref Cref RL þ ð1  RL Þ Ce =Cref adsorption studied was complete after 3 days. After adsorption
equilibrium, the aqueous-phase concentration of metals was
Eq. (7) is the dimensionless form of Langmuir equation in terms analyzed with a Shimadzu atomic absorption spectrophotometer
of RL. The characteristic equilibrium curves based on Eq. (7) are (Model AA68). Each experiment was duplicated under identical
shown in Fig. 3. It is seen that the curvature becomes smaller when conditions. The amount of adsorption at equilibrium qe was
RL is smaller. In the region of favorable adsorption (RL < 1), the obtained as follows:
equilibrium curve is subdivided into three types, slightly curved
(0.1 < RL < 1), largely curved (0.01 < RL < 0.1), and pseudo-rectan- ðC0  Ce ÞV
qe ¼ (8)
gular (RL < 0.01). W
F.-C. Wu et al. / Journal of Environmental Management 91 (2010) 798–806 801

3.5 Langmuir fit, r2, is 0.997 and 0.985, respectively). This is likely due
to different extents of interaction between SO2 4 and CTS chains
3.0 with NO 3 , leading to different types or extents of Cu(II) coordina-
tion (Mitani et al., 1991).
2.5 Literature results of Cu2þ adsorption on CTS and its derivatives
Cu(NO3)2 are compiled in Table 1, where they are classified into four groups
qe (mmol/g)

2.0 CuSO4 according to the nature of CTS; i.e., pristine CTS (Nos. 1–13), CTS
beads (Nos. 14–16), physically modified CTS (Nos. 17–19), and
1.5 chemically modified CTS (Nos. 20–32). In each group, it is listed in
the decreasing order of adsorption capacity qmon. In Nos. 1–3, qmon
1.0 is in the range 2.83–3.13 mmol/g, which roughly corresponds to the
theoretical value of 3.10 mmol/g for the coordination type of Fig. 3b.
0.5 Table 1 also lists the Langmuir constant KL, reference concentration
Cref, and approaching equilibrium factor RL. It is noticed that the
0.0 much small qmon in Nos. 10–13 is possibly a result of low Cref.
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0
The large qmon value of 3.21 mmol/g obtained using CTS bead
Ce (mmol/L) (No. 14) also reveals that one Cu2þ is coordinated with two –NH2
group. On the other hand, physically mixing of sand solids with CTS
Fig. 4. Isotherm data of Cu2þ adsorption on chitosan at 30  C (the curves are calculated to form irregular composite (No. 17) has no or even a positive effect
by the Langmuir equation).
on adsorption ability of CTS. However, this is not the case as the
added material has similar molecular structure or comparable
where V is the volume of the solution (L), and W is the weight of CTS
dimension such as cellulose or PVA (Nos. 18 and 19) because their
used (g).
presence leads to some steric hindrance for adsorption.
For chemical modification using complexing (crosslinking)
4. Results and discussion agents such as glutaraldehyde (GLA) (Cao et al., 2001; Juang et al.,
2001; Wan Ngah et al., 2002), glyoxal (GO) (Juang et al., 2001), N-
4.1. Adsorption of Cu2þ N0 -[bis(2-hydroxyl-3-formyl-5-methylbenzyl-dimethyl)]ethylene-
diamine (BHFMBME) (Vasconcelos et al., 2008), and ethylene glycol
Fig. 4 shows the isotherms of Cu2þ adsorption on pristine CTS at diglycidylether (EGDE) (Wan Ngah et al., 2002), the adsorption
30  C,
where the curves are obtained by Langmuir equation. The ability of CTS usually reduces since some –NH2 groups in CTS chain
amount of Cu(II) adsorbed from sulfate solution is larger than that are involved in crosslinking. However, the introduction of new
from nitrate solution (the correlation coefficient of linearized functional groups into CTS chain probably increases the interaction

Table 1
The Langmuir parameters and approaching equilibrium factor for Cu2þ taken from literature.

No. Adsorbent qmon (mmol/g) KL (L/mmol) Cref (mmol/L) RL (–) Reference


1 Pristine CTS vs. CuSO4 3.13 4.64 5.0 0.041 This work
2 Pristine CTS 2.85 2.61 5.0 0.071 Juang et al. (1999)
3 Pristine CTS 2.83 10.4 12.0 0.008 Ng et al. (2002)
4 Pristine CTS 2.36 106 31.5 0.0003 Wan et al. (2004)
5 Pristine CTS vs. Cu(NO3)2 2.14 0.44 5.0 0.310 This work
6 Pristine CTS 2.12 15.1 4.5 0.015 Hu et al. (2004)
7 Pristine CTS 1.94 6.99 5.0 0.028 Wu et al. (2000)
8 Pristine CTS 1.59 2.73 15.7 0.023 Alexandre et al. (2008)
9 Pristine CTS 1.42 12.3 4.0 0.020 Hu et al. (2004)
10 Pristine CTS 1.27 153 0.2 0.029 Wan Ngah et al. (2002)
11 Pristine CTS 0.53 142 0.2 0.031 Wan Ngah et al. (2004)
12 Pristine CTS 0.27 5.27 0.8 0.194 Chu (2002)
13 Pristine CTS 0.21 8.90 1.0 0.101 Saiano et al. (2005)
14 CTS bead 3.21 4.83 26.8 0.008 Zhao et al. (2007)
15 CTS bead 2.11a Juang et al. (2001)
16 CTS bead 1.64 15.3 5.0 0.013 Wu et al. (2000)
17 CTS/sand composite 3.59 17.2 31.5 0.002 Wan et al. (2004)
18 CTS/cellulose composite 0.84 27.8 1.6 0.022 Li and Bai (2005)
19 CTS/PVA composite 0.75 129 0.2 0.034 Wan Ngah et al. (2004)
20 GLA-CuCTSb 2.58a Cao et al. (2001)
21 GO-CTS 2.44a Juang et al. (2001)
22 BHFMBME-CTS 1.79 4.04 4.7 0.050 Vasconcelos et al. (2008)
23 GLA-CTS 1.29a Juang et al. (2001)
24 ECH-CTS 0.98 144 0.2 0.031 Wan Ngah et al. (2002)
25 GLA-CTS 0.94 309 0.2 0.014
26 EGDE-CTS 0.72 137 0.2 0.032
27 ECH-CTS 0.62 1.19 6.3 0.118 Coelho et al. (2007)
28 CM-CTS 2.53 492 20.0 0.0001 Sun and Wang (2006)
29 BPMAMFP-CTS 1.72 1.72 7.9 0.069 Justi et al. (2005)
30 2PM-CTS 1.64 2.16 5.0 0.085 Rodrigues et al. (1998)
31 RB-CTS 0.90 1.02 6.3 0.135 Vasconcelos et al. (2007)
32 4PM-CTS 0.71 0.89 5.0 0.183 Rodrigues et al. (1998)
a
The maximum amount of adsorption in that literature.
b
CTS was crosslinked with GLA after CTS was saturated with Cu2þ, and the loaded Cu(II) was then eluted by HCl prior to adsorption experiments.
802 F.-C. Wu et al. / Journal of Environmental Management 91 (2010) 798–806

with Cu2þ, thereby enhancing the adsorption (Wan Ngah et al., agent, the GO-CTS exhibits better adsorption ability (No. 21) likely
2002). Fig. 5 shows the structures of so-called GLA-CTS, BHFMBME- because GO has shorter molecular chain than GLA and thus most of
CTS, and EGDE-CTS. It is obvious that qmon decreases compared to –CHO groups in GO react with –NH2 groups in CTS at a 1:1 molar
pristine CTS and physically modified CTS. For example, they are 1.29 ratio (Juang et al., 2001).
and 0.94 mmol/g with GLA-CTS (Nos. 23 and 25). It is noted that for The second type of chemical modification is that the complexing
No. 20 system the CTS is crosslinked with GLA after CTS was agent is not crosslinked with the –NH2 groups of CTS chain. For
saturated with Cu2þ, and was then washed with HCl to remove example, epichlorohydrin (ECH) reacts with the –CH2OH groups,
loaded Cu(II) prior to adsorption. In this case, the –NH2 group that instead of –NH2 groups, of CTS, as shown in Fig. 5d (Coelho et al.,
remains uncrosslinked are sufficient for metal adsorption, leading 2007; Wan Ngah et al., 2002). Although the –NH2 group remains
to a larger qmon (2.58 mmol/g). Unlike the use of GLA as complexing unchanged in this case, its adsorption ability for Cu2þ still decreases
(Nos. 24 and 27) because the presence of ECH reduces the degree of
CH2OH freedom of the –NH2 groups.
O O Another modification way is substitution or grafting that is
essentially reacted with the –NH2 group of CTS (Nos. 28–32). For
OH example, reaction of CTS and monochloroacetic acid in NaOH
CH2OH solution produces N,O-carboxymethyl-CTS (CM-CTS) (Sun and
O O N Wang, 2006). This is the so-called carboxymethylation. CM-CTS
OH
contains hydroxyl (OH), carboxyl (–COOH), and amino (–NH2)
OH groups in the molecule, making it possible to offer sufficient chelate
N groups for increasing adsorption ability toward heavy metals
N (qmon ¼ 2.52 mmol/g). However, the grafting of CTS with 2[bis(-
CH pyridylmethyl)aminomethyl]-4-methyl-6-formylphenol
(CH2)3 (BPMAMFP) yielded qmon to be 1.72 mmol/g only (No. 29) (Justi
N et al., 2005). On the other hand, the substitution of CTS with N-(2-
CH pyridylmethyl) (2PM) exhibits larger qmon than that with N-(4-
pyridylmethyl) (4PM) as shown in Nos. 30 and 32 of Table 2 (1.64 vs.
N OH
N 0.71 mmol/g). This is because 2PM-CTS is chelated with Cu2þ than
4PM-CTS (Rodrigues et al., 1998).
OH OH As shown in Table 1, KL may vary in a few orders of magnitude
even for the same heavy metal and CTS material. This is because KL
O O O O
is strongly affected by the initial solute concentration C0 (Hall et al.,
CH2OH CH2OH 1966). Thus, KL is not applied for comparison here although it is also
an important parameter that can help us to estimate the selectivity
a GLA-CTS b BHFMBME-CTS of adsorption. This is the reason why the dimensionless form of RSF
(Wan Ngah et al., 2002) (Vasconcelos et al., 2008) (Eq. (5)) or RL (Eq. (6)) is used for further discussion.
Generally speaking, crosslinking or substitution of CTS inevi-
CH2OH tably results in a decrease in adsorption capacity, unless new
O O functional groups such as hydroxyl (OH), carboxyl (–COOH), and
amino (–NH2) are introduced by the complexing agents (Vascon-
OH
celos et al., 2008). The extent of capacity reduction strongly
HN
depends on the relative amount of complexing agent used (e.g.,
degree of crosslinking). This effect should be compensated with the
CH2 required properties of CTS such as the durability in acidic solutions
NH2
HC OH (pH < 2).
CH2 OH
O
Table 2
O O
CH2 The Langmuir parameters and approaching equilibrium factor for Zn2þ taken from
CH2O literature.
H2 C
CH2 No. Adsorbent qmon KL Cref RL (–) Reference
O
(mmol/g) (L/mmol) (mmol/L)
H 2C HOHC 1 Pristine CTS vs. in 4.42 0.20 5.0 0.498 This work
HC OH ZnSO4
CH2 CH2 2 Pristine CTS vs. 1.21 0.39 5.0 0.341 This work
Zn(NO3)2
HN CH2O 3 Pristine CTS 0.21a Lima and
O O Airoldi (2004)
OH 4 Pristine CTS 0.14 1.96 1.0 0.338 Paulino et al.
OH (2007)
5 CTS/carbon 0.92 175 0.5 0.012 Amuda et al.
O O
composite (2007)
CH2OH NH2 6 CTS/carbon 0.78 131 0.5 0.016 Amuda et al.
composite (2007)
c EGDE-CTS d ECH-CTS 7 HA-KA-CTS 0.58a Ding et al.
(2006)
(Wan Ngah et al., 2002) (Vieira and Beppu, 2006b)
8 KA-CTS 0.43a Ding et al.
(2007)
Fig. 5. Chemical structures of some chemically modified chitosan (a) GLA-CTS, (b)
a
BHFMBME-CTS, (c) EGDE-CTS, and (d) ECH-CTS. The maximum amount of adsorption in that literature.
F.-C. Wu et al. / Journal of Environmental Management 91 (2010) 798–806 803

4.2. Adsorption of Zn2þ Table 3


The Langmuir parameters and approaching equilibrium factor for Ni2þ taken from
literature.
The isotherms of Zn2þ adsorption on pristine CTS at 30  C are
shown in Fig. 6a, Like Cu2þ (Fig. 4), the amount of Zn(II) adsorbed No. Adsorbent qmon KL Cref RL (–) Reference
from sulfate solution is larger than that from nitrate solution (both (mmol/g) (L/mmol) (mmol/L)

r2 values are 0.997). Table 2 lists the literature results of Zn2þ 1 Pristine CTS vs. 1.13 0.37 5.0 0.353 This work
Ni(NO3)2
adsorption on CTS and its derivatives. According to the coordina-
2 Pristine CTS 1.11 0.38 1.7 0.611 Paulino et al.
tion type shown in Fig. 3b, the theoretical value is 3.10 mmol/g for (2007)
Zn2þ. The abnormal large qmon of 4.42 mmol/g in No. 1 of Table 2 is 3 Pristine CTS vs. 0.85 1.14 5.0 0.149 This work
possibly due to over-extrapolation from the Langmuir plot at such NiSO4
a high RL (0.498). Similarly, chemically modified CTS gives a much 4 Pristine CTS 0.37 5.93 1.0 0.144 Lima and
Airoldi (2004)
low capacity for Zn2þ than pristine CTS (Nos. 5–9). 5 Pristine CTS 0.10 1.76 1.0 0.362 Saiano et al.
(2005)
6 CTS coated silica 4.33 0.15 8.5 0.435 Vijaya et al.
4.3. Adsorption of Ni2þ (2008)
b a
7 GLA-CuCTS 2.33 Cao et al. (2001)
8 EDTA-CTS 2.10a Inoue et al. (1999)
Fig. 6b shows the isotherm of Ni2þ adsorption on pristine CTS at
9 DTPA-CTS 1.99a Inoue et al. (1999)
30  C, where r2 value are 0.996 and 0.997 in sulfate and nitrate 10 RB-CTS 0.19 1.06 1.7 0.357 Vasconcelos et al.
media, respectively. Literature results of Ni2þ adsorption on CTS (2007)
and its derivatives are also compiled in Table 3. Although the 11 BPMAMFP-CTS 0.16 44.6 0.9 0.026 Justi et al. (2005)
isotherm obtained in sulfate solution is curved more significantly a
The maximum amount of adsorption in that literature.
than that in nitrate solution, the latter system has a larger qmon b
CTS was crosslinked with GLA after CTS was saturated with Cu2þ, and the loaded
(Nos.1 and 3). This is because the latter system has a larger RL under Cu(II) was then eluted by HCl prior to adsorption experiments.
the conditions studied, leading to a larger qmon. Unlike other metals
such as Cu2þ and Zn2þ, the reasons caused such an unusual
counter-ion effect on Ni2þ adsorption remains unclear at this stage
(Mitani et al., 1991).
2.0 When CTS is physically coated on silica (No. 6), it gives an
abnormal large qmon of 4.33 mmol/g (Vijaya et al., 2008). It is
a Zn2+ system considerably larger than the theoretical value of 3.10 mmol/g
according to the coordination type of Fig. 3b, which is also due to
1.5
the large RL (0.435). In addition to aforementioned GLA-CuCTS (No.
ZnSO4 7), some commonly applied chelating agents such as ethyl-
qe (mmol/g)

Zn(NO3)2 enediaminetetraacetic acid (EDTA) and diethylenetriaminepenta-


1.0 acetic acid (DTPA) are used to crosslink with the –NH2 group of CTS
(Nos. 8–9). In contrast to pristine CTS (e.g., Nos. 1–2), a larger qmon
for Ni2þ are obtained (2.00–2.33 mmol/g) using these chemically
modified CTS (Nos. 7–9). Further studies should be made to clarify
0.5 the possible reasons why.

4.4. Adsorption of Cd2þ


0.0
0 1 2 3 4 5 The isotherms of Cd2þ adsorption on pristine CTS at 30  C are
Ce (mmol/L) illustrated in Fig. 7, indicating the higher amount of Cd(II) adsorbed

0.8 1.8

b Ni 2+
system Pb(NO3)2
1.5
Cd(NO3)2
0.6 CdSO4
1.2
qe (mmol/g)
qe (mmol/g)

0.4 0.9

NiSO4
Ni(NO3)2 0.6

0.2

0.3

0.0 0.0
0 1 2 3 4 5 0 1 2 3 4 5

Ce (mmol/L) Ce (mmol/L)

Fig. 6. Isotherm data of (a) Zn2þ and (b) Ni2þ adsorption on chitosan at 30  C (the Fig. 7. Isotherm data of Pb2þ and Cd2þ adsorption on chitosan at 30  C (the curves are
curves are calculated by the Langmuir equation). calculated by the Langmuir equation).
804 F.-C. Wu et al. / Journal of Environmental Management 91 (2010) 798–806

Table 4 in Fig. 8a, where CTS flakes is crosslinked with GLA (via –NH2
The Langmuir parameters and approaching equilibrium factor for Cd2þ taken from group) and then crosslinked with CS2 (via –CH2OH group) in NaOH.
literature.
Compared to the use of pristine CTS reported in the same article
No. Adsorbent qmon KL Cref RL (–) Reference (No. 3), the adsorption ability significantly increases (around 4.2
(mmol/g) (L/mmol) (mmol/L) times) after such modification. This means that the –OCS 2 group in
1 Pristine CTS 3.40 0.34 5.0 0.354 This work the so-called CS-GLA-CTS adsorbent plays a significant and positive
vs. CdSO4
role on adsorption and/or coordination of Cd2þ.
2 Pristine CTS 0.86 0.45 5.0 0.307 This work
vs. Cd(NO3)2
3 Pristine CTS 0.76 4.05 0.9 0.217 Sankararamakrishnan
et al. (2007)
4.5. Adsorption of Pb2þ
4 Pristine CTS 0.26 1.24 1.0 0.447 Saiano et al. (2005)
5 Pristine CTS 0.0084a Cao et al. (2001)
6 CTS/pectin 0.41a Debbaudt et al. Fig. 7 also shows the isotherm of Pb2þ adsorption on pristine CTS
composite (2004) at 30  C (r2 value is 0.985). Also, literature results of Pb2þ adsorp-
7 CTS/alginate 0.059 0.25 100 0.039 Gotoh et al. (2004) tion on CTS and its derivatives are compiled in Table 5. The present
composite
8 CS-GLA-CTS 3.18 7.42 0.9 0.132 Sankararamakrishnan
results reveal that the qmon for Pb2þ from nitrate solution
et al. (2007) (1.15 mmol/g) is larger than that for Ni2þ (0.85 mmol/g) from
9 BPMAMFP- 0.34 9.44 0.5 0.192 Justi et al. (2005) nitrate solution but smaller than those for Zn2þ (1.21 mmol/g) and
CTS Cu2þ (2.14 mmol/g) from nitrate solutions. The mechanism of metal
10 HQS-GLA- 0.29 2.47 8.9 0.044 Vitali et al. (2008)
adsorption from nitrate media remains unclear, while considering
CTS
11 GLA-CTS 0.077 2.59 8.9 0.042 Vitali et al. (2008) the differences in molecular weight. It is noted that the CTS
a
adsorbents used in Nos. 2 and 3 are derived from microorganisms
The maximum amount of adsorption in that literature.
(Paulino et al., 2007; Saiano et al., 2005). Moreover, the much small
Cref in Nos. 4 and 5 may lead to incorrect estimation of qmon value.
from sulfate solution than from nitrate solution (the r2 values are When CTS is entrapped in polyacrylamide (PAA), the capacity
0.997 and 0.993, respectively). Literature results of Cd2þ adsorption increases (about 2.2 times) in contrast to pristine CTS reported in
on CTS and its derivatives are compiled in Table 4. The qmon value the same article (No. 6 vs. No. 5).
obtained (3.40 mmol/g) for the adsorption of Cd2þ from sulfate
solution (No. 1) is slighter larger than the theoretical value
(3.10 mmol/g) when one Cd2þ is coordinated with two –NH2 4.6. Adsorption of Hg2þ
groups. Apparently, the type or nature of counter-ion (SO2 4 vs.

NO 3 ) has a significant effect on the adsorption of Cd . Literature results of Hg2þ adsorption on CTS and its derivatives
For chemically modified CTS, an abnormal large qmon are compiled in Table 6, indicating that CTS has excellent adsorp-
(3.18 mmol/g) was reported previously (No. 8) (Sankarar- tion ability only after chemical modification. This is mainly because
amakrishnan et al., 2007). The structure of that adsorbent is shown new groups are introduced, which are effective for adsorption. The
largest qmon of 3.06 mmol/g is obtained using the magnetic CTS
CH2OH resin modified with Schiff’s base derived from thiourea and GLA
O (Donia et al., 2008). Fig. 8b shows the structure of the so-called TU-
O
GLA-CTS adsorbent, where ]S is the main group for Hg2þ
OH adsorption. Similarly, other chemically modified CTS reveals
excellent adsorption ability as long as the complexing agents
OCS2Na+
N themselves contain more amino groups such as CR-amine-CTS and
CH2 CR-azole-CTS (Atia, 2005). In a series of experiments using chem-
O CH
O ically modified CTS (Nos. 3, and 7–10), the polyaminated CTS bead
(CH2)3 (via the reaction with polyethylenimine, PEI) was proved to give the
OH HC largest qmon (2.26 mmol/g). It is interesting to note that the
N commonly used adsorbents GLA-CTS and ECH-CTS (Vieira and
N Beppu, 2006b), effective for Cu2þ adsorption (Table 1), are not so
C S
CH efficient for this purpose (Nos. 12–13).
N
(CH2)3
HC
CH (CH2)3
Table 5
HC The Langmuir parameters and approaching equilibrium factor for Pb2þ taken from
N
literature.
N
OH No. Adsorbent qmon KL Cref RL (–) Reference
OH (mmol/g) (L/mmol) (mmol/L)

O O 1 Pristine CTS vs. 1.15 0.44 5.0 0.314 This work


O Pb(NO3)2
H2C O
2 Pristine CTS 0.72 40.4 0.5 0.049 Paulino et al. (2007)
OCS2Na+ CH2OH 3 Pristine CTS 0.68 11.6 1.0 0.079 Saiano et al. (2005)
4 Pristine CTS 0.56 271 0.005 0.043 Ng et al. (2003)
a CS-GLA-CTS b TU-GLA-CTS 5 Pristine CTS 0.38 914 0.008 0.120 Akkaya and
(Sankararamakrishnan et al., 2007) 6 CTS/PAA 0.84 1237 0.008 0.092 Ulusoy (2008)
(Donia et al., 2008)
composite
7 Pristine CTS 0.42 0.20 4.8 0.506 Alexandre et al.
Fig. 8. Chemical structures of some chemically modified chitosan (a) CS-GLA-CTS and
(2008)
(b) TU-GLA-CTS.
F.-C. Wu et al. / Journal of Environmental Management 91 (2010) 798–806 805

Table 6 PEI-CTS, Kawamura et al. (1993) have indicated that the selectivity
The Langmuir parameters and approaching equilibrium factor for Hg2þ taken from decreases in the order Hg2þ > Cd2þ > Zn2þ > Cu2þ > Ni2þ. Based on
literature.
the present results using pristine CTS, the adsorption capacity
No. Adsorbent qmon KL Cref RL (–) Reference decreases in the order Cu2þ > Zn2þ ¼ Pb2þ > Ni2þ. It should be
(mmol/g) (L/mmol) (mmol/L) noted that according to the present review such a comparison is
1 TU-GLA-CTS 3.06 1.00 20.0 0.047 Donia et al. convincible and meaningful only when the reference concentration
(2008)
of metal, Cref, in the experiments is kept sufficiently high.
2 CR-amine-CTS 2.30 0.68 2.0 0.423 Atia (2005)
3 PEI-CTS 2.26a Jeon and Holl
(2003) 5. Conclusions
4 CR-azole-CTS 2.21 0.54 2.0 0.480 Atia (2005)
5 CTS membrane 2.07 7.02 1.9 0.071 Vieira and
6 CTS spheres 1.87 4.21 1.9 0.113 Beppu (2006a) The abilities of pristine and modified CTS for adsorption of
7 Xanthated CTS 1.84a Jeon and Holl selected heavy metals from aqueous media were determined by
8 Phosphorlated 1.55a (2003) Langmuir equation. Generally speaking, crosslinking or substitu-
CTS
tion of CTS led to a significant decrease in adsorption ability. The
9 Carboxylated 1.51a
CTS
extent of capacity reduction strongly depended on the type and
10 GLA-CTS 1.47a degree of crosslinking. This effect should be compensated with the
11 CTS/pectin 0.50 5.62 1.0 0.151 Roberts and required properties of CTS such as the durability in acidic solutions
composite Domszy (1982) (pH < 2). However, there is a possibility to enhance adsorption
12 GLA-CTS 0.19 9.03 1.9 0.056 Vieira and
ability through the selection of complexing agents that contains
Beppu (2006b)
13 ECH-CTS 0.11 20.1 1.9 0.017 Vieira and new groups for effective adsorption of heavy metals such as
Beppu (2006b) hydroxyl (OH), carboxyl (–COOH), and amino (–NH2) groups. For
a
The maximum amount of adsorption in that literature.
example, the adsorption capacity of Cd2þ considerably increased
4.2 times compared to pristine CTS after CTS was crosslinked with
glutaraldehyde (via –NH2 group) and then with CS2 (via –CH2OH
4.7. Adsorption of Cr6þ group) in NaOH solution. The effective coordination site for Cd2þ
was the –OCS 2 group, instead of the –NH2 group. On the other
Table 7 compares the literature results of Cr6þ adsorption on CTS hand, the adsorption ability could be enhanced by introducing
and its derivatives. The pristine CTS gives an adsorption capacity of another adsorbent into CTS. For example, the adsorption capacity of
about 2.98 mmol/g (Nos. 1–2), but CTS bead (No. 3) has a much low Pb2þ increased about 2.2 times in contrast to pristine CTS when CTS
ability (0.66 mmol/g). Like Cd2þ adsorption, chemically modified chains were entrapped in polyacrylamide. The present review
CS-GLA-CTS (Fig. 8a) exhibits an excellent ability for Cr6þ adsorp- indicated that the comparison of adsorption selectivity of heavy
tion (qmon ¼ 12 mmol/g). Unlike Hg2þ adsorption, chemically metals among literature results was convincible only when the
modified TU-GLA-CTS (Fig. 8b) does not have high ability for Cr6þ reference concentration of metal in each experiment was kept high
(4.13 mmol/g), but it is still higher than those using pristine CTS. enough.

4.8. Comparisons on adsorption capacities of selected heavy metals References

Comparison among the adsorption capacities of heavy metals on Akkaya, R., Ulusoy, U., 2008. Adsorptive features of chitosan entrapped in poly-
acrylamide hydrogel for Pb2þ, UO2þ 4þ
2 , and Th . J. Hazard. Mater. 151, 380–388.
pristine and modified CTS is rather difficult from literature results Alexandre, A.T., Santos, L.B., Nozaki, J., 2008. Removal of Pb2þ, Cu2þ, and Fe3þ from
because of the complicated characteristics of CTS and solution battery manufacture wastewater by chitosan produced from silkworm chry-
environments. This is probably the reason why few studies discuss salides as a low-cost adsorbent. React. Funct. Polym. 68, 634–642.
Amuda, O.S., Giwa, A.A., Bello, I.A., 2007. Removal of heavy metal from industrial
metal adsorption selectivity unless the properties of CTS adsor-
wastewater using modified activated coconut shell carbon. Biochem. Eng. J. 36,
bents and solution conditions are well identified. However, Yang 174–181.
and Zall (1984) and Huang et al. (1996) have reported that the Atia, A.A., 2005. Studies on the interaction of mercury(II) and uranyl(II) with
selectivity decreases in the order Cu2þ > Cd2þ > Pb2þ > > Zn2þ and modified chitosan resins. Hydrometallurgy 80, 13–22.
Bailey, S.E., Olin, T.J., Bricka, R.M., Adrian, D.D., 1999. A review of potentially low-cost
Cu2þ ¼ Hg2þ > > Pb2þ ¼ Cd2þ > Ni2þ, respectively, using pristine sorbents for heavy metals. Water Res. 33, 2469–2479.
CTS (CTS used in these two studies is different). Using polyaminated Bhattacharya, A.K., Venkobachar, C., 1984. Removal of cadmium(II) by low cost
adsorbents. J. Environ. Eng. ASCE 110, 110–122.
Cao, Z., Ge, H., Lai, S., 2001. Studies on synthesis and adsorption properties of chi-
tosan cross-linked by glutaraldehyde and Cu(II) as template under microwave
Table 7 irradiation. Eur. Polym. J. 37, 2141–2143.
The Langmuir parameters and approaching equilibrium factor for Cr6þ taken from Chu, K.H., 2002. Removal of copper from aqueous solution by chitosan in prawn
literature. shell: adsorption equilibrium and kinetics. J. Hazard. Mater. 90, 77–95.
Coelho, T.C., Laus, R., Mangrich, A.S., de Favere, V.T., Laranjeira, M.C.M., 2007. Effect
No. Adsorbent qmon KL Cref RL (–) Reference of heparin coating on epichlorohydrin cross-linked chitosan microspheres on
(mmol/g) (L/mmol) (mmol/L) the adsorption of copper(II) ions. React. Funct. Polym. 67, 468–475.
Debbaudt, A.L., Ferreira, M.L., Gschaider, M.E., 2004. Theoretical and experimental
1 Pristine CTS 3.00 0.057 19.2 0.476 Sankararamakrishnan
study of M2þ adsorption on biopolymers III: comparative kinetic pattern of Pb,
et al. (2006) Hg and Cd. Carbohydr. Polym. 56, 321–332.
2 Pristine CTS 2.96 0.31 9.6 0.250 Udaybhaskar et al. Ding, P., Huang, K.L., Li, G.Y., Liu, Y.F., Zeng, W.W., 2006. Kinetics of adsorption of
(1990) Zn(II) ion on chitosan derivatives. Int. J. Biol. Macromol. 39, 222–227.
3 CTS bead 0.66 0.10 19.2 0.333 Sankararamakrishnan Ding, P., Huang, K.L., Li, G.Y., Zeng, W.W., 2007. Mechanisms and kinetics of
et al. (2006) chelating reaction between novel chitosan derivatives and Zn(II). J. Hazard.
4 CS-GLA-CTS 12.0 0.068 19.2 0.435 Sankararamakrishnan Mater. 146, 58–64.
et al. (2006) Donia, A.M., Atia, A.A., Elwakeel, K.Z., 2008. Selective separation of Hg(II) using
5 TU-GLA-CTS 4.13 10.9 0.1 0.488 Rojas et al. (2005) magnetic chitosan resin modified with Schiff’s base derived from thiourea and
6 Quaternary 1.15a Spinelli et al. (2004) glutaraldehyde. J. Hazard. Mater. 151, 372–379.
CTS Gotoh, T., Matsushima, K., Kikuchi, K.I., 2004. Preparation of alginate-chitosan
hybrid gel beads and adsorption of divalent metal ions. Chemosphere 55,
a
The maximum amount of adsorption in that literature. 135–140.
806 F.-C. Wu et al. / Journal of Environmental Management 91 (2010) 798–806

Guibal, E., Saucedo, I., Jansson-Charrier, M., Delanghe, B., Le Cloirec, P., 1994. Saiano, F., Ciofalo, M., Cacciola, S.O., Ramirez, S., 2005. Metal adsorption by Pho-
Uranium and vanadium sorption by chitosan and derivatives. Water Sci. Tech- mopsis sp. biomaterial in laboratory experiments and real wastewater treat-
nol. 30 (9), 183–190. ments. Water Res. 39, 2273–2280.
Hall, K.R., Eagleton, L.C., Acrivos, A., Vermeulen, T., 1966. Pore and solid diffusion Sankararamakrishnan, N., Dixit, A., Iyengar, L., Sanghi, R., 2006. Removal of hex-
kinetics in fixed bed adsorption under constant pattern conditions. Ind. Eng. avalent chromium using a novel cross linked xanthated chitosan. Bioresour.
Chem. Fundam. 5, 212–223. Technol. 97, 2377–2382.
Hu, K.J., Hu, J.L., Ho, K.P., Yeung, K.W., 2004. Screening of fungi for chitosan Sankararamakrishnan, N., Sharam, A.K., Sanghi, R., 2007. Novel chitosan derivative
producers and copper adsorption capacity of fungal chitosan and chitosana- for removal of cadmium in the presence of cyanide from electroplating
ceous materials. Carbohydr. Polym. 58, 45–52. wastewater. J. Hazard. Mater. 148, 353–359.
Huang, C.P., Chung, Y.C., Liou, M.R., 1996. Adsorption of Cu(II) and Ni(II) by pellet- Spinelli, V.A., Laranjeira, M.C.M., Favere, V.T., 2004. Preparation and characterization
ized biopolymer. J. Hazard. Mater. 45, 265–277. of quaternary chitosan salt: adsorption equilibrium of chromium(VI) ion. React.
Inoue, K., Yoshizuka, K., Ohto, K., 1999. Adsorptive separation of some metal ions by Funct. Polym. 61, 347–352.
complexing agent types of chemically modified chitosan. Anal. Chim. Acta 388, Steenkamp, G.C., Keizer, K., Neomagusb, H.W.J.P., Krieg, H.M., 2002. Copper(II)
209–218. removal from polluted water with alumina/chitosan composite membranes.
Jeon, C., Holl, W.H., 2003. Chemical modification of chitosan and equilibrium study J. Membr. Sci. 197, 147–156.
for mercury ion removal. Water Res. 37, 4770–4780. Sun, S., Wang, A., 2006. Adsorption kinetics of Cu(II) ions using N, O-carboxymethyl-
Juang, R.S., Ju, C.Y., 1997. Equilibrium sorption of copper(II)-ethylenediaminetetra- chitosan. J. Hazard. Mater. 131, 103–111.
acetic acid chelates onto cross-linked, polyaminated chitosan beads. Ind. Eng. Udaybhaskar, P., Iyengar, L., Rao, A.V.S.P., 1990. Hexavalent chromium interaction
Chem. Res. 36, 5403–5409. with chitosan. J. Appl. Polym. Sci. 39, 739–747.
Juang, R.S., Wu, F.C., Tseng, R.L., 1999. Adsorption removal of copper(II) using chi- Vasconcelos, H.L., Favere, V.T., Goncalves, N.S., Laranjeira, M.C.M., 2007. Chitosan
tosan from simulated rinse solutions containing chelating agents. Water Res. modified with reactive blue 2 dye on adsorption equilibrium of Cu(II) and Ni(II)
33, 2403–2409. ions. React. Funct. Polym. 67, 1052–1060.
Juang, R.S., Wu, F.C., Tseng, R.L., 2001. Solute adsorption and enzyme immobilization Vasconcelos, H.L., Camargo, T.P., Goncalves, N.S., Neves, A., Laranjeira, M.C.M.,
on chitosan beads prepared from shrimp shell wastes. Bioresour. Technol. 80, Favere, V.T., 2008. Chitosan crosslinked with a metal complexing agent:
187–193. synthesis, characterization and copper(II) ions adsorption. React. Funct. Polym.
Justi, K.C., Favere, V.T., Laranjeira, M.C.M., Neves, A., Peralta, R.A., 2005. Kinetics and 68, 572–579.
equilibrium adsorption of Cu(II), Cd(II), and Ni(II) by chitosan functionalized Vieira, R.S., Beppu, M.M., 2006a. Dynamic and static adsorption and desorption
with 2[bis-(pyridylmethyl) aminomethyl]-4-methyl-6-formylphenol. J. Colloid of Hg(II) ions on chitosan membranes and spheres. Water Res. 40,
Interface Sci. 291, 369–374. 1726–1734.
Kawamura, Y., Mitsuhashi, M., Tanibe, H., Yoshida, H., 1993. Adsorption of metal ions Vieira, R.S., Beppu, M.M., 2006b. Interaction of natural and crosslinked chitosan
on polyaminated highly porous chitosan chelating resin. Ind. Eng. Chem. Res. membranes with Hg(II) ions. Colloids Surf. A Physicochem. Eng. Aspects 279,
32, 386–391. 196–207.
Li, N., Bai, R., 2005. Copper adsorption on chitosan–cellulose hydrogel beads: Vijaya, Y., Popuri, S.R., Boddu, V.M., Krishnaiah, A., 2008. Modified chitosan and
behavior and mechanisms. Sep. Purif. Technol. 42, 237–247. calcium alginate biopolymer sorbents for removal of nickel(II) through
Lima, I.S., Airoldi, C., 2004. A thermodynamic investigation on chitosan–divalent adsorption. Carbohydr. Polym. 72, 261–271.
cation interactions. Thermochim. Acta 421, 133–139. Vitali, L., Laranjeira, M.C.M., Goncalves, N.S., Favere, V.T., 2008. Spray-dried chitosan
McKay, G., El-Geundi, M.S., Nassar, M.M., 1987. Equilibrium studies during the microspheres containing 8-hydroxyquinoline-5-sulfonic acid as a new adsor-
removal of dyestuffs from aqueous solutions using bagasse pith. Water Res. 21, bent for Cd(II) and Zn(II) ions. Int. J. Biol. Macromol. 42, 152–157.
1513–1520. Wan Ngah, W.S., Endud, C.S., Mayanar, R., 2002. Removal of copper(II) ions from
Mitani, T., Fukumuro, N., Yoshimoto, C., Ishii, H., 1991. Effect of the counterions aqueous solution onto chitosan and cross-linked chitosan beads. React. Funct.
(SO2 
4 and Cl ) on the adsorption of copper and nickel by swollen chitosan Polym. 50, 181–190.
beads. Agric. Biol. Chem. 55, 2419–2423. Wan Ngah, W.S., Kamari, A., Koay, Y.J., 2004. Equilibrium and kinetics studies of
Monteiro, O.A.C., Airoldi, C., 1999. Some thermodynamic data on copper-chitin and adsorption of copper(II) on chitosan and chitosan/PVA beads. Int. J. Biol. Mac-
copper–chitosan biopolymer interactions. J. Colloid Interface Sci. 212, 212–219. romol. 34, 155–161.
Ng, J.C.Y., Cheung, W.H., McKay, G., 2002. Equilibrium studies of the sorption of Wan, M.W., Petrisor, I.G., Lai, H.T., Kim, D., Yen, T.F., 2004. Copper adsorption
Cu(II) ions onto chitosan. J. Colloid Interface Sci. 255, 64–74. through chitosan immobilized on sand to demonstrate the feasibility for in situ
Ng, J.C.Y., Cheung, W.H., McKay, G., 2003. Equilibrium studies for the sorption of soil decontamination. Carbohydr. Polym. 55, 249–254.
lead from effluents using chitosan. Chemosphere 52, 1021–1030. Wang, Y.H., Lin, S.H., Juang, R.S., 2003. Removal of heavy metals from aqueous
Onsoyen, E., Skaugrud, O., 1990. Metal recovery using chitosan. J. Chem. Technol. solutions using various low-cost adsorbents. J. Hazard. Mater. 102,
Biotechnol. 49, 395–404. 291–302.
Paulino, A.T., Guilherme, M.R., Reis, A.V., Tambourgi, E.B., Nozaki, J., Muniz, E.C., Wu, F.C., Tseng, R.L., Juang, R.S., 1999. Role of pH in metal adsorption from aqueous
2007. Capacity of adsorption of Pb2þ and Ni2þ from aqueous solutions by chi- solutions containing chelating agents onto chitosan. Ind. Eng. Chem. Res. 38,
tosan produced from silkworm chrysalides in different degrees of deacetyla- 270–275.
tion. J. Hazard. Mater. 147, 139–147. Wu, F.C., Tseng, R.L., Juang, R.S., 2000. Comparative adsorption of metal and dye on
Roberts, G.A.F., Domszy, J.G., 1982. Determination of viscometric constants for flake- and bead-types of chitosans prepared from fishery wastes. J. Hazard.
chitosan. Int. J. Biol. Macromol. 4, 374–379. Mater. B73, 63–75.
Rodrigues, C.A., Laranjeira, M.C.M., de Favere, V.T., Stadler, E., 1998. Interaction of Yang, T.C., Zall, R.R., 1984. Adsorption of metals by natural polymers generated from
Cu(II) on N-(2-pyridylmethyl) and N-(4-pyridylmethyl) chitosan. Polymer 39, seafood processing wastes. Ind. Eng. Chem. Prod. Res. Dev. 23, 168–172.
5121–5126. Zhao, F., Yu, B., Yue, Z., Wang, T., Wen, X., Liu, Z., Zhao, C., 2007. Preparation of
Rojas, G., Silva, J., Flores, J.A., Rodriguez, A., Ly, M., Holger, M., 2005. Adsorption of porous chitosan gel beads for copper(II) ion adsorption. J. Hazard. Mater. 147,
chromium onto cross-linked chitosan. Sep. Purif. Technol. 44, 31–36. 67–73.

You might also like