You are on page 1of 9

Treatment of biodiesel wastewater by adsorption with commercial

chitosan akes: Parameter optimization and process kinetics


Wipawan Pitakpoolsil
a
, Mali Hunsom
b, c,
*
a
Department of Environmental Science, Faculty of Science, Chulalongkorn University, 254 Phayathai Rd., Bangkok 10330, Thailand
b
Department of Chemical Technology, Faculty of Science, Chulalongkorn University, 254 Phayathai Rd., Bangkok 10330, Thailand
c
Center of Excellence on Petrochemical and Materials Technology (PETRO-MAT), Chulalongkorn University, Bangkok 10330, Thailand
a r t i c l e i n f o
Article history:
Received 14 June 2013
Received in revised form
4 December 2013
Accepted 9 December 2013
Available online 8 January 2014
Keywords:
Adsorption
Biodiesel wastewater
Kinetics
Chitosan
a b s t r a c t
The possibility of using commercial chitosan akes as an adsorbent for the removal of pollutants from
biodiesel wastewater was evaluated. The effect of varying the adsorption time (0.5e5 h), initial waste-
water pH (2e8), adsorbent dose (0.5e5.5 g/L) and mixing rate (120e350 rpm) on the efciency of
pollutant removal was explored by univariate analysis. Under the derived optimal conditions, greater
than 59.3%, 87.9% and 66.2% of the biological oxygen demand (BOD), chemical oxygen demand (COD) and
oil & grease, respectively, was removed by a single adsorption. Nevertheless, the remaining BOD, COD
and oil & grease were still higher than the acceptable Thai government limits for discharge into the
environment. When the treatment was repeated, a greater than 93.6%, 97.6% and 95.8% removal of the
BOD, COD and oil & grease, respectively, was obtained. The reusability of commercial chitosan following
NaOH washing (0.05e0.2 M) was not suitable, with less than 40% efciency after just one recycling and
declining rapidly thereafter. The adsorption kinetics of all pollutant types by the commercial chitosan
akes was controlled by a mixed process of diffusion and adsorption of the pollutants during the early
treatment period (0e1.5 h) and then solely controlled by adsorption after 2 h.
2013 Elsevier Ltd. All rights reserved.
1. Introduction
Biodiesel is an alternative renewable fuel that has become
widely used as a partial replacement for diesel derived from non-
renewable fossil fuel due not only to its sustainably renewable
nature but also to its high cetane number, lubricity, ash point and
biodegradability, plus its low toxicity. Biodiesel is seen as a much
safer fuel than that derived from fossil fuel because it has a low
emission level of toxic compounds, such as SO
2
, hydrocarbons,
particulates, polycyclic aromatic hydrocarbons and CO (Knothe,
2005; Smith et al., 2009).
Biodiesel is currently produced from the transesterication re-
action between vegetable oils or animal fats and principally
methanol (although ethanol is also used to a lesser extent) in the
presence of an acidic or alkaline catalyst to formthe biodiesel (fatty
acid methyl esters (FAME) or fatty acid ethyl esters). The untreated
biodiesel contains several impurities, such as free glycerol, soap,
metals, methanol (or ethanol), free fatty acids (FFAs), catalyst, water
and glycerides, which impact negatively on the performance and
durability of the diesel engine (Ngamlerdpokin et al., 2011). Thus,
the removal of these components in a purication stage is essential.
The traditional purication method utilized is wet washing, which
involves using water or a weak acid to remove some of the excess
contaminants fromthe biodiesel. However, the addition of water or
a weak acid to the process leads to many disadvantages, including
an increased cost and production time, the generation of a highly
polluting efuent (wastewater) that needs to be treated prior to
environmental discharge (Berrios and Skelton, 2008) and the sig-
nicant loss of biodiesel into the wastewater phase (Canakci and
Gerpen, 2001). For example, more than 6.0 10
6
L/day of bio-
diesel was produced in Thailand in 2010 (DEDE, 2011), with the
formation of at least 1.2 10
6
L/day of biodiesel contaminated
wastewater. This wastewater is at a high pH, due to the signicant
levels of residual alkaline catalyst, and contains a high quantity of
hexane-extracted oil and a high solid content. Together, these
components inhibit the growth of most microorganisms making
this wastewater difcult to biodegrade naturally.
Currently, several processes have been developed to treat bio-
diesel wastewater, including biological (Kato et al., 2005; Nishiro
and Nakashimada, 2007; Siles et al., 2010; Suehara et al., 2005),
* Corresponding author. Department of Chemical Technology, Faculty of Science,
Chulalongkorn University, 254 Phayathai Rd., Bangkok 10330, Thailand. Tel.: 662
2187523 5; fax: 662 2555831.
E-mail address: mali.h@chula.ac.th (M. Hunsom).
Contents lists available at ScienceDirect
Journal of Environmental Management
j ournal homepage: www. el sevi er. com/ l ocat e/ j envman
0301-4797/$ e see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.jenvman.2013.12.019
Journal of Environmental Management 133 (2014) 284e292
chemical and physical processes. Although the biological processes
are the most efcient and economic way for reducing the envi-
ronmental impacts of biodiesel wastewater, they also generate
large amounts of low-density sludge that has low and slow
decomposition efciency. Thus, these approaches are time and
storage volume consuming. Accordingly, the chemical and physical
processes are alternative (but not mutually exclusive) potential
procedures for treating the wastewater from biodiesel production
plants. The electrocoagulation process, using an aluminum anode
and graphite cathode, is potentially suitable as a primary treatment
for FAME biodiesel wastewater because it can reduce the oil &
grease and total suspended solid (TSS) in biodiesel wastewater by
more than 95%, but it can only achieve a 55% reduction in the
chemical oxygen demand (COD) due to the poor removal of the
residual glycerol and methanol (Chavalparit and Ongwandee,
2009). A two-stage management of biodiesel wastewater
comprised of chemical recovery and electro-oxidation was found to
be effective in that it removed all of the COD and oil & grease, and
reduced the biological oxygen demand (BOD) by more than 95%
with pseudo-rst-order rate kinetics (Jaruwat et al., 2010). By using
chemical coagulation with either Al
2
(SO
4
)
3
or poly-aluminum
chloride (PAC) instead of electro-oxidation (Kumjadpai et al.,
2011), a similar removal efciency for BOD, COD and oil & grease
was obtained but over a broad wastewater pH range of 4.5e10.0
and 2.5e7.0, respectively. This process provided a 1.6-fold or 38%
lower operating cost (1.11 USD/m
3
) compared with the electro-
coagulation process (1.78 USD/m
3
). However, the electro-
coagulation process provided a better wastewater quality than the
chemical process with either Al
2
(SO
4
)
3
or PAC, with the exception
of the reduction of the BOD (Ngamlerdpokin et al., 2011). The use of
the dissolved air otation (DAF) process alone or with acidication
of the wastewater could not separate the residual oil & grease from
the biodiesel wastewater, and so the use of DAF with acidication
(with HCl or H
2
SO
4
) and coagulation (with Al
2
SO
4
) was then sug-
gested (Rattanapan et al., 2011).
Among the several chemical and physical methods utilized, the
adsorption process is one of the effective methods. The most
common adsorbents used are peat (Caldern et al., 2008; Ringqvist
et al., 2002; Xiong and Mahmood, 2010), bentonite clay (Koswojo
et al., 2010; Shi et al., 2011; Tahir and Naseem, 2007; Zhu and Ma,
2008), activated carbon (El-Naas et al., 2010; Huang and Su, 2010;
Soleimani and Kaghazchi, 2008) and agricultural waste (Moussavi
and Khosravi, 2010;

Sciban et al., 2007). Nevertheless, these sys-
tems still require physicochemical and microbiological pretreat-
ment to enhance the adsorption (Ahmad et al., 2005). Another
promising candidate adsorbent for wastewater treatment is chito-
san, a partially deacetylated form of the natural biopolymer chitin
(poly(b-(1-4)-2 D-acetamido-2-deoxy-b-D-glucosamine)) that is
found in the exoskeleton of crustaceans and fungal cell walls and is
a waste product from the marine seafood (especially shrimp) in-
dustry (Fereidoon and Jozef, 1991). The potential of chitosan lies in
part in that it has various excellent properties, including biode-
gradability, hydrophilicity, biocompatibility, good adsorption
properties, occulating ability, polyelectrolisity, and its capacity of
regeneration in a number of applications (Feng et al., 2000; Majeti,
2000). It can be used either in the original deacetylated form (with
varying degrees of deacetylation) or in a chemically modied form
depending on the properties of the wastewater (Chiou and Li,
2002). Both unmodied and chemically modied chitosan have
been used to adsorb heavy metals (Kumar et al., 2009; Paulino et al.,
2008), reactive dyes (Chen, 2008; Kyzas and Lazaridis, 2009;
Momenzadeh et al., 2011), organic compounds (Wan Ngah et al.,
2008; Wang et al., 2009) and oil residues from wastewater
(Ahmad et al., 2005; Piyamongkala et al., 2008; Sokker et al., 2011).
However, no work in the available literature has reported on the
use of either unmodied or modied chitosan to adsorb pollutants
from biodiesel wastewater. In this work, a series of adsorption ex-
periments were performed to evaluate the feasibility of using un-
modied commercial chitosan akes as an adsorbent for removing
pollutants from biodiesel wastewater. The effects of varying four
key operating parameters (adsorption time, initial wastewater pH,
dosage of adsorbent and the mixing rate) were investigated. The
reusability of the commercial chitosan akes following washing in
alkali, and the adsorption kinetics were also examined.
2. Materials and methods
The treatment of wastewater from a conventional biodiesel
production plant of our country, that uses waste used-vegetable oil
as the feedstock to form biodiesel with an alkali catalyst, was car-
ried out on a laboratory scale at ambient temperature. The original
wastewater with the properties demonstrated in Table 1 was rst
pre-treated by the addition of H
2
SO
4
(98.08% QReC Grade AR) to
reduce the pH to w2.0, as previously reported (Ngamlerdpokin
et al., 2011). The mixture was then shaken for 2 h before being
left for 30 min to allow the complete phase separation between the
upper FAME-rich phase and the lower acidic aqueous phase. Both
phases were then separated by slow decantation. The remaining
discharged wastewater (aqueous phase) after the extraction of
upper FAME-rich phase was then treated by the adsorption process
with the commercial chitosan akes (Taming Enterprises,
Thailand).
Prior to the adsorption stage, the moisture content in the
commercial chitosan akes was eliminated by drying at 105

C for
30 min. Meanwhile, the pH of the extracted aqueous wastewater
phase was adjusted to be within the preferred range (pH 2e8) by
the addition of NaOH (1 M, Earth Cheme Lab., Thailand). Subse-
quently, approximately 100 mL of wastewater was conventionally
treated by the addition of commercial chitosan akes (range 0.5e
5.5 g/L) as the adsorbent in a conical ask with a constant stirring
rate (range of 120e350 rpm) for a selected adsorption time (0.5e
5 h). Finally, the treated wastewater was separated from the
adsorbent by vacuumltration through lter paper No.1 (Whatman
70 mm # 1001070). The concentration of pollutants in the treated
wastewater was measured in terms of the BOD, COD, oil & grease,
total dissolved solids (TDS) and TSS according to standard methods
(APHA, 1998). In addition, to characterize the types of compounds
in the wastewater, gas chromatographyemass spectroscopy (GCe
MS) analysis was performed (6890N, Agilent of GC/Pegosees III,
Lego of MS).
To test the reusability of spend chitosan, 0.35 g of used-chitosan
akes were dipped in 100 mL NaOH at three different concentra-
tions (0.05, 0.10 and 0.2 M) and then shaken at a constant rate of
130 rpmfor 3 h. Afterward, the akes were separated fromthe basic
solution by ltration and rinsed several times with distilled water
to remove the excess base. The ready-to-use regenerated chitosan
was obtained after drying in an oven at 70

C for 24 h. Finally, it was


subjected to treat wastewater from biodiesel production plant at
optimum adsorption condition.
3. Results and discussion
3.1. Properties of the original biodiesel wastewater and pretreated-
wastewater
The original wastewater obtained from the waste used-oil bio-
diesel production plant had an opaque white color and a high pH
(range of 9.25e10.26). It contained very high BOD, COD, oil &
grease, TDS and TSS in comparison to the government standard
(Table 1). From the GCeMS analysis, the main components in the
W. Pitakpoolsil, M. Hunsom / Journal of Environmental Management 133 (2014) 284e292 285
contaminated biodiesel wastewater were glycerin, a by-product
from biodiesel production, un-reacted or residual FFAs, such as
palmitic acid, lauric acid and myristic acid, and unsaturated
aliphatic FFA compounds, such as oleic acid and their derivatives
(Fig.S.1(a)). After the pre-treatment stage (acid protonation),
various types of FFAs and FAMEs were removed or converted,
leaving only 2o-alcohol (3-methyl-2-hexanol) and saturated
dicarboxylic acids (pimelic acid, azelaic acid and suberic acid) as
contaminants together with glycerin (Fig.S.1(b)). This is because the
H

from the H
2
SO
4
quickly substitutes for the Na atom in the soap
molecule, forming the less polar and water insoluble FFA. In addi-
tion, the protons can also substitute the H
2
O molecule combining
biodiesel leading to the formation of the free FAMEs (Jaruwat et al.,
2010). The reduction in the level of these contaminating com-
pounds resulted in a signicant decrease in the BOD (1.24- to 1.27-
fold), COD (1.39- to 1.74-fold) and oil & grease (w1.5-fold) in the
wastewater, although they were still far in excess (22- to 30-fold,
53- to 78-fold, and 176- to 226-fold, respectively) of the Thai
standard for discharge to the environment (Table 1). In addition,
although the TDS and TSS were decreased 1.2- and 3.0-fold,
respectively, they were both still some 1.5-fold above the accept-
able Thai standard for environmental discharge (Table 1).
3.2. Effect of the four key adsorption parameters
3.2.1. Effect of the adsorption time
The effect of varying the adsorption time on the quantity of
pollutant removal, measured in terms of the BOD, COD and oil &
grease removal, was rst explored in the range of 0.5e5 h using
chitosan at 3.5 g/L, a mixing rate of 350 rpm and an initial waste-
water pH of 2. The removal efciency of all three monitored
pollutant types increased gradually over the rst hour (COD) to
1.5 h and (BOD and oil & grease) after the start-up time and was
then slightly faster up to 2 h where after it slowed and reached or
almost reached a plateau within 3 h (Fig. 1). The increase in the
pollutant removal efciency as a function of time during the early
adsorption period (0e1.5 h) might be due to the small chance of
interaction between oil molecules and chitosan particles over a
short adsorption period. As the adsorption time was increased, the
breaking up of the oil droplets into smaller ones was enhanced,
resulting in a greater adsorption of the oil droplets due to the
greater interfacial contact area between the oil molecules and
chitosan particles (Michael et al., 1994). Further increasing the
adsorption time over 3 h did not achieve a higher removal ef-
ciency of each pollutant category since the chitosan had already
reached its saturation point. The maximumobtained BOD, COD and
oil & grease removal was 37.3%, 50.2% and 46.4%, respectively. To
explain the role of pollutant removal by chitosan akes, it is
necessary to take into account the functional groups in the chitosan
structure as well as the residual pollutant contaminates found in
the biodiesel wastewater. As demonstrated in the GCeMS spectra,
the pollutants in the biodiesel wastewater were largely comprised
of FAMEs and FFAs with an anionic charge (Philipson and Ward,
1985). Thus, the positively charged amine functional group of chi-
tosan (at this pH) can bind and bridge with these anionic charges
(Osman and Arof, 2003) and consequently destabilize the emulsion
by charge neutralization (Jill et al., 1999).
SEM analysis of the chitosan akes before and after the
adsorption of pollutants from biodiesel wastewater revealed a
signicant difference in their surface morphologies. The fresh chi-
tosan akes showed an irregular and rough surface with small
pores distributed randomly in the structure (Fig. 2(a)), but after the
pollutant adsorption, a muddy-like compound was seen adhered to
the surface of the chitosan and covered the bumpy surface layer
and pores (Fig. 2(b)). Thus, the wastewater pollutants were likely to
have been adsorbed onto the pores and the surface of the chitosan
akes and to develop an oily layer. This is consistent with the
phenomena observed by Ahmad et al. (2005) on the adsorption of
pollutants from palm oil mill efuent by chitosan akes and
powder.
3.2.2. Effect of the initial wastewater pH
The pH of the wastewater solution affects not only the surface
chemistry of the adsorbent but also the chemistry of the pollutants
in the biodiesel wastewater. The effect of the initial wastewater pH
on the pollutant removal was evaluated over a pH range of 2e8
Table 1
Chemical and physical properties of the raw biodiesel wastewater and that after different treatments.
Characteristics Thai standard Fresh wastewater Pre-treated wastewater
a
Properties of treated wastewater
Single
b
Repetitive
c
pH 5.5e9 9.25e10.26 2.00e2.40 4.2e5.44 5.6e6.65
BOD (mg/L) 60 1492e2286 1312e1802 408e652 40e50
COD (mg/L) 400 29,595e54,362 21,242e31,304 1714e3822 382e384
Oil & grease (mg/L) 5 1040e1710 680e1130 160e200
40e60
TDS (mg/L) 3000 5340e5400 4530e4560 830e1130 290e370
TSS (mg/L) 150 670e690 220e230 50e90 20e40
a
Pretreatment by protonation with H
2
SO
4
at pH 2 and phase separation.
b
Pretreated wastewater (after protonation and phase separation) reset to pH 4 and adsorbed with commercial chitosan at 3.5 g/L with mixing at 300 rpm for 3 h.
c
Repetitive adsorption treatment (four times) of the pretreated wastewater (pH reset to 4.0) with the commercial chitosan at 3.5 g/L with mixing at 300 rpm for 3 h each
time.
0
20
40
60
80
100
0 1 2 3 4 5
P
o
l
l
u
t
a
n
t

r
e
m
o
v
a
l

(
%
)
Time (h)
BOD
COD
Oil & grease
Fig. 1. Effect of the adsorption time on BOD, COD and oil & grease removal by com-
mercial chitosan akes at 3.5 g/L, a mixing rate of 350 rpm and pretreated wastewater
with an initial pH of 2.
W. Pitakpoolsil, M. Hunsom / Journal of Environmental Management 133 (2014) 284e292 286
using commercial chitosan at 3.5 g/L, a mixing rate of 350 rpm and
an adsorption time of 3 h (taken as optimal, Section 3.2.1). The
interaction between the oil molecules and the adsorption site of
chitosan (NH
2
group) was expected to be increased under acidic
conditions (Sokker et al., 2011) because the adsorbent surface is
completely covered with hydronium ions (Laus et al., 2010) that
would promote a high removal efciency of the anionic pollutants.
However, this was not found to be the case here. Rather, the
removal efciency of COD, BOD and oil & grease were relatively low
at pH 2 (38.0%, 51.1% and 48.2% for BOD, COD and oil & grease,
respectively) and increased as the initial wastewater pH was
increased up to a maximum removal efciency at pH 4.0 (48.2%,
81.8% and 74.5% for BOD, COD and oil & grease, respectively). They
then decreased at pH values above 4.0 to a minimal level (over the
tested range of pH 2e8) at pH 8 (Fig. 3). Consequentially, a pre-
treated wastewater pH of 4.0 was taken to be optimal.
The low pollutant removal efciency at pH 2.0 might be
attributed to the swelling of the chitosan akes that caused themto
be very brittle and consequently reduce their adsorption capacity
(Wan Ngah and Fatinathan, 2010). As the pH increased up to pH 4.0
the pollutant removal efciency increased because of the decreased
chitosan swelling whilst the amine groups remained positively
charged and enhancing the adsorption. At pH values greater than
4.0 the removal efciency of all pollutants then decreased due to
the deprotonation of the amine groups of chitosan, leading to a
decrease in the positive charge density and a weakened electro-
static interaction between chitosan and the anionic pollutant
molecules (Nawi et al., 2010).
3.2.3. Effect of the chitosan dosage
The effect of the chitosan dosage on the removal efciency of
BOD, COD and oil & grease was evaluated with a chitosan ake dose
from0.5 to 5.5 g/L at pH 4.0 with a constant mixing rate of 350 rpm
for 3 h. The removal efciency of each pollutant type increased as
the chitosan ake dose increased from 0.5 to 3.5 g/L (Fig. 4), rising
from 12.9% to 52.9% for BOD, 29.5%e81.3% for COD, and 23.9%e
76.0% for oil & grease. This presumably simply reects the
increasing number of adsorption sites and so opportunities for
interaction of the pollutant type in the biodiesel wastewater with
chitosan. However, the removal efciency of all investigated pol-
lutants remained essentially the same as the dosage of chitosanwas
increased above 3.5 g/L to 5.5 g/L. This might be attributed to the
limitation of transport of pollutant particles to the chitosan surface
and the low adsorptive capacity utilization of the adsorbent (Jeon
and Park, 2005).
3.2.4. Effect of the mixing rate
Theoretically the mixing rate can inuence the adsorption
through improving the contact between the chitosan surface and
the pollutants in the wastewater and reducing the mass-transfer
limitation of pollutants adhering onto the surface of the chitosan.
In this study the effect of the mixing rate was investigated by
varying the agitation rate between 120 and 350 rpm, using an
initial wastewater pH of 4, a chitosan dosage of 3.5 g/L and an
Fig. 2. Representative SEM images (5000 magnication) of commercial chitosan akes (a) before, and (b) after use, and after regeneration in the presence of 0.1 M NaOH for the (c)
third and (d) fth times.
0
20
40
60
80
100
0 2 4 6 8 10
P
o
l
l
u
t
a
n
t

r
e
m
o
v
a
l

(
%
)
Initial wastewater pH
BOD
COD
Oil & grease
Fig. 3. Effect of the pretreated wastewater pH on the removal efciency of BOD, COD
and oil & grease by commercial chitosan akes at 3.5 g/L, a mixing rate of 350 rpm and
an adsorption time of 3 h.
W. Pitakpoolsil, M. Hunsom / Journal of Environmental Management 133 (2014) 284e292 287
adsorption time of 3 h. The removal efciency of the BOD, COD and
oil & grease increased sharply and almost linearly from17.4%, 38.8%
and 22.3%, respectively, to approximately 59.3%, 87.9% and 66.2%,
respectively, when the mixing rate was increased from 120 to
250 rpm, and then it increased slowly and attained or almost
attained a plateau value at 300 rpm and higher (Fig. 5). The
adsorption of all three pollutant types was, therefore, likely to be
limited by mass transport at low mixing rates. However, that
increasing the mixing rate above 300 rpm did not promote a
signicantly higher removal efciency of each pollutant type in-
dicates that their reduction at higher mixing rates was limited by
the kinetics of the pollutant adsorption by chitosan particles. The
rapid adsorption of pollutants onto chitosan in the early stages,
followed by a relatively slow rate at later stages has been reported
previously for the adsorption of metal ions by various types of
chitosan (Duzl et al., 2001; Jeon and Park, 2005; Wan Ngah et al.,
2002). Thus, the optimized conditions were set as an initial
wastewater pH of 4.0, a chitosan ake dose of 3.5 g/L and adsorbed
for 3 h with agitation at 300 rpm.
The properties of the obtained wastewater after a single treat-
ment with chitosan under the partially optimized conditions are
summarized in Table 1. The pH of the treated wastewater after the
adsorption with the unmodied commercial chitosan akes was a
bit acidic compared with the Thai government standard while the
TDS and TSS were reduced to within the Thai standard. Although
the BOD, COD and oil & grease were reduced signicantly, they
were still signicantly higher than the acceptable value (approxi-
mately 6.8- to 10.9- fold, 4.3- to 9.6-fold, and 32- to 40-fold,
respectively). Thus, further sequential cycles of adsorption were
performed as above except using fresh chitosan. The removal ef-
ciency of BOD, COD and oil & grease increased as the number of
treatment cycles increased up to four, and then reached a plateau or
only slightly increased with further repeat treatments (Fig. 6). The
GCeMS analysis revealed that most types of contaminants were
removed from the biodiesel wastewater except for glycerin
(Fig.S.1(c)). Only 40e50, 382e384 and 40e60 mg/L of BOD, COD
and oil & grease remained in the wastewater after four consecutive
treatments (Table 1). They were within the acceptable range (Thai
standard) for BOD and COD for discharge into the environment,
whilst the TDS and TSS were dramatically reduced and well within
the Thai standards. However, the oil & grease were still too high (8-
to 12-fold) and so would require additional treatment prior to
discharge.
3.3. Reusability of the commercial chitosan akes
To test the reusability of the commercial chitosan akes, they
were regenerated by washing in NaOH of various concentrations
(0.05, 0.10 and 0.20 M) for 3 h in order to remove the adsorbed
pollutants from the chitosan. The treated chitosan akes were then
re-used to adsorb pollutants from fresh pretreated biodiesel
wastewater at the partially optimized conditions (chitosan dosage
of 3.5 g/L, adsorption time of 3 h, an initial wastewater pHof 4 and a
mixing rate of 300 rpm). Treatment with NaOH regenerated the
adsorptive ability of the spent chitosan in an alkali concentration-
dependent manner, with the highest tested concentration of
NaOH (0.2 M) giving the highest subsequent level of pollutant
removal by the regenerated chitosan (Fig. 7). This is because higher
alkali concentrations result in greater repulsive electrostatic in-
teractions between the chitosan and pollutants (Wan Ngah et al.,
2008), and so a higher desorption from the chitosan. However,
the apparent adsorption capacity of the regenerated chitosan
decreased with increasing regeneration cycles, which is probably
0
20
40
60
80
100
0 1 2 3 4 5 6
P
o
l
l
u
t
a
n
t

r
e
m
o
v
a
l

(
%
)
Chitosan dosage (g/L)
BOD
COD
Oil & grease
Fig. 4. Effect of chitosan ake dosage on the removal efciency of BOD, COD and oil &
grease from pretreated wastewater, at an initial pH of 4.0, a mixing rate of 350 rpm and
an adsorption time of 3 h.
0
20
40
60
80
100
100 150 200 250 300 350 400
P
o
l
l
u
t
a
n
t

r
e
m
o
v
a
l

(
%
)
Mixing rate (rpm)
BOD
COD
Oil & grease
Fig. 5. Effect of the mixing rate on the removal efciency of BOD, COD and oil & grease
in the presence of chitosan akes at 3.5 g/L, pretreated wastewater with an initial pH of
4.0 and an adsorption time of 3 h.
0
20
40
60
80
100
0 1 2 3 4 5 6 7
P
o
l
l
u
t
a
n
t

r
e
m
o
v
a
l

(
%
)
Repetitive adsorption with fresh chitosan (times)
BOD
COD
Oil & grease
Fig. 6. The removal efciency of BOD, COD and oil & grease from pretreated waste-
water (initial pH of 4.0) as a function of the number of repetitive cycles of adsorption
with fresh chitosan at 3.5 g/L, mixing at 300 rpm and a 3 h adsorption cycle.
W. Pitakpoolsil, M. Hunsom / Journal of Environmental Management 133 (2014) 284e292 288
because the NaOH treatment could not remove all of the adsorbed
pollutants. Indeed, a signicant accumulation of a muddy-like
compound on the surface of chitosan was observed to increase
after each reuse, resulting in the marked decrease in the number of
available adsorption sites on the chitosan (Fig. 2(c) and (d)).
According to the obtained results, although NaOHcan be used to
regenerate the commercial chitosan akes, the reusability of the
regenerative chitosan was not good. The removal efciency of BOD,
COD and oil & grease decreased by almost 40% after just the rst
regeneration cycle and decreased further with each subsequent
regeneration cycle. Thus, the chemical modication of commercial
chitosan (e.g. by graft polymerization of chitosan with a polymer,
such as polyacrylamide) or other types of desorbing agents should
be carried out in order to obtain a better adsorption capacity and
reusability.
3.4. Adsorption kinetics
The adsorption of pollutants onto the commercial chitosan
akes consists of two processes; (a) the transport of pollutants from
the bulk solution to the surface of the chitosan and (b) the
adsorption of the pollutants onto the chitosan surface. In the
former stage, the surface of the chitosan is relatively free of pol-
lutants and the pollutant molecules that arrive at the surface may
attach instantly to the protonated amino groups NH

3
of the
chitosan. Hence, the adsorption rate may be dominated by the
diffusion of pollutant molecules from the bulk solution to the chi-
tosan surface.
Based on the Fickian diffusion law, the amount of adsorption by
diffusion-controlled dynamics as a function of time can be given by
Eq. (1) (Zhang and Bai, 2003);
q
t
2C
0
S

Dt=p
p
(1)
For analytical convenience, Eq. (1) can be rewritten in terms of
the intraparticle diffusion model, as shown in Eq. (2);
q
t
k
i
t
0:5
(2)
Under a diffusion-controlled transport mechanism, a plot of q
t
versus t
0.5
would show a linear relationship.
A linear relationship of q
t
against t
0.5
was observed for all
investigated pollutants at t
0.5
<1.4 (or t < 2 h), with a coefcient of
determination (R
2
) greater than 0.93 in each case (Fig. 8). This
supports the existence and the importance of diffusion-controlled
transport mechanisms in pollutant adsorption during the early
adsorption period. The rate constant of intraparticle diffusion,
determined from the slope of the plot between q
t
against t
0.5
, was
maximally observed in case of COD (0.4805 g/g-min
0.5
), which was
greater than those for BOD and oil & grease removal by approxi-
mately 16.3- and 36.1-fold, respectively. When t
0.5
>1.4 (or t >2 h),
the adsorption rate was almost constant and so the adsorption of
pollutants onto the chitosan surface may then be controlled by the
adsorption system.
To support this hypothesis, two chemical kinetic models were
used to determine the respective adsorption rate of BOD, COD and
oil & grease onto the commercial chitosan akes. The rst kinetic
model evaluated was the pseudo-rst-order Lagergrens equation 0
20
40
60
80
100
Fresh
chitosan
1 2 3 4 5
O
i
l

&

g
r
e
a
s
e

r
e
m
o
v
a
l

(
%
)
Number of chitosan regeneration
0.05 M
0.10 M
0.20 M
(c)
0
20
40
60
80
100
Fresh
chitosan
1 2 3 4 5
C
O
D

r
e
m
o
v
a
l

(
%
)
Number of chitosan regeneration
0.05 M
0.10 M
0.20 M
(b)
0
20
40
60
80
100
Fresh
chitosan
1 2 3 4 5
B
O
D

r
e
m
o
v
a
l

(
%
)
Number of chitosan regeneration
0.05 M
0.10 M
0.20 M
(a)
Fig. 7. Removal efciency of (a) BOD, (b) COD and (c) oil & grease from pretreated
wastewater (initial pH of 4.0) by adsorption with fresh chitosan and the regenerated
chitosan at 3.5 g/L, mixing at 300 rpm and a 3 h adsorption cycle.
0.00
0.05
0.10
0.15
0.20
0.25
0.30
0.5 1.0 1.5 2.0 2.5
t
0.5
q

(
g
/
g
)
0.0
1.0
2.0
3.0
4.0
5.0
6.0
q

(
g
/
g
)
BOD
Oil & grease
COD
R
2
= 0.9373
R
2
= 0.9588
R
2
= 0.9894
Fig. 8. Adsorption dynamics of the indicated pollutant categories onto commercial
chitosan akes at the partially optimized conditions (3.5 g chitosan/L pretreated
wastewater set to pH 4.0 and mixing at 300 rpm), showing the adsorbed amount of
pollutant (q
t
) as a function of t
0.5
.
W. Pitakpoolsil, M. Hunsom / Journal of Environmental Management 133 (2014) 284e292 289
(Annadurai and Krishnan, 1997; McKay and Ho, 1999), shown in
Eq. (3);
dq
t
dt
k
1
q
e
q
t
(3)
Integration of Eq. (3) with the initial condition of q
t
0 at t 0
provides Eq. (4);
logq
e
q
t
log q
e
k
1
t=2:303 (4)
Thus, a plot of log(q
e
q
t
) versus t provides the kinetic pa-
rameters k
1
and log q
e
from the slope and intercept, respectively.
Actually, as demonstrated in Eq. (4), the equilibrium adsorption
density q
e
is required to t the data, but in many cases q
e
remains
unknown due to the slow adsorption processes (Chiou and Li,
2002). This model is effective at describing the initial adsorption
reaction but it does not t well over the whole contact time, but
rather is applicable only over the initial stage of adsorption (Ahmad
et al., 2005).
The second kinetics model that was evaluated was the pseudo-
second-order equation (McKay and Ho, 1999), which is based on
the adsorption equilibriumcapacity and is expressed in terms of Eq.
(5);
dq
t
dt
k
2
q
e
q
t

2
(5)
Integrating Eq. (5) and applying the initial condition yields Eq.
(6),
1
q
e
q
t

1
q
e
k
2
t (6)
Further, rearrangement of Eq. (6) yields Eq. (7)
t
q
t

q
2
e
k
2

t
q
e
(7)
If second-order kinetics is applicable, a plot of t/q
t
versus t
should show a linear relationship and the rate constant of the
second-order adsorption k
2
and q
e
can be calculated from the slope
and intercept of this plot, respectively. This model is more likely to
predict the behavior over the whole range of adsorption and is in
agreement with chemical sorption being the rate-controlling step
(Chiou and Li, 2002), such as via valency forces through sharing or
exchange of electrons between pollutant anions and the adsorbent.
Fig. 9(a) shows the plot of log (q
e
q
t
) versus time for the
experimental data obtained under the optimized conditions (chi-
tosan at 3.5 g/L for 3 h with mixing at 300 rpm and an initial pre-
treatment wastewater pH of 4). It can be seen that a linear
relationship of log (q
e
q
t
) against time was observed over the
initial 1.5 h of adsorption, and then deviates strongly from linearity
from 2 h onwards. Thus, the attachment of pollutants to the chi-
tosan surface controlled the initial adsorption of pollutants. How-
ever, when the adsorption time was 2 h and greater the
experimental data tted very well with the pseudo-second-order
kinetics reaction (Fig. 9(b)). The derived adsorption rate constants
and equilibrium density values (D) obtained fromboth models, and
the coefcient of determination (R
2
) between the observed and
theoretical (model predicted) data are tabulated in Table 2. The rate
constants of both models, k
1
and k
2
, support the notion that the rate
of COD adsorption onto commercial chitosan akes for the whole
experiment was higher than that of BOD and oil & grease by
approximately 1.50- and 1.1 10
5
-fold for BOD and 1.57- and
1.1 10
5
-fold for oil & grease, respectively.
4. Conclusions
The treatment of biodiesel wastewater was carried out in lab-
oratory scale at ambient condition (30

C). It was found that greater


than 59.3%, 87.9% and 66.2% of BOD, COD and oil & grease was
respectively removed within 3 h at an initial wastewater pH of 4.0
in the presence of 3.5 g chitosan/L at a mixing rate of 300 rpm by a
single adsorption. The repetitive adsorption for four times can
enhance the removal of BOD, COD and oil & grease up to 93.6%,
97.6% and 95.8%, respectively. Only 40e50, 382e384 and 40e
-4.0
-3.0
-2.0
-1.0
0.0
1.0
0 1 2 3 4
l
o
g

(
q
e
-
q
t
)
Time (h)
BOD
COD
Oil & grease
(a)
R
2
= 0.9900
R
2
= 0.9834
R
2
= 0.9831
0.0
0.2
0.4
0.6
0.8
1.0
1.2
0
5
10
15
20
25
30
35
0 1 2 3 4 5 6
t
/
q
t
t
/
q
t
Time (h)
BOD
Oil & grease
COD
(b)
R
2
= 0.9857
R
2
= 0.9986
R
2
= 0.9979
Fig. 9. (a) Pseudo-rst-order reaction, and (b) pseudo-second-order reaction of the
adsorption of the indicated pollutant categories by commercial chitosan akes at the
optimized conditions (3.5 g chitosan/L pretreated wastewater set to pH 4.0 and mixing
at 300 rpm).
Table 2
Adsorption rate constants and equilibrium adsorption density estimated based on the rst- and second-order adsorption kinetics models.
Characteristics Pseudo-rst-order kinetics (during 0e1.5 h) Pseudo-second-order kinetics (during 2e5 h)
k
1
(h
1
) q
e,cal
(g/g) q
e,exp
(g/g) R
2
D
a
(%) k
2
(g/g-h) q
e,cal
(g/g) q
e,exp
(g/g) R
2
D (%)
BOD 0.5860 0.2480 0.2380 0.9834 4.2 0.044 0.2619 0.2380 0.9979 10.0
COD 0.8805 4.9203 4.9057 0.9900 0.3 467 5.2785 4.9057 0.9986 7.6
Oil & grease 0.5624 0.1485 0.1547 0.9831 4.0 0.008 0.1817 0.1547 0.9857 17.5
a
Absolute percentage deviation of the adsorption density at equilibrium solute concentration, as dened by: (q
e,exp
q
e,cal
)/q
e,exp
100.
W. Pitakpoolsil, M. Hunsom / Journal of Environmental Management 133 (2014) 284e292 290
60 mg/L of BOD, COD and oil & grease remained in the wastewater
after four repetitive treatments. The NaOH could partially regen-
erate the used commercial chitosan, but the reusability of such
regenerative chitosan was not good being at only 40% of the
adsorptive capacity after one use and deteriorating further with
each regeneration cycle. The adsorption of BOD, COD and oil &
grease was controlled by a mixed process of pollutant diffusion and
adsorption of pollutant according to a pseudo-rst order reaction
rate during the adsorption period (0e1.5 h) and thenwas controlled
by the adsorption process related to the pseudo-second order re-
action after 2 h adsorption time.
Acknowledgments
The authors would like to thank The Bangchak Petroleum Public
Co. Ltd. for material support and the Center for Petroleum, Petro-
chemicals and Advanced Materials for facility support. Also, we
thank the Publication Counseling Unit (PCU) of the Faculty of Sci-
ence, Chulalongkorn University and Dr. Robert D.J. Butcher for
comments, suggestions and checking the grammar.
Nomenclature
C
0
initial concentration of the pollutants in the bulk solution
in g/L
D diffusion coefcient in m
2
/h
k
1
rate constant of the pseudo-rst-order adsorption inper h
k
2
rate constant of the pseudo-second-order adsorption in g/
g-h
k
i
intraparticle diffusion rate constant in g/g-h
0.5
q
e
adsorption density at equilibrium in g/g
q
t
amount of pollutant adsorbed per unit weight of granules
or adsorption density at time t in g/g
S specic surface area of the chitosan measured in terms of
the BET surface area in m
2
/g
t adsorption time in h
Appendix A. Supplementary data
Supplementary data related to this article can be found at http://
dx.doi.org/10.1016/j.jenvman.2013.12.019.
References
Ahmad, A.L., Sumathi, S., Hameed, B.H., 2005. Adsorption of residue oil from palm
oil mill efuent using powder and chitosan ake: equilibrium and kinetic
studies. Water Res. 39, 2483e2494.
Annadurai, G., Krishnan, M.R.V., 1997. Adsorption of acid dye from aqueous solution
by chitin: batch kinetic studies. Ind. J. Chem. Technol. 4, 213e222.
APHA, 1998. American Water Works Association; Water Environment Federation.
Standards Methods for the Examination of Water and Wastewater, twentieth
ed. American Public Health Association, Washington, D.C.
Berrios, M., Skelton, R.L., 2008. Comparison of purication methods of biodiesel.
Chem. Eng. J. 144, 459e465.
Caldern, M., Moraga, C., Leal, J., Agouborde, L., Navia, R., Vidal, G., 2008. The use of
Magallanic peat as non-conventional sorbent for EDTA removal from waste-
water. Biores. Technol. 99, 8130e8136.
Canakci, M., Gerpen, J.V., 2001. Biodiesel production from oils and fats with high
free fatty acids. Trans. ASAE 44, 1429e1436.
Chavalparit, O., Ongwandee, M., 2009. Optimizing electrocoagulation process for
the treatment of biodiesel wastewater using response surface methodology.
J. Environ. Sci. 21, 1491e1496.
Chen, X., 2008. Study on treatment of printing and dyeing wastewater with chi-
tosan wrapping y-ash. J. Biotechnol. 136, S651.
Chiou, M.S., Li, H.Y., 2002. Equilibrium and kinetic modeling of adsorption of
reactive dye on cross-linked chitosan beads. J. Hazard. Mater. 93, 233e248.
DEDE. 2011. Ministry of Energy. http://www.dede.go.th/dede/images/stories/
Biodiesel/B100_Producer_Jan_11.pdf. (accessed 14.11.12.).
Duzl, E.M.S., Saucedo, M.T.I., Navarro, M.R., Avila, R.M., Guibal, E., 2001. Cadmium
sorption on chitosan sorbents: kinetic and equilibrium studies. Hydrometal-
lurgy 61, 157e167.
El-Naas, M.H., Al-Zuhair, S., Alhaija, M.A., 2010. Removal of phenol from petroleum
renery wastewater through adsorption on date-pit activated carbon. Chem.
Eng. J. 162, 997e1005.
Feng, C.W., Ru, L.T., Ruey, S.J., 2000. Comparative adsorption of metal and dye on
ake and bead-types of chitosans prepared from shery wastes. J. Hazard
Mater. B. 73, 63e75.
Fereidoon, S., Jozef, S., 1991. Isolation and characterization of nutrients and value-
added products from snow crab (Chionoecetes opilio) and shrimp (Pandalus
borealis) processing discards. J. Agri. Food Chem. 39 (8), 1527e1532.
Huang, C.C., Su, Y.J., 2010. Removal of copper ions from wastewater by adsorption/
electrosorption on modied activated carbon cloths. J. Hazard. Mater. 175 (1e3),
477e483.
Jaruwat, P., Kongjao, S., Hunsom, M., 2010. Management of biodiesel wastewater by
the combined processes of chemical recovery and electrochemical treatment.
Ener. Conver. Manage. 51, 531e537.
Jeon, C., Park, K.H., 2005. Adsorption and desorption characteristics of mercury (II)
ions using aminated chitosan bead. Water Res. 39, 3938e3944.
Jill, R.P., Chihpin, H.S.C., Ying, C.C., 1999. Evaluation of a modied chitosan
biopolymer for coagulation of colloidal particles. Colloid. Surf. 147, 359e364.
Kato, S., Yoshimura, H., Hirose, K., Amornkitbamurung, M., Sakka, M., Sugahara, I.,
2005. Application of Microbial Consortium System to Wastewater from Bio-
diesel Fuel Generator. IEA-Waterqual, Singapore. CD-ROM.
Knothe, G., 2005. Dependence of biodiesel fuel properties on the structure of fatty
acid alkyl esters. Fuel Process. Technol. 86, 1059e1070.
Koswojo, R., Utomo, R.P., Ju, Y.H., Ayucitra, A., Soetaredjo, F.E., Sunarso, J., Ismadji, S.,
2010. Acid Green 25 removal from wastewater by organo-bentonite from
Pacitan. Appl. Clay Sci. 48, 81e86.
Kumar, M., Tripathi, B.P., Shahi, V.K., 2009. Crosslinked chitosan/polyvinyl alcohol
blend beads for removal and recovery of Cd(II) from wastewater. J. Hazard.
Mater. 172, 1041e1048.
Kumjadpai, S., Ngamlerdpokin, K., Chatanon, P., Lertsathitphongs, P., Hunsom, M.,
2011. Management of fatty acid methyl ester (FAME) wastewater by a combined
two stage chemical recovery and coagulation process. Can. J. Chem. Eng. 89,
369e376.
Kyzas, G.Z., Lazaridis, N.K., 2009. Reactive and basic dyes removal by sorption onto
chitosan derivatives. J. Colloid Interf. Sci. 331, 32e39.
Laus, R., Costa, T.G., Szpoganicz, B., Fvere, V.T., 2010. Adsorption and desorption of
Cu(II), Cd(II) and Pb(II) ions using chitosan crosslinked with epichlorohydrin-
triphosphate as the adsorbent. J. Hazard. Mater. 183, 233e241.
Majeti, N.V.R.K., 2000. A review of chitin and chitosan applications. Reactive Funct.
Polym. 46, 1e27.
McKay, G., Ho, Y.S., 1999. The sorption of lead (II) on peat. Water Res. 33, 578e584.
Michael, S., Heike, K., Helmar, S., 1994. Adsorption kinetics of emulsiers at oile
water interfaces and their effect on mechanical emulsication. Chem. Eng.
Process 33, 307e311.
Momenzadeh, H., Tehrani-Bagha, A.R., Khosravi, A., Gharanjig, K., Holmberg, K.,
2011. Reactive dye removal from wastewater using a chitosan nanodispersion.
Desalination 271, 225e230.
Moussavi, G., Khosravi, R., 2010. Removal of cyanide from wastewater by adsorption
onto pistachio hull wastes: parametric experiments, kinetics and equilibrium
analysis. J. Hazard. Mater. 183, 724e730.
Nawi, M.A., Sabar, S., Jawad, A.H., Sheilatina, Wan Ngah, W.S., 2010. Adsorption of
Reactive Red 4 by immobilized chitosan on glass plates: towards the design of
immobilized TiO
2
echitosan synergistic photocatalyst-adsorption bilayer sys-
tem. Biochem. Eng. J. 49, 317e325.
Ngamlerdpokin, K., Kumjadpai, S., Chatanon, P., Tungmanee, U.,
Chuenchuanchom, S., Jaruwat, P., Lertsathitphongs, P., Hunsom, M., 2011.
Remediation of biodiesel wastewater by chemical- and electro-coagulation: a
comparative study. J. Environ. Manage. 92, 2454e2460.
Nishiro, N., Nakashimada, Y., 2007. Recent development of anaerobic digestion
processes for energy recovery from wastes. J. Biosci. Bioeng. 103, 105e112.
Osman, Z., Arof, A.K., 2003. FTIR studies of chitosan acetate based polymer elec-
trolytes. Electrochim. Acta 48, 939e999.
Paulino, A.T., Santos, L.B., Nozaki, J., 2008. Removal of Pb
2
, Cu
2
, and Fe
3
from
battery manufacture wastewater by chitosan produced from silkworm chry-
salides as a low-cost adsorbent. React. Funct. Polym. 68, 634e642.
Philipson, K.D., Ward, R., 1985. Effects of fatty acids on Na

-Ca
2
exchange and Ca
2
permeability of cardiac sarcolemmal vesicle. J. Biol. Chem. 260, 9666e9671.
Piyamongkala, K., Mekasut, L., Pongstabodee, S., 2008. Cutting uid efuent
removal by adsorption on chitosan and SDS-modied chitosan. Macromol. Res.
16, 492e502.
Rattanapan, C., Sawain, A., Suksaroj, T., Suksaroj, C., 2011. Enhanced efciency of
dissolved air otation for biodiesel wastewater treatment by acidication and
coagulation processes. Desalination 280, 370e377.
Ringqvist, L., Holmgren, A., born, I., 2002. Poorly humied peat as an adsorbent for
metals in wastewater. Water Res. 36, 2394e2404.
Shi, L., Zhang, X., Chen, Z., 2011. Removal of chromium (VI) from wastewater using
bentonite-supported nanoscale zero-valent iron. Water Res. 45, 886e892.
Siles, J.A., Martn, M.A., Chica, A.F., Martn, A., 2010. Anaerobic co-digestion of
glycerol and wastewater derived from biodiesel manufacturing. Biores. Technol.
101, 6315e6321.
Smith, P.C., Ngothai, Y., Nguyen, Q.D., ONeill, B.K., 2009. Alkoxylation of biodiesel
and its impact on low-temperature properties. Fuel 88 (4), 605e612.
Sokker, H.H., El-Sawy, N.M., Hassan, M.A., El-Anadouli, B.E., 2011. Adsorption of
crude oil from aqueous solution by hydrogel of chitosan based polyacrylamide
W. Pitakpoolsil, M. Hunsom / Journal of Environmental Management 133 (2014) 284e292 291
prepared by radiation induced graft polymerization. J. Hazard. Mater. 190, 359e
365.
Soleimani, M., Kaghazchi, T., 2008. Adsorption of gold ions from industrial waste-
water using activated carbon derived from hard shell of apricot stones-an
agricultural waste. Biores. Technol. 99, 5374e5383.
Suehara, K., Kawamoto, Y., Fujii, E., Kohda, J., Nakano, Y., Yano, T., 2005. Biological
treatment of wastewater discharged from biodiesel fuel production plant with
alkali-catalyzed transesterication. J. Biosci. Bioeng. 100, 437e442.
Tahir, S.S., Naseem, R., 2007. Removal of Cr(III) from tannery wastewater by
adsorption onto bentonite clay. Separ. Purif. Technol. 53, 312e321.
Wan Ngah, W.S., Fatinathan, S., 2010. Adsorption characterization of Pb(II) and
Cu(II) ions onto chitosan-tripolyphosphate beads: kinetic, equilibrium and
thermodynamic studies. J. Environ. Manage. 91, 958e969.
Wan Ngah, W.S., Endud, C.S., Mayanar, R., 2002. Removal of copper(II) ions from
aqueous solution onto chitosan and cross-linked chitosan beads. React. Funct.
Polym. 50, 181e190.
Wan Ngah, W.S., Hanaah, M.A.K.M., Yong, S.S., 2008. Adsorption of humic acid
from aqueous solutions on crosslinked chitosan-epichlorohydrinbeads: kinetics
and isotherm studies. Colloid Surf. B 65, 18e24.
Wang, J.P., Chen, Y.Z., Yuan, S.J., Sheng, G.P., Yu, H.Q., 2009. Synthesis and charac-
terization of a novel cationic chitosan-based occulant with a high water-
solubility for pulp mill wastewater treatment. Water Res. 43, 5267e5275.
Xiong, J.B., Mahmood, Q., 2010. Adsorptive removal of phosphate from aqueous
media by peat. Desalination 259, 59e64.
Zhang, X., Bai, R., 2003. Mechanisms and kinetics of humic acid adsorption onto
chitosan-coated granules. J. Colloid Interf. Sci. 264, 30e38.
Zhu, L., Ma, J., 2008. Simultaneous removal of acid dye and cationic surfactant from
water by bentonite in one-step process. Chem. Eng. J 139, 503e509.

Sciban, M., Radetic, B., Kevresan,



Z., Klasnja, M., 2007. Adsorption of heavy metals
from electroplating wastewater by wood sawdust. Biores. Technol. 98, 402e
409.
W. Pitakpoolsil, M. Hunsom / Journal of Environmental Management 133 (2014) 284e292 292

You might also like