You are on page 1of 8

Chemical Engineering Journal 174 (2011) 4148

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Adsorption of pesticides from aqueous solution onto banana stalk activated


carbon
J.M. Salman a , V.O. Njoku a,b , B.H. Hameed a,
a
b

School of Chemical Engineering, Engineering Campus, Universiti Sains Malaysia, 14300 Nibong Tebal, Penang, Malaysia
Department of Chemistry, Faculty of Science, Imo State University, P.M.B. 2000 Owerri, Nigeria

a r t i c l e

i n f o

Article history:
Received 24 May 2011
Received in revised form 8 August 2011
Accepted 9 August 2011
Keywords:
2,4-Dichlorophenoxyacetic acid
Bentazon
Isotherm
Kinetic
Thermodynamic

a b s t r a c t
Activated carbon was prepared from banana stalk by potassium hydroxide (KOH) and carbon dioxide
(CO2 ) activation and its ability to remove the pesticides, 2,4-dichlorophenoxyacetic acid (2,4-D) and
bentazon was explored. The banana stalk activated carbon (BSAC) was characterized by Fourier transform infrared spectroscopy (FT-IR) analysis. The adsorption kinetic data were analyzed using two kinetic
models: the pseudo-rst-order and pseudo-second-order models. The adsorption kinetics was better
represented by the pseudo-second-order model. The equilibrium adsorption data obtained at 30, 40 and
50 C were analyzed by Langmuir and Freundlich isotherm models and results showed that it was better
described by the Freundlich model. Higher adsorption capacity observed for 2,4-D than bentazon were
attributed to the presence of electron-withdrawing Cl groups on the aromatic ring and smaller molecular
size of 2,4-D. The thermodynamic parameters, G , H and S determined, showed that the adsorption of 2,4-D and bentazon onto BSAC was feasible, spontaneous and exothermic. The results showed that
BSAC is an efcient adsorbent for the adsorptive removal of 2,4-D and bentazon from aqueous solutions.
2011 Elsevier B.V. All rights reserved.

1. Introduction
The contamination of surface and ground water by pesticides
has become a serious environmental problem in recent years
due to the extensive application of these agrochemicals in crop
farms, orchards, elds and forest lands. This contamination arises
from surface runoff, leaching, wind erosion, deposition from aerial
applications, industrial discharges and various other sources. Consequently, pesticides have frequently been detected in water bodies
in different countries of the world [13]. Pesticides are harmful to
life because of their toxicity, carcinogenicity and mutagenicity [4].
The harmful inuence of pesticides on human health and the environment has resulted to the imposition of stringent legislation on
drinking water quality in many countries [5].
2,4-Dichlorophenoxyacetic acid (2,4-D) and bentazon are two
widely used anionic herbicides usually applied for the control
of broad-leaf weeds [6]. 2,4-D is moderately toxic but weakly
adsorbed by soil particles and therefore has high mobility and persistence in aqueous media [7,8]. Exposure to 2,4-D causes serious
eye and skin irritation, nausea, weakness, fatigue and in some cases
neurotoxic effects including inammation of nerve endings [9].
Bentazon is a contact post-emergence herbicide that can provide

Corresponding author. Tel.: +60 45996422; fax: +60 45941013.


E-mail address: chbassim@eng.usm.my (B.H. Hameed).
1385-8947/$ see front matter 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.cej.2011.08.026

effective control of broad-leaf weeds and sedges in many crops such


as, beans, corn, rice, peanuts and peas [10]. It is harmful if swallowed or absorbed through the skin and causes irritation to the eyes
[11]. The maximum allowable concentration of 2,4-D and bentazon
in drinking water is 100 and 30 g L1 respectively [11,12].
There are several procedures available for pesticides removal
from water which includes photocatalytic degradation [13,14],
ultrasound combined with photo-Fenton treatment [15], advanced
oxidation processes [16], aerobic degradation [17], electrodialysis
membranes [18], ozonation [19] and adsorption [20]. Adsorption
by activated carbon (AC) is one of the most widely used techniques
and has proven to be effective in the removal of pesticides from
aqueous solutions [21]. This is due to their mechanical stability,
low specicity, fast adsorption kinetics and high adsorption capacities [22,23]. ACs have high porosity and therefore have very high
surface area for adsorption. In addition, the chemical nature of their
surfaces enhances adsorption. However, one major restriction in its
use as adsorbent is the high cost of commercially available AC. Consequently, there is a growing interest in recent years, to produce AC
from renewable and inexpensive materials [24]. Researchers have
investigated the production of AC from many agricultural wastes
including pistachio shell, rambutan peel, mangosteen peel, corncob, olive-waste cakes, walnut shells and apricot stones [2531].
In Malaysia, banana (Musa sapientum) is massively cultivated and
consumed and remains the second most important fruit crop (after
durian). Banana stalk (BS) is therefore abundantly available and has

42

J.M. Salman et al. / Chemical Engineering Journal 174 (2011) 4148

Table 1
Some of properties and chemical structures of pesticides used.
Molecule weight (g mol1 )

Solubility in water (g L1 )

C8 H6 Cl2 O3

221.04

0.90

C10 H12 N2 O3 S

240.3

0.50

Pesticide

Formula

2,4-D

Bentazon

Chemical structure

no known industrial or commercial importance but rather constitutes a serious environmental problem. The use of banana stalk BS
as a precursor for AC will provide solution to environmental problems caused by this waste as well as produce a value-added product
from a low-cost material. In our previous study, AC prepared from
BS was found to be a good adsorbent for the removal of carbofuran
[32].
The purpose of this study is to determine the removal efciencies of AC prepared from BS for the pesticides, 2,4-D and bentazon
from articially contaminated solutions. The inuences of contact
time, initial concentration of pesticide, pH and temperature on the
adsorption capacities of BSAC for the pesticides were investigated.
In addition, kinetic modeling was carried out to understand the
mechanisms involved in the adsorption system. Isotherm models were also utilized in tting the equilibrium adsorption data
and adsorption performance of BSAC with respect to the two
pesticides were compared. Thermodynamic parameters were also
determined and the adsorption mechanisms further elucidated.
2. Materials and methods
2.1. Adsorbates
2,4-D and bentazon were obtained from SigmaAldrich (M) Sdn.
Bhd., Malaysia and were used separately as adsorbates. They were
of analytical grade and were used without further purication. Distilled water was used in the preparation of all solutions. Some
properties and chemical structures of the pesticides are given in
Table 1.
2.2. Preparation of activated carbon
Banana stalks were collected as wastes from local markets,
cut into small pieces of about 2 cm 2 cm, washed with water to
remove surface dirt and then dried to constant weight at 105 C.
The sample was carbonized and activated following the same
procedure described by Salman and Hameed [32]. The sample was
crushed, sieved to mesh size of 14 mm and carbonized in a stainless steel vertical tubular reactor placed in a tubular furnace under
puried nitrogen (99.995%) ow of 150 cm3 min1 at 600 C, with
a heating rate of 10 C min1 . The carbonization temperature was
maintained for 2 h before cooling still under nitrogen ow. The
carbonized material was soaked in potassium hydroxide (KOH)
solution with an impregnation ratio (KOH:char) of 2.5:1 (w/w) for
12 h and then dehydrated in an oven overnight at 105 C. The sample was then activated under the same conditions as carbonization
but to a nal temperature of 700 C. Once the nal temperature was
reached, the gas ow was switched from nitrogen to carbon dioxide
and this temperature was maintained for 2 h to achieve activation
[32]. The resulting AC was washed with 0.1 M HCl after cooling to

dissolve and remove any residual ash. This was followed by washing with hot distilled water until the pH of the supernatant solution
reached 6.57.0. This was followed by drying of the AC overnight
in an oven at 105 C.
2.3. Characterization of activated carbon
The surface functional groups of BSAC were determined using
Fourier transform infrared (FT-IR) spectroscopy (Perkin Elmer,
Model 2000 FT-IR, USA). FT-IR technique is an important technique
used in identifying characteristic surface functional groups on the
adsorbent, which in some cases are responsible for the binding of
the adsorbate molecules. The spectrum was recorded from 4000 to
400 cm1 adopting the KBr pellet method of sample handling.
2.4. Adsorption experiments
Stock solutions of 500 mg L1 2,4-D and bentazon were prepared
by dissolving appropriate amounts in distilled water. Solutions of
different initial concentrations of the pesticides were prepared by
serial dilution of the stock solutions. The kinetic and equilibrium
studies were undertaken at the pH of 3.5 and temperature of 30,
40 and 50 C. Batch adsorption studies were performed by shaking
200 mL of known initial concentrations of pesticide with 0.3 g BSAC
of 0.52.0 mm particle size in 250 mL conical asks. The effect of
solution initial pH on the adsorption of the pesticides by BSAC was
investigated by contacting 200 mL of 100 mg L1 2,4-D and bentazon solutions with 0.3 g BSAC in various asks of varying pH values
in the range 212 and at temperature of 30 C. The initial solution
pH was adjusted by drop-wise addition of 1.0 M HCl or 0.1 M NaOH
and solution pH was measured using a pH meter (EUTECH Instruments, Model Ecoscan, Singapore). The conical asks were then
covered with glass stoppers and placed inside a water-bath shaker
at shaking speed of 120 rpm. The samples were then withdrawn
at appropriate time intervals using glass syringes to determine the
residual concentrations of the adsorbates. The concentrations of
2,4-D and bentazon in the solutions before and after adsorption
were determined using a double beam UVvis spectrophotometer (Shimadzu UV-1700, Japan) at their maximum wavelength of
284 and 232 nm, respectively. The amounts of pesticide adsorbed
at time t, qt (mg g1 ) and at equilibrium, qe (mg g1 ) were calculated
using Eqs. (1) and (2).
qt =

(C0 Ct )V
W

(1)

qe =

(C0 Ce )V
W

(2)

where C0 , Ct and Ce (mg L1 ) are the liquid-phase concentrations of


pesticide at initial, time t and equilibrium, respectively. V (L) is the
volume of the solution and W (g) the mass of dry adsorbent used.

J.M. Salman et al. / Chemical Engineering Journal 174 (2011) 4148

43

781.72

Transmittance (%)

Banana stalk precursor


3850.79
3736.48

1160.25

2360.83

2929.75

1402.08
1320.04

1644.41

706.34

832.60

3446.38

1007.92

1621.99

Banana stalk char

617.85

1053.06

1406.16

2361.26
3736.10
3854.43

Banana stalk activated carbon


3736.69

1053.07

1537.20
2376.70

4000.0

3600

3200

2800

2400

2000

1800

1600

1400

1200

1000

800

600

400.0

Wavenumber (cm-1)
Fig. 1. FT-IR spectra of BS precursor, BS char and BSAC.

Fig. 1 shows the FT-IR spectra of the BS precursor, BS char


and BSAC, respectively. The gure shows that the peaks on the
BSAC spectrum are of low intensities compared to the corresponding peaks on the precursor and char spectra and many peaks
present in the precursor spectrum absolutely disappeared in BSAC
spectrum. This disappearance or weakening is consistent with
the breaking of many bonds in the KOH-impregnated BS precursor leading to the liberation and elimination of volatile species
and partial aromatization during carbonization [33]. Four peaks
were identied in the BSAC spectrum at 3736.69, 2376.70, 1537.20
and 1053.07 cm1 . The peak at 3736.69 cm1 could be ascribed
to OH stretching vibration in hydroxyl groups [34]. The band
at 2376.70 cm1 denotes the C C stretching vibrations in alkyne
groups while the peak around 1537.20 cm1 can be ascribed to
C C stretching vibration in aromatic rings [30]. The absorption at
1053.07 cm1 indicates the existence of CO stretching vibrations
in alcohols, phenols, acids, ethers or esters [35].

(a)

350
300
250

-1

3.1. Characterization of activated carbon

pesticide resulted to an increase in 2,4-D and bentazon uptake. This


is probably due to a higher mass transfer driving force as a result of
the increase in the number of molecules competing for the available binding sites on the adsorbent when the initial concentration

Ct (mg L )

3. Results and discussion

-1

Initial concentration of 2,4-D (mg L )


50
100
150
200
250
300

200
150
100
50
0
0

10

15

20

25

Time, t (h)

3.2. Effect of contact time and initial concentration of pesticide on


the adsorption

(b)

280

240
-1

Initial concentration of bentazon (mg L )


25
50
100
150
200
250

200

-1

Ct (mg L )

Fig. 2(a) and (b) shows the effect of contact time and initial
concentration of pesticide on the removal of 2,4-D and bentazon by BSAC. The results show that the residual concentration of
pesticides decrease with time, hence, the removal of 2,4-D and bentazon increase with time. Initially, adsorption increased sharply and
slowed gradually until equilibrium was attained in approximately
3 h for all initial concentrations of both pesticides studied. At this
point, the amount of pesticide desorbing from BSAC is in a state of
dynamic equilibrium with the amount of pesticide being adsorbed
on BSAC. This could be attributed to the larger number of vacant surface sites available for adsorption at the initial than the later stages
and the repulsive forces existing between pesticide molecules on
the BSAC surface and those in solution at the later stages.
Adsorption studies were conducted in the pesticide initial concentration ranges of 50300 and 25250 mg L1 for 2,4-D and
bentazon, respectively. Fig. 2(a) and (b) also shows the important
role of initial concentration of pesticide on the adsorption of 2,4D and bentazon onto BSAC. Increasing the initial concentration of

160

120

80

40

0
0

10

15

20

25

Time, t (h)
Fig. 2. Effect of contact time and pesticide initial concentration on adsorption onto
BSAC of (a) 2,4-D and (b) bentazon (30 C; agitation rate, 120 rpm; BSAC dose,
1000 mg L1 ).

44

J.M. Salman et al. / Chemical Engineering Journal 174 (2011) 4148

60

(a)

240
-1

200

50

-1

-1

50 mg g
100 mg g
-1
-1
200 mg g
250 mg g
Pseudo-first-order fit
Pseudo-second-order fit

150 mg g
-1
300 mg g

160

qe (mg/g)

-1

qt (mg g )

40

30

2,4-D
Bentazon

120

80

20

40
10

10

10

12

14

pH

(b)

Fig. 3. Effect of solution initial pH on 2,4-D and bentazon equilibrium adsorption at


30 C.

The effect of solution initial pH on 2,4-D and bentazon adsorption by BSAC was examined and the result is presented in Fig. 3. The
equilibrium adsorption of both pesticides decreased with increase
in initial pH of solution and this is attributable to the increased electrostatic repulsion between the pesticide ions and BSAC surface.
Both 2,4-D and bentazon are weak acids with pKa of 2.63.0 and 3.3,
respectively [8,37]. The FT-IR analysis revealed the presence of negatively charged functional groups on the BSAC surface. At pH above
the pKa values, the pesticides exist predominantly in the anionic
forms and as pH is increased, the extent of dissociation of these
molecules increased and they became more negatively charged.
This resulted to increased electrostatic repulsion or dispersion
between the pesticide ions and the BSAC surface and subsequently
decreased equilibrium adsorption as was observed in the pH range
studied. Similar observations were reported for the adsorption of
2,4-D using fertilizer and steel industry wastes and for the adsorption of bentazon onto AC cloth [11,37].
3.4. Adsorption kinetics
The modeling of adsorption kinetics was investigated using
two models, namely, the pseudo-rst-order (PFO) and the pseudosecond-order (PSO) models. The non-linear form of the PFO
equation [38] is expressed as Eq. (3).
(3)

where k1 (h1 ) is the PFO rate constant and t (h) the contact time.
The non-linear form of the PSO kinetic equation [39] is given by Eq.
(4):
qt =

q2e k2 t
1 + qe k2 t

25

-1

-1

25 mg L
50 mg L
-1
-1
150 mg L
200 mg L
Pseudo-first-order fit
Pseudo-second-order fit

-1

100 mg L
-1
250 mg L

100

-1

80
60
40
20

3.3. Effect of pH on pesticide adsorption

qt = qe (1 ek1 t )

20

140
120

qt (mg g )

of pesticide is increased [36]. However, the percent removal efciency of 2,4-D decreased from 98.4 to 85.4% as the 2,4-D initial
concentration increased from 50 to 300 mg L1 while for bentazon, the percent removal efciency decreased from 96.5 to 61.6% as
the bentazon initial concentration increased from 25 to 250 mg L1 .
Therefore, the adsorption of 2,4-D and bentazon by BSAC has strong
dependence on the initial concentration of the pesticides.

15

Time, t (h)

(4)

10

15

20

25

Time, t (h)
Fig. 4. Fit of pseudo-rst-order and pseudo-second-order kinetic models for the
adsorption onto BSAC of (a) 2,4-D and (b) bentazon (30 C; agitation rate, 120 rpm;
BSAC dose, 1000 mg L1 ).

where k2 (g mg1 h1 ) is the PSO rate constant. The kinetic data


were tted to the two models by non-linear procedure and the plots
are shown in Fig. 4(a) and (b). The kinetic model parameters as well
2 ) and root-mean squared
as the adjusted correlation coefcient (Radj
error (RMSE) values obtained are listed in Table 2. The applicability of the kinetic models was adjudged by the closeness to unity of
2 values and lowest RMSE values. In addition, the agreement
the Radj
between the qe values obtained from the models and the experimental values were utilized in conrming the suitability of the
models in tting the kinetic data. In view of these considerations,
the PSO model was more suitable for describing the adsorption
of 2,4-D and bentazon onto BSAC at all initial concentrations of
pesticides studied. A better t to the PSO model suggests that the
adsorption rate is dependent more on the availability of adsorption sites than the concentration of the pesticides in solution [40].
Similar observations have previously been reported [41,42].
3.5. Adsorption isotherms
Adsorption isotherms indicate how the adsorbate is distributed
between the liquid phase and the solid phase when the adsorption process reaches equilibrium state. In the present study, two
isotherm models have been used in tting the equilibrium data,
namely, Langmuir and Freundlich isotherm models. The Langmuir
model assumes monolayer adsorption onto a surface containing

J.M. Salman et al. / Chemical Engineering Journal 174 (2011) 4148

45

Table 2
Kinetics parameters for the adsorption of 2,4-D and bentazon onto BSAC at 30 C.
C0 (mg L1 )

2,4-D
50
100
150
200
250
300
Bentazon
25
50
100
150
200
250

qe,exp (mg g1 )

Pseudo-rst-order

Pseudo-second-order

k1 (h1 )

qe (mg g1 )

2
Radj

RMSE

k2 (g mg1 h1 )

qe (mg g1 )

2
Radj

RMSE

32.62
64.61
95.07
122.28
145.45
168.03

2.105
2.483
2.692
2.523
2.626
2.216

32.14
63.11
92.43
118.42
141.81
164.10

0.995
0.994
0.990
0.992
0.988
0.992

0.734
1.631
2.987
3.591
5.232
5.005

0.121
0.078
0.059
0.042
0.037
0.025

33.54
65.54
95.89
123.00
147.21
171.32

0.996
0.996
0.997
0.997
0.999
0.998

0.683
1.334
1.732
2.209
0.752
2.397

15.47
30.10
50.63
67.99
81.59
100.95

1.210
1.447
1.658
2.038
2.065
1.928

15.50
29.33
48.98
66.53
79.91
95.09

0.988
0.978
0.984
0.992
0.991
0.966

0.600
1.515
2.090
2.020
2.495
5.904

0.116
0.074
0.055
0.053
0.046
0.033

16.51
31.16
51.84
69.60
83.58
100.64

0.996
0.989
0.998
0.998
0.998
0.995

0.319
1.048
0.718
0.936
1.002
2.184

nite number of adsorption sites of uniform energies with no


transmigration of adsorbate in the plane of surface. The Langmuir
isotherm [43] is described by Eq. (5):
qe =

qm KL Ce
1 + KL Ce

used in previous studies. The high 2,4-D and bentazon adsorption


capacities exhibited by BSAC shows that BS is a good precursor
for the preparation of AC for the adsorptive removal of 2,4-D and
bentazon.

(5)

where qm (mg g1 ) is the monolayer adsorption capacity, and KL


(L mg1 ) the equilibrium adsorption constant related to the energy
of adsorption. The Freundlich isotherm [44], known to be satisfactory at low concentrations of adsorbate and based on the
assumption of adsorption on heterogeneous surfaces, is given by
Eq. (6):

(a)

200
180
160
140

-1

120
100
Experimental data
Langmuir fit
Freundlich fit

80
60
40
20
0
0

10

20

30

40

50

80

100

-1

Ce (mg L )

(b)

100

80

-1

where KF (mg g1 (L mg1 )1/n ) and n are Freundlich constants


related to the adsorption capacity and adsorption intensity, respectively. The equilibrium data for the adsorption of 2,4-D and
bentazon onto BSAC were tted to the two isotherm models by
non-linear regression. The non-linear ts of the adsorption equilibrium data at 30 C are depicted in Fig. 5(a) and (b) while the
isotherm parameter values and correlation coefcients obtained
2
at 30, 40 and 50 C are listed in Table 3. A comparison of the Radj
and RMSE values for the two models indicates that the Freundlich
isotherm model provides a better correlation of the experimental data for the adsorption of 2,4-D and bentazon onto BSAC.
The good t obtained with the Freundlich model suggests that
the adsorption of 2,4-D and bentazon onto BSAC may involve
multi-layer adsorption with interactions between the pesticide
molecules and the heterogeneous nature of the BSAC surface is
also implied. Similar observations have been made on the removal
of 2,4-dichlorophenol and 2,4,6-trichlorophenol by cattail berbased AC [21]. Comparison of the adsorption capacities of BSAC
shows higher adsorption in the case of 2,4-D than bentazon.
This could be attributed to the smaller molecular size of 2,4-D
which can easily penetrate the pores of the AC than the bentazon molecules. The higher maximum adsorption capacity for 2,4-D
could also be due to the presence of Cl groups on 2,4-D, which have
more pronounced electron withdrawing properties. The presence
of Cl groups in 2,4-D leads to more withdrawal of -electrons
from the aromatic ring, which constitutes an electron-acceptor
group and subsequently leads to enhancement of the adsorption
afnity [31].
Table 4 offers comparison of the Langmuir maximum adsorption
capacities of BSAC for 2,4-D and bentazon with other ACs reported
in literature. Though there were variations in some experimental
conditions, the table shows that the adsorption capacities of BSAC
for both pesticides are high compared to other activated carbons

qe (mg g )

(6)

qe (mg g )

1/n

qe = KF Ce

60

Experimental data
Langmuir fit
Freundlich fit

40

20

0
0

20

40

60
-1

Ce (mg L )
Fig. 5. Adsorption isotherm tting for adsorption onto BSAC at 30 C of (a) 2,4-D
and (b) bentazon.

46

J.M. Salman et al. / Chemical Engineering Journal 174 (2011) 4148

Table 3
Langmuir and Freundlich isotherm model parameters and correlation coefcients for adsorption of 2,4-D and bentazon onto BSAC at 30 C.
Pesticide

Isotherm model

Parameter

2,4-D

Langmuir

qm (mg g1 )
KL (L mg1 )
2
Radj

Temperature ( C)

RMSE
Freundlich

n
KF (L g1 )
2
Radj
RMSE

Bentazon

Langmuir

qm (mg g1 )
KL (L mg1 )
2
Radj

Freundlich

n
KF (L g1 )
2
Radj

RMSE

RMSE

30

40

50

196.33
0.279
0.961

178.60
0.158
0.978

166.7
0.135
0.973

7.815

4.236

4.310

2.60
42.01
0.979

2.77
40.85
0.998

3.13
40.05
0.998

2.632

1.104

1.218

115.07
0.067
0.943

92.60
0.067
0.972

90.90
0.098
0.975

6.952

4.384

4.065

2.44
14.94
0.961

2.87
16.36
0.997

2.97
18.14
0.977

4.908

2.540

2.856

Table 4
Comparison of 2,4-D and bentazon Langmuir adsorption capacities of various activated carbons.
Pesticide

Activated carbon type

Adsorption capacity (mg g1 )

Reference

2,4-D

Coal-based granular activated carbon F400


Modied coal-based granular activated carbon F400
Coal-based granular activated carbon F300
Date stone-based activated carbon
Banana stalk-based activated carbon

8.1310.48
11.7616.25
181.82
238.10
196.33

[41]
[41]
[45]
[46]
This study

Bentazon

Oxidized activated carbon cloth


Date seed-based activated carbon
Activated carbon cloth
Activated carbon cloth
Banana stalk-based activated carbon

17
86.26
115
151
115.07

[37]
[42]
[37]
[47]
This study

3.6. Adsorption thermodynamics

equilibrium concentration of pesticide in solution. Subsequently,


the values of G were calculated from Eq. (9).

Equilibrium experiments were performed at temperatures of


30, 40 and 50 C to determine the inuence of temperature on the
adsorption of 2,4-D and bentazon onto BSAC. The results showed
that increase in temperature resulted to decrease in adsorption
capacity of BSAC for both pesticides indicating the exothermic
nature of the adsorption. This could be attributed to the enhancement of the desorption step of the adsorption mechanism due to
the dislodgement of the adsorptive forces. Thermodynamic parameters related to the adsorption process, standard free energy change
(G ), standard enthalpy change (H ) and standard entropy
change (S ) were determined. The values of H and S were
determined from the slope and intercept of the vant Hoff plot of
ln Kd versus 1/T for the highest initial concentration of pesticide
using the expression given in Eq. (7).
ln Kd =

H
S

R
RT

(7)

where Kd is the equilibrium distribution coefcient, R


(8.314 J mol1 K1 ) the universal gas constant and T (K) the
absolute temperature. The value of Kd was obtained from the
expression in Eq. (8):

G = H TS

The values of G , H and S obtained are summarized in


2 values of the linear ts were higher than 0.997
Table 5. The Radj
for both pesticides, indicating high condence in the parameters
values obtained. The negative values of G at all temperatures
studied indicated that the adsorption processes were spontaneous
and feasible. The negative values of H conrm the exothermic
character of the adsorption of 2,4-D and bentazon onto BSAC and
suggests that the adsorption is of the diffusion type and hence
physisorption process. The physisorption mechanism inferred is
supported by the ndings from the isotherm study which indicated
multilayer adsorption. Similar results were reported for adsorption
of phenol onto activated carbon prepared from tobacco residues
and for the adsorption of basic dye on activated carbon prepared
from coconut husk [48,49]. The positive values of S showed
increased randomness at the solid-solution interfaces during the

Table 5
Thermodynamic parameters of the adsorption of 2,4-D and bentazon onto BSAC.
H (kJ mol1 )

Kd =

Cad,e
Ce

(9)

S (J mol1 K1 )

30 C

40 C

50 C

5.48

7.12

7.18

7.23

2.88

1.98

2.01

2.04

(8)

where
Cad,e
(mg L1 )
is
the
equilibrium
concentration of pesticide on BSAC and Ce (mg L1 ) is the

2,4-D
5.46
Bentazon
1.11

G (kJ mol1 )

J.M. Salman et al. / Chemical Engineering Journal 174 (2011) 4148

adsorption processes and conrmed high preference of 2,4-D and


bentazon adsorption onto BSAC.
4. Conclusion
The present study revealed that banana stalk (BS) can effectively be used as a raw material for preparation of activated carbon
for the removal of 2,4-D and bentazon from aqueous solutions.
The adsorption system followed the pseudo-second-order kinetic
model platform which suggests that the adsorption rate is dependent more on the availability of the adsorption sites than the
concentration of the pesticides in solution. The Freundlich isotherm
model provided a better correlation of the experimental equilibrium data suggesting that the adsorption of 2,4-D and bentazon
onto BSAC may involve multi-layer adsorption with interactions
between the pesticide molecules. Thermodynamic parameters
obtained indicated that the adsorption was feasible, spontaneous
and exothermic in nature. Higher adsorption capacity observed
for 2,4-D could be due to the presence of electron-withdrawing
Cl groups on the aromatic ring and smaller molecular size of
2,4-D. Adsorption capacities of BSAC for 2,4-D and bentazon were
high when compared with activated carbon from other precursors
showing that BS is a good precursor for the preparation of activated
carbon for the adsorptive removal of 2,4-D and bentazon.
Acknowledgments
The authors acknowledge the nancial support provided by Universiti Sains Malaysia under the Research University (RU) Scheme
(Project No. 1001/PJKIMIA/814072). The second author gratefully
acknowledges the award of USM-TWAS postdoctoral fellowship
(2009) in aid for research.
References
[1] Environment Agency, The Annual Report of the Environment Agency Pesticide
Monitoring Programme, Environment Agency, UK, 2002.
[2] G. Akcay, M. Akcay, K. Yurdakoc, Removal of 2,4-dichlorophenoxyacetic acid
from aqueous solutions by partially characterized organophilic sepiolite: thermodynamic and kinetic calculations, J. Colloid Interface Sci. 281 (2005) 2732.
[3] R.A. Rebich, R.H. Coupe, E.M. Thurman, Herbicide concentrations in the Mississippi River Basinthe importance of chloroacetanilide herbicide degradates,
Sci. Total Environ. 321 (2004) 189199.
[4] International Agency for Research on Cancer (IARC), Overall evaluations of carcinogenicity: An updating of IARC monographs
volumes 1 to 42, Supplement 7, WHO, Lyon, France, 1987.
http://monographs.arc.fr/ENG/Monographs/PDFs/index.php.
[5] A. Derylo-Marczewska, M. Blachnio, A.W. Marczewski, A. Swiatkowski, B. Tarasiuk, Adsorption of selected herbicides from aqueous solutions on activated
carbon, J. Therm. Anal. Calorim. 101 (2010) 785794.
[6] C. Tomlin, The Pesticide Manual, 10th ed., Crop Protection Publications, Boca
Raton, 1994.
[7] M. Arias-Estvez, E. Lpez-Periago, E. Martnez-Carballo, J. Simal-Gndara, J.-C.
Mejuto, L. Garca-Ro, The mobility and degradation of pesticides in soils and
the pollution of groundwater resources, Agric. Ecosyst. Environ. 123 (2008)
247260.
[8] J. Li, Y. Li, J. Lu, Adsorption of herbicides 2,4-D and acetochlor on
inorganicorganic bentonites, Appl. Clay Sci. 46 (2009) 314318.
[9] Extoxnet Data Sheet on 2,4-D, Pesticide Management Programme, Cornell University, US, 1994.
[10] R. Pourata, A.R. Khataee, S. Aber, N. Daneshvar, Removal of the herbicide bentazon from contaminated water in the presence of synthesized nanocrystalline
TiO2 powders under irradiation of UV-C light, Desalination 249 (2009) 301307.
[11] V.K. Gupta, I. Ali, Suhas, V.K. Saini, Adsorption of 2,4-D and carbofuran pesticides
using fertilizer and steel industry wastes, J. Colloid Interface Sci. 299 (2006)
556563.
[12] World Health Organisation (WHO), WHO Guidelines for Drinking Water Quality, World Health Organisation, Geneva, 2004.
[13] M. Ugurlu, M.H. Karaoglu, TiO2 supported on sepiolite: preparation, structural
and thermal characterization and catalytic behaviour in photocatalytic treatment of phenol and lignin from olive mill wastewater, Chem. Eng. J. 166 (2011)
859867.
[14] J. Gong, C. Yang, W. Pu, J. Zhang, Liquid phase deposition of tungsten doped
TiO2 lms for visible light photoelectrocatalytic degradation of dodecylbenzenesulfonate, Chem. Eng. J. 167 (2011) 190197.

47

[15] H. Katsumata, T. Kobayashi, S. Kaneco, T. Suzuki, K. Ohta, Degradation of linuron


by ultrasound combined with photo-Fenton treatment, Chem. Eng. J. 166 (2011)
468473.
[16] T. Zhou, T.-T. Lim, S.-S. Chin, A.G. Fane, Treatment of organics in reverse osmosis
concentrate from a municipal wastewater reclamation plant: feasibility test of
advanced oxidation processes with/without pretreatment, Chem. Eng. J. 166
(2011) 932939.
[17] H.M. Rajashekara Murthy, H.K. Manonmani, Aerobic degradation of technical
hexachlorocyclohexane by a dened microbial consortium, J. Hazard. Mater.
149 (2007) 1825.
[18] L.J. Banasiak, B. Van der Bruggen, A.I. Schfer, Sorption of pesticide endosulfan
by electrodialysis membranes, Chem. Eng. J. 166 (2011) 233239.
[19] M.I. Maldonado, S. Malato, L.A. Perez-Estrada, W. Gernjak, I. Oller, X. Domenech,
J. Peral, Partial degradation of ve pesticides and an industrial pollutant by
ozonation in a pilot-plant scale reactor, J. Hazard. Mater. 38 (2006) 363369.
[20] A.H. Al-Muhtase, K.A. Ibrahim, A.B. Albadarin, O. Ali-khashman, G.M. Walker,
M.N.M. Ahmad, Remediation of phenol-contaminated water by adsorption
using poly (methyl methacrylate) (PMMA), Chem. Eng. J. 168 (2011) 691699.
[21] L. Ren, J. Zhang, Y. Li, C. Zhang, Preparation and evaluation of cattail ber-based
activated carbon for 2,4-dichlorophenol and 2,4,6-trichlorophenolremoval,
Chem. Eng. J. 168 (2011) 553561.
[22] A.A. Halim, H.A. Aziz, M.A.M. Johari, K.S. Arifn, Comparison study of ammonia
and COD adsorption on zeolite, activated carbon and composite materials in
landll leachate treatment, Desalination 262 (2010) 3135.
[23] S.L Wang, Y.M. Tzou, Y.H. Lu, G. Sheng, Removal of 3-chlorophenol from water
using rice-straw-based carbon, J. Hazard. Mater. 147 (2007) 313318.
[24] D. Mohan, K.P. Singh, V.K. Singh, Wastewater treatment using low cost activated
carbons derived from agricultural byproductsa case study, J. Hazard. Mater.
152 (2008) 10451053.
[25] H. Dolas, O. Sahin, C. Saka, H. Demir, A new method on producing high surface
area activated carbon: the effect of salt on the surface area and the pore size
distribution of activated carbon prepared from pistachio shell, Chem. Eng. J.
166 (2011) 191197.
[26] M.A. Ahmad, R. Alrozi, Optimization of rambutan peel based activated carbon
preparation conditions for Remazol Brilliant Blue R removal, Chem. Eng. J. 166
(2011) 280285.
[27] M.A. Ahmad, R. Alrozi, Optimization of preparation conditions for mangosteen
peel-based activated carbons for the removal of Remazol Brilliant Blue R using
response surface methodology, Chem. Eng. J. 165 (2010) 883890.
[28] Y. Sun, P.A. Webley, Preparation of activated carbons from corncob with large
specic surface area by a variety of chemical activators and their application in
gas storage, Chem. Eng. J. 162 (2010) 883892.
[29] R. Baccar, P. Blnquez, J. Bouzid, M. Feki, M. Sarr, Equilibrium, thermodynamic and kinetic studies on adsorption of commercial dye by activated carbon
derived from olive-waste cakes, Chem. Eng. J. 165 (2010) 457464.
[30] J. Yang, K. Qiu, Preparation of activated carbons from walnut shells via vacuum
chemical activation and their application for methylene blue removal, Chem.
Eng. J. 165 (2010) 209217.
[31] B. Petrova, T. Budinova, B. Tsyntsarski, V. Kochkodan, Z. Shkavro, N. Petrov,
Removal of aromatic hydrocarbons from water by activated carbon from apricot stones, Chem. Eng. J. 165 (2010) 258264.
[32] J.M. Salman, B.H. Hameed, Removal of insecticide carbofuran from aqueous
solutions by banana stalks activated carbon, J. Hazard. Mater. 176 (2010)
814819.
[33] E. Yagmur, M. Ozmak, Z. Aktas, A novel method for production of activated
carbon from waste tea by chemical activation with microwave energy, Fuel 87
(2008) 32783285.
[34] A.A. Ahmad, B.H. Hameed, Reduction of COD and color of dyeing efuent from
a cotton textile mill by adsorption onto bamboo-based activated carbon, J.
Hazard. Mater. 172 (2009) 15381543.
[35] W. Tongpoothorn, M. Sriuttha, P. Homchan, S. Chanthai, C. Ruangviriyachai,
Preparation of activated carbon derived from Jatropha curcas fruit shell by simple thermo-chemical activation and characterization of their physico-chemical
properties, Chem. Eng. Res. Des. 89 (2011) 335340.
[36] J. Fan, J. Zhang, C. Zhang, L. Ren, Q. Shi, Adsorption of 2,4,6-trichlorophenol from
aqueous solution onto activated carbon derived from loosestrife, Desalination
267 (2011) 139146.
[37] C.O. Ania, F. Beguin, Mechanism of adsorption and electrosorption of bentazone on activated carbon cloth in aqueous solutions, Water Res. 41 (2007)
33723380.
[38] S. Lagergren, Zur theorie der sogenannten adsorption geloester stoffe, Kungliga
Svenska Vetenskapsakad Handlingar 24 (1898) 139.
[39] Y.S. Ho, G. McKay, Pseudo-second-order model for sorption processes, Process
Biochem. 34 (1999) 451465.
[40] Y. Liu, New insights into pseudo-second-order kinetic equation for adsorption,
Colloids Surf. A 320 (2008) 275278.
[41] P. Chingombe, B. Saha, R.J. Wakeman, Effect of surface modication of an engineered activated carbon on the sorption of 2,4-dichlorophenoxy acetic acid and
benazolin from water, J. Colloid Interface Sci. 297 (2006) 434442.
[42] J.M. Salman, V.O. Njoku, B.H. Hameed, Bentazon and carbofuran adsorption
onto date seed activated carbon: kinetics and equilibrium, Chem. Eng. J. 173
(2011) 361368.
[43] I. Langmuir, The adsorption of gases on plane surfaces of glass, mica and platinum, J. Am. Chem. Soc. 40 (1918) 13611403.
[44] H. Freundlich, ber die adsorption in lsungen (Adsorption in solution), Z. Phys.
Chem. 57 (1906) 384470.

48

J.M. Salman et al. / Chemical Engineering Journal 174 (2011) 4148

[45] J.M. Salman, B.H. Hameed, Adsorption of 2,4-dichlorophenoxyacetic acid and


carbofuran pesticides onto granular activated carbon, Desalination 256 (2010)
129135.
[46] B.H. Hameed, J.M. Salman, A.L. Ahmad, Adsorption isotherm and kinetic modeling of 2,4-D pesticide on activated carbon derived from date stones, J. Hazard.
Mater. 163 (2009) 121126.
[47] E. Ayranci, N. Hoda, Adsorption of bentazon and propanil from aqueous solutions at the high area activated carbon-cloth, Chemosphere 57 (2004) 755762.

[48] M. Kilic, E. Apaydin-Varol, A.E. Ptn, Adsorptive removal of phenol from


aqueous solutions on activated carbon prepared from tobacco residues:
equilibrium, kinetics and thermodynamics, J. Hazard. Mater. 189 (2011)
397403.
[49] I.A.W. Tan, A.L. Ahmad, B.H. Hameed, Adsorption of basic dye on
high-surface-area activated carbon prepared from coconut husk: equilibrium, kinetic and thermodynamic studies, Hazard. Mater. 154 (2008)
337346.

You might also like