Anda di halaman 1dari 9

Chemical Engineering Journal 174 (2011) 117125

Contents lists available at SciVerse ScienceDirect

Chemical Engineering Journal


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

NaOH-activated carbon of high surface area produced from coconut shell:


Kinetics and equilibrium studies from the methylene blue adsorption
Andr L. Cazetta a , Alexandro M.M. Vargas a , Eurica M. Nogami a , Marcos H. Kunita a ,
Marcos R. Guilherme a , Alessandro C. Martins a , Tais L. Silva b , Juliana C.G. Moraes a , Vitor C. Almeida a,
a
b

Department of Chemistry, Universidade Estadual de Maring, Av. Colombo 5790, CEP 87020-900, Maring, Paran, Brazil
Department of Civil Engineering, Universidade Estadual de Maring, Av. Colombo 5790, CEP 87020-900, Maring, Paran, Brazil

a r t i c l e

i n f o

Article history:
Received 5 July 2011
Received in revised form 20 August 2011
Accepted 22 August 2011
Keywords:
NaOH-activated carbon
Coconut shell
Methylene blue
Adsorption

a b s t r a c t
Activated carbons (ACs) of coconut shell produced by NaOH activation at impregnation ratios of
NaOH:char (w/w) equal to 1:1 (AC-1), 2:1 (AC-2) and 3:1 (AC-3) were prepared. The properties of these
carbons, including BET surface area, pore volume, pore size distribution, and pore diameter, were characterized from N2 adsorption isotherms. It was found that the ACs are essentially microporous and that
the BET surface area was in order of 783 m2 g1 for AC-1, 1842 m2 g1 for AC-2, and 2825 m2 g1 for
AC-3. Scanning electron microscopy images showed a high pore development while Boehm method and
Fourier-transform infrared spectroscopy spectra indicated the presence of acid functional groups which
was conrmed by pH drift method. The adsorption equilibrium and kinetics of methylene blue (MB) onto
AC-3 were carried out. Experimental data were tted to the four isotherm models (Langmuir, Freundlich,
Toth and RedlichPeterson), and was found that Langmuir model presented the best t, showing maximum monolayer adsorption capacity of 916 mg g1 . The kinetic studies showed that experimental data
follow pseudo-second-order model. The mechanism of the adsorption process was described from the
intraparticle diffusion model. The AC-3 has a high surface area and showed to be an efcient adsorbent
for removal of MB from aqueous solutions.
2011 Elsevier B.V. All rights reserved.

1. Introduction
The impacts caused by a variety of industrial pollutants and
growing concern for environmental issues have led to the search
for new methods of treatment, and development of new materials
that are able to reduce these environmental problems. Among the
various types of existing efuent treatment, the adsorption process using activated carbon is of easy application, good efciency
and economically viable [1].
Activated carbons (ACs) are porous materials that have a high
surface area and high adsorption capacity, which can remove a
wide variety of pollutants such as dyes, heavy metals, pesticides
and gases. Due to its adsorptive properties, the ACs are used to
purify, detoxify, deodorize, lter, discolor or alter the concentration of many liquid and gaseous materials. These applications are
of great interest in various industrial sectors such as food, pharmaceutical, chemical, oil, mining, and especially in treatment of
drinking water [2].
Because of the high cost and non-renewable source of commercially available AC, in recent years, researchers have studied the

Corresponding author. Tel.: +55 44 3261 3678; fax: +55 3261 4334.
E-mail address: vcalmeida@uem.br (V.C. Almeida).
1385-8947/$ see front matter 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.cej.2011.08.058

production of ACs from cheap and renewable precursors, such as


olive husk [3], coffee endocarp [4], cotton stalks [5], plum kernels
[6], r wood [7,8], pistachio shell [8], olive stone [9], bamboo [10]
and amboyant pods [11].
The coconut shell is a potential precursor for the production of
ACs, because it corresponds to 35% of the fruit mass. Brazil is the
fourth largest producer of coconut in the world: its current production is equivalent to 6% of world, which generates a large amount
of waste from this fruit. The use of coconut shell for the production
of ACs has been studied by some researchers, which reported the
chemical activation of the material from chemical reagents, such
as KOH [1214], H3 PO4 NaHCO3 [15] and ZnCl2 [16]. The production of ACs from coconut shell by chemical activation with NaOH,
so far, has not been reported in the literature yet. NaOH has been
used in studies for production of ACs with high surfacial area from
r wood [17], plum kernels [6] and amboyant pods [18]. Among
the basic reagents, Tseng [19] reported that NaOH activation in
comparison with KOH activation has advantages such as: (i) lower
dosage (weight measurement), (ii) cheaper, (iii) more environmentally friendly, and (iv) less corrosive.
Methylene blue (MB) is a cationic dye that is most commonly used for coloring. It is generally used for dyeing cotton,
wool, and silk. MB can cause eye burns in humans and animals,
methemoglobinemia, cyanosis, convulsions, tachycardia, dyspnea,

118

A.L. Cazetta et al. / Chemical Engineering Journal 174 (2011) 117125

irritation to the skin, and if ingested, irritation to the gastrointestinal tract, nausea, vomiting, and diarrhea [20]. This dye has been
studied because of its known strong adsorption onto solids, and it
often serves as a model compound for removing organic contaminants and colored bodies from aqueous solutions [10].
The focus of this research was to prepare ACs from coconut shell
by NaOH activation and evaluate its potential for adsorption in
removal of MB from the aqueous solution. The textural and chemical characterizations of the ACs were also performed. In order to
establish the removal capacity of this adsorbent, different models of
isotherms and adsorption kinetics were tted to the experimental
data.
2. Materials and methods
2.1. Raw material
Coconuts used for preparation of AC were obtained from local
market in Maring, Paran, Brazil. The fruits were washed with distilled water, and subsequently dried at 110 C for 48 h. After that,
the shells were removed from the fruits, dried at 110 C for 48 h,
ground and granulometrically separated. The proximate analysis of
the raw material used in this study by ASTM-D1762 Standards [21]
revealed moisture, ash, volatile matter, and xed carbon content
values of 3.52, 1.28, 74.60, and 20.60%, respectively.
2.2. Preparation and characterization of activated carbon
The raw material, which had particle size between 250 m
and 425 m, was placed in a horizontal stainless steel reactor and
heated in a furnace at the rate of 20 C min1 from room temperature to 500 C, and maintained at this temperature for 2.0 h. The
obtained char was mixed with varying amounts of NaOH pellets
and 10 mL of water, at the ratios of 1:1, 2:1, and 3:1 (NaOH:char) in
a vertical stainless steel reactor under magnetic stirring for 2 h and
then dried at 130 C for 4 h. The reactor containing the dry mixture
was set into a furnace under N2 ow of 100 cm3 min1 , and heated
at the rate of 20 C min1 to the nal temperature of 700 C, which
was maintained for 1.5 h. After cooling, the resulting mixture was
washed with a 0.1 M solution of HCl followed by hot distilled water
until pH 6.5 to eliminate activating agent residues and other inorganic species formed during the process. In the washing step, the
activated carbon was separated using 0.45-m membrane lters.
The carbon obtained was dried at 110 C for 24 h and kept in tightly
closed bottles for further analysis. The prepared activated carbon
at NaOH:char ratios of 1:1, 2:1, and 3:1 were labeled as AC-1, AC-2,
and AC-3, respectively.
The activated carbon yield was dened as the nal weight of
product after activation, washing, and drying. The percent yield
was determined from the relation:
wc
Yield (%) =
100
(1)
wo
where wc and wo are the nal activated carbon dry weight (g) and
the precursor dry weight (g), respectively.
Textural characterization of the ACs was carried out by N2
adsorption at 77 K using QuantaChrome Nova 1200 surface area
analyzer. The surface area, SBET , was determined from isotherms
using the BrunauerEmmettTeller equation (BET). The total pore
volume, VT , was dened as the volume of liquid nitrogen corresponding to the amount adsorbed at a relative pressure of
P/P0 = 0.99 [22]. The micropore volume, V , was determined with
the DubinineRadushkevich equation and the mesopore volume,
Vm , was calculated as the difference between VT and V . The average pore diameter, Dp , was calculated using the relation 4VT /SBET ,
and the pore size distribution, by the BJH method. Scanning electron

microscopy (SEM) analysis (Shimadzu, model SS 550) was carried


out for raw material, AC-1, AC-2 and AC-3 to study the development
of porosity.
The surfaces of ACs were chemically characterized by Boehm
titration [23] and pH drift method [24]. The surface organic structures of the raw material and AC were studied by FT-IR spectra
recorded at 4 cm1 of resolution and 20 scans min1 between 4000
and 400 cm1 using a Bomem MB-100 spectrometer.
2.3. Batch adsorption studies
The cationic dye, MB or Basic Blue 9 (B. Herzog, Germany), was
used as an adsorbate. It has the molecular formula C16 H18 N3 SCl and
the molecular weight of 319.85 g mol1 .
A stock solution of 1.0 g L1 was prepared by dissolving the
appropriate amount of MB in 100 mL and completing the volume of
1000 mL with distilled water. Batch adsorption was performed in
a set of 50 mL plastic asks containing 25 mL of MB solutions with
various initial concentrations (1001000 mg L1 ). The amount of
0.025 g of AC was added to each ask and kept at 25 C on a shaker.
For equilibrium studies, the experiment was carried out for 2.5 h
to ensure equilibrium was reached. Previous tests were performed
varying the solution pH from 2 to 10 and the MB removal of approximately 90% in all pH range was determined. Therefore, the pH 6.5
was selected because it favors the adsorption system MB-AC-3. All
samples were ltered prior to analysis (using 0.45-m membrane
lters) in order to minimize the interference of small particles of the
activated carbon. The MB concentrations in the supernatant solutions before and after adsorption were determined using a UVvis
spectrophotometer (Varian Cary 50 UV/Vis) at its maximum wavelength () of 664 nm. The MB concentration was determined by
comparing absorbance to a calibration curve previously obtained.
All experiments were duplicated and only the mean values were
reported. The amount of MB adsorbed onto AC, qe (mg g1 ) was
calculated by Eq. (2):
qe =

(C0 Ce )V
W

(2)

where C0 and Ce (mg L1 ) are the initial and equilibrium liquidphase concentrations of MB, respectively, V (L) is the volume of the
solution, and W (g) is the mass of dry adsorbent used. For batch
kinetic studies, the same procedure was followed, but the aqueous
samples were taken at preset time intervals. The concentrations of
MB were similarly measured. The amount of MB adsorbed at any
time, qt (mg g1 ), was calculated by Eq. (3):
qt =

(C0 Ct )V
W

(3)

where Ct (mg L1 ) is the liquid-phase concentration of MB at any


time. Initial concentrations of 800, 900, and 1000 mg L1 of the dye
and an adsorption time of 150 min were studied.
2.4. Adsorption isotherm and kinetic models
The application of adsorption isotherms is very useful to
describe the interaction between the adsorbate and the adsorbent
of any system. The parameters obtained from the different models provide important information on the sorption mechanisms
and the surface properties and afnities of the adsorbents. There
are several equations for analyzing experimental adsorption equilibrium data. The Langmuir and Freundlich models are the most
accepted surface adsorption model for single solute systems. On
the other hand, an interesting trend in the isotherm modeling is
the derivation in more than one approach, thus directing to the difference in the physical interpretation. In this study, the isotherms of

A.L. Cazetta et al. / Chemical Engineering Journal 174 (2011) 117125

-1

Langmuir

qe =

Qm Ka Ce
1+Ka Ce

Freundlich

RL =

1
1+Ka Ce

qe = KF Ce1/nF
RedlichPeterson

1+BRP Ce

Qmax bT Cf

Toth

Q =

Pseudo-rst order

qt = qe [1 eK1 t ]

[1+(bT Cf )1/nT ]

nT

h0 = K1 qe
Pseudo-second order

qt =
h0 =

K2 q2
t
e

700
600
AC-2

500
400
300
200

AC-1

100
0
0.2

K2 q2e

0.4

0.6

0.8

1.0

Relative Pressure (P/P )

qt = kid t0.5 + C

Intraparticle diffusion

two parameters (Langmuir and Freundlich [25]) and three parameters (RedlichPeterson [26] and Toth [27]) were applied. Table 1
shows the equations and parameters of such isotherms.
The kinetic models such as pseudo-rst order [28], pseudosecond order [29] and intraparticle diffusion model [25] were used
to understand the adsorption dynamics in relation to time for the
MB-AC-3 system. The equation and parameters of these models are
shown in Table 1.
Both adsorption isotherms and kinetic models of pseudo-rst
and pseudo-second order were tted employing the non-linear tting method, using the software Origin 6.0. The theoretical models
most appropriate that describe the experimental data of MB-AC-3
system were chosen from the correlation coefcient (R2 ). Additionally, the normalized standard deviation, qe (%), was calculated for
kinetic studies by Eq. (4).
qe (%) = 100

AC-3

800

1+K2 qe t

Qm = maximum adsorption capacity; Ka = Langmuir constant; RL = separation factor;


KF and nF = Freundlich constants; ARP , BRP and g = RedlichPeterson constants, bT
and nT = Toth constants, K1 and K2 = rate constants for the pseudo-rst and pseudosecond order adsorption, respectively; h0 = initial adsorption rate; kid = intraparticle
diffusion constant; C = intercept.

900

 [(qe,exp qe,cal )/qe,exp ]2


N1

(4)

where N is the number of data points, qe,exp and qe,cal (mg g1 ) are
the experimental and calculated equilibrium adsorption capacity
value, respectively.
3. Results and discussions
3.1. Yield
The yield values based on the original weight of the raw material
were of 33.80% for the char and of 28.94%, 23.26% and 18.80% for
the AC-1, AC-2 and AC-3, respectively. According to the results, the
increase in the impregnation ratio (NaOH:char) caused a decrease
in the yield values. The decrease in the yield for ACs is justied by
action of the dehydrating reagent (NaOH), which provided elimination and dehydration reactions, breaking the bonds COC and
CC of the raw material [22]. Tseng and co-workers [7] report that
the activation mechanism with NaOH is according to the following
reaction:
6NaOH + 2C  2Na + 2Na2 CO3 + 3H2

(5)

3.2. Textural characterization


Fig. 1(a) shows the N2 adsorption isotherms obtained for AC-1,
AC-2 and AC-3. According to the gure, a progressive increase in

0.9

AC-1
AC-2
AC-3

0.8
0.7
0.6

Kinetic

qe =

ARP Ce

1000

0.5

Isotherm

Expression

Names

P / n (P - P)

Models

a
Adsorbed Volume at STP (cm g )

Table 1
Non-linear forms of kinetic and isotherm models, and intraparticle diffusion model.

119

0.4
0.3
0.2
0.1
0.03

0.06

0.09

0.12

0.15

0.18

0.21

Relative Pressure (P/P )


Fig. 1. N2 adsorption (closed symbols) and desorption (open symbols) isotherms at
77 K (a) and linear ts for the adsorption isotherms (b) of the AC-1, AC-2 and AC-3.

the N2 volume was observed for whole range of relative pressure,


and that the higher initial volume was shown for AC-3. According to
IUPAC, the isotherms that show reversibility are classied as type I,
which also are known as Langmuir. This type of isotherm indicates
that the adsorbate and adsorbent have a high afnity, and that the
material in question consists mostly of micropores [30].
The BET equation requires a linear t that relates P/na (P0 P)
and P/P0 , where na is the amount in moles adsorbed at the relative
pressure P/P0 . Fig. 1(b) shows the linear ts for the isotherms of AC1, AC-2 and AC-3. The values of correlation coefcients obtained, R2 ,
were 0.9999, 0.9992 and 0.9998 for AC-1, AC-2 and AC-3, respectively.
The pore size distributions of the prepared ACs are shown in
Fig. 2. An increase in pore volume was observed for the ACs. Addi
tionally, the most of the pores had sizes smaller than 2.5 nm (25 A),
indicating a development of microporosity of the material
Table 2 lists the physical properties of activated carbon derived
from coconut shell by NaOH activation. The data show that the
SBET of the ACs increased from 783 m2 g1 to 2825 m2 g1 when the
NaOH:char ratio increased from 1 to 3. The differences between the
SBET values of AC-1 and AC-2, AC-2 and AC-3 were of approximately
1000 m2 g1 , showing a linear relationship of SBET with increasing
of the NaOH:char ratio. As can be observed, the increase in the
NaOH:char promotes an increase of the SBET and a decrease in the
AC yields. Researches, which used the r wood [17] and Siberian
anthracite [31] as a raw material, also showed that the increase of
the NaOH:char promoted and increase in the SBET value.
As can be observed in Table 2, as well as the value of SBET , the values of VT , V and Vm increased with the NaOH:char. VT is directly

120

A.L. Cazetta et al. / Chemical Engineering Journal 174 (2011) 117125

Table 2
Textural characteristics of the activated carbons in the NaOH:char ratio of 1:1 (AC-1), 2:1 (AC-2) and 3:1 (AC-3).

AC-1
AC-2
AC-3

SBET (m2 g1 )

VT (cm3 g1 )

V (cm3 g1 )

Vm (cm3 g1 )

V /VT (%)

Dp (nm)

Yield (%)

783
1842
2825

0.378
0.927
1.498

0.356
0.775
1.143

0.022
0.152
0.355

94.2
83.6
76.3

1.63
1.80
2.27

28.9
23.4
18.8

SBET = BET surface area; VT = total pore volume; V = micropore volume; Vm = mesopore volume; V /VT = percentage of micropores, Dp = average pore diameter.

-1

Incremental Pore Volume (cm g )

0.024
AC-1
AC-2
AC-3

0.021
0.018
0.015
0.012
0.009
0.006
0.003
0.000
15

20

25

30

35 40 45
Diameter ()

50

55

60

Fig. 2. Pore size distribution of AC-1, AC-2 and AC-3.

related to the development of porosity of the material. The activation process rearranges the carbon structure producing a more
ordered structural skeleton. The pore development occurs from
four stages: (i) opening of previously inaccessibly pores, (ii) creation of new pores, (iii) widening of the existing pores and (iv)
merger of the existing pores due to pore wall breakage [32].
The values of Vm and Dp for AC-1, AC-2 and AC-3 were
0.022 cm3 g1 , 0.152 cm3 g1 and 0.355 cm3 g1 and 1.63 nm,
1.80 nm and 2.27 nm, respectively (Table 2). According to the
results, there was a tendency of increase of Vm and Dp with the
increase of NaOH:char. This tendency can be justied by the micropore merge and collapse, which occurred when large amounts of
NaOH were used [3]. The micropores were the responsible for
the increase of VT . In addition, the increase of NaOH:char promoted a development of mesopore and consequently a decrease
in the micropore percentages (V /Vt (%)). The values of V /Vt (%)
were 94.2, 83.6 and 76.3 for the AC-1, AC-2 and AC-3, respectively.
The presence of mesopores and micropores in AC enhances their
adsorption capacities, especially for large molecules of adsorbates
as dye molecules [33].
Fig. 3(ac) shows the SEM images for the ACs. The activation
method with NaOH provided the development of many pores with
varying sizes and materials of high surface area.
3.3. Chemical surface characterization
The method described by Boehm [23] was used in chemical
surface characterization of the prepared ACs. The methodology
consists of a series of titrations in which it is possible to quantify chemical groups such as carboxyls, lactones and phenolics. The
Boehm method results are shown in Table 3.

Fig. 3. SEM images of the AC-1, AC-2 and AC-3.

Table 3
Results of the Boehm and pH drift methods for the AC-1, AC-2 and AC-3.

AC-1
AC-2
AC-3

Carboxylic (mmol g1 )

Lactonic (mmol g1 )

Phenolic (mmol g1 )

Acid (mmol g1 )

Basic (mmol g1 )

Total (mmol g1 )

pH drift

0.37
0.62
0.75

0
0
0

0.38
0.88
1.00

0.75
1.5
1.75

0.73
0.75
0.75

1.47
2.25
2.50

6.00
5.09
5.01

A.L. Cazetta et al. / Chemical Engineering Journal 174 (2011) 117125

32

121

20.0

31

19.5
Coconut shell

Transmitance (%)

Transmitance (%)

30
29
28
27

AC-3
19.0
18.5
AC-1
18.0

26

4000

AC-2

17.5

25
3500

3000

2500

2000

1500

1000

500

4000

3500

3000

-1

2500

2000

1500

1000

500

-1

Wavenumber (cm )

Wavenumber (cm )

Fig. 4. FT-IR spectra for the raw material (a) and ACs (b).

As can be observed in Table 3, there was a gradual increase in


the amount of acid groups (carboxylic and phenolic) for ACs, which
consequently caused an increase in the amount of total groups. The
lactonic groups have not been detected and basic groups had values
of approximately 0.75 mmol g1 for prepared ACs. Additionally, it
can be seen that acid characteristics are more evident for AC-2 and
AC-3.
The use of chemical reagents in the activation process provides
an increase in the amount of acid groups present in the ACs surface,
as observed in other studies [18,34]. On the other hand, the basic
character is presented for materials produced from thermal treatment. The basicity is due to Lewis basic sites, oxygen free, on the
graphene layer, and from some surface groups containing oxygen
such as carbonyls, pyrone, and chromene type structures [35].
The pH drift method provides important information about the
characteristics of acidity and basicity of activated carbons, as the
pHpzc value, which has helped in the understanding of the adsorption mechanism. According to Table 3, the pHpzc value of 6.0 for the
AC-1 was lower than for AC-2 (5.09) and AC-3 (5.01), respectively.
The acid characteristic of the ACs agrees with the results obtained
by Boehm method and by other researchers which produced ACs
from a chemical activation [36,37].
The FT-IR spectra were obtained to evaluate qualitatively the
chemical structures of the raw material and ACs. The spectra are
shown in Fig. 4. Fig. 4(a) shows the FT-IR spectrum of the raw
material, which indicated various surface functional groups. The
broad band at around 3404 cm1 is typically attributed to hydroxyl
groups. The band located at around 2900 cm1 corresponds to
CH stretching vibration. The region of the spectrum of 1612 cm1
is attributed to axial deformation of carbonyl groups (C O). The
stretching vibration of the molecular plane of C C bonds, characteristics of aromatic rings appear in the region of 1465 cm1
[38]. Axial and angular deformation of ketones arises in the region
of 1300 cm1 and 1100 cm1 . The broad band at 1058 cm1 is
attributed to the angular deformation symmetrical of ethers [39].
The band caused by OH out-of-plane bending vibrations band is
located at 580 cm1 [40]. The presence of hydroxyl groups, carbonyl group, ethers and aromatic compounds is an evidence of the
lignocellulosic structure of coconut shell as also observed in others materials such as Tunisian olive-waste cakes [41], jackfruit peel
waste [24], and cotton stalks [5].
The FT-IR spectra of the ACs obtained are shown in Fig. 4(b).
According to the results, it can be observed that there was disappearance of bands when comparing the raw material spectrum
with the ACs spectra, indicating that the chemical bonds were
broken during the carbonization process followed by the activation. In addition, other bands decreased drastically, indicating a

decrease in the functionality of the raw material. Montes-Morn


et al. [42] reported that the bands which were observed in the
region between 1700 cm1 and 1500 cm1 are attributed to C C
symmetrical stretching of pyrone groups and C O of carboxylic
groups. The spectra shown in Fig. 4(b) are similar to those reported
by Vargas et al. [18], which produced activated carbons from amboyant pods.
3.4. Adsorption isotherms
The adsorption isotherms describe how the adsorbate molecules
are distributed between the liquid phase and solid phase when the
system reaches the equilibrium [43]. The analysis of isotherm data
by tting them to different models is important to nd a sustainable
model that can be used [44].
In order to optimize the design of an adsorption system of
MB-AC-3, four adsorption isotherm models namely the Langmuir,
Freundlich, RedlichPeterson and Toth isotherms in their nonlinear forms, were applied to the equilibrium data, and are shown
in Fig. 5. The Langmuir and Freundlich isotherms are known as
two-parameter models, which provide information on the adsorption capacity and constants related to the activation energy. The
RedlichPeterson isotherm is known as three-parameter model,
which incorporates the factors of Langmuir and Freundlich equation in a single. The Toth isotherm is of three-parameters and was
developed to explain adsorption process in heterogeneous systems; it resembles the empirical proposal, which was developed
by Freundlich to explain the adsorption phenomenon of these systems [27]. The values of the maximum adsorption capacity (Qm ),
correlation coefcient (R2 ), and other constants obtained for the
models from experimental data are shown in Table 4.
The Langmuir model assumes that the adsorption is a process
which occurs in a homogeneous surface, in which the molecules
form a monolayer of adsorbate on the surface of the material, saturating the pores and preventing the transmigration [45].
According to the results presented in Table 4, the Qm and R2 values obtained from Langmuir isotherms were of 916.26 mg g1 and
0.8805, respectively. The Qm is close to the experimental value of
Qm that is equal to 955.25 mg g1 . The separation factor (RL ) dened
by Weber and Chakkravorti [46] is an important parameter of the
Langmuir isotherm that can be used to verify if the adsorption in the
system is unfavorable (RL > 1), linear (RL = 1), favorable (0 < RL < 1)
or irreversible (RL = 0). For the value range of concentrations studied (1001000 mg L1 ), the RL values decreased from 5.52 103
to 5.55 105 , indicating that the MB-AC-3 system is favorable.
The Freundlich model is an empirical equation based on the
adsorption on heterogeneous surface [47]. The heterogeneity fac-

122

A.L. Cazetta et al. / Chemical Engineering Journal 174 (2011) 117125

Table 4
Langmuir, Freundlich, RedlichPeterson and Toth isotherm model parameters and correlation coefcients for adsorption of MB on activated carbon.
Langmuir

Freundlich

Toth

RedlichPeterson

Qm = 916.26 mg g1
Ka = 1.80 L mg1
R2 = 0.8805

KF = 747.32 mg g1
nF = 20.08
R2 = 0.8796

Qm = 720.81 mg g1
bT = 0.031
nT = 0.95
R2 = 0.8796

ARP = 24.47 103 (L mg1 )g


BRP = 32.35 L g1
g = 0.95
R2 = 0.8796

a 1000

data t preferentially the Langmuir isotherm. The R2 value for the


Freundlich model was 0.8796.
The Toth model is a combination of the characteristics of
the Langmuir and Freundlich isotherms [45]. The Toth equation
(Table 1) reduces to Langmuir for nT = 1. In the present study,
nT = 0.95 is close to the unit value, which reduces the Langmuir
equation. Additionally, the model provided values of Qm and R2
equals to 720.32 mg g1 and 0.8796, respectively.
RedlichPeterson model is used as a compromise between Langmuir and Freundlich models [49]. The parameter g in the equation
indicates which of the two models, Freundlich or Langmuir, better ts the experimental data. For g = 1, the Langmuir isotherm
is favored, and g = 0 the equation of Freundlich is favored. In this
study, the g value is close to unity, which means that the isotherms
conform to Langmuir model better than Freundlich model.
Comparing the R2 values (Table 4), it can be observed that the
Langmuir model yielded the best t with the highest R2 value. The
suitability of the Langmuir model to experimental data was conrmed by the constant RedlichPeterson model and the value 1/n
described by the Freundlich model. The t of the experimental data
to the Langmuir model indicates a homogeneous nature of the AC3 surface. Additionally, it describes the formation of monolayer
coverage of dye molecule at the outer surface of the AC-3 [50].
The Qm value of 916.26 mg g1 describes the high adsorption
capacity of the AC-3, which is directly related to high surface are
(SBET ) and the average pore diameter (Dp ). Considering that width of
the MB molecules is 1.42 nm [25], and the Dp of AC-3 is 2.27 nm, the
MB molecules can diffuse from solution into the AC, which justies
the high value of Qm . In addition, the adsorption is favored due to
the acid characteristics of the AC-3 surface, which was described by
Boehm and pH drift methods. The negative surface of AC-3 interacts
effectively with the MB, due to cationic property of the dye.
Table 5 lists a comparison of the surface area (SBET ) and maximum adsorption capacity (Qm ) of ACs from coconut shell produced
from various activation methods. The NaOH-activated carbon prepared in this work presented a high surface area (SBET ), which was
proven by the high adsorption capacity of the MB dye.

900
800
Experimental values
Toth
Langmuir

600

-1

qe (mg g )

700
500
400
300
200
100
0
0

10

20

30

40

50

60

70

80

70

80

-1

Ce (mg L )

b 1000
900
800
Experimental values
Freundlich
Redlich-Peterson

600

-1

qe (mg g )

700
500
400
300
200
100
0
0

10

20

30

40

50

60

-1

Ce (mg L )
Fig. 5. Non-linear ts of the isotherm models. Langmuir and Toth (a),
RedlichPeterson and Freundlich (b).

3.5. Kinetic studies

tor (nF ) indicates whether the adsorption process is linear (nF = 1),
chemical (nF < 1) and or physical (nF > 1). Additionally, the value
of 1/nF < 1 indicates a normal Langmuir isotherm while 1/nF > 1
is an indicative of cooperative adsorption [48]. According to the
results shown in Table 4, the values of nF = 20.08 and 1/nF = 0.050
indicate that the adsorption is physical, and that the experimental

Kinetic studies are important to understand the dynamic of the


reaction in terms of order of the rate constant. Since the kinetics
parameters provide information for designing and modeling the
adsorption process. The data of adsorption kinetics for the AC-3-MB
system were analyzed by non-linear tting of two different kinetic
models: pseudo-rst order and pseudo-second order, which are

Table 5
The BET surface area (SBET ) and maximum adsorption capacity values (Qm ) of coconut activated carbons produced from various activation methods.
Activation methods

SBET (m2 g1 )

Adsorbate

Qm (mg g1 )

Reference

H3 PO4
KOH/CO2
KOH/CO2
KOH/CO2
ZnCl2 /steam
ZnCl2 /CO2
Steam
NaOH

902

1940
1026
2114
1884

2885

Zinc
2,4,6-Trichlorophenol
Methylene blue
Phenol
Methane
Methylene blue
Methylene blue
Methylene blue

60.41
191.73
434.78
205.8
122.85
14.36
277.90
916.26

[15]
[13]
[50]
[12]
[51]
[52]
[53]
This work

A.L. Cazetta et al. / Chemical Engineering Journal 174 (2011) 117125

1000

1000

800

800
600
qt (mg g )

600
400

Experimental values
Pseudo-first order
Pseudo-second order

-1

Experimental values
Pseudo-first order
Pseudo-second order

-1

qt (mg g )

123

200

400
200

0
0

20

40

60

80

100

120

140

160

20

40

Time (min)

60

80

100

120

140

160

Time (min)

c 1000
800
Experimental values
Pseudo-first order
Pseudo-second order

-1

qt (mg g )

600
400
200
0
0

20

40

60 80 100
Time (min)

120

140

160

Fig. 6. Non-linear ts of pseudo-rst order and pseudo-second order kinetics for the MB concentration of 800 (a), 900 (b) and 1000 (c) mg L1 .

shown in Fig. 6. According to Fig. 6, the AC-3-MB system reached


the equilibrium after 30 min, suggesting that interactions between
the adsorbent and adsorbate were favorable.
Table 6 shows the parameters obtained from the ts of the
adsorption kinetic models. The correlation coefcients (R2 ) for
all models were greater than 0.9705 in different concentrations,
being that the highest values were observed for the pseudo-second
kinetic model. The applicability of the pseudo-second-order kinetic
model was conrmed by the low value of normalized standard
deviation (qe = 2.59%). Additionally, the qe values obtained by the
tting agreed with qe,exp values. This suggests that the rate of the

adsorption process is preferably controlled by chemisorption; as


also reported by Tan et al. [50] which studied the methylene blue
adsorption onto KOH/CO2 -activated carbon from coconut.
To identify the diffusion mechanism, the intraparticle diffusion
model based on the theory proposed by Weber and Morris [54] was
applied to the AC-3-MB system.
Fig. 7 shows the plots of qt versus t1/2 . The value of the slope corresponds to intraparticle diffusion constant, kdi , and the intercept
value, Ci , at an approximate value of the thickness of boundary layer. The data from the three different initial concentrations
showed two stages of linearity, being the rst stage was completed

Table 6
Pseudo-rst order and pseudo-second order kinetic model parameters for different initial MB concentrations.
C0 (mg L1 )

qe,exp (mg g1 )

Pseudo-rst order
1

Pseudo-second order

800

795.15

qe = 774.54 (mg g )
K1 = 0.43 (min1 )
h0 = 333.05 (mg g1 min1 )
R2 = 0.9909

qe = 802.14 (mg g1 )
K2 = 0.00097 (g g1 min1 )
h0 = 624.13 ((mg g1 ) min1 )
R2 = 0.9998

900

868.50

qe = 826.39 (mg g1 )
K1 = 0.50 (min1 )
h0 = 413.20 (mg g1 min1 )
R2 = 0.9745

qe = 856.64 (mg g1 )
K2 = 0.00104 (g g1 min1 )
h0 = 763.18 ((mg g1 ) min1 )
R2 = 0.9932

1000

911.30

qe = 867.96 (mg g1 )
K1 = 0.47 (min1 )
h0 = 407.94 (mg g1 min1 )
R2 = 0.9758
qe (%) = 5.79

qe = 900.69 (mg g1 )
K2 = 0.00092 (g g1 min1 )
h0 = 746.34 ((mg g1 ) min1 )
R2 = 0.9959
qe (%) = 2.59

124

A.L. Cazetta et al. / Chemical Engineering Journal 174 (2011) 117125

Table 7
Intraparticle diffusion model constants and correlation coefcients for adsorption MB on prepared activated carbon.
C0 (mg L1 )

Intraparticle diffusion model

800
900
1000

kdi1 (mg g1 min1/2 )

kdi2 (mg g1 min1/2 )

C1 (mg g1 )

C2 (mg g1 )

R12

R22

119.09
89.07
104.57

4.97
12.59
10.08

345.14
471.86
463.74

739.37
730.79
795.71

0.9865
0.9834
0.9995

0.9087
0.9070
0.9112

Acknowledgement

1000
900

The authors acknowledge Fundaco Araucria and CAPES for the


nancial support.

800
700
-1

qt (mg g )

600

-1

References

800 mg L

500

-1

900 mg L
-1
1000 mg L

400
300
200
100
0
0

6
1/2

9 10 11 12 13

1/2

Time (min )
Fig. 7. Intraparticle diffusion plots for the adsorption at different initial MB concentrations.

within the rst 10 min. The rst stage is the instantaneous adsorption, and the second region is the gradual adsorption stage where
the intraparticle diffusion is the rate limiting. Both the linear line
did not pass through the origin, this suggests that the intraparticle
diffusion was not the only limiting mechanism in the adsorption
process. Table 7 shows the values of kdi , Ci and correlation coefcient R2 obtained for the plots. The (R2 )2 values are lower than
predicted by the pseudo-second order model, so the qexp value did
not agree well with the intraparticle diffusion model

4. Conclusion
The results of this study showed that the NaOH-ACs obtained
from coconut shell presented good development and high BET surface area. The ACs properties were related with NaOH:char ratio.
The textural characterization showed that the ACs are essentially
microporous and that BET surface area of 2825 m2 g1 was obtained
for AC-3. The well-developed porous structure was conrmed by
SEM analysis. The Boehm and pH drift methods and FT-IR analysis
showed that the ACs has acid characteristics and carboxylic groups,
phenolic groups and pyrone groups in its structure.
The adsorption potential of the AC-3 produced for the removal
of MB from the aqueous solution over a wide concentration range
was investigated. The equilibrium data were tted for the nonlinear models of Langmuir, Freundlich, Toth and RedlichPeterson,
being that the Langmuir model was which better tted to
the experimental data, showing a maximum adsorption capacity (Qm ) of 916.26 mg g1 . The kinetic data were tested for the
models of pseudo-rst order, pseudo-second order and intraparticle diffusion. The adsorption process follows the pseudo-second
order model, which suggests that the process was controlled by
chemisorption. Due to their chemical and textural characteristics,
the coconut shell activated carbon obtained in this work is potential
material in the water and wastewater treatment for the removal of
some organic pollutants.

[1] E. Demirbas, M. Kobya, M.T. Sulak, Adsorption kinetics of a basic dye from aqueous solutions onto apricot stone activated carbon, Bioresour. Technol. 99 (2008)
53685373.
[2] G. Crini, Non-conventional low-cost adsorbents for dye removal: a review,
Bioresour. Technol. 97 (2006) 10611085.
[3] C. Michailof, G.G. Stavropoulos, C. Panayiotou, Enhanced adsorption of phenolic
compounds, commonly encountered in olive mill wastewaters, on olive husk
derived activated carbons, Bioresour. Technol. 99 (2008) 64006408.
[4] J.V. Nabais, P. Carrott, M.M.L.R. Carrott, V. Luz, A.L. Ortiz, Inuence of preparation conditions in the textural and chemical properties of activated carbons
from a novel biomass precursor: the coffee endocarp, Bioresour. Technol. 99
(2008) 72247231.
[5] A.A. El-Hendawy, A.J. Alexander, R.J. Andrews, G. Forrest, Effects of activation
schemes on porous, surface and thermal properties of activated carbons prepared from cotton stalks, J. Anal. Appl. Pyrolysis 82 (2008) 272278.
[6] R.L. Tseng, Physical and chemical properties and adsorption type of activated
carbon prepared from plum kernels by NaOH activation, J. Hazard. Mater. 147
(2007) 10201027.
[7] F.C. Wu, R.L. Tseng, R.S. Juangm, Preparation of highly microporous carbons
from r wood by KOH activation for adsorption of dyes and phenols from water,
Sep. Purif. Technol. 47 (2005) 1019.
[8] F.C. Wu, R.L. Tseng, R.S. Juang, Comparisons of porous and adsorption properties
of carbons activated by steam and KOH, J. Colloid Interface Sci. 283 (2005)
4956.
[9] R. Ubago-Prez, F. Carrasco-Marn, D. Fairn-Jimnez, C. Moreno-Castilla, Granular and monolithic activated carbons from KOH-activation of olive stones,
Microporous Mesoporous Mater. 92 (2006) 6470.
[10] B.H. Hameed, A.T.M. Din, A.L. Ahmad, Adsorption of methylene blue onto
bamboo-based activated carbon: kinetics and equilibrium studies, J. Hazard.
Mater. 141 (2007) 819825.
[11] A.M.M. Vargas, C.A. Garcia, E.M. Reis, E. Lenzi, W.F. Costa, V.C. Almeida,
NaOH-activated carbon from amboyant (Delonix regia) pods: optimization of
preparation conditions using central composite rotatable design, Chem. Eng. J.
162 (2010) 4350.
[12] A.Z.M. Din, B.H. Hameed, A.L. Ahmad, Batch adsorption of phenol onto
physiochemical-activated coconut shell, J. Hazard. Mater. 161 (2008) 6572.
[13] I.A.W. Tan, A.L. Ahmad, B.H. Hameed, Preparation of activated carbon from
coconut husk: optimization study on removal of 2,4,6-trichlorophenol using
response surface methodology, J. Hazard. Mater. 153 (2008) 709717.
[14] Z. Hu, M.P. Srinivasan, Preparation of high-surface-area activated carbons from
coconut shell, Microporous Mesoporous Mater. 27 (1999) 1118.
[15] O.S. Amuda, A.A. Giwa, I.A. Bello, Removal of heavy metal from industrial
wastewater using modied activated coconut shell carbon, Biochem. Eng. J.
36 (2007) 174181.
[16] J.H. Tay, X.G. Chen, S. Jeyaseelan, N. Graham, Optimising the preparation of
activated carbon from digested sewage sludge and coconut husk, Chemosphere
44 (2001) 4551.
[17] F.C. Wu, R.L. Tseng, High adsorption capacity NaOH-activated carbon for
dye removal from aqueous solution, J. Hazard. Mater. 152 (2008) 1256
1267.
[18] A.M.M. Vargas, A.L. Cazetta, C.A. Garcia, J.C.G. Moraes, E.M. Nogami, E. Lenzi,
W.F. Costa, V.C. Almeida, Preparation and characterization of activated carbon
from a new raw lignocellulosic material: amboyant (Delonix regia) pods, J.
Environ. Manage. 92 (2011) 178184.
[19] R.L. Tseng, Mesopore control of high surface area NaOH-activated carbon, J.
Colloid Interface Sci. 303 (2006) 494502.
[20] S. Senthilkumaar, P.R. Varadarajan, K. Porkodi, C.V. Subbhuraam, Adsorption
of methylene blue onto jute ber carbon: kinetics and equilibrium studies, J.
Colloid Interface Sci. 284 (2005) 7882.
[21] ASTM-D1762-84, Ann. Book ASTM Stand. D1762-84 (1984) 292293.
[22] A.H. Basta, V. Fierro, H. El-Saied, A. Celzard, 2-Steps KOH activation of rice
straw: an efcient method for preparing high-performance activated carbons,
Bioresour. Technol. 100 (2009) 39413947.
[23] H.P. Boehm, Some aspects of the surface chemistry of carbon blacks and other
carbons, Carbon 32 (1994) 759769.

A.L. Cazetta et al. / Chemical Engineering Journal 174 (2011) 117125


[24] D. Prahas, Y. Kartika, N. Indraswati, S. Ismadji, Activated carbon from jackfruit
peel waste by H3 PO4 chemical activation: pore structure and surface chemistry
characterization, Chem. Eng. J. 140 (2008) 3242.
[25] B. Royer, N.F. Cardoso, E.C. Lima, J.C.P. Vaghetti, N.M. Simon, T. Calvete, R.C.
Veses, Applications of Brazilian pine-fruit shell in natural and carbonized forms
as adsorbents to removal of methylene blue from aqueous solutionskinetic
and equilibrium study, J. Hazard. Mater. 164 (2009) 12131222.
[26] J.C.P. Vaghetti, E.C. Lima, B. Royer, B.M. Cunha, N.F. Cardoso, J.L. Brasil, S.L.P. Dias,
Pecan nutshell as biosorbent to remove Cu(II), Mn(II) and Pb(II) from aqueous
solutions, J. Hazard. Mater. 162 (2009) 270280.
[27] K. Vijayaraghavan, T.V.N. Padmesh, K. Palanivelu, M. Velan, Biosorption of
nickel(II) ions onto Sargassum wightii: application of two-parameter and threeparameter isotherm models, J. Hazard. Mater. B133 (2006) 304308.
[28] G.F. Malash, M.I. El-Khaiary, Methylene blue adsorption by the waste of AbuTartour phosphate rock, J. Colloid Interface Sci. 348 (2010) 537545.
[29] R. Han, J. Zhang, P. Han, Y. Wang, Z. Zhao, M. Tang, Study of equilibrium, kinetic
and thermodynamic parameters about methylene blue adsorption onto natural
zeolite, Chem. Eng. J. 145 (2009) 496504.
[30] S.S. Brum, M.L. Bianch, V.L. Silva, M. Goncalves, M.C. Guerreiro, L.C.A. Oliveira,
Preparation and characterization of activated carbon produced from coffee
waste, Quim. Nova 5 (2008) 10481052.
[31] P. Nowicki, R. Pietrzak, H. Wachowska, Siberian anthracite as a precursor material for microporous activated carbons, Fuel 87 (2008) 20372040.
[32] K. Yang, J. Peng, C. Srinivasakannan, L. Zhang, H. Xia, X. Duan, Preparation of high
surface area activated carbon from coconut shells using microwave heating,
Bioresour. Technol. 101 (2010) 61636169.
[33] S.B. Hartono, S. Ismadji, Y. Sudaryanto, W. Irawaty, Utilization of teak sawdust
from the timber industry as a precursor of activated carbon for the removal of
dyes from synthetic efuents, J. Ind. Eng. Chem. 11 (2005) 864.
[34] Y.M. El-Sayed, T.J. Bandosz, Adsorption of valeric acid from aqueous solution
onto activated carbons: role of surface basic sites, J. Colloid Interface Sci. 273
(2004) 6472.
[35] M.F.R. Pereira, S.F. Soares, J.J.M. rfo, J.L. Figueiredo, Adsorption of dyes on
activated carbons: inuence of surface chemical groups, Carbon 41 (2003)
811821.
[36] A.M.M. Vargas, A.L. Cazetta, M.H. Kunita, T.L. Silva, V.C. Almeida, Adsorption of
methylene blue on activated carbon produced from amboyant pods (Delonix
regia): study of adsorption isotherms and kinetic models, Chem. Eng. J. 168
(2011) 722730.
[37] B.H. Hameed, B.K. Mahmound, A.L. Ahmad, Equilibrium modeling and kinetic
studies on the adsorption of basic dye by a low-coast adsorbent: coconut (Cocos
nucifera) bunch waste, J. Hazard. Mater. 158 (2008) 6572.
[38] S.F. Dyke, A.J. Floyd, M. Sainsbury, R.S. Theobald, Organic Spectroscopy: An
Introduction, 1st ed., Penguin Books, Ringwood, Victoria, Australia, 1971.

125

[39] R.M. Silverstein, F.X. Webster, D.J. Kiemle, Spectrometric Identication of


Organic Compounds, 7th ed., John Wiley & Sons, New York, 2005.
[40] J. Yang, K. Qiu, Preparation of activated carbon from walnut shells via vacuum
chemical activated and their application for methylene blue removal, Chem.
Eng. J. 165 (2010) 209217.
[41] R. Baccar, J. Bouzid, M. Feki, A. Montiel, Preparation of activated carbon from
Tunisian olive-waste cakes and its application for adsorption of heavy metal
ions, J. Hazard. Mater. 162 (2009) 15221529.
[42] M.A. Montes-Morn, D. Surez, J.A. Menndez, E. Fuente, On the nature of basic
sites on carbon surfaces: an overview, Carbon 41 (2004) 12191225.
[43] B.H. Hameed, I.A.W. Tan, A.L. Ahmad, Adsorption isotherm, kinetic modeling
and mechanism of 2,4,6-trichlorophenol on coconut husk-based activated carbon, Chem. Eng. J. 144 (2008) 235244.
[44] M. El-Guendi, Homogeneous surface diffusion model of basic dyestuffs on natural clay in batch adsorbers, Adsorpt. Sci. Technol. 8 (2) (1991) 255271.
[45] P.S. Kumar, S. Ramalingam, C. Senthamarai, M. Niranjanaa, P. Vijayalakshmi,
S. Sivanesan, Adsorption of dye from aqueous solution by cashew nut shell:
studies on equilibrium isotherm, kinetics and thermodynamics of interactions,
Desalination 261 (2010) 5260.
[46] T.W. Weber, R.K. Chakkravorti, Pore and solid diffusion models for xed-bed
adsorbers, AiChE J. 20 (1974) 228238.
[47] H. Freundlich, Ueber die adsorption in loesungen, Z. Phys. Chem. 57A (1907)
385470.
[48] K. Fytianos, E. Voudrias, E. Kokkalis, Sorptiondesorption behavior of 2,4dchlorophenol by marine sediments, Chemosphere 40 (2000) 36.
[49] O. Redlich, D.L. Peterson, A useful adsorption isotherms, J. Phys. Chem. 63 (1959)
10241026.
[50] I.A.W. Tan, A.L. Ahmad, B.H. Hameed, Adsorption of basic dye on high-surfacearea activated carbon prepared from coconut husk: equilibrium, kinetic and
thermodynamic studies, J. Hazard. Mater. 154 (2008) 337346.
[51] D.C.S. Azevedo, J.C.S. Arajo, M. Bastos-Neto, A.E.B. Torres, E.F. Jaguaribe, C.L.
Cavalcante, Microporous activated carbon prepared from coconut shells using
chemical activation with zinc chloride, Microporous Mesoporous Mater. 100
(2007) 361364.
[52] J.S. Macedo, N.B. Costa Jnior, L.E. Almeida, E.F.S. Vieira, A.R. Cestari, I.F.
L.S. Barreto, Kinetic and calorimetric study of the
Gimezez, N.L.V. Carreno,
adsorption of dyes on mesoporous activated carbon prepared from coconut
coir dust, J. Colloid Interface Sci. 298 (2006) 515522.
[53] N. Kannan, M.M. Sundaram, Kinetics and mechanism of removal of methylene
blue by adsorption on various carbons a comparative study, Dyes Pigment 51
(2001) 2540.
[54] W.J. Weber, J.C. Morris, Proc. Int. Conf. Water pollution symposium vol.2, Pergamon, Oxford, 1962, pp. 231-266.

Anda mungkin juga menyukai