You are on page 1of 5

Journal of Industrial and Engineering Chemistry 18 (2012) 780–784

Contents lists available at SciVerse ScienceDirect

Journal of Industrial and Engineering Chemistry


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

Treatment of pollutants in wastewater: Adsorption of methylene blue onto


olive-based activated carbon
Mónica Berrios, Marı́a Ángeles Martı́n *, Antonio Martı́n
University of Cordoba (Spain), Department of Inorganic Chemistry and Chemical Engineering, Campus Universitario de Rabanales, Edificio Marie Curie (C3),
Planta Baja, 14071 Cordoba, Spain

A R T I C L E I N F O A B S T R A C T

Article history: This study used olive stone-based activated carbon for the removal of methylene blue from wastewater in
Received 16 May 2011 order to evaluate the adsorption capacity of the carbon. The equilibrium and kinetics of adsorption were
Accepted 5 August 2011 examined at 258, 308, 358 and 40 8C and several agitation speeds. Type III adsorption isotherms
Available online 12 November 2011
corresponding to physical adsorption in a multilayer system were used for the methylene blue system.
The equilibrium data for methylene blue adsorption showed a good fit to the Freundlich equation. The
Keywords: kinetic data was analysed to determine kinetic constants and order of reaction. Kinetics was evaluated by
Activated carbon
means of an n-order model, showing that the reaction was a first-order reaction. The results indicated
Adsorption isotherms
that olive stone-based activated carbon could be used as a low-cost alternative to commercial activated
Kinetics studies
Methylene blue carbon for the removal of organic compounds from wastewater. However, due to its microporosity, the
Olive stones application of this type of activated carbon was found to be suitable for molecules smaller than
methylene blue.
ß 2011 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights
reserved.

1. Introduction due to its low price and suitability for the removal of organic and
inorganic pollutants from wastewater [7]. Although, these
The textile industry requires large amounts of water and agricultural wastes can be also used as biosorbents directly [8–
produces highly polluted wastewater containing different types of 10]. In this sense, Nieto et al. [11] studied the ability of crude olive
dyes [1]. The main problem involved in decontaminating textile stones, a residue of the olive-oil industry, for the adsorption of iron
wastewaters is the removal of colour, since no single process is present in the industrial wastewaters. Researchers have studied
currently capable of generating adequate effluents [2]. Most dyes the production of activated carbon from palm-tree cobs, plum
have an adverse impact on the environment as they are considered kernels, cassava peel, bagasse, jute fiber, rice husks, date pits,
toxic and have carcinogenic properties, which make the water nutshells, wood, maize cob, cotton seed shell, rubber seed coat,
inhibitory to aquatic life [3]. apricot stone, almond shell, pongam seed coat, coconut shell,
Biological treatment processes are reported to be efficient in orange peel, walnut stone, bamboo dust, sunflower seed hull and
chemical oxygen demand reduction, but are largely ineffective in peach stone as has been detailed in the literature [6,7,12–18]. Little
removing colour from wastewater. Hence, research has been information has been reported about the particular case of
conducted on physico-chemical methods for colour removal in activated carbon from olive stones [19–21]. In Mediterranean
textile effluent. These studies include the use of coagulants, countries, olive stones and residues are a cheap and quite abundant
oxidising agents, photocatalysis, ultrafiltration, electrochemical agricultural waste [21].
and adsorption techniques [4,5]. Activated carbon has a porous structure with a large internal
Among the various treatment technologies available, adsorp- surface area. Four consecutive mass transport steps are associated
tion onto activated carbon has proven to be one of the most with the adsorption of solute from solution by porous adsorbent as
effective and reliable physico-chemical treatments. However, follows: the adsorbate migrates through the solution to the
commercially available activated carbons are very expensive [6]. exterior surface of the adsorbent particles, molecular diffusion
The carbon derived from agricultural wastes is gaining importance takes place in the boundary layer, solute is moved from the particle
surface into the interior site by pore diffusion and finally the
adsorbate is adsorbed into the active sites at the interior of the
* Corresponding author. Tel.: +34 957 21 86 24; fax: +34 957 21 86 25. adsorbent particle. This phenomenon takes relatively long contact
E-mail address: iq2masam@uco.es (M.n). time [6,21].

1226-086X/$ – see front matter ß 2011 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights reserved.
doi:10.1016/j.jiec.2011.11.125
M. Berrios et al. / Journal of Industrial and Engineering Chemistry 18 (2012) 780–784 781

Determinations of surface area can be made by fitting the BET 8000 rpm for 5 min and filtered prior to analysis in order to remove
equation to the isothermal equilibrium data obtained. However, suspended particles of AC.
these values are not a true indication of the adsorption capacity of
an activated carbon applied in liquid-phase adsorption studies. It is 2.4. Kinetic experiments
therefore more logical to determine the porous structure by
combining both the gas-phase and the liquid-phase adsorption The kinetic tests were carried out in a similar way to the
equilibrium data. The literature indicates that the adsorption of previous equilibrium tests. Several amounts of AC (1, 2, 4 and 8 g)
phenol, methylene blue, caffeine and iodine from the aqueous were added to the MB solutions (initial concentration at 5 mg/L)
phase is a useful tool for product control in the manufacture of and kept in an isothermal shaker SI 50 GIRALT (STUART SCIENTIFIC,
activated carbon [22–24]. UK) at different temperatures (258, 308, 358 and 40 8C) and
In addition to determining the porous structure of activated agitation speeds (50, 100, 150, 200 and 250 rpm). The aqueous
carbon, methylene blue can be employed as a thiazine (cationic samples (10 mL) were taken at time intervals of up to 95 min. The
or basic) dye; the most commonly used dye for colouring among samples were centrifuged at 8000 rpm for 5 min and filtered prior
all other dyes of its category. It is generally used for dyeing to analysis in order to remove suspended particles of AC.
cotton, wool, and silk [25] and has a number of biological uses. The amount of adsorption or adsorption capacity q (mg/g) was
However, given that methylene blue has various harmful effects calculated by
on human beings, it is of utmost importance to remove it from
C0V 0  CiV i
wastewater. Methylene blue dissociates in aqueous solution as q¼ (1)
W
electrolytes into methylene blue cation and the chloride ion.
Because the coloured cation is retained at great length by several where C0 and Ci (mg/L) are the liquid-phase concentrations of MB
adsorbents [3,26], methylene blue was selected as the adsorbate at initial and any time (Ct) or equilibrium (Ce), respectively,
in this study. obtaining the amount of adsorption at time (qt) or the amount of
This research study aimed to evaluate the adsorption potential adsorption at equilibrium (qe). V0 is the initial volume of the
of olive stone-based activated carbon for methylene blue from a solution (0.5 L), Vi is the real volume when sampling and W is the
synthetic wastewater as olive stones are a very abundant and mass of the AC used (g).
inexpensive material in Mediterranean countries. The kinetic and
equilibrium data of adsorption studies were processed to 3. Results and discussion
understand the adsorption behaviour of the dye molecules onto
the activated carbon. Although this activated carbon is commer- 3.1. Adsorption equilibrium
cial, it was selected because no data about this behaviour has yet
been reported. The adsorption of dyes from the liquid to solid phase can be
considered a reversible reaction with equilibrium established
2. Materials and methods between the two phases [28]. The adsorption isotherm (qe vs. Ce)
indicates how the adsorption molecules distribute between the
2.1. Materials liquid phase and the solid phase when the adsorption process
reaches an equilibrium state. The analysis of the equilibrium data of
Methylene blue (MB) supplied by Panreac (Spain) was used as the isotherm models is very important for the use of adsorbents [12].
an adsorbate and was not purified prior to use. The molecular Fig. 1 shows the adsorption isotherms at four temperatures for
weight (g/mol), molecular volume (cm3/mol) and molecular the MB solution and olive stone-based AC system. According to
diameter (nm) of MB are 319.85, 241.9 and 0.8, respectively [27]. Brunauer et al. [29] and Hinz [30], the adsorption isotherms have
Olive stone-based activated carbon (AC) was supplied by Ibérica the same shape that the type III isotherms or S1 isotherms (Giles
de Carbones Activos S.A. Textural characterization of the AC was classification), which correspond to physical adsorption in a
carried out by N2 adsorption at 77 K using Micromeritic ASAP 2020 multilayer system where no difference is noticed between the
in our laboratory. The BET surface area, total pore volume and filling of the first layer and the other layers.
average pore diameter of the AC were found to be 587 m2/g, The Freundlich isotherm is the earliest known relationship
0.333 cm3/g and 2.27 nm, respectively. describing the adsorption equation. This fairly satisfactory

2.2. Analysis of methylene blue concentration 1.0

C0 (MB) = 0.508 mg/L


The MB concentration in the supernatant solution before and
after adsorption was determined using a UV–vis spectrophotome- 0.8
ter S-20 (BOECO, Germany) at 664 nm. The calibration curve was
very reproducible and linear over the concentration range used in
this work. 0.6
qe (mg/g)

2.3. Equilibrium experiments


0.4

The equilibrium tests were performed in 4 Erlenmeyer flasks


25ºC
(1 L) in which 500 mL of MB solution (initial concentration at 30ºC
0.2
0.5 mg/L) were placed. The agitation speed was fixed at 50 rpm and 35ºC
the AC doses (0.25, 0.50, 1.00 and 2.00 g) were added to the MB 40ºC

solutions and kept in an isothermal shaker SI 50 GIRALT (STUART


0.0
SCIENTIFIC, UK) at different temperatures (258, 308, 358 and 40 8C). 0.06 0.08 0.10 0.12
The contact time to reach equilibrium between the solid phase and Ce (mg/L)
the liquid phase was approximately 11.5 h (that was checked in
previous assays). The aqueous samples (10 mL) were centrifuged at Fig. 1. Adsorption isotherms for the MB solution and olive stone-based AC system.
782 M. Berrios et al. / Journal of Industrial and Engineering Chemistry 18 (2012) 780–784

Table 1 It is believed that adsorption of organics onto AC depends on


Freundlich constants for the MB solution and olive stone-based AC system at
both the pore structure and surface chemical properties of carbon
different temperatures.
as well as the adsorbate. Dye adsorption tests help to determine
Temperature (8C) the capacity of carbon to adsorb molecules of a particular size. The
25 30 35 40 MB molecule has a minimum molecular diameter of 0.8 nm and
KF  103 (mg/g)(L/g)1/n 936.9 734.1 1744.1 132.7
cannot enter pores with a diameter of less than 1.3 nm [28,31].
n 0.165 0.163 0.150 0.172 Therefore, it can only enter the larger micropores, but most of it is
r2 0.999 0.999 0.998 0.997 likely to be adsorbed in mesopores. Despite the high surface area of
olive stone-based AC (587 m2/g) as other authors have reported for
olive stone-based activated carbon [32], the adsorption capacity of
empirical isotherm can be used for non-ideal adsorption that MB in aqueous solution (for example qe = 0.858 mg/g at 25 8C with
involves heterogeneous surface energy systems and is expressed an AC dose of 0.5 g/L) was poor due to the molecular diameter of
by the following equation: MB and the AC pore size distribution as described in Section 2.
1=n
Although this AC has a high surface area, its application could be
qe ¼ K F Ce (2) more suitable for molecules smaller than MB due to the high
where KF is a rough indicator of the adsorption capacity and (1/n) is microporosity of AC.
the adsorption intensity. In general, the KF value increases as the
adsorption capacity of adsorbent for a given adsorbate increases. 3.2. Adsorption kinetics
The magnitude of the exponent (1/n) indicates easy uptake of
adsorbate from aqueous solution. A value for (1/n) below one The kinetic adsorption data was evaluated to understand the
indicates a normal Langmuir isotherm, while (1/n) above one is dynamics of the adsorption process. The MB molecules must
indicative of cooperative adsorption [12]. overcome three stages before coming into contact with the active
The Freundlich constants (KF and n) were calculated by sites of AC. These stages are the migration of the solute through the
nonlinear regression using Sigmaplot1 11.0 software. The results solution to the exterior surface of the adsorbent particles,
and the regression coefficients are shown in Table 1. molecular diffusion in the boundary layer and solute movement
As can be observed in the regression coefficients, the adsorption from particle surface into the interior site by pore diffusion.
of MB from wastewater on olive stone-based AC follow the Adsorption itself could be considered almost instantaneous if the
Freundlich isotherm at all tested temperatures. According to phenomenon was purely a physical process, while mass transfer
Hameed et al. [12], our (1/n) values above 1 were indicative of could be minimized by means of suitable agitation speeds.
cooperative adsorption. This means that the binding of a MB In order to evaluate the influence of agitation speed (rpm) on
molecule to one site on AC influences the affinity of other sites. the external mass transfer, several experiments were carried out in
Similar values of n were found by Avom et al. [24] for palm tree the range of 50–250 rpm for all the temperature conditions and AC
cobs-based AC, indicating the heterogeneity of the AC surface. doses selected. Fig. 3 shows the influence of agitation speed on the
When the temperature increased, the MB amount in solid phase adsorption capacity at 25 8C and an AC dose of 8 g/L. As can be
decreased in the equilibrium (Fig. 1). Hence, the effectiveness of observed, no differences were detected between 100 and 250 rpm.
the adsorption process decreased at a high temperature because However, when the agitation speed was increased from 50 to
the desorption process took place at higher temperatures. 100 rpm, the adsorption capacity (q) was enhanced. This increase
Therefore, the MB concentration in liquid phase increased. highlighted that the mass transfer through the solution and in the
Based on the experimental data, an empirical equation was boundary layer did not limit the adsorption process when an
obtained to relate qe to temperature (T) and Ce for the olive stone- agitation speed of 100 or higher was selected.
based AC. This relationship is shown in the three-dimensional Once the external mass transfer did not limit the adsorption
graph in Fig. 2. process, the kinetics was evaluated at 100 rpm. As can be observed
in Fig. 4, the q vs. t plots for all temperatures and AC doses were
found to rise exponentially to the maximum for an AC dose of 4 g/L.
Temperature did not significantly influence the adsorption

0.6

0.5

0.4
q (mg/g)

50 rpm
100 rpm
150 rpm
25ºC, 8 g AC/L 200 rpm
250 rpm

0.0
0 20 40 60 80 100

time (min)

Fig. 2. Three-dimensional graph of the relationship between qe, temperature (T) and Fig. 3. Effect of agitation speed on the adsorption capacity at 25 8C, 8 g AC/L and an
Ce . initial MB concentration of 5 mg/L.
M. Berrios et al. / Journal of Industrial and Engineering Chemistry 18 (2012) 780–784 783

Table 2
1.2 Kinetic constants of the n-order model.

AC dose (g/L)
1.0
2 4 8 16

0.8 K̄ (1/min) 0.020  0.002 0.030  0.009 0.057  0.004 0.050  0.003
q (mg/g)

0.6

Table 2. These kinetic constants were similar to the results


0.4 obtained by Santhy and Selvapathy [7] for the adsorption of
25ºC reactive dyes. The kinetic constant increased from 0.020 to 0.030
30ºC
0.2 and 0.057 (1/min) for AC doses of 2, 4 and 8 g/L, respectively. At
4 g AC/L, 100 rpm 35ºC
40ºC higher AC doses, the kinetic constant remained approximately
0.0 stable, indicating that the adsorption rate did not increase when
the AC dose was higher than 8 g/L.
0 20 40 60 80 100
time (min)
4. Conclusions
Fig. 4. Fitting of experimental data to exponential function at 100 rpm and AC dose
of 4 g/L.
This study showed that olive stone-based activated carbon can
be used for the removal of organic compounds from aqueous
solution under a wide range of conditions. Type III adsorption
capacity in the range studied for temperature and time (25–40 8C isotherms were used for the MB system. These isotherms
and 100 min). correspond to physical adsorption in a multilayer system where
Several mathematical expressions explain the increase in MB no difference is noticed between the filling of the first layer and
concentration in a dose of AC with greater adsorption time. Most the other layers. Adsorption behaviour was described by the
studies (i.e. Langergen and Svenska) use first- and second-order Freundlich isotherm with n values that demonstrate the
kinetic equations to model adsorption capacity [12]. Lee [33] heterogeneity of the AC surface. Kinetics was evaluated by means
applied pseudo second-order kinetic equations to study the of an n-order model, showing that the reaction was a first-order
adsorption of erythrosine dye from aqueous solution using reaction. The olive stone-based AC characterization showed high
activated carbon. In order to determine the exact order of reaction, microporosity, thus indicating that the poor adsorption capacity
the following generalised kinetic equation was employed: results for MB are due to the molecular properties of this
compound.
dq
¼ Kðqmax  qÞn (3)
dt
Acknowledgements
where K is the kinetic constant and qmax is the maximum
adsorption capacity for each temperature and AC dose. The kinetic The authors are very grateful to Iberica de Carbones and Junta
constant and the order of reaction can be calculated by linearising de Andalucia (PAIDI group RNM-271) for funding this research. We
Eq. (3): also wish to express our gratitude to laboratory technician
  Inmaculada Bellido Padillo for her help.
dq
log ¼ log K þ n log ðqmax  qÞ (4)
dt
References
as shown in Fig. 5 for an AC dose of 2 g/L. The order of reaction (n)
was found to be 1 or close to 1 in all the experiments. This result [1] F. Banat, N. Al-Bastaki, Desalination 170 (2004) 69.
[2] M.F.R. Pereira, S.F. Soares, J.J.M. Órfão, J.L. Figueiredo, Carbon 41 (2003) 811.
coincides with other authors [4,6,7,34]. The mean kinetic constants [3] E.N. El Qada, S.J. Allen, G.M. Walter, Chem. Eng. J. 124 (2006) 103.
(K, 1/min) for each AC dose (2, 4, 8 and 16 g/L) can be observed in [4] N. Kannan, M.M. Sundaram, Dyes Pigments 51 (2001) 25.
[5] V. Gomez, M.S. Larrechi, M.P. Callao, Chemosphere 69 (2007) 1151.
[6] P.K. Malik, J. Hazard. Mater. B113 (2004) 81.
[7] K. Santhy, P. Selvapathy, Bioresour. Technol. 97 (2006) 1329.
0.1 [8] A.R. Binupriya, M. Sathishkumar, D. Kavitha, K. Swaminathan, S. Yun, S. Mun,
Clean 35 (2007) 143.
[9] H. Parab, M. Sudersanan, N. Shenoy, T. Pathare, B. Vaze, Clean 37 (2009) 963.
[10] P. Saha, S. Chowdhury, S. Gupta, I. Kumar, R. Kumar, Clean 38 (2010) 437.
[11] L.M. Nieto, S.B.D. Alami, G. Hodaifa, C. Faur, S. Rodrı́guez, J.A. Giménez, J. Ochando,
Ind. Crops Prod. 32 (2010) 467.
(dq/dt) (mg/g·min)

[12] B.H. Hameed, A.T.M. Din, A.L. Ahmad, J. Hazard. Mater. 141 (2007) 819.
[13] D. Mohan, K.P. Singh, V.K. Singh, J. Hazard. Mater. 152 (2008) 1045.
[14] N. Thinakaran, P. Baskaralingam, K.V. Thiruvengada Ravi, P. Panneerselvam, S.
0.01 Sivanesan, Clean 36 (2008) 798.
[15] A. Dermibas, J. Hazard. Mater. 167 (2009) 1.
[16] M. Rafatullah, O. Sulaiman, R. Hashim, A. Ahmad, J. Hazard. Mater. 177 (2010) 70.
[17] S.B. Hartono, S. Ismadji, Y. Sudaryanto, W. Irawaty, J. Ind. Eng. Chem. 11 (2005)
864.
25ºC [18] M. Soleimani, T. Kaghazchi, J. Ind. Eng. Chem. 14 (2008) 28.
30ºC [19] M.H. Alaya, M.A. Hourieh, A.M. Youssef, F. El-Seiariah, Adsorpt. Sci. Technol. 18
35ºC
(2000) 27.
40ºC
[20] W.K. Lafi, Biomass Bioenerg. 20 (2001) 57.
2 g AC/L, 100 rpm
0.001 [21] G.G. Stavropoulos, A.A. Zabaniotou, Micropor. Mesopor. Mater. 82 (2005) 79.
0.01
[22] R.S. Juang, R.L. Tseng, F.C. Wu, Adsorption 7 (2001) 65.
0.1 1 10
[23] M.K.N. Yenkie, G.S. Natarajan, Sep. Sci. Technol. 28 (1993) 1177.
(qmax - q) (mg/g) [24] J. Avom, J.K. Mbadcam, C. Noubactep, P. Germain, Carbon 35 (1997) 365.
[25] S. Senthilkumaar, P.R. Varadarajan, K. Porkodi, C.V. Subbhuraam, J. Colloid Inter-
Fig. 5. Kinetic plots for the removal of MB by adsorption on olive stone-based AC. face Sci. 284 (2005) 78.
784 M. Berrios et al. / Journal of Industrial and Engineering Chemistry 18 (2012) 780–784

[26] S.S. Barton, Carbon 25 (1987) 343. [31] H. Valdes, M. Sanchez-Polo, J. Rivera-Utrilla, C.A. Zaror, Langmuir 18 (2002)
[27] E.N. El Qada, S.J. Allen, G.M. Walter, Chem. Eng. J. 135 (2008) 174. 2111.
[28] S. Wang, Z.H. Zhu, A. Coomes, F. Haghseresht, G.Q. Lu, J. Colloid Interface Sci. 284 [32] A. Wahby, Z. Abdelouahab-Reddam, R. El Mail, M. Stitou, J. Silvestre-Albero, A.
(2005) 440. Sepúlveda-Escribano, F. Rodrı́guez-Reinoso, Adsorption 17 (2011) 603.
[29] S. Brunauer, L.S. Deming, W.E. Deming, E. Teller, J. Am. Chem. Soc. 62 (1940) 1723. [33] J.J. Lee, J. Ind. Eng. Chem. 22 (2011) 224.
[30] C. Hinz, Geoderma 99 (2001) 225. [34] K.V. Kumar, J. Hazard. Mater. B137 (2006) 1538.

You might also like