You are on page 1of 9

Journal of the Taiwan Institute of Chemical Engineers 44 (2013) 5260

Contents lists available at SciVerse ScienceDirect

Journal of the Taiwan Institute of Chemical Engineers


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

Dried prickly pear cactus (Opuntia cus indica) cladodes as a low-cost and eco-friendly biosorbent for dyes removal from aqueous solutions
Noureddine Barka a,*, Khalid Ouzaouit b, Mohamed Abdennouri a, Mohammed El Makhfouk c
a

veloppement Durable (GEDD), Faculte Polydisciplinaire de Khouribga, Universite Hassan 1er, Hay Ezzaitouna, BP 145, Khouribga, Morocco Equipe de Recherche Gestion de lEau et De lectrochimie, Centre de recherche REMINEX, Site de Hajar, BP 469, Marrakech, Morocco Laboratoire de c le rieure de Technologie de Sa, BP 89, Route Dar Si Aissa, Sa, Morocco Equipe de Recherche Analyse Contro et Environnement (ERACE), Ecole Supe
b

A R T I C L E I N F O

A B S T R A C T

Article history: Received 10 May 2012 Received in revised form 4 August 2012 Accepted 2 September 2012 Available online 28 November 2012 Keywords: Prickly pear cactus Eco-friendly biosorbent Dyes Kinetics Equilibrium Thermodynamics

The biosorption of Methylene Blue (MB), Eriochrome Black T (EBT) and Alizarin S (AS) from aqueous solutions by dried prickly pear cactus cladodes as a low-cost, natural and eco-friendly biosorbent was investigated. The study was carried out under various parameters, such as average biosorbent particle size, pH, biosorbent dosage, contact time, initial dye concentration and temperature. The experimental results show that, the percentage of biosorption increases with an increase in the biosorbent dosage, and the decrease of particle size. The biosorption was pH dependent with a high biosorption of cationic dye (MB) in basic range and a high biosorption of anionic dyes (EBT and AS) in acidic range. The equilibrium uptake was increased with an increase in the initial dye concentration in solution. Biosorption kinetic data were properly tted with the pseudo-second-order kinetic model. The experimental isotherm data were analyzed using Langmuir, Freundlich, RedlichPeterson, Temkin and Toth isotherm equations. The best t was obtained by the RedlichPeterson model and the Langmuir model with a Langmuir maximum monolayer biosorption capacity of 189.83 mg/g for Methylene Blue, 200.22 mg/g for Eriochrome Black T and 118.35 mg/g for Alizarin S. The biosorption was exothermic in nature (DH8 = 31.035 kJ/mol for MB, 10.25 kJ/mol for EBT and 11.69 kJ/mol for AS). The reaction was accompanied by a decrease in entropy (DS8 = 94.76 J/K mol for MB, 38.44 J/K mol for EBT and 41.93 J/K mol for AS). The Gibbs energy (DG8) increased when the temperature was increased from 25 to 60 8C indicating a decrease in feasibility of biosorption at higher temperatures. 2012 Taiwan Institute of Chemical Engineers. Published by Elsevier B.V. All rights reserved.

1. Introduction Dye efuents discharged from the dyestuff manufacturing, dyeing, printing and textile industries represent a serious problem all over the world. They contain different types of synthetic dyes which are known to be a major source of environmental pollution in terms of both the volume of dye discharged and the efuent composition [1]. Most of these dyes are toxic, mutagenic and carcinogenic [25]. According to the survey of the Ecological and Toxicological Association of the Dyestuffs Manufacturing Industry (ETAD), over 90% of some 4000 dyes have LD50 values greater than 2000 mg/kg [6,7]. From an environmental point of view, the removal of synthetic dyes is of great concern. Several chemical and physical discoloration methods have been used for dye removal from wastewaters [814]. However application of these methods is somewhat restricted due to some limitations such as operational costs, formation of hazardous

* Corresponding author. Tel.: +212 661 66 66 22; fax: +212 523 49 03 54. E-mail address: barkanoureddine@yahoo.fr (N. Barka).

by-products, intensive energy requirement and limited adaptability to a wide range of efuents. Biological processes such as biosorption, bioaccumulation and biodegradation have been proposed as potential methods for the removal of dyes from textile wastewater. Unfortunately, bioaccumulation and biodegradation are time-consuming because most dye compounds are resistant to microbial activity and biological treatment alone. Moreover, the biodegradation of dye may produce by-products and/or other metabolites with higher toxicity than primary substrate [15]. Therefore, alternative technology must be developed to solve these problems. Biosorption process is one of the most widely used methods for the removal of pollutants from water with advantages of high treatment efciency and no contaminative by-product released into treated water. Currently, numerous studies have focused on the dye biosorption abilities; ndings in our previous study demonstrated that the Scolymus hispanicus L. could be employed as biosorbent material for the rapid removal of Methylene Blue and eriochrome black T from aqueous solution [16]; Akar et al. demonstrates that modied P. coccinea exhibited a good potential for decolorization process of Acid Red 44 dye [17]; Vucurovic et al. demonstrated that the Sugar

1876-1070/$ see front matter 2012 Taiwan Institute of Chemical Engineers. Published by Elsevier B.V. All rights reserved. http://dx.doi.org/10.1016/j.jtice.2012.09.007

N. Barka et al. / Journal of the Taiwan Institute of Chemical Engineers 44 (2013) 5260

53

beet pulp is a promising adsorbent for removal of Methylene Blue from aqueous solutions [18]; Waste P. mutilus was successfully used by Yeddou-Mezenner for the decolorization of Basic Blue 41 dye in synthetic wastewater conditions [19]; Gupta et al. used potato plant wastes for the removal of Methylene Blue and malachite green from aqueous solution [20]; Robinson et al. found that apple pomace and wheat straw are suitable adsorbents for the removal of ve reactive dyes from a synthetic efuent [21]. However, studies in this eld have not produced materials which meet all demands of biosorption activity. The prickly pear cactus (Opuntia cus indica; Opuntia spp., Cactaceae) is native to the United States, Mexico and South America, but it grows well in other areas, including Africa, Australia and the Mediterranean region. The cactus cladodes are mainly constituted by a heteropolysaccharide with a molecular weight from 2.3 104 to 3 106 g/mol [22,23]. The Opuntia cus indica mucilage is a mixture of acidic and neutral polysaccharides consisting primarily of 24.642% of arabinose; 2140.1% of galactose; 812.7% of galacturonic acid; 713.1% of rhamnose and 2222.2% of xylose [24]. Multiples uses have been found for this component, for instance as a food thickener, food emulsier, as a water purier (polyelectrolyte molecule), as an adhesive for lime [Ca(OH)2], as a natural super plasticizer in mortars and as a food product [22,23,25,26]. The composition of the cactus cladodes could confer to this biomass the possibility to establish bonds with organic molecules including dyes. At current time, there is no information in the literature about the biosorption potential of this biomass. The focus of the present study was to assess the potentiality of dried prickly pear cactus cladodes biomass as a low-cost, natural and eco-friendly biosorbent for the biosorption of Methylene Blue, Eriochrome Black T and Alizarin S from aqueous solution as an ideal alternative to the current expensive methods of removing dyes from wastewater. Biosorption studies were carried out under various parameters such as average biosorbent particle size, pH, biosorbent dosage, contact time, initial dye concentration and temperature. The biosorption kinetic data of the biomass were tested by pseudorst-order, pseudo-second-order kinetic and Elovich models. The equilibrium data were analyzed using Langmuir, Freundlich, RedlichPeterson, Temkin and Toth isotherm models. The thermodynamics of the biosorption was also evaluated.
Table 1 Physicochemical characteristics of used dyes. Name Methylene Blue (Basic blue 9) Molecular structure

2. Materials and methods 2.1. Preparation of the biosorbent The prickly pear cactus cladodes were naturally collected in July 2011 near Sa in Morocco. They were repeatedly washed with distilled water to remove dirt particles and were sun dried for 3 weeks, cutting into small pieces and then were dried at 60 8C for 24 h. The dried plant was then powdered using domestic mixer. The biosorbent was stored in a glass bottle for further use without any pre-treatment. 2.2. Preparation of dyes solutions Methylene Blue (Analytical Reagents), Eriochrome Black T (Chimica) and Alizarin S (Riedel de Haen) were of analytical grade. Their molecular structures are given in Table 1. Stock solutions were prepared by dissolving requisite quantity of each dye without further purication in distilled water, and the concentrations used were obtained by dilution of the stock solution. The pH was adjusted to a given value by addition of HCl (1 N) or NaOH (1 N) and was measured using a JENWAY pH-Meter 3305. 2.3. Biosorption experiments Biosorption experiments were conducted in 250 mL conical asks at a constant agitation speed. The experiments were carried out by varying the biosorbent particle size from particles less than 100 mm to particles bigger than 500 mm, contact time from 5 to 90 min, the biosorbent dosage from 0.2 to 4 g/L, the pH from 2 to 10, the initial concentration from 30 to 300 mg/L and the temperature from 25 to 60 8C. The temperature was controlled using an isothermal shaker. After the each biosorption procedure completed, the sample were centrifuged at 3000 rpm for 10 min to separate the solid phase from the liquid phase. 2.4. Analyses FTIR spectral analysis was recorded in a SP-FTIR-1 infrared spectrometer in the region of 4004000 cm1 and the samples were prepared as KBr pellets under high pressure. Surface area was determined by using N2 as the sorbate at 77 K in a Micromeritics

MW (g/mol)

lmax (nm)
661

N (CH3)2 N + S Cl
HO OH N N SO3Na

319.85

N (CH3)2
461.38 526

Eriochrome Black T (Mordant black 11)

NO2
Alizarin S (Mordant red 3)

OH OH SO3Na

342.26

520

54

N. Barka et al. / Journal of the Taiwan Institute of Chemical Engineers 44 (2013) 5260 Table 2 Surface area of dried prickly pear cactus cladodes with different particle sizes. Particle size d < 100 mm 100 mm < d < 200 mm 250 mm < d < 500 mm d > 500 mm Surface area m2/g 1.016 0.839 0.668 0.533

TriStar II 3020 sorptometer. Samples were outgassed prior to use at 473 K a night under vacuum. Specic total surface areas were calculated using the B.E.T. The point of zero charge was determined according to the method described by Ferro-Garcia et al. [27]. Dried cactus (50 mg) was suspended in 50 ml of NaNO3 solution (0.1 M). The initial pH of the solution was adjusted to dened values from 2.0 to 12.0 using HNO3 and NaOH solutions. The suspension was sealed and allowed to equilibrate at 100 rpm in a shaker at 25 8C for 6 h. After this period, the pH of the solution (pHf) is measured. pHpzc is derived from the curve pH = pHi pHf = f (pHi) as the intercept of the abscissa. Dye concentrations were determined from UVVis absorbance characteristic with the calibration curve method at the maximum of absorbance (Table 1). Before analysis, the pH of all the samples was readjusted to initial pH. A GBC UV/Vis 911 spectrophotometer was used. The adsorbed amounts and the biosorption yield were calculated by measuring the concentration of dye in solution before and after adsorption and using the following equations: q C 0 C R (1) (2)

C 0 C % Biosorption 100 C0

where q (mg/g) is the quantity of dye adsorbed per unit mass of biosorbent, % Biosorption is the biosorption yield, C0 (mg/L) is the initial dye concentration, C (mg/L) is the dye concentration after biosorption and R (g/L) is the mass of biosorbent per liter of aqueous solution. 3. Results and discussion 3.1. Analysis of the biosorbent The FT-IR spectra of dried prickly pear cactus cladodes in the range of 4004000 cm1 were taken to obtain information on the nature of functional groups at the surface of the biosorbent. The spectra presented in Fig. 1 show broad, strong and superimposed bands around 36003200 cm1 which may be due to the overlapping of OH and NH stretching vibration. The band at 2921 cm1 and the band at 2850 cm1 are due to asymmetric stretching vibration of CH2 and the symmetric stretching vibration of CH3, respectively, of aliphatic acids [28]. The peaks observed between 1730 and 1710 cm1, are indicative of stretching vibration of CO bonds due to non-ionic carboxyl groups (COOH, COOCH3) and may be assigned to hydrogen bonding between carboxylic acids or their esters [28]. The stretching vibration band at 1650 cm1 is due to asymmetric stretching of the carboxylic COO double bond of deprotonated carboxylate functional groups

[29]. A 1432 cm1 is of phenolic OH stretching. The peaks observed at 1370 cm1, which reect stretching vibrations of symmetrical or asymmetrical ionic carboxylic groups (COOH) of pectin [29]. The band at 1035 cm1 band could be due to the vibration of COC and OH of polysaccharides [30]. The peaks at 1254.6 and 1033.5 cm1 are due to the CO stretching vibration of ketones, aldehydes and lactones or carboxyl groups [31]. The band at 1324 cm1 assigned to CN groups on the biomass surface. The absorption peaks around 1155 and 1070 cm1 are indicative of P5 and POH stretching vibra5O tions. The SO stretching was identied at 952 cm1. Peaks in the region of wavenumbers lower than 800 cm1 could be attributed to N-containing bioligands. This results indicate that the dried cactus contain a variety of functional groups such as carboxyl, hydroxyl, sulfate, phosphate, aldehydes, ketones and other charged groups, suggesting that this biomaterial can be considered as an alternative for dyes removal. The B.E.T. surface areas of the biosorbent with different particle sizes are presented in Table 2. The table shows that dried prickly pear cactus has a very small surface area. This result suggested that this biosorbent do not have any dened holes on their morphology. The table also showed that the surface area decreases with the increase in the particle size. The pHzpc of the natural biosorbent was found to be at pH of 3.5. The zero point charge (pHzpc) of the biosorbent is one way to understand the biosorption mechanisms [32,33]. When the pH of the suspension is lower than 3.5, the surface of the dried cactus gets positively charged, and the surface of the dried cactus is negatively charged in pH values higher than 3.5. The value obtained shows that the dried cactus have an acidic surface which is in agreement with FTIR analysis which indicates that carboxylic acids are the main functional groups of this biomaterial. 3.2. Effect of biosorbent particle size Particle size is one of the important factors which affect the biosorption capacity. Kinetics of MB biosorption was carried out for different particle sizes of dried cactus cladodes varying from particles less than 100 mm to particles bigger than 500 mm. The results are illustrated in Fig. 2. The gure indicates that the amount of dye adsorbed at equilibrium was increased and the equilibrium time was decreased by decreasing the particle size. The amount of dye adsorbed at equilibrium increased from 54.91 to 118.80 mg/g for a decrease in biosorbent particle size from particles bigger than 500 mm to particles less than 100 mm. According to Aksu [34] the biosorption is related directly with surface area of biosorbent. The fast and higher adsorption capacity with smaller absorbent particles could be attributed to the fact that smaller particles provided a larger and easily accessible surface area as shown in Table 2. This situation was explained by larger total surface area of smaller particles, for the same amount of biomass. Similar behavior was found by Chu and Chen [35] in the biosorption of basic yellow 24 using dried activated sludge biomass. For further experiments, the fraction of particles less than 100 mm was selected.

Fig. 1. FT-IR spectra of dried prickly pear cactus cladodes biomass.

N. Barka et al. / Journal of the Taiwan Institute of Chemical Engineers 44 (2013) 5260

55

125

80

100

60
q (mg/g)

% Biosorption

75

40 Methylene blue 20 Eriochrome black T Alizarin S 0 0 1 2 R (g/L) 3 4

50 <100m 250-500m 0 0 15 30 45 60 75 90 Contact time (min)


Fig. 2. Effect of biosorbent particle size on the biosorption kinetics of MB by dried cactus cladodes: C0 = 100 mg/L, temperature = 25 8C, R = 0.5 g/L, and initial pH 5.2.

25

100-250m >500m

Fig. 4. Effect of biosorbent dosage on the biosorption of MB, EBT and AS by dried cactus: C0 = 100 mg/L, particle size <100 mm, contact time = 90 min, temperature = 25 8C, and initial pH.

3.3. Effect of pH 3.4. Effect of biosorbent dosage Since hydrogen ions affect the surface charge of the adsorbents and the adsorbate species [36,37], the sorption is greatly affected by the variation of solution pH. We have thus studied the efciency of the adsorption when the pH is varied in the range of 210 as shown in Fig. 3. The gure indicates that, for Methylene Blue, the biosorption is weak in acidic medium below pH 3. As the pH increase, the biosorption capacity for MB increases. For AS and EBT, the biosorption is high in acidic medium and decreases with the increase in pH of the solution. The isoelectric point of the natural biosorbent was found to be at pH of 3.5. At pH values above this point, there is a net negative charge on the cell surface and the ionic state of ligands such as carboxyl, phosphoryl, sulfhydryl, hydroxyl, and amino groups will be such so as to promote reaction with Methylene Blue as a cationic dye. As the pH is lowered, however, the overall surface charge on the cells will become positive, which will promote reaction with Eriochrome Black T as anionic dyes through electrostatic forces of attraction. This suggests that sorption via coulombic interaction might not be the main adsorption mechanism [38]. Biosorbent dosage is an important parameter because it determines the capacity of biosorbent for a given initial concentration of dye molecules. Data obtained from the experiments with varying biosorbent concentrations are presented in Fig. 4. It showed that increasing biosorbent dosage resulted in a sharply increase in the biosorption yield. The biosorption yield of MB increased from 37.74% to 68.92% when the biosorbent dosage was increased from 0.2 to 1 g/L, from 13.37 and 18.24% to 59.78 and 61.80% respectively, for EBT and AS, when the biosorbent dosage was increased from 0.2 to 3 g/L. According to Chowdhury et al. [39] the observed enhancement in biosorption yield with increasing biosorbent concentration could be due to an increase in the number of possible binding sites and surface area of the biosorbent. A further increase in biomass concentration over 1 g/L for BM and 3 g/L for EBT and AS did not lead to a signicant improvement in biosorption yield. It could be explained as a consequence of a partial aggregation of the biosorbent particles, which results in a decrease in effective surface area for the biosorption [40]. Similar behavior was found by Barka et al. [41] in the biosorption of basic yellow 24 using dried activated sludge biomass. 3.5. Biosorption kinetics

140 120 100 MB EBT AS

qe (mg/g)

80 60 40 20 0 0 2 4 6 pH 8 10

12

Biosorption rate gives important information for designing batch biosorption systems. Information on the kinetics of solute uptake is required for selecting optimum operating conditions for full-scale batch process. Fig. 5 shows the plot of MB, EBT and AS biosorption versus contact time. The biosorption was found to be rapid at the rst period of the process and then the rate of biosorption be slower and stagnates with the increase in contact time. The equilibrium time was 60 min for the three dyes. In order to characterize the kinetics involved in the process of biosorption, pseudo-rst order, pseudo-second order and Elovich equations were proposed and the kinetic data were analyzed based on the regression coefcient (r2) and the amount of dye adsorbed per unit weight of the biosorbent. The rst-order rate expression of Lagergren based on solid capacity is generally expressed as follows [42]: dq k1 qe q dt (3)

Fig. 3. Effect of pH on the biosorption of MB, EBT and AS by dried cactus cladodes: C0 = 100 mg/L, particle size <100 mm, R = 0.5 g/L, contact time = 90 min, and temperature = 25 8C.

56

N. Barka et al. / Journal of the Taiwan Institute of Chemical Engineers 44 (2013) 5260

125 100 MB EBT AS 75

50 25 0 0 15 30 45 60 75 90 Contact time (min)

Fig. 5. Effect of contact time on biosorption of MB, EBT and AS by dried cactus cladodes: C0 = 100 mg/L, R = 0.5 g/L, particle size <100 mm, temperature = 25 8C, and initial pH.

where a is the initial adsorption rate (mg/g min) and b is the desorption constant (g/mg). Parameters of the pseudo-rst-order, Pseudo-second-order and Elovich models were estimated with the aid of the non-linear regression. The obtained data and the correlation coefcients, r2, are given in Table 3. The table shows that the three models give a good correlation to experimental data. However, the correlation coefcients for the pseudo-second-order kinetic model are closer to one than those of the Lagergren rst order and Elovich model. Therefore, the sorption can be approximated more appropriately by the pseudo-secondorder kinetic. This result suggests that boundary layer resistance was not the rate limiting step [43]. The rate of dye adsorption may be controlled largely by a chemisorption process, in conjunction with the chemical characteristics of the biomass and dyes. Similar modeling results are also found in the kinetic studies on biosorption of Methylene Blue and Eriochrome Black T by Scolymus hispanicus L. [16], Direct Blue 199 by Aspergillus niger [45], Reactive Yellow 42 and Reactive Red 45 by chemically treated Citrus sinensis waste biomass [46], Acid Blue 225 and Acid Blue 062 by P. macerans [18] and Direct Blue 71 by palm ash [47]. 3.6. Mechanism of biosorption

After integrating and applying the boundary conditions, for q = 0 at t = 0 and q = q at t = t, the integrated form of Eq. (3) becomes: q qe 1 ek1 t (4)

q (mg/g)

where qe and q (both in mg/g) are respectively the amounts of dye adsorbed at equilibrium and at any time t, and k1 (1/min) is the rate constant of biosorption. The pseudo-second-order model proposed by Ho and McKay [43] was used to explain the sorption kinetics. This model is based on the assumption that the adsorption follows second order chemisorption. The pseudo-second-order model can be expressed as: dq k2 qe q2 dt (5)

For the interpretation of experimental kinetics data, from a mechanistic viewpoint, prediction of the rate-limiting step is an important consideration. The adsorbate transport from the solution phase to the surface of the adsorbent particles occurs in several steps. The overall adsorption process may be controlled either by one or more steps, e.g., lm or external diffusion, pore diffusion, surface diffusion and adsorption on the pore surface, or a combination of more than one step. The possibility of intra-particle diffusion was explored by using the intra-particle diffusion model [48]. q kid t 0:5 I (9)

After integrating for the similar boundary conditions, the following equation can be obtained: q k2 q2 t e 1 k2 qe t (6)

where k2 (g/mg min) is the rate constant of pseudo-second order adsorption. In reactions involving chemisorption of adsorbate on a solid surface without desorption of products, adsorption rate decreases with time due to an increased surface coverage. One of the most useful models for describing such chemisorption is the Elovich model. The Elovich model equation is generally expressed as [44]: dq aexpbq dt (7)

After integrating for the similar boundary conditions, the following equation can be obtained: q 1

ln1 a b t

(8)

where kid is the intra-particle diffusion rate constant (mg/g min1/2) and I (mg/g) is a constant. If the WeberMorris plot of qt versus t1/2 satises the linear relationship with the experimental data, then the sorption process is found to be controlled by intra-particle diffusion only. However, if the data exhibit multi-linear plots, then two or more steps inuence the sorption process. Fig. 6 shows the pore-diffusion plot of MB, EBT and AS biosorption on dried cactus cladodes. It is clear that the plots are multilinear, having at least three linear segments. The rst, sharper portion is attributed to the diffusion of adsorbate through the solution to the external surface of adsorbent or boundary layer diffusion of solute molecules. The second, linear portion is the gradual equilibrium stage with intra-particle diffusion dominating. The third portion is attributed to the nal equilibrium stage for which the intra-particle diffusion starts to slow down due to the extremely low adsorbate concentration left in the solution [49]. A decrease in the slope (which is equal to kid) of each segment was observed. This decrease for macro to mesopore diffusion is a normal consequence of the relative free path for diffusion that is available in each pore size. As pore size decreases, the path

Table 3 Kinetic constants for MB, EBT and AS biosorption onto dried cactus cladodes. Dye qexp Pseudo-rst-order qe MB EBT AS 118.80 47.75 56.69 116.63 47.06 55.51 k1 0.563 0.483 0.355 r2 0.989 0.989 0.991 Pseudo-second-order qe 120.77 48.92 58.42 K2 0.009 0.018 0.010 r2 0.998 0.997 0.999 Elovich

a
82470.96 7156.03 693.57

b
0.109 0.237 0.151

r2 0.986 0.979 0.976

N. Barka et al. / Journal of the Taiwan Institute of Chemical Engineers 44 (2013) 5260

57

125

100 MB EBT AS

q (mg/g)

75

50

25

0 0 2 4 6 8 10
Fig. 8. Effect of EBT concentration on its biosorption by dried cactus: R = 0.5 g/L, particle size <100 mm, contact time = 90 min, temperature = 20 8C, and initial pH.

time 1/2 (min1/2 )


Fig. 6. Intraparticle diffusion plots for the biosorption of MB, EBT and AS by dried cactus cladodes.

available for diffusion becomes smaller, which leads to a decrease in the rate of diffusion. 3.7. Biosorption isotherms The adsorption capacities of dried cactus for MB, EBT and AS were studied for different initial dye concentrations as shown in Figs. 79. The results indicated that the adsorption capacity increases with increasing the initial dye concentration. The increase in biosorption capacity with concentration is probably due to a high driving force for mass transfer. In fact, high concentration in solution implicates high molecules of dye xed at the surface of the biosorbent. Several theories of adsorption equilibrium were applied for the analysis of equilibrium sorption data obtained. 3.7.1. Langmuir model The Langmuir adsorption model [50] is established on the following hypotheses: (1) uniformly energetic adsorption sites, (2) monolayer coverage, and (3) no lateral interaction between adsorbed molecules. Graphically, a plateau characterizes the

Langmuir isotherm. Therefore, at equilibrium, a saturation point is reached where no further adsorption can occur. A basic assumption is that sorption takes place at specic homogeneous sites within the adsorbent. Once a dye molecule occupies a site, no further adsorption can take place at that site. A mathematical expression of the Langmuir isotherm is given by Eq. (10): qe qm K L C e 1 K LCe (10)

where qe (mg/g) is the adsorbed amount at equilibrium, Ce is the equilibrium dye concentration (mg/L), KL is Langmuir equilibrium constant (L/mg) and qm the maximum adsorption capacity (mg/g). 3.7.2. Freundlich model The Freundlich isotherm endorses the heterogeneity of the surface and assumes that the adsorption occurs at sites with different energy of adsorption. The energy of adsorption varies as a function of surface coverage [51]. This equation is also applicable to multilayer adsorption and is expressed by the following equation: qe K F Ce
1=n

(11)

Fig. 7. Effect of MB concentration on its biosorption by dried cactus: R = 0.5 g/L, particle size <100 mm, contact time = 90 min, temperature = 20 8C, and initial pH.

Fig. 9. Effect of AS concentration on its biosorption by dried cactus: R = 0.5 g/L, particle size <100 mm, contact time = 90 min, temperature = 20 8C, and initial pH.

58

N. Barka et al. / Journal of the Taiwan Institute of Chemical Engineers 44 (2013) 5260 Table 4 Adsorption isotherm constants for biosorption of MB, EBT and AS by dried cactus cladodes. Isotherm model MB 189.83 0.0255 0.986 EBT 200.22 0.0114 0.988 AS 118.35 0.0283 0.990

where KF is the Freundlich constant and n is the heterogeneity factor. The KF value is related to the adsorption capacity; while 1/n value is related to the adsorption intensity. 3.7.3. RedlichPeterson model As a combination of both Langmuir and Freundlich systems, Redlich and Peterson [52] have proposed an empirical equation, designated the three parameter equation, which may be used to represent adsorption equilibrium over a wide concentration range. qe K RCe 1 aR Ce
b

Langmuir qm KL r2 Freundlich KF n r2 RedlichPeterson KR aR

(12)

21.05 2.574 0.957

9.29 1.944 0.966

17.36 2.991 0.987

where b 1. This equation reduces to a linear isotherm at low surface coverage, to the Freundlich isotherm at high adsorbate concentration, and to the Langmuir isotherm when b = 1. 3.7.4. Temkin model Temkin and Pyzhev [53] considered the effects of some indirect adsorbate/adsorbate interactions on adsorption isotherms and suggested that because of these interactions the heat of adsorption of all the molecules in the layer would decrease linearly with coverage. The Temkin isotherm has been used in the following form: qe RT lnAC e b (13)

b
r2 Temkin A b r2 Toth qm KT t r2

4.92 0.0273 0.992 0.987

2.17 0.0077 1.061 0.989

5.80 0.1296 0.828 0.993

0.225 58.10 0.969

0.100 53.20 0.984

0.359 104.89 0.979

168.59 0.0043 1.434 0.982

168.01 0.0018 1.407 0.981

174.40 0.2490 0.488 0.984

where qe is the adsorbed amount at equilibrium (mg/g), Ce the equilibrium concentration of the adsorbate (mg/L), A and b are Temkin parameters. The constant b is related to the heat of adsorption. 3.7.5. Toth model The Toth isotherm [54] is derived from potential theory and is applicable to heterogeneous adsorption. It assumes a quasiGaussian energy distribution. Most sites have an adsorption energy lower than the peak or maximum adsorption energy. The application of his equation is best suited to multilayer adsorption similar to BET isotherms which is a special type of Langmuir isotherm and has very restrictive validity [55]. The Toth correlation is given as: qe qm C e
t 1=K T Ce 1=t

isotherms have the highest r2 values for both dyes whereas the Freundlich and Temkin values are considerably lower. A comparison of the r2 values for models suggested that the RedlichPeterson and Langmuir isotherms provide a better t for the experimental data compared to the Freundlich, Toth and Tempkin isotherms. The exponent b for RedlichPeterson model was close to 1, showing the closeness of the model to the Langmuir isotherm. This result suggests that the dyes are homogeneously adsorbed on a monolayer surface of the adsorbent. The maximum monolayer adsorption capacities are 189.83 mg/g for Methylene Blue, 200.22 mg/g for Eriochrome Black T and 118.35 mg/g for Alizarin S. 3.8. Biosorption thermodynamics Evaluation of temperature was carried out with the scope of testing the ability of dried cactus in dyes removal in the case of different kinds of efuents, bearing in mind the specic circumstances of dyestuff wastes. Data were collected at ve temperatures: from 25 to 60 8C. The variation of MB, EBT and AS adsorbed on dried cactus as function of solution temperature is shows in Fig. 10. A decrease of the amount of dye adsorbed was observed when the temperature increases. From these results, thermodynamic parameters including the change in free energy (DG8), enthalpy (DH8) and entropy (DS8) were used to describe thermodynamic behavior of the biosorption of MB, EBT and AS onto dried cactus. These parameters were calculated by considering the following reversible process:

(14)

where qe is the adsorbed amount at equilibrium (mg/g), Ce the equilibrium concentration of the adsorbate (mg/L), qm the Toth maximum adsorption capacity (mg/g), KT the Toth equilibrium constant, and t is the Toth model exponent. The model of Toth is reduced to the Langmuir model when the exponent t is equal to unity. 3.7.6. Analysis of adsorption isotherms The amounts of adsorbed quantities of each dye at the equilibrium (qe), versus equilibrium dye concentration were drawn in Figs. 79. The isotherms form was type L in Giles classication [56]. These types of isotherms are usually associated with ionic solute adsorption (e.g., metal cations and ionic dyes) with weak competition with the solvent molecules [57]. The experimental adsorption isotherms obtained were compared with the adsorption isotherm models and the constants appearing in each equation of those models were determined by nonlinear regression analysis. The results of these analyses are tabulated in Table 4. The correlation coefcients (r2) are also shown in this table. The table indicates that all the isotherms give reasonable t to experimental data. Overall, the RedlichPeterson and Langmuir

dye in solution biosorbed dye For such equilibrium reactions, KD, the distribution constant, can be expressed as: KD qe Ce (15)

The Gibbs free energy is:

DG RT lnK D

(16)

N. Barka et al. / Journal of the Taiwan Institute of Chemical Engineers 44 (2013) 5260

59

125 MB EBT AS

4. Conclusions In this study, the removal of Methylene Blue, Eriochrome Black T and Alizarin S from aqueous solutions by biosorption onto dried prickly pear cactus cladodes as a low-cost and natural available sorbent was investigated. The results show that the natural biomass of prickly pear cactus is an excellent biosorbent for the used dyes. The biosorption was rapid and increased by the decrease in biosorbent average particle size. The pH experiments showed that the signicant biosorption takes place in acidic range for Methylene Blue and in basic range for Eriochrome Black T and AS. The increase in mass biosorbent leads to increase in dye biosorption due to increase in number of biosorption sites. The equilibrium uptake was increased with increasing the initial concentration of dye in solution. Experimental data were better described by pseudo-second-order model. The adsorption isotherm could be well tted by the RedlichPeterson and Langmuir equations. The biosorption capacity decreases with an increase in solution temperature. Finally, the prickly pear cactus cladodes can be used as an effective natural biosorbent for the economic treatment of wastewater containing synthetic dyes. References

100

q (mg/g)

75

50

25

0 15 25 35 45 55 65 Temperature (C)
Fig. 10. Effect of temperature on the biosorption of MB, EBT and AS by dried cactus: C0 = 100 mg/L, R = 0.5 g/L, particle size <100 mm, contact time = 90 min, and initial pH.

where R is the universal gas constant (8.314 J mol/K) and T is solution temperature in K. The enthalpy (DH8) and entropy (DS8) of biosorption were estimated from the slope and intercept of the plot of ln KD versus 1/T yields, respectively. Ln K D

D G
RT

DH 
RT

DS 
R

(17)

The calculated constants are illustrated in Table 5. From the table it can be concluded that the biosorption was exothermic (DH8 = 31.035 kJ/mol for MB, 10.25 kJ/mol for EBT and 11.69 kJ/mol for AS). Also, the magnitude gives information on the type of biosorption, which can be either physical or chemical. The reaction was accompanied by a decrease in entropy (DS8 = 94.76 J/K mol for MB, 38.44 J/K mol for EBT and 41.93 J/K mol for AS). The negative DS8 value suggests a decrease in the randomness at the solid/solution interface during the biosorption. The Gibbs energy (DG8) increased when the temperature was increased from 25 to 60 8C indicating a decrease in feasibility of biosorption at higher temperatures.
Table 5 Thermodynamic parameters calculated for the biosorption of MB, EBT and AS by dried cactus. T (8C) Methylene Blue 25 30 40 50 60 Eriochrome black T 25 30 40 50 60 Alizarin S 25 30 40 50 60 KD 2.926 2.562 1.742 1.219 0.784 0.627 0.545 0.518 0.443 0.394 0.791 0.646 0.506 0.498 0.468

DG8 (J/mol)
2660.13 2369.97 1444.51 531.15 674.17 1155.56 1530.96 1709.03 2183.48 2575.72 580.10 1100.00 1772.42 1870.64 2104.31

DH8 (kJ/mol)
31.03

DS8 (J/K mol)


94.76

10.25

38.44

11.69

41.93

[1] Vandevivere PC, Bianchi R, Verstraete W. Treatment and reuse of wastewater from the textile wet-processing industry: review of emerging technologies. J Chem Technol Biotechnol 1998;72:289302. [2] Barka N, Assabbane A, Nounah A, Ait-Ichou Y. Photocatalytic degradation of indigo carmine in aqueous solution by TiO2-coated non-woven bres. J Hazard Mater 2008;152:10549. [3] Itoh K, Kitade Y, Yatote C. A pathway for biodegradation of an anthraquinone dye. C. I. Disperse Red 15, by a yeast strain Pichia anomala. Bull Environ Contam Toxicol 1996;56:4138. [4] Pinherio HM, Touraud E, Tomas O. Aromatic amines from azo dye reduction: status review with emphasis on direct UV spectrophotometric detection in textile industry wastewater. Dyes Pigments 2004;61:12139. [5] Suzuki T, Timofei S, Kurunczi L, Dietze U, Schuurmann G. Correlation of aerobic biodegradability of sulfonated azo dyes with the chemical structure. Chemosphere 2001;45:19. [6] Robinson T, McMullan G, Marchant R, Nigam P. Remediation of dyes in textile efuent: a critical review on current treatment technologies with a proposed alternative. Bioresour Technol 2001;77:24755. [7] Shore J. Advances in direct dyes. Indian J Fib Text Res 1996;21:129. [8] Gomathi Devi L, Anantha Raju KS, Girish Kumar S, Eraiah Rajashekhar K. Photodegradation of di azo dye Bismarck Brown by advanced photo-Fenton process: Inuence of inorganic anions and evaluation of recycling efciency of iron powder. J Taiwan Inst Chem Eng 2011;42:3419. [9] Barka N, Qourzal S, Assabbane A, Nounah A, Ait-Ichou Y. Photocatalytic degradation of an azo reactive dye. Reactive Yellow 84, in water using an industrial titanium dioxide coated media. Arab J Chem 2010;3:27983. [10] Barka N, Qourzal S, Assabbane A, Nounah A, Ait-Ichou Y. Adsorption of Disperse Blue SBL dye by synthesized poorly crystalline hydroxyapatite. J Environ Sci 2008;20:126872. [11] Fernandez-Sanchez C, Costa-Garcia A. Voltammetric studies of indigo adsorbed on pre-treated carbon paste electrodes. Electrochem Commun 2000;2:77681. [12] Jedidi I, Sadi S, Khmakem S, Larbot A, Elloumi-Ammar N, Fourati A, et al. New ceramic microltration membranes from mineralcoal y ash. Arab J Chem 2009;2:319. [13] Saratale RG, Saratale GD, Chang JS, Govindwar SP. Bacterial decolorization and degradation of azo dyes: a review. J Taiwan Inst Chem Eng 2011;42:138 57. [14] Aber S, Salari D, Parsa MR. Employing the Taguchi method to obtain the optimum conditions of coagulationocculation process in tannery wastewater treatment. Chem Eng J 2010;162:12734. [15] Ramsay JA, Nguyen T. Decoloration of textile dyes by Trametes versicolor and its effect on dye toxicity. Biotechnol Lett 2002;24:175660. [16] Barka N, Abdennouri M, Makhfouk ELM. Removal of Methylene Blue and Eriochrome Black T from aqueous solutions by biosorption on Scolymus hispanicus L.: kinetics, equilibrium and thermodynamics. J Taiwan Inst Chem Eng 2011;42:3206. [17] Akar T, Celik S, Tunali Akar S. Biosorption performance of surface modied biomass obtained from Pyracantha coccinea for the decolorization of dye contaminated solutions. Chem Eng J 2010;160:46672. [18] Vucurovic VM, Razmovski RN, Tekic MN. Methylene Blue (cationic dye) adsorption onto sugar beet pulp: equilibrium isotherm and kinetic studies. J Taiwan Inst Chem Eng 2012;43:10811. [19] Yeddou-Mezenner N. Kinetics and mechanism of dye biosorption onto an untreated antibiotic waste. Desalination 2010;262:2519.

60

N. Barka et al. / Journal of the Taiwan Institute of Chemical Engineers 44 (2013) 5260 [39] Chowdhury S, Chakraborty S, Saha P. Biosorption of Basic Green 4 from aqueous solution by Ananas comosus (pineapple) leaf powder. Colloids Surf B Biointerfaces 2011;84:5207. [40] Crini G, Peindy HN, Gimbert F, Robert C. Removal of C.I. Basic Green 4 (Malachite Green) from aqueous solutions by adsorption using cyclodextrin-based adsorbent: kinetic and equilibrium studies. Sep Purif Technol 2007;53:97110. [41] Barka N, Qourzal S, Assabbane A, Nounah A, Ait-Ichou Y. Removal of reactive yellow 84 from aqueous solutions by adsorption onto hydroxyapatite. J Saudi Chem Soc 2011;15:2637. [42] Lagergren S. Zur Theorie der Sogenannten Adsorption Geloster Stoffe, Kungliga Svenska Vetenskapsakademiens. Handlingar 1898;24:139. [43] Ho YS, McKay G. Pseudo-second-order model for sorption processes. Process Biochem 1999;34:45165. [44] Zeldowitsch J. Uber den mechanismus der katalytischen oxydation von CO an MnO2. Acta Physicochim URSS 1934;1:44964. [45] Xiong X-J, Meng X-J, Zheng T-L. Biosorption of C.I. Direct Blue 199 from aqueous solution by nonviable Aspergillus niger. J Hazard Mater 2010;175:2416. [46] Asgher M, Bhatti HN. Mechanistic and kinetic evaluation of biosorption of reactive azo dyes by free, immobilized and chemically treated Citrus sinensis waste biomass. Ecol Eng 2010;36:16605. [47] Ahmad AA, Hameed BH, Aziz N. Adsorption of direct dye on palm ash: kinetic and equilibrium modeling. J Hazard Mater 2007;141:706. [48] Weber Jr WJ, Morris JC. Kinetics of adsorption on carbon from solution. J Sanitary Eng Div ASCE 1963;89:159. [49] Crank J. The mathematics of diffusion, rst ed., London: Oxford Clarendon Press; 1965. [50] Langmuir I. The adsorption of gases on plane surfaces of glass, mica, and platinum. J Am Chem Soc 1916;40:1361403. [51] Freundlich HMF. Uber die adsorption in losungen. Zeitschrift fur Physikalische Chemie 1906;57A:385470. [52] Redlich O, Peterson DL. A useful adsorption isotherm. J Phys Chem 1959;63:10246. [53] Temkin MJ, Pyzhev V. Kinetics of ammonia synthesis on promoted iron catalysts. Acta Physiochim URSS 1940;12:21722. [54] Toth J. State equations of the solidgas interface layers. Acta Chem Acad Sci Hung 1971;69:31128. [55] Khan AR, Ataullah R, Al-Haddad A. Equilibrium adsorption studies for some aromatic pollutants from dilute aqueous solutions on activated carbon at different temperatures. J Colloid Interface Sci 1997;194:15465. [56] Giles CH, Smith D, Huitson A. A general treatment and classication of the solute adsorption isotherm. I. Theoretical. J Colloid Interface Sci 1974;47:755 65. [57] Giles CH, DSilva AP, Easton IA. A general treatment and classication of the solute adsorption isotherm part. II. Experimental interpretation. J Colloid Interface Sci 1974;47:76678.

[20] Gupta N, Kushwaha AK, Chattopadhyaya MC. Application of potato (Solanum tuberosum) plant wastes for the removal of Methylene Blue and malachite green dye from aqueous solution. Arab J Chem 2011. http://dx.doi.org/ 10.1016/j.arabjc.2011.07.021. [21] Robinson T, Chandran B, Nigam P. Removal of dyes from a synthetic textile dye efuent by biosorption on apple pomace and wheat straw. Water Res 2002;36:282430. [22] Cardenas A, Higuera-Ciapara I, Goycoolea FM. Rheology and aggregation of cactus (Opuntia cus-indica) mucilage in solution. J Prof Assoc Cactus Develop 1997;2:1527. [23] Medina-Torres L, Brito-De La Fuente E, Torrestiana-Sanchez B, Katthain R. Rheological properties of the mucilage gum (Opuntia cus indica). Food Hydrocolloids 2000;14:41724. [24] Trachtenberg S, Mayer AM. Composition and properties of Opuntia cus-indica mucilage. Phytochemistry 1981;20:26658. [25] Miller SM, Fugate EJ, Craver VO, Smith JA, Zimmerman JB. Toward understanding the efcacy and mechanism of Opuntia spp. As a natural coagulant for potential application in water treatment. Environ Sci Technol 2008;42:42749. [26] Saenz C, Sepulveda E, Matsuhiro B. Opuntia spp mucilages: a functional component with industrial perspectives. J Arid Environ 2004;57:27590. [27] Ferro-Garcia MA, Rivera-Utrilla J, Bantista-Toledd I, Moreno-Castilla AC. Adsorption of humic substances on activated carbon from aqueous solutions and their effect on the removal of Cr(III) Ions. Langmuir 1998;14:18806. [28] Li FT, Yang H, Zhao Y, Xu R. Novel modied pectin for heavy metal adsorption. Chin Chem Lett 2007;18:3258. [29] Farinella NV, Matos GD, Arruda MAZ. Grape bagasse as a potential biosorbent of metals in efuent treatments. Bioresour Technol 2007;98:19406. [30] Ibarra JV, Moliner R. Coal characterization using pyrolysis-FTIR. J Anal Appl Pyrol 1991;20:17184. [31] Chandrasekhar S, Pramada PN. Rice husk ash as an adsorbent for Methylene Blue-effect of ashing temperature. Adsorption 2006;12:2743. [32] Celekli A, Bozkurt H. Bio-sorption of cadmium and nickel ions using Spirulina platensis: kinetic and equilibrium studies. Desalination 2011;275:1417. [33] Hao X, Quach L, Korah J, Spieker WA, Regalbuto JR. The control of platinum impregnation by PZC alteration of oxides and carbon. J Mol Catal A Chem 2004;219:97107. [34] Aksu Z. Application of biosorption for the removal of organic pollutants: a review. Proc Biochem 2005;40:9971026. [35] Chu CH, Chen KM. Reuse of activated sludge biomass: II. The rate processes for the adsorption of basic dyes on biomass. Proc Biochem 2002;37:112934. [36] Gupta VK, Rastogi A, Saini VK, Jain N. Biosorption of copper(II) from aqueous solutions by Spirogyra species. J Colloid Interface Sci 2006;296:5963. [37] Gupta VK, Rastogi A. Equilibrium and kinetic modelling of cadmium(II) biosorption by nonliving algal biomass Oedogonium sp from aqueous phase. J Hazard Mater 2008;153:75966. [38] Zheng J-C, Feng H-M, Lam MHW, Lam PKS, Ding Y-W, Yu H-Q. Removal of Cu(II) in aqueous media by biosorption using water hyacinth roots as a biosorbent material. J Hazard Mater 2009;171:7805.

You might also like