You are on page 1of 9

Journal of Colloid and Interface Science 274 (2004) 371379 www.elsevier.

com/locate/jcis

Adsorption of Congo red from aqueous solution onto calcium-rich y ash


Bilal Acemio lu g
Department of Chemistry, Faculty of Science and Arts, Kahramanmaras Stc Imam University, Kahramanmaras 46100, Turkey Received 28 July 2003; accepted 9 March 2004

Abstract The adsorption of Congo red from solution was carried out using calcium-rich y ash with different contact times, concentrations, temperatures, and pHs. While the amount of dye adsorbed per unit weight of y ash increases with increasing concentration and temperature, it decreases slightly with increasing pH. The adsorption was between 93 and 98% under the conditions studied. Kinetic studies showed that the adsorption process obeyed the pseudo-second-order kinetic model. It was also determined that the adsorption isotherm followed Freundlich and DubininRadushkevich models. From thermodynamic studies, it was seen that the adsorption was spontaneous and endothermic. Desorption studies suggested that desorption was 29.18% in the presence of 0.1 N HCl and was 47.21% in the presence of CH3 COOH (50% v/v). This indicated that most of the dye was held by y ash via chemisorption as well as ion exchange. Furthermore, FTIR study also shows that a chemisorption process occurs between CR and y ash, probably indicating dye/y ash complexing. 2004 Elsevier Inc. All rights reserved.
Keywords: Adsorption; Kinetics; Adsorption isotherm; Adsorption thermodynamics; Desorption; Dye; Congo red; FTIR spectra

1. Introduction Wastewaters from the textile, cosmetics, printing, dying, food coloring, and papermaking industries are polluted by dyes. These colored efuents can be mixed in surface water and ground water systems, and then they may also transfer to drinking water. Dyes can cause allergic dermatitis, skin irritation, cancer, and mutations [1]. Therefore, it is necessary to remove the dye pollutions. For this purpose, many methods, such as activated carbon adsorption, chemical oxidation, reverse osmosis, coagulation and occulation, and biological treatments, have been developed for treating dyecontaining wastewater. The advantages and disadvantages of each technique have been extensively reviewed [2]. Of these methods, activated carbon adsorption is highly effective for the removal of dyes and pigments, as well as other organic and inorganic pollution. But the use of activated carbon is not suitable for developing countries because of its high cost. The use of low-cost adsorbents such as clay minerals
* Fax: +90-344-2512308.

E-mail address: bilacem@hotmail.com. 0021-9797/$ see front matter 2004 Elsevier Inc. All rights reserved. doi:10.1016/j.jcis.2004.03.019

[35], y ash [69], peat [10,11], wood powder [12,13], bark [14,15], and lignin [16] is more suitable. A large number of low-cost adsorbents have been treated for dye removal. For example, Asfour et al. have studied the adsorption of basic dye Astrazone blue FRR 69 on hardwood (Beech) sawdust [13]. Dogan and Alkan have investigated the kinetics of methyl violet adsorption onto perlite [2]. Liversidge et al. have studied the removal of Basic blue 41 dye by linseed cake [17]. Annadurai et al. have studied the adsorption of various dyes onto cellulose-based wastes [18]. Gupta et al. have studied the removal of Metomega Chrome Orange GL by y ash [19]. Moreover, Kannan and Meenakshsundaram have also studied the removal of Congo red by adsorption onto various activated carbons prepared from raw materials such as bamboo dust, coconut shell, groundnut shell, rice husk, and straw [20]. In another work, the removal of Congo red from aqueous solution by biogas waste slurry has been studied by Namasivayam and Yamuna [21]. The goal of this work is to study the adsorption of Congo red (CR) from aqueous solution by calcium-rich type C Turkish y ash. The effects of treatment time, initial dye concentration, pH and temperature on the adsorption are investigated in detail.

372

B. Acemio lu / Journal of Colloid and Interface Science 274 (2004) 371379 g

Table 1 Chemical composition of y ash Constituent SiO2 Al2 O3 CaO Fe2 O3 MgO Na2 O K2 O TiO2 SO3 Cra Mna Nia Pba Zna Cua Cda Cab Mgb Sib Alb Nab Kb
a The values measured as mg L1 . b Soluble concentration (mg L1 ).

Table 2 Some physical properties of y ash Percent weight 15.14 7.54 23.66 3.30 4.50 0.57 0.28 1.03 13.22 298.00 219.00 119.00 80.00 80.00 40.00 8.00 280.00 3.60 3.10 0.20 4.00 1.00 pH (in water) Particle size (mesh) Bulk density (g/cm3 ) Specic surface area (m2 /g) Specic gravity (g/cm3 ) pHZPC LOI 12.50 200 1.05 0.342 2.70 7.0 2.31

Fig. 1. Structure of Congo red.

received without any pretreatment in the adsorption experiments. 2.1.2. Congo red Congo red (CR), an anionic disazo direct dye, contains NH2 and SO3 functional groups. It has a maximum absorbency at wavelength 500 nm on a UVvis spectrophotometer. The color of Congo red changes from dark blue at pH 24 to red to pH 12. However, the degree of red color is different from the original red at pH 1012. Congo red has a molecular weight of 696.7 g mol1 . The structure of this dye is shown in Fig. 1. 2.2. Method Congo red was received from Merck (C.I. 22120) and used without further purication. All of the CR solutions were prepared with distilled water. For the adsorption experiments, 10 g of y ash sample was added into a liter of CR solution under each condition. The pH of dye solution was adjusted with 0.1 N NaOH and HCl solutions using a pH meter. The mixture was continuously agitated by a magnetic stirrer to reach equilibrium under experimental conditions desired. At the desired treatment time, the samples were pipetted from the reactor by means of a very thin point micropipette, which prevent the transition to solution of y ash particles [2]. The absorbencies of pipetted samples were measured using a Shimadzu UVvis 160 A spectrophotometer at wavelength 500 nm, which is a maximum absorbency wavelength. Then the concentrations of the samples were determined by using a standard curve. The amounts of CR adsorbed onto y ash were calculated by subtracting the nal solution concentration from the initial concentration of dye solutions. 2.2.1. Desorption studies The adsorbent utilized for the adsorption of an initial dye concentration of 7.17 105 M was separated from the

2. Materials and methods 2.1. Materials 2.1.1. Fly ash Particles captured on cyclones and electrostatic precipitators risen from coal and lignite burned in thermal power plants producing electrical energy are named y ash. The main components of y ash are alumina, silica, calcium oxide, iron oxide, and residual carbon [2224]. However, the properties and compositions of y ash from coal and lignite are different from each other. While the y ash from burning coal contains a large quantity of SiO2 , a small quantity of Al2 O3 , and a very small quantity of CaO, the y ash from burning lignite contains a large quantity of SiO2 and CaO, and a very small quantity of SO3 [22]. According to ASTM standards, the y ash is divided into two types [22]. If the sum of SiO2 , Al2 O3 , and Fe2 O3 is 70%, it is named type F. If the sum is 50%, it is named type C. The fact that the sum is in small quantities in type C y ash is due to high quantities of CaO [23]. Fly ash has been used in construction, agriculture, metal recovery, and water pollution control [24]. The lignite y ash used in this study was collected from the lignite-burning thermal power plant of Afsin-Elbistan, Kahramanmaras, Turkey (southeastern Anatolia). AfsinElbistan y ash is a type C ash with 23.66% CaO, and its constitution and some physical properties have been presented in Tables 1 and 2 [23]. Herein, the y ash particles were passed through a 200-mesh sieve, and it was used as

B. Acemio lu / Journal of Colloid and Interface Science 274 (2004) 371379 g

373

dye solution. The dye-loaded adsorbent was washed gently with water to remove any unadsorbed dye. Then dye-loaded y ash was stirred using a magnetic stirrer with 50 ml of distilled water at different pHs, 0.1 N HCl solution, and CH3 COOH solution (50% v/v), one by one. Desorbed dye was determined as mentioned before. 2.2.2. FTIR measurements Fly ash, Congo red, and air-dried dye-loaded y ash samples after adsorption were brought to constant weight in a drying oven at 60 C for 12 h. Then these samples were kept in the desiccator over calcium chloride. Afterward, pellets with 100 mg of KBr and 1.50 mg of samples were made. For pellets obtained, the infrared spectra in the range 4000400 cm1 were recorded on a Shimadzu FTIR-8000 spectrophotometer.

Fig. 2. Effect of initial pH of dye solution on adsorption of CR onto y ash for initial dye concentration of 7.17 105 M at 20 C.

3. Results and discussion 3.1. Effect of contact time on adsorption The effect of contact time on the amount of CR adsorbed per unit of adsorbent was investigated under all the experimental conditions, i.e., concentration, temperature, and pH. Very rapid adsorption is observed at 2 min, and thereafter a gradual increase in adsorption occurs with increasing contact time up to 50 min, after which a maximum value of adsorption is attained. After this time, the amount of dye adsorbed was not signicant: from time to time, very small decreases in the amount of the dye adsorbed were observed, indicating desorption. Therefore, the time of 50 min is xed as the optimum contact time. Similar results have been reported for the adsorption of CR on various activated carbons in work done by Kannan and Meenakshsundaram [20], and for the adsorption of an acid dye from aqueous solution by chitin in another work conducted by Annadurai and Krishnan [25]. 3.2. Effect of pH on adsorption The initial pH values of dye solutions affect the chemistry of both a dye molecule and an adsorbent. In this study, blank studies for CR were done at various pHs. The results obtained show that the molecular form of CR in solution medium changes markedly in the pH range 24, and at a high alkaline pH of 12, e.g., the color of CR changes from dark blue at pH 24 to red at pH 12. However, the red color is different from the original red in the pH range 1012. Therefore, in this section, the solution pHs were kept between 5 and 10. On the other hand, y ash used as an adsorbent has a different structure in acidic and alkaline media. For example, hydroxyl groups (e.g., silanol groups) risen from interaction with solvent (generally, water or solvent of dye solution) of inorganic groups on the y ash become more negative in alkaline media, and more positive in acidic media [23]. And

therefore, the adsorption degree changes as a function of interaction between adsorbent and adsorbate dyes at different pHs. In our study, pHZPC for y ash is 7 [23]. Therefore, the surface of the adsorbent will be positively charged below pHZPC , and beyond this it will be negatively charged. Since the CR is an anionic dye, maximum adsorption is carried out below pHZPC . Above pHZPC , a decrease in adsorption takes place due to repulsion between anionic dye molecules and negatively charged adsorbent surface. Fig. 2 illustrates effect of initial pH of dye solution on adsorption for initial dye concentration of 7.17 105 M as a function of contact time at 20 C. The amounts of dye adsorbed per unit y ash are 6.92 106 , 6.89 106 , 6.87 106 , and 6.80 106 mol g1 at pH values of 5, 6, 8, and 10, respectively. As also shown in Fig. 2, it is obvious that the amount of dye adsorbed with the rise in initial pH (i.e., above pHZPC 7) decreases slightly. The slight decrease in adsorption may be attributed to the degree of repulsion between anionic dye molecules and the negatively charged adsorbent surface. On the other hand, in Fig. 2, it is seen that the amount of the dye adsorbed with a decrease in initial pH (i.e., below pHZPC 7) increases signicantly. The fact that the increase in the adsorption of the dye onto y ash do not affect signicantly with changing of initial acidic pH of the dye solution is due to the alkaline character of y ash which has high calcium content. And thus, the y ash can neutralize the acidic pH of the dye solution, decreasing signicantly the effect of pH on the adsorption. Similar results have been reported for the removal of heavy metals by Australian y ash [24]. From the adsorbed amount at equilibrium, it is seen that the most effective pH for the adsorption of CR onto y ash is 5. And therefore, further experiments were conducted at pH 5. 3.3. Effect of initial dye concentration on adsorption Four different concentrations, i.e., 3.58 105 (25 mg dye L1 ), 7.17 105 (50 mg dye L1 ), 1.07 104 (75 mg

374

B. Acemio lu / Journal of Colloid and Interface Science 274 (2004) 371379 g

Fig. 3. Effect of initial dye concentration on the adsorption of CR onto y ash at pH 5 and 20 C. (E) 3.58 105 M; (1) 7.17 105 M; (P) 1.07 104 M; () 1.43 104 M.

Fig. 4. Effect of temperature on adsorption of CR onto y ash for initial dye concentration of 7.17 105 M at pH 5.

dye L1 ), and 1.43 104 M (100 mg dye L1 ), are selected to investigate the effect of initial dye concentration on the adsorption of CR onto y ash, and the results obtained at 20 C (room temperature) and pH 5 are shown in Fig. 3. As shown in Fig. 3, with increasing initial dye concentration from 3.58 105 to 1.43 104 M, the amount of dye adsorbed by y ash increases 3.34 106 to 1.38 105 mol g1 . Moreover, the percent removal of the dye was also increases 93.30 to 96.50 with increasing initial dye concentration from 3.58 105 to 1.43 104 M. The increase in the proportion of removed dye may be probably due to equilibrium shift during adsorption process. 3.4. Effect of temperature on adsorption In order to discuss the adsorption of Congo red onto y ash at different temperatures, it is important to know how to change the solubility of dye with temperature. Before adsorption, therefore, the solubility of Congo red at 2060 C was tested, and it was easily seen to be solubilized. Furthermore, it has been reported that Congo red has a very high solubility, 50100 g/L at 80 C [26]. Herein, the effect of temperature on adsorption was studied at 20, 35, 50, and 60 C, and the results were shown in Fig. 4. As shown in Fig. 4, the results indicate that the amount of dye adsorbed by y ash increases with an increase in temperature up to 50 C, and then a slight decrease in adsorption is observed. For example, for initial dye concentration of 7.17 105 M, when increasing initial solution temperature from 20 to 50 C, the amount of dye adsorbed per unit weight of y ash increases 6.92 106 to 7.08 106 mol g1 . At a temperature of 60 C, it was 6.99 106 mol g1 . Maximum percent removal by y ash was 98.74 at 50 C. The fact that the adsorption increases with an increase in temperature indicates the increase in the mobility of the large dye ions with increasing temperature. Moreover, increasing temperature may produce a swelling effect within the internal structure of the y ash, penetrating the large dye molecule

Fig. 5. Freundlich isotherm of CR onto y ash at pH 5 and 20 C.

further. Similar results were reported for the adsorption of Astrazone blue, a basic dye, onto hardwood sawdust [13] and for the adsorption of methylene blue onto unexpanded perlite [27]. 3.5. Adsorption isotherm The adsorption equilibrium data were tted for Langmuir, Freundlich, and DubininRadushkevich (DR) isotherms. The isotherm results indicate that the adsorption of CR onto y ash is consistent with the Freundlich and Dubinin Radushkevich isotherms. The Freundlich adsorption isotherm, which assumes that adsorption takes place on heterogeneous surfaces, can be expressed [28] as ln qe = ln kf + (1/n) ln Ce , (1)

where qe is the amount of dye adsorbed at equilibrium time (mol g1 ), Ce is the equilibrium concentration of the dye in solution (mol L1 ). kf and n are isotherm constants which indicate the capacity and intensity of the adsorption, respectively. Fig. 5 shows the plot of ln qe vs ln Ce at 20 C. The plot is in harmony with Freundlich adsorption model with a correlation coefcient of 0.99. The values of kf and n were

B. Acemio lu / Journal of Colloid and Interface Science 274 (2004) 371379 g

375

Fig. 6. DubininRadushkevich isotherm of CR onto y ash at pH 5 and 20 C. Table 3 Isotherm constants for the adsorption of Congo red onto y ash at pH 5 and 20 C Freundlich isotherm kf n (g/l) = 0.51 r 2 = 0.996 (mol/g) = 4.47 105 DubininRadushkevich isotherm Xm (mol/g) = 1.72 K (mol2 kJ2 ) = 1.32 102 r 2 = 0.996

Fig. 7. Isotherm of CR adsorption onto y ash at pH 5 and 20 C.

calculated from the slope and intercept of the plot of ln qe vs ln Ce . The DR isotherm assumes no homogeneous surface of the adsorbent. The DR equation can be expressed [29] as qe = Xm exp(K 2 ), (2)
Fig. 8. Vant Hoff plot of CR adsorption onto y ash for initial dye concentration of 7.17 105 M at pH 5.

where (Polanyi potential) is equal to RT ln(1 + 1/Ce ), qe is the amount of the dye adsorbed per unit y ash (mol g1 ), Xm is the adsorption capacity (mol g1 ), Ce is the equilibrium concentration of the dye in solution (mol L1 ), K is the constant of the adsorption energy (mol2 kJ2 ), R is the gas constant (kJ/mol K), and T is the temperature (K). The linear form of the DR isotherm is ln qe = ln Xm K 2 . (3) The plot of ln qe vs 2 at 20 C is presented in Fig. 6. The plot is in harmony with the DR adsorption model with a correlation coefcient of 0.99. The values of Xm and K were calculated from the intercept and slope of this plot. The constants obtained for Freundlich and DR isotherms are shown in Table 3. Moreover, the original adsorption isotherm (the graph of qe vs Ce ) is also presented in Fig. 7. As shown in Fig. 7, it can be mentioned that monolayer coverage does not occur on the heterogeneous surface of adsorbent. This situation is attributed that various active sites on y ash has different afnities to CR molecule. 3.6. Adsorption thermodynamics In understanding better the effect of temperature on the adsorption, it is important to study the thermodynamic parameters such as standard Gibbs free energy change G0 ,

standard enthalpy H 0 , and standard entropy S 0 . The Gibbs free energy of adsorption by using equilibrium constant (Kc ) is calculated from the following equation: G0 = RT ln Kc . (4) Standard enthalpy, H 0 , and standard entropy, S 0 , of adsorption can be estimated from vant Hoff equation given in
0 Hads S0 + , (5) RT R where R is the gas constant, Kc is adsorption equilibrium constant. The Kc value is calculated from the equation [6]

ln Kc =

Kc = CAe /CSe ,

(6)

where CAe is the equilibrium concentration of the dye ions on adsorbent (mg L1 ) and CSe is the equilibrium concentration of the dye ions in the solution (mg L1 ). The plot of ln Kc against 1/T (in Kelvin) should be lin0 ear. The slope of the vant Hoff plot is equal to Hads /R, and its intercept is equal to S 0 /R. The vant Hoff plot for the adsorption of CR onto y ash is given in Fig. 8. Thermodynamic parameters obtained are given in Table 4. As shown in the table, the negative values of G0 at different temperatures indicate the spontaneous nature of the adsorption

376

B. Acemio lu / Journal of Colloid and Interface Science 274 (2004) 371379 g

Table 4 Thermodynamic parameters for the adsorption of Congo red onto y ash Temperature ( C) 20 35 50 60 Kc 19.48 30.17 78.67 38.83 G0 (kJ/mol) 7.233 8.723 11.722 10.094 H0 (kJ/mol) 27.13 S0 (kJ/mol K) 0.118

process. Positive H 0 reveals endothermic adsorption. The positive value of S 0 suggests the increased randomness at the solid/solution interface during the adsorption of the dye onto y ash. A similar trend has been reported for the adsorption of Congo red onto coir pith carbon [30]. 3.7. Adsorption kinetics Several models have been proposed to express the adsorption mechanism of solute molecules onto an adsorbent: (a) pseudo-rst-order kinetic model, (b) intraparticle diffusion model, and (c) pseudo-second-order kinetic model. A pseudo-rst-order kinetic model of Lagergen [31] is given by log(qe qt ) = log qe k1 t, (7) 2.303 an intraparticle diffusion model of Weber and Morris [32] is shown as qt = ki t 1/2 , (8)

and a pseudo-second-order kinetic model of Ho and McKay [10,33] is 1 1 t = + t, 2 qt k2 qe qe (9)

where k1 is the rate constant for the pseudo-rst-order model, k2 is the rate constant for the pseudo-second-order model and ki is intraparticle diffusion rate constant. qe and qt are the amount of solute adsorbed per unit adsorbent at

equilibrium and any time, respectively. Herein, the initial ad2 sorption rate is h = k2 qe . The y ash/CR adsorption system in the present study was investigated in terms of the above-mentioned kinetic models for understanding the adsorption kinetics. First, the plots of log(qe qt ) vs t for the pseudo-rstorder model given in Eq. (7) for the adsorption of CR onto the y ash were drawn at different initial concentrations and solution temperatures. From the linear regression analysis obtained, it is determined that the values of r 2 are 0.64, 0.95, 0.81, and 0.58 for initial concentrations of 3.58 105, 7.17 105 , 1.07 104 , and 1.43 104 M at 20 C, respectively. Meanwhile, in the cases of the temperature effect, it is found that the values of r 2 are 0.85, 0.96, 0.93, and 0.89 for 20, 35, 50, and 60 C, respectively, at initial dye concentration of 7.17 104 M. Also, the values of qe from the pseudo-rst order kinetics are not in agreement with experimental data, qe (see Table 5). These plots obtained for the pseudo-rst-order model above are not shown, due to their lower correlation coefcients herein. Second, the plots of qt vs t 1/2 for intraparticle diffusion model given in Eq. (8) were obtained at different initial dye concentrations and solution temperatures. The r 2 values are found to be 0.65, 0.77, 0.59, and 0.39 for initial dye concentrations of 3.58 105, 7.17 105 , 1.07 104 , and 1.43 104 M at 20 C, respectively. And it is estimated that the values of r 2 are 0.75, 0.77, 0.68, and 0.63 for 20, 35, 50, and 60 C, respectively. No plots for the intraparticle diffusion model are also demonstrated due to their lower correlation coefcients. Then, the linear plots of t/qt vs t for the pseudo-secondorder model in Eq. (9) were obtained at different initial dye concentrations and solution temperatures. As shown in Fig. 9, the values of r 2 obtained for the linear plots of t/qt vs t are determined to equal 1 for the other initial dye concentrations (except for 1.07 104 M, which has a correlation of 0.9998) at an equilibrium contact time of 50 min.

Table 5 Comparison of kinetic parameters for the adsorption of Congo red onto y ash at various concentrations and pH 5 Cinit a 105 3.58 7.17 10.70 14.30
a b c d e f g h i j k

qe b 106 3.34 6.83 10.29 13.81

q2 c 106 3.34 6.90 10.30 13.81

k2 d 105 17.20 2.77 7.02 13.66

he 105 1.92 1.32 7.44 26.05

2 r2 f

q1 g 0.16 0.55 0.45 0.20

k1 h 0.0578 0.1680 0.0829 0.0960

2 r1 i

ki j 0.0414 0.2271 0.1165 0.0856

ri2 k 0.6569 0.7768 0.5959 0.3917

1 0.9998 1 1

0.6408 0.9607 0.8101 0.5804

Initial dye concentration (mol/l). Equilibrium sorption capacity obtained as experimental (mol/g). Equilibrium sorption capacity obtained from the pseudo-second-order equation (mol/g). The rate constant of the pseudo-second-order reaction (g/mol min). The initial sorption rate from the pseudo-second-order kinetics (mol/g min). Correlation coefcient from the pseudo-second-order equation. Equilibrium sorption capacity obtained from the pseudo-rst-order equation (mol/g). The rate constant of the pseudo-rst-order reaction (1/min). Correlation coefcient from the pseudo-rst-order equation. Intraparticle diffusion rate constant (mol/g min0.5 ). Correlation coefcient from intraparticle diffusion equation.

B. Acemio lu / Journal of Colloid and Interface Science 274 (2004) 371379 g

377

Fig. 11. Effect of pH on desorption of CR. Fig. 9. Effect of initial dye concentration on the pseudo-second-order kinetics of CR adsorption onto y ash at pH 5 and 20 C.

in Table 5. As shown in Table 5, with increasing concentration from 3.58 105 to 1.43 104 M at 20 C, the rate constant of pseudo-second-order, k2 , decreases from 17.20 105 to 13.66 105 g/mol min while the initial adsorption rate, h, increases from 1.92 105 to 26.05 105 mol/g min. Similar trend has also been reported by Ho and McKay [10] for the adsorption of basic dyes onto peat. 3.8. Desorption of Congo red Desorption process of the dye can help the researcher in elucidating the mechanism of an adsorption process. If the dye adsorbed onto the adsorbent can be desorbed by water, it can be said that the attachment of the dye onto the adsorbent is by weak bonds. If the acid, such as H2 SO4 and HCl, or higher alkaline can desorb the dye, it can be said that the attachment of the dye onto the adsorbent is by ion exchange. If organic acids, for example, such as CH3 COOH, can desorb the dye, it can be said that the adsorption of the dye onto the adsorbent is by chemisorption [1]. Herein, dye-loaded y ash was stirred with 50 ml of various alkaline waters, hydrochloric acid (0.1 N), and acetic acid (50% v/v), for 20 min, one by one. As shown in Fig. 11, a very low desorption was seen with various alkaline waters. Very low desorption of CR indicates that chemical activation or chemisorption might be taken place between active sites of y ash and functional groups of CR. In the case of 0.1 N HCl, while the desorption % was 29.18, the desorption % with CH3 COOH was 47.21. A desorption of 29.18% with HCl probably indicates an ion exchange mechanism. Furthermore, the fact that the most desorption occurred with CH3 COOH, an organic acid, conrms that adsorption of CR onto y ash carries out signicantly via a chemisorption mechanism. A similar result has also been reported for removal of Rhodamine B from aqueous solution by biogas waste slurry [35]. 3.9. FTIR study of CR adsorption Fourier transform infrared spectroscopy (FTIR) was used to examine the surface groups of the sorbents and to iden-

Fig. 10. Effect of temperature on the pseudo-second-order kinetics of CR adsorption onto y ash for initial dye concentration of 7.17 105 M at pH 5.

Also, it is found that r 2 values are >0.999 for all the studied temperatures, and the plots drawn for all the temperatures are overlapped due to negligible changes in the percent adsorption (1.025%) when increasing the temperature (Fig. 10). As a result, r 2 values obtained from pseudo-secondorder kinetic model are greater than those of other rate laws. Also, as shown in Table 5, the values of qe from the pseudo-second-order kinetics are in agreement with experimental data, qe , and the values of r 2 are also higher than 0.9998. These indicate that the adsorption perfectly complies with pseudo-second-order reaction and an activated sorption mechanism [10]. Similar kinetic results have also been reported for the adsorption of Congo red onto Aspergillus niger [34] and coir peat carbon [30] and of Basic blue 69 and Acid blue 25 onto peat [10]. The values of the rate constants and initial adsorption rates for the adsorption of CR onto the y ash were calculated from the slopes and intercepts of the pseudo-secondorder plots, respectively. The data obtained are presented

378

B. Acemio lu / Journal of Colloid and Interface Science 274 (2004) 371379 g

Fig. 12. FTIR spectra of y ash, CR, and dye-loaded y ash.

tify those groups responsible for dye adsorption. Adsorption in the IR region takes places because of the rotational and vibrational movements of the molecular groups and chemical band of a molecule. The two fundamental vibrations are stretching, where the atoms stay in the same bound axis but the distance between atoms increases or decreases, and deformation, where the positions of the atoms change relative to the original bound axis. FTIR spectra of CR on y ash before and after the adsorption are presented in Fig. 12. If is noticed the chemical structure of CR, it can be seen that an interaction occurs between active sites (between 613 and 1662 cm1 ) on y ash and SO3 and N=N groups on CR. As shown in Fig. 12, the strong bands in the 1062.7, 1178.4, and 1266.6 cm1 regions, attributed to S=O stretching [36], are diminished after adsorption with y ash. At the same time, the band at 1610.5 cm1 , attributed to N=N stretching [36,37], is also diminished after adsorption. Furthermore, the bands at 1450.4 and 1685.4 cm1 , assigned to aromatic skeletal vibrations [36], have been shifted, broadened, and reduced after adsorption. On the other hand, strong SiO bands at 8301130 cm1 [36] are markedly reduced and slightly shifted to higher frequency as a function of chemical interaction of CR with y ash. Also, the stretching absorption band of OH in the crystal structure of the adsorbent seemed at 3644.4 cm1 is diminished after adsorption with CR. All these ndings suggest that CR on y ash is held by chemical activation or chemisorption, probably indicating y ash/dye complexation.

References
[1] C. Namasivayam, N. Muniasamy, K. Gayatri, M. Rani, K. Ranganathan, Bioresour. Technol. 57 (1996) 37. [2] M. Dogan, M. Alkan, Chemosphere 50 (2003) 517.

[3] P. Nigam, G. Armour, I.M. Banat, D. Singh, R. Marchant, Bioresour. Technol. 72 (2000) 219. [4] R. Apak, K. Guclu, M.H. Turgut, J. Colloid Interface Sci. 243 (2001) 81. [5] G. Bereket, A.Z. Aroguz, M.Z. Ozel, J. Colloid Interface Sci. 187 (1997) 338. [6] K.P. Yadava, B.S. Tyagi, V.N. Singh, J. Chem. Technol. Biotechnol. 51 (1991) 47. [7] K.K. Panday, G. Prasad, V.N. Singh, Water Res. 19 (1985) 869. [8] T. Viraraghavan, K.R. Ramakrishna, Water Quality Res. J. Can. 34 (1999) 505. [9] G.P. Damasmahapatra, T.K. Pal, A.K. Bhadra, B. Bhattacharya, Sep. Sci. Technol. 31 (1996) 2001. [10] Y.S. Ho, G. McKay, Chem. Eng. J. 70 (1998) 115. [11] R. Gundogan, B. Acemioglu, M.H. Alma, J. Colloid Interface Sci. 269 (2004) 303. [12] Y.S. Ho, G. McKay, Pros. Safety Environ. Prot. 76 (1998) 183. [13] H.M. Asfour, O. Fadali, M.M. Nassar, M.S. El-Geundi, J. Chem. Technol. Biotechnol. A 35 (1985) 21. [14] M. Villaescusa, N. Miralles, J. Chem. Technol. Biotechnol. 75 (2000) 812. [15] A.M. Marganda, B.C.G. Ganzales, C.R. Guedes, Water Res. 27 (1993) 1333. [16] B. Acemioglu, A. Samil, M.H. Alma, R. Gundogan, J. Appl. Polym. Sci. 89 (2003) 1537. [17] R.M. Liversidge, G.J. Lloyd, D.A.J. Wase, C.F. Forster, Process Biochem. 32 (6) (1997) 473. [18] G. Annadurai, R.S. Juang, D.J. Lee, J. Hazard. Mater. B 92 (2002) 263. [19] G.S. Gupta, G. Prasad, V.H. Singh, J. Environ. Sci. Health A 23 (3) (1988) 205. [20] N. Kannan, M. Meenakshsundaram, Water Air Soil Pollut. 138 (2002) 289. [21] C. Namasivayam, R.T. Yamuna, J. Chem. Technol. Biotechnol. 53 (2) (1992) 153. [22] H. Ozkul, S. Koral, A. Bakoca, Afin-Elbistan uuu kllerinin tu la s s g blok eleman retiminde kullanlma olanaklarnn aratrlmas, s Istanbul Teknik University, Building Faculty, 1995, Istanbul, Turkey, in Turkish. [23] B. Bayat, J. Hazard. Mater. B 95 (2002) 251. [24] J. Ayala, F. Blanco, P. Garcia, P. Rodriguez, J. Sancho, Fuel 77 (11) (1998) 1147. [25] G. Annadurai, M.R.V. Krishnan, Indian J. Chem. Technol. 4 (1997) 217.

B. Acemio lu / Journal of Colloid and Interface Science 274 (2004) 371379 g

379

[26] Bumjin Industrial Co. Ltd., Korea. [27] M. Dogan, M. Alkan, Y. Onganer, Water Air Soil Pollut. 120 (2000) 229. [28] F.S.C. Anjos, E.F.S. Vieira, A.R. Cestari, J. Colloid Interface Sci. 253 (2002) 243. [29] O. Ceyhan, D. Baydas, Turk. J. Chem. 25 (2001) 193. [30] C. Namasivayam, D. Kavita, Dyes Pigments 54 (2002) 47. [31] S. Lagergen, Kungl. Svenska Vetenskapsakad. Handl. 24 (1898) 139.

[32] [33] [34] [35] [36]

W.J. Weber, J. Sanit. Eng. Div. ASCE 89 (SA2) (1963) 31. Y.S. Ho, G. McKay, Process Biochem. 34 (1999) 451. Y. Fu, T. Viraraghavan, Adv. Environ. Res. 7 (2002) 239. C. Namasivayam, R.T. Yamuna, Water Air Soil Pollut. 65 (1992) 101. R.M. Silverstein, G.C. Bassler, T.C. Morrill, Spectrometric Identication of Organic Compounds, Wiley, New York, 1991. [37] D.A. Skoog, Principles of Instrumental Analysis, fourth ed., Saunders, Fort Worth, TX, 1992.

You might also like