You are on page 1of 11

Removal of Reactive Textile Dyes (Remazol Brillant Blue R and Remazol Yellow) by Surfactant-Modied Natural Zeolite

Ays Kuleyin and Fulya Aydin e Environmental Engineering Department, Ondokuzmayis University, 55139 Samsun, Turkey; akuleyin@omu.edu.tr (for correspondence)
Published online 23 April 2010 in Wiley Online Library (wileyonlinelibrary.com). DOI 10.1002/ep.10454

In this study, surfactant-modied zeolite (SMZ) was used to remove Remazol Brillant Blue R and Remazol Yellow reactive dyes from aqueous solutions. The inuences of suspension pH, temperature, agitation rate, and the SMZ dosage on reactive dyes removal and adsorption capacity were investigated by conducting a series of batch adsorption experiments. The adsorption kinetics was tested for pseudo-rstorder, pseudo-second order, intraparticle diffusion model, Elovich, and Bangham models, and rate constants of the kinetic models were calculated. Equilibrium isotherms for the adsorption of reactive dyes were analyzed by the Freundlich, Langmuir, DubininRadushkevich, and Tempkin isotherm models. The Langmuir monolayer adsorption capacities of Remazol Brillant Blue R and Remazol Yellow were estimated as 13.9 and 38.31 mg/g, respectively. The adsorption rate was rapid; more than half of the adsorbed Remazol Brillant Blue R was removed in the rst 60 min and of Remazol Yellow in the rst 30 min. Thermodynamic parameters such as DH0, DS0, and DG0, at 208C, were found to be 5.2126 kJ/mol, 0.0273 kJ/mol K, and 22.7969 kJ/mol (Remazol Brillant Blue R), and 29.9747 kJ/mol, 0.10875 kJ/mol K, and 21.8900 kJ/mol (Remazol Yellow), respectively.
2010 American Institute of Chemical Engineers

2010 American Institute of Chemical Engineers Environ Prog, 30: 141151, 2011

Keywords: zeolite, adsorption, isotherm

modication,

reactive

dye,

INTRODUCTION

Dyes are among the major constituents of the wastewater produced by many industries related not only to textiles, paints and varnishes, inks, plastics, pulp and paper, cosmetics, and tanning but also of those that produce dyes. Colored dye efuents pose a major threat to the ecosystem. Many of the dyes are extremely toxic [1, 2]. The presence of color in water reduces its aquatic diversity by blocking the passage of light through it. In some cases, less than 1 ppm of dye concentration produces considerable water coloration. Consequently, light penetration in highly colored wastewater efuents is greatly reduced. The disposal of such wastes into receiving waters signicantly affects photosynthetic activity in aquatic life and causes damage to the environment [35]. Both biological and physical/chemical methods have been used for color removal from textile efuents. The former methods have not been very successful because of the essentially nonbiodegradable nature of most dyes [6]. The physical/chemical methods
July 2011 141

Environmental Progress & Sustainable Energy (Vol.30, No.2) DOI 10.1002/ep

that have been proven to be successful are adsorption [7], chemical oxidation [8], coagulation/occulation [9], membrane ltration [10, 11], and electrochemical methods [12, 13]. Adsorption, extensively used in industrial processes for a variety of separation and purication purposes, is rapidly becoming a prominent method of treating aqueous efuents. The removal of color from textile wastewater is considered an important application of the adsorption process using low-cost adsorbents such as shale oil ash [14], chitosan [15], y ash [16], bagasse y ash [17], waste red mud [18], sawdust [19], natural clay [2022], and zeolite [2325] in preference to expensive adsorbents such as activated carbon and polymer resins [24, 26]. Zeolites are hydrated aluminosilicate materials with cage-like structures having internal and external surface areas extending up to several hundred square meters per gram and cation exchange capacities of up to several milliequivalents per kilogram. At least 41 types of natural zeolites are known to exist, and many others have been synthesized. Both natural and synthetic zeolites are used in industry as adsorbents, soil modiers, ion exchangers, and molecular sieves [27]. The presence of 4.5 million tons of natural zeolites of high quality, mainly of clinoptilolite in Turkey, created an impetus for its utilization in wastewater treatment [25]. The typical unit cell formula of natural zeolite mineral, clinoptilolite, is stated to be Na6[(AlO2)6 (SiO2)30]24 H2O [28]. Natural zeolites possess permanent negative charges in their crystal structures, making them suitable for modication by cationic surfactants. Recent studies on the properties of surfactant-modied zeolite (SMZ) indicate that it is an effective sorbent for a variety of contaminants [2932]. SMZ can remove organic compounds and oxyanions from water. Partitioning is the mechanism responsible for organic adsorption by SMZ; while adsorption of transition metal cations remains generally unaffected [33]. Although the adsorption of oxyanions by SMZ was attributed to anion exchange on its positively charged surfactant bilayer, the adsorption of hydrophobic organic contaminants was due to partitioning of the organics in the organic phase created by surfactant tail groups [34]. The SMZ is stable in water and aggressive chemical solutions. Its relatively low cost makes it a viable alternative to other reactive media as activated carbon and ion exchange resins [35]. The aim of this study was to investigate the removal of two reactive dyes, namely Remazol Blue R and Remazol Yellow, from water by SMZ and of the effects of experimental parameters such as adsorbent dose, initial concentration, contact time, and temperature on the adsorption of reactive dyes.
MATERIALS AND METHODS

Figure 1. SEM photography of Manisa-Gordes Zeolite (magnication: 325,000).

Gordes region of Turkey. Zeolite was selected as an adsorbent for its relatively moderate surface area and exceptionally high selective ion-exchange capacity. Gordes zeolite has the following properties: a cationexchange capacity of 1.92.2 meq/g, pore diameter of 4 A, purity of 92%, bed porosity of 40%, density of 2.15 g/mL, apparent density of 1.30 g/mL, and suspension pH of 7.57.8 [2]. The surface area of this particular zeolite was 11.80 m2/g [2]. A SEM micro graph of natural Manisa-Gordes zeolite is shown in Figure 1. Zeolitized tuffs have a drusy texture and very high microporosity. The XRD spectrum of natural zeolite and SMZ is shown in Figure 2. Chemical composition of natural zeolite is given in Table 1.

Adsorbent The zeolite (clinoptilolite) samples used in this study were received from Incal Company in the
142 July 2011

Adsorbate The dyes (Remazol Brilliant Blue R and Remazol Yellow) marketed by Dyestar are of the reactive type. Our adsorption experiments were conducted using clinoptilolite and these two reactive dyes. The structural formulas of Remazol Brillant Blue R and Remazol Yellow are given as examples in Figure 3. The equilibrium concentrations of the dyes were determined at 590 and 413 nm for Remazol Brilliant Blue R and Remazol Yellow, respectively, using a Hitachi UVvis spectrophotometer. The calibration curves for each dye at the respective wavelengths were established as a function of dye concentration. Remazol Brilliant Blue R and Remazol Yellow dyes were used in the preparation of the synthetic adsorbate solutions. An aqueous solution of 1000 mg/L concentration was prepared by mixing appropriate amounts of the two dyes with distilled water. Fresh stock solution was prepared as required immediately before use in the experiments. The initial concentration was ascertained before the start of each experimental run.

Environmental Progress & Sustainable Energy (Vol.30, No.2) DOI 10.1002/ep

Figure 2. X-ray diffraction (XRD) measurements of zeolite used in this study. (a) Natural zeolite and (b) surfac-

tant-modied zeolite (SMZ).

Modication Natural zeolites have negatively charged surfaces that can be modied by cationic surfactants to enhance adsorption of nonionic organic compounds. Large surfactant molecules such as hexadecyltrimethyl ammonium bromide (HDTMA) are adsorbed to the negatively charged surface of zeolite forming a monolayer. As the concentration of surfactant solution increases, surfactant molecules with large tail groups may adsorb via hydrophobic (tailtail) interactions, forming a bilayer or admicelles. The adsorption or exchange of inorganic anions occurs subsequently on the surfactant-modied minerals. This is mainly due to replacement of less strongly adsorbed counterions of surfactant [35]. Before use, zeolite samples were sieved to a size of 0.5 mm. These samples were then treated with 1 mol/L NaCl solution to saturate the exchange sites

Table 1. EDXRF results of natural zeolite.

Component Na2O MgO Al2O3 SiO2 K2O CaO Fe2O3 Others

% 0.817 0.232 12.789 80.736 3.275 0.927 0.699 0.525

Component Al K O Si Na Mg Ca Others

% 6.769 2.719 50.432 37.739 0.606 0.140 0.663 0.932

with sodium ions. HDTMA supplied by Merck GmbH was used for modication of the zeolite (see Figure 4). One hundred milliliters of HDTMA solution of concentration 0.03 mol/L was poured into a ask,
July 2011 143

Environmental Progress & Sustainable Energy (Vol.30, No.2) DOI 10.1002/ep

Figure 5. Effect of solids concentration on the adsorp-

tion of reactive dyes on surfactant-modied zeolite (SMZ) and natural zeolite (C0: 100 mg/L, t: 24 h, and agitation speed: 150 rpm).

Figure 3. The structural formula of Remazol Brillant

Blue and Remazol Yellow.

and sorbent were separated by centrifugation at 5000 rpm and analyzed for dye concentrations. The amounts of Remazol Brilliant Blue R and Remazol Yellow adsorbed were calculated according to the following equation: qe C0 Ce V ; m (1)

Figure 4. Chemical structure of cationic surfactant

used to prepare modied zeolite.

where qe is the amount of dye adsorbed (mg/g), C0 and Ce are the initial and equilibrium concentrations of the liquid phase (mg/L), V is the volume of the solution (L), and m is the mass of the adsorbent (g).
RESULTS AND DISCUSSION

and 10 g of zeolite was added. The dispersions were shaken at room temperature with a mechanical shaker for 24 h followed by washing with distilled water. The SMZ was dried at 508C in an oven. Kinetic and Equilibrium Experiments Kinetic and equilibrium experiments were conducted using the batch equilibrium technique by mixing a known amount of SMZ with 50 mL of Remazol Brilliant Blue R and Remazol Yellow solution in conical asks. The zeolite sample was then weighed and poured into these asks containing solution of the two dyes at desired concentrations. Each adsorption process was repeated two times, and the average value was used. To evaluate kinetic parameter, the same amount of solution of dye was mixed with the same amount of adsorbent in each ask, and the adsorption process was ended after different contact time intervals for each one (5 min24 h). For each adsorption isotherm experiment, a known amount of the adsorbent was added to 50 mL of the solution in a concentration range from 50 to 1000 ppm for Remazol Brilliant Blue R and Remazol Yellow. The mixture was shaken in a temperaturecontrolled shaking water bath at a constant speed of 150 rpm for 24 h. In both experiments, the solution
144 July 2011

Effect of Adsorbent Dosage The adsorption experiments were carried out at different solid/liquid ratios for 24 h at an initial dye concentration of 100 mg/L. Adsorption results obtained for various adsorbent dosages are given in Figure 5. Figure 5 shows that removal efciency rises with increasing adsorbent dosages and remains almost constant after a concentration of 10 g/L is used. Adsorption improved slightly when more SMZ was added into the solution. Consequently, the optimum solid/liquid ratio chosen was 10 g/L (1%) and was retained constant for all the experiments. The increase in adsorption with adsorbent dosage can be attributed to the greater surface area and the availability of more adsorption sites [36]. Although the adsorbed amount increases with adsorbent concentration, this increase is not linear. This nding is consistent with that obtained by Yapar et al. [37]. Effect of Contact Time A series of experiments were carried out in this study to ascertain the optimum time needed to reach equilibrium. Figure 6 shows the removal of Remazol Brilliant Blue R and Remazol Yellow plotted against contact time at the initial concentration of 100 mg/L. As shown in Figure 6, as contact time increases, the

Environmental Progress & Sustainable Energy (Vol.30, No.2) DOI 10.1002/ep

dqt k2 qe qt 2 : dt The integrated form of Eq. 4 becomes: t 1 1 t; 2 q t k2 q e q e

(4)

(5)

Figure 6. Effect of contact time on the adsorption of

reactive dyes on surfactant-modied zeolite (C0: 100 mg/L, adsorbent dosage: 10 g/L, and agitation speed: 150 rpm).

where k2 is the rate constant of pseudo-second-order 2 adsorption (g/mg min) and h 5 k2qe , where h is initial adsorption rate (mg/g min). These constants can be determined by plotting t/qt against t.

Intraparticle Diffusion Model

The intraparticle diffusion model is expressed as [43]: amount of reactive dye adsorbed into the zeolite increases signicantly. The rate of removal of both the dyes was found to be very rapid during the initial 3 h and remained nearly constant thereafter. No signicant change in removal of either dye was observed after about 5 h. The optimum contact time was used as 24 h because of the initial concentrations of dyes increased to 1000 mg/L in adsorption isotherm experiments. Kinetics Parameters of Adsorption Studies conducted in this eld have used several kinetics models to examine the controlling mechanisms of adsorption processes such as chemical reaction, diffusion control, and mass transfer. Applications of these models are described in literature [3840]. Five kinetic models, namely pseudo-rst-order, pseudo-second-order, intraparticle diffusion, Elovich, and Bangham models, were used in this study to investigate the adsorption of Remazol Brilliant Blue R and Remazol Yellow on SMZ.
Pseudo-First-Order Model

log R log kid a log t;

(6)

where R is the percentage Remazol Brilliant Blue R and Remazol Yellow adsorbed and t is the contact time (h), a is the gradient of linear plots, and kid may be taken as a rate factor, that is, percentage Remazol Brilliant Blue R and Remazol Yellow adsorbed per unit time. Higher values of kid signify an enhancement in the rate of adsorption, whereas larger a values signify a better adsorption mechanism, which is related to improved bonding between the percentage of Remazol Brilliant Blue R and Remazol Yellow ions and the zeolite particles.
Elovich Model

The Elovich equation is as follows [38, 44]: dqt a:e bqt ; dt (7)

The pseudo-rst-order equation is expressed [41] as dqt k1 qe qt : dt The integrated form of Eq. 2 becomes: logqe qt log qe k1 t; 2:303 (3) (2)

where a is the initial rate (mg/g min) because (dqt/dt) approaches a when qt approaches zero, and parameter b is related to the extent of surface coverage and activation energy for chemisorptions (g/mg) [38]. Given that qt 5 0 at t 5 0, the integrated form of Eq. 7 is qt 1 1 lnt t0 ln t0 ; b b (8)

where t0 5 1/ab. If t is much larger than t0, Eq. 8 can be simplied as qt 1 1 lnab ln t: b b (9)

where qe and qt are amounts of dye adsorbed (mg/g) at equilibrium and time t (min), respectively, and k1 is the rate constant of pseudo-rst-order (min21).
Pseudo-Second-Order Model

The pseudo-second-order reaction model based on adsorption equilibrium capacity may be expressed as [42].

The equations constants can be obtained from the slope and intercept of a straight-line plot of qt versus ln t.
July 2011 145

Environmental Progress & Sustainable Energy (Vol.30, No.2) DOI 10.1002/ep

Table 2. Kinetic parameters for the removal of

Remazol Brilliant Blue R and Remazol Yellow by surfactant-modied zeolite. Remazol Brilliant Blue R 8.8982 0.9212 0.9795 10.1523 0.0026 0.2679 0.9982 48.2281 0.2753 0.8395 0.5507 1.5849 0.9566 5.7715 0.6362 0.9536

particle surface. This leads to higher adsorption of the two dyes. However, for both dyes, the percentage adsorption decreases as the initial concentration increases.

Kinetic parameters Pseudo-rst-order qe (mg/g) k1 (3102 min21) R2 Pseudo-second-order qe (mg/g) k2 (g/mg min) h (mg/g min) R2 Intraparticle diffusion kid a R2 Elovich a (mg/g min) b (g/mg) R2 Bangham k0 [mL/(g/L)] a R2

Remazol Yellow 4.7851 0.4836 0.9317 10.3520 0.0038 0.4072 0.9996 56.5718 0.2237 0.9113 1.2803 1.4246 0.9126 11.87985 0.5110 0.9317

Adsorption Isotherms The equilibrium adsorption isotherm is fundamentally crucial in the design of adsorption systems. The equilibrium adsorption is usually expressed by an isotherm equation characterized by certain parameters whose values represent the surface properties and afnity of the sorbent. The equilibrium relationships between sorbent and sorbate are described by adsorption isotherms, the ratio between the amount sorbed and that remaining in the solution at a xed temperature at equilibrium [27]. In this study, the adsorption of Remazol Brilliant Blue R and Remazol Yellow by SMZ was analyzed by the Langmuir, Freundlich, Tempkin, and Dubinin Radushkevich (D-R) isotherm models.

Langmuir Isotherm

Bangham Model

The Langmuir isotherm has been widely applied to adsorption processes of pollutants. A basic assumption of the Langmuir theory is that the adsorption takes place at specic homogenous sites in the sorbent. Moreover, when a site is occupied by a solute, no further adsorption can take place at that site. The Langmuir adsorption isotherm can be written as follows [45]: qe KL qm Ce =1 KL Ce ; (10)

Kinetic data were further used to know about the slow step occurring in the present adsorption system using Banghams equation [17].  log log C0 C0 q t m  log  k0 m 2:303V  a logt;

where C0 is the initial concentration of adsorbate in solution (mg/L), V is the volume of the solution (mL), m is the weight of adsorbent per liter of solution (g/ L), qt (mg/g) is the amount of adsorbate retained at time t, and a (<1) and k0 are constants. The values of the kinetic constants obtained for adsorption of Remazol Brilliant Blue R and Remazol Yellow onto SMZ are presented in Table 2. Good correlation coefcients were obtained for the pseudosecond-order kinetic model, which shows that the uptake process follows the pseudo-second-order rate expression with the correlation coefcients being higher than 0.99.

where qe is the amount adsorbed at equilibrium (mg/ g), Ce is the equilibrium concentration (mg/L), qm is the adsorption capacity (mg/g), and KL is the adsorption intensity or Langmuir coefcient (L/mg). Equation 7 can be rearranged into the following linear form: Ce =qe 1=KL qm Ce =qm : (11)

A plot of Ce/qe versus Ce gives a straight line with slope 1/qm and intercepts 1/KLqm.
Freundlich Isotherm

Effect of Initial Concentration The adsorption of Remazol Brilliant Blue R and Remazol Yellow by SMZ increases as the initial concentration is increased. Increasing the initial concentration of the two dyes also increases the mass transfer driving force and therefore the rate at which the molecules of the dyes pass from the solution to the
146 July 2011

The Freundlich isotherm assumes an exponentially decaying adsorption site energy distribution. This is an experimental model and is applicable to nonideal adsorption on heterogeneous surfaces as well as multilayer adsorption and is expressed by the following equation:
1=n qe KF Ce ;

(12)

where KF and 1/n are the constants that can be related to the adsorption capacity and the adsorption intensity, respectively.

Environmental Progress & Sustainable Energy (Vol.30, No.2) DOI 10.1002/ep

Table 3. Isotherm parameters for the removal of

Remazol Brilliant Blue R and Remazol Yellow by surfactant-modied zeolite at 208C. Remazol Brilliant Blue R 109.6479 0.1194 0.7547 13.4952 0.2487 0.9941 453.7909 1.1238 0.7841 13.2245 0.5467 0.9563 0.9120

Figure 7. Equilibrium isotherms for the removal of

Remazol Brilliant Blue R and Remazol Yellow on surfactant-modied zeolite (Ce: equilibrium dye concentration mg/L, contact time: 24 h, adsorbent dosage: 10 g/L, and agitation speed: 150 rpm).

This equation can be rearranged in the linear form by taking the logarithm of both sides as log qe log KF 1=n log Ce : (13)

Isotherm parameters Freundlich KF 1/n R2 Langmuir qm KL R2 Tempkin A B R2 DubininRadushkevich q0 B E R2

Remazol Yellow 183.9924 0.2386 0.8682 38.3141 0.0823 0.9969 3.8753 5.1132 0.9592 29.7997 1.0596 0.7375 0.9192

The values of KF and 1/n may be calculated by plotting log(qe) versus log(Ce). The slope is equal to 1/n, and the intercept is equal to log KF.
Tempkin Isotherm

monolayer and the D-R equation is applicable, the adsorption capacity per unit surface area of the adsorbent at equilibrium, q, can be written as [47]. q q0 exp Be2 e RT ln1 1=Ce ; (17)

Tempkin and Pyzhev assumed the effects of some indirect adsorbateadsorbate 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 Tempkin isotherm has been used in the following form: qe RT =bln ACe qe RT =b ln A RT =b ln Ce B RT =b: (14)

(18)

(15)

where q0 is the ultimate capacity per unit area in adsorbent micropores, B is the constant related to the adsorption energy, e is the Polenyi potential, and Ce is the equilibrium concentration of the adsorbate in units of gram per gram. The most probable energy of adsorption, E, has been chosen as E 2B1=2 : (19)

(16)

A plot of qe versus ln Ce enables the determination of the constants A and B. The constant B is related to the heat of adsorption [46].
DubininRadushkevich Isotherm

The empirical equation proposed by Dubinin and Radushkevich has been widely used to describe the adsorption of gases and vapors on microporous solids. In the case of liquid-phase adsorption, several studies have shown that the adsorption energy can be estimated according to the D-R equation. Assuming that the adsorption in micropores is limited to a

The adsorption isotherms of Remazol Brilliant Blue R and Remazol Yellow by SMZ are shown in Figure 7. Among the two dye compounds, Remazol Yellow is more hydrophobic than Remazol Brilliant Blue R, having the higher adsorption capacity on SMZ. The Langmuir equation represents the adsorption process accurately; the R2 value is higher for the Langmuir isotherm than for the Freundlich, Tempkin, and D-R isotherms. The values of isotherm constants for Remazol Brilliant Blue R and Remazol Yellow adsorption onto SMZ are presented in Table 3. The Freundlich exponent 1/n provides information on surface heterogeneity and surface afnity for the solute. The Freundlich exponent 1/n between 0.1194 and 0.2386 indicates favorable adsorption. As the
July 2011 147

Environmental Progress & Sustainable Energy (Vol.30, No.2) DOI 10.1002/ep

degree of favorability increases as 1/n approaches zero, both SMZ samples reveal a very high afnity for Remazol Brilliant Blue R and Remazol Yellow. Values of qm, which is dened as the maximum capacity of sorbent, have been calculated from the Langmuir plots. The maximum capacities of SMZ for Remazol Brilliant Blue R and Remazol Yellow were calculated as 13.4952 and 38.311 mg/g, respectively. These values compare favorably with some of those reported in the literature. The maximum adsorption capacities of other minerals for dye ions removal were also given 5 mg/g for Alunite (Turkey) [48], 6.3 mg/g for clay (Turkey) [49], and 19.94 mg/g for natural zeolite [50]. Effect of Temperature and Thermodynamic Parameters Temperature has a pronounced effect on the adsorption capacity of adsorbents. Figures 8 and 9 are plots of adsorption isotherms, plotting qe versus

Ce for Remazol Brilliant Blue R and Remazol Yellow at different temperatures ranging from 20 to 608C. The resulting graphs indicate that with increase in temperature the adsorption of Remazol Brilliant Blue R and Remazol Yellow increases. The increase in the temperature from 20 to 608C increases the adsorption capacity for Remazol Brilliant Blue R and Remazol Yellow from 13.85 to 25.97 mg/g and from 38.44 to 60.79 mg/g, respectively. These results conrm the endothermic nature of the adsorption process. The values of isotherm constants for Remazol Brilliant Blue R and Remazol Yellow adsorption onto SMZ at different temperatures are presented in Table 4. To estimate the effect of temperature on the adsorption of Remazol Brilliant Blue R and Remazol Yellow onto SMZ, the free energy change (DG0), enthalpy change (DH0), and entropy change (DS0) were determined. The adsorption process can be summarized as that which represents a heterogeneous equilibrium.

Figure 8. Effect of temperature on the removal of

Figure 9. Effect of temperature on the removal of

Remazol Brilliant Blue R on surfactant-modied zeolite (Ce: equilibrium dye concentration mg/L, contact time: 24 h, adsorbent dosage: 10 g/L, and agitation speed: 150 rpm).

Remazol Yellow on surfactant-modied zeolite (Ce: equilibrium dye concentration mg/L, contact time: 24 h, adsorbent dosage: 10 g/L, and agitation speed: 150 rpm).

Table 4. Isotherm parameters at different temperatures for Remazol Brilliant Blue R and Remazol Yellow.

Freundlich 208C 408C 608C Langmuir 208C 408C 608C Tempkin 208C 408C 608C D-R 208C 408C 608C

KF 109.6479 128.8546 204.4561 qm 13.4952 23.1481 25.7732 A 453.791 67.4193 123.059 qo 13.2275 20.5770 23.5423

Remazol Brilliant Blue R 1/n 0.1194 0.1695 0.1684 KL 0.2487 0.1068 1.6033 B 1.1238 2.1590 2.4795 B E 0.5467 0.9563 0.0506 3.143 0.1685 1.7226

R2 0.7547 0.9435 0.7164 R2 0.9941 0.9971 0.9995 R2 0.7841 0.9702 0.7971 R2 0.9210 0.9585 0.8859

KF 183.992 823.000 2027.68 qm 38.3141 45.8716 56.8182 A 3.8753 164.667 94.2100 qo 29.7997 32.3851 52.4783

Remazol Yellow 1/n 0.2386 0.1672 0.1268 KL 0.0823 0.2011 0.2403 B 5.1132 3.9883 4.9578 B E 1.0596 0.7375 0.0169 5.4393 0.0091 7.4125

R2 0.8682 0.8947 0.8942 R2 0.9969 0.9936 0.9841 R2 0.9592 0.8474 0.6500 R2 0.9192 0.7549 0.6132

148 July 2011

Environmental Progress & Sustainable Energy (Vol.30, No.2) DOI 10.1002/ep

Table 5. Thermodynamic parameters for Remazol Brilliant Blue R and Remazol Yellow.

DG0 (kJ/mol) Remazol Brilliant Blue R Remazol Yellow DH0 (kJ/mol) 5.2146 29.9747 DS0 (kJ/mol K) 0.02734 0.10875 208C 22.7969 21.8900 408C 23.3438 24.0651 608C 23.8907 26.2402 R2 0.9796 0.9331

Thermodynamic parameters such as free energy of adsorption (DG0), the heat of adsorption and standard enthalpy, and entropy changes (DH0 and DS0) can be evaluated by the following equations: Kd qe =Ce ; (20)

during adsorption. The low value of DS0 also indicated that no remarkable changes in entropy occur [50].
CONCLUSIONS

where Kd is the adsorption distribution coefcient. The Kd values are used in the following equation to determine the Gibbs free energy of the adsorption process at different temperatures. DG 0 RT ln Kd ; (21)

where DG0 is the free energy of adsorption (kJ/mol), R is the universal gas constant (8.314 J mol K), and T is the absolute temperature (K). The adsorption distribution coefcient may be expressed in terms of enthalpy change (DH0) and the entropy change (DS0) as a function of temperature: ln Kd DH 0 =RT DS 0 =R; (22)

where DH 0 is the heat of adsorption (kJ/mol) and DS 0 is the standard entropy changes (kJ/mol K). The values of DH 0 and DS 0 can be obtained from the slope and intercept of a plot of ln Kd against 1/T. Values of standard Gibbs free energy change for the adsorption process obtained by Eq. 21 are listed in Table 5. The negative DG0 values of Remazol Brilliant Blue R and Remazol Yellow onto SMZ were due to the fact that the adsorption processes were spontaneous with a high preference of dyes onto SMZ: the negative value of DG0 decreases with an increase in temperature, indicating that the spontaneous nature of the adsorption of the two dyes was inversely proportional to the temperature and higher temperature favored the adsorption [50]. The higher amount of the dyes adsorbed at higher temperatures may be attributed to the greater penetration of the dyes onto the micropores caused by a swelling effect in the internal structure of the zeolite, which increases with increasing temperature [51]. The standard enthalpy and entropy changes of adsorption for Remazol Brilliant Blue R and Remazol Yellow were 5.2146 kJ/mol, 0.0273 kJ/mol K and 29.9747 kJ/mol, 0.1087 kJ/mol K, respectively. The positive values of DH 0 conrmed the endothermic character of adsorption on the dye/SMZ system, whereas the positive DS 0 values conrmed the increased randomness at the solidsolute interface

This study shows that SMZ is an effective sorbent for the removal of Remazol Brilliant Blue R and Remazol Yellow from aqueous solutions. Adsorption kinetics was found to follow the second-order kinetic model. The increase in initial concentration of Remazol Brilliant Blue R and Remazol Yellow resulted in an increase in the adsorption by SMZ. The adsorption of the two reactive dyes onto SMZ is favorably inuenced by an increase in temperature. Equilibrium adsorption data for both the dyes were best represented by the Langmuir isotherm. The maximum capacities for Remazol Brilliant Blue R and Remazol Yellow were calculated as 13.4952 and 38.311 mg/g, respectively, from the Langmuir plots. It was also found that Remazol Yellow had a higher adsorption capacity than Remazol Brilliant Blue R on SMZ. The thermodynamic parameters showed that the process is spontaneous and endothermic. The rise in temperature favored adsorption.
ACKNOWLEDGMENTS

This work was supported by Ondokuz Mayis University by project no. MF073.
LITERATURE CITED

1. Adak, A., Bandyopadhyay, M., & Pal, A. (2006). Fixed bed column study for the removal of crystal violet (C.1. Basic Violet 3) dye from aquatic environment by surfactant-modied alumina, Dyes and Pigments, 69, 245251. 2. Armagan, B., Turan, M., & Celik, M.S. (2004). Equilibrium studies on the adsorption of reactive azo dyes into zeolite, Desalination, 170, 3339. 3. Oguz, E., & Keskinler, B. (2006). Determination of adsorption capacity and thermodynamic parameters of the PAC used for bomaplex red CR-L dye removal, Colloids and Surfaces, 268, 124 130. 4. Nigam, P., Armour, G., Banat, I.M., Singh, D., & Marchant, R. (2000). Physical removal of textile dyes from efuents and solid-state fermentation of dye-adsorbed agricultural residues, Bioresource Technology, 72, 219226. 5. Chen, K.C., Wu, J.Y., Huang, C.C., Liang, Y.M., & Hwang, J. (2003). Decolorization of azo dye using PVA-immobilized micro organisms, Biotechnology, 101, 241252.
July 2011 149

Environmental Progress & Sustainable Energy (Vol.30, No.2) DOI 10.1002/ep

6. Juang, R.S., Wu, F.C., & Tseng, R.L. (1997). The ability of activated clay for the adsorption of dyes from aqueous solutions, Environmental Technology, 18, 525531. 7. Armagan, B., Ozdemir, O., Turan, M., & Celik, M.S. (2003). The removal of reactive azo dyes by natural and modied zeolites, Journal of Chemical Technology and Biotechnology, 78, 725732. 8. Dutta, K., Mukhopadhyay, S., Bhattacharjee, S., & Chaudhuri, B. (2001). Chemical oxidation of methylene blue using a fenton-like reaction, Journal of Hazardous Materials Part B, 84, 5771. 9. Mohan, S.V., Sailaja, P.S., Srimurali, M., & Karthikeyan, J. (1999). Color removal of monoazo acid dye from aqueous solution by adsorption and chemical coagulation, Environmental Engineering and Policy, 1, 149154. 10. Kim, T.H., Park, C., & Kim, S. (2005). Water recyling from desalination and purication process of reactive dye manufacturing industry by combined membrane ltration, Journal of Cleaner Production, 13, 779786. 11. Capar, G., Yetis, U., & Yilmaz, L. (2006). Membrane based strategies for the pre-treatment of acid dye bath wastewaters, Journal of Hazardous Materials Part B, 135, 423430. 12. Golder, A.K., Hridaya, N., & Ray, S.S. (2005). Electrocoagulation of methylene blue and eosin yellowish using mild steel electrodes, Journal of Hazardous Materials Part B, 127, 134140. 13. Kobya, M., Can, M., & Bayramoglu, O.T. (2003). Treatment of textile wastewater by electrocoagulation using iron and aluminium electrodes, Journal of Hazardous Materials Part B, 100, 163 178. 14. Al-Qodah, Z. (2000). Adsorption of dyes using shale oil ash, Water Research, 34, 42954303. 15. Juang, R.S., Tseng, R.L., Wu, F.C., & Lee, S.H. (1997). Adsorption behaviour of reactive dyes from aqueous solutions on chitosan, Journal of Chemical Technology and Biotechnology, 31, 391399. 16. Wang, S., & Zhu, Z.H. (2005). Sonochemical treatment of y ash for dye removal from wastewater, Journal of Hazardous Materials Part B, 126, 91 95. 17. Mall, I.D., Srivastava, V.C., & Agarwal, N.K. (2006). Removal of Orange-G and Methyl Violet dyes by adsorption onto bagasse y ashKinetic study and equilibrium isotherm analyses, Dyes and Pigments, 69, 210223. 18. Namasivayam, C., Yamuna, R.T., & Arasi, D.J.S.E. (2001). Removal of acid violet from wastewater by adsorption on waste red mud, Environmental Geology, 41, 269273. 19. Batzias, F.A., & Sidiras, D.K. (2007). Dye adsorption by prehydrolysed beech sawdust in batch and xed-bed systems, Bioresource Technology, 98, 12081217. 20. Lee, S.H., Song, D.I., & Jeon, Y.W. (2001). An investigation of the adsorption of organic dyes onto organo-montmorillone, Environmental Technology, 22, 247254.
150 July 2011

21. Eren, E., & Afsin, B. (2007). Investigation of basic dye adsorption from aqueous solution onto raw and pre-treated sepiolite surfaces, Dyes and Pigments, 73, 162167. 22. Ghosh, D., & Bhattacharyya, K.G. (2002). Adsorption of methylene blue on kaolinite, Applied Clay Science, 20, 295300. 23. Ozdemir, O., Armagan, B., Turan, M., & Celik, M.S. (2004). Comparison of the adsorption characteristics of azo-reactive dyes on mezoporous minerals, Dyes and Pigments, 62, 4960. 24. Meshko, V., Markovska, L., Mincheva, M., & Rodrigues, A.E. (2001). Adsorption of basic dyes on granular activated carbon and natural zeolite, Water Research, 35, 33573366. 25. Benkli, Y.E., Can, M.F., Turan, M., & Celik, M.S. (2005). Modication of organozeolite surface for the removal of reactive azo dyes in xed-bed reactors, Water Research, 39, 487493. 26. Kannan, N., & Meenakshisundaram, M. (2002). Adsorption of congo red on various activated carbons, Water, Air and Soil Pollution, 138, 289305. 27. Ghiaci, M., Abbaspur, A., Kia, R., & SeyedeynAzad, F. (2004). Equlibirium isotherm studies for the sorption of benzene, toluene and phenol onto organo-zeolites and as-synthesized MCM-41, Separation and Purication Technology, 40, 217229. 28. Breck, D.W. (1974). Zeolite molecular sieves, New York: Wiley. 29. Li, Z., Roy, S., Zou, Y., & Bowman, R. (1998). Long-term chemical and biological stability of surfactant modied zeolite, Environmental Science and Technology, 32, 26282632. 30. Li, Z. (1999). Sorption kinetics of hexadecyltrimethylammonium on natural clinoptilolite, Langmuir, 15, 64386445. 31. Li, Z., & Bowman, R.S. (2001). Retention of inorganic oxyanions by organo-kaolinite, Water Research, 35, 37713776. 32. Haggerty, G.M., & Bowman, R.S. (1994). Sorption of chromate and other inorganic anions by organo-zeolite, Environmental Science and Technology, 28, 452458. 33. Sullivan, E.J., Hunter, D.B., & Bowman, R.S. (1998). Fourier transform Raman spectroscopy of sorbed HDTMA and the mechanism of chromate sorption to surfactant modied clinoptilolite, Environmental Science and Technology, 32, 19481955. 34. Li, Z., Burt, T., & Bowman, R.S. (2000). Sorption of ionisable organic solutes by surfactant modied zeolite, Environmental Science and Technology, 34, 37563760. 35. Bowman, R.S. (2003). Applications of surfactantmodied zeolites to environmental remediation, Microporous and Mesoporous Materials, 61, 43 56. 36. Srivastava, V.C., Swamy, M.M., Mall, I.D., Prasad, B., & Mishra, I.M. (2006). Adsorptive removal of phenol by bagasse y ash and activated carbon: Equilibrium, kinetics and thermodynamics, Colloids and Surfaces A: Physicochemical and Engineering Aspects, 272, 89104.

Environmental Progress & Sustainable Energy (Vol.30, No.2) DOI 10.1002/ep

37. Yapar, S., Ozbudak, V., Dias, A., & Lopes, A. (2005). Effect of adsorbent concentration on the adsorption of phenol on hexadecyl trimethyl ammonium-bentonite, Journal of Hazardous Materials Part B, 121, 135139. 38. Onal, Y. (2006). Kinetics of adsorption of dyes from aqueous solution using activated carbon prepared from waste apricot, Journal of Hazardous Materials Part B, 137, 17191728. 39. Namasivayam, C., & Kavitha, D. (2002). Removal of Congo Red from water by adsorption onto activated carbon prepared from coir pith, an agricultural solid waste, Dyes and Pigments, 54, 47 58. 40. Malik, P.K. (2003). Use of activated carbons prepared from sawdust and rice husk for adsorption of acid dyes; a case study of Acid Yellow 36, Dyes and Pigments, 56, 239249. 41. Lagergreen, S. (1898). About the theory of so called adsorption of soluble substance, Kungliga Svenska Vetenskapsakad Handl, 24, 139. 42. Ho, Y.S., & McKay, G. (1999). Pseudo-second order model for sorption processes, Process Biochemistry, 34, 451465. 43. Kobya, M. (2004). Removal of Cr(VI) from aqueous solution by adsorption onto hazelnut shell activated carbon: Kinetic and equilibrium studies, Bioresource Technology, 91, 317321. 44. Cheung, C.W., Porter, J.F., & Mckay, G. (2000). Sorption kinetics for the removal of copper and

45. 46.

47.

48.

49.

50.

51.

zinc from efuents using bone char, Separation and Purication Technology, 19, 5564. Seader, J.D., & Herley, E.J. (1998). Separation process principles, New York: Wiley. Allen, S.J., Mckay, G., & Porter, J.F. (2004). Adsorption isotherm models for basic dye adsorption by peat in single and binary component systems, Journal of Colloid and Interface Science, 280, 322333. Hsieh, C-T., & Teng, H. (2000). Liquid phase adsorption of phenol onto activated carbons prepared with different activation levels, Journal of Colloid and Interface Science, 230, 171175. Ozacar, M., & Sengil, A.I. (2003). Adsorption of reactive dyes on calcined alunite from aqueous solutions, Journal of Hazardous Materials Part B, 98, 211224. Gurses, A., Karaca, S., Dogar, C., Bayrak, R., Aci kyildiz, M., & Yalcin, M. (2004). Determination of adsorptive properties of clay/water system: Methylene blue sorption, Journal of Colloid and Interface Science, 269, 310314. Han, R., Zhang, J., Han, P., Wang, Y., Zhao, Z., & Tang, M. (2009). Study of equilibrium, kinetic and thermodynamic parameters about methylene blue adsorption onto natural zeolite, Chemical Engineering Journal, 145, 496504. Eren, E. (2009). Removal of basic dye by modied Unye Bentonite, Journal of Hazardous Materials Part B, 162, 13551363.

Environmental Progress & Sustainable Energy (Vol.30, No.2) DOI 10.1002/ep

July 2011 151

You might also like