You are on page 1of 8

Journal of Molecular Structure 827 (2007) 6774 www.elsevier.

com/locate/molstruc

A study of equilibrium and FTIR, SEM/EDS analysis of trimethoprim adsorption onto K10
Zehra Bekc i, Yoldas Seki, M. Kadir Yurdakoc
*
_ Dokuz Eylu Turkey l University, Faculty of Arts and Sciences, Department of Chemistry, 35160 Buca, Izmir, Received 30 November 2005; received in revised form 8 April 2006; accepted 17 April 2006 Available online 16 June 2006

Abstract The sorption behavior of K10, a type of montmorillonite for trimethoprim (TMP) drug, was studied by using batch technique under dierent pH and temperature. The interaction between K10 and TMP was investigated using SEM, and FTIR. It was observed that adsorption was increased between pH 2.5 and 6.3. By performing kinetic experiments, the pseudo-second-order kinetic model provides the best t for TMP adsorption onto K10 montmorillonite. The sorption of TMP reached the equilibrium state after 6 h sorption time and has been described by using Langmuir, Freundlich and DubininRadushkevich equations to obtain adsorption capacity values. The results indicate that the relative adsorption capacity values (Kf) are decreasing with the increase of temperature in the range of 298 318 K. The sorption energy values obtained from DR isotherm show that sorption of TMP onto K10 can be explained by ion exchange mechanism at 298, 308 and 318 K. The thermodynamic studies were conducted to nd the thermodynamic parameters DH, DS and DG. It was determined that adsorption process is spontaneous and exothermic in nature. 2006 Elsevier B.V. All rights reserved.
Keywords: Trimethoprim; K10 montmorillonite; Sorption; SEM; FTIR

1. Introduction In recent years, there has been a growing concern with environmental protection. This can be achieved either by decreasing the aux of pollutants to the environment or by their removal from contaminated media [1]. Antibiotics have been widely used in human treatment and as growth promoters as well as therapeutic agents in the livestock industry. Animal manure containing excreted antibiotics is frequently applied to agricultural elds where antibiotics may potentially contaminate groundwater and eventually enter the surface water, such as rivers and lakes. Antibiotics used in human treatment can also enter the environment, either by excretion or by disposal of surplus drugs into sewage systems [2]. Euent from wastewater treatment plant is released into the local aquatic surroundings [3].

Corresponding author. Tel.: +90 2324128695; fax: +90 2324534188. E-mail address: k.yurdakoc@deu.edu.tr (M. Kadir Yurdakoc ).

Trimethoprim (TMP) is among the most important synthetic antibiotics [4]. Its therapeutic action is inhibition of bacterial dihydroate reductase, an enzyme active in the folate metabolic pathway converting dihydrofolate to tetrahydrofolate. In respect of many articles, TMP was detected in hospital sewage water in Sweden [5], in wastewater euents from East Aurora and Holland [2], in manure and soil in a German farming area [6] and in two municipal wastewater treatment plants in USA [7] and in US Streams [8]. Release of antibiotics to the environment may lead to high, long-term concentration and promote unnoticed adverse eects on aquatic and terrestrial organisms. Prolonged exposure to low doses of antibiotics leads to the selective proliferation of resistant bacteria which could transfer the resistance genes to other bacterial species [9]. Eect can accumulate so slowly that changes remain undetected until they become irreversible [10]. Because it is impossible to stop antibiotic usage, precautions must be taken against continual input of antibiotics to environment

0022-2860/$ - see front matter 2006 Elsevier B.V. All rights reserved. doi:10.1016/j.molstruc.2006.04.054

68

Z. Bekc i et al. / Journal of Molecular Structure 827 (2007) 6774

by using adsorbents such as natural and synthetic zeolite, clay minerals including montmorillonite K10 for the goal of removing antibiotics from wastewater treatment plants. In the present study, adsorption of TMP onto montmorillonite K10 has been studied. The reasons why montmorillonite K10 was used as an adsorbent are providing high surface area, high adsorption capacity and mechanical stability. The Langmuir, Freundlich and DubininRadushkevich (DR) equations were used to t the equilibrium isotherm. Furthermore, the kinetic and thermodynamic parameters were calculated to nd out adsorption rate constant and adsorption mechanism. These results will be useful for further applications in removing antibiotics from wastewaters. 2. Materials and methods Montmorillonite K10 used in this study was supplied as a commercial product of Fluka (Cat. No. 69867) and was utilized as an adsorbent in the adsorption trials. The properties of montmorillonite K10 are summarized in Table 1. Trimethoprim (TMP) was purchased from Sigma (Cat. No. T 788) and used without further purication. For batch adsorption experiments, 0.1 g of K10 was added into 25 mL of TMP solution. TMP solution was prepared by using certain proportion of ethanol and distilled water to enhance the solubility of TMP and prepared in the concentration ranges of 0.41.3 mmol L1. The pH of drug solution was adjusted with NaOH and HCl solutions using a pH meter. The K10-TMP suspensions were agitated at 150 rpm in a temperature controlled shaking water bath (Memmert) for 6 h to reach equilibrium under experimental conditions. After the equilibrium, a portion of the supernatant layers of suspensions were taken and centrifuged at 5000 rpm for 15 min. Spectrophotometric method was applied to drug solutions. Spectrophotometric measurements were carried out using a Shimadzu UVVisible 1601 model spectrophotometer at kmax = 271 nm at ambient temperature. The amount of TMP adsorbed on K10 montmorillonite was calculated by the dierence of initial concentration and equilibrium concentration of TMP. The same procedures were carried out at solution temperatures of 298, 308, and 318 K to nd adsorption isotherms and thermodynamic parameters. In addition, some experiments were performed at various time intervals to determine the kinetic parameters.
Table 1 Chemical composition of K10 montmorillonite Component SiO2 Al2O3 Fe2O3 MgO CaO Na2O K2O Ignition loss Contents % 69.0 14.0 4.5 2.0 1.5 <0.5 <1.5 7.0

2.1. FTIR measurements For infrared analysis, TMP, K10 and TMP loaded K10 samples were brought to constant weight in a drying oven at 50 C for 24 h and kept in the desiccators. One hundred milligrams of ne KBr powders was dried at 110 C and mixed with 1 mg of sample, and then pellets were prepared as KBr disks. FTIR spectra in the range of 4000400 cm1 were obtained on a Perkin-Elmer FTIR Spectrophotometer (Spectrum BX-II). 2.2. SEM measurements In order to obtain information about the surface morphologies of K10, TMP and TMP loaded KSF, scanning electron microscopy, SEM was performed in a Jeol JSM 60 SEM at accelerating voltage of 20 kV attached an X-ray energy dispersive spectrometer, EDS. Before scanning process, all samples were dried and coated with gold to enhance the electron conductivity. SEM micrographs were taken at dierent magnications between 1000 and 8000. 3. Results and discussion 3.1. Adsorption isotherms The adsorption isotherms of TMP onto montmorillonite K10 at various temperatures were determined by plotting the amount of TMP adsorbed by K10 (Cs mmol g1) versus the equilibrium concentration of TMP (Ce mmol L1). Fig. 1 shows the trimethoprim sorption isotherms at 298, 308, and 318 K. In terms of the slope of initial portion of the curves, the shapes of these isotherms correspond to L type according to the Giles classication [11]. The L curve indicates that there is no strong competition between adsorbate and solvent for sites on the surface and K10 has a medium anity for the TMP molecules.
30

25

298K 308K 318K

20

Cs / mmolg

-1

15

10

0 0 5 10 15 20
-1

25

30

Ce / mmolL

Fig. 1. Adsorption isotherms of TMP on K10.

Z. Bekc i et al. / Journal of Molecular Structure 827 (2007) 6774

69

In the batch experiment, three well-known models, namely the Langmuir, Freundlich, and DubininRadushkevich (DR) adsorption equilibrium models, are used to analyze the experimental data. 3.1.1. Langmuir isotherm Langmuir adsorption isotherm is applicable in many adsorption processes. The basic assumption is the formation of a monolayer of adsorbate on the outer surface of the adsorbent and no further adsorption thereafter [12]. In order to nd sorption capacities of montmorillonite K10, linear form of Langmuir equation was applied; Ce 1 Ce C s LC m C m
1

3.1.3. DubininRadushkevich (DR) isotherm The linear form of DR isotherm (Eq. (3)) is used to evaluate the equilibrium data; ln C s ln X m be2 3

where Cs is the amount of ions sorbed onto the adsorbent (mol g1), Xm represents maximum adsorption capacity of the sorbent (mol g1), b is a constant related to sorption energy (mol2 kJ2) and e is Polanyi sorption potential, the amount of energy required to pull a sorbed molecule from its sorption site to innity which is equal to;   1 e RT ln 1 4 Ce where R is the gas constant in kJ mol1 K1; T is the temperature in Kelvin and Ce is the equilibrium concentration in solution (mol L1). Xm and b values can be determined from the intercept and the slope of the linear plot of experimental data of ln Cs versus e2. The sorption energy, E, for ions onto adsorbent calculated by using the expression is dened as below. E 2b
0:5

Ce is the liquid phase (mmol L ), and Cs is the solid phase concentration of adsorbate at equilibrium (mmol g1). Cm is the maximum amount that can be adsorbed by adsorbent as a monolayer (mmol g1); L is the Langmuir constant regarding the adsorption energy. Cm and L parameters obtained by plotting Ce/Cs versus Ce for the adsorption process of trimethoprim onto K10 corresponding to Langmuir equation are given in Table 2. It can be easily understood from Table 2 that maximum adsorption capacities of adsorbent for the monolayer adsorption increase by increasing temperature. 3.1.2. Freundlich isotherm A linear form of the Freundlich expression is as follows: ln C s ln K f nf ln C e 2 where Kf and nf are indicators of adsorption capacity and heterogeneity factor, respectively. The values of Kf and nf are estimated from the intercept and slope of the plot of ln Cs versus ln Ce. As can be seen apparently in Table 3 which demonstrates the Freundlich constants at various temperatures, Kf values are decreased by increasing temperature of the solutions. This result indicates that adsorption capacities of K10 increase when temperature decreases in a given temperature interval. It can be easily understood that trimethoprim is more adsorbed at 298 K, in comparison to 308 and 318 K. Correlation coecients are also calculated by tting the experimental adsorption equilibrium data for the TMPmontmorillonite K10, using both the Langmuir and Freundlich adsorption isotherms, and are shown in Tables 2 and 3, respectively. It was found that the adsorption isotherm for the TMP-K10 system is explained better by the Langmuir isotherm.

5
1

If the sorption energy is the range of 816 kJ mol , the adsorption type is designated for ion exchange mechanism. If sorption energy is smaller than 8 kJ mol1, the adsorption type is physical adsorption due to weak van der Waals forces [13]. Sorption parameters of DR equation at 298, 308, and 318 K are summarized in Table 4. All of the E values obtained at certain temperatures are in the range of 10.712.7 kJ mol1. This result indicates that sorption of TMP onto K10 can be explained by ion exchange mechanism at 298, 308, and 318 K. 3.2. Eect of pH on TMP adsorption onto K10 The pH of the aqueous solution is an important variable that inuences the adsorption of anions and cations at the solidliquid interfaces [14]. For this study, solutions of concentration 0.7 mmol L1 of the drug at ambient tempera-

Table 3 The sorption parameters of Freundlich equation for adsorption of TMP onto K10 Temperature (K) 298 308 318 R2 0.9789 0.9968 0.9877 Kf (mmol g1) 0.686 0.493 0.449 nf 0.548 0.421 0.408

Table 2 The sorption parameters of Langmuir equation for TMP onto K10 Temperature (K) 298 308 318 R
2

Table 4 The sorption parameters of DR equation for TMP onto K10 Temperature (K)
1

R2 0.989 0.998 0.991

Cm (mmol g ) 0.2606 0.2753 0.3133

L (g L ) 25.1150 22.2448 15.0780 298 308 318

Xm 103 (mol g1) 1.88 1.15 1.13

b 103 (mol2 kJ2) 4.37 3.27 3.12

E (kJ mol1) 10.7 12.4 12.7

0.9981 0.9898 0.9866

70

Z. Bekc i et al. / Journal of Molecular Structure 827 (2007) 6774


70 60 50 40 30 20 10 0 0 2 4 6 8 10

T: +H $ TH Protonated form of the drug is strongly adsorbed to the negatively charged sites of K10 surface by cation exchange process in acidic conditions. As the pH of the solution increased, a negatively charged surface site on the clay did not favor the adsorption of neutral T:, for that reason adsorption decrease in neutral and basic conditions. 3.3. Adsorption kinetics Fig. 4 shows the typical time changes of solid-phase TMP concentration. As can be seen apparently, optimum contact time was xed at about 6 h. The kinetic curve

The amount of drug adsorbed / mgg

ture were prepared at dierent pH values. After the equilibrium established, the solid-phase concentration of TMP versus equilibrium pH was plotted. The eect of pH on TMP adsorption onto montmorillonite K10 for the equilibrium pH between 2.5 and 8.1 is presented in Fig. 2. It was observed that with the increase in the pH of the solution, adsorption is increased between pH 2.5 and 6.3. Beyond the value of 6.3, there was a decrease in the adsorption. At low pH conditions, all of the TMP (T:) is in the protonated form, because TMP is a weak base. The chemical structure of TMP is shown in Fig. 3.

-1

Time /h
Fig. 4. Eect of contact time of TMP adsorption onto K10.

70

65

showed that the adsorption of TMP onto K10 is rapid in the rst minutes and when it reached equilibrium, the amount of TMP adsorbed remained constant. Several simplied kinetic models including the pseudo-rst-order equation, pseudo-second-order equation, and intraparticle diusion model were tested to check the validity [15,16]. A good correlation of kinetics data explains the adsorption mechanism [17]. The rst-order-equation is given as follows.   1 k1 1 1 6 qt q1 t qt where qt is the amount of TMP adsorbed at dierent times t (mg g1), k1 is the pseudo-rst-order rate constant for the adsorption process (min1). The value of q1 is the maximum adsorption capacity (mg g1) [18]. The values of the rst-order rate constant k1 and q1 are obtained from the slope and intercept of the straight line. The expression for the pseudo-second-order rate equation is given as [19].

60

Csx10 / molL

-1

55

50

45

40 2 3 4 5 6 7 8 9

pH
Fig. 2. pH eects on adsorption of TMP onto K10.

t 1 1 2 qt k 2 q2 q2

NH2 OCH3 N

H2N

N OCH3
Fig. 3. Chemical structure of TMP.

OCH3

where k2 is the pseudo-second-order rate constant (mg g1), q2 is the maximum adsorption capacity for the second order (g mg1min1). The straight line plots of t/ qt versus t for the pseudo-second-order adsorption of TMP onto K10 have been applied to determine rate constants k2 and q2 at various temperatures (R2 = 1). The equation is basically based on the sorption capacity; the description of sorption phenomenon suggests that the chemical reaction is rate controlling [20]. Pseudo-rst-order and pseudo-second-order cannot identify the diusion mechanism. To identify the diusion mechanism, intraparticle diusion model is applied to adsorption data. This equation is given as qt k i t1=2 C 8

Z. Bekc i et al. / Journal of Molecular Structure 827 (2007) 6774 Table 5 Kinetic parameters for the TMP adsorption onto K10 298 K Pseudo-second-order model R2 k2 (g mg1min1) q2 (mg g1) Intraparticle diusion model R2 ki (mg s1 g1) C 0.999 0.005 62.2 0.939 0.190 57.9 308 K 0.999 0.003 63.0 0.970 0.246 57.4 318 K 0.999 0.008 60.9 0.971 0.102 58.5

71

Table 6 Thermodynamic parameters for TMP adsorption onto K10 Temperature (K) 298 308 318 Kc 17.276 16.823 15.030 D G (kJ mol1) 7.06 7.23 7.17 DS (J mol1 K1) 5.56 DH (kJ mol1) 5.44

where ki is the intraparticle diusion rate constant (mg s1/2 g1), C is the intercept. The value of C and ki was obtained from the intercept and slope of qt versus t1/2. The pseudo-rst-order model gave poor slopes (data not shown here). The equation, which gave the best t for the experimental kinetics data, was the pseudosecond-order model. It can be clearly seen from Table 5, which shows kinetic parameters obtained from pseudo-second-order and intraparticle diusion models. The linear portions of the intraparticle diusion curves do not pass through the origin, so it was proposed that in addition to intraparticle diusion, other process may control the rate of adsorption. All of them may be operating simultaneously [21]. The pseudo-second-order kinetic model has indicated that these chemisorption systems involve vacancy forces through sharing or exchange of electrons between the sorbent and solute [22].

Fig. 5. SEM micrographs of (a) K10 montmorillonite (1900); (b) TMP loaded K10 (1500); (c) TMP loaded K10 (4000); (d) TMP loaded K10 (8000); (e) EDS analysis of K10; (f) EDS analysis of TMP-K10.

72

Z. Bekc i et al. / Journal of Molecular Structure 827 (2007) 6774

Fig. 6. FTIR spectra of K10, TMP and K10-TMP.

3.4. Thermodynamic parameters The amounts of TMP adsorbed onto montmorillonite K10 were measured in the temperature range 298318 K. The equilibrium distribution constant Kc is calculated by K c C s =C e 9 where Ce is the equilibrium concentration, Cs is the amount sorbed on solid at equilibrium. The following relationships have been used to evaluate the thermodynamic parameters Gibbs free energy change DG, standard enthalpy change DH and entropy change DS [23,24]. DG RT ln K d   DH  1 DS  ln K d T R R 10 11

vant Ho equation (Eq. (11)) can be used with the experimental data to evaluate entropy change of adsorption DS and enthalpy change of adsorption DH from the intercept and slope by plotting ln Kd versus 1/T. T represents the temperature in Kelvin. Additionally, the values of Gibbs free energy change DG are then calculated from Eq. (10) at 298, 308, and 318 K. The value of DH, DS and the values of DG are listed in Table 6, for the equilibrium concen-

tration of 2.0509 102 mmol L1 by using Langmuir equation to nd Cs and Kc values. It is known the absolute magnitude of the change in free energy for physisorption is between 20 and 0 kJ mol1; chemisorption has a range of 80 to 400 kJ mol1 [21,25]. DG values calculated are 7.06 kJ mol1 at 298 K, 7.23 kJ mol1 at 308 K and 7.17 kJ mol1 at 318 K, respectively. These results show that the adsorption of TMP onto K10 may occur as physically. Also, the negative values of DG at various temperatures indicate that the adsorption process is spontaneous. According to [21], the standard free energy change for multilayer adsorption was more than 20 kJ mol1 and less than zero. Additionally, it can be easily deduced that DG values at 298, 308, and 318 K are in the range of multilayer adsorption. The negative value of DH represents that the adsorption process is exothermic in nature. DH value obtained is 5.44 kJ mol1. Since, this value is lower than 40 kJ mol1, the type of adsorption can be accepted as physical process [17]. The positive value of DS suggests increased randomness during adsorption [26]. In order to examine the morphology of K10 and TMPK10, scanning electron microscopy (SEM) is equipped with an energy dispersive X-ray detector. Fig. 5a shows SEM

Z. Bekc i et al. / Journal of Molecular Structure 827 (2007) 6774

73

image of K10. The micrographs of TMP loaded K10 montmorillonite are presented in Fig. 5bd with dierent magnications. The fabric is formed of silicate particles with random orientation. Pore spaces can be seen. It can be clearly said that K10 has nonbrous morphology and diameter of particles that are in the range of about 10 2 nm. For the micrograph of TMP loaded K10, particles have dierent sizes and were dispersed in various directions. The orientation of clay layers is not obvious. As can be seen from Fig. 5d (8000 magnication), some agglomerations can be seen on silicate particles. The results of chemical analysis of the K10 clay and TMP loaded K10 were obtained using semiquantitative EDS data, presented in Fig. 5e and f, respectively. By the way of loading with TMP, carbon content of clay sample was increased from 19.865% to 38.894%. This result which conrms the loading of TMP successfully is compatible with previous ndings. The FTIR spectrum of TMP, K10-TMP and K10 is presented in Fig. 6. The band at 3620 cm1 is typical for smectites due to internal hydroxyls of clay minerals. The stretching vibration of water was observed at 3428 cm1. After TMP adsorption, this band was shifted to 3446 cm1. It can be inferred that hydrogen bonding can be seen between bounded water and TMP molecules. In the spectrum of K10, the band at 1635 cm1 is assigned to water deformation band. After TMP loading, a doublet was observed in that peak. This may be attributed to stretching vibration of C@NH2. One can conclude that the second peak at 1666 cm1 results from stretching vibration of C@NH2 which is the positively charged moiety of TMP. It can be reached that TMP may subject to ion exchange reaction with cations existing on clay surface. There are some absorption bands in the 15101450 cm1 region, due to aromatic group. The band at 1508 cm1 is possibly corresponding to 1,4-substituted aromatic compound. Therefore it is reasonable to assume that TMP was adsorbed onto K10 montmorillonite. 4. Conclusion The Langmuir, Freundlich and DR adsorption isotherms were used to obtain isotherm constants for evaluating the adsorption in various temperatures. The relative adsorption capacity and maximum adsorption capacity obtained from Freundlich and DR adsorption isotherms, respectively, are increased by decreasing the temperature of the solution. All E (sorption energy) values obtained from DR adsorption isotherm were between 8 and 16 kJ mol1, indicating that sorption of TMP onto K10 can be explained by ion exchange mechanism at 298, 308 and 318 K. The adsorptive removal of TMP from aqueous solution is pH dependent. In the range of pH 2.5 and 6.3, adsorption is high via the attraction between the negatively charged surface of clay and protonated form of the drug. In the range of pH 6.3 and 8.1, a negatively charged surface site on the

clay did not favor the adsorption of neutral TMP, for that reason adsorption decrease in neutral and basic conditions. From kinetic models applied to experimental data, pseudo-second-order model was found to be best model. Thermodynamic parameters; DH, DS and DG were estimated. DH value obtained is 5.44 kJ mol1, indicating physical and exothermic adsorption process. The adsorption process at 298, 308, and 318 K is spontaneous and multilayer. Since DG values are in the range of 0 and 20 kJ mol1, it can be concluded that the mechanism of adsorption is carried out with the combination of ion exchange mechanism and physical adsorption. To remove TMP euents from the wastewater treatment plants, adsorption is applicable via these kinetic and thermodynamic parameters. It is impossible and unpractical to stop antibiotic disposal. Hence, this study will give a route near future investigation relating to the removing of drugs from the environment and wastewater treatment plants. Acknowledgement lu, Dokuz Eylu The authors thank Murat Kus og l University, Faculty of Engineering, Department of Metallurgical and Materials Engineering. References
[1] H. Baker, F. Khalili, Anal. Chim. Acta 516 (2004) 179186. [2] A.L. Batt, D.S. Aga, Anal. Chem. 77 (2005) 29402947. [3] S.E. Jorgensen, B. Halling-Sorensen, Chemosphere 40 (2000) 691 699. [4] K. Florey, in: Analytical Prole of Drug Substances, vol. 7, Academic Press, New York, 1978, p. 459. [5] R. Lindberg, P.-A. Jarnheimer, B. Olsen, M. Johansson, M. Tysklind, Chemosphere 57 (2004) 14791488. [6] X-S. Miao, F. Bishay, M. Chen, C.D. Metcalfe, Environ. Sci. Technol. 38 (2004) 35333541. [7] J.E. Renew, C.H. Huang, J. Chromatogr. A 1042 (2004) 113121. [8] D.W. Kolpin, E.T. Furlong, M.T. Meyer, E.M. Thurman, S.D. Zaugg, L.B. Barber, H.T. Buxton, Environ. Sci. Technol. 36 (2002) 12021211. [9] S.B. Levy, in: D.J. Chadwick, J. Goode (Eds.), Antibiotic Resistance: Origins, Evolution, Selection and Spread, Ciba Foundation Symposium, vol. 207, Wiley, West Sussex, England, 1997, pp. 114. [10] M.S. Diaz-Cruz, M.J. Lopez de Alda, D. Barcelo, Trends Anal. Chem. 22 (2003) 6. [11] C.H. Giles, T.H. McEvan, S.N. Nakhwa, D. Smith, J. Chem. Soc. 4 (1960) 39733993. [12] M.K. Purkait, S. DasGupta, S. De, J. Environ. Manage. 76 (2005) 135142. [13] U.R. Malik, S.M. Hasany, M.S. Subhani, Talanta 66 (2005) 166173. [14] L. Zeng, L. Xiaomei, L. Jindun, Water Res. 38 (2004) 1318 1326. [15] R.S. Juang, S.H. Liu, K.H. Tsao, J. Colloid Interface Sci. 254 (2002) 234241. [16] Y.S. Ho, G. McKay, Proc. Biochem. 34 (1999) 451465. [17] K.C. Justi, M.C. Laranjeira, M.A. Neves, A.S. Mangrich, V.T. Favere, Polymer 45 (2004) 62856290. [18] M. Yurdakoc, Y. Seki, S. Karahan, K. Yurdakoc, J. Colloid Interface Sci. 286 (2005) 440446. [19] S. Motoyuki, Adsorption Engineering, Elsevier, Tokyo, 1990. [20] Y.S. Ho, G. McKay, Water Res. 33 (1999) 578584. zcan, A. O zcan, J. Colloid Interface Sci. 276 (2004) 3946. [21] A.S. O

74

Z. Bekc i et al. / Journal of Molecular Structure 827 (2007) 6774 [25] Y. Yu, Y.Y. Zhuang, Z.H. Wang, J. Colloid Interface Sci. 242 (2) (2001) 288293. [26] G.N. Manju, C. Raji, T.S. Aniruddhan, Water Res. 32 (1998) 3062 3070.

[22] R.S. Juang, F.C. Wu, R.L. Tseng, J. Colloid Interface Sci. 227 (2000) 437444. [23] I. Zhu, X. Ren, S. Yu, Environ. Sci. Technol. 32 (1998) 33743378. [24] B.S. Krishna, D.S.R. Murty, B.S.J. Prakash, J. Colloid Interface Sci. 229 (2000) 230236.

You might also like