Anda di halaman 1dari 12

Desalination 212 (2007) 282293

Modeling the effects of adsorbent dose and particle size on the adsorption of reactive textile dyes by fly ash
S. Karaa, C. Aydinera, E. Demirbasb*, M. Kobyaa, N. Dizgea
a

Gebze Institute of Technology, Department of Environmental Engineering, 41400, Gebze, Turkey Tel. +90 (262) 6053108; Fax +90 (262) 6053101; email: erhan@gyte.edu.tr b Gebze Institute of Technology, Department of Chemistry, 41400, Gebze, Turkey

Received 8 July 2006; accepted 18 September 2006

Abstract The adsorption of three reactive dyes, Remazol Red, Remazol Blue and Rifacion Yellow, from aqueous solutions using fly ash as an adsorbent was studied in an agitated batch system to investigate the influence of two parameters viz., adsorbent dosage and particle size on the removal efficiency of the reactive dyes. Firstly, the adsorbent was characterized with using several methods such as SEM, XRD and FTIR. The FTIR suggested that the dye on fly ash is probably indicating fly ash/dye complexation. XRD pattern of fly ash consisted of mainly quartz, mullite with some magnetite and calcite. Surface morphology of fly ash and dye loaded fly ash were obtained with SEM. Secondly, the percentage of dye removal at equilibrium, p, was determined with respect to these two parameters with a constant initial dye concentration of 100 mg/L, agitation speed of 250 rpm, pH 6 and temperature of 22C for a period of 48 h. The experimental data were treated with two simple empirical models used for predicting the percentage of the dyes adsorbed on the fly ash. Both models showed good correlation coefficients but the best model which determined the p values can be selected on the basis of the standard deviation of the calculated and experimental values. Keywords: Reactive dyes; Fly ash; Adsorbent particle size; Adsorbent dosage

1. Introduction The vinyl sulfone and chlorotriazine dyes are the most reactive and versatile of the fiber reactive dyes, which mean that the dye molecules actually react with fabric molecules. The dyes can
*Corresponding author.

be used on cotton, silk, wool, rayon, paper and wood but not on synthetic fibers [1,2]. These dyes discharge into water causes environmental pollution. Residues dyes are the major contributors to color in wastewaters generated from textile and dye manufacturing industries. Unless properly treated, these dyes can significantly affect the nature of the water and inhabits sunlight penetra-

0011-9164/07/$ See front matter 2007 Published by Elsevier B.V. doi:10.1016/j.desal.2006.09.022

S. Kara et al. / Desalination 212 (2007) 282293

283

tion into the stream and reduces photosynthetic action and may also be toxic to certain forms of aquatic life due to the presence of substituent metals and chlorine. Dyes can also cause allergic dermatitis, skin irritation, cancer and mutation [3]. The conventional methods for treating dyecontaining wastewaters are chemical coagulation, chemical oxidation, photochemical degradation, membrane filtration, including aerobic and anaerobic biological degradation [1,2]. The chemical coagulation process effectively decolorizes insoluble dyes, but it fails to work well with soluble dyes. Chemical oxidation is effective, but the oxidant requirements are very high and thus expensive. Photochemical degradation in aqueous solution is likely to progress slowly, as synthetic dyes are, in principle designed to exhibit high stability to light. Although biological treatment processes remove BOD, COD, and suspended solids to some extent, they are largely ineffective in removing color from wastewater, as most dyes are toxic to the organisms used in such processes. However, all of these methods suffer from one or other limitations, and none of them were successful in completely removing the color from wastewater. These technologies do not also show significant effectiveness or economic advantage. Low-cost treatment methods have, therefore, been investigated for a long time. Adsorption has been used extensively in industrial processes for separation and purification. In wastewater treatment, commercially activated carbon has long been used as a standard adsorbent for color removal. In spite of its widespread use in various cleaning procedures, activated carbon remains expensive; therefore, the development of low-cost alternative adsorbents has been the focus of recent research [4,5]. Contributions in this regard have been made by many researchers who have utilized a number of substances such as agricultural wastes: coir pith, banana pith, sugar cane dust, sawdust, activated carbon fibers and rice hulls [610], industrial solid wastes: fly ash, red mud and shale oil ash [1118], and so forth.

During coal-fired electric power generation, two main types of coal combustion by-products are obtained, fly ash and bottom ash. The current annual worldwide production of coal ash is estimated about 700 million tons of which at least 70% is fly ash [19]. Although, significant quantities are being used in a range of applications and particularly as a substitute for cement in concrete, large amounts are not used and this requires disposal. Making a more productive use of fly ash would have considerable environmental benefits, reducing air and water pollution. About 55 million tons of coal and lignite is combusted annually in Turkey resulting in more than 15 million tons of fly ash. In the present study, the adsorbent was characterized with a number of methods includes SEM, XRD and FTIR. Effects of adsorbent dosage and particle size on the adsorption of three reactive azo dyes [Remazol Blue (RB), Remazol Red RB 133 (RR) and Rifacion Yellow HED (RY)] from aqueous solutions using the fly ash under equilibrium conditions were investigated. The data were treated with two empirical models to predict the percentage of the dyes adsorbed on the fly ash.

2. Materials and methods 2.1. Materials The fly ash was obtained from Afsin-Elbistan Thermal Power Station in Turkey. The AfsinElbistan power plant consumes 18106 metric tons of coal per year and generates about 3.24106 metric tons of fly ashes returning to the dumping area of the mine as combustion waste. The particle size distribution of the fly ash was found between 3.6 and 181 m. Higher percentage of the fly ash consists of particles with diameter 40 125 m (determined by the method of laser beam dispersion using the Malvern 2000 particle size analyzer). The particle size of the adsorbent was determined by sieve analysis. The fly ash was

284

S. Kara et al. / Desalination 212 (2007) 282293

sieved by using a sieve set and then was collected in the range of 4050, 5060, 6080, 80100, and 100125 m, respectively. The surface area of the fly ash was measured by BET (Brunauer EmmettTeller nitrogen adsorption technique). The bulk density of the adsorbent was determined with a densitometer. The fly ash was used as received without any pretreatment in the adsorption experiments. Three reactive dyes obtained from Dystar and Itochu were used in the adsorption study. Their chemical structures and characterizations are presented in Fig. 1 and Table 1, respectively.

2.2. Methods Batch adsorption tests were conducted by varying adsorbent particle sizes and adsorbent doses on a rotary shaker using 100 ml screw-cap conical flasks containing 50 ml of dye solution having a concentration of 100 mg/L, pH 6, agitation speed of 250 rpm and temperature of 22C for a period of 48 h. After this period, the final equilibrium concentrations of the dye in solution were measured spectrophotometrically at a wavelength of 518, 585 and 411 nm for RR, RB and RY, respectively using a Perkin-Elmer UV-visible spec-

NH2 SO3Na SO2CH2CH2OSO3Na

N H

(a)
Cl NaSO3O O S O SO3Na N=N N OH NH N N H SO3Na SO3Na N

(b)

H N SO3H N=N SO3H

N N

Cl N N H SO3H SO3H NH

Cl N N N N H

SO3H SO3H

N=N SO3H SO3H

SO3H

SO3H

(c) Fig. 1. Chemical structure of the reactive dyes (a) Remazol Brilliant Blue (RB), (b) Remazol Red 133 (RR) and (c) Rifacion Yellow HED (RY).

S. Kara et al. / Desalination 212 (2007) 282293 Table 1 Characterizations of the reactive dyes
Parameters Color index name Chromophore group Reactive anchor systems* Molar mass, g/mol Max absorbance, m, nm Purity, % Water solubility at 293 K, g/l Acute oral toxicity LD50, mg/ kg Fish toxicity LC0, mg/L pH value, 10 g/l water COD, mg/g BOI5, mg/g DOC, mg/g Company Remazol Brillant Blue (RB) Reactive Blue 19 Anthraquinone VS 506.5 585 50 100 2000 55.5 Dystar Remazol Red 133 (RR) Reactive Red 198 Azo MCT+VS 984.2 518 63 70 2000 >500 7 540 <10 120 Dystar Rifacion Yellow HED (RY) Reactive Yellow 84 Disazo ACT 1922 411 80 70 6.5 160 Itochu

285

*MCT monochlortriazine; ACT bisaminochlorotriazine, VS vinylsulfone

trophotometer model 550S. The percentage removal of dye, p, was calculated using the following equation
p=

( C0 Ce ) 100
C0

(1)

where C0 and Ce are the initial and equilibrium dye concentration (mg/L), respectively. All experiments were replicated and the average results were used in data analysis. 3. Results and discussion 3.1. Characterization of the adsorbent material Since adsorption is a surface phenomenon, the rate and extent of adsorption are functions of the specific surface area of the adsorbent used, i.e., the portion of the total surface area that is available for adsorption. In fact, the amount of adsorption per unit weight of fly ash depends on its composition, texture and porosity. The fly ash sample was characterized through physico-chemical

analysis together with SEM, XRD pattern and FTIR. Chemical composition of the fly ash by chemical analysis was given as SiO2, 15.14; Fe2O3, 3.30; Al2O3, 7.54; CaO, 23.66; MgO, 4.5; SO3, 13.22; K2O, 0.28; Na2O, 0.57 and TiO2, 1.03 and lost on ignition, 2.31 wt % [11]. The specific surface area and bulk density of the fly ash were determined as 3.62 m2/g and 1.05 g/cm3. The surface of adsorbent was characterized by scanning electron microscopy (SEM, Philips XL30S-FEG) before and after the adsorption experiments using RR dye (Fig. 2). The SEM image clearly shows that the fly ash particles are mainly composed of irregular and porous particles. The pores in Fig. 2b are more densely packed with dyes than that of Fig. 2a which is no-adsorption of dye on the adsorbent. 3.2. XRD measurements X-ray diffraction (XRD) has long been used as a definitive technique for identifying minerals and other crystalline phases in a wide range of materials. The XRD for the adsorbents were mea-

286

S. Kara et al. / Desalination 212 (2007) 282293

(a) Fly ash before adsorption process

(b) Fly ash particle with dye adsorbed

Fig. 2. Typical SEM micrograph of fly ash (magnification: 2000) (a) before dye adsorption (b) with dye adsorbed. The arrows show no-adsorption of dye on the adsorbent.

sured with an automated Rigaku X-ray diffractometer D-Max Rint 2200 Series instrument using Cu K radiation at 40 kV and 40 mA over the range (2 of 570). The XRD pattern of fly ash is shown in Fig. 3. The XRD pattern indicates that the major phases for the sample are quartz, mullite with some magnetite and calcite. According to ASTM standards, the fly ash is divided into two types. If the sum of SiO2, Al2O3, and Fe2O3 is 70%, it is named type F. If the sum is upto 50%, it is named type C. The fly ash is a C class fly ash with 23.6% lime CaO as a major constituent. 3.3. FTIR measurements FTIR technique was used to examine the surface groups of the adsorbent and to identify those groups responsible for dye adsorption. Adsorption in the IR region takes place because of rotational and vibrational movements of the molecular groups and chemical band of a molecule. The IR spectra of the samples were recorded on a Bio Rad FTS 175 C spectrophotometer using a pellet (pressed-disk) technique. For this, the adsorbent was intimately mixed with approximately 100 mg of dry, powdered KBr. The mixture was pressed

under a pressure of 10,00015,000 psi into a transparent disk. The infrared spectra of fly ash and dye-loaded fly ash samples before and after the adsorption process were recorded in the range 4000400 cm1 (Fig. 4). The adsorption bands in the region 1637595 cm1 were assigned to SO3 and N=N groups on the dyes. The strong bands at 993.3, 1120.9 and 1149.4 cm1 regions are attributed to S=O stretching and the bands at 1626 1637.8 cm1 to N=N stretching. Strong SiO bands at 8751121 cm1 are shifted to higher frequencies as a function of chemical interaction of the dyes with fly ash. Adsorption band at 3645 cm1 is assigned to free hydroxyl. All these findings suggest that the dye on fly ash is held by chemical activation or chemisorption, probably indicating fly ash/dye complexation [11,20]. 3.4. Modeling of the percentage removal The adsorption experiments were carried out using various particle sizes (40125 m) and adsorbent dosages (50015000 mg/L) of the fly ash at pH 6, initial concentration of 100 mg/L, temperature of 22C and agitation speed of 250 rpm. A plot of p for any adsorbent dosage and geomet-

S. Kara et al. / Desalination 212 (2007) 282293

287

Fig. 3. XRD pattern of fly ash.

Fig. 4. FTIR spectra of fly ash and dye-loaded fly ash.

288

S. Kara et al. / Desalination 212 (2007) 282293

ric mean size of the adsorbent particle as a function of adsorbent dosage for various adsorbent sizes is shown in Figs. 5a and 5b. The figure shows that the percent dye removal at equilibrium increased with increasing adsorbent dosage while decreasing with the particle size. The relatively higher adsorption with smaller adsorbent particle may be attributed to the fact that smaller particles yield large surface areas and the availability of more adsorption sites. The minimum amount of fly ash corresponding to the maximum adsorption is declared as the optimum dosage. The optimum dose observed in the present study is 10 g/L. The removal efficien(RR) (a)
0.021
dp (m) 45 56.5 71.5 90 112.5

cies of RR, RB and RY reactive dyes with the particle sizes ranged from 45 to 112.5 m at the optimum dosage were calculated as 97%, 87% and 82%, respectively (Table 2). Plots in Fig. 5a suggest an empirical model form as [21,22]:

p=

Ws a + bWs

(2)

which can be linearized as


1 1 +b =a Ws p
(RB) (RY)
0.30
dp (m) 45 56.5 71.5 90 112.5

(3)

0.16
dp (m) 45 56.5 71.5 90 112.5

0.018

0.12

0.24 0.18 0.12

1/p

0.015

0.08

0.012

0.04 0.06

0.009 0.0000 0.0006 0.0012 0.0018 0.0024 0.00 0.0000 0.0006 0.0012 0.0018 0.0024 0.00 0.0000 0.0006 0.0012 0.0018 0.0024

1/WS (b)
100 90 80 70 60 50 30
WS (mg/L) 500 1000 3000 6000 10000 15000 150 180

100 80 60 40 20 0 30
WS (mg/L) 500 1000 3000 6000 10000 15000 150 180

100 80 60 40 20 0 30
WS (mg/L) 500 1000 3000 6000 10000 15000

60

90

120

60

90

120

60

90

120

150

180

dP (m)

Fig. 5. Plots of (a) the dosage and (b) particle size of the fly ash on the percentage adsorption of RB, RR and RY reactive dyes from aqueous solutions [Eqs. (2) and (4)].

S. Kara et al. / Desalination 212 (2007) 282293

289

Table 2 The removal efficiency and adsorbent capacity of the reactive dyes on the adsorbent with respect to particle size at various adsorbent dosages
Particle diameter, Adsorbent dosage, dP (m) WS (mg/L) 45 500 1000 3000 6000 10000 15000 500 1000 3000 6000 10000 15000 500 1000 3000 6000 10000 15000 500 1000 3000 6000 10000 15000 500 1000 3000 6000 10000 15000 RR Removal (%) qe (mg/g) 64.5 75.1 87.5 93.8 97.2 98.4 56.9 75.1 87.0 93.7 97.5 98.8 49.5 75.0 86.5 93.6 97.9 99.0 43.7 74.9 86.1 93.6 98.3 99.3 35.2 74.8 85.7 93.4 98.6 99.9 129 75 29 16 10 7 114 75 29 16 10 7 99 75 29 16 10 7 87 75 29 16 10 7 70 75 29 16 10 7 RB Removal (%) qe (mg/g) 8.4 16.9 44.6 75.2 94.5 97.4 6.7 14.7 43.1 73.1 93.7 97.2 5.6 12.5 40.9 71.1 92.4 96.7 4.9 11.1 41.1 70.2 90.9 95.3 3.7 9.4 37.6 68.0 88.3 94.8 17 17 15 13 9 6 13 15 14 12 9 6 11 12 14 12 9 6 10 11 14 12 9 6 7 9 13 11 9 6 RY Removal (%) qe (mg/g) 14.1 26.6 46.1 67.5 84.6 98.3 12.6 24.0 45.1 67.3 83.8 97.9 11.1 22.4 44.2 66.4 83.2 97.4 9.9 21.3 43.1 65.9 82.9 96.8 7.0 17.9 41.4 64.3 81.7 96.6 28 27 15 11 8 7 25 24 15 11 8 7 22 22 15 11 8 6 20 21 14 11 8 6 14 18 14 11 8 6

56.5

71.5

90

112.5

where a and b are model coefficients. The values of a and b along with the correlation coefficients for various adsorbent sizes can be obtained from the regression analysis (Table 3). The values of a and b as a function of mean particle size of the adsorbent (dp) is then substituted in equations shown in Table 5 to obtain the predictive model for p which is illustrated in Table 6. p values calculated from the model equation shown in Table 6 are plotted against experimental values (Fig. 6a).

The conformity between the experimental data and the model-predicted values was expressed by the correlation coefficients (r2). The best fit is obtained for RR dye (r2 = 0.997) as compared the rest of the reactive dyes (Fig. 6). In addition to that, the deviation between the theoretical and the experimental data is expressed with the standard deviation. RR dye loaded fly ash has lowest in value (SD = 4.11). To provide an alternative predictive model for

290

S. Kara et al. / Desalination 212 (2007) 282293

Table 3 Values of the coefficients a and b from Eq. (2) for various particle sizes, dp
Dye dp (m) 45.0 56.5 71.5 90.0 112.5 45.0 56.5 71.5 90.0 112.5 45 56.5 71.5 90 112.5 a 3.021 3.206 3.362 3.591 3.958 31.141 35.548 40.795 47.424 57.191 38.214 42.565 50.633 60.363 91.063 b 0.01002 0.01007 0.01011 0.01016 0.01020 0.0096 0.0089 0.0080 0.0069 0.0052 0.0095 0.0091 0.0079 0.0049 0.0016 r2 0.991 0.990 0.991 0.991 0.993 0.997 0.999 0.999 0.997 0.994 0.961 0.963 0.978 0.995 0.997

RB

RR

RY

Table 4 Values of the coefficients c and d from Eq. (4) for various adsorbent dosages, Ws
Dye Ws (mg/l) 500 1000 3000 6000 10000 15000 500 1000 3000 6000 10000 15000 500 1000 3000 6000 10000 15000 c 67.114 79.837 89.821 96.879 100.390 101.400 17.620 26.974 48.900 69.828 86.195 99.353 17.195 21.439 46.830 78.764 95.748 101.250 d 0.1059 0.0749 0.0580 0.0505 0.0370 0.0363 0.0866 0.0802 0.0711 0.0474 0.0397 0.0263 0.1062 0.1004 0.0896 0.0831 0.0621 0.0491 r2 0.997 0.992 0.976 0.984 0.990 0.987 0.982 0.987 0.944 0.965 0.990 0.987 0.989 0.994 0.995 0.999 0.999 0.992

RB

RR

RY

S. Kara et al. / Desalination 212 (2007) 282293 Table 5 Correlations between the coefficients a and b from Eq. (2) and c and d from Eq. (4) with respect to adsorbent dosage and particle size
Dye RB RR RY Dye RB RR RY Model equation: p = (Ws/a + bWs) a = 2.4161 + 0.0135 dp b = 0.00992 2.62106 dp a = 13.788 + 0.3812 dp b = 0.01256 6.45105 dp a = 0.7971 + 1.697 dp b = 0.0157 1.216104 dp Model equation: p = c + d (dp) c = 77.459 + 0.002 Ws d = 0.0834 + 3.9106 Ws c = 25.467 + 0.0055 Ws d = 0.0823 + 4.1106 Ws c = 24.16 + 0.0061 Ws d = 0.1046 + 3.87106 Ws r2 0.994 0.978 0.998 0.997 0.973 0.964 r2 0.703 0.692 0.923 0.938 0.877 0.810

291

p, plots depicted in Fig. 5b were subjected to linear regression analysis, according to a model of the form [21,22]
p = c + d (d p )

(4)

where c and d are coefficients. The values of c and d along with the regression analysis to obtain equations with respect to the various adsorbent dosages are presented in Tables 46. Fig. 6b shows the best fit between experimental and model predicted values obtained for RR dye. The standard deviation between calculated and experimental values is 4.94. Finally, to select the best predictive model, the simple statistic is to determine the standard deviation between calculated and experimental values. From this point, model equation provided in

100 80 60

RR

100 90 80 70 60 50 50

RB

100 80 60 40 20 0

RY

a)
40 20

Theoretical

20
RR

40

60

80

100

60
RB

70

80

90

100

20
RY

40

60

80

100

100 80 60

100

100 80 60

90

b)
40 20 0

80

40
70

20 0

20

40

60

80

100

60 60

70

80

90

100

20

40

60

80

100

Experimental

Fig. 6. Plots of the calculated values of p derived from (a) Eq. (2) and (b) Eq. (4) for three reactive dyes against the corresponding experimental values.

292

S. Kara et al. / Desalination 212 (2007) 282293

Table 6 Results from two alternative generalized predictive models for p


Dyes RB Model equations

p=

Ws

( 2.4161 + 0.0135 d p ) + ( 0.0092 2.62 106 d p )Ws

p = 77.459 + 0.002 Ws 0.0834 + 3.9 106 Ws d p


RR

p=

Ws

(13.788 + 0.3812 d p ) + ( 0.01256 6.45 105 d p )Ws

p = 25.467 + 0.0055 Ws 0.0823 + 4.1 106 Ws d p


RY

p=

Ws

( 0.7971 + 1.697 d p ) + ( 0.0157 1.216 10 4 d p ) Ws

p = 24.160 + 0.0061 Ws 0.1046 + 3.87 106 Ws d p

Table 5 with a lower standard deviation was the more successful in predicting the p values.
[2]

4. Conclusions Fly ash, a waste residue generated in a substantial amount in during coal-fired electric power generation, was used as adsorbent in this study. The adsorbent was characterized with a number of techniques such as FTIR, XRD and SEM. The percentage removals of dyes using predictive models as a function of the particle size and adsorbent dosage were calculated. Both models showed high values of coefficients (r2) but the best model which determines the p values can be expressed in terms of the standard deviation. The lowest value of standard deviation determines for predicting of the p values. The adsorption of the reactive dyes increased with increasing adsorbent dosage, and increased with decreasing the particle size. References
[1] C.ONeill, F.R. Hawkes, D.L. Hawkes, N. Lourenco, H.M. Pinheiro and W. Delee, Colour in textile ef[3] [4]

[5]

[6]

[7]

[8]

[9]

fluents sources, measurement, discharge consents and simulation: a review. J. Chem. Technol. Biotechnol., 74(11) (1999) 10091018. Y.M. Slokar and A.M. Le Marechal, Methods of decoloration of textile wastewaters. Dyes Pig., 37(4) (1997) 335356. P.K. Ray, Environmental-pollution and cancer. J. Sci. Ind. Res., 45(78) (1986) 370371. T. Viraraghavan and K.R. Ramakrishna, Fly ash for color removal from synthetic dye solutions. Water Qual. Res. J. Can., 34(3) (1999) 505517. S.J.T. Pollard, G.D. Fowler, C.J. Sollars and R. Perry, Low-cost adsorbents for waste and wastewater treatment: a review. Sci. Total Environ., 116(12) (1992) 3152. P.K. Malik, Dye removal from wastewater using activated carbon developed from sawdust: adsorption equilibrium and kinetics. J. Hazard. Mater., 113(13) (2004) 8188. C. Namasivayam, R. Radhika and S. Suba, Uptake of dyes by a promising locally available agricultural solid waste: coir pith. Waste Manag., 21(4) (2001) 381387. S.D. Khattri and M.K. Singh, Colour removal from dye wastewater using sugar cane dust as an adsorbent. Adsorp. Sci. Technol., 17(4) (1999) 269282. K. Kadirvelu, M. Palanival, R. Kalpana and S. Rajeswari, Activated carbon from an agricultural byproduct, for the treatment of dyeing industry waste-

S. Kara et al. / Desalination 212 (2007) 282293 water. Biores. Technol., 74(3) (2000) 263265. [10] Y.S. Ho and G. McKay, A kinetic study of dye sorption by biosorbent waste product pith. Res. Conser. Rec., 25(34) (1999) 171193. [11] B. Acemioglu, Adsorption of Congo red from aqueous solution onto calcium-rich fly ash. J. Colloid Interf. Sci., 274 (2004) 371379. [12] D. Mohan, K.P. Singh, G. Singh and K. Kumar, Removal of dyes from wastewater using flyash, a lowcost adsorbent. Ind. Eng. Chem. Res., 41(15) (2002) 36883695. [13] P. Janos, H. Buchtova and M. Ryznarova, Sorption of dyes from aqueous solutions onto fly ash. Water Res., 37(20) (2003) 49384944. [14] S.K. Khare, K.K. Panday, R.M. Srivastava and V.M. Singh, Removal of Victoria blue from aqueous solution by fly-ash. J. Chem. Technol. Biotechnol., 38(2) (1987) 99104. [15] K.V. Kumar, V. Ramamurthi and S. Sivanesan, Modeling the mechanism involved during the sorption of methylene blue onto fly ash. J. Colloid Interf. Sci., 284 (2005) 1421. [16] A.E. Muftuoglu, B. Karakelle, M. Ergin, A.Y. Erkol

293

[17]

[18] [19]

[20]

[21]

[22]

and F. Yilmaz, The removal of basic Blue 41 dye from aqueous solutions by bituminous shale. Adsorp. Sci. Technol., 21(8) (2003) 751760. S. Wang, Y. Boyjoo, A. Choueib and Z.H. Zhu, Removal of dyes from aqueous solution using fly ash and red mud. Water Res., 39(1) (2005) 129138. Z. Al-Qodah, Adsorption of dyes using shale oil ash. Water Res., 34(17) (2000) 42954303. M.L. Hall and W.R. Livingston, Fly ash quality, past, present and future, and the effect of ash on the development of novel products. J. Chem. Technol. Biotechnol., 77 (2002) 234239. P.M. Silverstein, G.C. Bassler and T.C. Morrill, Spectrometric Identification of Organic Compounds. Wiley, New York, 1991. D.S. Bhargava and S.B. Sheldarkar, Use of TNSAC in phosphate adsorption studies and relationships. Effects of adsorption operating variables and related relationships. Water Res., 27(2) (1993) 313324. M. Kobya, E. Demirbas and M. Bayramoglu, Modeling the effects of adsorbent dose and particle size on the adsorption Cr(VI) ions from aqueous solutions. Adsorp. Sci. Technol., 22(7) (2004) 583594.

Anda mungkin juga menyukai