You are on page 1of 8

Environmental Pollution 107 (2000) 391398

www.elsevier.com/locate/envpol

Adsorption of phenol by bentonite


F.A. Banat a,*, B. Al-Bashir b, S. Al-Asheh a, O. Hayajneh b
a

Department of Chemical Engineering, Jordan University of Science and Technology, Irbid 22110, Jordan b Department of Civil Engineering, Jordan University of Science and Technology, Irbid 22110, Jordan Received 9 February 1999; accepted 14 June 1999

``Capsule'': Factors aecting the adsorption of phenol by bentonite were investigated.


Abstract The potential of bentonite for phenol adsorption from aqueous solutions was studied. Batch kinetics and isotherm studies were carried out to evaluate the eect of contact time, initial concentration, pH, presence of solvent, and the desorption characteristics of bentonite. The adsorption of phenol increases with increasing initial phenol concentration and decreases with increasing the solution pH value. The adsorption process was signicantly inuenced by the solvent type in which phenol was dissolved. The anity of phenol to bentonite in the presence of cyclohexane was greater than that in water and was lowest in the presence of methanol. Methanol was used to extract phenol from bentonite. The degree of extraction was dependent on the amount of phenol adsorbed by bentonite. X-ray diraction analysis showed that the crystalline structure of bentonite was destroyed when cyclohexane was used. The ability of bentonite to adsorb phenol from cyclohexane decreased as the water to cyclohexane ratio was increased. Furthermore, hysteresis was observed in phenol desorption from bentonite in aqueous solutions. The equilibrium data in aqueous solutions was well represented by the Langmuir and Freundlich isotherm models. The removal of phenol from aqueous solutions was observed without surface modication. # 2000 Elsevier Science Ltd. All rights reserved.
Keywords: Bentonite; Phenol; Adsorption; Desorption; X-ray diraction

1. Introduction Phenolic compounds are common contaminants in wastewaters, being generated from petroleum and petrochemical, coal conversion, and phenol-producing industries. Phenols are widely used for the commercial production of a wide variety of resins including phenolic resins, which are used as construction materials for automobiles and appliances, epoxy resins and adhesives, and polyamide for various applications (Fang and Chen, 1997). Phenols are considered as priority pollutants since they are harmful to organisms at low concentrations and many of them have been classied as hazardous pollutants because of their potential harm to human health. Stringent US Environmental Protection Agency (EPA) regulations call for lowering phenol content in the wastewater to less than 1 mg/l (Dutta et al., 1992). Dutta et al. (1998) classied the treatment processes for phenolic wastewater into two principal

* Corresponding author. Fax: +962-2-295123. E-mail address: banatf@just.edu.jo (F.A. Banat).

categories: destructive process such as destructive oxidation with ozone (Hoigne, 1985), hydrogen peroxide (Kochany and Bolton, 1992), or manganese oxides (Ukrainczyk and McBride, 1992); and recuperative processes such as adsorption into porous solids (Danis et al., 1998), membrane separation (McGray and Ray, 1987), and solvent extraction (Eahart et al., 1977). Activated carbons are among the most eective adsorbents, having high surface area per unit mass. Activated carbon exhibits high adsorption capacity for phenolic compounds. Literature on the adsorption of phenolic compounds onto activated carbon is abundant (Cooney and Xi, 1994; Bercic and Pintar, 1996; Mollah and Robinson, 1996; Halhouli et al., 1997; Khan et al., 1997; Kildu and King, 1997). Due to the relatively high cost of activated carbons there have been attempts to utilize low cost, naturally occurring adsorbents to remove trace organic and inorganic contaminants from wastewaters (Darwish et al., 1996). In recent years, there has been increasing interest in utilizing natural clay minerals like montmorillonite, kaolinite, and illite for the removal of toxic metals and some organic pollutants from aqueous solutions (Boyd et al., 1988; Brigatti et

0269-7491/00/$ - see front matter # 2000 Elsevier Science Ltd. All rights reserved. PII: S0269-7491(99)00173-6

392

F.A. Banat et al. / Environmental Pollution 107 (2000) 391398

al., 1995; Gutierrez and Fuents, 1996; Lo et al., 1997). Bentonite consists essentially of clay minerals of the smectite (montmorillonite) group and has a wide range of industrial applications including clarication of edible and mineral oils, paints, cosmetics, and pharmaceuticals (Christidis, 1998). The abundance of bentonite in most continents of the world and its low cost make it a strong candidate as an adsorbent for the removal of many pollutants from wastewaters. Research studies have shown its ability to bind and remove pathogenic viruses, pesticides, herbicides, and other toxins (Hartman and Martin, 1984; Lipson and Stotzky, 1985). Other studies were carried out to investigate the possible use of natural bentonite as an eective adsorbent for the removal of rare earth elements and heavy metals from aqueous solutions (Chegrouche et al., 1997; Mellah and Chegrouche, 1997). Chegrouche et al. (1997) investigated the eectiveness of bentonite in removing lanthanum from aqueous solutions, reporting that natural bentonite is a promising adsorbent for lanthanum removal. In another study, Mellah and Chegrouche (1997) examined the feasibility of using natural bentonite for the removal of zinc from aqueous solutions. The eect of operating parameters such as agitation speed, solidliquid ratio, temperature, and initial zinc concentration, together with equilibrium isotherms were studied. They found that zinc adsorption onto natural bentonite is physical in nature and that the equilibrium adsorption isotherms are well described by the Langmuir and Freundlich model. Viaraghavan and Kapoor (1994) studied the potential of bentonite for adsorbing mercury from wastewater. However, they found that for an initial concentration of approximately 1 mg/l, only 34% of mercury was adsorbed by bentonite. The aim of this work was to investigate, experimentally, the potential of natural bentonite to adsorb phenolic pollutants using phenol as a model component. Laboratory batch kinetics and isotherm studies were conducted to evaluate the adsorption capacity of natural bentonite. The eects of contact time, pH, and initial adsorbate concentration were studied. In addition, the eect of solvent type on adsorption of phenol to bentonite was explored using three solvents: water, methanol, or cyclohexane. X-ray diraction analysis was carried out to study the eect of phenol adsorption on the crystalline structure of bentonite. 2. Materials and methods 2.1. Sorbent The bentonite used in this study was purchased from Halewood Chemicals Ltd. (Middlesex, UK). The bentonite was used in the batch experiments without any pretreatment.

2.2. Batch adsorption test Batch adsorption experiments were carried out by allowing an accurately weighted amount of bentonite to reach equilibrium with phenol solutions of known concentrations. Initial concentrations of phenol were held between 25 and 500 mg/l. The pH was adjusted using dilute HCl or NaOH solutions. Known weights of bentonite (2.5 g) were added to narrow-neck bottles each containing 50 ml solution. The bottles were darkly brown-colored to prevent photoxidation. The bottles were subsequently capped with screw caps tted with Teon liners and shaken in a temperature-controlled water bath shaker (B. Braun, model Infors AG, Bottmingen, Germany). Preliminary kinetic experiments showed that adsorption equilibrium was reached within 48 h. At the end of the equilibrium period the contents of the bottles were ltered, centrifuged for 1 h at 1200g using a 4124 Bestell centrifuge (Haraeus-Christ, Germany) and the supernatant was subsequently analyzed for residual concentration of phenol. 2.3. Analysis of phenol When water was the solvent, the concentration of phenol was determined following the method of Gales and Booth (1976) which is based on spectrophotometric analysis of the developed color resulting from the reaction of phenol with 4-aminoantipyrine. The nal equilibrium concentrations were determined spectrophotometrically using a Spectronic 21 UVD spectrophotometer (Milton Roy Company, USA). When methanol and cyclohexane were the solvents, phenol concentrations were measured using a Varian 3400 gas chromatography apparatus (Varian Inc., USA). The amount adsorbed was determined from the initial and nal concentration of phenol in the liquid phases. All experiments were run in triplicates to ensure reproducibility. 2.4. Desorption test After adsorption experiments, the phenol-laden bentonite was agitated with 50 ml distilled water for 48 h. The suspension was then centrifuged and the amount of desorbed phenol was estimated by measuring phenol in the supernatant solution. Repetitive cycles of phenol desorption were performed in batch experiments. 2.5. X-ray diraction The eect of phenol adsorption from water, methanol, and cyclohexane solvents on basal spacing of bentonite was studied using the X-ray diraction analysis technique. The analysis was carried out using water solvent system with three initial phenol concentrations

F.A. Banat et al. / Environmental Pollution 107 (2000) 391398

393

(0, 250, and 500 mg/l), methanol solvent system with two initial phenol concentrations (0 and 500 mg/l), and cyclohexane solvent system with two initial phenol concentrations (0 and 500 mg/l). The samples were prepared, agitated continuously for 48 h, then the oriented samples were prepared by allowing a few drops of each suspension to be dried slowly on a glass slide. The oriented samples were analyzed using an X-ray diractometer (Philips PW 1729). Then basal spacing of each sample was calculated using Bragg's law: I where d is the basal spacing (A),  is the angle of diffraction ( ), l is the wavelength (nm), and n is the path dierences between the reected waves which equals an integral number of wavelengths (l). 3. Results and discussion 3.1. Eect of operating variables 3.1.1. Contact time The adsorption data for the uptake of phenol versus contact time at dierent initial concentrations is presented in Fig. 1. The results show that equilibrium time required for the adsorption of phenol on bentonite is almost 6 h. It is also seen that the remaining concentration of phenol becomes asymptotic to the time axis after 6 h of shaking. However, for subsequent experiments, the samples were left for 48 h to ensure equilibrium. These results also indicate that the sorption process can be considered very fast because of the largest amount of phenol attached to the sorbent within the rst 6 h of 2d sin  nlY

adsorption. It was also seen that an increase in initial phenol concentration results in increased phenol uptake. The kinetics results of Fig. 1 can be used to determine if particle diusion is the rate-limiting step for phenol adsorption onto bentonite. Weber and Morris (1963) reported that if particle diusion is involved in the sorption process then a plot of adsorbate uptake versus the square root of time would result in a linear relationship and that particle diusion would be the ratecontrolling step if this line passes through the origin. As elucidated in Fig. 2, the results can be represented by such a linear relationship but they do not pass through the origin. This indicates that particle diusion is involved in the sorption process but it is not the only rate-limiting mechanism (Poots et al., 1976) and that some other mechanisms are involved. 3.1.2. Eect of pH The adsorption of phenol by bentonite was studied at various pH values. Dierent initial concentrations of phenol were prepared (in the range of 25500 mg/ml) and adjusted to dierent pH values of 5, 8, and 11. Bentonite was added to make its concentration 50 mg/ ml and sorption was carried out until equilibrium. The results are displayed in Fig. 3. The pH was measured before and after adsorption process and it has been found that the dierence between the two measured values of pH was less than 0.2 for all samples. As was expected, the adsorbed amount decreases with increasing the pH value (Fig. 3). This can be attributed to the dependency of phenol ionization on the pH value. The ionic fraction of phenolate ion (9ions ) can be calculated from: 1 X P 9ions 1 10pK pr

Fig. 1. Dynamics of phenol uptake by bentonite for various initial phenol concentrations.

Fig. 2. Plots of phenol adsorption versus square root of time for various initial phenol concentrations.

394

F.A. Banat et al. / Environmental Pollution 107 (2000) 391398

Obviously, 9ions increases as the pH value increased. Accordingly, phenol, which is a weak acid (pKa=10), will be adsorbed to a lesser extent at higher pH values due to the repulsive forces prevailing at higher pH values. Similar behavior has been reported by Halhouli et al. (1997) for the adsorption of phenol by activated carbon. 3.1.3. Initial phenol concentration As has been shown previously (Fig. 1) the adsorption of phenol by bentonite increases as the initial phenol concentration increased. This is also inconsistent with the results of Fig. 3. Increasing the initial phenol concentration would increase the mass transfer driving force and therefore the rate at which phenol molecules pass from the bulk solution to the particle surface. This would result in higher phenol adsorption. On a relative basis, however, the percentage adsorption of phenol decreases (Fig. 4) as the initial phenol concentration increases. 3.2. Adsorption isotherms Several models have been published in the literature to describe experimental data of adsorption isotherms. The Langmuir and Freundlich models are the most frequently employed models. In this work, both models were used to describe the relationship between the amount of phenol adsorbed and its equilibrium concentration in solutions at dierent pH values. The linear form of the Langmuir isotherm model can be represented by the following relation: 1 1 1 1 Y qe Qo bQo Ce Q

where qe is the amount adsorbed at equilibrium (mg/g), Ce is the equilibrium concentration of the adsorbate (mg/l), and Qo (mg/g) and b (l/mg) are the Langmuir constants related to the maximum adsorption capacity and the energy of adsorption, respectively. These constants can be evaluated from the intercept and the slope of the linear plot of experimental data of 1/qe versus 1/ Ce. The linear form of the Freundlich isotherm model is given by the following equation: ln qe ln kp 1 ln Ce Y n R

where kF (mg/g)(l/mg)1/n and 1/n are Freundlich constants related to adsorption capacity and adsorption intensity, respectively, of the sorbent. The values of kF and 1/n can be obtained from the intercept and slope, respectively, of the linear plot of experimental data of lnqe versus lnCe. The three equilibrium curves that were obtained at the three pH values in this study are well represented by the Freundlich isotherm model (Fig. 5). When the Langmuir isotherm model was applied to these data, a very good t was obtained at pH 5 and 8, while some deviation from experimental data was observed at pH 11 (Fig. 6). But in general, this model is also applicable to describe the experimental equilibrium data for all pH values. The Langmuir constants Q0 and b and Freundlich constants kF and 1/n at various pH values are displayed in Table 1. It is obvious that the parameters Q0 and kF, which are related to the sorption capacity, increase with a decrease in the pH values. This is consistent with the experimental observation. R2 values, which are a measure of goodness-of-t (Table 1), show

Fig. 3. Relationship between equilibrium phenol concentration and its adsorption for various pH values.

Fig. 4. Relationship between initial phenol concentration and its per cent adsorption for various pH values.

F.A. Banat et al. / Environmental Pollution 107 (2000) 391398

395

that both the Langmuir and Freundlich isotherm models can adequately describe the adsorption data. 3.3. Eect of solvents The feasibility of any adsorption process depends greatly on the cost of regeneration of spent adsorbents. Regeneration can be achieved either by thermal regeneration or solvent regeneration. Tamon et al. (1996) discussed the importance of clarifying the inuence of solvents on the adsorption process. If solvent regeneration is to be used, the amount adsorbed in the presence of the solvent should be much less than that by water (Tamon et al., 1990). Fig. 7 shows the adsorption isotherms of phenol onto bentonite using water, methanol, and cyclohexane solutions at 20 C. The adsorption isotherm for the case of water and methanol solvents conformed to an L-type isotherm, while for cyclohexane solvent the isotherm conformed to an S-type isotherm. The adsorption of phenol using cyclohexane solvent was greater than that using aqueous and methanol solvents. This phenomena could be attributed to the interaction of phenol with both solvent and sorbent. Phenol is known to have good solubility in water (8.2 g/100 ml) resulting from the
Table 1 Parameters of Langmuir and Freundlich isotherm models at various pH values pH Langmuir constants Q0 (mg/g) 5 7 11 1.712 0.842 0.433 b (l/mg) 0.0141 0.0158 0.0096 R2 0.963 0.985 0.941 Freundlich constants kF (mg/g)(l/g)1/n 0.100 0.075 0.0214 1/n 0.473 0.390 0.478 R2 0.996 0.904 0.948

hydrogen bonding between water and phenol molecules which is promoted by the dipole moment of the two molecules which is 1.8 D for water and 1.6 D for phenol (Francis, 1987). Methanol, which has a dipole moment of 1.7 D, exhibits similar interaction with phenol as a result of the hydrogen bond initiated between phenol and methanol. Because of the non-polar characteristic property of cyclohexane, its interaction with phenol is very weak. Accordingly, the solubility of phenol in cyclohexane is limited relative to that in water and methanol. To achieve adsorption, the interaction between phenol and bentonite should be stronger than that between phenol and solvent. Phenol is adsorbed on negatively charged surfaces through hydrogen bonding and/or charge transfer complexes (Hamaker and Thompson, 1972; Boyd, 1982). The higher phenol

Fig. 6. Relationship between equilibrium phenol concentration and its adsorption at various pH values. Symbols are experimental and lines are predicted data using Langmuir model.

Fig. 5. Relationship between equilibrium phenol concentration and its adsorption by bentonite at various pH values. Symbols are experimental and lines are predicted data using Freundlich model.

Fig. 7. Adsorption isotherms of phenol by bentonite using dierent solvent systems.

396

F.A. Banat et al. / Environmental Pollution 107 (2000) 391398

adsorption by bentonite using water than that using methanol solvent is in agreement with the ndings of Mingelgrin and Gerstl (1983). According to those authors, when phenol was adsorbed from polar solution such as water and methanol, phenol adsorption from a highly polar solvent occurred preferentially. The results obtained in this study conrmed their conclusion. The results of Fig. 7 can also be explained in terms of the competition between phenol and the solvent. For example, in the presence of methanol, methanol could be adsorbed by bentonite and then block those sites that are available for the adsorption of phenol. However, the anity of cyclohexane to bentonite could be very low, so there are a larger number of sites on bentonite available for the adsorption of phenol. Because adsorption of phenol from aqueous system was greater than that from methanol solvent, methanol could be used as a regenerative solvent. This was done by carrying adsorption of phenol by bentonite, using dierent initial phenol concentrations, until equilibrium. After that, the cake which was loaded with phenol was separated from each sample by centrifugation and resuspended in methanol solution. Then, phenol, which was extracted from each of these samples, was measured in the supernatant. Consequently, a relationship between the amount of phenol extracted (%) and phenol uptake by bentonite can be obtained (Fig. 8). As shown (Fig. 8), the amount of phenol extracted by methanol increased as phenol uptake decreased. The eect of water-to-cyclohexane ratio on the adsorption of phenol by bentonite was conducted at a constant initial phenol concentration of 200 mg/l. As shown in Fig. 9, the ability of phenol to be adsorbed from cyclohexane decreased as the amount of water increased. This could be attributed to the formation of a diusing layer surrounding bentonite particles which prevents adsorption of phenol from cyclohexane. The

thickness of this diusing layer increases as the amount of water increases, causing greater resistance to phenol adsorption from cyclohexane. Clay minerals in the layered-silicate group, such as bentonite, generally have a plate-like shape. The distance separating two successive sheets is referred to as the basal spacing. X-ray diraction analysis was performed to investigate the eects of phenol adsorption and solvent system on the basal spacing of bentonite, and the subsequent eect on the interlamellar surfaceexposure for adsorption. The variation in the basal spacing of bentonite structure resulting from adsorption process is a strong indicator of the extent of sorbate reaching the interlamellar sorption sites (Timothy et al., 1985). Phenol adsorption from aqueous systems using initial concentrations of 250 and 500 mg/l increased basal spacing of bentonite by 5.54 and 9.12%, respectively (Table 2). Phenol adsorption, with initial concentration of 500 mg/l in the methanol system, showed an increase in basal spacing of bentonite by 1.92% (Table 2), which is lower than that of using aqueous systems. This conrms the results of lower phenol uptake using methanol than that using a water system. It can also be concluded from these results that there is

Fig. 9. Eect of moisture content on the adsorption of phenol by bentonite.

Table 2 Summary of the X-ray diraction analysis results System Bentonite+water Bentonite+water+250 mg/l phenol Bentonite+water+500 mg/l phenol Bentonite+methanol Bentonite+methanol+500 mg/l phenol Bentonite+cyclohexane Bentonite+cyclohexane+500 mg/l phenol 2y ( ) 6.095 5.775 5.585 5.845 5.735 ) d001 (A 14.501 15.304 15.824 15.121 15.411

Fig. 8. Desorption of phenol from bentonite using methanol.

F.A. Banat et al. / Environmental Pollution 107 (2000) 391398

397

a correlation between the basal spacing of bentonite and its phenol uptake. The presence of cyclohexane destroys the crystalline structure of bentonite (basal spacing equals innity). Because cyclohexane is non-polar, it will not be attracted to the charged surface functional groups. Therefore, it diuses between the silicate sheets of the bentonite particles weakening the attraction forces between binding silicate sheets in a crystalline structure and, consequently, converts it to an amorphous form. This results in a substantial increase in the surface area of bentonite which results in an S-shaped isotherm. 3.4. Adsorption/desorption The adsorption/desorption isotherms of phenol using aqueous solution were obtained in 48-h batch equilibrated experiment at 20 C. For each initial concentration investigated, triplicate samples were prepared and analyzed. The desorption isotherms were generated for each initial concentration spanning three consecutive equilibrium states. As shown in Fig. 10, the desorption isotherms are concave down (hyperbolic) which suggests that desorption increases with decreasing aqueous concentration (Al-Bashir, 1994). The hysteresis between adsorption and desorption indicates the presence of irreversible adsorption. These results are in good agreement with Tamon and Okazaki's (1996) ndings regarding the desorption of phenol from activated carbon. The authors measured adsorption and desorption equilibria of 11 kinds of aromatic compounds in their aqueous solutions using activated carbon. They reported that desorption was greatly dependent on substituent groups of aromatic compounds. In the desorption of aromatic compounds with electrondonating groups, hysteresis between adsorption and desorption was observed. On the other hand, com-

pounds with an electron-attracting group showed no hysteresis and the desorption was reversible. Phenol which has hydroxyl group belongs to the electrondonating group. 4. Conclusions The potential of bentonite to adsorb phenol from aqueous solution was assessed. In spite of bentonite's ability to adsorb phenol and, consequently, its possible utilization in the treatment of phenol-contaminated wastewaters, its adsorptive capacity was limited (of an order of magnitude of 1 mg/g). Bentonite adsorption capability was strongly dependent on the pH of the solution. The adsorption capacity was increased with a decrease in the pH and an increase in the initial phenol concentration. The possible regeneration of bentonite using dierent solvents was investigated by studying the adsorption equilibrium isotherms of phenol in water, methanol and cyclohexane. Because phenol adsorption using methanol was the lowest of the three solvents, methanol was used to extract the adsorbed phenol. The degree of extraction was dependent on the uptake of phenol bentonite. The basal spacing of bentonite was greatly aected by the solvent nature and the initial phenol concentration. The adsorption/desorption isotherms showed considerable hysteresis indicating irreversible adsorption occurs in aqueous solution. The Langmuir and Freundlich isotherms were found to be applicable for the adsorption equilibrium data. References
Al-Bashir, B., 1994. Bioavailability of nitrogen-substituted polycyclic aromatic hydrocarbons in ooded soil systems. PhD thesis, McGill University, Canada. Bercic, G., Pintar, A., 1996. Desorption of phenol from activated carbon by hot water regeneration, desorption isotherms. Industrial Engineering and Chemistry Research 35, 46194625. Boyd, S.A., 1982. Adsorption of substituted phenols by soil. Soil Science 13, 337343. Boyd, S.A., Shaobai, S., Lee, J.F., Mortland, M., 1988. Pentachlorophenol sorption by organoclays. Clay and Clay Minerals 36, 125130. Brigatti, M., Corrodini, F., Franchini, G., Mazzoni, S., Medici, L., Poppi, L., 1995. Interaction between montmorillonite and pollutants from industrial wastewaters: exchange of Zn2+ and Pb2+ from aqueous solutions. Applied Clay Science 9, 383395. Chegrouche, S., Mellah, A., Telmoune, S., 1997. Removal of lanthanum from aqueous solutions by natural bentonite. Water Research 31, 17331737. Christidis, G., 1998. Physical and chemical properties of some bentonite deposits of Kimolos Island, Greece. Applied Clay Science 13, 7998. Cooney, D.O., Xi, Z., 1994. Activated carbon catalyzes reactions of phenolics during liquid phase adsorption. AIChE Journal 40, 361364. Danis, T.G., Albanis, T.A., Petrakis, D.E., Pomonis, P.J., 1998. Removal of chlorinated phenols from aqueous solutions by adsorption on alumina pillared clays and mesoporous alumina aluminum phosphates. Water Research 32, 295302.

Fig. 10. Sorption/desorption isotherm of phenol by/from bentonite using aqueous solution.

398

F.A. Banat et al. / Environmental Pollution 107 (2000) 391398 Kochany, J., Bolton, J.R., 1992. Mechanism of photodegredation of aqueous organic pollutants. Environmental Science and Technology 26, 262265. Lipson, S., Stotzky, G., 1985. Specicity of virus adsorption to clay minerals. Canadian Journal of Microbiology 31, 5053. Lo, B., Mak, R., Lee, S., 1997. Modied clays for waste containment and pollutant attenuation. Journal of Environmental Engineering 123, 2532. McGray, S.B., Ray, R.J., 1987. Concentration of Synfuel processes condensate by reverse osmosis. Separation Science and Technology 22, 745. Mellah, A., Chegrouche, S., 1997. The removal of zinc from aqueous solutions by natural bentonite. Water Research 31, 621629. Mingelgrin, U., Gerstl, Z., 1983. Reevaluation of partitioning as a mechanism of nonionic chemicals adsorption in soils. Journal of Environmental Quality 12, 111. Mollah, A., Robinson, C.W., 1996. Pentachlorophenol adsorption and desorption characteristics of granular activated carbon: kinetics. Water Research 30, 29012907. Poots, V., McKay, G., Healy, J.J., 1976. The removal of acid dye from euent using natural adsorbents: peat. Water Research 10, 10611066. Tamon, H., Okazaki, M., 1996. Desorption characteristics of aromatic compounds in aqueous solutions on solid adsorbents. Journal of Colloid and Interface Science 179, 181187. Tamon, H., Abi, M., Okazaki, M., 1996. Solvent dependence of anity coecient for phenol adsorption on activated carbon. Journal of Chemical Engineering of Japan 29, 384386. Tamon, H., Saito, T., Kishimura, M., Okazaki, M., Toei, R., 1990. Solvent regeneration of spent activated carbon in wastewater treatment. Journal of Chemical Engineering of Japan 23, 426432. Timothy, A.W., Turgute, D., Robert, B.E., 1985. Interaction of aliphatic amines with montmorillonite to enhance adsorption of organic pollutants. Clay and Clay Minerals 33, 301. Ukrainczyk, L., McBride, M.B., 1992. Oxidation of phenol in acidic aqueous suspensions of manganese oxides. Clay and Clay Minerals 40, 157166. Viaraghavan, T.V., Kapoor, A., 1994. Adsorption of mercury from wastewaters by bentonite. Applied Clay Science 9, 3149. Weber, W.J., Morris, J.C., 1963. Kinetics of adsorption on carbon from solution. Journal Sanitory Engineering Division, American Society of Chemical Engineering 89, 3159.

Darwish, N., Halhouli, K., Al-Dhoon, N., 1996. Adsorption of phenol from aqueous systems onto spent oil shale. Separation Science and Technology 31, 705714. Dutta, N.N., Brothakur, S., Baruah, R., 1998. A novel process for recovery of phenol from alkaline wastewater: laboratory study and predesign cost estimate. Water Environmental Research 70, 49. Dutta, N.N., Patil, G.S., Brothakur, S., 1992. Phase transfer catalyzed extraction of phenolic substances from aqueous alkaline stream. Separation Science and Technology 27, 1435. Eahart, J.P., Won, K., Wang, H.Y., Prausnitz, J.M., 1977. Recovery of organic pollutants via solvent extraction. Chemical Engineering Progress 73, 67. Fang, H.H., Chen, O., 1997. Toxicity of phenol towards aerobic biogranules. Water Research 31, 22292242. Francis, C.A., 1987. Organic Chemistry. McGraw-Hill, New York. Gales, M.E., Booth, R.L., 1976. Automated 4AAP phenolic method. American Water Works Association 68, 540. Gutierrez, M., Fuents, H., 1996. A mechanistic modeling of montmorillonite contamination by cesium sorption. Applied Clay Science 11, 1124. Halhouli, K.A., Darwish, N.A., Al-Jahmany, Y., 1997. Eects of temperature and inorganic salts on the adsorption of phenol from multicomponent system on a decolorizing carbon. Separation Science and Technology 32, 30273036. Hamaker, J.W., Thompson, J.M., 1972. Adsorption. In: Goring, C.A., Hamaker, J.W., (Eds.), Organic Chemicals in the Soil Environment, Vol. 1. Marcel Dekker, New York, pp. 49143. Hartman, W.A., Martin, D.B., 1984. Eect of suspended bentonite clay on the acute toxicity of glyphosate to Daphnia pulex and Lemna minor. Bulletin of Environmental Contamination and Toxicology 33, 355361. Hoigne, J., 1985. Organic micropollutants and treatment processes: kinetics and nal eects of ozone and chlorine dioxide. Science of the Total Environment 47, 169185. Khan, A.R., Al-Bahri, T.A., Al-Haddad, A., 1997. Adsorption of phenol based organic pollutants on activated carbon from multicomponent dilute aqueous solutions. Water Research 31, 2102 2112. Kildu, J.E., King, C.J., 1997. Eect of carbon adsorbent surface properties on the uptake and solvent regeneration of phenol. Industrial Engineering and Chemistry Research 36, 16031613.

You might also like