Anda di halaman 1dari 9

J Porous Mater (2013) 20:619627 DOI 10.

1007/s10934-012-9635-5

Activated carbon derived from macadamia nut shells: an effective adsorbent for phenol removal
Liana Alvares Rodrigues Loriane Aparecida de Sousa Ribeiro nio Thim Rafael Reinaldo Ferreira Gilmar Patroc Manoel Orlando Alvarez-Mendez Aparecido dos Reis Coutinho

Published online: 5 October 2012 Springer Science+Business Media New York 2012

Abstract In this study, activated carbon based on the waste macadamia nut shells (MAC) was investigated for potential use as an adsorbent for phenol removal. The pseudo second-order kinetic model best described the adsorption process. The extent of the phenol adsorption was affected by the pH solution and the adsorbent dosage. Equilibrium data tted well to the Langmuir model with a maximum adsorption capacity of 341 mg g-1. The calculated thermodynamic parameters suggested that the phenol adsorption onto MAC was physisorptive, spontaneous and exothermic in nature. Phenol desorption from loaded adsorbent was achieved by using 0.1 mol L-1 NaOH, ethanol (100 %) and deionized water. Keywords Phenol Adsorption Activated carbon Macadamia

1 Introduction The chemical pollutant removal from wastewater is essential to reduce the harmful effect on the environment and human health [1]. Phenol has already been listed as one of the top priority contaminants and also the most important substructure of potentially carcinogenic pollutants
L. A. Rodrigues (&) L. A. de Sousa Ribeiro G. P. Thim gico de Aerona utica-ITA/CTA, Prac Instituto Tecnolo a Mal. dos Campos, Sa o Jose o Paulo CEP Eduardo Gomes 50, Sa 12228-900, Brazil e-mail: lika_eng@yahoo.com.br R. R. Ferreira M. O. Alvarez-Mendez A. d. R. Coutinho rio de Materiais Carbonosos, Universidade Metodista Laborato rbara d0 Oeste, de Piracicaba, Rod. SP 306, Km 01, Santa Ba o Paulo 13450-971, Brazil Sa

discharged from ne chemical plants [2, 3]. Their presence in water supplies is noticed by its bad taste and odor [4]. The concern increase for the public health and environmental issues have led to establish rigid limits on the acceptable environmental levels of specic pollutants [4]. The permissible concentration of phenolic contents in potable water is 1 lg L-1 according to the recommendation of the World Health Organization [5]. Consequently, water contaminated with phenol must be treated before discharge into water streams to avoid legal problems. Concern about the removal of phenol from wastewater is driven considerable efforts in the past years. Several methods have been proposed to achieve this proposal, including catalytic oxidation [6], photo-oxidation [7], electrochemical oxidation [8], biological degradation [9], ultraltration [10] and adsorption [1, 3]. Among these methods, adsorption is the only one who does not need special process requirements or catalyst. Adsorption is a simple and economical alternative for phenol removing [11]. Activated carbon is considered to be an effective adsorbent for the phenol removal from aqueous solution [12] due to its extended surface area, high adsorption capacity, microporous structure and specially the surface reactivity [4]. Activated carbon reactivity is directly related to the presence of carboxylic groups, hydroquinone free radicals, quinones, metallic ions and nitrogenous impurities on its surface [13]. However, the activated carbon use based on expensive starting materials (wood or coal) is unjustied for most pollution control applications [12]. The use of low cost starting materials (industrial or agricultural residues) as source for the activated carbon has emerged as a potential alternative to get the cost reduction, the increase in the biomass worth and remedy for the waste disposal as well [11]. Disposal of the macadamia nut waste shells has created a serious problem for the nut processing industries due to

123

620

J Porous Mater (2013) 20:619627

the global increase of the macadamia nut production [14]. One important and attractive property of the macadamia shell is its high cracking pressure, which allows the preparation of a durable activated carbon for adsorption [14]. Some volatile species are released during the pyrolysis process, which may lead to a material structure collapse if shell is not strength enough. On the other hand, materials with high cracking pressure are more able to maintain its structure during the gas releasing step. Nguyen and Do [15] prepared activated carbon from macadamia nut shell (MNS) using physical activation. Their results showed that MNS is a good starting raw material for the activated carbon production, having a well-developed structure and high surface area. Tam and co-workers compared activated carbon from macadamia nut coconut shells considering the pore structure and cost production. They concluded that macadamia nut shells are inherently better suited for the production of activated carbon by air oxidation than coconut shells [16]. Poinern and co-workers also compared activated carbon from macadamia nut coconut shells taking into account their adsorption capacity of gold and they settled that their capacities are very similar [14]. Therefore the proposal of this work is to convert macadamia nut shell into activated carbon, as an alternative solution to the shell disposal problem produced by the nut processing industries. Therefore, the focus of this research was to evaluate the potential use of the MAC as a phenol remover from aqueous solution in order to make a better use of this abundant agricultural waste and avoid the harmful effects of phenol on the human health.

area through the Brunauer, Emmett, and Teller (BET) equation [17]. The micropore area was obtained by applying the t-plot method [18]. The total pore volume was determined by converting the volume adsorbed at the saturation point P/P0 * 0.99 into liquid volume [11], whilst the micropore volume was calculated at the point of interception of the linear region of the t-plot after saturation of the micropores [18]. The mesopore volume was calculated from the difference between the total pore volume and the micropore volume [11]. The pore size distribution of MAC was determined based on the non-local density theory [19]. The content of the oxygenated chemical groups on the MAC surface were determined according to the Boehm method [20]. One gram of MAC was placed in 25 cm3 of the following aqueous solutions: sodium hydroxide, sodium carbonate, sodium bicarbonate, sodium ethoxide and hydrochloric acid. The vials were sealed and their contents were shaken for 24 h and then ltered. The ltrate was titrated with either HCl and NaOH to determine the base or acid excess, respectively. The point of zero charge (pHPZC)of the MAC sample was determined using a batch equilibration technique previously described [21]. The initial pH values (pH-i) of 25 cm3 of KNO3 aqueous solutions (concentrations 10-1 and 10-2 M) were adjusted in pH range of 28 using 0.1 M of HCl or NaOH. Then, 0.05 g of MAC was added to each sample. Equilibration was carried out in a thermostated orbital shaker for 24 h at 298 K. The dispersions were then ltered and the nal pH of the solutions (pHf) was determined. The point of zero charge was found from a plot of (pHipHf) versus pHi. 2.3 Adsorption studies The adsorption studies were carried out in amber glass asks of 100 cm3, where MAC was mixed to a phenol aqueous solution (50 cm3) under stirring in a thermostatic shaker. The supernatant solution was separated from the adsorbent by ltration. The phenol concentration in the supernatant was spectrophotometrically determined using UVVIS absorbance spectrophotometry at 269 nm [22]. Concentration higher than that detected by spectrophotometer was determined by dilution. These batch adsorption tests were conducted in independent duplicate, using two different lots of MAC. The results were the average between the duplicated experiments. The adsorption studies with phenol synthetic solutions were estimated using the following conditions. 2.3.1 Kinetic study MAC dose 2 g dm-3; initial phenol concentration 1500 mg dm-3, pH without adjustment; shaking time 0400 min; T = 298 K.

2 Materials and methods 2.1 Preparation of activated carbon MNS from Limeira-SP, Brazil, was used as a precursor in the present work. The precursor was mechanically crushed, milled and sieved to obtain particles of around 3.0 mm in diameter. MNS (500 g) was carbonized in a stainless steel tube reactor (1,000 mm length and 100 mm internal diameter) installed inside of a vertical electric furnace under nitrogen ow. Temperature was rise from room temperature to 1,123 K at a heating rate of 5 K min-1 and maintained for 1 h. Then, the nitrogen ow was turned off; CO2 ow was switched to 600 cm3 min-1 and held for 10 h. 2.2 Characterization Nitrogen adsorption isotherm at 77 K was determined in an Autosorb-1MP gas analyzer (Quantachrome Instruments). Adsorption data was used to determine the BET surface

123

J Porous Mater (2013) 20:619627

621

2.3.2 Effect of adsorbent dosage MAC dose 08 g dm-3, initial phenol concentration 1,500 mg dm-3, pH without adjustment; shaking time 24 h; T = 298 K. 2.3.3 Effect of pH MAC dose 2 g dm-3, initial phenol concentration 1500 mg dm-3, pH 2-13; shaking time 24 h; T = 298 K. 2.3.4 Adsorption isotherms MAC dose 2 g dm-3, initial phenol concentration 3001,500 mg dm-3, pH without adjustment; shaking time 24 h; temperature range 298328 K. 2.3.5 Thermodynamic study MAC dose 2 g dm-3, initial phenol concentration 1,500 mg dm-3, pH without adjustment; shaking time 24 h; temperature range 298328 K.

Table 1 Surface functional groups for MAC Functional group Carboxyl Lactones and lactols Phenols Total basic sites Total acid sites mmol g-1 0.00 0.00 0.15 0.71 0.15

Table 2 Physical properties of MAC SBET (m2 g-1) 1,083 Smicro (m2 g-1) 1,033 VT (cm3 g-1) 0.476 Vmicro (cm3 g-1) 0.413 rp (nm) 1.23

SBET BET surface area, Smicro micropore surface area, VT total pore volume, Vmicro micropore volume, rp average pore radius

350

300

250

Adsorption (cc/g)

2.3.6 Desorption The adsorbent used for the adsorption of 1,500 mg dm-3 of phenol solution was separated from the solution by ltration and was gently washed with deionized water to remove phenol molecules that were not adsorbed on MAC. Then this adsorbent was mixed with 50 cm3 of different desorbing and agitated in time intervals no longer than the equilibrium time. Different desorbing agents such as: deionized water, NaOH (0.1 M) and ethanol (100 %) were used in order to determine the most appropriate eluant or desorbing solution. These desorbing agents were chosen due to difference in pH value of solution (water and NaOH 0.1 M) and the nature of solvent (water and ethanol).

200

150

100

50

Adsorption Dessorption
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

P/Po

Fig. 1 N2 Adsorption isotherm of MAC

3 Results and discussion 3.1 Characterization Boehm titration results are presented in Table 1. Analysis of the data indicates that MAC has a considerable amount of surface groups, whereas, as shown in Table 1, the sample is predominantly basic. The pHPZC value is measured as 9.8, which conrms a basic surface. Table 2 shows the physical properties of MAC. This activated carbon shows a large BET surface area (1,083 m2 g-1). The micropore surface area, calculated by t-plot method, is 1,033 m2 g-1, which means that micropore surface area corresponds to 95 % of the BET surface

area. Figure 1 shows the isotherm and hysteresis loop, which can be classied as types 1 and H4, respectively, according to IUPAC classication [23]. Figure 2 shows the pore size distribution of the MAC determined by density functional theory as described in Sect. 2; one can see that it consists mainly of a microporous pore structure. The sample has micropore volume equal to 0.413 cm3 g-1, total pore volume as high as 0.476 cm3 g-1 and is 1.23 nm. 3.2 Kinetic study According to Fig. 3 the adsorption process of phenol is very fast. The phenol removal (qe) increases up to 15 min and then it is kept constant up to 360 min. The adsorbed amount at equilibrium (15 min) is 280 mg g-1, which corresponds to 35 % of initial phenol (Fig. 3).

123

622
Pore size distribution function (cc/g.nm)
0.08

J Porous Mater (2013) 20:619627

0.06

Adsorption kinetic model is often described by one of the following models: pseudo-rst order, pseudo-second order or intra particle diffusion. The pseudo rst-order model can be described by Eq. (1) [24]: logqe qt log qe k1 t 2:303 1

0.04

0.02

0.00 1 2 3 4 5 6 7 8 910 20 30 40 50

where, qt and qe are the mass of the adsorbed phenol at time t and at equilibrium, respectively, and k1 is the rate constant of pseudo rst-order adsorption process. The pseudo second-order model can be described by Eq. (2) [24]: t 1 t qt k 2 q2 q e e 2

Pore size (nm)

Fig. 2 Pore size distribution of MAC


290 280 270 260

250 240 230 220 210 -50 0 50 100 150 200 250 300 350 400

where, k2 is the pseudo second-order rate constant and qt and qe were described previously. The correlation coefcient of the linear tting of experimental data to the rstorder kinetic model is very low (Fig. 4). The tting process related to the pseudo second-order kinetic model (Fig. 5) gives a relatively high R2, indicating that the pseudo second-order kinetic model successfully describes the kinetics of phenol adsorption onto MAC. In addition, the value of qe calculated using the pseudo second-order model is very close to the experimental one, which conrms the adsorption system belongs to a second-order kinetic model. The intra particle diffusion model can be expressed by Eq. (3) [24]:
1500 mg dm
-3

qe (mg g )

-1

qt kin t0:5 cb

t (min)

Fig. 3 The variation of adsorption capacity with adsorption time at various initial phenol concentrations (T = 298 K; adsorbent dose = 2 g dm-3)
1.9 1.8 1.7 1.6 1.5 1.4 1.3 1.2 1.1 1.0 0 1 2 3 4 5

log (qe-qt) (mg g )

where qt is the mass of adsorbed phenol at time t, kin is a kinetic constant and cb is the thickness of the boundary layer. Figure 6 shows that the plot of qt versus t0.5 is multi linear, indicating that three steps took place. According to Cheung et al. [25] the rst portion was attributed to phenol diffusion from bulk solution to the external surface of the adsorbent. The second portion described the gradual adsorption stage, where intra particle diffusion was the rate limiting. The third portion was attributed to the nal equilibrium stage, where the intra particle diffusion started to slow down due to the low phenol concentration in the solution [25]. The intra particle diffusion has been reported to be the rate-limiting factor and in this study kin in a phenol concentration of 1,500 mg g-1 was found to be 5.59 mg g-1 min-1/2 with R2 values of 0.98 (Table 3). 3.3 Effect of adsorbent dosage The phenol removal was studied by varying the amount of adsorbent from 0.025 to 0.4 g. Figure 7 presents the removal efciency of the MAC. The amount of adsorbent signicantly inuenced on the extent of phenol adsorption. It can be observed that the removal efciency increased

-1

t (min)
Fig. 4 Pseudo-rst-order kinetics for adsorption of phenol onto MAC

123

J Porous Mater (2013) 20:619627


0.00 400 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40 0.45 80 70

623

0.020 0.018 0.016 0.014

350 60 300 50 40 30 200 20 150 10 0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40 0.45

0.010 0.008 0.006 0.004 0.002 0.000 0 1 2 3 4 5

qe (mg g )

0.012

250

t (min)

dosage (g)

Fig. 5 Pseudo-second-order kinetics for adsorption of phenol onto MAC


290 280 270 260

Fig. 7 Effect of adsorbent dose on the adsorption of phenol onto MAC (C0 = 1,500 mg dm-3; T = 298 K; t = 60 min)

adsorption sites, which is associated to the adsorbent overcrowding and resulting in a lower qe value [26]. Thus, an optimum dose of 0.1 g is selected for all the experiments. 3.4 Effect of pH The adsorption process is dependent on the pH due to the electric charge of adsorbate and adsorbent [11]. The phenol adsorption on MAC was studied at several initial pH (from 2 to 13) and is illustrated in Fig. 8. Phenol removal remained constant under acidic conditions and a slight decrease can be observed in the pH of 78. The phenol adsorption decreased abruptly when pH exceeded 10. The point of zero charge (pHpzc) of MAC was here determined and it is equal to 9.8. Therefore, the surface of MAC at pH values lower than 9.8 is positively charged and at pH values higher than 9.8 are negatively charged. Therefore the decrease in the phenol amount adsorbed as the pH exceeds 10 is attributed to both: an increase in the solubility of dissociated phenol at pH [ pKa and an increase in the repulsion forces between the dissociated form of the adsorbate and the negatively charged activated carbon surface [27]. From the experimental results, pH 4.6 (solution without adjustment) was selected as an optimum pH value.

qt (mg g )

-1

250 240 230 220 210 0 5 10 15


0.5

20

0.5

(min )

Fig. 6 Intraparticle diffusion kinetics for adsorption of phenol onto MAC

with the increase in the quantity of MAC due to an increase in the amount of the available adsorption sites [26]. The extent of phenol removal was found to be 11 % using 0.025 g of MAC and greatly increased to 73 % using 0.4 g. However, higher adsorbent dose results in a decrease in the adsorption capacity as shown in Fig. 7. The amount of the available adsorption sites decreases in the higher adsorbent dose due to the aggregation or overlapping of the

Table 3 Kinetic parameters for phenol adsorption onto MAC Co (mg dm-3) First-order-model k1 (min-1) 1,500 0.031 q1 (mg g-1) 46.01 R2 0.55 Second-order-model k2 (g mg-1 min-1) 0.005 q2 (mg g-1) 272 R2 0.99 Intraparticle diffusion kin (mg g-1 min-1/2) 5.592 R2 0.98

123

removal efficiency (%)

t qt (min g mg )

-1

-1

-1

624
300

J Porous Mater (2013) 20:619627 Table 4 Adsorption isotherms constants for phenol onto MAC Temperature (K)
250

Langmuir isotherm b (dm3 mg-1) Qm (mg g-1) 341 248 223 200 R2 0.98 0.97 0.97 0.99

200

298 308 318 328

0.015 0.012 0.007 0.007

qe (mg g )

-1

150

100

50

Table 5 Summary of adsorption capacity values of phenol on different activated carbons reported in literature Adsorbent
2 4 6 8 10 12 14

Qm (mg g-1) 341 90 149 333 50 34 32 64

Reference This work [11] [12] [27] [35] [35] [36] [37]

pH
Fig. 8 Effect of solution pH on the adsorption of phenol onto MAC (C0 = 1500 mg dm-3; T = 298 K; adsorbent dose = 2 g dm-3; t = 60 min)

Macadamia activated carbon Avocado activated carbon Rattan sawdust based activated carbon Organobentonite Acid treated coconut shells activated carbon Coconut shells activated carbon Commercial activated carbon Eucalyptus wood activated carbon

350

300 250

qe (mg g )

-1

200 150

100 50

25C 35C 45C 55C

0 0 200 400 600


-3

800

1000

1200

Ce (mg dm )
Fig. 9 Effect of temperature on adsorption of phenol onto MAC

3.5 Equilibrium isotherms The equilibrium adsorption isotherm is essential for describing the interactive behavior between adsorbate and adsorbent [28]. The adsorption data are shown in Fig. 9. Langmuir isotherm model was used to treat the adsorption data. The Langmuir model is [29]: Langmuir isotherm : qe Qm bCe 1 qe C e 4

Table 4 shows Langmuir isotherm model t the adsorption data well and the maximum adsorption capacity is strongly correlated to temperature. Experimental results indicated that the adsorption capacity decreases from 341 to 200 mg g-1, when the solution temperature increases from 298 to 328 K, which shows that the adsorption process is exothermic. Table 5 shows the maximum capacity values reported in literature for phenol adsorption on different activated carbons. One can see in Table 5 that the best material for phenol adsorption reported in literature was organobentonite, which showed a phenol retention capacity of 333 mg g-1. However, this value of 333 is lower than 341 mg g-1 achieved in this work using MAC, which can be associated to its higher surface area (1,083 m2 g-1). The value of removal attained by MAC appears to be high enough to consider it in the purication industrial process of phenol wastewaters. 3.6 Thermodynamic study Generally, the Gibbs free energy (DG), enthalpy (DH) and entropy (DS) changes of adsorption were calculated from interpolated data using the best tting isotherm. In this study, the Langmuir isotherm was used to calculate the thermodynamic parameters using the following equations [11]:

where qe is the adsorbed amount at equilibrium, Ce is the equilibrium concentration, Qm (mg g-1) is the maximum adsorption capacity and b is the binding constant which is related to the heat of adsorption. The nonlinear least-square regression was employed to analyze the results and they are presented in Table 4.

123

J Porous Mater (2013) 20:619627

625 Table 6 Thermodynamic parameters for phenol adsorption onto MAC T (K) 298 308 318 328 DG (kJ mol-1) -21.20 -20.38 -20.25 -19.77 DH (kJ mol-1) -34.4 DS (J mol-1 K-1) -44.82

ln kc

DS DH R RT

5 6

DG DH T DS

where DG is the Gibbs standard free energy change, DH is the enthalpy change, DS is the entropy change, R gas constant, T absolute temperature, and Keq is equilibrium constant obtained from qe Ce-1. The values of the DH and DS were graphically determined from the plot in the Fig. 10. Thermodynamic parameters of phenol adsorption onto MAC are listed in Table 6. The negative values of DH and DG indicate that the adsorption process is exothermic, favorable and spontaneous. In addition, the negative value of DS means a reduction of the phenol molecule movement when it goes from water to MAC surface [2]. 3.7 Desorption Three different solutions were used for studying the adsorbent regeneration. The rst one was deionized water, the second was ethanol (100 % v/v) and the third was a sodium hydroxide 0.1 M solution. Figure 11 shows the results of the desorption experiments and one can conclude that deionized water was the worst desorbent agent, releasing only 5 % of phenol, ethanol was able to recover about 36 % and sodium hydroxide solution showed a desorption efciency of 55 %. Cooney et al. [30] and Ozkaya and co-workers [31] also studied the phenol desorption by water and ethanol and they found out that ethanol was a better agent than water, agreeing with our results. These authors attributed their conclusions to phenol solubility, which is better in ethanol than in water. In addition, ethanol can probably be a better

solvent to weaken the adsorptive interaction between phenol and adsorbent surface [31]. The results related to sodium hydroxide 0.1 M can be associated to the negative electric charges created on the MAC surface in this pH, as previously discussed. The electrostatic repulsion between the negative surface charge and the phenolate anions could be responsible for the good results obtained using NaOH solution. Furthermore, the reaction between NaOH and phenol, forming a phenolsodium salt, should facilitate the desorption of phenol from the carbon surfaces [32]. It is well known that the regeneration efciency depends on the electron-donating property of adsorbed molecule, where molecules with strong electron-donating ability are difcult to be desorbed [33]. Consequently, the low desorption percentage reached in this study may be attributed to the strong electron-donating ability nature of the phenol molecule [34].

4 Conclusion The results indicated that the pseudo second-order equation provided the better correlation for the adsorption data. The optimum adsorbent dose obtained for phenol adsorption

% phenol desorbed
8.6 8.4 8.2 8.0

50

40

% phenol desorbed

30

ln Kc

7.8 7.6 7.4 7.2 0.00300 0.00305 0.00310 0.00315 0.00320 0.00325 0.00330 0.00335 0.00340

20

10

0
Deionized water EtOH (100%) NaOH (0.1 M)

T (K )
Fig. 10 Plot of ln Kc versus T-1 for phenol adsorption onto MAC (C0 = 1,500 mg dm-3; adsorbent dose = 2 g dm-3; t = 60 min) Fig. 11 Fraction of desorbed phenol from MAC by different desorbing solutions (adsorbent dose = 2 g dm-3; t = 60 min, T = 298 K)

-1

-1

123

626

J Porous Mater (2013) 20:619627 14. G.E.J. Poinern, G. Senanayake, N. Shah, X.N. Thi-Le, G.M. Parkinson, D. Fawcett, Adsorption of the aurocyanide, complex on granular activated carbons derived from macadamia nut shellsA preliminary study. Miner. Eng. 24, 16941702 (2011) 15. C. Nguyen, D. Do, Preparation of carbon molecular sieves from macadamia nut shells. Carbon 33, 17171725 (1995) 16. M. Tam, M. Antal Jr, Preparation of activated carbons from macadamia nut shell and coconut shell by air activation. Ind. Eng. Chem. Res. 38, 42684276 (1999) 17. S. Brunauer, P. Emmett, E. Teller, Adsorption of gases in multimolecular layers. J. Am. Soc. 60, 309319 (1938) 18. B. Lippens, J.D. Boer, Studies on pore systems in catalysts: V. The t method. J. Catal. 323, 319323 (1965) a, Microstructure and quantitative estimation 19. A. Gil, L.M. Gand of the micropore-size distribution of an alumina-pillared clay from nitrogen adsorption at and carbon dioxide adsorption at. Chem. Eng. Sci. 58, 30593075 (2003) 20. H. Boehm, Some aspects of the surface chemistry of carbon blacks and other carbons. Carbon 32, 759769 (1994) astvan, D. Jovanovic , I. Jankovic -C , S. Milonjic , 21. S. Lazarevic kovic , R. Petrovic , Adsorption of Pb2?, Cd2? and Sr2? D. Janac ions onto natural and acid-activated sepiolites. Appl. Clay Sci. 37, 4757 (2007) 22. O. Ylmaz, I. Cem Kantarli, M. Yuksel, M. Saglam, J. Yanik, Conversion of leather wastes to useful products. Resour. Conserv. Recycl. 49, 436448 (2007) 23. K.S.W. Sing, Reporting physisorption data for gas/solid systems with special reference to the determination of surface area and porosity (Recommendations. Pure Appl. Chem. 57(1985), 603619 (1984) 24. L.A. Rodrigues, L.J. Maschio, R.E. da Silva, M.L.C.P. da Silva, Adsorption of Cr(VI) from aqueous solution by hydrous zirconium oxide. J. Hazard. Mater. 173, 630636 (2010) 25. W.H. Cheung, Y.S. Szeto, G. McKay, Intraparticle diffusion processes during acid dye adsorption onto chitosan. Bioresour. Technol. 98, 28972904 (2007) 26. U.K. Garg, M.P. Kaur, V.K. Garg, D. Sud, Removal of hexavalent chromium from aqueous solution by agricultural waste biomass. J. Hazard. Mater. 140, 6068 (2007) 27. H.B. Senturk, D. Ozdes, A. Gundogdu, C. Duran, M. Soylak, Removal of phenol from aqueous solutions by adsorption onto organomodied Tirebolu bentonite: equilibrium, kinetic and thermodynamic study. J. Hazard. Mater. 172, 353362 (2009) 28. L. Li, S. Liu, T. Zhu, Application of activated carbon derived from scrap tires for adsorption of Rhodamine B. J. Environ. Sci. 22, 12731280 (2010) 29. M. Caetano, C. Valderrama, A. Farran, J.L. Cortina, Phenol removal from aqueous solution by adsorption and ion exchange mechanisms onto polymeric resins. J. Colloid Interface Sci. 338, 402409 (2009) 30. D.O. Cooney, A. Nagerl, A.L. Hines, Solvent regeneration of activated carbon. Water Res. 17, 403410 (1983) 31. W. Tanthapanichakoon, P. Ariyadejwanich, P. Japthong, K. Nakagawa, S.R. Mukai, H. Tamon, Adsorption-desorption characteristics of phenol and reactive dyes from aqueous solution on mesoporous activated carbon prepared from waste tires. Water Res. 39, 13471353 (2005) 32. B. Ozkaya, Adsorption and desorption of phenol on activated carbon and a comparison of isotherm models. J. Hazard. Mater. 129, 158163 (2006) 33. H. Tamon, M. Atsushi, M. Okazaki, On irreversible adsorption of electron-donating compounds in aqueous solution. J. Colloid Interface Sci. 390, 384390 (1996) 34. H. Tamon, M. Okazaki, Inuence of surface oxides on ethanol regeneration of spent carbonaceous adsorbents. J. Colloid Interface Sci. 196, 120122 (1997)

onto MAC is 0.1 g and the optimum pH range is 28. The monolayer adsorption capacity of MAC was found to be 341 mg g-1 from Langmuir model equation. The adsorption of phenol dependence on temperature was investigated and the thermodynamic parameters DG, DH and DS were determined. The results show a feasible, spontaneous and exothermic adsorption process. Desorption experiments demonstrate that the MAC achieve the maximum recovery (55 %) using a 0.1 M NaOH solution.
Acknowledgments The authors gratefully acknowledge FAPESP (06/55450-7; 08/55774-2; 2011/20153-0), CNPq and CAPES for nancial support.

References
1. S. Chen, Z.P. Xu, Q. Zhang, G.Q.M. Lu, Z.P. Hao, S. Liu, Studies on adsorption of phenol and 4-nitrophenol on MgAl-mixed oxide derived from MgAl-layered double hydroxide. Sep. Purif. Technol. 67, 194200 (2009) 2. W. Zhang, Q. Du, B. Pan, L. Lv, C. Hong, Z. Jiang et al., Adsorption equilibrium and heat of phenol onto aminated polymeric resins from aqueous solution. Colloids Surf. A 346, 3438 (2009) 3. Z.W. Ming, C.J. Long, P.B. Cai, Z.Q. Xing, B. Zhang, Synergistic adsorption of phenol from aqueous solution onto polymeric adsorbents. J. Hazard. Mater. 128, 123129 (2006) 4. O. Hamdaoui, E. Naffrechoux, Modeling of adsorption isotherms of phenol and chlorophenols onto granular activated carbon. Part I. Two-parameter models and equations allowing determination of thermodynamic parameters. J. Hazard. Mater. 147, 381394 (2007) 5. World Health Organization, Health criteria and other supporting information, in Guidelines for Drinking-Water Quality, vol. 2, 2nd edn. (Geneva, 1996) guez, F. Garc a6. A. Santos, P. Yustos, A. Quintanilla, S. Rodr Ochoa, Route of the catalytic oxidation of phenol in aqueous phase. Appl. Catal. B 39, 97113 (2002) 7. R. Gerdes, D. Whrle, W. Spiller, G. Schneider, G. Schnurpfeil, G. Schulzekloff, Photo-oxidation of phenol and monochlorophenols in oxygen-saturated aqueous solutions by different photosensitizers. J. Photochem. Photobiol. A 111, 6574 (1997) nguez, J. Garc a-Go mez, M.A. 8. P. Canizares, M. Dias, J.A. Dom Rodrigo, Electrochemical oxidation of aqueous phenol wastes on synthetic diamond thinlm electrodes. Ind. Eng. Chem. Res. 41874194 (2002) 9. M.H. El-Naas, S. a Al-Muhtaseb, S. Makhlouf, Biodegradation of phenol by Pseudomonas putida immobilized in polyvinyl alcohol (PVA) gel. J. Hazard. Mater. 164, 720725 (2009) 10. G.-M. Zeng, K. Xu, J.-H. Huang, X. Li, Y.-Y. Fang, Y.-H. Qu, Micellar enhanced ultraltration of phenol in synthetic wastewater using polysulfone spiral membrane. J. Membr. Sci. 310, 149160 (2008) 11. L.A. Rodrigues, M.L.C.P. da Silva, M.O. Alvarez-Mendes, A.R. Coutinho, G.P. Thim, Phenol removal from aqueous solution by activated carbon produced from avocado kernel seeds. Chem. Eng. J. 174, 4957 (2011) 12. B.H. Hameed, A.A. Rahman, Removal of phenol from aqueous solutions by adsorption onto activated carbon prepared from biomass material. J. Hazard. Mater. 160, 576581 (2008) s, M. Sa nchez-Polo, J. Rivera-Utrilla, C. Zaror, Effect of 13. H. Valde ozone treatment on surface properties of activated carbon. Langmuir. 18, 21112116 (2002)

123

J Porous Mater (2013) 20:619627 35. K.P. Singh, A. Malik, S. Sinha, P. Ojha, Liquid-phase adsorption of phenols using activated carbons derived from agricultural waste material. J. Hazard. Mater. 150, 626641 (2008) 36. G.G. Stavropoulos, P. Samaras, G.P. Sakellaropoulos, Effect of activated carbons modication on porosity, surface structure and phenol adsorption. J. Hazard. Mater. 151, 414421 (2008)

627 riz, C. Plada, T. Cordero, ller, J. P 37. N. Tancredi, N. Medero, F. Mo Phenol adsorption onto powdered and granular activated carbon, prepared from Eucalyptus wood. J. Colloid Interface Sci. 279, 357363 (2004)

123

Anda mungkin juga menyukai