You are on page 1of 7

Applied Surface Science 255 (2008) 19751981

Contents lists available at ScienceDirect

Applied Surface Science


journal homepage: www.elsevier.com/locate/apsusc

Thermodynamics of pentachlorophenol adsorption from aqueous solutions by oxidized multi-walled carbon nanotubes
M. Abdel Salam a,*, R.C. Burk b
a b

Chemistry Department, Faculty of Science, King Abdulaziz University, P.O. Box 80200, Jeddah 21589, Saudi Arabia Ottawa-Carleton Chemistry Institute, Department of Chemistry, Carleton University, 1125 Colonel By Drive, Ottawa, ON K1S 5B6, Canada

A R T I C L E I N F O

A B S T R A C T

Article history: Received 12 December 2007 Received in revised form 6 June 2008 Accepted 19 June 2008 Available online 3 July 2008 Keywords: Multi-walled carbon nanotubes Oxidation Pentachlorophenol Adsorption Isotherm Thermodynamics

Chemical modication of MWCNTs via oxidation proved to be a useful tool to improve the suspension stability and solubility of MWCNTs in aqueous solution. Multi-walled carbon nanotubes (MWCNTs) were oxidized using different oxidizing agents and the produced oxidized MWCNTs were characterized using different techniques. IR measurements showed the presence of carboxylic acid function groups especially for the MWCNTs oxidized with nitric acid and hydrogen peroxide. Oxidation of MWCNTs increased their solubility in aqueous solution and hence enhances the contact between the carbon nanotubes and the water and pentachlorophenol molecules. The oxidized MWCNTs were used to study the removal of pentachlorophenol (PCP) from aqueous solutions. The equilibrium adsorption of PCP on the oxidized MWCNTs at various temperatures was studied and the adsorption equilibrium was well described using different adsorption models. The thermodynamic of adsorption was studied at different temperatures and the results showed that the adsorption process was product favored, and becomes more so at higher temperature, since the adsorption is endothermic in general. 2008 Elsevier B.V. All rights reserved.

1. Introduction Recently, releases of different pollutants to the environment have drawn the attention of scientists due to their toxic effects. Most of these pollutants have limited solubility in water and as a result, they exist only at very low concentrations. Many research studies have showed the ability of carbon nanotubes (CNTs) to adsorb different pollutants from various aqueous samples [119]. This ability of CNTs is due to strong interaction between the CNTs surface and the pollutants due to the unique structure of CNTs as a result of the delocalized p-electrons on the hexagonal arrays of carbon atoms in grapheme sheets of CNTs surface. There have been many studies devoted to the adsorption of PCP using different adsorbents including carbon based materials [8,2032], but there was only one study exploring the potentiality of using pristine MWCNTs [8] for the removal of chlorophenols (CPs) from aqueous solution. Up to the authors knowledge, there is no study on the adsorption behavior of PCP, as an example of CPs, on oxidized MWCNTs. Oxidized MWCNTs were used previously for the adsorption of lead, copper and cadmium ions [33,34] and trihalomethanes [15] from aqueous solutions. These research studies concluded that oxidized CNTs

show higher adsorption capacities for the target pollutants compared with other adsorbents such as activated carbon. Thermodynamic studies of adsorption of different analytes by CNTs are scarce. More work is needed to understand the adsorption behavior (thermodynamics) of various analytes on pristine and oxidized MWCNTs. Thermodynamics calculation of the adsorption process at equilibrium is required understand the mechanism of adsorption and to understand the spontaneity, and heat of adsorption by calculating different thermodynamic parameters of the adsorption. In this paper, the adsorption of PCP to different oxidized form of multi-walled carbon nanotubes (MWCNTs) was studied to determine the effect of introducing oxygen containing functional group on the adsorption of PCP on MWCNTs. The study focuses on the equilibrium of the adsorption and the evaluation of different thermodynamic parameters. 2. Experimental details 2.1. Oxidation and characterization of MWCNTs MWCNTs were purchased from Sun Nanotech (China) and were used as received. The MWCNTs (2.0 g) were added to 200 mL solutions of 8 M HNO3, 18% H2O2 and 1 M KMnO4 (acidied), separately. The HNO3 suspension was reuxed at 140 8C for 4 h. The H2O2 and KMnO4 suspensions were heated at 80 8C for 4 h.

* Corresponding author. Tel.: +966 541 88 6660. E-mail address: masalam16@hotmail.com (M. Abdel Salam). 0169-4332/$ see front matter 2008 Elsevier B.V. All rights reserved. doi:10.1016/j.apsusc.2008.06.168

1976

M. Abdel Salam, R.C. Burk / Applied Surface Science 255 (2008) 19751981

MWCNTs oxidized with KMnO4 were further treated with H2SO4 in order to remove MnO2 from the product. The three suspensions were then washed with distilled water until the wash water was pH neutral and a centrifuge (3500 rpm for 30 min) was used for solid liquid separation. The solid products were dried in air at 50 8C. The specic surface area of the pristine and oxidized MWCNTs were measured by nitrogen adsorption at 77 K using the single point BET method with a Micromeretics model 2200 surface area analyzer. Infrared spectra measurements of the oxidized MWCNTs were performed on an ABB Bomem MB Series FTIR spectrometer using KBr pellets. The acidic functional groups on the oxidized MWCNTs were qualitatively and quantitatively determined by titration with base [33]. The acidbase neutralization method involved the dispersing of 6.0 mg of the oxidized MWCNTs in 15.0 mL of 1.0 mM aqueous solutions of NaOH, NaHCO3 or Na2CO3. The mixtures were shaken, sonicated for 4 h, and then centrifuged for 30 min at 3500 rpm. 4.0 mL of the supernatant of each solution was added to 10.0 mL of 1.0 mM HCl to neutralize the unreacted base, and the solutions were back-titrated with 1.0 mM NaOH using a potentiometer to determine the endpoint. The titration of pristine MWCNTs was used as a blank. SEM was done using a JEOL JSM-6400 Digital Scanning Electron Microscope. The XRD prole was acquired using a Philips 1710 X-ray diffraction system (35 kV and 50 mA) with a Cu Ka target. 2.2. Analytical procedure PCP concentration was determined using gas chromatograph with an electron capture detector (GC-ECD), which required an acetylation step for the polar PCP prior the analysis. For acetylation of PCP, 3.0 mL of 5% K2CO3 solution was added to the aqueous solution of PCP and stirred for 1 min, followed by the addition of 2.0 mL of acetic anhydride and stirred vigorously for 2 min to ensure complete acetylation of the PCP. The acetylated PCP was extracted into 4.0 mL n-hexane with 2 min of shaking. The acetylated PCP was analyzed using a Varian 3600 GC-ECD and a 30 m 0.25 mm id 0.25 mm lm thickness DB-5 silica capillary column (Varian). The GC conditions were as follows: helium as the carrier gas at a ow rate of 1 mL/min; nitrogen as the make-up gas at a ow rate of 27 mL/min; injector temperature, 300 8C; column initial temperature, 70 8C; and nal temperature, 280 8C; with heating rate, 15 8C/min; detector temperature, 300 8C. The sample size injected was 1 mL. 2.3. Adsorption equilibrium studies Stock solutions were prepared by dissolving pentachlorophenol (Aldrich 99.0+%) in deionized water followed by serial dilutions. 2.0 mg of MWCNTs were introduced into 14 mL glass vials. 10 mL aqueous PCP solution with different concentration ranging from 0.030 to 1.2 mg/L was added to the MWCNTs. The vials were then shaken for 24 h at different temperatures. Each suspension was then centrifuged at 3000 rpm and the supernatant solution was analyzed for PCP after derivatization using GC-ECD. PCP adsorbed by MWCNTs at equilibrium was calculated as follows: q C 0 C t V m (1)

a temperature controlled water bath. All samples were equilibrated for 24 h. 2.4. Fitting the adsorption isotherms An adsorption isotherm is important in order to be able to predict the equilibrium loading, or to determine the external adsorbent surface loading of PCP during the mass transfer rate experiments of PCP adsorption. The adsorption equilibrium data were tted with various adsorption isotherm models, usually used for adsorption from aqueous solutions. The two-parameter Langmuir [35] and Freundlich [36] models are two of the most commonly used isotherms. The three-parameter Radke and Prausnitz model [37], and the four-parameter FritzSchlunder model [38], were also tted to the data. The Langmuir isotherm can be derived by assuming that only monolayer coverage of the adsorbent surface is possible. qe qm K L C e 1 K LCe (2)

where Ce is the equilibrium PCP concentration (mg/L), qe is the amount adsorbed (mg/g), KL is a Langmuir constant related to the adsorption energy, and qm is a Langmuir constant related to the adsorption capacity, which is related to the amount of adsorbate adsorbed at monolayer coverage. The Freundlich isotherm provides an empirical description of a single-solute adsorption equilibrium: qe K F Ce
1=n

(3)

where KF and 1/n are Freundlich constants related to adsorption capacity and energy distribution of the adsorption sites, respectively. Radke and Prausnitz suggested an empirical correlation to describe equilibrium for the adsorption of organic solutes from dilute aqueous solutions and their equation is: 1 1 1 1= qe KC e kCe p (4)

where K, k and p are Radke and Prausnitz isotherm constants. Finally, for a single solute the FritzSchlunder equation may be expressed as: qe
b a 1 Ce 1 b2 1 a2 Ce

(5)

where a1, a2, b1, and b2 are FritzSchlunder isotherms constants. Statistical software (Polymath version 5.1) was used to t the data to these four adsorption isotherm models. 3. Results and discussion 3.1. Acidic functional group quantitation The concentrations of the different acidic functional groups on the surface of the oxidized MWCNTs were determined qualitatively and quantitatively using an acidbase neutralization method. According to this method, NaOH can neutralize all the acidic functional groups (carboxylic groups, lactones and phenols), Na2CO3 can neutralize some of the acidic function groups (carboxylic groups and lactones), whereas, NaHCO3 can neutralize only strong acidic function groups (carboxylic groups). The results of the titration are shown in Table 1. Each titration was performed three times and the pristine MWCNTs were used as a blank. The concentrations of acidic functional groups were considerable, and were in agreement with other values reported in the literatures

where q is the amount of PCP adsorbed (mg/g), C0 is the initial aqueous concentration of PCP (mg/L), Ct is the concentration of PCP after shaking for a certain period of time (mg/L), V is the solution volume (L) and m is the mass of MWCNTs (g). The effect of temperature on the equilibrium adsorption of PCP was studied at different temperatures; 5 8C, 15 8C, 25 8C, and 35 8C using

M. Abdel Salam, R.C. Burk / Applied Surface Science 255 (2008) 19751981 Table 1 The concentrations of different acidic groups on the surface of MWCNTs after oxidation Oxidant Carboxylic groups (mmol g1) 0.13 0.00 0.47 0.03 0.35 0.14 Lactone groups (mmol g1) 0.23 0.08 0.52 0.07 0.44 0.31 Phenolic groups (mmol g1) 0.21 012 0.27 0.24 0.24 0.33 Total (mmol g1)

1977

KMnO4 HNO3 H2O2

0.56 0.04 1.27 0.20 1.04 0.16

The total acidic surface group concentration (neutralized carboxylic groups, lactones and phenols) was determined by titration with NaOH, carboxylic groups by titration with NaHCO3, whereas lactones concentration was determined by the difference between Na2CO3 titration (carboxylic groups and lactones) and NaHCO3 titration.

[34,39]. The total concentration of the acidic functional groups are higher for the MWCNTs oxidized with HNO3 and H2O2, 1.27 mmol g1 and 1.04 mmol g1, respectively, whereas the MWCNTs oxidized with KMnO4 have the lowest concentration, 0.56 mmol g1. This is likely the reason that the MWCNTs oxidized with HNO3 and H2O2 are well suspended in water whereas those oxidized by KMnO4 were poorly suspended in water. 3.2. Specic surface area measurements The specic surface areas of the pristine and oxidized MWCNTs were measured using the BET method by nitrogen adsorption at 77 K. The specic surface area for the pristine MWCNTs was found to be 148 5 m2/g, which was in good agreement with the other reported values [14,34,40,41]. The specic surface area of the oxidized MWCNTs was slightly lower; 144 6 m2/g and 140 3 m2/g for the MWCNTs oxidized with H2O2 and HNO3, respectively. For the MWCNTs oxidized with KMnO4 the specic surface area of the oxidized tube was smaller; 106 3 m2/g. This may suggest that KMnO4 as an oxidizing agent was very strong and effective enough to remove most of the amorphous carbon as well as damaging some of the MWCNTs. That is mainly due to the introduction of new functional groups at the CNTs surface (OH, COOH, and >C O), which decreased the amount of active sites available for nitrogen adsorption. Meanwhile, the MWCNTs intrinsic morphology of the MWCNTs was improved due to the removal of the impurities; such as amorphous carbon and metal catalysts, which associated with the pristine of MWCNTs. This may explain why the aqueous suspension of the MWCNTs oxidized with KMnO4 was not stable enough in comparison with the other MWCNTs oxidized with H2O2 and HNO3. Also, it is well known that oxidation of carbon nanotubes mainly occurred at the defected sites. Due to the oxidation, the microporosity of the carbon nanotubes enhanced and the adsorption occurs at the outer surface and inside the tubes simultaneously. This should enhance the adsorption greatly. 3.3. Infrared spectroscopy The IR measurements were performed in order to conrm the formation of acidic functional groups upon oxidation. Fig. 1 shows the IR spectra of pristine and oxidized MWCNTs. It is clear from Fig. 1 that the oxidized MWCNTs spectra show absorption peaks at 1730 cm1 corresponding to the stretching vibration of C O from the carboxylic acid groups (COOH) [4247]. This absorption peak was very clear and strong for the MWCNTs oxidized with HNO3, whereas it was much weaker with the MWCNTs oxidized with H2O2 and KMnO4. A broad peak at approximately 3500 cm1 was much stronger in the H2O2 and HNO3 treated MWCNTs. This peak is characteristic of an OH stretch. It is either due to moisture, alcohol or phenol, OH or carboxylic groups in these samples. The aromatic C C stretch is observed at 1645 cm1 and at 1450 cm1

Fig. 1. IR spectra of MWCNTs before and after oxidation.

for all of the spectra with comparable intensities. The peak at 1450 cm1 is more intense for the MWCNTs oxidized with HNO3 and H2O2 than in the pristine and MWCNTs oxidized with KMnO4. The weak, broad peaks at 1130 cm1 are found in all the oxidized MWCNTs it could be due to a CO stretch from either a phenol or lactone. These results and the titration data show clearly that the MWCNTs can be oxidized and that carboxylic, phenolic and lactone groups are likely obtained. 3.4. Scanning electron microscopy SEM imaging was used to study the surface of the MWCNTs before and after oxidation. Representative images are presented in Fig. 2. It is clear from Fig. 2(a) that the pristine MWCNTs are highly tangled tubes with diameters of 100200 nm. The MWCNTs oxidized with HNO3 (Fig. 2(b)) are uniform in size, but have smaller diameters than the untreated CNTs. Those oxidized with H2O2 (Fig. 2(c)) appear to have similar diameters to the untreated CNTs. However, MWCNTs oxidized with KMnO4 (Fig. 2(d)), were mostly destroyed and SEM showed the appearance of large pieces of carbon. This may explain the fact that suspension of KMnO4 treated CNTs are not stable, and that the concentration of acidic groups in these samples is lower than the others. 3.5. XRD analysis Oxidation of MWCNTs mainly leads to the removal of amorphous carbon as well as many other impurities associated with MWCNTs. According to our study, there was not a great change in the degree of graphitization of MWCNTs as a result of oxidation process and this was conrmed with the XRD analysis [48,49]. The XRD analysis showed that the d-spacing; for the (0 0 2) plane, before and after the oxidation did not changed signicantly. The values of the d-spacing (in angstrom) were 3.406, 3.408, 3.06, and 3.398, for the pristine MWCNTs, MWCNTs oxidized with HNO3, H2O2, and KMnO4, respectively. 3.6. Adsorption isotherm of PCP on oxidized MWCNTs The adsorption of PCP to the pristine and oxidized MWCNTs at different temperatures; 5 8C, 15 8C, 25 8C, and 35 8C, was studied and the adsorption curves are presented in Fig. 3. The Figure shows that PCP adsorbs strongly onto the pristine and oxidized MWCNTs at all temperatures. The adsorption of both the oxidized MWCNTs was less than the pristine MWCNTs due to nature of the adsorbent used. From the physical nature of the adsorbent point of view, one

1978

M. Abdel Salam, R.C. Burk / Applied Surface Science 255 (2008) 19751981

Fig. 2. SEM images of (a) pristine MWCNTs, (b) MWCNTs after oxidation with HNO3, (c) MWCNTs oxidized with H2O2, and (d) MWCNTs oxidized with KMnO4.

of the important features for most of carbon adsorbent is their aromaticity, due to the presence of the delocalized p electrons. Pristine MWCNTs, mostly, have a uniform surface with many delocalized p electrons, which increases their adsorption capability in comparison with other carbonueous materials. It was found that the adsorptive properties of carbonueous material are determined mostly by its surface chemical composition. The presence of oxygen and hydrogen within the surface groups crucially affected the adsorptive properties of the adsorbent. Functional groups and delocalized electrons of the MWCNTs structure determine the chemical character of the MWCNTs surface [50]. It was reported that the presence of the carboxyl and hydroxyl groups inhibited the adsorption of phenolic compounds. This effect can be explained by PCP adsorption that is governed by pp dispersion interaction between the MWCNTs and the aromatic ring of the PCP. Oxygen containing functional groups can localize the p electrons, and consequently, remove them from the p-electron system of the MWCNTs [51], lowering the dispersive forces with PCP p-electron. Another explanation is the adsorption of water molecules to the oxygen containing function groups on the surface of the oxidized MWCNTs by the means of hydrogen bonding [52] which is the solvent effect. It can be concluded here that the oxidation of MWCNTs and the introduction of oxygen containing functional groups inhibit the adsorption of PCP due to the localization of the p-electron of the MWCNTs as well as the solvent effect. The adsorption of PCP is slightly lower for the MWCNTs oxidized with HNO3 compared with the MWCNTs

oxidized with H2O2. This may be due to oxidation with HNO3 introduced more oxygen containing function groups; carboxylic, phenolic and lactonic, which decreased the adsorption of PCP to their surfaces. The equilibrium concentrations were in the range of 3.19 mg PCP/g at 35 8C and 1.6 mg PCP/g at 5 8C for the MWCNTs oxidized with HNO3 and of 3.49 mg PCP/g at 35 8C and 2.06 mg PCP/g at 5 8C for the MWCNTs oxidized with H2O2, whereas it was of 3.80 mg PCP/g at 35 8C and 6.28 mg PCP/g at 5 8C for the pristine MWCNTs. The equilibrium adsorption data of the PCP on oxidized MWCNTs at different temperatures were tted to the different isotherms models; Langmuir, and Freundlich models, Radke Prausnitz model and FritzSchlunder model and Table 2 shows the values of the isotherm parameters for the different models. In general, it can be seen from Table 2 that the adsorption models t the data well, as evidenced by the good regression coefcients (R2). At 25 8C and using Freundlich model tting data, the adsorption efciency, Kf, decreased sharply from 46.3, to 4.58 and 4.13, for the pristine and MWCNTs oxidized with H2O2, and HNO3, respectively. At the same temperature; 25 8C, and using Langmuir model tting data, the same effect was found. The adsorption capacity, qm, decreased markedly from 11.6, to 4.35 and 3.86, for the pristine and MWCNTs oxidized with H2O2, and HNO3, respectively. This was expected due to the introduction of the oxygen containing functional groups on the surface of the MWCNTs upon oxidation using either H2O2 or HNO3 as the oxidizing agent. Comparing the adsorption capacity at different temperatures using the same isotherm model reveals the following:

M. Abdel Salam, R.C. Burk / Applied Surface Science 255 (2008) 19751981

1979

- For the MWCNTs oxidized with H2O2 and using Freundlich model tting data, the adsorption capacity, Kf, decreased gradually from 5.44, to 4.58, 3.98, 3.21 by decreasing the temperature from 35 8C, to 25 8C, 15 8C, and 5 8C, respectively. The same effect was found when the Langmuir model was adopted, the adsorption capacity, Kf, decreased gradually from 5.04, to 4.35, 3.69, 3.05 by decreasing the temperature from 35 8C, to 25 8C, 15 8C, and 5 8C, respectively. - For the MWCNTs oxidized with HNO3 and using Freundlich model tting data, the adsorption capacity, Kf, decreased gradually from 5.07, to 4.13, 3.32, 2.45 by decreasing the temperature from 35 8C, to 25 8C, 15 8C, and 5 8C, respectively. The same effect was found when the Langmuir model was adopted, the adsorption capacity, Kf, decreased gradually from 4.81, to 3.86, 3.00, 2.15 by decreasing the temperature from 35 8C, to 25 8C, 15 8C, and 5 8C, respectively. However, the low values of R2, compared with the other two isotherm models, may be due to limitations of Freundlich and Langmuir models. Both RadkePrausnitz and FritzSchlunder models were in a good agreement with the experimental data; good R2 values, as they often used to represent solute adsorption data on heterogeneous surfaces such as MWCNTs and AC. However, the change in the values of their tting parameters was not signicant and it was not clear to withdraw any conclusions. 3.7. Temperature studies and evaluation of the thermodynamic data The effect of temperature on the adsorption equilibrium was studied in order to understand the adsorption mechanism. The adsorption equilibrium depends on the temperature in two different ways. High temperature, generally, increases the rate of diffusion of the adsorbate molecules through the solution to the external and internal surface of the adsorbent, and may change the equilibrium adsorption capacity of the adsorbent for the particular adsorbate. Thermodynamic parameters were calculated from the variation of the thermodynamic equilibrium constant K0 with a change in temperature [53,54]. K0 for the adsorption reaction can be dened as follows:
Fig. 3. Adsorption isotherms for PCP on pristine MWCNTs (a), MWCNTs oxidized with H2O2 (b), and MWCNTs oxidized with HNO3 (c), at different temperatures.

K0

0 as g C s =Cs s 0 ae g e C e =Ce

(6)

where as and ae are the activities of adsorbed solute and the solute in solution at equilibrium, respectively. Cs is the surface
Table 2 Results of PCP adsorption isotherm analysis using pristine and oxidized MWCNTs at different temperatures Temperature adsorbent Freundlich qe K f Ce Kf 35 8C MWCNTs H2O2 HNO3 25 8C MWCNTs H2O2 HNO3 15 8C MWCNTs H2O2 HNO3 5 8C MWCNTs H2O2 HNO3 16.9 5.44 5.07 1/n R2
1=n

Langmuir (1 + KLCe) qm 5.07 5.04 4.81 KL

qe = qmKLCe/ R2

Radke and Prausnitz 1=qe 1=KC e 1=b 1=kCe K k 1/b R2

FritzSchlunder qe a1 Ce =1 a2 Ce

b1

b2

a1
1007 36.8 28.0

a2
631 1.27 1.26

b1
1.44 11.9 10.8

b2
1.77 1.79 2.00

R2

0.57 0.55 0.57

0.885 0.912 0.908

32.9 5.06 4.46

0.943 0.966 0.960

2.56 0.35 0.36

93.5 3.08 10.3

1.29 0.18 0.67

0.968 0.992 0.991

0.983 0.990 0.995

26.2 4.58 4.13

0.56 0.52 0.55

0.814 0.902 0.964

8.18 4.35 3.86

33.4 5.06 4.75

0.893 0.961 0.993

1.64 0.38 0.42

105.3 10.4 4.0

1.92 0.72 0.33

0.956 0.990 0.983

5.36 16.8 21.4

0.01 1.06 1.09

0.03 6.60 6.31

1.99 2.00 1.20

0.973 0.991 0.995

46.3 3.98 3.32

0.77 0.49 0.52

0.844 0.912 0.961

11.6 3.69 3.00

13.1 5.89 5.41

0.879 0.963 0.985

1.42 0.44 0.575

64.3 9.59 3.96

2.06 0.72 0.39

0.914 0.980 0.980

12.3 9.36 5.01

0.01 0.85 0.70

0.25 4.14 3.47

1.99 2.18 2.87

0.957 0.983 0.993

52.0 3.21 2.45

0.68 0.50 0.44

0.943 0.918 0.941

19.9 3.05 2.15

11.1 5.07 7.47

0.952 0.952 0.961

1.17 0.56 0.84

138.5 7.90 3.53

1.56 0.81 0.40

0.967 0.970 0.953

1.38 6.31 3.62

0.002 0.83 0.61

0.71 5.21 3.80

1.98 2.82 2.92

0.987 0.978 0.978

1980

M. Abdel Salam, R.C. Burk / Applied Surface Science 255 (2008) 19751981

concentration of PCP in mg per gram of MWCNTs, Ce, is the aqueous 0 concentration of PCP at equilibrium (mg/L), Cs is the surface 0 concentration of PCP at monolayer coverage of the adsorbent, Ce is molar concentration of PCP at standard conditions; equal 1 M, gs is the activity coefcient of the adsorbed solute and ge is the activity coefcient of the solute in solution. In dilute solution and at low surface coverage, the activity coefcients approach unity, reducing Eq. (6) to the form: K0
0 C s =Cs 0 C e =Ce

(7)

The standard free energy change (DG8) for adsorption was calculated from the relationship:

DG RT ln K 0

(8)
Fig. 4. Temperature dependency plots for the adsorption of PCP on pristine and oxidized MWCNTs.

where R is the universal gas constant and T is the temperature in Kelvin. Using the Vant Hoff equation, the average standard enthalpy change (DH8) can be calculated from the relation between K0 and T: ln K 0 DH const RT


(9)

and standard entropy changes (DS8) can be calculated from:

DG DH T DS

(10)

Using the above relations, the thermodynamic parameters, the standard free energy change, DG8, the standard enthalpy change, DH8, and the change in the entropy, DS8, were calculated from the variation of the thermodynamic equilibrium constant K0 with a change in temperature. Table 3 summarizes the variation of the thermodynamic parameters for the adsorption of PCP on the pristine and oxidized MWCNTs. In general, the adsorption equilibrium constant, K0, increased by decreasing the temperature using pristine MWCNTs, whereas its values was decreases as the temperature decreases for the oxidized MWCNTs. The standard free energy change, DG8, is negative for both oxidized MWCNTs, as would be expected for a product favored reaction, and become more negative as the temperature was increased. Fig. 4 shows the Vant Hoff plot for the adsorption of PCP onto the pristine and the oxidized MWCNTs. It is clear from Fig. 4 and Table 3 that the positive value of DH8 suggests that the PCP and oxidized MWCNTs interaction is endothermic, which was the reason for the decrease in adsorption at lower temperature. The magnitude of DH8
Table 3 Values of thermodynamic parameters for the adsorption of PCP on pristine and oxidized MWCNTs ln Ko 35 8C Pristine MWCNTs MWCNTs-H2O2 MWCNTs-HNO3 25 8C Pristine MWCNTs MWCNTs-H2O2 MWCNTs-HNO3 15 8C Pristine MWCNTs MWCNTs-H2O2 MWCNTs-HNO3 5 8C Pristine MWCNTs MWCNTs-H2O2 MWCNTs-HNO3 13.2 1 12.2 0.3 12.5 0.6 13.6 2 12.0 0.5 12.2 0.1 13.8 1 11.9 0.1 12.0 0.2

suggests a weak type of bonding between the PCP and the MWCNTs. The decrease in DG8 (become less negative) with decreasing the temperature is consistent with the endothermic process. In the case of the adsorption of PCP on the oxidized MWCNTs, the change in entropy was found to be positive. The positive entropy value may suggest an-entropy driven process, especially with the endothermic reaction (+ve DH8) and spontaneous adsorption (ve DG8) due to the hydrophobic bonding between PCP and the oxidized MWCNTs surface [53]. The positive entropy, generally, comes from the loss of structured water surrounding the PCP molecules. 4. Conclusions Although the adsorption of PCP onto the oxidized MWCNTs was less than the pristine MWCNTs, but they can consider to be a good adsorbents for the removal of PCP from aqueous solutions. The equilibrium adsorption of PCP on the oxidized MWCNTs at various temperatures was studied and the adsorption equilibrium is well described using different adsorption models. The thermodynamic parameters showed that the adsorption process is product favored, and becomes more so at higher temperature, since the adsorption is endothermic. The magnitude of the enthalpy suggests a weak type of bonding between the PCP and the MWCNTs. The entropy values were positive, implying that the PCP is more ordered on the aqueous phase than at the oxidized MWCNTs surface mainly due to loss of structured water surrounding the PCP molecules. References
[1] G.Z. Fang, J.X. He, S. Wang, J. Chromatogr. A 1127 (2006) 1217. [2] Q. Zhou, J. Xiao, W. Wang, J. Chromatogr. A 1125 (2006) 152158. [3] C. Basheer, A.A. Alnedhary, B.S.M. Rao, S. Valliyaveettil, H.K. Lee, Anal. Chem. 78 (2006) 28532858. [4] Q. Zhou, W. Wang, J. Xiao, Anal. Chem. Acta 559 (2006) 200206. [5] Q. Zhou, J. Xiao, W. Wang, G. Liu, Q. Shi, J. Wang, Talanta 68 (2006) 13091315. [6] Q. Zhou, W. Wang, J. Xiao, J. Wang, G. Liu, Q. Shi, G. Guo, Microchim. Acta 152 (2006) 215224. [7] P. Liang, Q. Ding, F. Song, J. Sep. Sci. 28 (2005) 23392343. [8] Y. Cai, S. Mou, Y. Lu, J. Chromatogr. A 1081 (2005) 245247. [9] W.F. Lu, M.Y. Ding, R. Zheng, J. Chromatogr. Sci. 43 (2005) 383387. [10] A. Ambrosi, R. Antiochia, L. Campanella, R. Dragone, I. Lavagnini, J. Hazard. Mater. 122 (2005) 219225. [11] C. Pan, S. Xu, H. Zou, Z. Guo, Y. Zhang, B. Guo, J. Am. Soc. Mass. Spectrom. 16 (2005) 263270. [12] G. Liu, J. Wang, Y. Zhu, X. Zhang, Anal. Lett. 37 (2004) 30853104. [13] P. Liang, Y. Liu, L. Guo, J. Zeng, H. Lu, J. Anal. At. Spectrom. 19 (2004) 14891492. [14] Y. Cai, G. Jiang, J. Liu, Q. Zhou, Anal. Chem. 75 (2003) 25172521. [15] C. Lu, Y.L. Chung, K.F. Chang, Water Res. 39 (2005) 11831189. [16] A. Staej, K. Pyrzynska, Sep. Purif. Technol. 58 (2007) 4952. [17] C. Ye, Q.M. Gong, F.P. Lu, J. Liang, Sep. Purif. Technol. 58 (2007) 26.

DG8 (kJ/mol)
33.9 0.4 31.2 0.1 32.0 0.2 33.7 2 29.8 0.1 30.3 0.1 33.1 0.3 28.5 0.1 28.8 0.2

DH8 (kJ/mol)
23.7 6 +10.8 1 +16.4 0.9 23.7 6 +10.8 1 +16.4 0.9 23.7 6 +10.8 1 +16.4 0.9

DS8 (J/K mol)


33.1 2 136 2 157 3 33.5 2 137 1 157 1 32.6 2 137 3 156 0.2

14.3 1 11.7 0.2 11.8 0.1

32.9 0.2 27.1 0.1 27.3 0.3

23.7 6 +10.8 1 +16.4 0.9

33.4 10 136 6 157 2

M. Abdel Salam, R.C. Burk / Applied Surface Science 255 (2008) 19751981 [18] X. Xie, L. Gao, J. Sun, Colloids Surf. A 308 (2007) 5459. [19] M. Tuzen, M. Soylak, J. Hazard. Mater. 147 (2007) 219225. [20] B.N. Estevinho, N. Ratola, A. Alves, L. Santos, J. Hazard. Mater. 137 (2006) 1157 1181. [21] M. Cea, J.C. Seaman, A.A. Jara, B. Fuentes, M.L. Mora, M.C. Diez, Chemosphere 67 (2007) 13541360. [22] S. Srivastava, A.H. Ahmed, I.S. Thakur, Bioresour. Technol. 98 (2007) 11281132. [23] P.E. Diaz_ores, R. Leyva-Ramos, R.M. Guerrero-Coronado, J. Mendoza-Barron, Ind. Eng. Chem. Res. 45 (2006) 330336. [24] C. Cui, X. Quan, S. Chen, H. Zhao, Sep. Purif. Technol. 47 (2005) 7379. [25] X. Liu, X. Quan, L. Bo, S. Chen, Y. Zhao, Carbon 42 (2004) 415. [26] W. Jianlong, Q. Yi, N. Horan, E. Stentiford, Bioresour. Technol. 75 (2000) 157161. [27] S. Zheng, Z. Yang, D.H. Jo, Y.H. Park, Water Res. 38 (2004) 23152322. [28] J.P. DiVincenzo, D.L. Sparks, Arch. Environ. Contam. Toxicol. 40 (2001) 445450. [29] X. Liu, X. Quan, L. Bo, S. Chen, Y. Zhao, Carbon 42 (2004) 415422. [30] D.S. Shen, X.W. Liu, Y.H. He, J. Hazard. Mater. 125 (2005) 231236. [31] I. Bras, L.T. Lemos, A. Alves, M.F.R. Pereira, Manage. Environ. Qual. 15 (2004) 491501. [32] D.M. Han, G.Z. Fang, Z.P. Yan, J. Chromatogr. A 1100 (2005) 131136. [33] Y. Li, J. Ding, Z. Luan, Z. Di, Y. Zhu, C. Xu, D. Wu, B. Wei, Carbon 41 (2003) 2787 2792. [34] Y. Li, S. Wang, Z. Luan, J. Ding, C. Xu, D. Wu, Carbon 41 (2003) 10571062. [35] J. Langmuir, J. Am. Chem. Soc. 40 (1918) 13611403. [36] H.Z. Freundlich, J. Phys. Chem. 57 (1907) 385.

1981

[37] C.J. Radke, J.M. Prausnitz, Ind. Eng. Chem. Fundam. 11 (1972) 445451. [38] W. Fritz, E.U. Schlunder, Chem. Eng. Sci. 29 (1974) 12791282. [39] B.C. Satishkumar, A. Govindaraj, J. Mofokeng, G.N. Subbanna, C.N.R. Rao, J. Phys. B: At., Mol. Opt. Phys. 29 (1996) 49254934. [40] Y. Cai, G. Jiang, J. Liu, Q. Zhou, Anal. Chim. Acta 494 (2003) 149156. [41] Q. Li, D. Yuan, J. Chromatogr. A 1003 (2003) 203209. [42] W. Sun, Z. Huang, L. Zhang, J. Zhu, Carbon 41 (2003) 16851687. [43] J. Chen, M.A. Hamon, H. Hu, Y. Chen, A.M. Rao, P.C. Eklund, R.C. Haddon, Science 282 (1998) 9598. [44] M.S.P. Shaffer, X. Fan, A.H. Windle, Carbon 36 (1998) 16031612. [45] M.A. Hamon, H. Hu, R.C. Haddon, Chem. Phys. Lett. 347 (2001) 812. [46] J. Chen, R.E. Smalley, R.C. Haddon, J. Phys. Chem. B 105 (2001) 25252528. [47] A. Koshio, M. Yudasaka, M. Zhang, S. Iijima, Nano Lett. 1 (2001) 361363. [48] J. Yan, L. Yi, W.R. Zhong, J. Serb. Chem. Soc. 70 (2005) 277282. [49] S.L. Zhang, X.H. Hu, H.D. Li, Z.J. Shi, K.T. Yue, J. Zi, Z.N. Gu, X.H. Wu, Z.L. Lian, Y. Zhan, F.M. Huang, L.X. Zhou, Y.G. Zhang, S. Iijima, Phys. Rev. B 66 (2002) 035413. [50] C.A. Leon, L.R. Radovic, in: P.A. Thrower (Ed.), Chemistry and Physics of Carbon, vol. 24, Marcel Dekker, New York, 1994, p. 214. [51] R.W. Couglin, F.S. Ezra, Environ. Sci. Technol. 2 (1968) 291297. cienly, Z. Hubicki, M. Barczak, Chemosphere 58 (2005) [52] A. Dabrowski, P. Podkos 10491070. [53] X. Peng, Y. Li, Z. Luan, Z. Di, H. Wang, B. Tian, Z. Jia, Chem. Phys. Lett. 376 (2003) 154158. [54] R. Niwas, U. Gupta, A.A. Khan, Colloids Surf. A 164 (2000) 115119.

You might also like