1-s2.0-S0008622312009505-main

12
The effects of the surface oxidation of activated carbon, the solution pH and the temperature on adsorption of ibuprofen Hanen Guedidi a,b , Laurence Reinert a , Jean-Marc Le ´ve ˆque a , Yasushi Soneda c , Nizar Bellakhal b , Laurent Duclaux a, * a Laboratoire de Chimie Mole ´culaire et Environnement, Universite ´ de Savoie, Le Bourget du Lac 73376, France b Laboratoire de Chimie Analytique et Electrochimie, Universite ´ de Tunis, 2092 Manar II, Tunisia c National Institute of Advanced Industrial Science and Technology, Energy Technology Research Institute, 16-1 Onogawa, Tsukuba, Ibaraki 305-8569, Japan ARTICLE INFO Article history: Received 3 August 2012 Accepted 28 November 2012 Available online 7 December 2012 ABSTRACT A commercial microporous–mesoporous granular activated carbon was modified by oxida- tion with either H 2 O 2 in the presence or absence of ultrasonic irradiation, or NaOCl or by a thermal treatment under nitrogen flow. Raw and modified materials were characterized by N 2 adsorption–desorption measurements at 77 K, Boehm titrations, pH measurements and X-ray photoelectron spectroscopy. Ibuprofen adsorption kinetic and isotherm studies were carried out at pH 3 and 7 on raw and modified materials. The thermodynamic parameters of adsorption were calculated from the isotherms obtained at 298, 313 and 328 K. The pore size distribution of carbon loaded with ibuprofen brought out that adsorption occurred preferentially into the ultramicropores. The adsorption of ibuprofen on pristine activated carbon was found endothermic, spontaneous (DG° = 1.1 kJ mol 1 ), and promoted at acidic pH through dispersive interactions. All explored oxidative treatments led mainly to the for- mation of carbonyl groups and in a less extent to lactonic and carboxylic groups. This then helped to enhance the adsorption uptake while decreasing adsorption Gibbs energy (nota- bly 7.3 kJ mol 1 after sonication in H 2 O 2 ). The decrease of the adsorption capacity after bleaching was attributed to the presence of phenolic groups. Ó 2012 Elsevier Ltd. All rights reserved. 1. Introduction Pharmaceutically active compounds (PhACs) are frequently detected in treated wastewater arising from sewage treat- ment plants, from which they are not completely removed [1,2]. Moreover, recent studies have reported the presence of traces or ultra traces (ng L 1 ) of PhACs in drinking water [3]. Many of these compounds are also suspected to have adverse impact on humans and wild life [4]. According to literature, ibuprofen (IBP), a non steroidal anti-inflammatory, is one of the most common drugs found in water [5]. Indeed, the latter has been detected in the effluents from several sewage-treatment plants at concentra- tions up to 24.6 lgL 1 [6], possibly presenting a potential hazard for human health [7]. The removal of IBP was already reported using different techniques such as ultrasonic irradiation [8], combined ad- vanced oxidation processes (AOPs) such as sonophotocatalyt- ic degradation in the presence of Fe 3+ and TiO 2 [9], electrochemical degradation [10], recognizing extraction using molecularly imprinted polymers (MIPs) [11,12], adsorp- tion on mesoporous silica [13] or activated carbons [14,15]. Activated carbon has several applications depending on its porous structure and adsorption capacity [16]. In the 0008-6223/$ - see front matter Ó 2012 Elsevier Ltd. All rights reserved. http://dx.doi.org/10.1016/j.carbon.2012.11.059 * Corresponding author: Fax: +33 479758805. E-mail address: [email protected] (L. Duclaux). CARBON 54 (2013) 432 443 Available at www.sciencedirect.com journal homepage: www.elsevier.com/locate/carbon

Transcript of 1-s2.0-S0008622312009505-main

  • ar

    Ma

    a Laboratoire de Chimie Moleculaire et Environnement, Universite de Savoie, Le Bourget du Lac 73376, Franceb Laboratoire de Chimie Analytique et Electroc National Institute of Advanced Industrial Sc

    Tsukuba, Ibaraki 305-8569, Japan

    A R T I C L E I N F O

    Article history:

    Received 3 August 2012

    Accepted 28 November 2012

    Available online 7 December 2012

    e adsorption uptake while decreasing adsorption Gibbs energy (nota-

    Pharmaceutically active compounds (PhACs) are frequently

    Many of these compounds are also suspected to have adverse

    impact on humans and wild life [4].

    According to literature, ibuprofen (IBP), a non steroidal

    anti-inflammatory, is one of the most common drugs found

    in water [5]. Indeed, the latter has been detected in the

    hazard for human health [7].

    electrochemical degradation [10], recognizing extraction

    using molecularly imprinted polymers (MIPs) [11,12], adsorp-

    tion on mesoporous silica [13] or activated carbons [14,15].

    Activated carbon has several applications depending on

    its porous structure and adsorption capacity [16]. In the

    C A R B O N 5 4 ( 2 0 1 3 ) 4 3 2 4 4 3

    Avai lab le at www.sc iencedi rect .com

    els* Corresponding author: Fax: +33 479758805.detected in treated wastewater arising from sewage treat-

    ment plants, from which they are not completely removed

    [1,2]. Moreover, recent studies have reported the presence of

    traces or ultra traces (ng L1) of PhACs in drinking water [3].

    The removal of IBP was already reported using different

    techniques such as ultrasonic irradiation [8], combined ad-

    vanced oxidation processes (AOPs) such as sonophotocatalyt-

    ic degradation in the presence of Fe3+ and TiO2 [9],bly 7.3 kJ mol1 after sonication in H2O2). The decrease of the adsorption capacity afterbleaching was attributed to the presence of phenolic groups.

    2012 Elsevier Ltd. All rights reserved.

    1. Introduction effluents from several sewage-treatment plants at concentra-tions up to 24.6 lg L1 [6], possibly presenting a potentialmation of carbonyl gr

    helped to enhance th0008-6223/$ - see front matter 2012 Elsevihttp://dx.doi.org/10.1016/j.carbon.2012.11.059

    E-mail address: Laurent.duclaux@univ-sachimie, Universite de Tunis, 2092 Manar II, Tunisia

    ience and Technology, Energy Technology Research Institute, 16-1 Onogawa,

    A B S T R A C T

    A commercial microporousmesoporous granular activated carbon was modified by oxida-

    tion with either H2O2 in the presence or absence of ultrasonic irradiation, or NaOCl or by a

    thermal treatment under nitrogen flow. Raw and modified materials were characterized by

    N2 adsorptiondesorption measurements at 77 K, Boehm titrations, pH measurements and

    X-ray photoelectron spectroscopy. Ibuprofen adsorption kinetic and isotherm studies were

    carried out at pH 3 and 7 on raw and modified materials. The thermodynamic parameters

    of adsorption were calculated from the isotherms obtained at 298, 313 and 328 K. The pore

    size distribution of carbon loaded with ibuprofen brought out that adsorption occurred

    preferentially into the ultramicropores. The adsorption of ibuprofen on pristine activated

    carbon was found endothermic, spontaneous (DG = 1.1 kJ mol1), and promoted at acidicpH through dispersive interactions. All explored oxidative treatments led mainly to the for-

    oups and in a less extent to lactonic and carboxylic groups. This thenNizar Bellakhal b, Laurent Duclaux a,*The effects of the surface oxidsolution pH and the temperatu

    Hanen Guedidi a,b, Laurence Reinert a, Jean-

    journal homepage: www.er Ltd. All rights reservedvoie.fr (L. Duclaux).tion of activated carbon, thee on adsorption of ibuprofen

    rc Leveque a, Yasushi Soneda c,

    evier .com/ locate /carbon.

  • literature, numerous studies have shown that adsorption by

    activated carbon is effective for the removal of various pollu-

    tants in water and waste water such as heavy metals [1719],

    (2)

    C A R B O N 5 4 ( 2 0 1dyes [2025], pesticide [26,27], fluoride [28] and PhACs [29].

    Mestre et al. [15] previously reported that IBP adsorption

    on activated carbons was enhanced by the presence of supe-

    rmicropores and acidic pH values but was not by the variation

    of temperature in the range of 298313 K. In addition, they

    emphasized the favorable effect of basic carbon surface [14]

    on the IBP adsorption process and hypothesized that the pres-

    ence of mesopores was crucial to ensure the accessibility to

    the inner microporosity which had to be large enough

    (0.7 nm) to accommodate IBP moieties. More recently, the

    adsorption of IBP on a mesoporous activated carbon prepared

    from Artemisia Vulgaris leaves [30] was reported to be endo-

    thermic and the maximum uptake was observed at pH 2 sug-

    gesting a physisorption mechanism involving the carboxylic

    function of the adsorbate molecule.

    The modification of the surface chemistry of activated car-

    bons by treatments in liquid media with oxidative agents

    such as HNO3, (NH4)2S2O8, and H2O2 is known to generate

    an increase in the number of acidic surface sites. However,

    this also leads to a drastic decrease in the total specific sur-

    face area and micropore volume, explained by the presence

    of oxygen groups at the entrance and the walls of the pores

    [31]. For all cited treatments, except the one with H2O2, a

    destruction of the basic sites was observed [31].

    Functional oxygen groups play an important role in the

    adsorption of aromatic molecules. Haydar et al. [32] reported

    that oxidation by HNO3 greatly increased the number of car-

    boxylic groups, withdrawing electrons from the graphene lay-

    ers and thus reducing the p-nitrophenol uptake. Mattson

    et al. [33] suggested that an interaction between carbonyl sur-

    face groups (as electron donors) and aromatic rings of phenol

    (acceptors) lead to the formation of donoracceptor com-

    plexes. Other authors explained the adsorption of phenol

    through the formation of hydrogen bonds (of the OH function

    of phenol) with carboxylic groups [34].

    This present work aims at investigating the IBP adsorption

    profile on chemically modified surfaces-bearing activated car-

    bons. Thus, a commercial microporousmesoporous acti-

    vated carbon was treated in several different ways in order

    to obtain adsorbents with different surface properties. The

    adsorption properties of the modified and pristine activated

    carbons were studied as a function of concentration, pH,

    and temperature to investigate the impact of the surface

    chemistry on the IBP adsorption.

    2. Experimental

    2.1. Ibuprofen

    2-[4-(2-Methylpropyl) phenyl] propanoic acid (Fig. 1), also

    named ibuprofen, was purchased from the nationalFig. 1 Molecular structure of Ibuprofen.an ultrasonic probe (20 kHz, 43 lm amplitude, 19 mm

    diameter, 350 W electrical power, Sonics and Materials

    Ultrasonic Processor - VC505) dipped into the carbon

    suspension, maintained at 298 K in a cylindrical dou-

    bled-jacket cooled reactor.

    All the oxidized samples were then recovered by

    filtration at 0.45 lm on membrane filters (Durapore-

    Millipore), rinsed with distilled water until reaching a

    neutral pH (1 L per g) and further dried at 353 K for 24 h.

    The bleached AC was named AC[NaOCl]. The samples

    oxidized with H2O2, and sonicated in H2O2 were

    referred to AC[H2O2] and AC[H2O2]US, respectively.

    (3) Sample named AC700N2 was prepared by heating at

    973 K (10 K min1, 1 h at 973 K), under a nitrogen flowand kept under this inert atmosphere until cooling to

    room temperature.

    2.3. N2 adsorptiondesorption at 77 K and CO2 adsorptionat 273 K

    N2 adsorptiondesorption isotherms of the activated carbons

    were measured using an automatic sorptometer (ASAP 2020,

    Micromeritics) at 77 K. The CO2 adsorption isotherms were

    measured at 273 K on two samples (AC and AC loaded with

    56.7 mg g1 IBP at pH 7). Prior to measurements, samples (ex-cept IBP loaded AC) were degassed for 12 h at 523 K under vac-

    uum. IBP loaded AC was outgassed for 4 days at 323 K, below

    ibuprofen melting point (348350 K) [39].

    The specific surface areas of the activated carbons were

    calculated using the BrunauerEmmettTeller (BET) equation,

    assuming the area of the nitrogen molecule to be 0.162 nm2.

    As negative unrealistic C factors were obtained by applyingthe B

    0.3, thFor the bleaching treatment, AC (1 g) was dispersed in

    200 mL of a 0.13 mol L1 NaOCl solution and stirred atroom temperature during 24 h.

    AC samples (2.5 g) were also oxidized by impregnation

    at room temperature for 5 h with 50 mL of 35% H2O2,

    or (3) by ultrasound irradiation for 5 h in H2O2 (35%,

    50 mL). The sonication treatment was performed usingthe filtrate [38]. After drying at 383 K for 24 h, this sample, re-

    ferred to AC, was further chemically modified by four differ-

    ent treatments as described here below:

    (1)laboratory of the control of drugs (LNCM, Tunisia) and used

    as received (99.1% purity). This molecule is very slightly

    soluble in water (21 mg L1) [35] but soluble in a majority oforganic solvents [36,37]. The dimensions were estimated

    using Chemsketch 3D Viewer assuming a parallelepiped

    shape: 1.13 0.50 0.31 nm3.

    2.2. Raw and modified activated carbons

    A commercial granulated activated carbon (SigmaAldrich,

    Ref. 292591, 414 mesh) was milled and washed with a hydro-

    chloric acid solution (0.1 mol L1) for 24 h under stirring to re-move metal salt impurities. The resulting suspension was

    then filtered off and the recovered carbon was rinsed with

    distilled water (200 mL/g) until reaching a constant pH of

    3 ) 4 3 2 4 4 3 433ET model in the relative pressure range from 0.05 to

    e BET specific surface areas were then preferentially

  • 0 1computed in the relative pressure range from 0.01 to 0.05, as

    for microporous materials.

    The total pore volume was estimated as the liquid vol-

    ume of N2 adsorbed at a relative pressure of 0.995. In addi-

    tion, the micropore volume was determined by using the

    Dubinin-Radushkevich equation for N2 adsorption (P/P0from 0.01 to 0.05) and for CO2 adsorption [40]. The meso-

    pore volume was calculated by difference of the total pore

    volume and the micropore volume. The pore size distribu-

    tions (PSD) of the pristine AC and IBP loaded AC (pH = 7

    and 56.7 mg g1 of IBP) were determined by using NLDFT(non local density functional theory) models applied on

    both the adsorption isotherms of N2 at 77 K and CO2 at

    273 K. For that, infinite slit pores were assumed for CO2adsorption (pores diameter from 0.35 to 1.1 nm), while finite

    slit pores having a diameter-to-width aspect ratio of 12

    were used for N2 adsorption simulations (pores diameter

    from 3.5 to 250 A) [41]. N2 adsorption data at P/P0 < 0.01

    were obtained using incremental fixed doses of 10 cm3 g1(STP), setting the equilibration interval at 300 s [42]. CO2adsorption data were obtained at P/P0 ranging from

    4 104 and 3.5 102, using 45 s equilibration interval.

    2.4. pHs of carbons suspended in water

    The pH of each activated carbon sample (0.5 g) was measured

    in a distilled water suspension (12.5 mL) after heating at 90 Cand then cooling to room temperature [43].

    The pHPZC (pH value at which the surface has a zero net

    charge) was determined by the so-called pH drift method

    [34]. The pH of deoxygenized NaCl aqueous solutions (50 mL

    at 0.01 mol L1) were adjusted to successive initial values be-tween 2 and 12 by adding either 0.1 mol L1 HCl or NaOH. Sus-pensions were prepared by adding activated carbon (0.15 g) to

    each solution. After stirring for 48 h under N2, the final pH

    was measured and plotted versus the initial pH. The pHPZCwas determined at the value for which pHfinal = pHinitial.

    2.5. Boehm titrations

    Boehm titrations quantify the basic and oxygenated acid sur-

    face groups on activated carbons [44]. Surface functional

    groups such as carboxyl (RCOOH), lactone (ROCO), phenol

    (ArOH), carbonyl or quinone (RR 0C@O) and basic groups weredetermined using different reactants, assuming that: NaOC2-H5 reacted with all groups; NaOH did not react with RR 0C@Ogroups; Na2CO3 did not react with RR 0C@O nor ROH groupsand that NaHCO3 only reacted with RCOOH groups. The pro-

    cedure was as follows: about 0.15 g of each carbon sample

    was mixed in a closed erlenmeyer with 50 mL of a 0.1 mol L1

    aqueous reactant solution (NaOH, Na2CO3 or NaHCO3). In the

    case of NaOC2H5, only 0.1 g of carbon sample was added to

    50 mL of 0.01 mol L1 solutions. The mixtures were stirredfor 24 h at constant speed (650 rpm) and room temperature

    and then filtered off. Back-titrations of the filtrate (10 mL)

    were then achieved with standard HCl (0.01 mol L1) to deter-mine the oxygenated groups contents. Basic groups contents

    were also determined by back titration of the filtrate with

    434 C A R B O N 5 4 ( 2NaOH (0.01 mol L1) after stirring of the activated carbon(0.15 g) in HCl (50 mL, 0.01 mol L1) for 24 h.2.6. X-ray photoelectron spectroscopy

    X-ray photoelectron spectroscopy (XPS) measurements were

    performed using an ESCALAB 250 spectrometer (Thermo Fish-

    er Scientific) at monochromated Al-Ka anode X-ray radiation,

    on a 150 800 lm2 analysis region, under 2 109 mbar vac-uum. The high-resolution scans (0.1 eV) were obtained over

    the 280.1299.9 eV (C1s) and 523.1539.9 eV (O1s) energy

    ranges with a pass energy of 20 eV. After baseline subtraction,

    the curve fitting was performed assuming a mixed Gaussian

    Lorentzian peak shape (the ratio of Gaussian to Lorentzian

    form was equal to 0.3). The carbon 1s electron binding energy

    corresponding to graphitic carbon was referenced at 284.6 eV

    for the calibration [45].

    2.7. Adsorption experiments: kinetics and isotherms

    The IBP solutions were prepared from UHQ water (Ultra High

    Quality, 18.2 MX purity) containing 10 vol.% of methanol

    (99.9%, SigmaAldrich) in order to increase the solubility.

    The kinetics were studied at room temperature at pH 3 and

    7. Suspensions of 10 mg of the differently treated ACs in

    30 mL of 100 ppm IBP solution were put to stirring at

    650 rpm and then filtered at different intervals between

    15 min and 7 days. The IBP concentrations were subsequently

    determined by UVvisible spectrometry (Varian, Cary50). The

    maximum absorbance was observed at 220 nm at pH = 3, and

    222 nm at pH = 7.

    The isotherms of adsorption of IBP on the various acti-

    vated carbons were studied at three constant temperatures:

    298, 313 and 328 K, at pH 3 and 7. Activated carbons (10 mg)

    were introduced in IBP solutions (30 mL) of varying concentra-

    tions (5100 mg L1) and stirred for 5 days.The equilibrium IBP uptake Qads (mg g

    1) was calculatedfrom equation:

    Qads Ce Ci V

    m

    where V is the solution volume (mL), Ci is the initial IBP con-

    centration (mg L1), Ce is the equilibrium IBP concentration(mg L1) and m is the weight of the dry activated carbon (mg).

    3. Results and discussion

    3.1. Characterization of activated carbons

    3.1.1. Surface chemistryThe Boehm titrations of AC revealed the overall presence of

    more basic than acidic functional groups (Table 1), confirmed

    by the pH value (9.05). When AC was oxidized respectively by

    NaOCl or by H2O2 with and without ultrasonic irradiation, the

    pH values decreased due to the increase of the oxygenated

    surface groups concentrations from 0.37 to 2.1, 3.9 and3.5 meq g1. The value of pHPZC slightly decreased for thebleached activated carbon (from 7.76 to 7.60) but increased

    for both AC[H2O2] and AC[H2O2]US (Table 1). This slight in-

    crease in pHPZC could be attributed to the higher amount of

    carbonyl groups (2.10 and 2.25 meq g1 for AC[H2O2] and

    3 ) 4 3 2 4 4 3AC[H2O2]US, respectively) and lactonic groups (0.77 and

    1.32 meq g1 for AC[H2O2] and AC[H2O2]US, respectively)

  • deconvolution and brought out the highest content in ad-

    sorbed O detected by XPS (Table 2).

    Whatever the oxidized activated carbons, their contents in

    graphitic carbon (at 284.6 eV) were lower than in pristine AC,

    and the percentages of oxygen bonded to carbon increased

    ti

    AC

    Fig. 2 O1s XPS spectra of raw and modified activated

    carbons.

    C A R B O N 5 4 ( 2 0 1 3 ) 4 3 2 4 4 3 435which possess high pKa values (pKa-carbonyl 1620, pKa-lactonic 8.2).

    The oxidation by ultrasonic irradiation in H2O2 (AC[H2O2]-

    US sample) compared to the silent analog (AC[H2O2] sample)

    yielded to a slight decrease of the carbonyl groups. By con-

    trast, additional phenolic groups (0.27 meq g1) were foundunder ultrasound irradiation, whose origin can be attributed

    to the reaction between atoms of carbon and free OH radicalsgenerated by the sonolytic degradation of H2O2 [46]. Ultra-

    sonic and silent treatments with hydrogen peroxide gener-

    ated the highest contents in carbonyl and lactonic groups

    (Table 1) while the bleaching treatment yielded to the highest

    amount of phenolic groups (0.5 meq g1). The carbonyl andlactonic contents were lower in bleached samples

    (1 meq g1) than in samples treated with H2O2 (>2 meq g1).The ultrasonic treatment also induced a significant particle

    size reduction due to the cavitation phenomenon [47]. Laser

    granulometry characterization (not shown) confirmed that

    AC[H2O2]US was mainly formed of submicrometric particles

    (median size d50 = 1 lm) while oxidation in H2O2 without

    ultrasound preserved the micrometric particles

    (d50 = 100 lm), such as for the raw AC (d50 = 700 lm).

    Boehm titrations (Table 1) revealed quite different surface

    groups contents for a fresh sample characterized just after

    annealing under N2 atmosphere, in comparison to a one-year

    air atmosphere aged sample. The fresh AC700N2 sample

    was found to be the most basic sample, which is in agreement

    Table 1 Surface groups (in meq g1) obtained from Boehmcarbons.

    Sample AC AC[NaOCl]

    Carboxylic groups 0 0.15Lactonic groups 0.03 0.35Phenolic groups 0.10 0.50Carbonyl groups 0.24 1.08Total oxygenated groups 0.37 2.08Total basic groups 0.99 0.99pH 9.05 7.18pHPZC 7.76 7.60a Fresh AC700N2.b Aged AC700N2.with its measured pH and pHPZC values. Although the anneal-

    ing under N2 atmosphere removed many oxygen surface

    groups, the sample underwent again oxidation with time

    when exposed to air atmosphere (Table 1), leading to a similar

    carbonyl content (1 meq g1) than the bleached sample. Infact, a N2 thermal treatment is known to be only temporarily

    efficient for removing the oxygen containing surface groups,

    as it also generates very reactive surface sites that further re-

    act with oxygen from air, enabling the formation of new oxy-

    genated surface groups [48].

    Surface chemical groups of the different activated carbons

    were also characterized by XPS. Fig. 2 presents the O1s spec-

    trum of the raw and treated activated carbons. Deconvolution

    of the C1s and O1s spectra gave respectively five and four

    peaks (Table 2).

    XPS analysis of the aged AC700N2 sample confirmed a

    slight oxidation compared to raw AC, through C1s spectrumtrations, pH and pHPZC values of raw and modified activated

    [H2O2] AC[H2O2]US AC700N2a AC700N2

    b

    0.32 0.37 0 01.32 0.77 0 0.230 0.27 0 0.032.25 2.10 0.10 1.023.89 3.51 0.10 1.280.65 0.62 1.14 0.778.38 8.03 10.14 8.747.80 8.03 9.08 8.44after oxidation (Table 2), in agreement with the Boehm

    titrations.

    The C1s spectra (Table 2) displayed higher carbonyl (or qui-

    none) and carboxylic (or ester or lactone) groups contents in

    AC[H2O2] (8.05% of C@O group; 5.82% of OC@O groups) andAC[H2O2]US (9.82% of C@O group; 7.69% of OC@O groups)compared to the bleached AC[NaOCl] sample (4.45% of O

    C@O, and 7.02% of C@O), for which the levels were quite sim-ilar to the AC sample. By contrast, Cagnon et al. [49] reported

    an increase in carboxylic and carbonyl groups XPS contents

    after bleaching treatment of an activated carbon but without

    mentioning the used concentration of NaOCl. Perrad et al. [50]

    also showed by XPS an increase in the formation of carboxylic

    groups after oxidation of an activated carbon cloth by NaOCl

    (0.53 M) for 90 min.

    XPS O1s (Fig. 2) analyses confirmed that oxidation was

    more efficient under ultrasonic irradiation in hydrogen

  • This reduction after oxidation treatment was the highest for

    the AC[NaOCl] sample. A blockage of the micropores due to

    the formation of surface functional groups might be the cause

    of this slight reduction in the surface area as already stated

    before [53]. By contrast, the very small increase in the BET

    specific area (SBET = 809 m2 g1) and in the micropore volume

    (0.32 cm3 g1) for AC700N2 compared to raw AC, might beattributed to the removal of the functional groups from the

    inner pores through the thermal treatment (Table 1).

    The ultramicropore volumes obtained from the DFT PSD

    s)

    42042

    5977

    8

    P/P0

    Fig. 3 N2 adsorptiondesorption isotherms at 77 K of the

    pristine AC (h: adsorption, j: desorption) and AC[NaOCl] (4:adsorption, m: desorption).

    0 1peroxide (14.37 at.0% in sample AC[H2O2]US) compared to the

    silent analog reaction (11.97 at.0% in sample AC[H2O2]).

    XPS analyses (Table 1), indicated higher carbonyl and lac-

    tonic groups amount in AC[H2O2]US compared to AC[H2O2],

    disagreeing with the contents determined by Boehm titra-

    tions that are known to be only semi-quantitative.

    After NaOCl treatment, the increase of the number of C-O

    bonds from phenols and lactones (16.77% of CO for AC[NaO-

    Cl] compared to 10.22% for AC) was consistent with the evolu-

    tion observed by Boehm titrations.

    Whatever the activated carbon, the O1s analyses (Table 2)

    showed OC at.% lower than the CO content obtained from

    the C1s spectra, which could be the signature of the presence

    of ether bridges in all of the activated carbons.

    The O1s analyses after deconvolution (Table 2) indicated

    also O@C at.% lower than the C@O content obtained fromthe C1s spectra, particularly for raw AC and samples treated

    with H2O2 as already reported for other oxidized activated car-

    bons [51,52]. This revealed the difficulties for XPS analysis to

    take into account all the surface oxygen groups species in

    particular ethers and anhydrides, etc.

    The peak at the lowest energies (530.42531.01 eV) related

    to the presence of metal oxides, increased after oxidation by

    H2O2. Indeed, the X-ray micro-analysis showed the presence

    of metals, such as Al and Fe on the surface of AC which might

    have formed metal oxides after their oxidation by hydrogen

    Table 2 Atomic percentages of carbon (C1s) and oxygen (O1

    Signal attribution E(eV) AC

    C1s CC (284.6) 62.0CO (285.13285.55) 10.2C@O (285.99286.75) 6.5OC@O (286.83289.14) 4.4Carbonate, pp transition (288.57291.48) 9.9

    O1s Metal oxide (530.42531.01) 0.8O@C (531.8532.29) 1.8OC (532.96533.77) 2.8Chemisorbed O (534.7535.22) 1.2

    Total atomic O% 6.8

    436 C A R B O N 5 4 ( 2peroxide.

    3.1.2. Porosity characterizationThe nitrogen adsorptiondesorption isotherms for the raw

    (Fig. 3) and modified carbons are quite similar (not shown),

    except for AC[NaOCl] sample (Fig. 3). They were typical of

    microporousmesoporous adsorbents (type I and IV in the IU-

    PAC classification) with a hysteresis loop (H4 types) in the

    desorption branch at relative pressures above 0.5. The adsorp-

    tiondesorption hysteresis on AC isotherm showed clearly

    that liquid nitrogen was condensed in slit-shaped mesopores.

    The N2 isotherms at 77 K showed a slight decrease of the

    BET surface area (Table 3, SBET = 709 m2 g1 for AC[NaOCl],

    781 m2 g1 for AC[H2O2] and 762 m2 g1 for AC[H2O2]US) of

    the oxidized carbons compared to the pristine AC (SBET = 800 -

    m2 g1). The decrease in the specific surface areas after oxida-tion was related only to the decrease in the micropore

    volumes observed either by N2 or CO2 adsorption (Table 3).obtained by XPS analysis of raw and modified AC.

    AC[NaOCl] AC[H2O2] AC[H2O2]US AC700N2

    53.02 51.77 46.74 54.7416.77 12.07 18.84 11.577.02 8.05 9.82 7.214.45 5.82 7.69 4.986.49 10.38 2.53 10.51

    0.83 2.31 2.48 1.026.28 3.51 5.40 4.914.63 3.93 5.71 2.400.51 2.16 0.78 2.67

    12.25 11.91 14.37 11

    0

    50

    100

    150

    200

    250

    300

    350

    0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

    Vads

    (cm

    3 g-1

    STP

    )

    3 ) 4 3 2 4 4 3analysis of CO2 isotherms (Table 3) were found quite similar

    (about 0.14 cm3 g1) for all the modified activated carbonsand slightly lower than to the one of AC (0.18 cm3 g1).

    The N2 DFT pore size distribution of the pristine AC (not

    shown) confirmed that the AC material was both of micropo-

    rous and mesoporous nature. The micropore and mesopore

    volumes computed from the N2 DFT model were 0.31 and

    0.21 cm3 g1, respectively.The N2 adsorptiondesorption isotherm of raw AC and IBP-

    loaded AC (Fig. 4 left) had a very similar profile but the iso-

    therm of the loaded sample was shifted towards lower

    adsorption volumes. This difference is due to a decrease of

    the microporous volume concomitant to IBP adsorption. The

    volume occupied by IBP calculated from loading content:

    0.05 cm3 g1 was similar than to the volume calculated fromthe difference of microporous volumes before after loading

    (0.05 cm3 g1 from N2 adsorption) (Table 3). An activated car-

    bon (AC or treated AC) saturated by IBP at pH 3, where the up-

  • rption and CO2 adsorption studies.

    l] AC[H2O2] AC[H2O2]US AC700N2 IBP loaded AC

    781 762 809 6550.31 0.30 0.32 0.260.25 0.25 0.23 0.220.56 0.55 0.55 0.480.26 0.25 0.27 0.180.14 0.14 0.15 0.12

    0 1 3 ) 4 3 2 4 4 3 437Table 3 Textural properties obtained by N2 adsorption/deso

    Sample AC AC[NaOC

    BET surface area SBET (m2 g1) 800 709

    Micropore volumea (N2) (cm3 g1) 0.31 0.28

    Mesoporous volume (cm3 g1) 0.21 0.21Total pore volumeb (cm3 g1) 0.52 0.49Micropore volumea (CO2) (cm

    3 g1) 0.27 0.23Ultramicropore volumec (CO2) (cm

    3 g1) 0.18 0.14a From Dubinin-Radushkevich.b Single point adsorption at P/P0 = 0.995.c From NLDFT slit pore model.

    Isotherms

    0

    50

    100

    150

    200

    250

    300

    350

    400

    0.000001 0.00001 0.0001 0.001 0.01 0.1P/P0

    Vads

    (cm

    3 g-1

    STP

    )

    C A R B O N 5 4 ( 2take is maximum, contains 130165 mg g1 of adsorbate,which means a volume of 0.11 cm3 g1 for the uptake volumeof IBP (taking into account the volume of IBP : 0.23 nm3 men-

    tioned in Section 2.1). As the IBP uptake volume is lower than

    the microporous volumes of the studied activated carbons,

    this molecule can be accommodated in the micropores (Vol-

    ume of micropores 0.31 cm3 g1 for AC). The ultramicroporevolume values, investigated by CO2 adsorption before and

    after loading (Table 3), have confirmed that these smaller

    micropores were occupied by IBP (IBP uptake volume deduced

    from ultramicropore volumes: 0.06 cm3 g1). The PSD plots of

    pristine AC and IBP-loaded AC (Fig. 4 right) clearly displayed

    that the adsorption occurred in the ultramicropores of diam-

    eter lower than 0.6 nm. Similarly we have studied the micro-porosity of a IBP-loaded AC700N2 by CO2 adsorption (Volume

    of ultramicropores 0.09 cm3 g1 for IBP-loaded AC700N2),and we have confirmed that IBP was also adsorbed into ultra-

    micropores. Mestre et al. [14] studied a series of activated car-

    bons with various surface chemistry and showed that a large

    fraction of ultramicropores were not appropriate to effectively

    remove ibuprofen. In our activated carbon, IBP is preferen-

    tially adsorbed in the slit shape ultramicropores which are

    accessible to the molecule in terms of dimensions.

    3.2. Adsorption kinetics of raw activated carbon

    Fig. 5 shows the effect of contact time on the adsorption of IBP

    onto AC. The equilibrium adsorption time was found at about

    4000 min. The kinetics of adsorption of IBP by AC was

    Fig. 4 Adsorption isotherms of N2 at 77 K (s: AC, h: 56.7 mg g

    loaded AC), and NLDFT pore size distributions of AC (4) and IBP1

    PSD

    0

    0.01

    0.02

    0.03

    0.04

    0.05

    0.06

    3 5 7 9Pore width ()

    Incr

    emen

    tal v

    olum

    e (c

    m3 g

    -1)investigated by four models, namely, the Lagergren pseudo-

    first-order model (1) [54], the pseudo-second-order model (2)

    [55], the Elovich model (3) [56] and the intraparticle diffusion

    model (4) [57].

    The first-order equation is:

    dqtdt

    k1qe qt 1

    where k1 is the rate constant of pseudo-first-order adsorption

    (min1), qe and qt (mg g1) are the amounts of adsorbed IBP at

    equilibrium and at time t (min), respectively.

    The second-order equation is:

    dqtdt

    k2qe qt2 2

    where k2 is the rate constant of pseudo-second-order adsorp-

    tion (g mg1 min1), qe and qt (mg g1) are the amounts of ad-

    sorbed IBP at equilibrium and at time t (min), respectively.

    1 IBP loaded AC at pH 7), and of CO2 at 273 K (4: AC, : IBPloaded AC (), from CO2 isotherms.

    00.10.20.30.40.50.60.70.80.9

    1

    0 1000 2000 3000 4000 5000 6000Time (min)

    C/C0

    Fig. 5 Kinetic result of Ibuprofen adsorption onto AC at

    25 C and pH 3 ([IBP] = 100 mg L1, madsorbent = 10 mg,VIBP = 30 mL, stirring speed = 650 rpm).

  • The Elovich equation is:

    qt 1b

    lna b 1b

    lnt t0 3

    where a is the adsorption initial rate (mg g1 min1, t0 = 1/(a b)) and b is a constant (g mg1) related to the external sur-face area and activation energy of adsorption

    (chemisorption).

    The intraparticle diffusion model is formulated by:

    qt kpt0:5 4

    where kp (mg g1 min1/2) is the intraparticle diffusion rate

    constant.

    adsorption uptake at pH = 7, the latter increased with the oxi-

    dative treatment by H2O2 with or without ultrasound (Fig. 7).

    In addition, IBP adsorption is claimed to operate mainly

    through interactions with the p-electron system of the

    relation coefficient > 0.91) for AC sample (pseudo-first order,

    pH = 3 pH = 7

    132.90 67.140.087 0.1210.949 0.912138.10 69.75

    ) 0.000436 0.0011470.985 0.9522148 14920.092 0.183

    0

    20

    40

    60

    80

    100

    120

    140

    160

    180

    0 1000 2000 3000 4000 5000 6000

    Time (min)

    Ibup

    rofe

    n up

    take

    (mg

    g-1

    )

    AC at pH=3

    pseudo 1storderpseudo 2ndorderIntraparticlediffusionElovich

    438 C A R B O N 5 4 ( 2 0 1 3 ) 4 3 2 4 4 3The fitted parameters of adsorption kinetics at pH 7 and 3

    (Table 4) were calculated from the non linear regressions of

    the integrated Eqs. (1)(3). The profiles of each fitted curve dis-

    played in Fig. 6, in comparison with the experimental kinetics

    at pH 3, were very similar to the ones of fitted kinetic plots at

    pH 7 (not shown). Whatever the pH value, the intraparticle

    diffusion model (Fig. 6) was not in good agreement with the

    experimental kinetics. The best fit of adsorption kinetics at

    pH 3 (R2 = 0.985) was obtained for the pseudo-second-order

    model. The IBP uptake decreased while the pH increased from

    3 to 7. The pseudo-second-order constant k2 obtained at

    pH = 7 was higher than at pH = 3 meaning that the IBP adsorp-

    tion at pH = 7 required higher amount of AC than at pH = 3. As

    a consequence, activated carbon AC has more affinity for IBP

    at acidic pH. On the other hand, the Elovich model showed a

    satisfactory correlation coefficient (R2 = 0.982 at pH = 3). Thus,

    both pseudo-second-order and Elovich models are applicable

    to reproduce the adsorption kinetics of IBP on AC at pH = 3.

    Besides, the Elovich model was found to be the best model

    at pH = 7 (R2 = 0.997). The adsorption initial rate of ibuprofen

    (a) decreased as the pH of the ibuprofen solution became ba-

    sic. The value of 1/b, indicative of the number of sites avail-

    able for adsorption, was higher at pH = 3 (1/b 11) than atpH = 7 (1/b 5).

    3.3. Adsorption isotherms

    3.3.1. Effect of pHpH remains one of the most important factors controlling the

    adsorption capacity of activated carbons, by influencing both

    the chemistry of the surface groups and the IBP charge (pKa of

    IBP equals to 4.91 [58]). Its impact was subsequently studied at

    Table 4 Results of the best IBP kinetic fittings (R-square corpseudo-second order and Elovich models).

    Kinetic model Parameters

    Pseudo-first-order qe (mg g1)

    k1 (min1)

    R2

    Pseudo-second-order qe (mg g1)

    k2 (mg g1 min1

    R2

    Elovich a (mg g1 min1)b (g mg1)

    R2pH pKa-2 and pH pKa+2 (i.e. pH = 3 and 7). For pH valueshigher than pKa(IBP), the molecule was present under its an-

    ionic form. The pHPZC of AC was equal to 7.76 (Table 1). At

    pH < pHPZC, the activated carbon was positively charged.

    Thus, at pH = 7, the negatively charged IBP species were at-

    tracted by the positive AC surface (electrostatic interactions),

    which was not the case at pH = 3. At such a value, the surface

    of AC remained positively charged (pH < pHPZC) whereas IBP

    moieties existed under their neutral form disabling the favor-

    able electrostatic interactions between both. However, the

    dominant dispersive interactions explained the higher

    adsorption capacity at this low pH. Indeed, the uptake of

    IBP by AC was found higher at pH = 3 (Fig. 7, qmax = 130 -

    mg g1) than at pH = 7 (Fig. 7, qmax = 75 mg g1).

    Whatever the explored pH, adsorption capacities of the

    raw and the modified carbons were different. As the studied

    carbons probed by N2 adsorption at 77 K and CO2 adsorption

    at 273 K were very similar in term of microporous and ultra-

    microporous accessible volumes (Section 3.1.2, Table 3), nota-

    bly lower than the uptake volume of IBP (0.11 cm3 g1,Section 3.1.2); the difference in uptakes of the various carbons

    can only be attributed to the variety of their surface

    chemistry.

    While the bleaching treatment led to a decrease of the

    Fig. 6 Fitted profiles of ibuprofen adsorption kinetics of AC

    sample at pH 3 () by pseudo-first-order, pseudo-secondorder, Elovich and intraparticle diffusion models.0.982 0.997

  • pH=7

    0

    20

    40

    60

    80

    100

    120

    140

    160

    180

    0 20 40 60 80

    Ibup

    rofe

    n up

    take

    (mg/

    g)

    an

    P]

    0 1 3 ) 4 3 2 4 4 3 439graphitic rings. One thus would expect that additional acidic

    surface oxygen groups might have withdrawn the electrons

    from the p-system. However, this explanation is not fully sat-

    isfactory here as the oxidation by H2O2 treatment led to an in-

    crease in the adsorption uptake.

    Indeed, the knee of the adsorption isotherm of AC[H2O2]US

    was more pronounced compared to any other samples indi-

    cating that strong interactions with this sample occurred

    even at low concentration. As both XPS characterization and

    Boehm titrations (Section 3.1.1) showed that the main differ-

    ence in terms of resulting surface chemistry, between the

    H2O2 and NaOCl oxidation treatments, laid in the amount of

    carbonyl and lactonic groups; the higher IBP uptake could

    be thus related to the higher amount of carbonyl (also from

    quinones) and lactonic groups, created after H2O2 oxidation.

    These carbonyl groups could have interacted with the aro-

    pH=3

    0

    20

    40

    60

    80

    100

    120

    140

    160

    180

    0 20 40 60 80Equilibrium solution concentration (mg/L)

    Ibup

    rofe

    n up

    take

    (mg/

    g)

    Fig. 7 Adsorption isotherms of ibuprofen at 25 C onto rawAC[H2O2]US, *: AC[NaOCl] and m: AC700N2) at pH 3 and 7 ([IBtime = 5 days, stirring speed = 650 rpm).

    C A R B O N 5 4 ( 2matic ring of ibuprofen in a donoracceptor bond as previ-

    ously reported [40]. In fact, the oxygen lone pair electrons

    are known to be responsible of the formation of donoraccep-

    tor complexes via a dipole moment interaction; the carbonyl

    groups acting as the electron donor and the aromatic ring act-

    ing as the acceptor. As the carbonyl oxygen groups have more

    intense dipole moments than carboxylic acid oxygens, they

    have been reported as stronger donors [33]. In their molecular

    form, the carboxylic acid could also interact in a different

    manner by Van der Waals, hydrogen bonds with the carbox-

    ylic groups of IBP, enabling the formations of dimers [34].

    The functionalization of activated carbons by oxidation with

    NaOCl yielded to specific oxygen surface groups (mainly phe-

    nolic). Some of these groups might be inappropriate for ben-

    eficial interaction with IPB, leading to a decrease of the

    dispersive interactions, through decreasing the p-electron

    density of the graphene layers [59]. Thus the uptake of

    AC[NaOCl] remained rather similar to AC at pH 3 and lower

    at pH 7.

    A significant increase of the adsorption capacity compared

    to AC was observed for AC700N2 aged sample: the uptake,

    equal to 92 mg g1 at pH = 7, reached a maximum of

    160 mg g1 at pH = 3, showing that the adsorption was alsopromoted on slightly oxidized activated carbons. Moreover,

    at pH 7, the electrostatic attraction contribution might in-

    crease slightly in the AC700N2 sample due to the shift of

    pHPZC towards basic values (pHPZC of AC700N2 equals to

    9.08). At pH 3, the increase in adsorption uptake for

    AC700N2, was attributed to beneficial interactions of IBP with

    the oxygenated surface groups (mainly carbonyl groups).

    3.3.2. Effect of temperatureThe effect of temperature was studied at pH 3 in order to

    determine the thermodynamic parameters of adsorption of

    IBP on three different activated carbons (Figs. 810: samples

    AC, AC[H2O2]US and AC700N2, respectively) and to compare

    the interactions energies.

    As the temperature increased from 298 to 328 K, the

    adsorption capacities increased from 130 to 180 mg g1 for

    Equilibrium solution concentration (mg/L)

    d modified activated carbons (: AC, d: AC[H2O2], j:= 5100 mg L1, madsorbent = 10 mg, VIBP = 30 mL, contactAC (Fig. 8), from 140 to 160 mg g1 for AC[H2O2]US (Fig. 9)and from 163 to 230 mg g1 for AC700N2 (Fig. 10).

    To investigate the temperature dependency of the adsorp-

    tion capacity, four equilibrium models were analyzed, including

    Langmuir, Freundlich, LangmuirFreundlich and Redlich

    Peterson. All the isotherms were simulated using an iterative

    0

    50

    100

    150

    200

    250

    0 20 40 60 80 100Ce (mg L-1)

    Qad

    s(m

    g g-

    1 )

    Fig. 8 Experimental adsorption isotherms of ibuprofen on

    AC at pH 3, and at 298 K (), 313 K (m) and 328 K (j) fitted bythe LangmuirFreundlich model (lines).

  • 0 10

    50

    100

    150

    200

    250

    0 10 20 30 40 50 60 70 80Ce (mg L-1)

    Qad

    s(m

    g g-

    1 )

    Fig. 9 Experimental adsorption isotherms of ibuprofen on

    AC[H2O2]US at pH 3, and at 298 K (), 313 K (m) and 328 K (j)fitted by the LangmuirFreundlich model (lines).

    150

    200

    250

    s(m

    g g-

    1 )

    440 C A R B O N 5 4 ( 2procedure based on a non-linear least-squares algorithm. The

    Langmuir adsorption isotherm equation, expressed as follows

    requires for its applicability a mono-layered coverage on the

    surface of adsorbent [60]:

    qe qmaxKLCe1 KLCe

    where qe, KL, qmax and Ce are the ibuprofen uptake at equilib-

    rium (mg g1), the Langmuir constant (L mg1), the monolayeradsorption capacity (mg g1) and the solution concentrationat equilibrium (mg L1), respectively.

    The Freundlich equation is applicable for multi component

    adsorption. The Freundlich isotherm is expressed by:

    qe KF C1=ne

    where KF is the Freundlich constant (mg1-1/n L1/n g1) and n is

    the Freundlich exponent.

    The Freundlich equation was successfully used to fit the

    experimental data for many fluidsolid systems, but one of

    the main drawbacks of this equation is that it doesnt tend

    to a limiting value while the equilibrium concentration in-

    creases. The LangmuirFreundlich equation is a way to over-

    come this problem [61]. It is expressed as follows:

    qe qmaxKLC

    ne

    1 KLCne

    0

    50

    100

    0 10 20 30 40 50 60 70 80Ce (mg L-1)

    Qad

    Fig. 10 Experimental adsorption isotherms of ibuprofen on

    AC700N2 at pH 3, and at 298 K (), 313 K (m) and 328 K (j)fitted by the LangmuirFreundlich model (lines).The RedlichPeterson isotherm, proposed by considering

    the limitations of Freundlich and Langmuir isotherms [62],

    is given by:

    qe KRCe

    1 aRCbewhere KR (L g

    1) and aR (L mg1) are the RedlichPeterson iso-

    therm constants and b is the exponent, ranging between 0

    and 1.

    The experimental data of IBP adsorption on AC and modi-

    fied carbons at pH 3 were well fitted by the LangmuirFreund-

    lich and RedlichPeterson models. In terms of R2, the

    LangmuirFreundlich model provided better fitting

    (R2 > 0.96). Figs. 810 present the fittings by the Langmuir

    Freundlich equation of isotherms at pH 3.

    The estimated parameters for the LangmuirFreundlich

    model are reported in Table 5.

    The adsorption capacities, qmax increased as temperature

    increased for each carbon indicating an endothermic adsorp-

    tion. The LangmuirFreundlich constants KL were the highest

    for AC[H2O2]US suggesting that this activated carbon has a

    higher adsorption affinity to IBP than AC and AC700N2. Com-

    petitive adsorption of water molecules and IBP might have

    also occurred as the carbon became more hydrophilic upon

    oxidation by H2O2 with ultrasonic irradiation [34]. Moreover,

    as n value is near to unity for AC[H2O2]US, the adsorption pro-

    cess of IBP rather follows the Langmuir model. The n increase

    for AC and AC700N2 together with temperature suggested

    that adsorption affinity of IBP for the substrates increased

    as heating.

    3.4. Thermodynamic parameters

    The isosteric Gibbs free energy DG of adsorption of IBP ontothree activated carbons (AC, AC[H2O2]US and AC700N2) was

    calculated using the results obtained from the simulations

    (Table 5) and the following equation:

    LnKd LnCa=Ce DG=RTwhere Kd is the equilibrium constant, Ca is the amount of IBP

    adsorbed at equilibrium (mg L1), Ce is the concentration ofIBP remaining in the solution at equilibrium (mg L1), T isthe solution temperature (K) and R is the perfect gas constant.

    The isosteric enthalpy DH and entropy DS of adsorptionwere also calculated from the slope and the intercept of the

    plot of Ln(Kd) versus 1/T, respectively, using the equation:

    DG = DH TDS. Thermodynamic parameters were calcu-lated for the adsorption of IBP onto raw and modified acti-

    vated carbons at constant adsorption uptake value equal to

    100 mg g1 and at pH = 3 (Table 6).At pH = 3, the adsorption capacity of IBP onto all studied

    activated carbons increased as the temperature increased

    from 298 K to 328 K, that was consistent with the positive val-

    ues obtained for DH. This suggested that the adsorption ofIBP by the three activated carbons was an endothermic pro-

    cess at this pH in agreement with the work of Dubey et al.

    [30] for the adsorption of IBP at natural pH on an activated

    carbon prepared from Artemisia vulgaris. At pH 3, the esti-

    3 ) 4 3 2 4 4 3mated values for DG were 1.1, 7.3 and 4.5 kJ moL1,respectively for AC, AC[H2O2]US and AC700N2, indicating a

  • pti

    A

    f ad

    0 1spontaneous physisorption process [63]. The DH values wereless endothermic as increasing the amount of oxygenated

    groups, suggesting an exothermic interaction of some of

    them with IBP.

    The positive values of DS showed an increase of the disor-der at the solid-solution interface, probably due to the desorp-

    tion of solvent molecules (water and methanol) prior to IBP

    adsorption. The positive values of DS could indicate the com-petitive adsorption between IBP and solvent molecules.

    The highest absolute value of Gibbs energy (7.3 kJ mol1)for the AC[H2O2]US sample suggested the presence of high en-

    ergy adsorption sites related to particular oxygenated groups

    formed by sonication in hydrogen peroxide. The characteriza-

    tion of the surface chemistry suggested that these sites might

    be carbonyl (or lactonic) groups mainly formed through

    hydrogen peroxide oxidation. The increase in the adsorption

    Table 5 Parameters of the LangmuirFreundlich fits of adsorat temperature in the range 298328 K (qmax in mg g

    1).

    T (K) AC

    qmax KL n qmax

    298 160.0 0.61 0.25 146.6313 179.1 0.45 0.83 152.0328 185.2 0.75 0.86 159.9

    Table 6 Isosteric Gibbs free energy, enthalpy and entropy oQads = 100 mg g

    1) at pH 3.

    Sample AC

    DG (kJ mol1) 1.1DH (kJ mol1) 74.6DS (kJ K1 mol1) 254.1

    C A R B O N 5 4 ( 2energy compared to pristine AC (4.5 kJ mol1 compared to1.1 kJ mol1) was also observed for aged AC700N2, maybethrough the interaction with quite similar beneficial oxygen-

    ated groups, but in a lower amount.

    4. Conclusion

    We showed that the adsorption of ibuprofen occurred prefer-

    entially into the ultramicropores of the raw and N2 heat trea-

    ted activated carbons. The adsorption of IBP is strongly

    dependent on pH and is promoted by dominant dispersive

    interactions which are increasing when the molecule is pres-

    ent on its neutral form. For this reason at pH 7, as the electro-

    static interaction predominates, the adsorption uptake

    subsequently decreases compared to pH 3 where the adsorp-

    tion was found endothermic and entropically promoted as

    the competition with solvent molecules occurs. We brought

    out that some particular surface chemistry nature, such as

    carbonyl or lactonic or quinone groups can influence posi-

    tively the adsorption uptake of IBP.

    As an example, oxidation treatment in hydrogen perox-

    ide combined or not with ultrasonic irradiation enables

    increasing slightly the adsorption capacity, and the Gibbsenergy of adsorption at acidic pH. This might be attributed

    to a donoracceptor mechanism between the created car-

    bonyl groups (also in lactonic and carboxylic groups) and

    the p aromatic ring of ibuprofen. Both appropriate oxidation

    or reduction treatments can increase the adsorption uptake,

    but the reduction of the surface functionality (by annealing

    at 973 K under N2) also gives a higher adsorption capacity

    whatever the pH value as the carbon becomes slightly oxi-

    dized after aging, favoring dispersive interactions. However

    treatment with NaOCl does not benefit to ibuprofen uptake

    because it preferentially leads to the formation of phenolic

    groups and a lack of carbonyl adsorption sites [32]. The

    LangmuirFreundlich model allows describing well the ibu-

    profen adsorption at pH 3. DH values indicate that theadsorption of ibuprofen on the studied activated carbons

    is an endothermic process, in relation with solvent

    AC[H2O2]US AC700N2

    7.3 4.58.1 31.9

    51.6 122.2sorption of ibuprofen (at constant adsorption uptake valueon isotherms of ibuprofen on the different activated carbons

    C[H2O2]US AC700N2

    KL n qmax KL n

    1.23 0.95 190.7 0.30 0.700.98 1.52 212.5 0.46 0.871.22 1.08 256.2 0.35 0.87

    3 ) 4 3 2 4 4 3 441interaction.

    Acknowledgments

    Generous MIRA grant of the Region Rhone-Alpes (France) is

    thankfully acknowledged.

    R E F E R E N C E S

    [1] Tambosi JL, Yamanaka LY, de Fatima Peralta Muniz MoreiraHJJR, Schroder HF. Recent research data on the removal ofpharmaceuticals from sewage treatment plants (STP). QuimNova 2010;33(2):41120.

    [2] Fent K, Weston A, Caminada D. Ecotoxicology of humanpharmaceuticals. Aquat Toxicol 2006;76(2):12259.

    [3] Jones OA, Lester JN, Voulvoulis N. Pharmaceuticals: a threatto drinking water. Trends Biotechnol 2005;23(4):1637.

    [4] Stackelberg PE, Furlong ET, Meyer MT, Zaugg SD, HendersonAK, Reissman DB. Persistence of pharmaceutical compoundsand other organic wastewater contaminants in aconventional drinking-water treatment plant. Sci TotalEnviron 2004;329(13):99113.

  • 442 C A R B O N 5 4 ( 2 0 1 3 ) 4 3 2 4 4 3[5] Verlicchi P, Galletti A, Petrovic M, Barcelo D. Hospital effluentsas a source of emerging pollutants: an overview ofmicropollutants and sustainable treatment options. J Hydrol2010;389(34):41628.

    [6] Miege C, Choubert JM, Ribeiro L, Eusebe M, Coquery M.Removal efficiency of pharmaceuticals and personal careproducts with varying wastewater treatment processes andoperating conditions conception of a database and firstresults. Water Sci Technol 2008;51(1):4956.

    [7] Kot-Wasik A, Debska J, Namiesnik J. Analytical techniques instudies of the environmental fate of pharmaceuticals andpersonal-care products. TrAC Trends Anal Chem2007;26(6):55768.

    [8] Mendez-Arriaga F, Torres-Palma RA, Petrier C, Esplugas S,Gimenez J, Pulgarin C. Ultrasonic treatment of watercontaminated with ibuprofen. Water Res 2008;42:42438.

    [9] Madhavan J, Grieser F, Ashokkumar M. Combined advancedoxidation processes for the synergistic degradation ofibuprofen in aqueous environments. J Hazard Mater2010;178(13):2028.

    [10] Ciraco L, Anjo C, Correia J, Pacheco MJ, Lopes A.Electrochemical degradation of Ibuprofen on Ti/Pt/PbO2 andSi/BDD electrodes. Electrochim Acta 2009;54(4):146472.

    [11] Haginaka J, Sanbe H. Uniform-sized molecularly imprintedpolymers for 2-arylpropionic acid derivatives selectivelymodified with hydrophilic external layer and theirapplications to direct serum injection analysis. J Am ChemSoc 2000;72(1):520610.

    [12] Farrington K, Regan F. Investigation of the nature of MIPrecognition: the development and characterisation of a MIPfor Ibuprofen. Biosens Bioelectron 2007;22(6):113846.

    [13] Zhu Y-F, Shi J-L, Li Y-S, Chen H-R, Shen W-H, Dong X-P.Storage and release of ibuprofen drug molecules in hollowmesoporous silica spheres with modified pore surface.Micropor Mesopor Mater 2005;85(12):7581.

    [14] Mestre AS, Pires J, Nogueira JMF, Parra JB, Carvalho AP, AniaCO. Waste derived activated carbons for removal of ibuprofenfrom solution: role of surface chemistry and pore structure.Biores Technol 2009;100(5):17206.

    [15] Mestre AS, Pires J, Nogueira JMF, Carvalho AP. Activatedcarbons for the adsorption of ibuprofen. Carbon2007;45(10):197988.

    [16] Gupta VK, Carrott PJM, Ribeiro Carrott MML, Suhas TL. Low-cost adsorbents: growing approach to wastewater treatment a review. Environ Sci Technol 2009;39:783842.

    [17] Gupta VK, Rastogi A. Biosorption of lead from aqueoussolutions by green algae Spirogyra species: kinetics andequilibrium studies. J Hazard Mater 2008;152(1):40714.

    [18] Gupta VK, Rastogi A, Nayak A. Biosorption of nickel ontotreated alga (Oedogonium hatei): application of isotherm andkinetic models. J Colloid Interface Sci 2010;342(2):5339.

    [19] Gupta VK, Rastogi A, Nayak A. Adsorption studies on theremoval of hexavalent chromium from aqueous solutionusing a low cost fertilizer industry waste material. J ColloidInterface Sci 2010;342(1):13541.

    [20] Gupta VK, Jain R, Mittal A, Mathur M, Sikarwar S.Photochemical degradation of the hazardous dye Safranin-Tusing TiO2 catalyst. J Colloid Interface Sci 2007;309:4649.

    [21] Gupta VK, Mittal A, Kurup L, Mittal J. Adsorption of ahazardous dye, erythrosine, over hen feathers. J ColloidInterface Sci 2006;304(1):527.

    [22] Gupta VK, Ali I, Saini VK. Adsorption studies on the removalof Vertigo Blue 49 and Orange DNA13 from aqueous solutionsusing carbon slurry developed from a waste material. JColloid Interface Sci 2007;315(1):8793.[23] Gupta VK, Mittal A, Gajbe V, Jyoti M. Removal and recovery ofthe hazardous azo dye acid orange 7 through adsorption overwaste materials: bottom ash and de-oiled soya. Ind EngChem Res 2006;45(4):144653.

    [24] Gupta VK, Jain R, Malathi S, Nayak A. Adsorptiondesorptionstudies of indigocarmine from industrial effluents by usingdeoiled mustard and its comparison with charcoal. J ColloidInterface Sci 2010;348:62833.

    [25] Gupta VK, Jain R, Varshney S. Removal of Reactofix goldenyellow 3 RFN from aqueous solution using wheat huskanagricultural waste. J Hazard Mater 2007;142:4438.

    [26] Gupta VK, Ali I. Removal of endosulfan and methoxychlorfrom water on carbon slurry. Environ Sci Technol2008;42:76670.

    [27] Gupta VK, Gupta B, Rastogi A, Agarwal S, Nayak A. Pesticidesremoval from waste water by activated carbon prepared fromwaste rubber tire. Water Res 2011;45:404755.

    [28] Gupta VK, Ali I, Saini VK. Defluoridation of wastewatersusing waste carbon slurry. Water Res 2007;41(15):330716.

    [29] Redding AM, Cannon FS, Snyder SA, Vanderford BJ. A QSAR-like analysis of the adsorption of endocrine disruptingcompounds, pharmaceuticals, and personal care products onmodified activated carbons. Water Res 2009;43(15):384961.

    [30] Dubey SP, Dwivedi AD, Sillanpaa M, Gopal K. Artemisiavulgaris-derived mesoporous honeycomb-shaped activatedcarbon for ibuprofen adsorption. Chem Eng J2010;165(2):53744.

    [31] Santiago M, Stuber F, Fortuny A, Fabregat A, Font J. Modifiedactivated carbons for catalytic wet air oxidation of phenol.Carbon 2005;43:213445.

    [32] Haydar S, Ferro-Garca MA, Rivera-Utrilla J, Joly JP. Adsorptionof p-nitrophenol on an activated carbon with differentoxidations. Carbon 2003;41(3):38795.

    [33] Mattson JS, Mark Jr HB, Malbin MD, Weber Jr WJ, CrittendenJC. Surface chemistry of active carbon: specific adsorption ofphenols. J Colloid Interface Sci 1969;31(1):11630.

    [34] Franz M, Arafat HA, Pinto NG. Effect of chemical surfaceheterogeneity on the adsorption mechanism of dissolvedaromatics on activated carbon. Carbon 2000;38:180719.

    [35] Yalkowsky SH, Dannenfelser RM. Aquasol database ofaqueous solubility. Version 5, College of Pharmacy, Universityof Arizona, Tuscon, AZ; 1992.

    [36] Lee T, Kuo CS, Chen YH. Solubility, polymorphism,crystallinity, and crystal habit of acetaminophen andibuprofen by initial solvent screening. Pharm Technol2006;30(10):7292.

    [37] Manrique J, Martinez F. Solubility of Ibuprofen in someethanol + water cosolvent mixtures at several temperatures.Lat Am J Pharm 2007;26(3):34454.

    [38] Gaid K, Cavelier C, Martin G. Mechanisms of biologicalpurification on activated carbon. Water Res 1982;16(1):717.

    [39] Lerdkanchanaporn S, Dollimore D. The evaporation ofIbuprofen from Ibuprofen starch mixtures usingsimultaneous TG-DTA. Thermochim Acta 2000;357358:718.

    [40] Gregg SJ, Sing KSW. Adsorption surface area and porosity. 2nded. New York: Academic Press; 1982.

    [41] Jagiello J, Olivier JP. A simple two-dimensional nldft model ofgas adsorption in finite carbon pores. application to porestructure analysis. J Phys Chem C 2009;113:193825.

    [42] Jagiello J, Thommes M. Comparison of DFT characterizationmethods based on N2, Ar, CO2, and H2 adsorption applied tocarbons with various pore size distributions. Carbon2004;42:122732.

    [43] Reffas A, Bernardet V, David B, Reinert L, Bencheikh LehocineM, Dubois M, et al. Carbons prepared from coffee grounds byH3PO4 activation: Characterization and adsorption ofmethylene blue and Nylosan Red N-2RBL. J Hazard Mater2010;175(1):77988.[44] Boehm HP. Surface oxides on carbon and their analysis: acritical assessment. Carbon 2002;40(2):1459.

  • [45] Xie Y, Sherwood PMA. X-ray photoelectron-spectroscopicstudies of carbon fiber surfaces. 11. Differences in the surfacechemistry and bulk structure of different carbon fibers basedon poly(acrylonitrile) and pitch and comparison with variousgraphite samples. Chem Mater 1990;2(3):2939.

    [46] Voncina DB, Majcen-Le-Marechal A. Reactive dyedecolorization using combined ultrasound/H2O2. Dyes Pigm2003;59(2):1739.

    [47] Doktycz SJ, Suslick KS. Interparticle collisions driven byultrasound. Science 1990;247(4946):10679.

    [48] Menendez JA, Phillips J, Xia B, Radovic LR. On themodification and characterization of chemical surfaceproperties of activated carbon: in the search of carbons withstable basic properties. Langmuir 1996;12(18):440410.

    [49] Cagnon B, Py X, Guillot A, Joly JP, Berjoan R. Pore modificationof pitch-based activated carbon by NaOCl and air oxidation/pyrolysis cycles. Microp Mesop Mater 2005;80:18393.

    [50] Perrard A, Retailleau L, Berjoan R, Joly JP. Liquid phaseoxidation kinetics of an ex-cellulose activated carbon clothby NaOCl. Carbon 2012;50:222634.

    [51] Salgado JRC, Duarte RG, Ilharco LM, Botelho do Rego AM,Ferraria AM, Ferreira MGS. Effect of functionalized carbon asPt electrocatalyst support on the methanol oxidationreaction. Appl Catal B 2011;102:496504.

    [52] Park S-J, Jung WY. Effect of KOH activation on the formationof oxygen structure in activated carbons synthesized frompolymeric precursor. J Colloid Interface Sci 2002;250:938.

    [53] Valdes H, Sanchez Polo M, Rivera Utrilla J, Zaror CA. Effect ofozone treatment on surface properties of activated carbon.Langmuir 2002;18:21116.

    [54] Lagergren S. Zur theorie der sogenannten adsorption geloster stoffe. Kungliga Svenska VetenskapsakademiensHandlingar 1898;24(4):139.

    [55] Ho Y-S, McKay G. Pseudo-second order model for sorptionprocesses. Process Biochem 1999;34(5):45165.

    [56] Chein SH, Clayton WR. Application of Elovich equation to thekinetics of phosphate release and sorption in soils. Soil SciSoc Am J 1980;44(2):2658.

    [57] Weber WJ, Morris JC. Kinetics of adsorption of carbon fromsolutions. J Sanit Eng Div Amer Soc Civ Eng 1963;89(2):3163.

    [58] Lindqvist N, Tuhkanen T, Kronberg L. Occurrence of acidicpharmaceuticals in raw and treated sewages and in receivingwaters. Water Res 2005;39(11):221928.

    [59] Donnet JB. Structure and reactivity of carbons: from carbonblack to carbon composites. Carbon 1982;20(4):26782.

    [60] Zhu S, Yang N, Zhang D. Poly(N, N-dimethylaminoethylmethacrylate) modification of activated carbon for copperions removal. Mater Chem Phys 2009;113(23):7849.

    [61] Yao C. Extended and improved Langmuir equation forcorrelating adsorption equilibrium data. Sep Purif Technol2000;19(3):23742.

    [62] Vasanth Kumar K, Porkodi K. Comments on adsorption of 4-chlorophenol from aqueous solutions by xad-4 resin:isotherm, kinetic, and thermodynamic analysis. J HazardMater 2007;143(12):5989.

    [63] Silva SP, Sousa S, Rodrigues J, Antunes H, Porter JJ, GoncalvesI, et al. Adsorption of acid orange 7 dye in aqueous solutionsby spent brewery grains. Sep Purif Technol 2004;40(3):30915.

    C A R B O N 5 4 ( 2 0 1 3 ) 4 3 2 4 4 3 443

    The effects of the surface oxidation of activated carbon, the solution pH and the temperature on adsorption of ibuprofen1 Introduction2 Experimental2.1 Ibuprofen2.2 Raw and modified activated carbons2.3 N2 adsorptiondesorption at 77K and CO2 adsorption at 273K2.4 pHs of carbons suspended in water2.5 Boehm titrations2.6 X-ray photoelectron spectroscopy2.7 Adsorption experiments: kinetics and isotherms

    3 Results and discussion3.1 Characterization of activated carbons3.1.1 Surface chemistry3.1.2 Porosity characterization

    3.2 Adsorption kinetics of raw activated carbon3.3 Adsorption isotherms3.3.1 Effect of pH3.3.2 Effect of temperature

    3.4 Thermodynamic parameters

    4 ConclusionAcknowledgmentsReferences