fotocatalisis nitrofenol grafito_tio2

Upload: angel-logan

Post on 14-Apr-2018

218 views

Category:

Documents


0 download

TRANSCRIPT

  • 7/30/2019 Fotocatalisis Nitrofenol Grafito_TiO2

    1/12

    Photochemical &Photobiological Sciences

    PAPER

    Cite this: Photochem. Photobiol. Sci., 2013,

    12, 1091

    Received 24th November 2012,

    Accepted 25th March 2013

    DOI: 10.1039/c3pp25398h

    www.rsc.org/pps

    Photoelectrochemical oxidation of p-nitrophenol on anexpanded graphiteTiO2 electrode

    B. Ntsendwana,a S. Sampath,a,b B. B. Mambaa and O. A. Arotiba*a

    In the quest for more efficient photoanodes in the photoelectrochemical oxidation processes for organic

    pollutant degradation and mineralisation in water treatment, we present the synthesis, characterisation

    and photoelectrochemical application of expanded graphiteTiO2 composite (EGTiO2) prepared using

    the solgel method with organically modified silicate. The BrunauerEmmettTeller surface area analyser,

    ultravioletvisible diffuse reflectance, scanning electron microscopy, energy dispersive spectroscopy, X-ray

    diffractometry, Raman spectrometry and X-ray photoelectron spectroscopy were employed for the

    characterisation of the composites. The applicability of the EGTiO2 as photoanode material was investi-

    gated by the photoelectrochemical degradation of p-nitrophenol as a target pollutant in a 0.1 M Na2SO4

    (pH 7) solution at a current density of 5 mA cm2. After optimising the TiO2 loading, initial p-nitrophenol

    concentration, pH and current density, a removal efficiency of 62% with an apparent kinetic rate constant

    of 10.4 103 min1 was obtained for the photoelectrochemical process as compared to electrochemical

    oxidation and photolysis, where removal efficiencies of 6% and 24% were obtained respectively after

    90 min. Furthermore, the EGTiO2 electrode was able to withstand high current density due to its high

    stability. The EGTiO2 electrode was also used to degrade 0.3 104 M methylene blue and 0.1 104 M

    Eosin Yellowish, leading to 94% and 47% removal efficiency within 120 reaction time. This confirms the

    suitability of the EGTiO2 electrode to degrade other organic pollutants.

    1. Introduction

    Advanced oxidation processes (AOPs) have been identified as

    promising alternative techniques for water treatment due to

    their ability to effectively degrade a wide range of organic pol-

    lutants by means of in situ generated powerful oxidants such

    as hydroxyl radicals.19 Among the AOPs, electrochemical oxi-

    dation of aqueous wastes containing biorefractory organics

    such as phenolic compounds has been shown to possess some

    advantages such as ease of automation, high efficiency and

    environmental compatibility.1012 However, this process is

    limited by mass transfer of organic matter in the bulk solution

    to the surface of the electrode and is generally accompanied by

    side reactions such as oxygen or chlorine evolution due to the

    high voltage required to destroy the organic molecules inaqueous solutions. These shortcomings can be improved by

    the combination of electrochemical oxidation systems with

    other processes.

    The application of heterogeneous photocatalysis involving

    the use of a semiconducting photocatalyst (e.g. TiO2) in the

    degradation of organic pollutants has been extensively studied.

    TiO2 is capable of destroying more than 3000 types of refractory

    organic compounds under ultraviolet radiation.13 Upon light

    irradiation with energies greater than its band gap, a TiO2semiconductor will be excited to generate electronhole pairs.

    The electrons and holes will migrate from the conduction and

    valence bands to the solid surface to respectively initiate reduc-

    tive and oxidative reactions. The holes will oxidize the surface

    adsorbed water or hydroxyl ions to form hydroxyl radicals

    (OH), while the electron can reduce the dissolved O2 mole-

    cules to form various species, such as O2, H2O, H2O2 and

    OH.14 Such oxygen-containing species can be photocatalyti-

    cally active in the mineralization of organic contaminants andthe inactivation of microorganisms such as bacteria and

    viruses. However, photocatalytic oxidation technology suffers

    from low photo efficiency owing to the slow rate of electron

    transport and high rates of recombination of the photogene-

    rated electronhole pairs.11,12 It has been reported that the

    application of electrical energy can alleviate the problems

    associated with electronhole recombination by driving away

    the photoelectrons from the semiconductorelectrolyte inter-

    face, leading to more efficient degradation of organic contami-

    nants due to the increased concentration of OH radicals.10Electronic supplementary information (ESI) available. See DOI:

    10.1039/c3pp25398h

    aDepartment of Applied Chemi stry, University of Johannesburg, P.O. Box 17011,

    Doornfontein, 2028, South Africa. E-mail: [email protected] of Inorganic and Physical Chemistry, Indian Institute of Science,

    Bangalore, 560012, India

    This journal is The Royal Society of Chemistry and Owner Societies 2013 Photochem. Photobiol. Sci., 2013, 12, 10911102 | 1091

    View Article OnlineView Journal | View Issue

    http://www.rsc.org/ppshttp://pubs.rsc.org/en/journals/journal/PP?issueid=PP012006http://pubs.rsc.org/en/journals/journal/PPhttp://dx.doi.org/10.1039/c3pp25398hhttp://www.rsc.org/pps
  • 7/30/2019 Fotocatalisis Nitrofenol Grafito_TiO2

    2/12

    The combination of heterogeneous photocatalysis and elec-

    trochemical oxidation techniques generally termed photoelec-

    trochemical processes is fast becoming a plausible route to

    organic waste degradation in water treatment.1517 A typical

    combination of heterogeneous photocatalysis and electroche-

    mical oxidation involves the immobilization of TiO2 powder

    (a photoactive material) on a conductive material upon which

    a bias anodic potential is applied.17 Photoelectrochemical oxi-

    dation is more effective than photocatalytic oxidation since thephotoactive and the conducting materials can simultaneously

    generate hydroxyl radicals which are responsible for degra-

    dation. The main advantages of the immobilization of the

    catalyst are the avoidance of losses in the recovery process and

    the enhanced photocatalytic efficiency through the use of an

    external electric field.16 The anodic applied potential should

    be kept low to prevent the photocatalyst from falling off and

    thus prolong the service life of the photoelectrode. Conversely,

    the use of very low anodic potentials can reduce the efficiency

    of the electrochemical oxidation process. Therefore, the photo-

    electrochemical oxidation technique for the degradation and

    mineralization of organic pollutants should be optimised forenhanced photoefficiency and high stability.

    The use of TiO2 photo anodes for degradation of pollutants

    has been explored.1822 The preparation of these TiO2 photo-

    anodes involves the use of expensive conducting materials such

    as Indium doped Tin Oxide (ITO), Fluorine doped Tin Oxide

    (FTO) glass and Ti sheet. ITO and FTO glass substrates suffer

    from low conductivity. The scarcity and increasing price of

    indium and low production volume of high quality FTO limit

    the wide application of ITO and FTO glass as suitable sub-

    strates.23,24 The highly conductive Ti is also limited due to high

    cost.25 Recently, carbon materials such as activated carbon,

    carbon nanotubes and graphite have been used as substrates

    for immobilization of TiO2 due to their low cost and high con-

    ductivity. These carbon nanoTiO2 composites, such as carbon

    coating of photoactive anatase-type TiO2, carbonTiO2 photo-

    catalyst, TiO2-mounted activated carbon, and other modified

    nanoTiO2 particles, have shown good photocatalytic activity

    for the decomposition of organic compounds.2630 Palmisano

    et al.25 also reported the preparation of a photoanode by sup-

    porting TiO2 onto graphite rods for photoelectrochemical oxi-

    dation of p-nitrophenol. However, these forms of carbon suffer

    some challenges such as difficulty in preparation on a large

    scale. In addition, the processing of these composites into elec-

    trodes requires substrates such as Ti or ITO glass or the use of

    an adhesive. The use of substrates affects the loading of thecatalyst and the performance of these composites tends to

    deteriorate due to the use of adhesive during fabrication.31,32

    Contrary to its counterparts, expanded graphite (EG) can be

    easily processed to electrodes without using an adhesive or a

    binder that can be degraded during treatment. Expanded graph-

    ite is a low density carbon material which exhibits a series of

    unique properties such as stability to aggressive media, develo-

    ped specific surface, electrical conductivity, high temperature

    resistance, compatibility and flexibility.3336 It is produced when

    graphite intercalation compounds (GICs) are given a thermal

    shock, the intercalates vaporise and tear the layers apart. This

    leads to an expansion in the c direction, resulting in a puffed-

    up material. EG exhibits a porous structure which can trap or

    accommodate foreign compounds leading to the formation of

    composite materials. The graphite particles in EG can be recom-

    pressed or re-stacked without any binder and the restacking

    mechanism results in the interlocking of the layers during com-

    pression.37 EG has a preferential orientation of the basal plane

    when compressed. It is a very good adsorption substrate due toits near perfect crystallographic face and has better homogen-

    eity of the surface than any other form of graphite. EG is used

    for seals, high temperature gaskets, catalyst supports and elec-

    tromagnetic shielding materials.38,39

    In addition, EG has been used as a substrate to trap titania

    powder for the removal of heavy oils and photocatalytic degra-

    dation of chlorinated phenoxyacetic acids.40,41 The composite

    material demonstrated high stability and enhanced efficiency.

    However, to the best of our knowledge no attempt has been

    made at using EG material as a support anode (substrate) for

    photoelectrochemical oxidation of organic pollutants in water.

    This paper focuses on the preparation and characterisationof expanded graphiteTiO2 composite (EGTiO2) photoanode

    using the solgel derived organically modified silicate binder

    method and its application in the photoelectrochemical degra-

    dation of p-nitrophenol in aqueous medium. The EGTiO2composite preparation method affords a high stability of the

    catalyst.42 The enhanced stability can enable the photoelec-

    trode to withstand high current density and anodic potential,

    which will then increase the production of hydroxyl radicals

    and reduce the recombination rate of photogenerated elec-

    trons and holes respectively.

    2. Experimental procedure

    2.1 Materials and apparatus

    Methyltrimethoxysilane (MTMOS), p-nitrophenol, methyl blue,

    eosin yellowish, methanol and HCl were purchased from

    Sigma-Aldrich. Titania, P-25 was supplied by Degussa Corp.

    The characterisation of expanded graphite and its composite

    was carried out using SEM/EDS (ESEM quanta), specific surface

    areas were measured by BrunauerEmmettTeller (BET) nitro-

    gen adsorptiondesorption (Shimadzu, Micromeritics ASAP

    2010 Instrument), UV-Vis diffuse reflectance spectroscopy (Shi-

    madzu 2450) of dry powders using BaSO4 was used as a reflec-

    tance standard, and X-ray photoelectron spectra were recordedwith an ESCA-3 Mark II spectrometer (VG Scientific, UK) using

    Al K radiation (1486.6 eV), an X-ray diffractometer (Bruker D8

    with Cu-K) and Raman spectroscopy (Lab RAM HIR, Horida

    Jobin Xvon, France; using 514.5 nm air cooled with Ar+ laser)

    with 50 objective and a laser intensity of 1.3 mW.

    The experiments were carried out in a photoreactor with an

    approximate volume of 100 mL. EG and EGTiO2 with 1 cm2

    area were used as working electrodes, Ag/AgCl (3.0 M KCl) and

    platinum foil as a reference and a counter electrode respect-

    ively. A potentiostat/galvanostatic with a voltage range of 13 V

    Paper Photochemical & Photobiological Sciences

    1092 | Photochem. Photobiol. Sci., 2013, 12, 10911102 This journal is The Royal Society of Chemistry and Owner Societies 2013

    View Article Online

    http://dx.doi.org/10.1039/c3pp25398h
  • 7/30/2019 Fotocatalisis Nitrofenol Grafito_TiO2

    3/12

    and a current range of 2.515 mA was employed as the power

    supply for electrochemical degradation of p-nitrophenol

    (60 mL) in 0.1 M sodium sulphate supporting electrolyte. A

    250 W quartz tungsten-halogen (QTH) lamp equipped with a

    liquid filter was employed as a light source and the light inten-

    sity at the outside surface of the lamp was 0.27 W. The dis-

    tance between the reactor and the light source was 2 cm.

    The effects of the TiO2 loading, the pH of the solution, the

    current density (2.5 mA15 mA), initial concentration(00.8 mM) and applied potential 1 V3 V on the degradation

    efficiency ofp-nitrophenol were investigated.

    Aliquots were withdrawn from the electrochemical cell at

    30 min intervals. The decay of nitrophenol concentration was

    studied using a UV-Vis spectrophotometer (Perkin Elmer

    model Lambda 35). Absorption bands at 310 nm were chosen

    to convert the absorbance to concentration using linear

    regression: y = 8.5x + 0.13667 obtained from a plot of absor-

    bance vs. concentration (00.8 mM).

    2.2 Preparation of supported titania films

    Methanol (about 10 mL), titania P-25 (80 mg), and methyltri-methoxysilane (MTMOS) (100 mg) were mixed in a vial (20 mL)

    and sonicated for 15 min to ensure uniform dispersion of

    titania. EG (100 mg) was subsequently added and the mixture

    was shaken for 5 min. After the addition of 1 M HCl (0.1 mL),

    the mixture was allowed to hydrolyze for 3 h in a closed vessel

    and poured into a Petri dish to dry. The contents of the Petri

    dish were constantly mixed for uniform distribution of titania

    on the EG. After drying for 24 h at room temperature, the float-

    ing catalyst was incubated at 90 C for 24 h to remove residual

    methanol. The percentages of EG and TiO2 in the EGTiO2composite material were 56% and 44% respectively.

    2.3 Fabrication of the EG

    TiO2 electrode

    The prepared composite material was processed into pellets,

    which were then used for the fabrication of the EGTiO2 elec-

    trode using a glass rod, copper wire and conduction silver

    paint (ESI). The clean (oxygen-free) Cu wire was coiled on one

    end to form a flat surface where the EGTiO2 pellets were

    placed. The conduction between the Cu wire and EGTiO2 was

    achieved by using a silver paste. The EGTiO2 was then left to

    air-dry for several minutes. The edges of the EGTiO2 were

    then covered using epoxy resin so that the current contribution

    is only from the basal plane. The copper wire is to transmit

    the electrons to the external circuit since the electrode was

    immersed in the solution. Owing to the compressibility of EGand its ability to interlock, it is possible to fabricate the elec-

    trode into various shapes, thicknesses and sizes.

    3. Results and discussion

    3.1 Characterisation of expanded graphiteTiO2 composite

    material

    3.1.1 BET measurements. The commercially available P-25

    titania was modified using a methyl silicate binder (MTMOS)

    to improve its hydrophobicity and thus avoid detachment of

    the film when wetted. The MTMOS serves as a binder for TiO2,

    both inside and outside the graphite particle. It also interpene-

    trates the catalyst, resulting in high loading of TiO2, and

    increases its rigidity. Table 1 presents the BET surface areas of

    the EG, P-25 titania, unmodified and methylsilane modified

    EGTiO2. The unmodified EGTiO2 showed a high surface area

    and pore volume in comparison with the methyl silane modi-

    fied EGTiO2. The decrease in pore volume is due to interpene-tration of TiO2 into the porous matrix of EG facilitated by the

    MTMOS.

    3.1.2 UV-Vis diffuse reflectance. The effect of methyl sili-

    cate binder on the photocatalytic properties of TiO2 was

    studied and is given in Fig. 1. The characteristic band of TiO2was observed below 390 nm. The methyl silane modified EG

    TiO2 showed high absorbance than the unmodified EGTiO2composites. The high absorbance of methyl silane modified

    EGTiO2 is attributed to the presence of a higher amount of

    TiO2 loaded on graphite particles. The high amount of TiO2results in enhanced photocatalytic efficiency. This result also

    shows that the binder has no negative or quenching eff

    ect onthe absorbance of TiO2. As supported by the BET data, the

    primary role of the binder is to enhance TiO2 loading. Thus

    methyl silane modified EGTiO2 was selected as suitable com-

    posites for degradation ofp-nitrophenol.

    3.1.3 SEM/EDS. The expanded graphite (EG) consists of

    both closed and open pores which provide surface roughness

    and porous cavities for titania incorporation (Fig. 2a). As seen

    Table 1 BET data of EG, TiO2, unmodified and methyl silane modified EGTiO2

    Sample BET surface area/m2 g1 Pore volume/cm3 g1

    EG 14.66 0.06584TiO2 13.89 0.08641EGTiO2 10.36 0.07265EGTiO2 (MTMOS) 4.745 0.02139

    Fig. 1 UV-Vis diffuse reflectance spectra of (a) TiO2, (b) methyl silane modified

    EGTiO2, (c) unmodified EGTiO2 and (d) EG.

    Photochemical & Photobiological Sciences Paper

    This journal is The Royal Society of Chemistry and Owner Societies 2013 Photochem. Photobiol. Sci., 2013, 12, 10911102 | 1093

    View Article Online

    http://dx.doi.org/10.1039/c3pp25398h
  • 7/30/2019 Fotocatalisis Nitrofenol Grafito_TiO2

    4/12

    in Fig. 2b, the TiO2 was well dispersed on the surface of EG

    without clogging the pores. This dispersion allows the possibi-

    lities of both photocatalysis and adsorption on the EGTiO2.

    The presence of Ti K, O K and Si K in the EDX spectrum of the

    composite materials confirms a successful insertion of TiO2on the matrix of EG.

    3.1.4 X-ray diffractometry. Prominent characteristic peaks

    of TiO2 occurring at 2= 25.4 and 27.4 are due to the reflec-

    tions for anatase and rutile phases of the commercial P25respectively as shown in Fig. 3a. Since commercial TiO2-P25 is

    a mixture of rutile (20%) and anatase (80%) crystallite, the

    peaks corresponding to both phases are expected.

    An estimate of the particle size from the broadening of the

    main (101) anatase peak observed at 2= 25.4 can be done by

    using the following Scherrer formula:

    dK

    cos

    where K is the shape factor which is dimensionless and has a

    typical value of about 0.9, is the Cu K radiation wavelength, is the line broadening at half the maximum intensity

    (FWHM) in radians and is the Bragg angle. The particle size

    of the TiO2-P25 was found to be 25 nm.

    Fig. 2 SEM/EDS of (a) EG and (b) EGTiO2 materials.

    Fig. 3 XRD pattern of (a) TiO2 and (b) EGTiO2 materials. R: rutile; A: anatase.

    Paper Photochemical & Photobiological Sciences

    1094 | Photochem. Photobiol. Sci., 2013, 12, 10911102 This journal is The Royal Society of Chemistry and Owner Societies 2013

    View Article Online

    http://dx.doi.org/10.1039/c3pp25398h
  • 7/30/2019 Fotocatalisis Nitrofenol Grafito_TiO2

    5/12

    The XRD pattern of the EGTiO2 composite material con-

    sists of a characteristic peak of EG occurring at 2 = 26.7

    (Fig. 3b). The anatase and rutile peaks, which show the incor-

    poration of TiO2 into the matrix of EG, were also observed on

    the spectrum of the composite. Since the region at which the

    peaks of the composites appear is narrow, the EG peak over-

    laps the peak corresponding to the rutile phase in Fig. 3b.

    However, the intensity of the XRD peaks of the P25 in the com-

    posite materials decreased remarkably. This is inferred to be adilution effect caused by the presence of EG.

    3.1.5 Raman spectroscopy. Raman spectra of the anatase

    phase of TiO2 has six Raman active modes (A1g + 2B1g + 3Eg) at

    147, 198, 398, 515, 640 and 796 cm1 while rutile has four

    active modes (A1g + B1g + B2g + Eg) situated at 144, 448, 612

    and 827 cm1, respectively.43 The TiO2-P25 used in the study is

    a mixture of anatase (80%) and rutile (20%). Thus the Raman

    peaks of titania (Fig. 4a) at 148, 522 and 644 cm1 are assigned

    to the anatase phase while one peak at 400 cm1 corresponds

    to the rutile phase. These signature peaks were also observed

    in the Raman spectrum of EGTiO2 (Fig. 4b) with the G band

    at 1587 cm1

    corresponding to the sp2

    hybridized carbon ofthe EG.

    3.1.6 X-Ray photoelectron spectroscopy. The high resolu-

    tion spectra of the prepared EGTiO2 composite materials and

    the location of binding energies are given in Fig. 5(ad) and

    Table 2.

    The spectrum of EGTiO2 exhibits Ti 2p, O 1s, Si 2p and

    C 1s. The Ti 2p1/2 and Ti 2p3/2 spin orbital splitting photo-

    electrons is located at binding energies of 465.360 and 459.613

    eV respectively as shown in Fig. 5(a). The prominent peak at

    459.61 eV indicates that the Ti element mainly existed as

    Ti(IV).44,45

    The deconvolution of the C 1s spectrum (Fig. 5b) revealed

    the presence of graphitic (sp2) a n d CH from the methyl

    portion of the MTMOS structure at 284.57 and 285.750 eV

    respectively. The curve resolution of the O 1s signal (Fig. 5c)

    indicated the presence of peaks located at 530.37 eV and

    531.13 eV which are assigned to bulk oxide and hydroxyl

    species respectively. The O1s binding energy of the SiOTi

    species observed at 532.39 eV confirms the bonding of the

    MTMOS to the TiO2 surface. The SiOSi, SiOCH3 and Si 2p

    peaks observed at 533.19, 533.39 and 103.75 eV are character-

    istic peaks of the silane binder.

    3.2 Electrochemical characterisation of the EGTiO2

    electrodeThe electrochemical properties of EG and EGTiO2 electrodes

    were studied using 2 mM [Fe(CN)6]3/4 in 0.1 M KNO3 solu-

    tion as a redox probe at a scan rate of 20 mV s1 as shown in

    Fig. 6. The EG electrode exhibited enhanced current in com-

    parison with the EGTiO2 electrode. The electron transfer peak

    to peak separations (Ep) of 117 mV and 170 mV for EG and

    EGTiO2 electrodes respectively were obtained. It is known

    that the closer the Ep to 59 mV for a reversible system, the

    faster the electron transfer rate. Hence the EG electrode exhi-

    bits faster electron transfer kinetics than the EGTiO2 elec-

    trode. The slow electron transfer in the EGTiO2solution

    interface is due to electrochemical inactiveness of the TiO2film which reduces the conductivity of EG.

    The charge transfer resistance of the redox couple at EG

    and EGTiO2 electrodes was studied using electrochemical

    impedance spectroscopy (Fig. 7). The charge transfer resist-

    ances (Rct) of EG and EGTiO2 electrodes in 5 mM [Fe(CN)6]3/4

    were found to be 0.326 k cm2 and 1.47 k cm2 respectively

    using the simple Randles equivalence circuit fitting. The

    lower charge transfer resistance at the EG electrode indicates

    faster electron transfer kinetics than at the EGTiO2. The

    charge transfer rate constant (Kapp) was obtained from the

    Nyquist plot using the following equation:

    Kapp RT=F2

    RctC

    where Kapp is the apparent rate constant at the EG and EG

    TiO2 electrodes, R is the gas constant, T is the temperature in

    Kelvin, F is the Faraday constant, Rct is the charge transfer

    resistance and C is the concentration of the redox couple.46

    Fig. 4 Raman spectra of (a) TiO2 and (b) expanded graphiteTiO2 materials.

    Photochemical & Photobiological Sciences Paper

    This journal is The Royal Society of Chemistry and Owner Societies 2013 Photochem. Photobiol. Sci., 2013, 12, 10911102 | 1095

    View Article Online

    http://dx.doi.org/10.1039/c3pp25398h
  • 7/30/2019 Fotocatalisis Nitrofenol Grafito_TiO2

    6/12

    Fig. 5 High resolution curves of (a) Ti 2p, (b) C 1s, (c) O 1s and (d) Si 2p found in EG TiO2 composite material.

    Table 2 Location of binding energies for the EGTiO2 photoemissions

    Ti B.E./eV O 1s B.E./eV C 1s B.E./eV Si B.E./eV

    2p3/2 465.360 Bulk O2 530.37 Graphitic (sp2) 284.57 Si 2p 103.75

    2p1/2 459.613 OH 531.13 CH 285.75SiOSi 533.19SiOTi 532.39SiOCH3 533.39

    Fig. 6 Cyclic voltammograms of (a) EG and (b) EGTiO2 electrodes in 5 mM [Fe (CN)6]3/4 at 20 mV s1.

    Paper Photochemical & Photobiological Sciences

    1096 | Photochem. Photobiol. Sci., 2013, 12, 10911102 This journal is The Royal Society of Chemistry and Owner Societies 2013

    View Article Online

    http://dx.doi.org/10.1039/c3pp25398h
  • 7/30/2019 Fotocatalisis Nitrofenol Grafito_TiO2

    7/12

    The Kapp of 16.3 105 cm s1 and 3.61 105 cm s1 were

    obtained for EG and EGTiO2 respectively. These kinetic data

    show that the EGTiO2 has a slower kinetics owing to the elec-

    trochemical inactiveness of TiO2 as explained using the cyclic

    voltammetric data.

    3.3 Photoelectrochemical oxidation ofp-nitrophenol

    3.3.1 Degradation kinetics. The Langmuir Hinshelwood

    model is usually employed to describe the kinetics of electro-

    chemical degradation of aquatic organics.43,44 It basically

    relates the degradation rate (r) and reactant concentration in

    water at time t (C), which is expressed as follows:

    r dC

    dt

    krKadC

    1KadC

    where kr is the rate constant and Kad is the adsorption equili-

    brium constant. When the solution is highly diluted and the

    adsorption is relatively weak, KadCbecomes1; the reaction is

    essentially an apparent first-order reaction. The equation can

    then be simplified to the pseudo-first-order kinetics with an

    apparent first-order rate constantKapp:

    lnC0

    Ct krKadt Kappt

    where C0 is the initial concentration (mM) and Ct is the con-

    centration after a period of time. A plot of ln(C0/Ct) versus t

    gives a linear relationship, confirming that the oxidationprocess displays pseudo-first-order kinetic behaviour. The

    apparent kinetic rate constant is obtained from the magnitude

    of the slopes of the straight lines and their corresponding half-

    life values (Table 3).

    The degradation of p-nitrophenol under the three different

    processes (Fig. 8) showed pseudo-first-order kinetic behaviour

    and the kinetic parameters are given in Table 3. Electrochemi-

    cal oxidation exhibited low removal efficiency of the electroche-

    mical process (6%) in comparison with the photocatalytic

    process. This is attributed to the fact that electrochemical

    oxidation involves adsorption of hydroxyl group on the active

    surface of the electrode before the formation of hydroxyl radi-

    cals. These adsorbed hydroxyl groups have a much stronger

    tendency to combine with each other and generate oxygen

    molecules. The hydroxyl radicals responsible for oxidation of

    organic pollutant face the competition from the oxygen evol-

    ution processes and this result in low degradation efficiency.

    These parasitic oxygen molecules can also prevent contact

    between hydroxyl radicals and organic pollutants and this

    result in poor mass transfer of the organic pollutant to the

    electrode surface.

    In photocatalysis, the photogenerated electron is used to

    reduce the oxygen molecules to form superoxide radicalswhich can later form hydroxyl radicals. Therefore, in photoca-

    talysis there are no parasitic oxygen side reactions that can

    affect the degradation efficiency.

    Under the same conditions, the photoelectrochemical oxi-

    dation process resulted in an enhanced degradation efficiency

    ofca. 62% with a fast kinetics rate of 10.4 103 min1 signify-

    ing a synergic combination of electrochemical oxidation and

    photolysis processes. It has been reported that a combination

    of the oxygen released by the electrochemical method can be

    reutilized as the electron acceptor for the oxidation in the pho-

    tolysis process, promoting the degradation rate by indirect oxi-

    dation.47

    In addition to this possibility, the TiO2 used in ourcomposite is capable of generating hydroxyl radicals upon

    Fig. 7 Nyquist plot of (a) EG and (b) EGTiO2 electrodes in 5 mM [Fe(CN)6]3/4.

    Table 3 Degradation kinetic parameters of p-nitrophenol

    ProcessKapp/ 10

    3

    min1t1/2/min

    Removalefficiency/%

    Photolysis 1.56 444 24Electrochemical oxidation 0.72 962 6Photoelectrochemical oxidation 10.4 66.7 62

    Fig. 8 Normalized concentration decay with time of 0.4 mM p-nitrophenol in

    a 0.1 M Na2SO4 solution (pH 7) at a current density of 5 mA cm2 under (a)

    direct photolysis, (b) electrochemical oxidation and (c) photoelectrochemical

    oxidation.

    Photochemical & Photobiological Sciences Paper

    This journal is The Royal Society of Chemistry and Owner Societies 2013 Photochem. Photobiol. Sci., 2013, 12, 10911102 | 1097

    View Article Online

    http://dx.doi.org/10.1039/c3pp25398h
  • 7/30/2019 Fotocatalisis Nitrofenol Grafito_TiO2

    8/12

    irradiation with light of energy higher than its band gap and

    thus enhance the overall efficiency.

    Therefore, the advantage of combining the electrochemical

    oxidation system with photolysis is that the by-product of elec-

    trolysis (newly generated oxygen species) could be utilized to

    elevate the photolysis efficiency and in this way the enhanced

    removal efficiency ofp-nitrophenol was achieved.

    In order to optimize the photoelectrochemical degradation

    of the p-nitrophenol and to further probe the photoelectrocata-lytic process, a series of studies discussed below were

    conducted.

    3.3.2 Optimisation of experimental conditions. The effects

    of TiO2 loading, pH of the solution, current density, initial

    concentration of the p-nitrophenol and applied potential were

    studied and given in Fig. 9.

    Upon increasing the amount of TiO2 content from 80 mg to

    180 mg (Fig. 9a), the degradation rate increased from 7.93

    103 min1 to 10.4 103 min1 respectively.

    The enhanced reaction rate is attributed to the generation

    of more electrons and hole pairs and thus higher concen-

    tration of hydroxyl radicals for degradation of p-nitrophenol.However, the enhancement of degradation efficiency was not

    higher upon increasing the TiO2 content as would be expected.

    This is attributed to the fact that at high TiO2 concentration,

    some TiO2 particles are shielded from the applied UV light

    rendering them inactive.

    The pH values of the reaction solution can affect the

    surface charge of the active species and thus alteration of the

    pH of the solution leads to different interactions between the

    organic pollutant and the electrode depending on the surface

    charge of the electrode material. According to Wang et al.48

    the functional groups from hydrated titanium dioxide may be

    TiO, TiOH and/or TiOH2+. The point of zero charge ( pHpzc) ofthe Degussa P25 TiO2 normally ranges from 6.25 to 6.6.

    49,50 At

    pH values higher than pHpzc, TiO is the predominant group

    on the TiO2 surface.

    Most phenolic compounds act like weak acids and thus dis-

    sociation of hydrogen ions from the phenolic compounds

    depends on the pH of the solution. It is known that p-nitro-

    phenol exists as a nitrophenolate anion when the pH of the

    solution is greater than its pKa value (7.2) and as a molecule

    when the pH of the solution is lower than its pKa value. There-

    fore, the neutral molecular form of p-nitrophenol could be

    adsorbed easily on the surface of TiO2 by a weak electrostatic

    attraction with TiO

    via hydrogen bonding.47

    Thus, at neutralpH, the degradation was more efficient than at low pH as

    shown in Fig. 9b. The low degradation efficiency at low pH is

    attributed to the fact that in a highly acidic medium, there is

    Fig. 9 Dependence of photoelectrochemical degradation of 0.4 mM p-nitrophenol in Na2SO4 at pH 7, 5 mA cm2 and 0.27 W output power on (a) TiO2 loading,

    (b) pH of the solution, (c) initial concentration of p-nitrophenol and (d) current density.

    Paper Photochemical & Photobiological Sciences

    1098 | Photochem. Photobiol. Sci., 2013, 12, 10911102 This journal is The Royal Society of Chemistry and Owner Societies 2013

    View Article Online

    http://dx.doi.org/10.1039/c3pp25398h
  • 7/30/2019 Fotocatalisis Nitrofenol Grafito_TiO2

    9/12

    high production of hydrogen ions (H+) which can compete

    with molecules of p-nitrophenol for adsorption sites on the

    TiO surface, leading to low photoelectrocatalytic degradation

    ofp-nitrophenol.

    The influence of the initial p-nitrophenol concentration on

    the photodegradation of p-nitrophenol is given in Fig. 9c. An

    increase of the initial concentration from 0.2 mM to 0.4 mM

    p-nitrophenol resulted in enhanced degradation efficiency.

    However further increase of the concentration to 8 mM led to

    diminished degradation efficiency because the generated elec-

    tronhole pairs are insufficient to decompose every p-nitro-

    phenol molecule as the number of molecules increases with

    an increase ofp-nitrophenol concentration.Fig. 9d reveals the effect of current density during photo-

    electrochemical oxidation of p-nitrophenol. The degradation

    efficiency increased with an increase of current density from

    2.5 mA cm2 to 5 mA cm2. The increase of degradation

    efficiency is attributed to higher production of hydroxyl radical

    upon an increase of the current density. An increase of current

    density from 5 mA cm2 to 15 mA cm2 did not really increase

    the degradation efficiency as would have been expected. This

    is because at higher current density, the rate of formation of

    intermediates also increases. These rapidly formed intermedi-

    ate products scavenge the in situ generated hydroxyl radicals

    that are also needed by the parent organic pollutant, thusslowing down the degradation kinetics of the entire process

    relatively. Based on the point of energy cost, it is favourable to

    Fig. 10 Dependence of photoelectrochemical degradation of 0.4 mM p-nitro-phenol in Na2SO4 at pH 7 and 0.27 W output power on applied potential for

    the EGTiO2 photoelectrode.

    Fig. 11 UV-Vis spectra during degradation of 0.3 104 M MB at (a) photolysis, (b) electrochemical oxidation (1.5 V), (c) photoelectrochemical oxidation (1.5 V)

    and (d) their normalized concentration decay versus time.

    Photochemical & Photobiological Sciences Paper

    This journal is The Royal Society of Chemistry and Owner Societies 2013 Photochem. Photobiol. Sci., 2013, 12, 10911102 | 1099

    View Article Online

    http://dx.doi.org/10.1039/c3pp25398h
  • 7/30/2019 Fotocatalisis Nitrofenol Grafito_TiO2

    10/12

    be informed that an increase of the current density may not

    necessarily increase the degradation efficiency, and for this

    work, 5 mA cm2 can be said to be the optimum. Furthermore,

    the ability of the EGTiO2 photoanode to withstand a high

    current density suggests that this composite electrode exhibits

    high stability and thus can be considered for real practical

    applications.

    3.3.3 Effect of the electrode potential. The importance of

    selecting a suitable potential for the degradation of 0.4 mMp-nitrophenol solutions was studied by the application of

    different electrode potentials ranging from +1 V to +3 V as

    shown in Fig. 10.

    The increase in biased potential helps in the separation of

    photogenerated electrons and holes, leading to higher concen-

    tration of hydroxyl radicals and thus increased degradation.

    The degradation efficiency increased as a function of biased

    potential. At an applied potential of 3 V, about 62% removal

    efficiency of 0.4 mM p-nitrophenol (60 mL) was achieved

    within 90 min reaction time. In comparison with a report by

    Palmisano et al.,25 where 70% removal efficiency of 0.072 mM

    p-nitrophenol (130 mL) was obtained at 3 V applied potentialfor 6.5 h degradation time using a graphite rodTiO2 anode,

    the EGTiO2 composite electrode used in this study demon-

    strated improved efficiency. The enhanced efficiency of this

    composite electrode can be attributed to the porous matrix of

    EG materials which allows high loading of TiO2. The high

    loading of TiO2 is beneficial for the production of hydroxyl

    radical and thus enhanced efficiency at high concentration of

    p-nitrophenol at a shorter reaction time was achieved. In

    addition, no detachment of TiO2 from expanded graphite was

    observed upon increasing the potential. Therefore, the EG

    TiO2 composite electrode exhibited high stability.

    3.4 Suitability of the EGTiO2 electrode towards degradation

    of other organic pollutants

    To demonstrate the photoelectrochemical degradation versati-

    lity of the EGTiO2 electrode, two model organic dyes methyl-

    ene blue (a cationic dye) and eosin yellowish (an anionic dye)

    were employed as target pollutants as shown in Fig. 11 and 12

    respectively. Since the photoanode is positively charged in

    acidic solutions and negatively charged in alkaline solutions,

    the efficiency of methylene blue (MB) photodegradation is

    expected to increase with pH owing to the electrostatic inter-

    actions between the negative surface of the photoanode and

    the cationic nature of MB. Thus the degradation of MB wascarried out in a highly alkaline (pH 12) medium. The photo-

    electrochemical degradation of MB was compared to photo-

    lysis and electrochemical degradation as shown in Fig. 11(a)(c).

    Fig. 12 UV-Vis spectra during degradation of 0.1 104 M Eosin Y at (a) photolysis, (b) electrochemical oxidation (1.5 V), (c) photoelectrochemical oxidation (1.5

    V) and (d) their normalized concentration decay versus time.

    Paper Photochemical & Photobiological Sciences

    1100 | Photochem. Photobiol. Sci., 2013, 12, 10911102 This journal is The Royal Society of Chemistry and Owner Societies 2013

    View Article Online

    http://dx.doi.org/10.1039/c3pp25398h
  • 7/30/2019 Fotocatalisis Nitrofenol Grafito_TiO2

    11/12

    The absorbance of the maximum absorption band at 680 nm

    decreased with time for the photolysis process reaching 29%

    removal efficiency at a rate constant of 3.08 103 min1. The

    electrochemical oxidation process resulted in an improved

    removal efficiency of 80% and a faster kinetics 11.9 103

    min1 over photolysis. The combination of these two processes

    resulted in the highest degradation efficiency of 94% and the

    most facile kinetic rate of 20.9 103 min1 when compared to

    the individual processes (Fig. 11c). These results demonstratethe possible applicability of the EGTiO2 photoanode in the

    photoelectrochemical degradation of organic dyes such as MB.

    For the degradation of eosin yellowish (Eosin Y), UV-Vis

    reflectance spectra obtained before and after light irradiation

    are observed at 400 nm600 nm absorption band intervals. As

    an anionic dye, Eosin Y interacts more efficiently with the EG

    TiO2 photoanode at low pH, where the surface of the photoa-

    node is positively charged. Thus the degradation of Eosin Y

    was conducted at pH 4. The absorbance of Eosin Y decreased

    with irradiation time using the photolysis process resulting in

    37% removal efficiency with a rate constant of 10.4 102

    min3

    (Fig. 12a) while a resistance to electrochemical oxi-dation was observed owing to a poor removal efficiency of 6%

    with a rate constant of 1.55 102 min3 (Fig. 12b). However,

    the combination of the two processes improved the removal

    efficiency to 47% with a 1.55 102 min3 rate constant under

    the same conditions (Fig. 12c). The improvement of the

    removal efficiency is attributed to the synergistic character of

    the photoelectrochemical process and also supports the wider

    applicability of the EGTiO2 photoanode. The low removal

    efficiency recorded for Eosin Y suggests that the photoanode

    may be more suitable for cationic pollutants than for anionic

    pollutants.

    4. Conclusion

    This work successfully reports the synergic effect of photoelec-

    trochemistry on the degradation of p-nitrophenol in compari-

    son with individual electrochemical oxidation and photolysis

    processes. The removal efficiency of 62% after just 90 minutes

    under the optimum conditions of neutral pH, a relatively mild

    3 V potential and the low current density suggest the suit-

    ability of the prepared EGTiO2 as a photoanode in the photo-

    electrochemical degradation of phenolic compounds, which

    are toxic and relatively resistant to biological degradation. The

    potential of the EGTiO2 photoanode for wider photodegrada-tion applications is evidenced by its ability to degrade two

    other classes of organic pollutants, namely: methyl blue and

    Eosin Y. The low cost of graphite and the ease of preparation

    of the EGTiO2 also lead to possible large-scale applications.

    References

    1 E. Neyens and J. Baeyens, J. Hazard Mater. B, 2003, 98(13),

    3350.

    2 K. C. Namkung, A. E. Burgesse and D. H. Bremner, Environ.

    Technol., 2003, 26, 341352.

    3 O. Legrini, E. Oliveros and A. M. Braun, Chem. Rev., 1993,

    93(2), 671698.

    4 S. Malato, J. Blanco, Ch. Richter and M. I. Maldonado,

    Appl. Catal., B, 2000, 25, 3138.

    5 D. Robert and S. Malato, Sci. Total Environ., 2002, 291, 85

    97.

    6 Y. Sun and J. J. Pignatello, Environ. Sci. Technol., 1993, 27,304310.

    7 S. Gelover, P. Mondragon and A. Jimenez, J. Photochem.

    Photobiol., A, 2004, 165, 241246.

    8 C. H. Comninellis and C. Pulgarin, J. Appl. Electrochem.,

    1991, 21, 703708.

    9 F. Fresno, C. Guillard, J. M. Coronado, S. M. Chovelon,

    D. Tudela, J. Soria and J. M. Herrmann, J. Photochem.

    Photobiol., A, 2005, 173(1), 1320.

    10 K. Vinodgopa and P. V. Kamat, Sol. Energy Mater. Sol. Cells,

    1995, 38, 401410.

    11 O. Carp, C. L. Huisman and A. Reller, Prog. Solid State

    Chem., 2004, 32, 33177.12 A. Fujishima, X. Zhang and D. A. Tryk, Surf. Sci. Rep., 2008,

    63(12), 515582.

    13 C. Hachem, F. Bocquillon, O. Zahraa and M. Bouchy, Dyes

    Pigm., 2001, 49(2), 117125.

    14 J. Rathousk, V. Kalousek, M. Kol and J. Jirkovsk, Photo-

    chem. Photobiol. Sci., 2011, 10, 419424.

    15 T. C. An, W. B. Zhang, X. M. Xiao, G. Y. Sheng, J. M. Fu and

    X. H. Zhu, J. Photochem. Photobiol., A, 2004, 161, 233242.

    16 T. An, Y. Xiong, G. Li, C. Zha and X. Zhu, J. Photochem.

    Photobiol., A, 2002, 152, 155165.

    17 I. M. Butterfield, P. A. Christensen, A. Hamnett, K. E. Shaw,

    G. M. Walker and S. A. Walker, J. Appl. Electrochem., 1997,

    27, 385395.

    18 C. He, Y. Xiong, J. Chen, C. Zha and X. Zhu, J. Photochem.

    Photobiol., A, 2003, 157, 7179.

    19 J. M. Kesselman, N. S. Lewis and M. R. Hoffmann, Environ.

    Sci. Technol., 1997, 31, 22982302.

    20 P. S. Shinde, P. S. Patil, P. N. Bhosale, A. Brger, G. Nauer,

    M. Neumann-Spallart and C. H. Bhosale, Appl. Catal., B,

    2009, 89, 288294.

    21 J. Li, J. Wang, L. Haung and G. Lu, Photochem. Photobiol.

    Sci., 2010, 9, 3946.

    22 S. Yang, X. Quan, X. Li and C. Sun, Photochem. Photobiol.

    Sci., 2006, 5, 808814.

    23 T. Minami, Thin Solid Films, 2008, 516, 13141321.24 J. Hpkes, J. Mller and B. Rech, Transparent Conductive

    Zinc Oxide, ed. K. Ellmer, A. Klein and B. Rech, Springer,

    Berlin, 2008, pp. 359413.

    25 G. Palmisano, V. Loddo, H. H. El Nazer, S. Yurdakal,

    V. Augugliaro, R. Ciriminna and M. Pagliaro, Chem. Eng. J.,

    2009, 155, 339346.

    26 T. Tsumura, N. Kojitani, I. Izumi, N. Iwashita, M. Toyoda

    and M. Inagaki, J. Mater. Chem., 2002, 12, 13911396.

    27 M. Janus, B. Tryba, M. Inagaki and A. W. Morawski, Appl.

    Catal., B, 2004, 52, 6167.

    Photochemical & Photobiological Sciences Paper

    This journal is The Royal Society of Chemistry and Owner Societies 2013 Photochem. Photobiol. Sci., 2013, 12, 10911102 | 1101

    View Article Online

    http://dx.doi.org/10.1039/c3pp25398h
  • 7/30/2019 Fotocatalisis Nitrofenol Grafito_TiO2

    12/12

    28 B. Tryba, A. W. Morwski and M. Ingaki, Appl. Catal., B,

    2003, 46, 203208.

    29 A. Orlov, D. A. Jefferson, N. Macleod and R. M. Lambert,

    Catal. Lett., 2004, 92, 4147.

    30 T. Ohno, T. Tsubota, S. Miyayama and K. Sayama, Catal.

    Lett., 2005, 102, 207210.

    31 Y. Q. Chu, C. A. Ma and Y. H. Zhu, Acta Phys.-Chim. Sin.,

    2004, 20(3), 331335.

    32 Y. H. Wen, G. P. Cao and J. Cheng, J. Inorg. Mater., 2006, 21(2), 441447.

    33 J. H. Li, L. L. Fengand and Z. X. Jia, Mater. Lett., 2006, 60,

    746.

    34 I. M. Afanasovl, D. V. Savchenko, S. G. Ionov,

    D. A. Rusakov, A. N. Seleznevand and V. V. Avdeev, Inorg.

    Mater., 2009, 45, 486.

    35 J. H. Li, Q. Liu and H. F. Da, Mater. Lett., 2007, 61, 1832.

    36 F. Y. Kang, Y. P. Zheng, H. N. Wang, Y. Nishi and

    M. Inagaki, New Carbon Mater., 2002, 40, 1575.

    37 M. B. Dowell and R. A. Howard, Carbon, 1986, 24, 311.

    38 D. D. L. Chung, J. Mater. Sci., 1987, 22, 4190.

    39 E. P. Gilbert, P. A. Reynolds and J. W. White, J. Chem. Soc.,Faraday Trans., 1998, 94, 1861.

    40 T. Tsumura, N. Kojitania, H. Umemur, M. Toyod and

    M. Inagaki, Appl. Surf. Sci., 2002, 196, 429436.

    41 A. Modestov, V. Glezer, I. Marjasin and O. Lev, J. Phys.

    Chem. B, 1997, 101, 46234629.

    42 K. Ramanathan, D. Avnir, A. Modestov and O. Lev, Chem.

    Mater., 1997, 9, 25332540.

    43 J. Zhang, M. Li, Z. Feng, J. Chen and C. Li, J. Phys. Chem. B,

    2006, 110, 927935.

    44 H. Liu, W. Yang, Y. Ma, Y. Cao, J. Yao, J. Zhang and T. Hu,Langmuir, 2003, 19, 30013005.

    45 W. Que, Y. Zhou, Y. L. Lam, Y. C. Chan and C. H. Kam,

    Appl. Phys. A, 2001, 73, 171176.

    46 V. Ganesh, S. K. Pal, S. Kumar and V. Lakshminarayanan,

    J. Colloid Interface Sci., 2006, 296, 195203.

    47 L. Wei, H. Zhu, X. Mao and F. Gan, Sep. Purif. Technol.,

    2011, 77, 1825.

    48 K. H. Wang, Y. H. Hsieh and L. J. Chen, J. Hazard. Mater.,

    1998, 59, 251260.

    49 C. Chiou, C. Wu and R. Juang, Sep. Purif. Technol., 2008,

    62, 559564.

    50 K. H. Wang, Y. H. Hsieh, C. H. Wu and C. Y. Chang, Chemo-sphere, 2000, 40, 389394.

    Paper Photochemical & Photobiological Sciences

    1102 | Photochem. Photobiol. Sci., 2013, 12, 10911102 This journal is The Royal Society of Chemistry and Owner Societies 2013

    View Article Online

    http://dx.doi.org/10.1039/c3pp25398h