effects of titanate nanotubes synthesized by a microwave hydrothermal method on photocatalytic...

Upload: christian-salas-suarez

Post on 19-Oct-2015

23 views

Category:

Documents


0 download

TRANSCRIPT

  • Accepted Manuscript

    Title: Effects of Titanate Nanotubes Synthesized by a Microwave HydrothermalMethod on Photocatalytic Decomposition of Perfluorooctanoic Acid

    Authors: Ying-Chu Chen, Shang-Lien Lo, Jeff Kuo

    PII: S0043-1354(11)00283-1

    DOI: 10.1016/j.watres.2011.05.020

    Reference: WR 8608

    To appear in: Water Research

    Received Date: 13 December 2010

    Revised Date: 7 May 2011

    Accepted Date: 21 May 2011

    Please cite this article as: Chen, Y.-C., Lo, S.-L., Kuo, J. Effects of Titanate Nanotubes Synthesized bya Microwave Hydrothermal Method on Photocatalytic Decomposition of Perfluorooctanoic Acid, WaterResearch (2011), doi: 10.1016/j.watres.2011.05.020

    This is a PDF file of an unedited manuscript that has been accepted for publication. As a service toour customers we are providing this early version of the manuscript. The manuscript will undergocopyediting, typesetting, and review of the resulting proof before it is published in its final form. Pleasenote that during the production process errors may be discovered which could affect the content, and alllegal disclaimers that apply to the journal pertain.

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    1

    Effects of Titanate Nanotubes Synthesized by a Microwave Hydrothermal 1

    Method on Photocatalytic Decomposition of Perfluorooctanoic Acid 2

    3

    Ying-Chu Chena,b, Shang-Lien Loa,*, and Jeff Kuoc 4

    5

    aGraduate Institute of Environmental Engineering, National Taiwan University, Taipei, 6

    106, Taiwan, 7

    bDepertment of Environmental Monitoring & Information Management, 8

    Environmental Protection Agency, Taiwan (R.O.C.), 9 cDepartment of Civil and Environmental Engineering, California State University, 10

    Fullerton, CA 11

    12

    *Corresponding author: Shang-Lien Lo 13

    14

    E-mail address: [email protected] 15

    Phone: +886-2-2362-5373 16

    Fax: +886-2-2392-8821 17

    Mailing address: Graduate Institute of Environmental Engineering, National Taiwan 18

    University, 71, Chou-Shan Rd., Taipei 106, Taiwan (R.O.C.) 19

    20

    Abstract 21

    Titanate nanotubes (TNTs) were used to remove perfluorooctanoic acid (PFOA) 22

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    2

    from aqueous solutions in this study. Direct photolysis of PFOA by a 254-nm UV light 23

    (400 W) was found effective to decompose PFOA without presence of photocatalysts. 24

    Shorter-chain perfluorocarboxylic acids (PFCAs) and fluoride ions were formed 25

    during photodecomposition. Addition of TNTs as photocatalysts did not greatly 26

    enhance photocatalytic decomposition of PFOA. TNTs mainly act as adsorbents to 27

    adsorb PFOA and form TNT-PFOA complexes. It suggested that sodium ions and 28

    oxygen atoms on the surfaces of TNTs play important roles in PFOA adsorption. X-ray 29

    Photoelectron Spectroscopy (XPS) and Fourier Transform Infrared Spectroscopy 30

    (FTIR) analyses indicated that ion-exchange, electrostatic interaction, and 31

    hydrophobic interaction all participated in the photocatalytic reaction of PFOA by 32

    TNTs. 33

    34

    Keyword: Microwave; Perfluorooctanoic acid; Photocatalysis; Titanate nanotube; 35

    X-ray Photoelectron Spectroscopy 36

    37

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    3

    1. Introduction 1

    Perfluorocarboxylic acids (CnF2n+1COOH, PFCAs) and their derivatives are 2

    anthropogenic organic compounds that have a wide range of applications. PFCAs can 3

    be used as industrial surfactants, additives, firefighting foams, and lubricants (Lin et al., 4

    2010a). Some perfluorinated acids, especially perfluorooctanoic acid (PFOA, 5

    C7H15COOH), have garnered concerns because of their ubiquitous usage, persistence, 6

    and bioaccumulation in the environment. PFOA is exceptionally inert and persists 7

    indefinitely in the environment (Musijowski et al., 2007), even in remote polar areas 8

    (Qu et al., 2010). Some evidences show that PFOA can accumulate in organisms and 9

    pose a potential human health risk (Hinderliter et al., 2006; Potera, 2009). The Science 10

    Advisory Board of the United States Environmental Protection Agency (U.S. EPA) 11

    reported that PFOA is likely a carcinogen (Hogue, 2006). Furthermore, U.S. EPA 12

    launched a program to reduce 95 % PFOA emissions from manufacturers by 2010 and 13

    to eliminate the use of the chemical by 2015 (Qu et al., 2009). Development of 14

    effective treatment methods to convert PFOA into harmless species is desirable. 15

    Due to its inherent resistance to chemical and microbiological treatment, many 16

    technologies have been developed to decompose PFOA, including advanced oxidation 17

    processes (AOPs), such as UV-H2O2, persulfate, photo-Fenton, ozonation and 18

    sonolysis (Cheng et al., 2008; Hori et al., 2004; Hori et al., 2008b; Moriwaki et al., 19

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    4

    2005; Vicitis et al., 2008). Direct photolysis (Cao et al., 2010; Chen and Zhang, 2006; 20

    Chen et al., 2007; Hori et al., 2004) and photocatalysis (Qu et al., 2010) in aqueous 21

    solutions are alternatives that are being researched in recent years. PFOA is sensitive to 22

    light with wavelengths from deep UV-region to 200 nm (Cao et al., 2010), thus can be 23

    efficiently decomposed by direct UV irradiation. TiO2-mediated heterogeneous 24

    photocatalysis of PFOA was found to involve generation of electron/hole pairs that led 25

    to better photocatalytic efficiencies (Dillert et al., 2008; Doll and Frimmel, 2005; 26

    Estrellan et al., 2010; Panchangam et al., 2009; Ravichandran et al., 2006; 27

    Ravichandran et al., 2009). The -Gallium oxide (Zhao and Zhang, 2009) and ZnO 28

    (Ravichandran et al., 2007) were also used as photocatalysts to assist 29

    photodecomposition of PFOA. Decomposition of PFOA into fluoride ions, which can 30

    readily react with Ca2+ to form CaF2, a raw material in global demand for production of 31

    hydrofluoric acid (Hori et al., 2009; Hori et al., 2010; Ravichandran et al., 2006). 32

    In addition to direct photolysis and photocatalysis, adsorption has been 33

    demonstrated to be an effective and economical method to remove polar organic 34

    pollutants from aqueous solutions (Senevirathna et al., 2010; Yu et al., 2009), but only 35

    a few papers were found in our literature survey that studied the removal of PFOA by 36

    adsorbents such as active carbon (Lampert et al., 2007; Ochoa-Herrera et al., 2008; 37

    Senevirathna et al., 2010; Yu et al., 2009; Qu et al., 2009), polymers (Senevirathna et 38

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    5

    al., 2010), zeolite and sludge (Ochoa-Herrera et al., 2008). PFOA is moderately 39

    adsorbable with active carbon, and the adsorption efficiency is greatly influenced by 40

    particle sizes of adsorbents. The powdered activated carbon (PAC) achieved an 41

    adsorption equilibrium within 4 h (Senevirathna et al., 2010; Yu et al., 2009), while 42

    granular activated carbon (GAC) required over 168 h to reach the equilibrium (Yu et al., 43

    2009). 44

    Titanate nanotubes (TNTs) synthesized by microwave hydrothermal (M-H) 45

    methods have some desired properties, such as high specific surface area, 46

    photocatalytic properties, and ion-exchangeable capabilities (Ou and Lo, 2007). In our 47

    previous study, the specific surface area of microwave-assisted TNTs was 150 m2 g-1 48

    and could adsorb 2,000 mg Pb(II) per gram of TNTs at pH of 4 (Chen et al., 2010a). 49

    Compositing TNTs with semiconductors, such as cadmium sulfide (CdS), can enhance 50

    their photocatalytic capability to eliminate 52.3 % of ammonia in water (Chen et al., 51

    2010b). TNTs synthesized by M-H methods have great potentials to be photcatalysts 52

    and adsorbents in various applications. 53

    This study aimed at investigating adsorption and photocatalytic behaviors of TNTs 54

    synthesized by a M-H method for their removal of PFOA from aqueous solutions. The 55

    sorption isotherms were developed and effects of solution pH were evaluated; while 56

    effects of UV light, loading concentrations, and solution pH on photocatalysis were 57

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    6

    also studied. A detailed reaction mechanism was proposed based on the experimental 58

    results and the instrumental (X-ray Photoelectron Spectroscopy (XPS) and Fourier 59

    Transform Infrared Spectroscopy (FTIR)) analyses. Findings of this study provide 60

    insights of using TNTs to remediate environmental media that are contaminated by 61

    PFOA. 62

    63

    2. Materials and methods 64

    2.1. Materials 65

    All chemicals used in the experiments were of reagent grade. Perfluorooctanoic 66

    acid (PFOA, C7F15COOH, 96 % purity) was from Aldrich (USA). Perfluoroheptanoic 67

    acid (PFHpA, C6F13COOH, 98 % purity), perfluoropentanoic acid (PFPeA, 68

    C4F9COOH, 97 % purity), and heptafluorobutyric acid (HFBA, C3F7COOH, 99 % 69

    purity) were from Alfa Aesar (USA). Perfluorohexanoic acid (PFHxA, C5F11COOH, 70

    97 % purity) was purchased from Fluka (USA). Methyl alcohol anhydrous (CH3OH, 71

    99.9 % purity, Mallinckrodt Chemicals, USA), boric acid (H3BO3, 99.5 % purity, 72

    Nacalai Tesque Inc., Japan), and sodium hydroxide (NaOH, 99.7 % purity, Alps 73

    Chem Co. Ltd., Taiwan) were used to prepare mobile phases for high performance 74

    liquid chromatography (HPLC) analyses. Sulfuric acid (H2SO4, 97 % purity, Nacalai 75

    Tesque Inc., Japan) was used to regenerate the suppressor column of HPLC. Also, 76

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    7

    sodium bicarbonate (NaHCO3, 99.6 % purity, Nacalai Tesque Inc., Japan), sodium 77

    carbonate (Na2CO3, 99.0 % purity, Nacalai Tesque Inc., Japan), and H2SO4 were used 78

    to prepare the mobile phase for ion chromatography. 79

    Titanium dioxide (TiO2 Degussa P25, 99.5 % purity, Simakyu Pure Chemicals, 80

    Japan), NaOH, and hydrochloric acid (HCl, 70 % purity, Fisher Scientific, USA) 81

    were used to fabricate TNTs. Nitric acid (HNO3, 60.0 % purity, Yakuri Pure 82

    Chemicals Co. Ltd., Japan) and NaOH were used for pH adjustment. All solutions 83

    were prepared with Milli-Q ultrapure water (18 M-cm resistivity). 84

    2.2. Synthesis of TNTs 85

    Titanate nanotubes were synthesized by the M-H method described elsewhere 86

    (Chen et al., 2010a). The microwave digestion system (Ethos Touch Control, 87

    Milestone Corporation, Italy) consists of a double-walled vessel with an inner Teflon 88

    liner and an outer shell of high strength ULTEMpolyetherimide. The Teflon liner 89

    can resist reaction conditions employed in this study and no cross-contamination from 90

    the Teflon liner was observed. Briefly, a mixture of 600 mg TiO2 and 70 mL NaOH (10 91

    N) was stirred for 40 min in a Teflon container which was then sealed before the 92

    microwave heating process. The container was then placed into the microwave 93

    digestion system under 400-W irradiation at 403 K for 3 h. After the treatment, the 94

    solution was washed three times with 0.5 N HCl and four times with deionized water, 95

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    8

    followed by centrifugation at 800 rpm for 5 min (Kubota 6800, Kubota Co., Japan), 96

    and vacuum dried at 214 K for 24 h (FD3-12P, Kingmech Co., Ltd., Taiwan). 97

    2.3. Batch experiments 98

    All the experiments were carried out in polypropylene (PP) bottles to avoid 99

    potential interferences from sample containers. The experimental apparatus for 100

    photocatalysis contains a cooling water jacket to maintain the temperature at 298 K 101

    (B204, Firstek Scientific Co., Ltd., Taiwan) and a UV lamp (254 nm, 400 W, Philips, 102

    Holland). Known amounts of photocatalysts (TNTs or TiO2) were dispersed into 2-liter 103

    PFOA solutions (50 mg L-1). During UV irradiation, 20 mL aliquots were taken at 0, 1, 104

    2, 3, 6, 9, 24 h, and photocatalysts were immediately removed by filtration through 105

    0.22-m filters (MillexGS, Millipore, Ireland) for subsequent analyses. The 106

    experiments were carried out twice and mean values were reported here. 107

    Adsorption experiments were performed under ambient conditions. Analytical 108

    grade PFOA was used to prepare a stock solution of 1,000 mg L-1, which was further 109

    diluted with deionized water to different concentrations for adsorption experiments. 110

    Adsorption isotherm studies were conducted with initial PFOA concentration ranging 111

    from 5 100 mg L-1 together with 20 mg TNTs or TiO2 in 100 mL solutions. Solutions 112

    were then shaken at 298 K for 24 h at a speed of 150 rpm in a reciprocal shaker bath 113

    (BT-350, Bersing Co., Ltd., Taiwan). The pH values of initial solutions were adjusted 114

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    9

    by using HNO3 or NaOH to reach the desired values. Samples were taken from each 115

    solution and then analyzed by HPLC. All experiments were conducted twice and 116

    average values were used in the data analyses. 117

    2.4. Analytical Methods 118

    Concentrations of PFOA and shorter-chain PFCAs were analyzed by HPLC 119

    (Dionex, UltiMate 3000, USA) coupled with a conductivity detector (ED-50, Dionex, 120

    USA) and an anion self-regenerating suppressor (ASRS 300 2-mm, USA). The 121

    conductivity detector was used in conjunction with a Dionex micromembrane 122

    suppressor column. The suppressor column was constantly regenerated using 0.5 mM 123

    H2SO4 at a flow rate of 1 mL min-1. The PFCAs were extracted by a 150 2.1 mm, 124

    3.5m column (AcclainPolar AdvantageII C18, Dionex, USA) maintained at 303 K. 125

    A trinary gradient was employed with mobile phase A containing 100 % methanol, 126

    mobile phase B containing Milli-Q water, and mobile phase C containing 9 mM NaOH 127

    and 100 mM H3BO3 in deionized water. The flow rate of the solvents was 0.3 mL min-1. 128

    The gradient was operated at 20 % phase A, 40 % phase B for the initial 5 min; 20 % 129

    phase A increased to 60 % and 40 % phase B decreased to 0 % for the next 10 min (5 130

    15 min); Maintaining 60 % phase A, 0 % phase B during 15 20 min; 60 % phase A 131

    decreased to 20 %, 0 % phase B increased to 40% for final 5 min. The phase C was 132

    maintained at 40 % during the total running time, 25 min. All calibration curves for 133

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    10

    PFCAs were linear over the 0.5 50 mg L-1 range. The limits of detection (LODs) 134

    using 50 L samples, based on a signal-to-noise (S/N) ratio of 3, were 1 mg L-1 for 135

    PFCAs. Degradation ratios were calculated as follows: 136

    1000

    0

    =

    CCCR % (1) 137

    where C is concentration of PFOA (mM) and 0C is initial concentration of PFOA 138

    (mM). 139

    Concentrations of aqueous fluoride ions were determined by an ion 140

    chromatography (Metrohm 790 Personal IC, Metrohm Ltd., Switzerland) with a 141

    column of CH-9101 Herisan (Metrohm Ltd., Switzerland). A mixture of 1.8 M 142

    Na2CO3/1.7 M NaHCO3 (1:1) was used as the mobile phase at a flow rate of 1.0 mL 143

    min-1. Defluorination ratios were calculated as follows: 144

    100150

    =

    CC

    R FF % (2) 145

    Where FC is concentration of fluoride ions (mM), 0C is initial concentration of 146

    PFOA (mM), and the factor of 15 corresponds to the number of fluorine atoms in one 147

    PFOA molecule. 148

    2.5. Characterization Methods for TNTs 149

    The TNTs used in this study were fabricated in our research lab. The physchemcial 150

    properties, such as BET surface area and zeta potential, of TNTs synthesized by the 151

    M-H method can be found in Chen et al. (2010a). Surface functional groups of TNTs 152

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    11

    were characterized by FTIR and XPS. The IR scanning range was 4000 650 cm-1 153

    with 4 cm-1 resolution using a mercury-cadmium-telluride detector in a Nexus 470 154

    spectrometer (Thermo Nicolet, USA). KBr powder was used to record the reference 155

    spectrum. Chemical bonding of PFCAs onto surfaces of TNTs/TiO2 was investigated 156

    by XPS using an ESCA PHI 1600 (Physical Electronics, USA). The XPS spectra were 157

    recorded with a monochromatized Mg (K) source of 1253.6 eV energy (15 kV, 400 158

    W). The bonding energy scale for final calibration was corrected by the C1s peak of 159

    284.6 eV. 160

    161

    3. Results and discussion 162

    3.1. Direct photolysis 163

    Aqueous PFOA solutions (50 mg L-1) were irradiated with a 254-nm UV light (400 164

    W), and the extents of photolysis were shown in Fig. 1(a). Control tests (without UV 165

    irradiation) were carried out and no significant changes in PFOA concentration were 166

    observed. The degradation rate was the fastest at pH 4, followed by pH 7, and then pH 167

    10; 98% of initial PFOA was photodecomposed at pH 4 after 48-h irradiation. It was 168

    reported that PFOA has a weak absorption of UV light longer than 200 nm, and only 169

    9% of initial PFOA was decomposed with a 254-nm UV light after 2-h irradiation (23 170

    W low-pressure mercury lamp) (Cao et al., 2010), and 89.5% of initial PFOA was 171

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    12

    decomposed by a xenon-mercury lamp after 72-h irradiation (wavelengths were 172

    mainly in the range of 220 460 nm) (Hori et al., 2004). It implies that irradiation 173

    intensity of lights plays an important role in affecting photolytic efficiency of PFOA. 174

    PFOA can be efficiently photodecomposed under a light emitting higher irradiation 175

    intensity than that with weaker irradiation intensity. Photodecomposition of PFOA and 176

    formation of shorter-chain PFCAs, bearing C6-C7 perfluoroalkyl groups were 177

    quantified by HPLC. The C5-C4 perfluoroalkyl groups were hardly detected within 178

    24-h irradiation, while C3-C2 perfluoroalkyl groups were known to be easily 179

    evaporated under ambient conditions. At pH 4, the concentration of PFHpA increased 180

    with irradiation up to 24 h and then decreased, concentration of PFHxA increased up to 181

    36 h and then decreased, while fluoride concentrations increased continuously with 182

    extended irradiation time (Fig. 1(b)). It indicates that photodecomposition of PFOA 183

    was achieved by a step-by-step removal of the CF2 units. Besides irradiating with a 184

    254-nm UV light, photolysis of PFOA was also effective in VUV system (185 nm) 185

    (Cao et al., 2010; Chen and Zhang, 2006; Chen et al., 2007). Therefore, direct 186

    photolysis of PFOA was mainly depended on irradiation intensity of lights when the 187

    wavelength were the same; otherwise, longer wavelength of lights needs higher 188

    irradiation intensity to assist photolytic efficiency of PFOA without catalysts. 189

    After 24 h of irradiation, defluorination of PFOA at pH 4, 7, and 10 were 85 %, 68 190

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    13

    %, and 55 %, respectively. The acid condition (pH 4) was found more favorable for 191

    photodegradation (98 %) and defluorination (85 %) of PFOA. Although initial pH 192

    value of solutions were adjusted by HNO3 and/or NaOH to reach the desired values, 193

    pH value of solutions would gradually decrease during 24-h UV irradiation and finally 194

    reached the equilibrium at pH 4, which is the natural pH of 50 mg L-1 PFOA under 195

    ambient conditions. HNO3 has been demonstrated to have minor effects on degradation 196

    and defluorination of PFOA (Ravichandran et al., 2006), whereas presume of sodium 197

    ions may have retardation effects on defluorination and decomposition of PFOA since 198

    dissolved metal ions were found to affect photocatalytic efficiencies of PFCAs 199

    (Ravichandran et al., 2006). Therefore, using NaOH to increase initial pH value of 200

    solutions to 7 or 10 might not be favorable in photodecomposition of PFOA. Released 201

    fluoride ions can bind with sodium ions to form sodium fluoride (NaF), causing a 202

    lower defluorination at pH 7 and 10. In addition, more anionic fluoride ions released 203

    into solutions would also decrease pH value of solutions during photolysis. Solution 204

    pH can be an indicator to observe if experiments reached the equilibrium when it 205

    becomes steady. Consequently, this study suggests not to adjust pH value of the 206

    solutions and maintain it at pH 4 (the highest photolytic efficiency) in the subsequent 207

    experiments. 208

    3.2. Adsorption of PFOA onto TNTs 209

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    14

    Adsorption isotherms of PFOA onto TNT or TiO2 are shown in Fig. 2. TiO2 has a 210

    much lower adsorption capacity for PFOA than TNTs at pH 4; it is as expected be due 211

    to the fact that TNTs have a surface area that is three times larger, 140 v.s. 50 m2 g-1 212

    (Chen et al., 2010a). The length of PFOA molecules is 1 nm (Yu et al., 2009), thus 213

    PFOA can easily diffuse into inner pores of TNTs (pore diameters of TNTs were 9 nm 214

    (Chen et al., 2010a)), and adsorb onto inner as well as outside walls of TNTs. 215

    Fig. 2 also shows the extents of adsorption decrease with an increase in solution pH. 216

    Due to the low pKa of PFOA, 2.5 (US EPA, 2002), more PFOA will be in ionic form as 217

    solution pH increases. Electrostatic attraction between negatively charged PFOA and 218

    positively charged TNT surfaces would promote surface reactions with valence band 219

    holes (h+), which are favorable in PFOA photodecomposition (Bahnemann et al., 1997; 220

    Estrellan et al., 2010). Higher adsorption efficiency achieved at lower solution pH was 221

    due to electrostatic forces of attraction (Panchangam et al., 2009). Besides electrostatic 222

    interaction, hydrophobic interaction was also important for adsorption of PFOA 223

    (Higgins et al., 2005; Higgins and Luthy, 2006; Yu et al., 2009). Hydrophobic PFOA 224

    and TNTs had a hydrophobic interaction during the adsorption process (Yu et al., 2009). 225

    The formation of PFOA-TNTs complexes could enhance oxidative power of TNTs 226

    (Dillert et al., 2008). 227

    3.3. Photocatalysis induced by TNTs 228

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    15

    First-order kinetics were used to simulate photodecomposition of PFOA at pH 4 by 229

    a 254-nm UV light, and the results were shown in Fig. 3. The pseudo first-order rate 230

    constants were calculated to be 0.0230 h-1, 0.0337 h-1 and 0.0271 h-1, for 0.05, 0.125 231

    and 0.25 g L-1 TNTs, respectively. The remaining PFOA concentration at 24-h 232

    photocatalytic reactions decreased as amounts of TNTs increased from 0.05 to 0.125 g 233

    L-1; when amounts of TNTs further increased 0.25 g L-1, the remaining PFOA 234

    concentration started to increase. It is apparent that amounts of TNTs have great 235

    influences on photodecomposition of PFOA and a suggested loading amount was 236

    0.125 g L-1. An adequate loading concentration increases the generation rate of 237

    electron/hole pairs that promote photodecomposition of PFOA, while excessive 238

    amounts of TNTs in solutions would decrease the light penetration (Zhu et al., 2005). 239

    The superior photocatalytic performance of TNTs than TiO2 might be due to the larger 240

    surface area of TNTs to receive more UV irradiation to excite PFOA adsorbed on the 241

    surfaces than TiO2 particles (Estrellan et al., 2010; Panchangam et al., 2009). 242

    Consequently, TNTs synthesized by raw materials of TiO2 can enhance their 243

    photocatalytic efficiency to decompose PFOA. 244

    Concentrations of remaining PFOA gradually decreased while those of fluoride 245

    ions increased as the photocatalytic reaction proceeded; however, the reaction rates 246

    with the presence of TNTs were much slower than that of direct photolysis. In the 247

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    16

    presence of 0.125 g L-1 TNTs, only 59% of the initial PFOA was degraded, whereas the 248

    corresponding yield for direct photolysis was almost 100% after 24-h irradiation. Thus, 249

    an addition TNTs or TiO2 as photocatalysts would retard photodecomposition of PFOA. 250

    This phenomenon can be explained in terms of heterogeneous reactions. TNTs, which 251

    are intended as photocatalysts, act as adsorbents to adsorb PFOA before mineralization 252

    occurs. Photocatalysts were also found to scatter UV lights (Zhu et al., 2005) and 253

    suppress formation of shorter-chain PFCAs during photocatalytic reactions (Hori et al., 254

    2004). Generating active hydroxyl radicals ( OH ) in TiO2 system has poor 255

    photoreactivity for PFCAs (Cao et al., 2010; Hori et al., 2004). Slower degradation 256

    rates of PFOA in a TNT system may also be attributed to release sodium ions, that 257

    were reported to affect photodecomposition of PFCAs. 258

    Multi-walled TNTs were synthesized by scrolling up four layers of TiO6 259

    octahedrons (Scheme S1 in the online Supplementary Material) (Chen et al., 2002), 260

    thus the inner surface area of TNTs is much larger than outside walls. Researchers 261

    found that the non-planar structure provided a scattering center for incident light, thus 262

    elongated the path to promote the photocatalytic capability of photocatalysts (Lin et al., 263

    2010b). Outer surfaces of TNTs could harvest the lights from any directions in the 264

    surrounding that could be helpful for increasing the photocatalytic efficiency; therefore, 265

    when PFOA diffused into pores and adsorbed onto inner walls of TNTs, its opportunity 266

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    17

    to receive UV lights was greatly reduced. 267

    Defluorination ratios of PFOA were much smaller than the PFOA decomposition 268

    ratios. It implies that intermediate products formed during photocatalysis (Chen and 269

    Zhang, 2006). Fluoride ions can adsorb onto surfaces of photocatalysts, leading to less 270

    available surface active sites and smaller defluorination ratios (Estrellan et al., 2010; 271

    Panchangam et al., 2009). TNTs have a high affinity to adsorb fluoride ions (Fig. A1 in 272

    the online Supplementary Material). Positive charges on TNT surfaces under acidic 273

    conditions (pH 4) cause more electrostatic attractions of fluoride ions. From the above 274

    results, it can be deduced that smaller defluorination ratios of photocatalytic reactions 275

    was caused by adsorption of fluoride ions onto TNT surfaces since fluoride ions were 276

    continuously released during photodecomposition of PFOA; the other possibility was 277

    that PFOA was directly adsorbed onto inner surfaces of TNTs without mineralization. 278

    3.4. FTIR and XPS analyses of PFOA-TNTs/TiO2 279

    Figure 4 shows XPS results of TNT or TiO2 before and after photocatalytic 280

    reactions. The O1s, C1s, F1s, Na1s, and Ti2p3 peaks were investigated in XPS 281

    wide-scans (Fig. 4(a)). Intensities O1s peaks at 532 eV were obviously decreased after 282

    photocatalytic reactions (Fig. 4(b)). It reveals that oxygen-terminated functional 283

    groups on TNT surfaces react with PFOA during photocatalytic reactions that led to a 284

    decrease in oxygen concentrations after photocatalytic reactions (Guan et al., 2007). 285

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    18

    The exciting peaks at 287 eV and 292 eV corresponded to bulk carbon signals (Fig. 286

    4(c)) (Guan et al., 2007). Enhanced intensities of C1s peaks suggest that PFOA and 287

    shorter-chain PFCAs were directly adsorbed onto TNTs/TiO2 surfaces. If 288

    mineralization of PFOA continued, peaks corresponded to fluoride ions should appear 289

    at 689 eV (Fig. 4(d)) (Guan et al., 2007; Hori et al., 2006; Hori et al., 2008a); however, 290

    no peaks corresponded to fluoride species were found. Most of PFOA molecules were 291

    directly adsorbed onto surfaces of TNTs without mineralization during UV irradiation 292

    since the adsorption rate of TNTs is very fast (Chen et al., 2010a). Intensities of Na1s 293

    peaks apparently decreased after photocatalysis (Fig. 4(e)). These results support the 294

    previous ones that TNTs mainly act as adsorbents to adsorb PFOAs rather than as 295

    photocatalysts. The adsorption mechanism of TNTs is that sodium ions on TNT 296

    surfaces react with PFOA and less sodium ions would be detected on the surfaces of 297

    TNTs. No sodium ions were detected while the reactions were photocatalyzed by TiO2. 298

    The binding energies for Ti(IV) (467 eV) and Ti(III) (461 eV) on TNTs or TiO2 were 299

    similar after photocatalytic reactions (Fig. 4(f)). The XPS spectra support the proposed 300

    photocatalytic mechanisms that PFOA firstly adsorbed onto surfaces of TNTs, and the 301

    remaining PFOA in solutions or PFOA adsorbed on the outside walls of TNTs are 302

    mineralized by UV irradiation and formed shorter-chain PFCAs. 303

    Figure 5 shows FTIR spectra of TNTs/TiO2 photocatalysts and PFOA- TNTs/TiO2 304

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    19

    complexes. The band at 1138 cm-1 was attributed to C-F bonds stretching vibrations 305

    in OOCC7F15 groups on the surfaces of PFOA-TNTs/TiO2 complexes (Kutsuna and 306

    Hori, 2008; Men et al., 2008; Xu et al., 2009). The bands at 1438 cm-1 and 1647 cm-1 307

    were attributed to COO asymmetrical stretching vibrations and symmetrical 308

    stretching vibrations, respectively (Xu et al., 2009). The wide band at 3400 cm-1 region 309

    was attributed to presence of hydroxyl groups ( OH ) and traces of water in the KBr 310

    pellet (Men et al., 2008). Intensities of hydroxyl groups ( OH ) after photocatalytic 311

    reactions indicated that catalysts were excited by UV lights to form hydroxyl radicals 312

    ( OH ), which can degrade PFOA and prevent rapid recombination with electron/hole 313

    pairs (Estrellan et al., 2010). However, intensities of hydroxyl groups ( OH ) after 314

    photocatalytic reaction by TiO2 were ascribed to be the occurrence of oxygen 315

    vacancies on the surfaces, which were also found to be beneficial to its photocatalytic 316

    activity (Ismail et al., 2007). The above FTIR and XPS results indicated that the 317

    removal of PFOA mainly occurred with the functional groups on the surface of 318

    photocatalysts. 319

    3.5. Mechanisms of PFOA decomposition 320

    Based on our experimental results, it can propose the photodecomposition 321

    mechanisms of PFOA by TNTs under UV irradiation (Fig. 6). PFOA exists as an 322

    anionic compound when emerging in solutions (eq. (3)) (Burns et al., 2008). 323

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    20

    +++ OHCOOFCOHCOOHFC 3-

    1572157 (3) 324

    Positive charges on TNT surfaces are favorable to adsorb anionic PFOA with or 325

    without UV irradiation (eq. (4)). 326

    COOFCTNTsCOOFCTNTsTNTsTNTs h 157157)( + + (4) 327

    PFOA adsorbed on the outer surfaces of TNTs and the remaining PFOA in solutions 328

    were excited by UV irradiation to form excited state of PFOA and photolyzed into 329

    157FC and COOH radicals (eq. (5)). 330

    + COOHFCCOOHFCCOOHFC hh 157157157

    (5) 331

    The 157FC radical continue adsorb onto surfaces of excited TNTs to reach the 332

    maximum adsorbility (eq. (6)). 333

    157157 FCTNTsFCTNTsTNTsTNTsh

    + (6) 334

    The C7F15 radicals also react with water (water acts as a nucleophilic reagent to attrack 335

    the end carbon atom) to form OHFC 157 (Chen et al., 2007), an alcohol, which 336

    undergoes HF elimination to form C6F13COF (eq. (7)) (Chen and Zhang, 2006; Hori et 337

    al., 2004). C6F13COF was then hydrolyzed to form PFHpA with one less the CF2 unit 338

    (Hori et al., 2004) and fluoride ions were releases into aqueous solutions (eq. (8)). 339

    + ++ FHCOFFCOHFC 136157 (7) 340

    + ++ FHCOOHFCCOFFC 136136 (8) 341

    The C-C bond between C6F13 and COOH can be further cleaved to form C6F13 radicals, 342

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    21

    which undergoes similar reaction routes as C7F15 radicals to give PFHxA. The bond 343

    cleavage routes of PFOA were similar to the photo-Kolbe mechanism (Dillert et al., 344

    2008; Ravichandran et al., 2006). In the same manner, PFOA bearing shorter-chain 345

    PFCAs, such as PFHpA, PFHxA, PFPeA, and HFBA, were formed in a stepwise 346

    manner as time progressed. 347

    348

    4. Conclusions 349

    Titanate nanotubes (TNTs) synthesized by the M-H method with large surface 350

    specific surface areas make them good candidates for removal of PFOA from aqueous 351

    solutions. Direct photolysis, photocatalysis were conducted under different solution 352

    pH, coupled with XPS and FTIR spectroscopies. The findings obtained in this study 353

    have demonstrated the following: 354

    (1) PFOA could efficiently be photodecomposed by a 254-nm UV light (400 W) 355

    within 48 h without presence of photocatalysts. Also, shorter-chain 356

    perfluorocarboxylic acids (PFCAs) and released fluoride ions were formed during 357

    photodecomposition. 358

    (2) TNTs are good adsorbents than TiO2 for PFOA removal. The maximum adsorption 359

    capacity can be as high as 50 mg PFOA /g TNT at acidic pH of 4. 360

    (3) XPS and FTIR analyses indicated that interactions for the removal of PFOA mainly 361

    occurred on the surfaces of TNTs. 362

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    22

    363

    Acknowledgements 364

    The authors would like to thank the National Science Council of the Republic of 365

    China, for their financial support under Contract No. NSC 98-2221-E-002-040-MY3. 366

    367

    Appendix. supplementary data 368

    Supplementary data associated with this article can be found in the online versions, 369

    at doi:10.1016/j.watres.2011.00.000. 370

    371

    References 372

    Bahnemann, D.W., Hilgendorff, M., and Memming, R. (1997) Charge carrier 373

    dynamics at TiO2 particles: reactivity of free and trapped holes. J. Phys. Chem. B 374

    101(21), 4265-4275. 375

    Burns, D.C., Ellis, D.A., Li, H., Mcmurdo, C.J., and Webster, E. (2008) Experimental 376

    pKa determination for perfluorooctanoic acid (PFOA) and the potential impact of 377

    pKa concentration dependence on laboratory-measured partitioning phenomena 378

    and environmental modeling. Environ. Sci. Technol. 42(24), 9283-9288. 379

    Cao, M.H., Wang, B.B., Yu, H.S., Wang, L.L., Yuan, S.H., and Chen, J. (2010) 380

    Photochemical decomposition of perfluorooctanoic acid in aqueous periodate with 381

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    23

    VUV and UV light irradiation. J. Hazard. Mater. 179(1-3), 1143-1146. 382

    Chen, J. and Zhang, P. (2006) Photodegradation of perfluorooctanoic acid in water 383

    under irradiation of 254 nm and 185 nm light by use of persulfate. Water Sci. 384

    Technol. 54(11-12), 317-325. 385

    Chen, J., Zhang, P., and Liu, J. (2007) Photodegradation of perfluorooctanoic acid by 386

    185 nm vacuum ultraviolet light. J. Environ. Sci. 19(4), 387-390. 387

    Chen, Q., Du, G.H., Zhang, S., and Peng, L.M. (2002) The structure of trititanate 388

    nanotubes. Acta. Cryst. Sect. B 58(4), 587-593. 389

    Chen, Y.C., Lo, S.L., and Kuo, J. (2010a) Pb(II) adsorption capacity and behavior of 390

    titanate nanotubes made by microwave hydrothermal method. Colloids Surf. A: 391

    Physicochem. Eng. Asp. 361(1-3), 126-131. 392

    Chen, Y.C., Lo, S.L., Ou, H.H., and Chen, C.H. (2010b) Photocatalytic oxidation of 393

    ammonia by cadmium sulfide/titanate nanotubes synthesized by microwave 394

    hydrothermal method. Water Sci. Technol., article in press. 395

    Cheng, J., Vecitis, C.D., Park, H., Mader, B.T., and Hoffmann, M.R. (2008) 396

    Sonochemical degradation of perfluorooctane sulfonate (PFOS) and 397

    perfluorooctanoate (PFOA) in landfill groundwater: environmental matrix effects. 398

    Environ. Sci. Technol. 42(21), 8057-8063. 399

    Dillert, R., Bahnemann, D., and Hidaka, H. (2008) Light-induced degradation of 400

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    24

    perfluorocarboxylic acids in the presence of titanium dioxide. Chemosphere 67(4), 401

    785-792. 402

    Doll, T.E. and Frimmel, F.H. (2005) Cross-flow microfiltration with periodical 403

    back-washing for photocatalytic degradation of pharmaceutical and diagnostic 404

    residues-evaluation of the long-term stability of the photocatalytic activity of TiO2. 405

    Water Res. 39(5), 847-854. 406

    Estrellan, C.R., Salim, C., and Hinode, H. (2010) Photocatalytic decomposition of 407

    perfluorooctanoic acid by iron and niobium co-doped titanium dioxide. J. Hazard. 408

    Mater. 179(1-3), 79-83. 409

    Guan, B., Zhi, J., Zhang, X., Murakami, T., and Fujishima, A. (2007) Electrochemical 410

    route for fluorinated modification of boron-doped diamond surface with 411

    perfluorooctanoic acid. Electrochem. Commun. 9(12), 2871-2821. 412

    Higgins, C.P. and Luthy, R.G. (2006) Sorption of perfluorinated surfactants on 413

    sediments. Environ. Sci. Technol. 40(23), 7251-7256. 414

    Higgins, C.P., Field, J.A., Criddle, C.S., and Luthy, R.G. (2005) Quantitative 415

    determination of perfluorochemicals in sediments and domestic sludge. Environ. 416

    Sci. Technol. 39(11), 3946-3956. 417

    Hinderliter, P.M., DeLorme, M.P., and Kennedy, G.L. (2006) Perfluorooctanoic acid: 418

    relationship between repeated inhalation exposures and plasma PFOA 419

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    25

    concentration in the rat. Toxicology 222(1-2), 80-85. 420

    Hogue, C. (2006) POFA called likely human carcinogen. Chem. Eng. News 83, 5. 421

    Hori, H., Hayakawa, E., Einaga, H., Kutsuna, S., Koike, K., Ibusuki, T., Kitagawa, H., 422

    and Arakawa, R. (2004) Decomposition of environmentally persistent 423

    perfluorooctanoic acid in water by photochemical approaches. Environ. Sci. 424

    Technol. 38(22), 6118-6124. 425

    Hori, H., Murayama, M., and Kutsuna, S. (2009) Oxygen-induced efficient 426

    mineralization of perfluoroalkylether sulfonates in subcritical water. Chemosphere 427

    77(10), 1400-1405. 428

    Hori, H., Murayama, M., Inoue, N., Ishida, K., and Kutsuna, S. (2010) Efficient 429

    mineralization of hydroperfluorocarboxylic acids with persulfate in hot water. 430

    Catal. Today 151(1-2), 131-136. 431

    Hori, H., Nagaoka, Y., Sano, T., and Kutsuna, S. (2008a) Iron-induced decomposition 432

    of perfluorohexanesulfonate in sub- and supercritical water. Chemosphere 70(5), 433

    800-806. 434

    Hori, H., Nagaoka, Y., Yamamoto, A., Sano, T., Yamashita, N., Taniyasu, S., and 435

    Kutsuna, S. (2006) Efficient decomposition of environmentally persistent 436

    perfluorooctanesulfonate and related fluorochemicals using zerovalent iron in 437

    subcritical water. Environ. Sci. Technol. 40(3), 1049-1054. 438

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    26

    Hori, H., Yamamoto, A., Koike, K., Kutsuna, S., Murayama, M., Yoshimoto, A., and 439

    Arakawa, R. (2008b) Photocatalytic decomposition of a perfluoroether carboxylic 440

    acid by tungstic heteropolyacids in water. Appl. Catal. B: Environ. 82(1-2), 58-66. 441

    Ismail, A.A., Abboudi, M., and Holloway, P. (2007) Photoluminence from terbium 442

    doped silica-titania prepared by a sol-gel method. Mater. Res. Bull. 42(1), 137-142. 443

    Kutsuna, S. and Hori, H. (2008) Experimental determination of Henrys law constant 444

    of perfluorooctanoic acid (PFOA) at 298 K by means of an inert-gas stripping 445

    method with a helical plate. Atmos. Environ. 42(39), 8883-8892. 446

    Lampert, D.J., Frisch, M.A., and Speitel Jr., G.E. (2007) Removal of perfluorooctanoic 447

    acid and perfluorooctane sulfonate from wastewater by ion exchange. Periodical 448

    Haz., Toxic, and Radioactive Waste Mgmt. 11(1), 60-68. 449

    Lin, A.Y.C., Panchangam, S.C., and Ciou, P.S. (2010a) High levels of 450

    perfluorochemicals in Taiwans wastewater treatment plants and downstream 451

    rivers pose great risk to local aquatic ecosystems. Chemosphere 80(10), 452

    1167-1174. 453

    Lin, C.J., Chen, S.Y., and Liou Y.H. (2010b) Wire-shaped electrode of CdSe-sensitized 454

    ZnO nanowire arrays for photoelectrochemical hydrogen generation. Electrochem. 455

    Commun. 12(11), 1513-1516. 456

    Men, X.H., Zhang, Z.Z., Song, H.J., Wang, K., and Jiang, W. (2008) Fabrication of 457

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    27

    superhydrophobic surfaces with poly(furfuryl alcohol)/multi-walled carbon 458

    nanotubes composites. Appl. Surf. Sci. 254(9), 2563-2568. 459

    Moriwaki, H., Takagi, Y., Tanaka, M., Tsuruho, K., Okitsu, K., and Maeda, Y. (2005) 460

    Sonochemical decomposition of perfluorooctane sulfonate and perfluorooctanoic 461

    acid. Eviron. Sci. Technol. 39(9), 3388-3392. 462

    Musijowski, J., Trojanowicz, M., Szostek, B., Lima, J.L.F.D.C., Lapa, R., Yamashita, 463

    H., Takayanagi, T., and Motomizu, S. (2007) Flow-injection determination of total 464

    organic fluorine with off-line defluorination reaction on a solid sorbent bed. Anal. 465

    Chim. Acta 600(1-2), 147-154. 466

    Ochoa-Herrera, V. and Sierra-Alvarez, R. (2008) Removal of perfluorinated 467

    surfactants by sorption onto granular activated carbon, zeolite, and sludge. 468

    Chemosphere 72(10), 1588-1593. 469

    Ou, H.H. and Lo, S.L. (2007) Review of titania nanotubes synthesized via the 470

    hydrothermal treatment: fabrication, modification, and application. Sep. Purif. 471

    Technol. 58(1), 179-191. 472

    Panchangam, S.C., Lin, A.Y.C., Tsai, J.H., and Lin, C.F. (2009) Sonication-assisted 473

    photocatalytic decomposition of perfluorooctanoic acid. Chemosphere 75(5), 474

    654-660. 475

    Potera, C. (2009) Reproductive toxicology: study associates PFOS and PFOA with 476

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    28

    impaired fertility. Environ. Health Perspect. 117(4), A148. 477

    Qu, Y., Zhang, C., Li, F., Bo, X., Liu, G., and Zhou, Q. (2009) Equilibrium and kinetics 478

    study on the adsorption of perfluorooctanoic acid from aqueous solution onto 479

    powdered activated carbon. J. Hazard. Mater. 169(1-3), 146-152. 480

    Qu, Y., Zhang, C., Li, F., Chen, J., and Zhou, Q. (2010) Photo-reductive defluorination 481

    of perfluorooctanoic acid in water. Water Res. 44(9), 2939-2947. 482

    Ravichandran, L., Selvam, K., and Swaminathan, M. (2007) Effect of oxidants and 483

    metal ions on photodefluoridation of pentafluorobenzoic acid with ZnO. Sep. Purif. 484

    Technol. 56(2), 192-198. 485

    Ravichandran, L., Selvam, K., Krishnakumar, B., and Swaminathan, M. (2009) 486

    Photovalorisation of pentafluorobenzoic acid with platinum doped TiO2. J. Hazard. 487

    Mater. 167(1-3), 763-769. 488

    Ravichandran, L., Selvam, K., Muruganandham, M., and Swaminathan, M. (2006) 489

    Photocatalytic cleavage of C-F bond in pentafluorobenzoic acid with titanium 490

    dioxide-P25. J. Fluorine Chem. 127(9), 1204-1210. 491

    Senevirathna, S.T.M.L.D., Tanaka, S., Fujii, S., Kunacheva, C., Harada, H., Shivakoti, 492

    B.R., and Okamoto, R. (2010) A comparative study of adsorption of 493

    perfluorooctane sulfonate (PFOS) onto granular activated carbon, ion-exchange 494

    polymers and non-ion-exchange polymers. Chemosphere 80(6), 647-651. 495

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    29

    US EPA (2002) Revised draft hazard assessment of perfluorooctanoic acid and its salts. 496

    United States Environmental Protection Agency Office of Pollution Prevention and 497

    Toxics Risk Assessment Division: Washington, DC. Available from: 498

    http://www.ewg.org/files/EPA_PFOA_110402.pdf. 499

    Vecitis, C.D., Park, H., Cheng, J., Mader, B.T., and Hoffmann, M.R. (2008) Kinetics 500

    and mechanism of the sonolytic conversion of the aqueous perfluorinated 501

    surfactants, perfluorooctanoate (PFOA), and perfluorooctane sulfonate (PFOS) 502

    into inorganic products. J. Phys. Chem. 112(18), 4261-4270. 503

    Xu, X.H., Zhang, Z.Z., and Liu, W. (2009) Fabrication of superhydrophobic surfaces 504

    with perfluorooctanoic acid modified TiO2/polystyrene nanocomposites coating. 505

    Colloid. Surf. A: Phys. Eng. Asp. 341(1-3), 21-26. 506

    Yu, Q., Zhang, R., Deng, S., Huang, J., and Yu, G. (2009) Sorption of perfluorooctane 507

    sulfonate and perfluorooctanoate on active carbons and resin: kinetic and isotherm 508

    study. Water Res. 43(4), 1150-1158. 509

    Zhao, B. and Zhang, P. (2009) Photocatalytic decomposition of perfluorooctanoic acid 510

    with -Ga2O3 wide bandgap photocatalyst. Catal. Commun. 10(8), 1184-1187. 511

    Zhu, X., Castleberry, S.R., Nanny, M.A., and Butler, E.C. (2005) Effects of pH and 512

    catalyst concentration on photocatalytic oxidation of aqueous ammonia and nitrite 513

    in titanium dioxide suspersions. Environ. Sci. Technol. 39(10), 3784-3791. 514

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    31

    0 10 20 30 40 50

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    0 10 20 30 40 50

    0

    1

    2

    3

    4

    5

    6(b)

    C/C 0

    Time (h)

    pH 4 pH 7 pH 10 No UV

    (a)

    pH4 pH7 pH10 No UV0

    20

    40

    60

    80

    100

    Perc

    enta

    ge

    Degradation Defluorination

    Fluor

    ide

    io

    n (m

    M)

    Time (h)

    Fluoride ion PFOA PFHpA PFHxA Total fluoride ion

    1

    Fig. 1. (a) Direct photolysis of PFOA at different solution pH by 254 nm UV 2

    irradiation (b) and formed shorter-chain PFCAs during photodecomposition. The 3

    reaction conditions are [PFOA]initial = 50 mg L-1, T = 298 K in a 2 L solution. 4

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    32

    20 40 60 80

    -10

    0

    10

    20

    30

    40

    50

    TiO2 (pH 4) TNTs (pH 4) TNTs (pH 7) TNTs (pH 10)

    Q e (m

    g g-1

    )

    Ce (mg L-1)

    5

    Fig. 2. Adsorption isotherms of PFOA onto TNTs/TiO2. The reaction conditions are 6

    [PFOA]initial = 5 100 mg L-1, T = 298 K, reacted with 20 mg of TNTs or TiO2 in a 100 7

    mL solution. Ce is the equilibrium concentration in solutions and qe is the amount of 8

    PFOA adsorbed per unit weight of the adsorbent. 9

    10

    11

    12

    13

    14

    15

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    33

    0 5 10 15 20 25

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    R2=0.9939

    R2=0.9598

    R2=0.9655

    R2=0.9754

    TiO2 (0.50 g) TNTs (0.50 g) TNTs (0.25 g) TNTs (0.10 g)

    ln (C

    0/C)

    Time (h)

    0.50g TiO2 0.50g TNTs 0.25g TNTs 0.10g TNTs0

    10

    20

    30

    40

    50

    60

    49%

    59%

    44%

    Perc

    enta

    ge

    Degradation Defluorination

    19%

    16

    Fig. 3. Pseudo first-order kinetics of photocatalytic decomposition of PFOA by 17

    TNTs/TiO2 by 254 nm UV irradiation. The reaction conditions are [PFOA]initial = 50 18

    mg L-1, pH 4, T = 298 K in a 2 L solution. The first-order rate reaction can be expressed 19

    as 0t ln[C]-ktln[C] += , where t[C] is the concentration of the chemical of interest at 20

    a particular time and 0[C] is the initial concentration. 21

    22

    23

    24

    25

    26

    27

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    34

    1000 800 600 400 200 0 545 540 535 530 525

    300 295 290 285 280 700 695 690 685 680

    1085 1080 1075 1070 475 470 465 460 455

    (b)

    TNTs + PFOA

    TNTs + PFOA

    TNTs + PFOA

    (a)

    (e)

    (d)(c)

    Inte

    nsi

    ty (a.

    u.)

    (f)

    TNTs + PFOA

    TNTs + PFOA

    TiO2 + PFOATiO2 + PFOA

    TNTsTNTs

    TNTs

    TNTs

    TNTs

    TNTs

    TNTs + PFOA

    Binding Energy (eV)

    Ti2p3Na1s

    F1sC1s

    O1s

    TiO2 + PFOA

    TiO2 + PFOA

    TiO2 + PFOA

    TiO2 + PFOA

    28

    Fig. 4. XPS spectra of virgin TNTs and PFOA-TNTs/TiO2 complexes after 29

    photocatalysis. 30

    31

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    35

    4000 3500 3000 2500 2000 1500 1000-1.0

    -0.5

    0.0

    0.5

    1.0

    1.5

    2.0

    4000 3500 3000 2500 2000 1500 1000-1.0

    -0.5

    0.0

    0.5

    1.0

    1.5TiO2+PFOA

    TiO2

    TNTs+PFOA

    (b)

    (a)TNTs

    Abso

    rbe

    nce

    Wavenumber (cm-1)

    32

    Fig. 5. The Fourier-transformed infrared spectra of virgin TNTs/TiO2 and 33

    PFOA-TNTs/TiO2 complexes after photocatalysis. 34

    35

    36

  • MAN

    USCR

    IPT

    ACCE

    PTED

    ACCEPTED MANUSCRIPT

    36

    Ionization

    C7F15COOH

    C7F15COO-

    Adsorption Photocatalysis

    TNT+ + C7F15COO-

    TNTs

    Ionization

    TNT - C7F15COO

    C7F15 + COOAdsorption Photocatalysis

    TNT - C7F15

    TNT - PFOA

    C7F15OH

    H2O

    C6F13COF

    H+F-

    C6F13COOHF-

    PFHpACF2

    CF2

    Mineralization

    37

    38

    Fig. 6. Proposed PFOA decomposition mechanism by TNTs. 39

    40

    41

    42

    43

    44

    45