contents journal of photochemistry and photobiology a: chemistrycamj/pdf/2013/hu_2013_photo.pdf ·...

9
Journal of Photochemistry and Photobiology A: Chemistry 256 (2013) 7–15 Contents lists available at SciVerse ScienceDirect Journal of Photochemistry and Photobiology A: Chemistry journa l h o me pa g e: www .elsevier.com/locate/jphotochem Enhanced photocatalytic degradation of dyes by TiO 2 nanobelts with hierarchical structures A. Hu a,, R. Liang a , X. Zhang b , S. Kurdi a , D. Luong b , H. Huang a , P. Peng a , E. Marzbanrad a , K.D. Oakes b , Y. Zhou a , M.R. Servos b,∗∗ a Department of Mechanical and Mechatronics Engineering, University of Waterloo, 200 University Avenue West, Waterloo, Ontario, Canada b Department of Biology, University of Waterloo, 200 University Avenue West, Waterloo, Ontario, Canada a r t i c l e i n f o Article history: Received 17 October 2012 Received in revised form 11 January 2013 Accepted 31 January 2013 Available online 19 February 2013 Keywords: TiO2 nanowires Hierarchical structure Surface adsorption Photocatalytic degradation a b s t r a c t TiO 2 nanobelt membranes with hierarchical structure were successfully synthesized by sequentially autoclaving Ti at 190 C for 3 d in the presence of 10 M NaOH and 10 M KOH solutions. Microstructural characterization revealed TiO 2 nanoparticles joining together and on the surface of nanobelts. These hierarchical structures form a three dimensional porous membrane which significantly enhances both surface specific area and light absorption, resulting in improved chemical adsorption capacity and pho- tocatalytic degradation efficiency relative to nanobelts with smooth surfaces using methylene blue as a model molecule. The adsorption of methylene blue to these structures follows a pseudo-second order kinetics chemisorption mechanism with rate-limited diffusion correlated to pore structure and size. The dominant reactive oxygen species are identified as hydroxyl radicals and valence band holes through the scavenging reaction. The synergistic enhancement of filtration through surface adsorption and pho- tocatalytic degradation is also demonstrated in a prototype photocatalytic membrane reactor with UV excitation at 365 nm. The reuse of nanobelt membranes after annealing shows the excellent recovery of TiO 2 catalyst. These studies may contribute to additional applications of hierarchical TiO 2 nanobelt membranes, including those harnessing sunlight for water treatment. © 2013 Elsevier B.V. All rights reserved. 1. Introduction The dominant causes of surface water and groundwater pol- lution are industrial discharges, landfill leachates, agricultural runoff, and municipal wastewater effluents which contain a myriad of contaminants including nutrients, pesticides as well as pharmaceuticals and personal care products (PPCPs) [1–3]. Conventional wastewater treatment utilizes various mechanical, chemical, biological and/or physical processes. Wastewater treat- ment plants (WWTPs) frequently combine multiple processes in their operations, such as coagulation–flocculation–sedimentation, biological degradation, filtration, chlorination, and ozonation during water treatment. Many of these methods require high oper- ating/maintenance costs or, in the case of chlorination, release disinfection by-products or produce unintended consequences in receiving environment fauna [4]. Biological treatment, while inex- pensive, cannot degrade many toxic or bio-recalcitrant organic compounds, such as pharmaceuticals or dyes [5–7]. Dyes are Corresponding author. Tel.: +1 519 8884567 35464. ∗∗ Corresponding author. E-mail addresses: [email protected], [email protected] (A. Hu), [email protected] (M.R. Servos). environmentally-relevant contaminants found in wastewater from textile, printing and paper manufacturing industries. The need for economically-feasible approaches to remove these waterborne contaminants promoted the development of innovative water treatment technologies. Many environmental remediation techniques employ simple physical processes such as adsorption to effectively remove non- biodegradable, environmentally hazardous substances. Activated carbon and a suite of other low-cost, high-sorbing materials such as biomass, agricultural and industrial solid waste, clays and zeo- lites have been extensively investigated as alternative adsorbents for color removal [8,9]. Conversely, advanced oxidation processes (AOPs) are increasingly applied as an energy intensive but highly effective means of non-selectively degrading various organic pol- lutants including PPCPs and dyes. AOPs rely upon the generation of highly reactive short-lived transitory species (principally hydroxyl radicals), to efficiently degraded or even mineralize a wide range of organics into smaller biodegradable compounds or inorganic molecules like CO 2 and H 2 O [10,11]. Some AOPs employ wideband semiconductive catalysts such as TiO 2 , ZnO, Fe 2 O 3 , SnO 2 , WO 3 , and CdS. TiO 2 in particular has attracted considerable research attention as a non-toxic, chemically inert and photostable catalyst capable of degrading waterborne dyes among other pollutants. TiO 2 nanopar- ticles have been successfully applied within prototype WWTPs 1010-6030/$ see front matter © 2013 Elsevier B.V. All rights reserved. http://dx.doi.org/10.1016/j.jphotochem.2013.01.015

Upload: others

Post on 26-Jan-2020

22 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Contents Journal of Photochemistry and Photobiology A: Chemistrycamj/pdf/2013/Hu_2013_Photo.pdf · 2013-06-21 · A. Hu et al. / Journal of Photochemistry and Photobiology A: Chemistry

Es

AKa

b

a

ARRAA

KTHSP

1

lrmaCcmtbdadrpc

(

1h

Journal of Photochemistry and Photobiology A: Chemistry 256 (2013) 7– 15

Contents lists available at SciVerse ScienceDirect

Journal of Photochemistry and Photobiology A:Chemistry

journa l h o me pa g e: www .e lsev ier .com/ locate / jphotochem

nhanced photocatalytic degradation of dyes by TiO2 nanobelts with hierarchicaltructures

. Hua,∗, R. Lianga, X. Zhangb, S. Kurdia, D. Luongb, H. Huanga, P. Penga, E. Marzbanrada,

.D. Oakesb, Y. Zhoua, M.R. Servosb,∗∗

Department of Mechanical and Mechatronics Engineering, University of Waterloo, 200 University Avenue West, Waterloo, Ontario, CanadaDepartment of Biology, University of Waterloo, 200 University Avenue West, Waterloo, Ontario, Canada

r t i c l e i n f o

rticle history:eceived 17 October 2012eceived in revised form 11 January 2013ccepted 31 January 2013vailable online 19 February 2013

eywords:iO2 nanowiresierarchical structure

a b s t r a c t

TiO2 nanobelt membranes with hierarchical structure were successfully synthesized by sequentiallyautoclaving Ti at 190 ◦C for 3 d in the presence of 10 M NaOH and 10 M KOH solutions. Microstructuralcharacterization revealed TiO2 nanoparticles joining together and on the surface of nanobelts. Thesehierarchical structures form a three dimensional porous membrane which significantly enhances bothsurface specific area and light absorption, resulting in improved chemical adsorption capacity and pho-tocatalytic degradation efficiency relative to nanobelts with smooth surfaces using methylene blue as amodel molecule. The adsorption of methylene blue to these structures follows a pseudo-second orderkinetics chemisorption mechanism with rate-limited diffusion correlated to pore structure and size. The

urface adsorptionhotocatalytic degradation

dominant reactive oxygen species are identified as hydroxyl radicals and valence band holes throughthe scavenging reaction. The synergistic enhancement of filtration through surface adsorption and pho-tocatalytic degradation is also demonstrated in a prototype photocatalytic membrane reactor with UVexcitation at 365 nm. The reuse of nanobelt membranes after annealing shows the excellent recoveryof TiO2 catalyst. These studies may contribute to additional applications of hierarchical TiO2 nanobelt

ose h

membranes, including th

. Introduction

The dominant causes of surface water and groundwater pol-ution are industrial discharges, landfill leachates, agriculturalunoff, and municipal wastewater effluents which contain ayriad of contaminants including nutrients, pesticides as well

s pharmaceuticals and personal care products (PPCPs) [1–3].onventional wastewater treatment utilizes various mechanical,hemical, biological and/or physical processes. Wastewater treat-ent plants (WWTPs) frequently combine multiple processes in

heir operations, such as coagulation–flocculation–sedimentation,iological degradation, filtration, chlorination, and ozonationuring water treatment. Many of these methods require high oper-ting/maintenance costs or, in the case of chlorination, releaseisinfection by-products or produce unintended consequences in

eceiving environment fauna [4]. Biological treatment, while inex-ensive, cannot degrade many toxic or bio-recalcitrant organicompounds, such as pharmaceuticals or dyes [5–7]. Dyes are

∗ Corresponding author. Tel.: +1 519 8884567 35464.∗∗ Corresponding author.

E-mail addresses: [email protected], [email protected]. Hu), [email protected] (M.R. Servos).

010-6030/$ – see front matter © 2013 Elsevier B.V. All rights reserved.ttp://dx.doi.org/10.1016/j.jphotochem.2013.01.015

arnessing sunlight for water treatment.© 2013 Elsevier B.V. All rights reserved.

environmentally-relevant contaminants found in wastewater fromtextile, printing and paper manufacturing industries. The needfor economically-feasible approaches to remove these waterbornecontaminants promoted the development of innovative watertreatment technologies.

Many environmental remediation techniques employ simplephysical processes such as adsorption to effectively remove non-biodegradable, environmentally hazardous substances. Activatedcarbon and a suite of other low-cost, high-sorbing materials suchas biomass, agricultural and industrial solid waste, clays and zeo-lites have been extensively investigated as alternative adsorbentsfor color removal [8,9]. Conversely, advanced oxidation processes(AOPs) are increasingly applied as an energy intensive but highlyeffective means of non-selectively degrading various organic pol-lutants including PPCPs and dyes. AOPs rely upon the generation ofhighly reactive short-lived transitory species (principally hydroxylradicals), to efficiently degraded or even mineralize a wide rangeof organics into smaller biodegradable compounds or inorganicmolecules like CO2 and H2O [10,11]. Some AOPs employ widebandsemiconductive catalysts such as TiO2, ZnO, Fe2O3, SnO2, WO3, and

CdS. TiO2 in particular has attracted considerable research attentionas a non-toxic, chemically inert and photostable catalyst capable ofdegrading waterborne dyes among other pollutants. TiO2 nanopar-ticles have been successfully applied within prototype WWTPs
Page 2: Contents Journal of Photochemistry and Photobiology A: Chemistrycamj/pdf/2013/Hu_2013_Photo.pdf · 2013-06-21 · A. Hu et al. / Journal of Photochemistry and Photobiology A: Chemistry

8 nd Ph

iTndeaanwchmsdnCcfmh

crp[cpnasT[nasnfitwi

mcscAtta

2

2

ontNts[f2

A. Hu et al. / Journal of Photochemistry a

ncluding those focussed on dye decoloration applications [5].wo photocatalyst loading methods have been investigated: TiO2anoparticle slurries and immobilized TiO2 nanoparticles embed-ed on a reaction surface. TiO2 nanoparticle slurries are a highlyfficient means of treating water by photocatalytic degradations the nanoparticles mix well with waterborne contaminants andllow for relatively unimpeded light exposure [1]. However, TiO2anoparticles must subsequently be separated from the treatedater, which represents a logistically challenging and energetically

ostly step [1,12]. Recently, TiO2 nanowire/nanotube membranesave been developed as a multifunctional material for water treat-ent, which promises to combine three removal mechanisms:

urface adsorption [9], ultrafiltration [11–13] and photocatalyticegradation [11,12,14]. However, the adsorption kinetics of TiO2anowire membranes have not been systematically established.ontaminant degradation by TiO2 nanowire membranes used asatalysts in photocatalytic membrane reactors (PMRs) may benefitrom synergies arising from combinations of two or three removal

echanisms; to date, the potential of such synergistic interactionsas not been explored.

Photocatalytic efficiency can be influenced by many physi-al and chemical features of catalysts, such as large and highlyeactive surface areas, enhanced light adsorption, and the sup-ressed recombination of photo-generated electrons and holes15–17]. Specially-designed nanowires with structural hierarchyan effectively suppress electron–hole recombination and improvehotocatalytic efficiency. Mixtures of anatase and rutile phase TiO2anoparticles (in 4:1 ratios; P25, Degassa) display a high photocat-lytic efficiency [18] as does a TiO2(B) nanowire with an anatasehell [19]. Fu et al. found that dispersived Au nanoparticles oniO2 films is also effectively improved the photocatalytic efficiency20]. Liu et al. reported another highly efficient heterogeneousanostructure in which anatase nanoparticles are layer-by-layerssembled on rutile nanorods [21]. Recently, Zhou et al. [17]ynthesized long TiO2 nanobelts embedded with surface anataseanoparticles using the acid hydrothermal method. Charge trans-

ers between nanobelts and nanoparticles leads to dramaticallymproved charged pair separation. From an engineering perspec-ive, simple methods for synthesizing TiO2 nanowire membranesith heterostructures in scalable volumes would significantly facil-

tate adoption of these materials in numerous applications.In this work, we developed a facile method of fabricating TiO2

embranes with structural hierarchy, in which TiO2 nanoparti-les can be grown on the surface of nanobelts. These hierarchicaltructures demonstrated enhanced surface adsorption and photo-atalytic degradation efficiencies towards the dye methylene blue.

prototype photocatalytic membrane reactor was designed andested with the newly developed TiO2 nanobelt membranes, wherehe synergistic degradation of methylene blue by both surfacedsorption and photocatalytic degradation was confirmed.

. Experimental procedure

.1. Chemicals and materials

Unless otherwise specified, all reagents and chemicals weref analytical grade and purchased from Sigma–Aldrich. TiO2anobelts were synthesized using modified hydrothermal growthechniques. TiO2 anatase nanobelts were grown for 48 h in 10 MaOH solution within 125 mL Teflon®-lined acidic digestion reac-

ors (Parr Instruments) at 190 ◦C using Ti foil as the titanium

ource and a small amount of carboxyl solvents as oxygen source11,22]. Subsequently, the suspension nanobelts were separatedrom the NaOH solution by centrifugation, potentially generating

g of nanobelt within 60 mL of solution. The fabricated nanobelts

otobiology A: Chemistry 256 (2013) 7– 15

were then dispersed into 10 M KOH solution and grown for another24 h at 190 ◦C. Following the KOH autoclaving, nanobelts wereagain centrifugally separated and washed in 0.1 M HCl solution anddeionized water several times until a neutral pH was obtained. Thefabricated nanobelts were deposited into membranes or annealedat 700 ◦C for 1 h for experimental characterization. Commerciallyavailable TiO2 P25 nanoparticles (Degussa, P25, specific area:50.964 m2/g) were used for comparison.

2.2. Membrane preparation

The TiO2 nanobelt membranes were fabricated by a filtrationand hot-press method [12]. Briefly, the TiO2 nanobelt suspensionwas deposited on a Whatman® #1001042 paper filter within avacuum filtration apparatus, with the deposited thick films subse-quently sandwiched between two Al2O3 ceramic discs and pressedfor 30 min at 200 ◦C by a uniaxial pressure machine exerting a forceof 200 N. The resulting TiO2 thick films were fired at 700 ◦C for 1 hto burn the filter paper away and shift the phase to more photocat-alytic activated one (see Fig. S1† batch-processing of TiO2 nanowiremembranes).

2.3. Adsorption and photocatalytic degradation

The surface adsorption experiments were carried out by dispers-ing a slurry of prepared nanomaterials into 100 mL of 0.038 mMmethylene blue solution (in ultrapure water, pH 6.8) in the dark at22 ◦C with adsorption accelerated by magnetic stirring at 300 rpm.Photocatalytic degradation was assessed in the presence or absenceof UV irradiation provided by a 100 W middle pressure mercurylamp (Cole Parmer) producing an intensity of 2.1 mW/cm2 with themain emission wavelength of 365 nm positioned 5 cm from the sur-face of the solution. To saturate the surface adsorption capacity ofthe nanomaterials, each solution was first stirred in the dark for 1 h.To determine the active oxygen species, KI, isopropanol (i-PrOH)and catalase were used as quenchers during degradation analysis.The concentration of KI, i-PrOH, and catalase in the initial reactionsolution were 0.1 mmol/L, 0.1 mmol/L and 500 mg/L according toprevious studies [23,24].

2.4. Photocatalytic membrane reactor

A photocatalytic membrane reactor (PMR) was machined fromstainless steel with two sealed chambers allowing for the secureinsertion of a TiO2 nanowire membrane between (see Fig. S2†Photocatalytic degradation experimental set-up: photocatalyticmembrane reactor). A 3 mm thick quartz window was installedon the top surface of the PMR and sealed by an O-ring, with theTiO2 membrane fixed parallel to the quartz window with a gap of1 cm. Waterborne contaminants are introduced into the top cham-ber between the quartz window and TiO2 membrane by an externalperistaltic pump. After filtration the treated water was collectedin the bottom chamber where it could be analyzed after a singlepass (or recirculated for multiple passes) across the TiO2 mem-brane. The TiO2 membrane is underlain by a stainless steel meshto avoid breakthrough from excessive pressure or flow; all watermust pass through the membrane as it is sealed by a Teflon® O-ringon top and bottom surfaces. The recovery of TiO2 membrane was

tested by post-annealing at 250 ◦C for 10 min. After post-annealing,the adsorbed methylene blue was totally removed and membranesrecovered into white. The annealed membrane was put back to thereactor and examined for reusability.
Page 3: Contents Journal of Photochemistry and Photobiology A: Chemistrycamj/pdf/2013/Hu_2013_Photo.pdf · 2013-06-21 · A. Hu et al. / Journal of Photochemistry and Photobiology A: Chemistry

nd Photobiology A: Chemistry 256 (2013) 7– 15 9

2

e(wdfmNwiEToupTwmAssgm6

3

3

i(feftawtapNtsTiawssdrsb(rNTaAtfa

A. Hu et al. / Journal of Photochemistry a

.5. Characterization and analysis

The phase and microstructure of fabricated nanowires werexamined by X-ray diffractometry, scanning electron microscopySEM) and transmission electron microscopy (TEM). TEM samplesere prepared by suspension of TiO2 nanobelts in ethanol andripping into carbon Lacey film coated Cu grids. The specific sur-ace area was determined by the Brunauer–Emmett–Teller (BET)

ethod using N2 adsorption data (Quantachrome InstrumentsOVA 2200). X-ray photoelectron spectrum (XPS) measurementsere carried out using a multi-technique ultra-high vacuum

maging XPS microprobe spectrometer (Thermo VG ScientificSCALab 250) with a monochromatic Al K� 1486.6 eV X-ray source.he spectrometer was calibrated by Au 4f7/2 (Binding energyf 84.0 eV) with respect to the Fermi level. The chamber vac-um level was maintained below 2 × 10−10 Torr. The constantass energy was set at 160 eV for survey scans at 1 eV step.he pass energy of 20 eV was used for Na1s/K2p/Ti2p/O1s scans,ith an energy resolution of 0.05 eV. Visual Raman spectra wereeasured using a Renishaw micro-Raman spectrometer with anr ion laser at an excitation wavelength of 488 nm. For bothpectrometers the resolution was 1 cm−1. The diffuse reflectancepectra were characterized by UV-NIR spectroscopy with an inte-rated spherical detector using BaTiO3 as reference material. Theethylene blue concentration was determined by adsorption at

63 nm.

. Results and discussion

.1. Phase identification and microstructural characterization

Fig. 1 shows typical SEM images of TiO2 nanobelts grown (a)n 10 M NaOH solution for 72 h, (b) in 10 M KOH for 72 h andc) in 10 NaOH solution for 48 h and then in 10 M KOH solutionor 24 h. These experiments illustrate that NaOH solution is nec-ssary for the formation of TiO2 nanobelts with widths rangingrom 30 to 100 nm, lengths in the range of tens of microme-ers, but of pretty thin thickness. However, when KOH solutionslone are employed, the nanowires become even narrower, withidths shrinking to a few nanometers, but retaining lengths up

o a few micrometers. Such variation in nanowire morphologyrising from substitutions in alkaline solutions is consistent withrevious studies [13]. However, nanowires grown in consecutiveaOH and KOH alkaline solutions displayed an innovative struc-

ure, in which 20–50 nm nanoparticles were synthesized on theurface of the nanobelts, with some connecting different nanobelts.hese hierarchical structures form complex 3D networks differ-ng from the entangled nanobelts that arise from growth in singlelkaline solutions since these hierarchical nanobelts are joined orelded via the nanoparticles. It is evident that these hierarchical

tructures are not a simple mixture of nanobelts and nanoparticlesince extended ultra-sonication treatment of suspension solutionoes not yield the separation of nanobelts and nanoparticles. X-ay diffraction patterns for nanowires grown in NaOH, KOH, orequential NaOH/KOH solutions are shown in Fig. 2, and peaks cane indexed by the anatase phase for nanowires grown in NaOHJCPDS 46-1238). For KOH solution the peaks are much wide withemarkable noise although the dominant phase is still anatase. InaOH/KOH solution anatase peaks are slightly sharper than in KOH.his can be understood due to a small nanowire grown in KOHddition and hierarchic structures appeared in sequential solutions.

lthough the morphology of these nanowires differs dramatically,

he dominant phase is the same after annealing at 700 ◦C. In order tourther elucidate the chemical purity three kinds of TiO2 nanowiresre examined by X-ray Photoelectron Spectroscopy (XPS, shown in

Fig. 1. Scanning electron microscopy images of TiO2 nanobelts grown in (a) 10 MNaOH solution at 190 ◦C for 72 h, (b) in 10 M KOH at 190 ◦C for 72 h and (c) in 10 MNaOH at 190 ◦C for 48 h and then in 10 M KOH at 190 ◦C for 24 h.

Fig. 3). The survey unveils that Ti and O are the dominant com-positions co-existing with minor C, Na and/or K. The Ti 2p3/2 andTi 2p1/2 peaks can be identified at 464.3 eV and 458.5 eV, respec-tively, which are consistent with the typical values for TiO2 [25,26].The almost identical Ti2p doublet in all three compositions indicatethat Ti keeps the similar tetravalent bonding state [26,27]. From thepeak areas of C1s at 284.6 eV, Na1s at 1070.7 eV and K2p at 294.7 eVone can deduce their atomic percentages in the corresponding com-

positions. The concentration of C yields 6% for nanowires fromNaOH solution, 5.2% from NaOH–KOH and 5% from KOH solution.The carbon is induced by carbonyl solvents added as an oxidizerduring hydrothermal growth [11,22]. Na is founded as 3.7% and
Page 4: Contents Journal of Photochemistry and Photobiology A: Chemistrycamj/pdf/2013/Hu_2013_Photo.pdf · 2013-06-21 · A. Hu et al. / Journal of Photochemistry and Photobiology A: Chemistry

10 A. Hu et al. / Journal of Photochemistry and Photobiology A: Chemistry 256 (2013) 7– 15

Fga

4tNtectp[di[fm

e

F“a

NaOH

NaOH-KOH

200 300 400 500 600 700 800 900 1000

KOH

Inte

nsity

/a.u.

Raman shift (cm-1)

ig. 2. X-ray diffractometry patterns of three TiO2 nanomaterials: (a) nanobelts

rown in NaOH solution, (b) nanobelts fabricated in KOH solution, and (c) hier-rchical nanobelts grown sequentially in NaOH and KOH solutions.

.3% in nanowires separately grown in NaOH and NaOH–KOH solu-ions while the K fraction is founded at 0.9% and 2.4% in wires fromaOH–KOH and KOH solutions, respectively. Such low concentra-

ions show that the ion exchange by diluted HCl is very effective toliminate alkaline ions [28,29]. The high purity of TiO2 nanowiresan be also evident by Raman spectra. Fig. 4 shows the Raman spec-ra for three kinds of nanowires. Typical modes of TiO2 anatasehase at 197, 399, 515, and 639 cm−1 can be clearly observed26,28,30]. The low mode at 144 cm−1 is out of the effective range ofevice detection. The small peak at 247 cm−1 of nanowires grown

n sequential solution could be relative to minor amorphous TiO231,32]. Therefore, three kinds of nanowires display similar Ramaneatures. This indicates all of them are TiO2 anatase although their

orphologies are significantly different.The diffuse reflection adsorption spectra demonstrate their

nergy gaps are around 3.0 eV for nanowires from NaOH solutions

ig. 3. X-ray Photoelectron spectra of three kinds of TiO2 nanowires. (a) Survey scan.In MNN” comes from Indium substrate. (b) Ti2p doublets. (c) Na1s peak and (d) C1snd K2p peaks.

Fig. 4. Raman spectra of TiO2 nanowires grown from three kinds of alkaline solu-tions. Excitation wavelength is 488 nm.

and 2.95 eV for nanowires either from NaOH–KOH or KOH solu-tion (Fig. 5). It is important to point out that hierarchical nanobeltmembranes (NaOH–KOH) display the enhanced absorbance (1.3compared to 0.9 for NaOH and KOH nanowires) within UV wave-lengths, confirming the unusual light absorption capacities of thesenovel structures. Such an absorption enhancement, which hasbeen observed in a variety of hierarchical structures [33–35], arsesfrom the multi-scattering of mesoporous structures [33–36]. Thedetailed lattice hierarchical structure of both surface TiO2 nanopar-ticles and long nanobelts shown in Fig. 1(c) can be seen in theTEM images of Fig. 6. The lattice orientation can be attributed tothe materials being in the anatase phase, which is in agreementwith the XRD results. Fig. 6a shows two welded multi-layerednanobelts joined through shared layers. Surface nanoparticles may

be attached at angles to the orientation of the core nanowires,sometimes separated by an amorphous layer between the ori-ented surface nanoparticles and underlying nanobelts (Fig. 6b). In

200 250 300 350 400 450 500 550 600 650 700 750 800

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

2 3 4 5 60

1

2

3NaOH-KOH

KOH

NaOH

Ab

s

Wavelength (nm)

NaOH-KOH

KOH

NaOH

E (eV)

(h

v)1

/2

Fig. 5. UV-NIR diffuse reflection adsorption spectrum of TiO2 nanobelt membranesmeasured by an integrated sphere detector. Inset is the analysis of optical gap byTauc formula.

Page 5: Contents Journal of Photochemistry and Photobiology A: Chemistrycamj/pdf/2013/Hu_2013_Photo.pdf · 2013-06-21 · A. Hu et al. / Journal of Photochemistry and Photobiology A: Chemistry

A. Hu et al. / Journal of Photochemistry and Photobiology A: Chemistry 256 (2013) 7– 15 11

F anopaa tase p(

swoaaosvahs11ac(

t

TCt

ig. 6. TEM images of hierarchical TiO2 nanobelts: (a) two joined nanobelts; (b) a nnd surface nanoparticles. The lattice distance corresponds to (1 0 1) planes of anad) an amorphous nanoparticle grown on the nanobelt.

ome instances, the surface particles are almost perfectly alignedith the orientation of their basal nanobelts (Fig. 6c), while in

thers the surface nanoparticles themselves contain amorphousreas with little defined orientation as shown in Fig. 6(d). Thereppears to be no general relationship coordinating the orientationf surface nanoparticles and their basal nanobelts. Nevertheless, noodium titanate or potassium titanate are evident in TEM obser-ation in various nanowires [28,29,37]. This displays that TiO2natase is the dominant composition in current nanowires withierarchical structures, consistent with XRD and Raman data. Thepecific surface areas induced by the differing alkaline solutions are6.709 m2/g for anatase nanowires obtained from NaOH solution,70.211 m2/g for anatase nanowires autoclaved from KOH solution,nd an enhanced specific surface area of 47.614 m2/g for hierarchi-

al nanobelts produced by sequential exposure to NaOH and KOHTable 1).

To better understand the conditions favoring the growth ofhe current hierarchical structures, the use of sequential alkaline

able 1haracteristics of TiO2 nanomaterials as adsorbents and photocatalysts. ki is the intra-parhe macro- and meso-pore structure of the adsorbent. Kap is the apparent reaction rate in

Nanomaterials S (m2/g, BET) qe (mg/g)

P25 50.236 65.2

NaOH 16.709 88.6

KOH 170.211 170

NaOH–KOH 47.614 159

rticle adhered onto a nanobelt. Note the amorphous shell between crystalline corehase; (c) a crystalline nanoparticle orientated parallel to the nanobelt orientation;

solutions was contrasted with results from mixtures of two alkalinesolutions. The growth morphology of TiO2 nanobelts with increas-ing KOH concentrations is provided in supplementary materials(see Fig. S3† TiO2 nanowires growth with various KOH/NaOH ratiosin solution), which demonstrates a similar hierarchical structurecan be generated with a proper NaOH:KOH ratio of approximately5:2. Higher ratios of KOH suppress both the nanobelt width and themass of nanobelts obtained. This result is consistent with previousreport [13], where TiO2 nanobelts with a width of 20–100 nm arefabricated in a NaOH solution while narrow TiO2 nanobelts of 10 nmin width are grown in a KOH solution. As nanobelt shape dependson different growth rates at variable planes, the present hierarchi-cal structure of adhering nanoparticles on nanobelt surfaces are thenet result of competing growth induced by the respective NaOH and

KOH solutions [11,13]. It is notable that in the sequential growth,although there are only potassium ions in the 2nd growth stagethe initial composite is Na2Ti3O7 [11]. There are still a competitivecrystalline favorite by Na and K ions, respectively.

ticle diffusion rate, and the i is relative to continuous diffusion stages correlated to photocatalyic degradation according to pseudo-first model.

K1 (mg/g min0.5) K2 (mg/g min0.5) Kap (min−1)

8.1 4.5 0.04512.7 2.8 0.05730.4 0.56 0.02126.7 2.8 0.089

Page 6: Contents Journal of Photochemistry and Photobiology A: Chemistrycamj/pdf/2013/Hu_2013_Photo.pdf · 2013-06-21 · A. Hu et al. / Journal of Photochemistry and Photobiology A: Chemistry

12 A. Hu et al. / Journal of Photochemistry and Photobiology A: Chemistry 256 (2013) 7– 15

0 20 40 60 80 100

0.01

0.02

0.03

0.04

0.05

0.06

Co

nce

ntr

atio

n (

mM

)

Time (min)

KOH

NaOH

P25

NaOH-KOH

Fig. 7. Adsorption kinetics of methylene blue onto 4 TiO2 nanomaterials: P25 andthree nanobelts autoclaved from NaOH, KOH and sequential NaOH–KOH solutions,rpp

3

fswttsituattN

mtmntoi[momenaNbTnt

att[

Fig. 8. Kinetic analysis of rates of degradation by photocatalysis with four kinds ofTiO2 catalysts. The solid lines fit the pseudo first-order model. The apparent degra-

espectively. The solution volume is 100 mL, pH 6.8. Inset: the time profile fitted toseudo-second order kinetics. The obtained equilibrium adsorption capacity, qe , isrovided in Table 1.

.2. Chemical adsorption

The adsorption kinetics of methylene blue in the presence ofour different TiO2 nanomaterials at room temperature is demon-trated in Fig. 7. While the adsorption capacity of the nanowiresere high (especially those grown in KOH or NaOH/KOH owing to

heir large specific surface areas), the commercial P25 nanopar-icles displayed a low adsorption rate, despite their fairly highpecific surface area of 50.694 m2/g. While specific surface area ismportant, adsorption is also dependent on the surface charge ofhe nanoparticle/nanowire, the adsorbent mass, ambient pH val-es and the temperature of the solution. For all TiO2 nanomaterials,dsorptive capacity is nearly saturated after 30 min, but notably,he NaOH–KOH hierarchical nanobelts display enhanced adsorp-ion capacities compared to smooth TiO2 nanowires grown fromaOH solution.

With respect to kinetic modeling, the pseudo-second-orderodel has been evaluated to reveal the adsorption mechanism,

/qt = 1/(kq2e ) + t/qe where qt = (C0 − Ct)(v/w) is the mass of

ethylene blue adsorbed (g/g) at time t by 1 g of hierarchicalanobelts, C0 and Ct are the initial concentration and concentra-ion at time, t, v is the volume of the solution, and w is the massf the adsorbent used, qe is the equilibrium adsorption capac-ty, and k is the rate constant of pseudo-second-order adsorption38]. The linear fit (inset on Fig. 7) indicates that the adsorption

echanism is based on chemisorption since the pseudo-secondrder model is based on a chemical reaction between dimeric dyeicelles and sorbent [39,40]. Application of best fit determined the

quilibrium adsorption capacities of P25 nanoparticles and TiO2anowires autoclaved in NaOH, KOH, NaOH–KOH as 65.2, 88.6, 170nd 159 mg/g, respectively. The adsorption capacities of KOH andaOH–KOH nanobelts approach those of commercial activated car-on (200–400 mg/g [41]) and titanate nanowires (196 mg/g) [9].his shows a possibility to using the chemical adsorption of theseanowire membranes to remove the micro-pollutants during waterreatment.

To optimize the TiO2 nanowire membrane microstructure for

dsorption, it is necessary from a mechanistic viewpoint, to identifyhe steps involved in adsorption. It is well known that adsorp-ion incorporates external mass transfer and intra-particle diffusion38,42]. Intra-particle diffusion, which can be described by Weber

dation rate is given in Table 1. Inset: SEM images TiO2 nanowires synthesized inKOH solution. SEM samples are fabricated by drying the aquatic suspension of TiO2

nanowires.

and Morris’ model, qt = ki t1/2 + c where c is the constant and ki isthe intra-particle diffusion rate [43], may be further separated intoa few continuous diffusion stages correlated to the macro-, meso-and micropore structure of the adsorbent [42,44]. The plotting of qt

as a function of t1/2 reveals two linear ranges in the curve reflectingthe transition from adsorption to one kind of pore to another (seeFig. S4† Analysis with Weber and Morris’ model). The slope of thelinear relationship, ki, characterizes the rate of intra-particle dif-fusion; the value of ki given by the best fit is shown in Table 1. Itis obvious that for the three nanobelts, ki has two different valueswhile P25 displays similar values over two diffusion regimes. Theseresults are consistent with the adsorption reported for sphagnumpeat [44], where the continuous branching pore distribution wasproposed. It is reasonable to deduce that a similar pore distribu-tion is evident within the current membranes. However, modifiedpeat amended with polyvinylalcohol and formaldehyde displayeda mitigated variation in pore size [39], similar to that of P25. Itseems indicating that the chemical adsorption is not a simple sur-face adsorption. The further discrimination of different adsorptionmechanics is interesting but beyond the scope of this study.

3.3. Photocatalytic degradation

Photocatalytic degradation of methylene blue in the presenceof various TiO2 nanomaterial suspensions is illustrated in Fig. 8.Although nanobelts produced from KOH possess strong surfaceadsorption properties, these nanomaterials only weakly photode-graded methylene blue molecules, displaying similar propertiesto those of titanate nanotubes [11]. Nanobelts from NaOH areslightly more effective at photocatalytically degrading methyleneblue than P25, while hierarchical TiO2 nanobelts autoclaved fromNaOH–KOH solution display the strongest degradation. Photocat-alytic degradation can be described by pseudo first-order kinetics:−dC/dt = KapC, where the C is the concentration and Kap is the appar-ent reaction rate. A time function ln(C0/C) = Kapt can be deduced byintegrating the above equation where C0 is the initial concentra-tion. The linear relationship in Fig. 8 is the best fit to the pseudo

first order model, yielding the value of Kap as 0.021, 0.045, 0.057 and0.089 min−1 for nanobelts from KOH, P25, NaOH and NaOH–KOH,respectively. These degradation rates are better than that of hierar-chical TiO2 nanoflakes [35] and hollow TiO2 anatase microspheres
Page 7: Contents Journal of Photochemistry and Photobiology A: Chemistrycamj/pdf/2013/Hu_2013_Photo.pdf · 2013-06-21 · A. Hu et al. / Journal of Photochemistry and Photobiology A: Chemistry

A. Hu et al. / Journal of Photochemistry and Photobiology A: Chemistry 256 (2013) 7– 15 13

Fmo

wftcNiFibaodnapdNarhtchcedhm0

3

wwpiaaiOf

200 300 400 500 600 700

0.0

0.2

0.4

0.6

0.8

1.0

1.2

Ab

so

rba

nce

Wavelength(nm)

0 min

10 min

20 min

30 min

40 min

Malachite Green

0.0

0.4

0.8 0 min

10 min

20 min

30 min

40 min

Congo Red

62.3%, 31.1% and 6.6%, respectively.The mechanism by which hierarchical structures enhance

photocatalytic efficiency might be insightful for expropriatingthese properties in additional applications. Initially, hierarchical

0.00

0.03

0.06

0.09

De

gra

da

tio

n R

ate

(m

in-1)

ig. 9. Adsorption and photocatalytic removal of methylene blue by a TiO2 nano-aterials suspended as a slurry. The dark regime illustrates removal by adsorption

ver 30 min while the 365 nm UV regime illustrates photocatalyic degradation.

ith more reactive (0 0 1) facets as photocatalysts [45]. The per-ormance of current hierarchical TiO2 nanowires can be comparedo nanoheterostructure Ag/TiO2 nanowires [17], where the con-entration of methyl orange is degraded in 2 orders within 30 min.evertheless, the photocatalytic performance is determined by var-

ous factors, as mentioned in the introduction. Shown in the inset toig. 8, we find that nanowires from KOH solution tend to aggregaten aquatic media and forms secondary micro-sized bundles. It haseen shown that an aggregated semiconducting catalyst results in

lower photocatalytic capacity [46]. The synergistic decolorationf methylene blue by both surface adsorption and photocatalyticegradation under 365 nm UV illumination is shown in Fig. 9. It isotable that the chemical adsorption and degradation are two sep-rated color removal mechanisms. Nanowires from KOH solutionsossess a superior surface adsorption while their photocatalyticegradation mechanism is relatively weak. While nanowires fromaOH have an improved degradation performance their chemicaldsorption capacity is limited. The chemical adsorption is cor-elated with the specific surface area (see Table 1). In general,ierarchical nanobelts generated with NaOH/KOH alkaline solu-ions demonstrate a significant improvement in color removal byombining two mechanisms. This is an attractive feature of currentierarchical nanobelts. Shown in Fig. 10, two typical dyes, Mala-hite green and Congo red are chosen for examine the extensiveffectiveness of hierarchical nanowires. Both concentrations can beramatically degraded with exposing to 365 nm UV light with oneour. The analysis shows both of them follow a pseudo-first orderodel with degradation rates of 0.057/min for Malachite green and

.034/min for Congo red.

.4. Identification of reaction oxygen species

The active oxygen species have been extensively investigatedith commercial P25 TiO2 nanoparticles and TiO2 coated non-oven paper [23,24], where hydroxyl, positive holes and hydrogeneroxide are identified as dominant oxygen species. In this exper-

ment, we employ the similar techniques to identify the presencend roles of these species. Shown in Fig. 11, degradation rates with

dding three kinds of specie quenchers are examined. Isopropanols used to scavenge OH•, KI can remove valence band holes andH• radicals while catalase is a quencher of H2O2 [23,47]. There-

ore, 60.4% decrease of k due to i-PrOH can be attributed to the

Fig. 10. Photocatalytic degradation of Malachite green and Congo red solutions withhierarchic TiO2 nanowires exposing to 365 nm UV light.

removal of OH•. Furthermore, it can deduce that 21.6% of degrada-tion is based on valence band holes from an 82% decrease of thedegradation rate due to KI addition. The scavenging experimentwith catalase shows the significant role of H2O2 during dye degra-dation. There are two origins to generate H2O2, either from TiO2photocatalysis or dye direct photolysis [23,24]. The removal H2O2results in 52.3% of decrease of K. However, H2O2 does not directlyoxidize MB rather works as intermediate species to generate OH•.Since both H2O2 and holes can form OH•, we can deduce about thehole contribution due to photolysis is 8.1%. This value is obtainedby subtracting 52.3% from 60.4%. One can easily figure out the totalhole contribution for dye degradation is 29.7% by adding 8.1% to21.6%. Hence, the left 9.9% can be attributed to the mixture effectof secondary species, such as, O2

•−, HO2•− and O3. The obtained

degradation fractions, 60.4% from OH•, 29.7% from holes and 9.9%from other oxidative species are approximately consistent with theprevious report with P25 nanoparticles [24], where the values are

TiO2/UV i-PrOH KI Catalase

Fig. 11. Degradation rates of methylene blue with hierarchical TiO2 nanowires withoxygen species quenchers: isopropanol, KI and catalase.

Page 8: Contents Journal of Photochemistry and Photobiology A: Chemistrycamj/pdf/2013/Hu_2013_Photo.pdf · 2013-06-21 · A. Hu et al. / Journal of Photochemistry and Photobiology A: Chemistry

14 A. Hu et al. / Journal of Photochemistry and Ph

0 5 10 15 20 25 30 35 40 45 50 55 60

0.020

0.025

0.030

0.035

0.040

Co

nce

ntr

atio

n(

mM

)

Filtration time

NaOH

NaOH-KOH

reuse

Dark 365 nm

Fig. 12. Synergistic (adsorption/photocatalytic degradation) decoloration of methy-lene blue by TiO2 nanobelt membranes in a prototype photocatalytic membranereactor. The filtration flow rate was 50 mL/min with the dark regime correspond-ing to the surface adsorption and the 365 nm UV regime illustrating the combinedeffects of surface adsorption and UV photocatalytic degradation. Two nanomem-branes membranes are compared: membranes autoclaved in NaOH solution alone,and hierarchical nanobelt membranes sequentially autoclaved in NaOH and KOH.The reuse curve displays the performance of TiO2 nanobelt membranes after 3 con-tar

spptltrpmsba[racee

3

TbtntNbritberb

inuing experimental cycles. Each cycle includes 5 continuing adsorptions in darknd then 5 continuing degradations under UV. After each cycle, the membrane isecovered by annealing at 250 ◦C for 10 min.

tructures exhibit heightened UV light adsorption due to their com-licated surface which can enhance light harvesting to improvehotocatalytic efficiency (Fig. 5). As both OH• and holes are veryransient species (half-life ∼10−9 s), photocatalytic degradation isimited to the vicinity of the nanomaterial surface, reinforcinghe importance of surface adsorption behavior by the nanomate-ial. Hierarchical nanobelts displayed enhanced surface adsorptionroperties, which are beneficial for degrading methylene blueolecules in the vicinity of the nanobelts. Finally, hierarchical

tructures may also facilitate the separation of electron–hole pairsy a charge-transfer mechanism between crystalline nanobelts andmorphous surfaces similar to surface modification by vitamin C41]. The amorphous surface area can trap charge carriers, therebyeducing the recombination of electron–hole pairs, which mayccount for the enhanced photocatalytic efficiency of hierarchi-al nanobelts relative to those derived from NaOH alone. However,lucidating the underlying mechanism behind this photocatalyticnhancement is beyond the scope of the current work.

.5. Prototype TiO2 nanowire membrane reactors

We further characterized the decoloration of methylene blue byiO2 nanobelt membranes in the prototype photocatalytic mem-rane reactor using NaOH and NaOH/KOH derived membranes inhe presence and absence of UV (Fig. 12). Notably, hierarchicalanobelts demonstrate both enhanced surface adsorption and pho-ocatalytic degradation relative to TiO2 nanobelts autoclaved fromaOH solution alone. However, when using hierarchical mem-ranes mounted in the PMR, both the adsorption and degradationates dramatically decreased relative to the use of nanobelt slurryn solution. The decline in surface adsorption can be attributedo the significant reduction of contact area between the water-

orne dye and membranes since only the top membrane surface isxposed to the dye. Similarly, the lower photocatalytic degradationate is also a function of the limited upper membrane surface areaeing exposed to UV light. Enlarging the illumination area should be

otobiology A: Chemistry 256 (2013) 7– 15

considered to optimize the membrane performance. Nevertheless,the preliminary experiments with the PMR demonstrate successfulmethylene blue removal by adsorption and photocatalytic degra-dation. Since the present membranes we utilized in the PMR canalso filter nanowires, it would be possible to combine the use ofthese nanomaterials in a slurry configuration with their filtrationby the same material configured into a membrane in the PMR toenhance water treatment. Under this scenario, adsorption, photo-catalytic degradation and nanocatalyst separation could be realizedin the same PMR.

After filtration the color of TiO2 nanobelt membranes turnedfrom white to blue due to the surface adsorbed methylene blue.The recovery of TiO2 nanobelt membranes is tested by anneal-ing the membrane at 250 ◦C for 10 min. Such a post-annealingcan effectively remove the adsorbed dye. The membrane is reusedto examine the reusability of membranes. Shown in Fig. 12, themembrane displays almost similar adsorption in the dark and pho-tocatalytic degradation performance with 365 nm excitation for 4repeated experimental cycles (for clearance only the data for thelast repeated experimental cycle are shown). Each cycle includes5 times of continuing adsorption followed by 5 times of continu-ing photocatalytic degradation. After each experimental cycle, theTiO2 nanobelt membrane is recovered by annealing. Thus, the cur-rent experiment displays an excellent stability and recovery of TiO2nanobelt membranes.

Currently, up to 10 pieces of membranes with a mass of 0.2 gcan be fabricated from one-pot synthesis (2 g, Fig. S1† batch-processing of TiO2 nanowire membranes) while one furnace cansimultaneously process 6–8 high pressure reactors in parallel. Con-sequently, over 50 individual membranes can be produced withone furnace, dramatically reducing fabrication costs. Compared toother methods to generate hierarchical structures, where eitherspecific synthesis facility is required [35] or a complicated acidicpost-treatment is employed [13], the current method presents apractical and simple protocol to fabricate hierarchical nanobeltsscalable to a large mass production with superior photocatalyticdegradation and surface adsorption performance. The current PMRcan be easily modified to combine a compound parabolic solar col-lector for economical solar disinfection of water [1,48], potentiallybroadening application of these materials to water treatment sce-narios within remote locations or developing countries.

4. Conclusions

Hierarchical TiO2 nanobelts with surface TiO2 nanoparticleshave been synthesized by sequentially autoclaving in 10 M NaOHand 10 M KOH solutions at 190 ◦C. The resultant hierarchicalnanomaterials displayed enhanced chemical adsorption and photo-catalytic degradation properties towards the model dye methyleneblue. The chemical adsorption of methylene blue followed apseudo-second order kinetic degradation displaying rate-limiteddiffusion dependent on pore structure and size distribution. Thedominant oxidative species were identified as hydroxyl radicalsand valence band holes. In a prototype photocatalytic membranereactor, adsorption, filtration and photocatalytic degradation canbe combined to generate synergies in the decoloration of methy-lene blue. The present results demonstrate the proof-in-conceptapplication of hierarchical TiO2 nanobelt membranes for contami-nant removal during water treatment.

Acknowledgements

This work has been financially supported by the Natural Sci-ences and Engineering Research Council of Canada through astrategic project grant, the Canadian Water Network Innovative

Page 9: Contents Journal of Photochemistry and Photobiology A: Chemistrycamj/pdf/2013/Hu_2013_Photo.pdf · 2013-06-21 · A. Hu et al. / Journal of Photochemistry and Photobiology A: Chemistry

nd Ph

TRC&tC

A

fj

R

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

A. Hu et al. / Journal of Photochemistry a

echnologies for Water Treatment Program, and the Canadaesearch Chairs Program. Technical support from Trojan UV, theity of Guelph Wastewater Services, Deep Blue NRG, and GE Water

Process Technologies are highly appreciated. X. Zhang would likeo thank the Ontario Ministry of Research & Innovation and theanadian Institutes of Health Research for financial supports.

ppendix A. Supplementary data

Supplementary data associated with this article can beound, in the online version, at http://dx.doi.org/10.1016/j.photochem.2013.01.015.

eferences

[1] S. Malato, J. Blanco, A. Vidal, C. Richter, Photocatalysis with solar energy at apilot-plant scale: an overview, Applied Catalysis B 37 (2002) 1–15.

[2] C.D. Metcalfe, B.G. Koenig, D.T. Bennie, M. Servos, T.A. Ternes, R. Hirsch, Occur-rence of neutral and acidic drugs in effluents of Canadian sewage treatmentplants, Environmental Toxicology and Chemistry 22 (2003) 2872–2880.

[3] M.N. Chong, B. Jin, C.W.K. Chow, C. Saint, Recent developments in photocatalyticwater treatment technology: a review, Water Research 44 (2010) 2997–3927.

[4] S. Malato, P. Fernandez-Ibanez, M. Maldonado, I.M. Blanco, W. Gernjak, Decon-tamination and disinfection of water by solar photocatalysis: recent review andtrends, Catalysis Today 147 (2009) 1–59.

[5] L. Bousselmi, S.U. Geissen, H. Schroeder, Textile wastewater treatment andreuse by solar catalysis: results from a pilot plant in Tunisia, Water Scienceand Technology 49 (2004) 331–337.

[6] S. Mozia, M. Tomaszewska, A.W. Morawski, Photocatalytic membrane reac-tor (PMR) coupling photocatalysis and membrane distillation-effectiveness ofremoval of three azo dyes from water, Catalysis Today 129 (2007) 3–8.

[7] L. Rizzo, S. Meric, M. Guida, D. Kassinos, V. Belgiorno, Heterogeneous pho-tocatalytic degradation kinetics and detoxification of an urban wastewatertreatment plant effluent contaminated with pharmaceuticals, Water Research43 (2009) 4070–4078.

[8] M. Rafaltullah, O. Sulaiman, R. Hashim, A. Ahmad, Adsorption of methylene blueon low-cost adsorbents: a review, Journal of Hazardous Materials 177 (2010)70–80.

[9] K. Hu, X. Xiao, X. Cao, R. Hao, X. Zuo, X. Zhang, J. Nan, Adsorptive separation andphotocatalytic degradation of methylene blue dye on titanate nanotube pow-ders prepared by hydrothermal process using metal Ti particles as a precursor,Journal of Hazardous Materials 192 (2011) 515–520.

10] V. Augugliaro, M. Litter, L. Palmisano, J. Soria, The combination of hetero-geneous photocatalysis with chemical and physical operations: a tool forimproving the photoprocess performance, Journal of Photochemistry and Pho-tobiology C 7 (2006) 127–144.

11] A. Hu, X. Zhang, K.D. Oakes, P. Peng, Y. Zhou, M. Servos, Hydrothermal growthof free standing TiO2 nanowire membranes for photocatalytic degradation ofpharmaceuticals, Journal of Hazardous Materials 189 (2011) 278–285.

12] N. Barka, S. Qourzal, A. Assabbane, A. Nounah, Y. Ait-Ichou, Factors influencingthe photocatalytic degradation of Rhodamine B by TiO2-coated non-wovenpaper, Journal of Photochemistry and Photobiology A 195 (2008) 346–351.

13] X. Zhang, J. Pan, A. Du, W. Fu, D. Sun, J. Leckie, Combination of one-dimensionalTiO2 nanowire photocatalytic oxidation with microfiltration for water treat-ment, Water Research 43 (2009) 1179–1186.

14] S.P. Albu, A. Ghicov, J.M. Macak, R. Hahn, P. Schmuki, Self-organized, free-standing TiO2 nanotube membrane for flow through photocatalytic application,Nano Letters 7 (2007) 1286–1289.

15] H.G. Yang, C.H. Sun, S.Z. Qiao, J. Zou, G. Liu, S.C. Smith, H.M. Cheng, G.Q. Lu,Anatase TiO2 single crystals with a large percentage of reactive facets, Nature453 (2008) 638–641.

16] M. Liu, L. Piao, W. Lu, S. Ju, L. Zhao, C. Zhou, H. Li, W. Wang, Flower-likeTiO2 nanostructures with exposed {0 0 1} facets: facile synthesis and enhancedphotocatalysis, Nanoscale 2 (2010) 1115–1117.

17] W. Zhou, G. Du, P. Hu, G. Li, D. Wang, H. Liu, J. Wang, R. Boughton, D. Liu, H.Jiang, Nanoheterostructures on TiO2 nanobelts achieved by acid hydrothermalmethod with enhanced photocatalyic and gas sensitive performance, Journalof Materials Chemistry 21 (2011) 7937–7945.

18] Z. Zheng, H. Liu, J. Ye, J. Zhao, E.R. Waclawik, H. Zhu, Structure and contributionto photocatalytic activity of the interfaces in nanofibers with mixed anataseand TiO2 (B) phases, Journal of Molecular Catalysis A 316 (2010) 75–82.

19] D. Yang, H. Liu, Z. Zheng, Y. Yuan, J. Zhao, E. Waclawik, X. Ke, H. Zhu, An efficientphotocatalyst structure: TiO2 (B) nanofibers with a shell of anatase nanocrys-tals, Journal of the American Chemical Society 131 (2009) 17885–17893.

20] P. Fu, P.Y. Zhang, Uniform dispersion of Au nanoparticles on TiO2 film via elec-

tronstatic self-assembly for photocatalytic degradation of bisphenol A, AppliedCatalysis B 96 (2010) 176–184.

21] Z. Liu, X. Zhang, S. Nishimoto, M. Jin, D.A. Tryk, T. Murakam, A. Fujishima,Anatase TiO2 nanoparticles on rutile TiO2 nanorods: a heterogeneous nano-structure via layer-by-layer assembly, Langmuir 23 (2007) 10916–10919.

[

otobiology A: Chemistry 256 (2013) 7– 15 15

22] A. Hu, X. Zhang, D. Luong, K.D. Oakes, M.R. Servos, R. Liang, S. Kurdi, P. Peng,Y. Zhou, Adsorption and photocatalytic degradation kinetics of pharmaceut-icals by TiO2 nanowires during water treatment, Waste Biomass Valorization3 (2012) 443–449.

23] X. Zhang, T. Zhang, J. Ng, D. Sun, High-performance multifunctional TiO2

nanowire ultrafiltration membrane with a hierarchical layer structure for watertreatment, Advanced Functional Materials 19 (2009) 3731–3736.

24] X. Zhang, D.D. Sun, G. Li, Y. Wang, Investigation of the roles of active oxygenspecies in photodegradation of azo dye AO7 in TiO2 photocatalysis illuminatedby microwave electrodeless lamp, Journal of Photochemistry and PhotobiologyA 199 (2008) 311–315.

25] G. Wang, H. Wang, Y. Ling, Y. Tang, X. Yang, R.C. Fitzmorris, C. Wang, J.Z. Zhang,Y. Li, Hydrogen-treated TiO2 nanowire arrays for photoelectrochemical watersplitting, Nano Letters 11 (2011) 3026–3033.

26] X. Chen, L. Liu, P.Y. Yu, S.S. Mao, Increasing solar absorption for photocatalysiswith black hydrogenated titanium dioxide nanocrystals, Science 331 (2011)746–750.

27] A. Foli, I. Pochard, A. Nonat, U.H. Jackbsen, A.M. Shepherd, D.E. Macphee,Engineering photocatalytic cements: understanding TiO2 surface chemistry tocontrol and modulate photocatalytic performances, Journal of the AmericanCeramic Society 93 (2010) 3360–3369.

28] R.A. Zarate, S. Fuentes, J.P. Wiff, V.M. Fuenzalida, A.L. Cabrera, Chemical compo-sition and phase identification of sodium titanate nanostructures grown fromtitania by hydrothermal processing, Journal of Physics and Chemistry of Solids68 (2007) 628–637.

29] Q. Chen, W. Zhou, G. Du, L.M. Peng, Trititanate nanotubes made via a singlealkali treatment, Advanced Materials 14 (2002) 1208–1211.

30] L.C. Sikuvhihulu, N.J. Coville, T. Ntho, M.S. Scurrell, Potassium titanate: an alter-native support for gold catalysed carbon monoxide oxidation? Catalysis Letters123 (2008) 193–197.

31] Q. Li, B. Liu, L. Wang, D. Li, R. Liu, B. Zou, T. Cui, G. Zou, Pressure-induced amorphization and polyamorphism in one-dimensional singlecrystal TiO2 nanomaterials, Journal of Physical Chemistry Letters 1 (2010)309–314.

32] T. Mazza, E. Barborini, P. Piseri, P. Milani, D. Cattaneo, A. Li Bassi, C.E. Bottani,C. Ducati, Raman spectroscopy characterization of TiO2 rutile nanocrystals,Physical Review B 75 (2007), 045416-1-5.

33] Z. Zhang, P. Wang, Optimization of photoelectrochemical water splitting per-formance on hierarchical TiO2 nanotube arrays, Energy and EnvironmentalScience 5 (2012) 6506–6512.

34] W. Yang, F. Wan, Q. Chen, J. Li, D. Xu, Controlling synthesis of well-crystallized mesoporous TiO2 microspheres with ultrahigh surface area forhigh-performance dye-sensitized solar cells, Journal of Materials Chemistry 20(2010) 2870–2876.

35] Y. Tang, P. Wee, Y. Lai, X. Wang, D. Gong, P.D. Kanhere, T. Lim, Z. Dong,Z. Chen, Hierarchical TiO2 nanoflakes and nanoparticles hybrid structure forimproved photocatalytic activity, Journal of Physical Chemistry C 116 (2012)2772–2780.

36] A. Usami, Theoretical study of application of multiple scattering of light to a dye-sensitized nanocrystalline photoelectrochemical cell, Chemical Physics Letters277 (1997) 105–108.

37] X. Meng, D. Wang, J. Liu, S. Zhang, Preparation and characterization of sodiumtitanate nanowires from brookite nanocrystallites, Materials Research Bulletin39 (2004) 2163–2170.

38] Q. Sun, L. Yang, The adsorption of basic dyes from aqueous solution on modifiedpeat–resin particles, Water Research 37 (2003) 1535–1544.

39] K. Vasanth Kumar, V. Ramamurthi, S. Sivanesan, Modeling the mechanisminvolved during the sorption of methylene blue onto fly ash, Journal of Colloidand Interface Science 284 (2005) 14–21.

40] Y. Ho, G. McKey, The kinetics of sorption of basic dyes from aqueous solutionby sphagnum moss peat, Canadian Journal of Chemical Engineering 76 (1998)822–827.

41] T. Rajh, J.M. Nedeljkovic, L.X. Chen, O. Poluektov, M.C. Thurnauer, Improvingoptical and charge separation properties of nanocrystalline TiO2 by surfacemodification with vitamin C, Journal of Physical Chemistry B 103 (1999)3515–3519.

42] G.M. Walker, L. Hansen, J.A. Hanna, S.J. Allen, Kinetics of a reactive dye adsorp-tion onto dolomitic sorbents, Water Research 37 (2003) 2081–2089.

43] W.J. Weder, J.C. Morris, Kinetics of adsorption on carbon from solution, Journalof the Sanitary Engineering Division ASCE 89 (1963) 31–60.

44] S.J. Allen, G. Mckay, K.Y.H. Khader, Intraparticle diffusion of a basic dyeduring adsorption onto sphagnum peat, Environmental Pollution 56 (1989)39–50.

45] S. Liu, J. Yu, M. Jaroniec, Tunable photocatalytic selectivity of hollow TiO2 micro-sphere composed of anatase polyhedral with exposed {0 0 1} facets, Journal ofthe American Chemical Society 132 (2010) 11914–11916.

46] A. Mills, S.L. Hunte, An overreview of semiconductor photocatalysis, Journal ofPhotochemistry and Photobiology A 108 (1997) 1–35.

47] K. Ishibashi, A. Fujishima, T. Watanabe, K. Hashimoto, Quantum yields of activeoxidative species formed on TiO2 photocatalyst, Journal of Photochemistry andPhotobiology A 134 (2000) 139–142.

48] J. Blanco, S. Malato, P. Fernandez, A. Vidal, A. Morales, P. Trincado, J.C. Oliveira,C. Minero, M. Musci M., C. Casalle C., M. Brunotte, S. Tratzky, N. Dischinger, K.H.Funken, C. Sattler, M. Vincent, M. Collares-Pereira, J.F. Mendes, C.M. Rangel,Compound parabolic concentrator technology development to commercialsolar detoxification applications, Solar Energy 67 (1999) 317–330.