Effect of tungsten-doping on the properties and photocatalytic performance of titania thin films on glass substrates

https://doi.org/10.1016/j.jtice.2016.07.016Get rights and content

Highlights

  • Increase in W-doping enhanced anatase recrystallization.

  • Increase in W-doping increased grain sizes and changed grain morphology.

  • W6+ ions enhanced Ti-vacancy formation, leading to lattice contraction.

  • W-doping lowered the band gap owing to a positive shift in the conduction band.

  • The best photocatalytic performance was observed at 0.01 mol% W.

Abstract

Highly-transparent TiO2 thin films with varying W-dopant levels (0.010–0.100 mol%) were deposited on soda-lime-silica glass substrates by spin coating, followed by annealing in air at 450 °C for 2 h. Characterization was by GAXRD, AFM, SIMS, XPS, UV–vis, and MB degradation. The films were comprised of single-crystal anatase grains. Increasing doping levels increased the crystallite sizes, grain sizes, and film thicknesses; they also changed the grain morphology from anhedral to euhedral. W6+ ions formed a substitutional solid solution in anatase and probably exhibited charge compensation by Ti vacancy formation, which led to lattice contraction. Contamination from the glass substrates was homogeneously distributed on the grain boundaries. Doping decreased the optical indirect band gap. However, at the highest doping level, it is not certain if this was an artificial effect of precipitation from overdoping or a true effect of lattice stress. The 0.010 mol% W sample showed the best performance probably owing to its formation of a homogeneous solid solution within the range of semiconducting properties, which could have narrowed the band gap or introduced mid-gap states. The 0.100 mol% W sample may have been overdoped and caused WO3 precipitation. If so, the presence of the TiO2–WO3 couple may have enhanced charge separation and hence photocatalytic performance.

Introduction

Photocatalytic materials [1], [2], [3] are being used currently for self-sterilizing surfaces [4], air purification [5], [6], water purification [7], [8], [9], self-cleaning surfaces [10], [11], [12], and solar cells [13], [14], [15]. Titanium dioxide (TiO2) is one of the most promising photocatalysts owing to its low cost, non-toxicity, and chemical and thermal stability [16], [17], [18]. Work by Fujishima and Honda [19] on water splitting using single-crystal TiO2 showed that high-energy radiation, such as UV, is required to excite electrons from the valence to the conduction band owing to the relatively large band gap (∼3.2 eV [20]). This limits the efficiency of TiO2 as a photocatalyst since UV radiation comprises only ∼5% of the solar spectrum [21].

Several methods have been used to enhance both the spectral response of TiO2 in the visible region and its photocatalytic activity [22], [23], [24]. The most common is the addition of dopants, which are intended to create mid-gap states in the electronic structure, thereby narrowing the band-gap [25], [26]. The dopants also act as trapping centers that separate the hole and electron pairs, thereby extending the lifetime of these charge carriers and enhancing photocatalytic activity [26]. Several dopants, including metals [27], [28], [29], [30] and non-metals [31], [32], [33], have been trialed for this purpose.

Doping with transition metal ions can help to retard the recombination of photogenerated electrons and holes, which thus increases the lifetime of charge carriers [34]. Moreover, tungsten (W) additions are believed to extend the spectral response of TiO2 into the visible region through a positive shift in the conduction band [12], [13]. However, the limited studies on W-doped TiO2 have focused on powders. León-Ramos et al. [35] reported that 0.03 mol% W-doped TiO2, produced hydrothermally, displayed higher photodegradation compared to undoped TiO2. Li et al. [36] demonstrated that 3 mol% WO3-doped TiO2 powders, prepared by sol–gel, were most effective in degrading organic solutions, while Couselo et al. [37] suggested that the optimal performance of 3 mol% W-doped TiO2 powders resulted from two competing factors, namely increasing surface area and decreasing specific activity per adsorption site with increased doping. Recently, Hua et al. [38] reported that sol–gel-derived TiO2 nanoparticles doped with 0.4–1.0 wt% W resulted in enhanced capability of degradation of methyl orange solution owing to the role of W ions in shifting the absorption edge of TiO2 to the visible range. Wu et al. [39] showed that W-doped TiO2 particles produced by the solvothermal method displayed enhanced near-infrared absorption performance and similarly enhanced photocatalytic activity to that of commercial Degussa P25 TiO2 in the degradation of NOx gas. Further, Sathasivam et al. [40] demonstrated that 2.25 at% W-doped TiO2 films prepared by aerosol-assisted chemical vapor deposition were effective in degrading resazurin redox dye since the dopants contributed to an increase in the surface area and limitation in the relative grain boundary regions, which act as charge carrier recombination sites.

Prior studies have shown that doping with W6+ has the potential to modify the photocatalytic properties of TiO2 since the similarity of its crystal radius to that of Ti4+ [41] suggests the potential for substitutional solid solubility. Further, the fact that its crystal radius is smaller than that of the two interstices adjacent to the central Ti ion in the elongated TiO6 octahedron [42] suggests the potential for interstitial solid solubility. Both types of solid solubility would require charge compensation in the form of titanium vacancies or oxygen injection (with ionic compensation) or electrons (with electronic compensation); Frenkel defect formation also is possible. All of these would be expected to have significant effects on the photocatalytic performance.

Most prior work has focused on reducing the optical band gap of TiO2 through W additions, without extended consideration of the effects of doping on the mineralogical, microstructural, and photocatalytic properties. Further, in other studies, the doping levels used have tended to be high (2–20 mol%) and well above the typical semiconducting addition range. This is particularly the case for W-doped thin films, as shown in Table 1 [35], [36], [37], [38], [39], [40], [43], [44], [45]. Previous work by the authors [46] has led to the observation that additions of transition metal dopants typically in excess of ∼1 mol% (metal basis) are capable of exceeding the solubility limits, resulting in precipitation on the grain boundaries. This effect lowers the band gap of thin films but it cannot alter that of the doped TiO2 further than that of the TiO2 at the solubility limit. Hence, the band gap of the thin film and that of the TiO2 comprising the film are different. Further effects may include lattice distortion and partial amorphization. All of these effects can alter the photocatalytic efficiency of the doped films, generally deleteriously.

There are many variables that affect the photocatalytic performance of doped TiO2, including crystal structure, lattice deformation, defect equilibria (structural variables), grain size, surface area, and topography (microstructural variables) [46]. Since previous studies have taken a more focused approach to this range of variables, the present work aims to describe a more comprehensive investigation into the effects of doping on the mineralogical, microstructural, optical, chemical, and photocatalytic performance of TiO2 thin films deposited on the common substrate soda-lime-silica glass. In the present case, doping with W in a compositional range (0.010–0.100 mol%) that is consistent with the formation of a homogeneous solid solution and semiconducting dopant levels is designed to systematize the structural variables while avoiding the effects of mechanistic changes in the microstructural variables, all of which affect the photocatalytic properties. In particular, since the band gap is used widely in the assessment of photocatalytic performance, in the absence of a WO3-TiO2 phase diagram or other solubility data, the low levels of dopant are intended to avoid precipitation, significant modification of the microstructure from precipitation and liquid formation, and artificial alteration of the band gap. In the present work, the recognized potential to influence the photocatalytic performance by W-doping of TiO2, summarized above, has been investigated by the fabrication of thin films by spin coating, followed by annealing at 450 °C for 2 h. The thin films were characterized using different techniques to investigate the effects of W-doping on the mineralogical, morphological, topographical, chemical, optical, and photocatalytic properties.

Section snippets

TiO2 thin film fabrication

To prepare sol-gel solutions for coating, titanium isopropoxide (TIP, reagent-grade, 97 wt%, Sigma-Aldrich) and isopropanol (reagent plus ≥99 wt%, Sigma-Aldrich) were mixed to obtain a titanium concentration of 0.1 M (2.84 g of TIP diluted in 100 mL isopropanol). The solution then was magnetically stirred on a hot plate at 90 °C for 10 min. Different amounts of tungsten hexachloride (WCl6, ≥99.9 wt%, Sigma-Aldrich) were added to obtain W-dopant concentrations of 0.010, 0.025, 0.050, 0.075, and 0.100 

Mineralogical characteristics

GAXRD patterns of the annealed thin films (Fig. 1a) showed that the films consisted only of anatase (JCPDS 21-1272). No rutile or W-containing phases were observed, but these phases possibly could be present at levels below the detection limit [47]. However, these data are consistent with the presence of unsaturated solid solutions of W in TiO2 since the peak intensities show a consistent increasing trend. Had saturation solubility been reached, then the intensities would have become consistent

Summary

W-doped TiO2 thin films were fabricated using a sol-gel method followed by deposition by spin coating on soda-lime-silica glass substrates and annealing at 450 °C for 2 h. The major conclusions are as follows:

  • (1)

    GAXRD Data: W formed a substitutional solid solution in the TiO2 (anatase) lattice. This resulted in lattice contraction, largely owing to Ti vacancy formation.

  • (2)

    AFM Data: Undoped anhedral TiO2 of grain size ∼25–30 nm converted to euhedral grains of size ∼30-40 nm upon doping. The grains were

Acknowledgments

The authors acknowledge the financial support from the Australian Research Council (ARC) and the characterisation facilities provided by the Australian Microscopy & Microanalysis Research Facilities (AMMRF) node at UNSW Australia.

References (71)

  • M. Pelaez

    A review on the visible light active titanium dioxide photocatalysts for environmental applications

    Appl Catal A Environ

    (2012)
  • E. Aubry et al.

    Effect of Na diffusion from glass substrate on the microstructural and photocatalytic properties of post-annealed TiO2 films synthesised by reactive sputtering

    Surf Coat Technol

    (2012)
  • YuJ. et al.

    Effect of surface structure on photocatalytic activity of TiO2 thin films prepared by sol–gel method

    Thin Solid Films

    (2000)
  • LiXZ. et al.

    Photocatalytic activity of WOx-TiO2 under visible light irradiation

    J Photochem Photobiol B Chem

    (2001)
  • J. Tauc et al.

    State in the gap

    J Non-Cryst Solids

    (1972)
  • ZhangJ. et al.

    Enhanced photocatalytic hydrogen production activities of Au-loaded ZnS flowers

    ACS Appl Mater Interfaces

    (2013)
  • TianC. et al.

    Cost-effective large-scale synthesis of ZnO photocatalyst with excellent performance for dye photodegradation

    Chem Commun

    (2012)
  • BaiX. et al.

    Performance enhancement of ZnO photocatalyst via synergic effect of surface oxygen defect and graphene hybridization

    Langmuir

    (2013)
  • J. Šícha et al.

    High-rate low-temperature dc pulsed magnetron sputtering of photocatalytic TiO2 films: the effect of repetition frequency

    Nanoscale Res Lett

    (2007)
  • LiD. et al.

    Visible-light-driven N−F−codoped TiO2 photocatalysts. 2. Optical characterization, photocatalysis, and potential application to air purification

    Chem Mater

    (2005)
  • T.N. Obee et al.

    TiO2 photocatalysis for indoor air applications: effects of humidity and trace contaminant levels on the oxidation rates of formaldehyde, toluene, and 1,3-butadiene

    Environ Sci Technol

    (1995)
  • L. Cermenati et al.

    Probing the TiO2 photocatalytic mechanisms in water purification by use of quinoline, photo-fenton generated OH• radicals and superoxide dismutase

    J Phys Chem B

    (1997)
  • LiX.Z. et al.

    Study of Au/Au3+-TiO2 photocatalysts toward visible photooxidation for water and wastewater treatment

    Environ Sci Technol

    (2001)
  • Y. Ohko et al.

    Photoelectrochemical anticorrosion and self-cleaning effects of a TiO2 coating for type 304 stainless steel

    J Electrochem Soc

    (2001)
  • WuZ. et al.

    Structural color in porous, superhydrophilic, and self-cleaning SiO2/TiO2 bragg stacks

    Small

    (2007)
  • ZhangX. et al.

    Double-layered TiO2−SiO2 nanostructured films with self-cleaning and antireflective properties

    J Phys Chem B

    (2006)
  • B. O'Regan et al.

    A low-cost, high-efficiency solar cell based on colloidal TiO2 films

    Nature

    (1991)
  • U. Bach et al.

    Solid-state dye-sensitized mesoporous TiO2 solar cells with high photon-to-electron conversion efficiencies

    Nature

    (1998)
  • M.K. Nazeeruddin et al.

    Investigation of sensitizer adsorption and the influence of protons on current and voltage of a dye-sensitized nanocrystalline TiO2 solar cell

    J Phys Chem B

    (2003)
  • U.M.K. Shahed et al.

    Efficient photochemical water splitting by a chemically modified n-TiO2

    Science

    (2002)
  • YinS. et al.

    Preparation of visible light-activated titania photocatalyst by mechanochemical method

    Chem Lett

    (2003)
  • WangJ. et al.

    Mechanochemical synthesis of SrTiO3−xFx with high visible light photocatalytic activities for nitrogen monoxide destruction

    J Mater Chem

    (2003)
  • K. Honda et al.

    Electrochemical photocatalysis of water at a semiconductor electrode

    Nature

    (1972)
  • D.A.H. Hanaor et al.

    Review of the anatase to rutile phase transformation

    J Mater Sci

    (2011)
  • S.G. Kumar et al.

    Review on modified TiO2 photocatalysis under UV/Visible light: selected results and related mechanisms on interfacial charge carrier cransfer dynemics

    J Phys Chem A

    (2011)
  • Cited by (16)

    • Towards full-spectrum photocatalysis: Successful approaches and materials

      2021, Applied Catalysis A: General
      Citation Excerpt :

      For cations with oxidation state above Ti, tungsten is present as W6+ up to levels of ca. 25 at. %, with concomitant existence of Ti vacancies [112,113]. In this case, band gap energy mirrors the changes occurring in the local W-O-Ti local entities [112].

    • Highly efficient tungsten-doped hierarchical structural N-Enriched porous carbon counter electrode material for dye-sensitized solar cells

      2020, Electrochimica Acta
      Citation Excerpt :

      The load of transition metal compounds on carbon material substrates can significantly enhance the electrocatalytic activity and further improve the photovoltaic performance. In the past decade, among non-noble metal electrocatalysts including transition-metal oxides/chalcogenides/pnictides, tungsten species have widely highlighted great attention of researchers as the outstanding electrocatalysts owing to their Pt-like electronic behavior, excellent electrocatalytic activity, relatively low cost, and electrochemical stability [24,25,44–47]. Therefore, tungsten species are considered as the potential substitutes for noble metal Pt for I3− reduction reaction electrocatalyst.

    • Properties and performance of photocatalytic CeO<inf>2</inf>, TiO<inf>2</inf>, and CeO<inf>2</inf>–TiO<inf>2</inf> layered thin films

      2019, Ceramics International
      Citation Excerpt :

      Since the photocatalytic data do not show an overt trend, then both the simultaneous effect of the extent of recrystallization of the thin film phases and the associated defectivenesses of their nanostructures play significant but converse roles, although the former is greater. The preparation of the precursor solutions was essentially identical to that used for the sol-gel synthesis of CeO2 [14] and TiO2 [2,15–19] films. The precursor solution was prepared using cerium nitrate hexahydrate (99% purity, Sigma-Aldrich) as the source of Ce dissolved in ethanol.

    • Photocatalytic performance of TiO<inf>2</inf> impregnated polyester for the degradation of Reactive Green 12: Implications of the surface pretreatment and the microstructure

      2017, Journal of Photochemistry and Photobiology A: Chemistry
      Citation Excerpt :

      To overcome this problem, researchers supported the photocatalysts on different substrates. A variety of substrates had been tested as support for the catalyst such as glass, carbon fibres, polymers … etc. [12,13]. In the present study, polyester as support for TiO2 was used because of its flexibility, relative stability, resistance to UV, and low-cost [14].

    View all citing articles on Scopus
    View full text