Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Photocatalytic WO3/TiO2 nanowires: WO3 polymorphs influencing the atomic layer deposition of TiO2

Dávidné Nagy *ab, Tamás Firkala b, Eszter Drotár c, Ágnes Szegedi c, Krisztina László d and Imre Miklós Szilágyi be
aInstitute for Materials and Processes, School of Engineering, The University of Edinburgh, The King's Buildings, Mayfield Road, Edinburgh, EH9 3JL, UK. E-mail: davidne.nagy@gmail.com
bDepartment of Inorganic and Analytical Chemistry, Budapest University of Technology and Economics, Szent Gellért tér 4., Budapest, H-1111, Hungary
cHungarian Academy of Sciences, Institute of Materials and Environmental Chemistry, Magyar tudósok körútja 2., Budapest, H-1117, Hungary
dDepartment of Physical Chemistry and Materials Science, Budapest University of Technology and Economics, Budafoki út 8., Budapest, H-1111, Hungary
eMTA-BME Technical Analytical Research Group of the Hungarian Academy of Sciences, Szent Gellért tér 4., Budapest, H-1111, Hungary

Received 25th July 2016 , Accepted 30th September 2016

First published on 30th September 2016


Abstract

50–70 nm hexagonal (h-) and 70–90 nm monoclinic (m-) WO3 nanoparticles (NPs) were prepared by controlled annealing of (NH4)xWO3 in air at 470 and 600 °C, respectively. In addition, 5–10 nm thick and several micrometer long h-WO3 nanowires (NWs) were obtained by microwave hydrothermal synthesis at 160 °C with Na2WO4, HCl and (NH4)2SO4 as starting materials. TiO2 was deposited on h-WO3 NWs by atomic layer deposition (ALD) at 300 °C using Ti(iOPr)4 and H2O as precursors. The as-prepared materials were studied by TG/DTA-MS, XRD, Raman, SEM-EDX, TEM, ellipsometry, UV-Vis, and their photocatalytic activity was also tested by the photodecomposition of aqueous methyl orange. Our study is the first evidence of diverse ALD nucleation on various WO3 polymorphs, since on h-WO3 NWs TiO2 nucleated only as particles, whereas on m-WO3 conformal TiO2 film was formed, explained by the different surface OH coverage of h- and m-WO3. The h-WO3 NWs had significantly higher photocatalytic activity compared to h-WO3 NPs, and similar performance as m-WO3 NPs. By adding TiO2 to h-WO3 NWs by ALD method, the photocatalytic performance increased by 65%, showing clearly the uniqueness of ALD to obtain superior oxide composite photocatalysts.


Introduction

Water scarcity and the purity of available water resources around the world are concerning issues of our century.1,2 Semiconductor photocatalysis received substantial attention recently to address environmental remediation challenges such as purifying water sources.3–7 It was presented that using solar energy the purification of water is possible from toxic compounds such as certain pharmaceutical drugs, bacteria or herbicids etc.8,9 Among semiconductor oxides TiO2 is one of the most studied photocatalyst since its valence and conductance energy levels are suitable for both oxidation and reduction of water molecules (water splitting).10–17 However, the utilization of TiO2 is hugely limited by its restricted light absorption properties to the UV range. Therefore, besides TiO2, the photocatalytic activity of several other semiconductor oxides was also studied (e.g. V2O5, WO3, ZnO, ZrO2).18–24 Nanostructured WO3 can absorb part of the Visible spectrum, making it the second most studied oxide for visible-responsive photocatalysis.25

Another limiting factor of TiO2 is the relatively fast recombination of the photo-induced charges. One approach to improve the photocatalytic performance of a semiconductor oxide is forming composite photocatalysts with other semiconductors. An important benefit of forming composite nanostructures could be the reduced recombination rate of the photo-generated charges by effective charge transfer between the two semiconductor oxide. In the case of type II band alignment where one of the nanostructure has both the valence and conduction band energies lower positioned than the respective bands of the counterpart effective separation of the charge carriers can take place.26 In addition to the production of longer-lived charges, the heterostructure can benefit from Vis activity if a UV active photocatalyst such as TiO2 is coupled with semiconductor oxides absorbing in the visible region. TiO2/WO3 is such a system and several studies showed that such nanocomposite heterostructures have superior photocatalytic activities over single semiconductors.27–30

It is also known that the photocatalytic activity is a complex function of several factors such as the crystal structure, morphology or optical properties of the material etc.31,32 In order to achieve high performance photocatalysts maximizing the surface area of the material is a general approach. Therefore, several techniques have been adopted to fabricate nanostructured oxide photocatalysts. Among the reported nanostructures, nanofibers have received immense attention due to their high surface-to-volume ratio, good optical, electric and chemical properties.33,34

Previous studies revealed that WO3/TiO2 1D photocatalysts have huge potential.35–38 Unfortunately, WO3 is quite difficult to prepare with very small dimensions. The smallest forms of the most studied WO3 polymorph, i.e. monoclinic (m-) WO3 have been 10–100 nm as particles, or 50–200 nm as nanofibers.39–41 Nevertheless, m-WO3 is a widespread photocatalyst. In contrast, the second most important WO3 modification, i.e. hexagonal (h-) WO3, has been studied in photocatalysis only once, and its photocatalytic activity was lower compared to m-WO3.42 It was explained by that unlike the completely oxidized m-WO3, h-WO3 always contain some cation (e.g. Na+, K+, NH4+) impurities in its hexagonal channels, which are vital for stabilizing the metastable structure.43,44 Besides, h-WO3 contains also partially reduced W atoms, which may serve as recombination centers during the photocatalytic reaction, decreasing the activity.43 However, h-WO3 has been prepared recently in the form of very fine nanowires with ca. 5–10 nm thickness, which is a very promising morphology for photocatalysis, but unfortunately their photocatalytic properties have not been tested yet.

Recently it was reported that m-WO3/TiO2 core/shell nanofibers, where the shell layer was deposited by atomic layer deposition (ALD), had excellent photocatalytic properties due to the synergy of WO3 and TiO2.45 ALD is based on successive, alternating surface controlled reactions from the gas phase to produce highly conformal and uniform thin films with thickness control of sub-nanometer precision. Thus, ALD provides new strategies in modifying the properties of nanoscaled materials and new synthetic routes to novel nanostructures.46–52

In the above example the m-WO3 nanofibers were 200–300 nm thick. Thus, it is logical to expect better photocatalytic properties, if WO3 1D nanostructures with even smaller dimensions can be used as substrates. The now available 5–10 nm thick h-WO3 fibers are good candidates for employing in WO3/TiO2 photocatalysts with improved properties.

In this study, we obtained h-WO3 NWs by microwave assisted hydrothermal synthesis at 160 °C with Na2WO4, HCl and Na2SO4 as starting materials. ALD was the method of choice to put TiO2 onto the h-WO3 nanowires, since it has the capability to deposit very thin layers on highly structured surfaces. The TiO2 layer was deposited at 300 °C using Ti(iOPr)4 and H2O as precursors, and it was designed to have around 3 nm thickness because previous studies showed that thin TiO2 nanolayers could significantly enhance the photo-efficiency of the nanosized WO3.45 It must be noted that up to now ALD growth has been done only on m-WO3, but not on h-WO3. Since the two polymorphs have different surface properties,42 this allowed us to study the characteristics of ALD nucleation on the surface of various WO3 crystalline structures, which was not yet reported in the literature to the best of our knowledge.

For comparison, we also prepared hexagonal (h-) and monoclinic (m-) WO3 nanoparticles (NPs) by controlled annealing of (NH4)xWO3 in air at 470 and 600 °C, respectively. In addition, we prepared an m-WO3/TiO2 composite by annealing the h-WO3/TiO2 sample at 600 °C in air. It is well established that at this temperature the h-WO3 structure transforms into m-WO3. Usually when m-WO3 is formed thermally from nanostructured precursors, it will be present in the form of 50–200 nm particles.53 However, we aimed to test whether the TiO2 coating might prevent the morphology change, and m-WO3/TiO2 with nanowire morphology could be obtained.

The obtained materials were studied by TG/DTA-MS, XRD, Raman, SEM-EDX, TEM, ellipsometry, UV-Vis, and their photocatalytic activity was tested by UV-Vis through the photo-bleaching of aqueous methyl orange.

Experimental

Preparation methods

Hexagonal (h-) and monoclinic (m-) WO3 nanoparticles (NPs) were prepared by annealing hexagonal ammonium tungsten bronze, (NH4)xWO3−y in air at 470 °C and 600 °C, respectively.34,54

For the preparation of h-WO3 nanowires (NWs) a microwave-assisted hydrothermal synthesis method was applied.44 1.5 g Na2WO4·2H2O was dissolved in 33.75 ml H2O, then under stirring 3.75 ml 3 M HCl was added dropwise, and finally 22.5 ml 0.5 M (NH4)2SO4 was introduced. The as-prepared solution was heated to 160 °C in 20 min, and then kept at 160 °C for 3 hours in a Synthos 3000 Anton Paar microwave reactor. The solid reaction product was centrifuged, washed two times with water, once with ethanol, and again two times with water. In each washing step, 45 ml solvent was poured over the crystals; the dispersion was stirred for 1 min, and centrifuged for 5 min at 6000 min−1. Finally, the as-prepared h-WO3 NWs were dried at 80 °C for 12 hours. The yield was 64%.

The h-WO3/TiO2 nanowire composite was prepared by depositing a TiO2 nanolayer onto the surface of the as-prepared h-WO3 NWs by atomic layer deposition (ALD). The ALD reaction was carried out at 300 °C in a Picosun SUNALE R-100 type reactor. 100 ALD cycles were applied using Ti(OiPr)4 (60 °C evaporation temperature) and H2O as precursors, with 1 s pulse and 30 s purge times for both precursors.

The m-WO3/TiO2 composite sample was obtained by annealing the h-WO3/TiO2 sample at 600 °C in air.

Characterization methods

Thermal analysis was used to investigate the influence of washing steps during the h-WO3 NW preparation and to check the thermal stability of as-synthesized h-WO3 NWs. For this an STD 2960 simultaneous TG/DTA device (TA Instruments Inc.) was used, which was on-line coupled to a Thermostar GSD 200 quadruple mass spectrometer (Balzers Instruments) through a heated (200 °C), 100% silanated quartz capillary (D = 0.15 mm). During the TG/DTA-MS measurements, samples (ca. 100 mg) were heated in a Pt crucible at 10 °C min−1 heating rate in flowing air (130 ml min−1). The fragments of the evolved gases were monitored by Multiple Ion Detection (MID) mode in the 1–64 m/z range.

The crystal phases were determined by recording the XRD pattern of the samples by a PANalytical X'pert Pro MPD X-ray diffractometer using Cu Kα radiation.

The Raman spectra provided further information on the crystal structure and the bonds. A Jobin–Yvon Labram type spectrophotometer equipped with an Olympus BX-41 microscope was used. For excitation source a frequency-doubled Nd–YAG laser (532 nm) was employed. The scattered photons were collected by a CCD type detector.

To investigate the optical properties of the catalysts, UV-Vis diffuse reflectance spectra were recorded by a Cary 100 UV-Vis spectrophotometer equipped with a DRA-CA-30I type integration sphere.

For the study of the morphology and composition, SEM-EDX and TEM measurements were applied. A LEO 1540XB type RÖNTEC FEG SEM microscope equipped with a Quantax EDX detector was used for the SEM-EDX recordings. The TEM images were provided by a FEI Morgagni 268D type TEM microscope employing a tungsten cathode.

For determining the apparent surface area of the photocatalysts, the BET model was used. The low temperature nitrogen adsorption measurement was conducted in a NOVA 2000E type (Quantachrome, USA) device at the temperature of liquid nitrogen.

For estimating the ALD TiO2 film thickness, a TiO2 film was deposited on a Si wafer by ALD, using 300 cycles and the same precursors, pulse and purge times, as in the case of h-WO3 NW substrates. The film thickness was determined by ellipsometry. The spectra were recorded by a Woollam M-2000DI ellipsometer between 400 and 1500 nm.

Photocatalysis

The photocatalytic efficiencies of the as-prepared catalysts were tested in the photo-bleaching reaction of methyl-orange (MO). The concentration of the catalyst and the methyl orange were 100 mg/350 ml and 10 mg/350 ml, respectively. A Heraus TQ 150 mercury immersion lamp (radiation flux, Φ: 200–600 nm 47 W, strongest spectral lines: 254 nm: 4 W, 313 nm: 4.3 W, 366 nm: 6.4 W, 405 nm: 3.2 W, 436 nm: 4.2 W, 546 nm: 5.1 W, 577 nm: 4.7 W) was applied in a Heraus type cylindrical glass photo-reactor. The experiments were conducted under oxygen bubbling at room temperature assured by the continuous water circulation through the cooling jacket of the lamp. At first the solution was kept in dark for 15 min, to reach the adsorption equilibrium between the catalyst and the dye. Then the lamp was switched on and 3 ml samples were taken in every 30 min using a syringe equipped with filter (Sigma Aldrich, Iso-Disc, PTFE, 0.45 μm pore size membrane). The decline of the MO concentration was followed by a Jasco V-550 type UV-Vis spectrophotometer at 465 nm.

Results and discussion

Formation mechanism of h-WO3 nanowires

The fabrication of nano-sized WO3 by wet-chemical synthesis approaches such as hydrothermal (HT) synthesis is a popularly applied method for cheap and versatile nanostructure formulation.44,55 To direct crystal growth in HT synthesis various structure directing agents were presented in the literature from simple inorganic salts like Na2SO4, NH4NO3 to complex organic structures like oxalic acid or poly-ethylene glycol.33,55,56 To fabricate delicate h-WO3 nanowires in this paper (NH4)2SO4 was used following an earlier report.44 It is believed that the capping agents such as (NH4)2SO4 dictate anisotropic 1D growth by selectively binding to specific crystal facets of the initial crystal nuclei of WO3. The adsorption of capping agents increases the surface energy of the selected facets leading to inhibited crystal growth rates and resulting in the formation of 1D nanostructures.44 It is also known that several other factors may contribute to the morphology and crystal phases of the final nanomaterials such as time, temperature or the pH.31,55,57 This latter one plays an especially important role in the formation of WO3 since H+ is essential to form tungstic acid from the selected tungsten precursor which is typically a tungstate. Then WO3 is formed according to eqn (1).
 
image file: c6ra18899k-t1.tif(1)

Thermal analysis

The role of washing steps in reaching pure h-WO3 nanowires after the microwave hydrothermal reactions was studied by TG/DTA-MS measurement. The results (presented in detail in the ESI) confirmed that washing steps were effective and removed all precursors and reactants. The as-prepared, non-washed, only dried h-WO3 NW samples contained significant amount from not used reagents, i.e. (NH4)2SO4, HCl, and water as well, from which various gaseous products were detected in air, e.g. H2O, NH3, from the decomposition of NH4+, NO and N2O from the combustion of NH3, HCl, SO and SO2 from the decomposition of SO42− (ref. 58) (Table S1). When the sample was annealed, until 100 °C small amount of adsorbed H2O was released in an endothermic reaction (Fig. S1 and S2). Then between 150–300 °C H2O and NH3 evolved accompanied by an endothermic heat effect, whereas between 300–450 °C H2O, NH3, NOx and SOx gases were detected. The combustion of NH3 was an endothermic reaction, which changed the DTA curve to endothermic.59 HCl was released above 450 °C, accompanied by further release of SO2.

After the washing steps (Fig. S3 and S4), all impurities were removed, and accordingly the total mass loss until 900 °C decreased from 18.2% to 5.8%. Now only the release of water was observed in two endothermic reactions between 25–200 and 200–450 °C. The metastable hexagonal WO3 framework was transformed into the thermodynamically stable m-WO3 at 503 °C in an exothermic reaction. The results confirmed the importance of proper washing steps after the microwave hydrothermal reaction in order to obtain pure h-WO3 nanowires. The thermal analysis also showed that the cleaned, pure h-WO3 was stable until 500 °C, therefore performing the ALD reaction at 300 °C was safe.

After the ALD reaction performed at 300 °C-on and 103 Pa vacuum the h-WO3 NWs lost considerable amount of adsorbed and structural water (only 2.5% mass loss) (Fig. S6 and S7). The exothermic DTA peak belonging to the hexagonal–monoclinic WO3 transformation was at 506 °C.

XRD

The XRD patterns (Fig. 1) served information about the crystalline structure of the samples. No crystalline impurities were detected in any of the samples. All reflection peaks of h-WO3 NPs were assigned to pure h-WO3 (ICDD 33-1387). The h-WO3 NWs exhibited well-defined and intense diffraction peaks supposing high crystallinity of the sample. The crystalline phase was confirmed to be pure h-WO3. The XRD pattern of m-WO3 NPs was interpreted as pure monoclinic WO3 (ICDD 43-1035).
image file: c6ra18899k-f1.tif
Fig. 1 Powder diffraction pattern of the m-WO3 NPs, h-WO3 NPs, m-WO3 NWs, and h-WO3/TiO2 NWs samples.

In the diffraction pattern of h-WO3/TiO2 nanocomposite (Fig. 1) only the h-WO3 phase could be identified (ICDD 33-1387). The characteristic reflections of TiO2 could not be detected in the XRD pattern due to the very thin TiO2 layer (aimed to be less than 3 nm).

Raman

The crystalline phases of the as-prepared WO3 nanostructures were confirmed by their Raman spectra as well (Fig. 2), in agreement with the XRD results. The h-WO3 NPs showed well-defined bands characteristic of h-WO3. The most intense peak at 785 cm−1 along with the bands at 650 and 691 cm−1 can be associated with the stretching vibration modes of (O–W–O).60,61 The bands at 263 and 320 cm−1 can be assigned to bending vibrations of W–O–W, whereas at 186 cm−1 the lattice vibration mode can be found.60,62
image file: c6ra18899k-f2.tif
Fig. 2 Raman spectra of the m-WO3 NPs, h-WO3 NPs, m-WO3 NWs, and h-WO3/TiO2 NWs samples.

The Raman spectrum of h-WO3 NWs was substantially different from that of the h-WO3 NPs. One of the deviances was that the h-WO3 NW exhibited overlapping bands in the region of 600 and 850 cm−1. Another considerable difference was that exclusively in the Raman spectra of h-WO3 NWs terminal W[double bond, length as m-dash]O stretching modes were recognizable in the interval of 925–965 cm−1.60,63–65 These peaks are common for all types of WO3 hydrates, and in some cases the appearance of these bands were attributed to surface humidity.43 The absence of these bands in the spectra of h- and m-WO3 NPs can be ascribed to the high temperature treatment at 500 and 650 °C, respectively. Santato et al. reported that the W[double bond, length as m-dash]O stretching modes in the latter wavenumber interval gradually disappeared by applying increasing annealing temperature in the synthesis process.63

The Raman bands of m-WO3 NPs were in good correspondence with literature values.66 The peaks were sharp and well-developed supposing good crystallinity of the sample. The bands positioned at 806 and 720 cm−1 can be assigned to stretching (O–W–O) vibrational modes. The bands at lower wavelength values belong to (O–W–O) deformation modes.60

The Raman bands of h-WO3/TiO2 nanocomposite (Fig. 2) were broadened, compared to h-WO3 NWs. Probably during the ALD deposition some partially reduced W atoms were produced, either due to the conditions (300 °C, 103 Pa vacuum) or to the reaction between the precursors and the substrate, and therefore the structure became less ordered. The Raman spectrum is sensitive to the different oxidation states of the atoms due to the change of the chemical bond strength, which resulted in a small shift of the peak positions involving W atoms of different oxidations states.42 In the h-WO3/TiO2 sample the bands involving completely oxidized (+6) and partially reduced (+5, +4) W atoms overlapped, making the peaks broader. The presence of partially reduced tungsten atoms was also indicated by the color change of the sample (it became blue) and by its optical behavior, discussed later. In addition, the terminal W[double bond, length as m-dash]O stretching modes (above 900 cm−1) of the h-WO3/TiO2 nanocomposite had much lower intensities compared to the substrate h-WO3 NW. The occurred ALD nucleation and the applied temperature and pressure could be accounted for the reduced band intensities.

Characteristic bands of TiO2 could not be observed as the bands of TiO2 usually become perceptible above 10 nm film thickness.45 Similar findings were made by others who found that 1–5% WO3 loading in a WO3/TiO2 nanostructure did not show bands for WO3 due to low content of WO3.67

SEM-EDX, TEM

Fig. 3 shows the SEM and TEM images of the prepared catalyst. It was found that the h-WO3 NPs (Fig. 3a and b) consisted of 50–70 nm nanocrystals. The SEM and TEM images confirmed the formation of uniform and good quality nanowires for the h-WO3 NWs (Fig. 3c and d) being several micrometer long and ca. 5–10 nm thick. On Fig. 3e and f the m-WO3 NPs can be viewed, consisting of irregularly-shaped nanoparticles with a characteristic dimension of 60–90 nm.
image file: c6ra18899k-f3.tif
Fig. 3 TEM and SEM images, respectively, of the prepared samples (a, b) h-WO3 NPs, (c, d) h-WO3 NWs, (e, f) m-WO3 NPs.

The EDX elemental analysis revealed that h-WO3 NWs contained 2.82 atom% of Na beside W and O atoms, which can be attributed to stabilizing Na ions in the hexagonal channels.43

SEM-EDX results were also obtained about the effectiveness of the washing steps. On the SEM image of the as-prepared, non-washed, dried h-WO3 NW sample micrometer scale particles of the not used reactants can be also seen, and EDX analysis confirmed the presence of N, Cl and S, besides O, Na and W (Fig. S5).

In the h-WO3/TiO2 NWs the SEM and TEM images (Fig. 4a and b) revealed that the TiO2 was successfully deposited onto the surface of h-WO3 NWs in the form of individual nanoparticles. The TiO2 nanoparticles had the diameter of around 5–10 nm, which corresponds to double thickness obtained by ellipsometry data, and is rationalized by the 3D growth of the TiO2 particles. This is in agreement with expectations, as in the case of nanoparticle nucleation a 3D ALD growth is taking place. EDX showed 2.64 atom% Ti in the h-WO3/TiO2 sample, and confirmed the successful deposition of TiO2 onto the h-WO3 NWs.


image file: c6ra18899k-f4.tif
Fig. 4 TEM and SEM images of the (a and b) h-WO3/TiO2 NWs catalyst, (c and d) m-WO3/TiO2 samples obtained by annealing h-WO3/TiO2 NWs.

It was also investigated whether an m-WO3/TiO2 composite with nanowire morphology could be obtained by annealing the h-WO3/TiO2 sample. According to SEM and TEM images (Fig. 4c and d), the fine nanowire morphology was lost when h-WO3 transformed into m-WO3. In the m-WO3/TiO2 composite m-WO3 was present in the form of 30–50 nm particles, some of them being connected to each other in a line, pointing to their h-WO3 nanowire origin. The TiO2 could be observed as 10 nm particles on the surface of m-WO3.

Apparent surface area

The BET nitrogen adsorption measurements were carried out to compare the surface area of the synthesized WO3 nanostructures. For h-WO3 NPs it was found to be 11 m2 g−1.68 Whereas the h-WO3 NWs exhibited a much higher surface area of 101 m2 g−1 which is in good agreement with earlier literature results.44 The almost an order of magnitude higher surface area of the h-WO3 NWs relative to the h-WO3 NPs showed the significant effect of the morphology in the surface area which is considered to be an important factor in photocatalysis. The BET surface area of m-WO3 nanostructure was found to be the lowest among the WO3 nanostructures with 6.5 m2 g−1.42

Ellipsometry

The estimation of the equivalent TiO2 layer thickness was determined by ellipsometry on a reference Si wafer. As the native oxide layer on the Si wafer is approximately 3 nm thick, three times more ALD cycle (300 ALD cycle) were run under the same condition on the reference wafer than during the preparation of the nanocomposite in order to make the estimation more accurate. The approximated equivalent TiO2 thickness prepared by 100 ALD cycles was calculated as 3.65 nm.

ALD nucleation determined by the surface properties of WO3 polymorphs

Based on SEM and TEM results, it was found that the TiO2 nucleation followed a distinct pattern on WO3 polymorphs. Earlier it was reported that the ALD-deposited TiO2 formed a continuous layer on the surface of electrospun m-WO3 nanofibers resulting in a core–shell nanocomposite.45 Our results revealed that for h-WO3 nanowires, instead of a continuous layer, the deposited TiO2 formed nanoparticle islands. The surface chemistry of the nanostructures plays a key role in the ALD reaction as the native functional group coverage on the surface serves as first binding sites for the ALD reaction.69–71 Most probably, the distinct ALD nucleation could be accounted for the different surface OH group density of the WO3 polymorphs. Szilágyi et al. reported that based on the XPS spectra the amount of surface OH group was 0.52% for the h-WO3 NPs and 0.70% for m-WO3 NPs.42 The nucleation of TiO2 nanoparticles on the surface of h-WO3 as opposed to continuous film construction, is probably due to the less available nucleation sites.

UV-Vis absorption

All WO3 nanostructures regardless of the crystal phase exhibited pale yellow colour, however it was found that h-WO3/TiO2 NWs nanocomposite turned into blue after the ALD reaction. Solid phase UV/Vis diffuse reflectance (DR) spectra were recorded and analyzed to reveal the optical behavior of the nanostructures and help to understand the reason for the colour differences. The DR spectra is shown in Fig. 5.
image file: c6ra18899k-f5.tif
Fig. 5 UV-Vis diffuse reflectance spectra of h-WO3 NWs, h-WO3 NPs, m-WO3 NPs and h-WO3/TiO2 NWs samples.

It was observed that the h-WO3 NWs and NPs exhibited a sharp drop of the absorption around 460 nm and 478 nm respectively. Similarly, m-WO3 NPs showed an absorption threshold at about 485 nm. These values are typical band edges for nanostructured WO3.56,72 It is believed that h-WO3 nanostructures typically exhibit higher energy band values (therefore appears at lower wavelength) due to structural differences.73

For the h-WO3/TiO2 nanocomposite it was noted that although TiO2 has a typical absorption threshold around 390–410 nm,74 the nanocomposite still exhibited absorption in the whole Vis range. Presumably, the enhanced visible absorption arose from the newly generated in-gap states in the band gap due to the presence of partially reduced W atoms. The reduced W atoms were indicated by Raman spectroscopic results and it is believed to be caused by the high temperature and vacuum used over the course of ALD reaction. The blue appearance of the sample is also an indication of the presence of in-gap states in the far-red region of the visible spectrum.

Photocatalysis

The photo-efficiency of the catalysts was modelled in the photo-bleaching reaction of methyl orange under UV-Vis light irradiation (Fig. 6). The relative absorbance was calculated and plotted against time to follow the dye degradation (see on Fig. 6a). Additionally, first-order kinetics could be used to describe the behavior of our data which can be expressed in the following form:
 
image file: c6ra18899k-t2.tif(2)
where C is the dye concentration, C0 is the initial dye concentration, k is the first-order rate constant and t is time. To test the repeatability of the photocatalytic results multiple experiments were performed. It was found that the experimental error has never exceeded ±4% and was typically 2–3%. The effect of photolysis was negligible under the experimental conditions. The catalysts after the photocatalytic test did not show any visible change e.g. in their color.

image file: c6ra18899k-f6.tif
Fig. 6 (a) Photocatalytic degradation curves and (b) linearized first rate kinetic curves of the m-WO3 NPs, h-WO3 NPs, h-WO3 NWs, and h-WO3/TiO2 NWs samples.

It was found that the nanowire morphology could enhance significantly the photoactivity of h-WO3 NWs. Compared to the reference h-WO3 NPs, the h-WO3 NWs photo-bleached more than twice as much dye within 4 hours. The enhanced efficiency was reflected on the first-order rate constant values as well, exhibiting 1.72 × 10−3 min−1 and 4.93 × 10−4 min−1, respectively. It is generally expected that 1D nanostructures could improve the photocatalytic performance by providing enhanced charge carrier properties and therefore reduced charge recombination.75 However, in some cases it was revealed that other factors such as the oxidation state of the WO3 could significantly reduce the photo-bleaching performance despite the 1D morphology.31 Our results confirmed that there was no such inhibiting factor therefore the nanowire morphology provided an enhanced efficiency over the nanoparticle morphology in the photocatalytic test.

Additionally, the photo-efficiency of the h-WO3 NWs was comparable to that of the m-WO3 NPs showing similar rate constant values of 1.72 × 10−3 min−1 and 1.79 × 10−3 min−1, respectively. The monoclinic phase WO3 was reported to exhibit higher photo-efficiency relative to hexagonal WO3 due to the more oxidized composition,42 which was confirmed by our results when h-WO3 and m-WO3 NPs with similar morphologies were compared. Clearly, beside the prominent effect of the crystal phase, the morphology could play an important role in the determination of the photo-efficiency. The highest photo-degradation was presented by the h-WO3 NW/TiO2 nanocomposite. It decomposed 56% of the original dye concentration by the end of the 4 hour reaction, which represents a higher than four times improvement relative to the h-WO3 NPs and almost two times better performance relative to the h-WO3 NWs. Therefore, the highest rate constant of 4.22 × 10−3 min−1 was calculated for the h-WO3 NW/TiO2 nanocomposite. The enhanced photoactivity can be attributed to the better light utilization due to absorption in the complete UV-Vis spectrum, and reduced recombination of the photo-generated charges through effective charge separation between h-WO3 and TiO2. In the literature it was earlier reported that m-WO3–TiO2 nanocomposite prepared by depositing TiO2 onto nanofibers of m-WO3 by ALD showed improved activities relative to bare WO3 and TiO2.45 However, successful ALD reaction onto h-WO3 NWs to prepare h-WO3 NW/TiO2 nanocomposite has not yet been presented to the best of our knowledge. The advantage of coupling WO3 with TiO2 in various crystal phases and morphologies was shown to have a great potential in improving photocatalytic performances.35,76,77 This was not straightforward, since previously it was observed that in the case of blue h-WO3 samples with partially reduced W atoms the photocatalytic activity was significantly lower, even compared to yellow h-WO3 NPs.42 The reason could be that although the blue h-WO3 had absorption in the complete UV-Vis spectrum, this was overcome by that the partially reduced W atoms served as recombination centers for photo-generated electrons and holes.

Nevertheless, in the case of the h-WO3/TiO2 NW sample, the addition of TiO2 strongly influenced the photocatalytic activity, and resulted in a more effective photocatalyst compared to pure h-WO3 or m-WO3. It clearly shows the efficiency of ALD to reprogramme the surface properties of nanostructures by depositing nanolayers or nanoparticles on them, and thus to prepare photocatalysts with superior properties.

Mechanism of photocatalytic activities

It was revealed that m-WO3 nanostructures typically overperform h-WO3 nanostructures.42 It is believed that structural differences play a key role in that. Photoluminescence studies (PL) confirmed that although h-WO3 and m-WO3 has similar emission pattern, h-WO3 typically shows lower intensities in the luminescence spectrum.42 This was previously attributed to the partially reduced state of h-WO3 compared to m-WO3 which arose from the presence of stabilizing cations in its hexagonal crystal structure. Between the reduced W atoms polaron transition could take place induced by the absorbed light.78 Due to this, fewer light portion can actually play role in the excitation of electrons from the valence band to the conduction band of the h-WO3 which explains the somewhat limited photocatalytic activities of h-WO3 relative to m-WO3.

One approach to mitigate e–h+ pair recombination in single nanostructures is to form heterojunctions. In the event of beneficially aligned energy levels, the photo-generated charges can accumulate on different semiconductors which was found to produce longer-lived charges.26 For a WO3/TiO2 system the electrons would favor to accumulate on the conduction band of WO3 whereas holes would be preferably injected from the valence band of WO3 to the valence band of TiO2. This way the separation of the photo-generated charges can take place which supports photocatalytic surface reactions to take place. PL studies were also applied earlier to confirm the effective charge-separation in WO3–TiO2 systems.67 It was found that the emission intensities for the nanocomposite versus e.g. TiO2 was much lowered which was explained by the improved charge separation in the nanocomposite material.

Conclusions

In this paper h-WO3 has been prepared by controlled annealing of (NH4)xWO3 and by microwave hydrothermal synthesis to produce nanoparticles and nanowires, respectively. The two distinct morphologies were tested in photocatalytic application against m-WO3 nanoparticles prepared also from (NH4)xWO3. The h-WO3 and m-WO3 NPs were pure and consisted of 50–70 nm and 70–90 nm particles, respectively. The h-WO3 NWs were single crystalline, 5–10 nm thick and several micrometer long. It was found that the h-WO3 NWs had double as high photocatalytic activity compared to h-WO3 NPs, and their performance was similar to m-WO3 NPs.

The h-WO3 nanowires were then used as substrates for TiO2 deposition by atomic layer deposition to study the nucleation characteristics of TiO2 on hexagonal phase WO3 for the first time. It was shown that the nucleation on the surface of h-WO3 was substantially different from that of the m-WO3. When TiO2 was deposited by ALD onto h-WO3 NWs, it did not form a continuous layer as on the surface of m-WO3, but rather 5–10 nm particles. The unique nucleation pattern was understood by the lower surface OH density of h-WO3 compared to m-WO3 which did not serve dense enough nucleation sites for the ALD reactions to occur in a continuous manner. It was also presented that in contrast to yellow h-WO3 NWs, the h-WO3/TiO2 NW turned to blue, due to the appearance of partially reduced W atoms. Our study is the first investigation of ALD nucleation on various WO3 polymorphs and also the first example that different crystalline modifications of the same oxide material have so diverse effect on ALD growth.

When TiO2 was grown on h-WO3 NWs, it significantly increased the photocatalytic degradation rate, and resulted in a 65% increase in photocatalytic performance. Our results show that ALD is an outstanding tool to prepare composite photocatalysts.

Acknowledgements

D. Nagy acknowledges the Principal Career Development Scholarship awarded by the University of Edinburgh. I. M. Szilágyi acknowledges a János Bolyai Research Fellowship of the Hungarian Academy of Sciences and an OTKA-PD-109129 grant. Z. Baji, A. L. Tóth are acknowledged for contributing to the ALD depositions, ellipsometry and SEM-EDX measurements (Hungarian Academy of Sciences, Centre for Energy Research, Institute of Technical Physics and Materials Science, Budapest, Hungary).

References

  1. H. Zhu and T. Lian, Energy Environ. Sci., 2012, 5, 9406–9418 CAS.
  2. R. Das, S. B. A. Hamid, M. E. Ali, A. F. Ismail, M. S. M. Annuar and S. Ramakrishna, Desalination, 2014, 354, 160–179 CrossRef CAS.
  3. H. Tong, S. X. Ouyang, Y. P. Bi, N. Umezawa, M. Oshikiri and J. H. Ye, Adv. Mater., 2012, 24, 229–251 CrossRef CAS PubMed.
  4. J. Byrne, P. Dunlop, J. Hamilton, P. Fernández-Ibáñez, I. Polo-López, P. Sharma and A. Vennard, Molecules, 2015, 20, 5574 CrossRef CAS PubMed.
  5. N. Zhou, V. Lopez-Puente, Q. Wang, L. Polavarapu, I. Pastoriza-Santos and Q.-H. Xu, RSC Adv., 2015, 5, 29076–29097 RSC.
  6. W. Wu, J. Changzhong and V. A. L. Roy, Nanoscale, 2015, 7, 38–58 CAS.
  7. W. Zhong, D. Li, S. Jin, W. Wang and X. Yang, Appl. Surf. Sci., 2015, 356, 1341–1348 CrossRef CAS.
  8. S. Yurdakal, V. Loddo, V. Augugliaro, H. Berber, G. Palmisano and L. Palmisano, Catal. Today, 2007, 129, 9–15 CrossRef CAS.
  9. S. Malato, P. Fernández-Ibáñez, M. I. Maldonado, J. Blanco and W. Gernjak, Catal. Today, 2009, 147, 1–59 CrossRef CAS.
  10. X. B. Chen, L. Liu, P. Y. Yu and S. S. Mao, Science, 2011, 331, 746–750 CrossRef CAS PubMed.
  11. G. Halasi, I. Ugrai and F. Solymosi, J. Catal., 2011, 281, 309–317 CrossRef CAS.
  12. M. Gratzel, Nature, 2001, 414, 338–344 CrossRef CAS PubMed.
  13. Q. Gu, J. L. Long, Y. G. Zhou, R. S. Yuan, H. X. Lin and X. X. Wang, J. Catal., 2012, 289, 88–99 CrossRef CAS.
  14. A. Fujishima, X. T. Zhang and D. A. Tryk, Surf. Sci. Rep., 2008, 63, 515–582 CrossRef CAS.
  15. O. Czakkel, E. Geissler, I. M. Szilágyi and L. Krisztina, 2013, 1, 23–30.
  16. S. Banerjee, D. D. Dionysiou and S. C. Pillai, Appl. Catal., B, 2015, 176–177, 396–428 CrossRef CAS.
  17. Y. Zhang, Z. Jiang, J. Huang, L. Y. Lim, W. Li, J. Deng, D. Gong, Y. Tang, Y. Lai and Z. Chen, RSC Adv., 2015, 5, 79479–79510 RSC.
  18. K. Teramura, T. Tanaka, M. Kani, T. Hosokawa and T. Funabiki, J. Mol. Catal. A: Chem., 2004, 208, 299–305 CrossRef CAS.
  19. T. Arai, M. Yanagida, Y. Konishi, Y. Iwasaki, H. Sugihara and K. Sayama, Catal. Commun., 2008, 9, 1254–1258 CrossRef CAS.
  20. Z.-G. Zhao and M. Miyauchi, Angew. Chem., Int. Ed., 2008, 47, 7051–7055 CrossRef CAS PubMed.
  21. D. S. Bohle and C. J. Spina, J. Am. Chem. Soc., 2009, 131, 4397–4404 CrossRef CAS PubMed.
  22. K. Tang, J. Zhang, W. Yan, Z. Li, Y. Wang, W. Yang, Z. Xie, T. Sun and H. Fuchs, J. Am. Chem. Soc., 2008, 130, 2676–2680 CrossRef CAS PubMed.
  23. G. Panthi, M. Park, H.-Y. Kim, S.-Y. Lee and S.-J. Park, J. Ind. Eng. Chem., 2015, 21, 26–35 CrossRef CAS.
  24. W. W. Zhong, Y. F. Lou, S. F. Jin, W. J. Wang and L. W. Guo, Sci. Rep., 2016, 6, 1–7 Search PubMed.
  25. Z.-F. Huang, J. Song, L. Pan, X. Zhang, L. Wang and J.-J. Zou, Adv. Mater., 2015, 27, 5309–5327 CrossRef CAS PubMed.
  26. Y. J. Wang, Q. S. Wang, X. Y. Zhan, F. M. Wang, M. Safdar and J. He, Nanoscale, 2013, 5, 8326–8339 RSC.
  27. R. Bogue, Sens. Rev., 2014, 34, 1–8 CrossRef.
  28. D. L. Liao, C. A. Badour and B. Q. Liao, J. Photochem. Photobiol., A, 2008, 194, 11–19 CrossRef CAS.
  29. H. Y. Wang, Y. Yang, X. Li, L. J. Li and C. Wang, Chin. Chem. Lett., 2010, 21, 1119–1123 CrossRef CAS.
  30. M. R. Gholipour, C.-T. Dinh, F. Beland and T.-O. Do, Nanoscale, 2015, 7, 8187–8208 RSC.
  31. D. Nagy, D. Nagy, I. M. Szilágyi and X. Fan, RSC Adv., 2016, 6, 33743–33754 RSC.
  32. F. Amano, E. Ishinaga and A. Yamakata, J. Phys. Chem. C, 2013, 117, 22584–22590 CAS.
  33. Y. Xia, P. Yang, Y. Sun, Y. Wu, B. Mayers, B. Gates, Y. Yin, F. Kim and H. Yan, Adv. Mater., 2003, 15, 353–389 CrossRef CAS.
  34. T. Firkala, B. Fórizs, E. Drotár, A. Tompos, A. Tóth, K. Varga-Josepovits, K. László, M. Leskelä and I. M. Szilágyi, Catal. Lett., 2014, 144(5), 831–836 CrossRef CAS.
  35. F. Wang, X. Chen, X. Hu, K. S. Wong and J. C. Yu, Sep. Purif. Technol., 2012, 91, 67–72 CrossRef CAS.
  36. M. Shu-Quian, Z. De-Feng, Z. Xiao-Fei, Y. Guo-Cheng and L. Zhao-Hui, J. Inorg. Mater., 2014, 29, 605–613 Search PubMed.
  37. B. Lu, X. Li, T. Wang, E. Xie and Z. Xu, J. Mater. Chem. A, 2013, 1, 3900–3906 CAS.
  38. P. Guo, L. T. Meng and C. H. Wang, Adv. Mater. Res., 2013, 850–851, 78–81 CrossRef.
  39. E. Kamali Heidari, E. Marzbanrad, C. Zamani and B. Raissi, Nanoscale Res. Lett., 2010, 5, 370–373 CrossRef PubMed.
  40. H. Long, W. Zeng and H. Zhang, J. Mater. Sci.: Mater. Electron., 2015, 26, 4698–4707 CrossRef CAS.
  41. I. M. Szilágyi, E. Santala, M. Heikkilä, M. Kemell, T. Nikitin, L. Khriachtchev, M. Räsänen, M. Ritala and M. Leskelä, J. Therm. Anal. Calorim., 2011, 105, 73–81 CrossRef.
  42. I. M. Szilágyi, B. Fórizs, O. Rosseler, Á. Szegedi, P. Németh, P. Király, G. Tárkányi, B. Vajna, K. Varga-Josepovits, K. László, A. L. Tóth, P. Baranyai and M. Leskelä, J. Catal., 2012, 294, 119–127 CrossRef.
  43. I. M. Szilágyi, J. Madarász, G. Pokol, P. Király, G. Tárkányi, S. Saukko, J. Mizsei, A. L. Tóth, A. Szabó and K. Varga-Josepovits, Chem. Mater., 2008, 20, 4116–4125 CrossRef.
  44. A. Phuruangrat, D. J. Ham, S. J. Hong, S. Thongtem and J. S. Lee, J. Mater. Chem., 2010, 20, 1683–1690 RSC.
  45. I. M. Szilágyi, E. Santala, M. Heikkilä, V. Pore, M. Kemell, T. Nikitin, G. Teucher, T. Firkala, L. Khriachtchev, M. Räsänen, M. Ritala and M. Leskelä, Chem. Vap. Deposition, 2013, 19(4–6), 149–155 CrossRef.
  46. V. Miikkulainen, M. Leskela, M. Ritala and R. L. Puurunen, J. Appl. Phys., 2013, 113, 021301 CrossRef.
  47. M. N. Liu, X. L. Li, S. K. Karuturi, A. I. Y. Tok and H. J. Fan, Nanoscale, 2012, 4, 1522–1528 RSC.
  48. I. M. Szilágyi and D. Nagy, J. Phys.: Conf. Ser., 2014, 559, 012010 CrossRef.
  49. H. Kim, H. B. R. Lee and W. J. Maeng, Thin Solid Films, 2009, 517, 2563–2580 CrossRef CAS.
  50. I. M. Szilágyi, G. Teucher, E. Härkönen, E. Färm, T. Hatanpää, T. Nikiti, L. Khriachtchev, M. Räsänen, M. Ritala and M. Leskelä, Nanotechnology, 2013, 24, 245701 CrossRef PubMed.
  51. M. Knez, K. Nielsch and L. Niinistö, Adv. Mater., 2007, 19, 3425–3438 CrossRef CAS.
  52. S. Boyadjiev, V. Georgieva, L. Vergov, B. Zs, F. Gáber and I. M. Szilágyi, J. Phys.: Conf. Ser., 2014, 559, 012013 CrossRef.
  53. I. M. Szilágyi, J. Pfeifer, C. Balázsi, A. L. Tóth, K. Varga-Josepovits, J. Madarász and G. Pokol, J. Therm. Anal. Calorim., 2008, 94, 499–505 CrossRef.
  54. I. M. Szilágyi, S. Saukko, J. Mizsei, A. L. Tóth, J. Madarász and G. Pokol, Solid State Sci., 2010, 12, 1857–1860 CrossRef.
  55. V. B. Patil, P. V. Adhyapak, S. S. Suryavanshi and I. S. Mulla, J. Alloys Compd., 2014, 590, 283–288 CrossRef CAS.
  56. T. Peng, D. Ke, J. Xiao, L. Wang, J. Hu and L. Zan, J. Solid State Chem., 2012, 194, 250–256 CrossRef CAS.
  57. J. Yu, L. Qi, B. Cheng and X. Zhao, J. Hazard. Mater., 2008, 160, 621–628 CrossRef CAS PubMed.
  58. D. Hunyadi, A. V. M. Ramos and I. M. Szilágyi, J. Therm. Anal. Calorim., 2015, 120, 209–215 CrossRef CAS.
  59. D. Hunyadi, I. Sajó and I. M. Szilágyi, J. Therm. Anal. Calorim., 2014, 116, 329–337 CrossRef CAS.
  60. A. Sonia, Y. Djaoued, B. Subramanian, R. Jacques, M. Eric, B. Ralf and B. Achour, Mater. Chem. Phys., 2012, 136, 80–89 CrossRef CAS.
  61. I. M. Szilágyi, S. Saukko, J. Mizsei, P. Király, G. Tárkányi, A. Tóth, A. Szabó, K. Varga-Josepovits, J. Madarász and G. Pokol, Mater. Sci. Forum, 2008, 589, 161–165 CrossRef.
  62. I. M. Szilágyi, L. Wang, P.-I. Gouma, C. Balázsi, J. Madarász and G. Pokol, Mater. Res. Bull., 2009, 44, 505–508 CrossRef.
  63. C. Santato, M. Odziemkowski, M. Ulmann and J. Augustynski, J. Am. Chem. Soc., 2001, 123, 10639–10649 CrossRef CAS PubMed.
  64. A. Rey, P. García-Muñoz, M. D. Hernández-Alonso, E. Mena, S. García-Rodríguez and F. J. Beltrán, Appl. Catal., B, 2014, 154–155, 274–284 CrossRef CAS.
  65. I. M. Szilágyi, I. Sajó, P. Király, G. Tárkányi, A. Tóth, A. Szabó, K. Varga-Josepovits, J. Madarász and G. Pokol, J. Therm. Anal. Calorim., 2009, 98, 707–716 CrossRef.
  66. J. Zhang, W. Zhang, Z. Yang, Z. Yu, X. Zhang, T. C. Chang and A. Javey, Sens. Actuators, B, 2014, 202, 708–713 CrossRef CAS.
  67. H. Shen, I.-R. Ie, C.-S. Yuan and C.-H. Hung, Appl. Catal., B, 2016, 195, 90–103 CrossRef CAS.
  68. D. Nagy, T. Firkala, X. Fan and I. M. Szilágyi, Eur. Chem. Bull., 2016, 5, 40–42 Search PubMed.
  69. C. S. Hwang and A. Y. Yoo, Atomic Layer Deposition for Semiconductors, Springer, New York, Heidelberg, Dordrecht, London, 2014 Search PubMed.
  70. N. Pinna and M. Knez, Atomic Layer Deposition of Nanostructured Materials, Wiley-VCH Verlag & Co. KGaA, Weinheim, Germany, 2011 Search PubMed.
  71. M. Ritala and M. Leskelä, in Handbook of Thin Films, ed. H. S. Nalwa, Academic Press, Burlington, 2002, pp. 103–159,  DOI:10.1016/B978-012512908-4/50005-9.
  72. S. Adhikari and D. Sarkar, RSC Adv., 2014, 4, 20145–20153 RSC.
  73. J. Su, X. Feng, J. D. Sloppy, L. Guo and C. A. Grimes, Nano Lett., 2010, 11, 203–208 CrossRef PubMed.
  74. G. Liu, L.-C. Yin, J. Wang, P. Niu, C. Zhen, Y. Xie and H.-M. Cheng, Energy Environ. Sci., 2012, 5, 9603–9610 CAS.
  75. H. Dong, Z. Wu, F. Lu, Y. Gao, A. El-Shafei, B. Jiao, S. Ning and X. Hou, Nano Energy, 2014, 10, 181–191 CrossRef CAS.
  76. J. K. Yang, X. T. Zhang, H. Liu, C. H. Wang, S. P. Liu, P. P. Sun, L. L. Wang and Y. C. Liu, Catal. Today, 2013, 201, 195–202 CrossRef CAS.
  77. L. Yang, Z. Si, D. Weng and Y. Yao, Appl. Surf. Sci., 2014, 313, 470–478 CrossRef CAS.
  78. S. H. Lee, H. M. Cheong, C. E. Tracy, A. Mascarenhas, D. K. Benson and S. K. Deb, Electrochim. Acta, 1999, 44, 3111–3115 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra18899k

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.