Next Article in Journal
Plasma-Enhanced Chemical Looping Oxidative Coupling of Methane through Synergy between Metal-Loaded Dielectric Particles and Non-Thermal Plasma
Next Article in Special Issue
Exceptional Photocatalytic Performance of the LaFeO3/g-C3N4 Z-Scheme Heterojunction for Water Splitting and Organic Dyes Degradation
Previous Article in Journal
Lignin Valorization: Production of High Value-Added Compounds by Engineered Microorganisms
Previous Article in Special Issue
Recent Advances in g-C3N4-Based Photocatalysts for NOx Removal
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Vertical Growth of WO3 Nanosheets on TiO2 Nanoribbons as 2D/1D Heterojunction Photocatalysts with Improved Photocatalytic Performance under Visible Light

1
College of Chemistry and Life Science, Zhejiang Normal University, Jinhua 321004, China
2
School of Biology and Chemical Engineering, Jiaxing University, Jiaxing 314001, China
*
Author to whom correspondence should be addressed.
Catalysts 2023, 13(3), 556; https://doi.org/10.3390/catal13030556
Submission received: 12 January 2023 / Revised: 6 March 2023 / Accepted: 7 March 2023 / Published: 9 March 2023
(This article belongs to the Special Issue Advances in Heterojunction Photocatalysts)

Abstract

:
We report the construction of 2D/1D heterojunction photocatalysts through the hydrothermal growth of WO3 nanosheets on TiO2 nanoribbons for the first time. Two-dimensional WO3 nanosheets were vertically arrayed on the surface of TiO2 nanoribbons, and the growth density could be simply controlled by adjusting the concentration of the precursors. The construction of WO3/TiO2 heterojunctions not only decreases the band gap energy of TiO2 from 3.12 to 2.30 eV and broadens the photoresponse range from the UV region to the visible light region but also significantly reduces electron–hole pair recombination and enhances photo-generated carrier separation. Consequently, WO3/TiO2 heterostructures exhibit improved photocatalytic activity compared to pure WO3 nanosheets and TiO2 nanoribbons upon visible light irradiation. WO3/TiO2-25 possesses the highest photocatalytic activity and can remove 92.8% of RhB pollutants in 120 min. Both further increase and decrease in the growth density of WO3 nanosheets result in an obvious reduction in photocatalytic activity. The kinetic studies confirmed that the photocatalytic degradation of RhB follows the kinetics of the pseudo-first-order model. The present study demonstrates that the prepared WO3/TiO2 2D/1D heterostructures are promising materials for photocatalytic removal of organic pollutants to produce clean water.

Graphical Abstract

1. Introduction

The demand for fresh water increases exponentially with global population growth and industrial development [1,2,3]. Clean water scarcity is one of the most pressing issues for human beings to confront nowadays, especially in developing countries [4]. According to World Wildlife, two-thirds of the world’s population will face water shortages by 2025 [5]. To address this challenge, many advanced technologies have been developed to remove aqueous pollutants for clean water production [6,7], including adsorption [8,9], membrane filtration [10], and semiconductor photocatalytic technology [11]. Among various techniques, semiconductor photocatalysis offers a green and sustainable water purification technology because harmful pollutants can be completely decomposed to CO2, H2O, and other small molecules without the involvement of any chemicals [11,12,13,14]. Mainly, photocatalysts absorb photons as an energy source to (1) generate electron−hole pairs, (2) separate the photoexcited charges, (3) transfer electron and hole to the photocatalyst’s surface, and (4) utilize the active charges on the photocatalyst’s surface for catalytic decomposition of pollutants [12,13,14,15,16,17].
Heterojunction photocatalysts, particularly systems consisting of a low-dimensional 1D and 2D semiconductor, are recognized as prospective building blocks for next-generation advanced photon harvesting and conversion technologies [18,19]. The considerable potential of 1D/2D heterojunctions arises from high charge carrier mobility along the 1D nanostructures and effective charge recombination prevention ability of 2D nanostructures due to the strong electron confinement effect in atomic-thick layers. Among four different configurations based on interfacial contact, heterojunctions with vertically aligned 2D nanosheets on 1D nanostructures are highly desirable for photocatalytic systems [19]. That is because the vertically aligned 2D nanosheets on 1D nanostructures not only allow maximum exposure of the entire surface, thereby maximally providing the edge active sites, but also create better electronic contacts and optimized electron transport pathways, leading to enhanced photo-generated electron–hole separation efficiency [18,19]. This unique 2D array on 1D structures widely exists in nature, such as leaf array on a tree branch and the bony plates along stegosaur backs, and their advancement in light absorption and conversion have been proven in nature for millions of years.
Since the discovery of the Honda–Fujishima effect in 1972 [20], TiO2 has been considered to be the most promising photocatalyst because of its high photosensitivity, excellent stability, nontoxic nature, and low cost [13,21,22]. In particular, 1D TiO2 nanoribbons have received much more attention for advanced photocatalyst design because of the following features [22,23,24]. First, the well-defined 1D geometry facilitates fast and long-distance electron transport, leading to long-time photocatalytic stability. Second, 1D nanoribbons possess larger specific surface areas than corresponding bulk materials, providing more active catalytic sites. Third, the large length-to-diameter ratio of 1D nanoribbons increases light absorption and scattering, enhancing light use efficiency [19,20]. Beside TiO2, WO3 is another extensively investigated photocatalyst driven by its narrow band gap and visible light absorption ability [25,26,27]. Among various morphologies, 2D WO3 nanosheets possess an extremely high percentage of exposed surfaces and a strong quantum confinement effect in the thinnest dimension, leading to abundant active sites and enhanced light conversion efficiency, which are highly desirable for high-performance photocatalysts [25,26,27,28,29,30]. Inspired by the booming progress in 1D TiO2 nanoribbon and 2D WO3 nanosheet photocatalysts, the rational design of 2D/1D WO3/TiO2 multidimensional heterojunction photocatalysts is expected to integrate the merits of both 1D and 2D nanogeometry and lead to enhanced photocatalytic performance. Particularly, a 2D WO3 array on a 1D TiO2 heterojunction is of special interest for highly efficient photocatalyst design. Coupling narrow band gap WO3 (2.4–2.8 eV) can not only broaden the absorption wavelength of TiO2 from UV to visible light regions, resulting in the formation of visible light-responded heterojunction photocatalysts, but also effectively suppresses electron−hole pair recombination, thus improving the photocatalytic activity [31]. To date, substantial efforts have been devoted to construct WO3/TiO2 heterojunction systems, including WO3 nanorod/TiO2 nanofiber composite [32], WO3 nanoparticle/TiO2 nanoparticle composite [33,34], heterostructured TiO2/WO3 porous microspheres [35,36], a WO3 nanoparticle/TiO2 nanotubes system [37,38,39], heterostructured WO3/TiO2 nanosheets [40,41], and WO3 nanosheet/TiO2 nanoparticle composite [42]. However, to the best of our knowledge, 2D/1D heterojunction photocatalysts composed of TiO2 nanoribbons and WO3 nanosheets have not been reported. Demonstrating the vertical growth of WO3 nanosheets on TiO2 nanoribbons for 2D/1D heterojunction photocatalysts is especially challenging.
Herein, we report the vertical growth of WO3 nanosheets on TiO2 nanoribbons as 1D/2D heterojunction photocatalysts with improved photocatalytic performance under visible light. The WO3/TiO2 heterojunctions were constructed through the introduction of TiO2 nanoribbons into the hydrothermal growth system of WO3 nanosheets. The as-obtained WO3/TiO2 sample was characterized by using XRD, SEM, TEM, XPS, UV–vis, and PL analysis. WO3 nanosheets vertically arrayed on the surface of TiO2 nanoribbons lead to the formation of 1D/2D heterojunctions with maximum exposure of the entire surface, which not only broadens the light adsorption spectrum into the visible region but also inhibits the recombination of photoinduced carriers, resulting in enhanced photocatalytic degradation of aqueous pollutants.

2. Results and Discussion

2.1. XRD Analysis

The hydrothermal growth process is accompanied by an obvious color change in TiO2 from white to yellow-green (Figure 1a), indicating the successful and uniform growth of WO3 on TiO2. In order to confirm WO3 growth, XRD patterns of the samples before and after hydrothermal reaction were recorded (Figure 1b). Before growth of WO3 nanosheets, all the diffraction peaks located at 25.3°, 37.8°, 48.0°, 53.9°, 55.0°, 62.1°, 62.7°, and 68.8° can be indexed to anatase TiO2 (JCPDS file no. 21-1272), which is in good agreement with previous reports [43,44]. After the growth of the WO3 nanosheets, except for the peaks originating from TiO2 nanoribbons (marked with red boxes), new peaks (marked with blue circles) centered at 16.5°, 25.6°, 30.5°, 33.4°, 34.1°, 34.8°, 38.9°, 44.1°, 45.9°, 46.3°, 49.6°, 52.6°, 56.2°, 57.2°, 58.4°, 61.2°, 64.3°, and 66.0° were observed, which match well with WO3•H2O (JCPDS file no. 43-0679) [45], indicating the formation of WO3/TiO2-25 composites [32,33,34].

2.2. FTIR Analysis

In order to further confirm the formation of WO3/TiO2 composites, FTIR spectra of all samples have been collected and shown in Figure 1c. For the spectrum of pure TiO2 nanoribbons (green line), the broad adsorption band centered at 3430 cm−1 could be ascribed to the stretching vibrations of the surface hydroxyl groups and molecularly chemisorbed water [46]. The peak at 1630 cm−1 results from -OH bending of molecularly physisorbed water [46]. The large absorption in the range 492 cm−1 can be attributed to the strong stretching vibrations of the Ti-O bond in TiO2 nanoribbons [46]. For the spectrum of pure WO3 nanosheets (red line), the broad adsorption band center at 3410 cm−1 is related to the stretch region of the surface hydroxyl groups with hydrogen bonds and crystal water of WO3•H2O, while the peak centered at 1633 cm−1 is from O–H bending of physisorbed water. The weak band observed around 600–750 cm−1 is attributed to the O–W–O stretching modes of WO3, and the peak located at 941 cm−1 is associated with W=O stretching [46]. For WO3/TiO2 composites with different ratios, the broad peak around 3412–3420 cm−1 of each sample can be ascribed to the stretching vibrations of -OH groups [46]. These characteristic absorption peaks of WO3 and TiO2 located in the range of 400–1000 cm−1 are all observed in the spectra of WO3/TiO2 composites, indicating the successful combination of WO3 and TiO2. From WO3/TiO2-5 to WO3/TiO2-75, the intensity of peaks from TiO2 nanoribbons decreases, and the intensity of peaks ascribed to WO3 nanosheets increases, indicating the increasing growth density of WO3 nanosheets on TiO2 nanoribbons, which is in good agreement with the raw material proportioning.

2.3. SEM and TEM Analysis

In order to reveal the configuration of WO3/TiO2 heterojunctions, SEM and TEM were applied to observe the pure TiO2 nanoribbons and the WO3/TiO2-25 heterostructure. Before WO3 growth, pure TiO2 nanoribbons exhibit typical 1D morphology with a length from several tens to several hundreds of micrometers (Figure 2a). TEM images (Figure 2b–d) of an individual nanoribbon show that the width of the nanoribbons is around 100 nm and the surface of the nanoribbons are flat and clean, which provides comfortable platforms for the nucleation and growth of WO3 nanosheets. After WO3 growth, the 1D typical morphology is still maintained, while the surface of the nanoribbons is vertically arrayed with tetragonal nanosheets at a length of 100 nm to 400 nm and thickness of 40 nm (Figure 2e). A similar structure has also been reported in BiOBr/TiO2 systems [43]. A TEM image (Figure 2f) of a single heterostructured nanoribbon demonstrates that 2D nanosheets were grown on both sides of the nanoribbons, leading to the formation of vertically aligned 2D/1D heterostructures with maximum exposure of junctions and the entire surface. High-magnification TEM images (Figure 2g,h) show very clear crystalline planes with a d-spacing of 0.347 nm, which closely matched the inter-planar spacing of the (111) facets of WO3 [45], further confirming the successful formation of WO3/TiO2 heterojunctions. Figure 2i–l show HAADF-STEM-EDS mapping images of a typical nanoribbon arrayed with sheet structures. Ti and O elements can be found in the nanoribbons, while W and O elements existed in the nanosheets, which directly proves that the nanoribbons were TiO2 and the nanosheets arrayed on nanoribbons were WO3. These results visually verified that the flat TiO2 nanoribbons served as the backbones and the WO3 nanosheets with a tetragonal profile were arrayed on them, leading to the formation of typical vertically aligned 2D on 1D heterostructures. This unique 2D array on 1D structures widely exist in nature and their advancement in light absorption and conversion have been proved in nature for millions of years. For example, trees adopt a leaf array on branch structures to obtain enough access to maximum light. Stegosaurs possessed bony plates along their backs to absorb more heat from the sun to warm their blood on cool days. Therefore, 2D WO3/1D TiO2 heterostructures are anticipated to be highly desired for the design of advanced photocatalytic systems.
The growth density of WO3 nanosheets can be regulated simply via control over the precursor concentration. When the concentration of the precursors was decreased to 5 mM, few nanosheets could be observed on the surfaces of the nanoribbons (Figure S1a). At 10 mM precursor concentration, only a small amount of nanoribbons was decorated with WO3 nanosheets (Figure S1b). On increasing the precursor concentration to 50 mM, a much higher density of nanosheets could be grown on the nanoribbon backbones, and some isolate nanosheets (Figure S1c) were found between nanoribbons. Further increasing the precursor concentration to 75 mM would lead to a large amount of free nanosheets (Figure S1d), indicating the formation of the mixture of WO3/TiO2 heterojunctions and WO3 nanosheets (Figure S2).

2.4. XPS Analysis

The element composition and valence of WO3/TiO2-25 composite were analyzed by X-ray photoelectron spectroscopy (XPS). Figure 3a shows the Ti 2p spectra of pure TiO2 nanoribbons and WO3/TiO2-25 heterojunctions. The pure TiO2 nanoribbons exhibit two XPS peaks at 458.6 eV and 464.3 eV, which correspond to Ti 2p1/2 and Ti 2p3/2, respectively. The energy gap between Ti 2p1/2 and Ti 2p3/2 peaks is 5.7 eV, revealing the oxidation state of Ti in TiO2 nanoribbons is +4 [36,47], which is also consistent with the XRD result. After the growth of WO3 nanosheets on TiO2, both the Ti 2p1/2 and Ti 2p3/2 peaks exhibit obvious red shifts from 458.6 to 459.1 eV and from 464.3 to 464.8 eV, respectively. This result suggests the conversion of Ti-O-Ti bonds to Ti-O-W bonds, which confirms the strong interaction between WO3 and TiO2 in the WO3/TiO2-25 heterojunctions [34,36]. Furthermore, the +4 oxidation state of Ti is unchanged after the formation of WO3/TiO2-25 heterojunctions as the energy gap between the peaks at 459.1 and 464.8 eV remains 5.7 eV. The W 4f XPS spectra of pure WO3 nanosheets and WO3/TiO2-25 heterojunctions are shown in Figure 3b. The two XPS peaks of pure WO3 nanosheets at 35.7 eV and 37.8 eV are attributed to W 4f7/2 and W 4f5/2, respectively. The energy gap between W 4f7/2 and W 4f5/2 peaks is 2.1 eV, which suggests that the W element existed in the form of W6+ in WO3 nanosheets [34,36]. For WO3/TiO2-25 heterojunctions, the binding energies of W 4f7/2 and W 4f5/2 peaks were both shifted to low energy by 0.1 eV, while the energy gap between these two peaks is unchanged, indicating the W6+ oxidation state in WO3/TiO2-25 heterojunctions [34,36,48], which is in good agreement with the XRD data. The high-resolution O 1s XPS spectrum of the samples is shown in Figure 3c,d. It can be seen that the Ti-O bond peak in TiO2 nanoribbons is located at 530.0 eV (Figure 3c), while the W-O bond peak in WO3 nanosheets is centered at 530.5 eV (Figure 3d). For WO3/TiO2-25 heterojunctions, the binding energy at 530.4 eV corresponded to the lattice oxygen of Ti4+-O or W6+-O, indicating that W-O and Ti-O shared the orbital O 1s in the W-O-Ti bond [32,43,44]. The peak exhibited a 0.4 eV up-shift from Ti-O bonds (from 530.0 to 530.4 eV) and 0.1 eV down-shift from W-O bonds (from 530.5 to 530.4 eV). The above XPS results show that the binding energy of Ti 2p shifted to high energy, while the binding energy of W 4f and O 1s shifted to low energy after the formation of WO3/TiO2-25 heterojunctions, confirming the electron density decrease on TiO2 and electron density increase on WO3, respectively [36,49,50]. The electron density change indicates that the photo-generated carriers have been successfully transferred between WO3 and TiO2 in the composite, which further proved the formation of WO3/TiO2-25 heterojunction structures [34,36].

2.5. Optical Analysis

The UV–vis diffuse reflection spectra of TiO2, WO3, and WO3/TiO2 heterojunctions were recorded to understand the light absorption property and are shown in Figure 4a. The TiO2 nanoribbons show strong absorbance in the ultraviolet region, and the optical absorption edge was found to be around 400 nm owing to the large band gap of anatase TiO2 [21,22]. After the combination of WO3, the optical absorption edge exhibits a 130 nm red shift compared to pure TiO2 nanoribbons. Thus, the hybrid WO3/TiO2 heterostructures can utilize a larger fraction of the solar spectrum for photocatalytic reactions. In order to quantitatively determine their band gaps, (αhν)1/2 vs. photon energy (hν) are generated from the diffuse reflectance spectra (Figure 4b). Therein, α is the Kubelka–Munk function of the diffuse reflectance spectra (α = (1 − R)2/2R, and R is the reflectance). The apparent band gaps of pristine TiO2 and WO3 were calculated to be 3.12 eV and 2.26 eV, respectively, while the estimated band gap of WO3/TiO2 heterostructures was calculated to be 2.30 eV. The band gap reduction in TiO2 nanoribbons after WO3 sheet growth suggests the strong interaction between TiO2 and WO3 in WO3/TiO2 heterostructures, which can effectively hinder the recombination of e-h+ through the WO3/TiO2 heterojunction [27,29,35].

2.6. Photoluminescence Analysis

Photoluminescence measurements were further undertaken to unveil the charge recombination behavior of TiO2 and WO3/TiO2-25 heterostructures. It is widely known that fluorescence emission signals are derived mainly from the recombination of photo-generated electron–hole pairs, and a lower fluorescence intensity signifies a higher photocatalytic efficiency [12,31,43]. Figure 5a shows the fluorescence spectra of TiO2 and WO3/TiO2-25 heterostructures in the wavelength range of 475–725 nm. It can be seen clearly that the emission peaks of TiO2 and WO3/TiO2-25 heterostructures are in similar shapes, and the fluorescence emission intensity of TiO2 shows an obvious reduction after modification with WO3 nanosheets, indicating that growth of WO3 nanosheets on TiO2 nanoribbons can effectively suppress recombination of electron–hole pairs and increase the lifetime of the charge carriers, which provides solid basis for photocatalytic activity improvement. In order to further confirm the enhanced photo-generated electron–hole separation efficiency, photocurrent responses of all samples have been recorded and shown in Figure S4. The photocurrent of all the WO3/TiO2 composite samples is significantly higher than that of the pure TiO2 and WO3, further confirming the higher charge separation induced by WO3 nanosheet growth, which is consistent with photoluminescence spectra. It is worth noting that WO3/TiO2-25 exhibits the highest photocurrent density of 7.9 μA·cm−2, which is roughly 7.1, 6.6 times higher than that of pure TiO2 nanowires and pure WO3 nanosheets. The best photocurrent response manifests that the WO3/TiO2-25 heterostructure composite possesses the highest photoelectrons–holes transfer efficiency and thus should be a promising candidate as a high-performance photocatalyst.

2.7. Photocatalytic Performance

UV–vis diffuse reflection, photoluminescence, and photocurrent measurements demonstrate that construction of 2D/1D heterojunctions through vertical growth WO3 nanosheets on TiO2 nanoribbons not only broaden the light utilization region but also effectively suppresses electron–hole pair recombination; thus, theoretically, WO3/TiO2 heterostructures should possess higher photocatalytic activity than pure TiO2 and WO3 [11,12,31]. Figure 5b shows the photocatalytic degradation behavior of RhB over various photocatalysts under visible light irradiation. In the dark condition, all the samples exhibit similar RhB removal efficiency, owing to the similar BET surface area (Figure S3). All five kinds of WO3/TiO2 heterostructures prepared under different precursor concentrations show higher photocatalytic activity than pure WO3 nanosheets and TiO2 nanoribbons. Of the five heterojunctions, the WO3/TiO2-25 system shows the highest photocatalytic activity, which can remove 92.8% of RhB pollutants in 120 min. The photocatalytic degradation ration of RhB over WO3/TiO2-50 and WO3/TiO2-10 heterojunctions is 86.3% and 71.7%, respectively, indicating that too much higher and lower growth density of WO3 nanosheets causes photocatalytic activity decrease. The reason is that over-dense WO3 nanosheets result in lower exposure of junctions, while sparse growth density leads to few junctions. Figure 5c shows the first-order reaction kinetic curves for the RhB degradation reaction over various photocatalysts. It can be seen that the slope of WO3/TiO2 heterostructures is bigger than that of pure WO3 nanosheets and TiO2 nanoribbons, suggesting the enhanced photocatalytic rate constant after growth of WO3 nanosheets on TiO2 nanoribbons. The enhanced photocatalytic activity of the WO3/TiO2 heterostructures is attributed to the synergetic effects of the improved visible light utilization and enhanced electron–hole separation, which have been proven by UV–vis diffuse reflection and photoluminescence spectra. Over five consecutive cycles, there was no notable change for the apparent photocatalytic degradation ratio, indicating the excellent durability of WO3/TiO2 heterostructures (Figure 5d).
The effect of WO3/TiO2-25 catalyst loading on the photocatalytic degradation ratio of RhB is shown in Figure 6a. The photodegradation ratio increased with the increase in catalyst loading in the range of 5–20 mg and reached a maximum when 20 mg of the catalyst was used. Further increase in catalyst loading from 20 mg to 40 mg led to a decline in the photodegradation ratio from 92.8% to 82.6%. The observation can be explained as follows. Increasing catalyst loading from 5 mg to 20 mg would provide more active photodegradation sites, thus leading to an enhancement of the degradation ratio. When the catalyst was increased to 40 mg, the photodegradation reaction system became thick colloid, reducing the light penetration depth, which means only the catalysts in the solution surface layer can be photo activated as degradation sites, while the catalyst in the solution bottom layer make little contribution to RhB photodegradation, thus resulting in a decline in the degradation ratio [51]. Another reason might be due to catalyst aggregation resulting from high catalyst concentration, which would lead to a decrease in total surface area available for degradation, reduction in site density for surface holes and electrons, and an increase in the diffusion path length [51].
In order to reveal the photocatalytic mechanism, several trapping agents, including ascorbic acid (AC), ammonium oxalate (AO), isopropanol (IPA), and hydrogen peroxide (H2O2), were applied as scavengers for probing the active radicals in photocatalytic degradation. The scavengers AC, AO, IPA, and H2O2 functioned as trapping agents for superoxide anions (O2), holes (h+), hydroxyl radicals (OH), as well as electrons (e), respectively [52]. As shown in Figure 6b, various sacrificial agents have a great impact on the RhB degradation efficiency. After adding AO, AC, and IPA, the RhB photocatalytic degradation efficiency decreased obviously compared with the blank. The photodegradation ratios of RhB corresponding to AC, AO, and IPA were 40.2%, 20.7%, and 32.2%, respectively, indicating that h+ and OH constituted the major active species for RhB photodegradation [52]. The photodegradation ratio of RhB reached nearly 100% in the presence of H2O2. The addition of H2O2 resulted in the consumption of e in the conduction band (CB) and the enhancement of photoinduced e/h+ pair separation. Furthermore, H2O2 was reduced by e to generate OH, as indicated by H2O2 + e → OH + OH [53]. The generation of extra ·OH further promoted photocatalytic degradation. Therefore, active species detection reveals that photoinduced h+ and ·OH were the main active species in the RhB photodegradation process.
The conduction band (CB) and the valence band (VB) edge level potentials of TiO2 and WO3 were measured by applying the following formula [54]: EVB = X − Ec + 0.5Eg, ECB = EVB − Eg. Where EVB is the VB edge level potential and ECB is the CB edge level potential, and X is the absolute electronegativity of the semiconductor material (electronegativity of TiO2 and WO3 is 5.8 eV and 6.59 eV, respectively) [54]. Ec is the energy of free electrons (4.5 eV vs. NHE), and Eg is the band gap of TiO2 and WO3. The CB and VB edge level potentials of TiO2 were determined as −0.26 eV and 2.86 eV, respectively, while the CB and VB edge level potentials of WO3 were calculated to be 0.96 eV and 3.22 eV, respectively. The photocatalytic degradation mechanism of the WO3/TiO2 heterojunctions is schematically demonstrated in Figure 6c. When WO3/TiO2 heterojunction photocatalysts are exposed to visible light irradiation, the photo-generated electrons in the WO3 valence band will be excited to the conduction band, generating holes in the valence band. However, it cannot occur for TiO2 due to the band gap energy of TiO2 (3.12 eV) being higher than the visible photon energy. The valence band of WO3 is higher than that of TiO2 (3.22 eV vs. 2.86 eV), so the photo-generated holes can be easily transferred to the valence band of TiO2 through the 2D/1D junction interface [31,32,33,34,35,36]. The photo-generated electrons on the WO3 conduction band cannot react with O2 to produce •O2 radicals because the conduction edge potential of WO3 was higher than the redox potential of O2 to •O2 (0.96 eV vs. −0.33 eV). Thus, the radicals of •O2 can hardly participate in the photodegradation reaction, which is in good agreement with the scavenger experiment results. Furthermore, photo-generated holes (h+) can react with adsorbed H2O or OH to generate •OH radicals, which are the main active species for organic pollutant photo-degradation confirmed by the scavenger experiment. The above results showed TiO2 acted as a photo-generated hole acceptor in the heterojunction, effectively reducing the recombination of photo-generated carriers and enhancing photocatalytic performance, and h+ and •OH are responsible for the photodegradation of RhB.

3. Materials and Methods

3.1. Preparation of WO3/TiO2

TiO2 nanoribbons were synthesized according to our previous report [55,56]. Then, WO3 nanosheets were vertically grown on TiO2 nanoribbons using a hydrothermal process. Typically, 0.1 g of TiO2 nanoribbons were dispersed in 30 mL of distilled water dissolved by 0.25 g of Na2WO4•2H2O (25 mM), 2 mL of 3 mol/L HCl aqueous solution, and 0.3 g of citric acid. Then, the resulting suspension was transferred into a 50 mL Teflon-lined stainless steel autoclave and kept at 120 °C for 24 h. After the hydrothermal reaction was completed, the resulting products were collected by centrifugation, washed several times with deionized water, and dried at 60 °C overnight. For comparison, WO3/TiO2 composites with different ratios were prepared in a similar manner. The samples were denoted as WO3/TiO2-5, WO3/TiO2-10, WO3/TiO2-25, and WO3/TiO2-50, WO3/TiO2-75 in which the number represents the concentration of Na2WO4·2H2O.

3.2. Material Characterization

The morphology and structure of the photocatalysts were examined by a HITACHlS-4800 field emission scanning electron microscope (SEM, Hitachi, Tokyo, Japan). Transmission electron microscope (TEM, Thermo Fisher Scientific, Waltham, MA, USA) and high-resolution TEM (HRTEM, Thermo Fisher Scientific, Waltham, MA, USA) images were taken with an FEI Tecnai G20 electron microscope. For TEM samples, the photocatalyst was dispersed in ethanol using ultrasonic treatment and the TEM samples were prepared by depositing a drop of diluted suspension on a carbon-coated copper grid. UV-vis-near-infrared (NIR, Agilent, Polo Alto, CA, USA) reflection spectra were recorded on an Agilent Australia Carry-5000 spectrophotometer. The UV–Visible absorption spectra were recorded on a Shimadzu UV-2550 spectrophotometer. The X-ray photoelectron spectroscopy (XPS, Thermo Fisher Scientific, Waltham, MA, USA) analysis was carried out on an Escalab 250Xi spectrometer with monochromatic Al Kα radiation (hν = 1486.7 eV). Phase identification of the photocatalysts was measured by a Shimadzu powder difractometer with Cu-Kα radiation. The solid-state fluorescence spectra were characterized using a FLS980 fluorescence spectrophotometer.

3.3. Photodegradation and Photocurrent Test

Rhodamine B (RhB) solution with a concentration of 9 mg/L was used as the model waste water. An amount of 20 mg of samples were dispersed in 100 mL of RhB solution. The solution was kept in the dark for 20 min and then exposed to visible light irradiation. A 500 W halogen lamp was used as the visible light source, and the average light intensity was 60 mW•cm−2. The distance between the lamp and the solution was 10 cm. During the measurement process, a certain amount of the solution was sampled at certain time intervals and centrifuged (8000 rpm, 3 min) to remove the photocatalyst particles. The supernatant was analyzed to measure the concentration of RhB by using a Shimadzu UV-2550 UV–vis spectrometer (peak center: 554 nm). To analyze the photocatalytic degradation mechanism, ascorbic acid (AC), ammonium oxalate (AO), isopropanol (IPA), and hydrogen peroxide (H2O2) reagents were selected as scavengers for anions (O2), holes (h+), hydroxyl radicals (·OH), as well as electrons (e), respectively. Next, 5 mg of AC, 5 mg of AO, 0.2 mL of IPA, and 0.2 mL of H2O2 were added into the above photodegradation system while keeping other conditions unchanged.
Photocurrent studies were performed on a CHI 660D electrochemical workstation using a three-electrode configuration where ITO electrodes were deposited with the samples as a working electrode, Pt as a counter electrode, and a saturated calomel electrode as reference. The electrolyte was 0.35 M/0.25 M Na2S-Na2SO3 aqueous solution. For the fabrication of the working electrode, 0.25 g of the sample was grinded with 0.06 g polyethylene glycol (PEG, molecular weight: 20,000) and 0.5 mL ethanol to make a slurry. The slurry was spread onto a 1 cm × 4 cm ITO glass using the doctor blade technique and then allowed to air-dry. A 500 W halogen lamp was used as the visible light source and the average light intensity was 60 mW•cm−2.

4. Conclusions

In conclusion, a novel WO3/TiO2 2D/1D heterojunction photocatalyst was constructed via the introduction of TiO2 nanoribbons into a WO3 nanosheet growth system. Two-dimensional WO3 nanosheets were vertically arrayed on the surface of TiO2 nanoribbons, and the growth density could be easily controlled by adjusting the concentration of the precursors. The UV–vis diffuse reflection result reveals that vertical growth of WO3 nanosheets on TiO2 nanoribbons successfully decreases the band gap energy of TiO2 from 3.12 to 2.30 eV and broadens the photoresponse range from the UV region to the visible light region. Furthermore, the photoluminescence measurement demonstrates that construction of WO3/TiO2 heterojunctions significantly reduces electron–hole pair recombination. Consequently, WO3/TiO2 heterostructures with different WO3 nanosheet growth density show higher photocatalytic activity than pure WO3 nanosheets and TiO2 nanoribbons. WO3/TiO2-25 possesses the highest photocatalytic activity, which can remove 92.8% of RhB pollutants in 120 min and maintain high removal efficiency after five consecutive cycles. The present study demonstrates that the prepared WO3/TiO2 2D/1D heterostructures are promising materials for photocatalytic removal of organic pollutants to purify water.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal13030556/s1, Figure S1: SEM images of WO3/TiO2 heterostructures prepared under different precursor concentrations: (a) 5 mM, (b) 10 mM, (c) 50 mM, (d) 75 mM. Figure S2: SEM image of pure WO3 nanosheets. Figure S3: Nitrogen adsorptione-desorption isotherm: (a) TiO2 nanoribbons; (b) WO3 nanosheets; (c) WO3/TiO2 heterostructures and Figure S4. Photocurrent responses of all samples.

Author Contributions

Conceptualization, L.Z.; methodology, L.Z., L.W., K.X., and H.T.; validation, L.W. and H.T.; formal analysis, L.W., K.X., and H.T.; investigation, L.Z.; data curation, L.W., K.X., and L.W.; writing—original draft preparation, L.Z.; writing—review and editing, L.Z. and L.W.; visualization, L.W. and K.X.; supervision, L.Z.; project administration, L.Z.; funding acquisition, L.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Jiaxing City Public Welfare Research Project of China, grant number 2021AY10059 and Innovation Jiaxing Excellent Talents Support Program, grant number 84321005.

Data Availability Statement

The data presented in this study are available on request from the corresponding authors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. McDonald, R.I.; Green, P.; Balk, D.; Montgomery, M. Urban growth, climate change, and freshwater availability. Proc. Natl. Acad. Sci. USA 2011, 108, 6312–6317. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Oki, T.; Kanae, S. Global hydrological cycles and world water resources. Science 2006, 313, 1068–1072. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Vörösmarty, C.J.; McIntyre, P.B.; Gessner, M.O.; Dudgeon, D.; Prusevich, A.; Green, P.; Glidden, S.; Bunn, S.E.; Sullivan, C.A.; Liermann, C.R.; et al. Global threats to human water security and river biodiversity. Nature 2010, 467, 555–561. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Gadgil, A. Drinking Water in Developing Countries. Annu. Rev. Energy Environ. 1998, 23, 253–286. [Google Scholar] [CrossRef] [Green Version]
  5. Water Scarcity Threats. Available online: https://www.worldwildlife.org/threats/water-scarcity (accessed on 12 January 2023).
  6. Shannon, M.A.; Bohn, P.W.; Elimelech, M.; Georgiadis, J.G.; Mariñas, B.J.; Mayes, A.M. Science and technology for water purification in the coming decades. Nature 2008, 452, 301–310. [Google Scholar] [CrossRef]
  7. Macedonio, F.; Drioli, E.; Gusev, A.A.; Bardow, A.; Semiatf, R.; Kuriharag, M. Efficient technologies for worldwide clean water supply. Chem. Eng. Process. 2012, 51, 2–17. [Google Scholar] [CrossRef]
  8. Tee, G.T.; Gok, X.Y.; Yong, W.F. Adsorption of pollutants in wastewater via biosorbents, nanoparticles and magnetic biosorbents: A review. Environ. Res. 2022, 212, 113248. [Google Scholar] [CrossRef]
  9. Ouachtak, H.; Akhouairi, S.; Haounati, R.; Addi, A.A.; Jada, A.; Taha, M.L.; Douch, J. 3,4-Dihydroxybenzoic acid removal from water by goethite modified natural sand column fixed-bed: Experimental study and mathematical modeling. Desalin. Water Treat. 2020, 194, 439–449. [Google Scholar] [CrossRef]
  10. Pendergast, M.T.M.; Eric, M.V.; Hoek, E.M.V. A review of water treatment membrane nanotechnologies. Energy Environ. Sci. 2011, 4, 1946–1971. [Google Scholar] [CrossRef] [Green Version]
  11. Andrew, M.; Davies, R.H.; Worsley, D. Water purification by semiconductor photocatalysis. Chem. Soc. Rev. 1993, 22, 417–425. [Google Scholar]
  12. Parvulescu, V.I.; Epron, F.; Garcia, H.; Granger, P. Recent progress and prospects in catalytic water treatment. Chem. Rev. 2021, 122, 2981–3121. [Google Scholar] [CrossRef] [PubMed]
  13. Schneider, J.; Matsuoka, M.; Takeuchi, M.; Zhang, J.; Horiuchi, Y.; Anpo, M.; Bahnemann, D.W. Understanding TiO2 Photocatalysis: Mechanisms and Materials. Chem. Rev. 2014, 114, 9919–9986. [Google Scholar] [CrossRef] [PubMed]
  14. Kumar, A.; Choudhary, P.; Kumar, A.; Camargo, P.H.; Krishnan, V. Recent advances in plasmonic photocatalysis based on TiO2 and noble metal nanoparticles for energy conversion, environmental remediation, and organic synthesis. Small 2022, 18, 2101638. [Google Scholar] [CrossRef] [PubMed]
  15. Weng, B.; Lu, K.Q.; Tang, Z.C.; Chen, H.M.; Xu, Y.J. Stabilizing ultrasmall Au clusters for enhanced photoredox catalysis. Nat. Commun. 2018, 9, 1543. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Pan, X.; Chen, W.J.; Cai, H.; Li, H.; Sun, X.J.; Weng, B.; Yi, Z. A redox-active support for the synthesis of Au@SnO2 core-shell nanostructure and SnO2 quantum dots with efficient photoactivities. RSC Adv. 2020, 10, 33955–33961. [Google Scholar] [CrossRef] [PubMed]
  17. Haounati, R.; Alakhras, F.; Ouachtak, H.; Saleh, T.A.; Al-Mazaideh, G.; Alhajri, E.; Jada, A.; Hafid, N.; Addi, A.A. Synthesized of Zeolite@Ag2O Nanocomposite as Superb Stability Photocatalysis toward Hazardous Rhodamine B Dye from Water. Arab. J. Sci. Eng. 2023, 48, 169–179. [Google Scholar] [CrossRef]
  18. Nawaz, A.; Goudarzi, S.; Asghari, M.A.; Pichiah, S.; Selopal, G.S.; Rosei, F.; Wang, Z.M.; Zarrin, H. Review of Hybrid 1D/2D Photocatalysts for Light-Harvesting Applications. ACS Appl. Nano Mater. 2021, 4, 11323–11352. [Google Scholar] [CrossRef]
  19. Hou, H.; Zhang, X. Rational design of 1D/2D heterostructured photocatalyst for energy and environmental applications. Chem. Eng. J. 2020, 395, 125030. [Google Scholar] [CrossRef]
  20. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef]
  21. Ren, Y.; Dong, Y.; Feng, Y.; Xu, J. Compositing two-dimensional materials with TiO2 for photocatalysis. Catalysts 2018, 8, 590. [Google Scholar] [CrossRef] [Green Version]
  22. Guo, Q.; Zhou, C.; Ma, Z.; Yang, X. Fundamentals of TiO2 photocatalysis: Concepts, mechanisms, and challenges. Adv. Mater. 2019, 31, 1901997. [Google Scholar] [CrossRef] [PubMed]
  23. Feng, T.; Feng, G.S.; Yan, L. One-dimensional nanostructured TiO2 for photocatalytic degradation of organic pollutants in wastewater. Int. J. Photoenergy 2014, 2014, 563879. [Google Scholar] [CrossRef] [Green Version]
  24. Ge, M.; Cao, C.; Huang, J. A review of one-dimensional TiO2 nanostructured materials for environmental and energy applications. J. Mater. Chem. A 2016, 4, 6772–6801. [Google Scholar] [CrossRef]
  25. Szilágyi, I.M.; Fórizs, B.; Rosseler, O.; Szegedi, Á.; Németh, P.; Király, P.; Tárkányi, G.; Vajna, B.; Katalin, V.-J.; László, K. WO3 photocatalysts: Influence of structure and composition. J. Catal. 2012, 294, 119–127. [Google Scholar] [CrossRef] [Green Version]
  26. Dutta, V.; Sharma, S.; Raizada, P.; Thakurb, V.K.; Parvaz, K.A.A.; Sainie, V.; Asiri, A.M.; Singha, P. An overview on WO3 based photocatalyst for environmental remediation. J. Environ. Chem. Eng. 2021, 9, 105018. [Google Scholar] [CrossRef]
  27. Wang, F.; Valentin, C.D.; Pacchioni, G. Rational band gap engineering of WO3 photocatalyst for visible light water splitting. Mater. Sci. ChemCatChem. 2012, 4, 476–478. [Google Scholar] [CrossRef]
  28. Liu, D.; Zhang, S.; Wang, J.; Peng, T.; Li, R. Direct Z-scheme 2D/2D photocatalyst based on ultrathin g-C3N4 and WO3 nanosheets for efficient visible-light-driven H2 generation. ACS Appl. Mater. Interfaces 2019, 11, 27913–27923. [Google Scholar] [CrossRef]
  29. Lei, B.; Cui, W.; Chen, P.; Chen, L.; Li, J.; Dong, F. C-doping induced oxygen-vacancy in WO3 nanosheets for CO2 activation and photoreduction. ACS Catal. 2022, 12, 9670–9678. [Google Scholar] [CrossRef]
  30. Liang, Y.; Yang, Y.; Zou, C.; Xu, K.; Luo, X.; Luo, T.; Li, J.; Yang, Q.; Shi, P.; Yuan, C. 2D ultra-thin WO3 nanosheets with dominant {002} crystal facets for high-performance xylene sensing and methyl orange photocatalytic degradation. J. Alloy. Compd. 2019, 783, 848–854. [Google Scholar] [CrossRef]
  31. Lai, C.W. WO3-TiO2 Nanocomposite and its applications: A review. Nano Hybrids Compos. 2018, 20, 1–26. [Google Scholar] [CrossRef]
  32. Zhang, L.; Qin, M.; Yu, W.; Zhang, Q.; Xie, H.; Sun, Z.; Shao, Q.; Guo, X.; Hao, L.; Zheng, Y.; et al. Heterostructured TiO2/WO3 nanocomposites for photocatalytic degradation of toluene under visible light. J. Electrochem. Soc. 2017, 164, H1086–H1090. [Google Scholar] [CrossRef]
  33. Liu, Y.; Xie, C.; Li, J.; Zou, T.; Zeng, D. New insights into the relationship between photocatalytic activity and photocurrent of TiO2/WO3 nanocomposite. Appl. Catal. A 2012, 433–434, 81–87. [Google Scholar] [CrossRef]
  34. Prabhu, S.; Cindrella, L.; Kwon, O.J.; Mohanrajubet, K. Photoelectrochemical and photocatalytic activity of TiO2-WO3 heterostructures boosted by mutual interaction. Mater. Sci. Semicond. Process. 2018, 88, 10–19. [Google Scholar] [CrossRef]
  35. Yang, J.; Zhang, X.; Liu, H.; Wang, C.; Liu, S.; Sun, P.; Wang, L.; Liu, Y. Heterostructured TiO2/WO3 porous microspheres: Preparation, characterization and photocatalytic properties. Catal. Today 2013, 201, 195–202. [Google Scholar] [CrossRef]
  36. Wang, Q.; Zhang, W.; Hu, X.; Xua, L.; Chen, G.; Li, X. Hollow spherical WO3/TiO2 heterojunction for enhancing photocatalytic performance in visible-light. J. Water Process. Eng. 2021, 40, 101943. [Google Scholar] [CrossRef]
  37. Zhang, Y.; Xu, M.; Li, H.; Ge, H.; Bian, Z. The enhanced photoreduction of Cr (VI) to Cr (III) using carbon dots coupled TiO2 mesocrystals. Appl. Catal. B 2018, 226, 213–219. [Google Scholar] [CrossRef]
  38. Song, Y.Y.; Gao, Z.D.; Wang, J.H.; Xia, X.H.; Lynch, R. Multistage coloring electrochromic device based on TiO2 nanotube arrays modified with WO3 nanoparticles. Adv. Funct. Mater. 2011, 21, 1941–1946. [Google Scholar] [CrossRef]
  39. Paramasivam, I.; Nah, Y.C.; Das, C.; Shrestha, N.K.; Schmuki, P. WO3/TiO2 nanotubes with strongly enhanced photocatalytic activity. Chem. Eur. J. 2010, 16, 8993–8997. [Google Scholar] [CrossRef]
  40. Xu, T.; Wang, Y.; Zhou, X.; Zheng, X.; Xu, Q.; Chen, Z.; Ren, Y.; Yan, B. Fabrication and assembly of two-dimensional TiO2/WO3·H2O heterostructures with type II band alignment for enhanced photocatalytic performance. Appl. Surf. Sci. 2017, 403, 564–571. [Google Scholar] [CrossRef]
  41. Lee, J.Y.; Jo, W.K. Heterojunction-based two-dimensional N-doped TiO2/WO3 composite architectures for photocatalytic treatment of hazardous organic vapor. J. Hazard. Mater. 2016, 314, 22–31. [Google Scholar] [CrossRef]
  42. Liu, J.; Yang, L.; Li, C.; Chen, Y.; Zhang, Z. Optimal monolayer WO3 nanosheets/TiO2 heterostructure and its photocatalytic performance under solar light. Chem. Phys. Lett. 2022, 804, 139861. [Google Scholar] [CrossRef]
  43. Mei, Y.; Su, Y.; Li, Z.; Bai, S.; Yuan, M.; Li, L.; Yan, Z.; Wu, J.; Zhu, L.W. BiOBr nanoplates@TiO2 nanowires/carbon fiber cloth as a functional water transport network for continuous flow water purification. Dalton Trans. 2017, 46, 347–354. [Google Scholar] [CrossRef] [PubMed]
  44. Chen, J.Z.; Ko, W.Y.; Yen, Y.C.; Chen, P.H.; Lin, K.J. Hydrothermally processed TiO2 nanowire electrodes with antireflective and electrochromic properties. ACS Nano 2012, 6, 6633–6639. [Google Scholar] [CrossRef] [PubMed]
  45. Li, F.; Li, C.; Zhu, L.; Guo, W.; Shen, L.; Wen, S.; Ruan, S. Enhanced toluene sensing performance of gold-functionalized WO3· H2O nanosheets. Sens. Actuators B Chem. 2016, 223, 761–767. [Google Scholar] [CrossRef]
  46. Liu, W.; Du, T.; Ru, Q.; Zuo, S.; Cai, Y.; Yao, C. Preparation of graphene/WO3/TiO2 composite and its photocathodic protection performance for 304 stainless steel. Mater. Res. Bull. 2018, 102, 399–405. [Google Scholar] [CrossRef]
  47. Idriss, H.; Barteau, M.A. Characterization of TiO2 surfaces active for novel organic syntheses. Catal. Lett. 1994, 26, 123–139. [Google Scholar] [CrossRef]
  48. Chen, S.; Chen, L.; Gao, S.; Ca, G. The preparation of coupled WO3/TiO2 photocatalyst by ball milling. Powder Technol. 2005, 160, 198–202. [Google Scholar]
  49. Li, S.; Wang, P.; Wang, R.; Liu, Y.; Jing, R.; Zi, Z.; Meng, L.; Liu, Y.; Zhang, Q. One-step co-precipitation method to construct black phosphorus nanosheets/ZnO nanohybrid for enhanced visible light photocatalytic activity. Appl. Surf. Sci. 2019, 497, 143682. [Google Scholar] [CrossRef]
  50. Zhang, Z.; Liu, K.; Feng, Z.; Bao, Y.; Dong, B. Hierarchical sheet-on-sheet ZnIn2S4/g-C3N4 heterostructure with highly efficient photocatalytic H2 production based on photoinduced interfacial charge transfer. Sci. Rep. 2016, 6, 19221. [Google Scholar] [CrossRef] [Green Version]
  51. Paul, D.R.; Sharma, R.; Nehra, S.P.; Sharma, A. Effect of calcination temperature, pH and catalyst loading on photodegradation efficiency of urea derived graphitic carbon nitride towards methylene blue dye solution. RSC Adv. 2019, 9, 15381–15391. [Google Scholar] [CrossRef] [Green Version]
  52. Gao, J.; Hu, J.; Wang, Y.; Zheng, L.; He, G.; Deng, J.; Liu, M.; Li, Y.; Liu, Y.; Zhou, H. Fabrication of Z-Scheme TiO2/SnS2/MoS2 ternary heterojunction arrays for enhanced photocatalytic and photoelectrochemical performance under visible light. J. Solid State Chem. 2022, 307, 122737. [Google Scholar] [CrossRef]
  53. Zhu, B.; Xia, P.; Li, Y.; Ho, W.; Yu, J. Fabrication and photocatalytic activity enhanced mechanism of direct Z-scheme g-C3N4/Ag2WO4 photocatalyst. Appl. Surf. Sci. 2017, 391, 175–183. [Google Scholar] [CrossRef]
  54. Vembuli, T.; Thiripuranthagan, S.; Sureshkumar, T.; Erusappan, E.; Kumaravel, S.; Kasinathan, M.; Sivakumar, A. Degradation of Harmful Organics Using Visible Light Driven N-TiO2/rGO Nanocomposite. J. Nanosci. Nanotechnol. 2021, 21, 3081–3091. [Google Scholar] [CrossRef] [PubMed]
  55. Zhu, L.; Gu, L.; Zhou, Y.; Cao, S.; Cao, X. Direct production of a free-standing titanate and titania nanofiber membrane with selective permeability and cleaning performance. J. Mater. Chem. 2011, 21, 12503–12510. [Google Scholar] [CrossRef]
  56. Zhou, Y.; Zhu, L.; Gu, L.; Cao, S.; Wang, L.; Cao, X. Guided growth and alignment of millimetre-long titanate nanofibers in solution. J. Mater. Chem. 2012, 22, 16890–16896. [Google Scholar] [CrossRef]
Figure 1. (a) Photograph of TiO2 before and after hydrothermal growth of WO3. (b) XRD patterns of TiO2, WO3, and WO3/TiO2−25. (c) FTIR spectra of TiO2, WO3, and WO3/TiO2 composites with different ratios.
Figure 1. (a) Photograph of TiO2 before and after hydrothermal growth of WO3. (b) XRD patterns of TiO2, WO3, and WO3/TiO2−25. (c) FTIR spectra of TiO2, WO3, and WO3/TiO2 composites with different ratios.
Catalysts 13 00556 g001
Figure 2. (a) SEM image of TiO2 nanoribbons. (bd) TEM images of TiO2 nanoribbons. (e) SEM image of WO3/TiO2-25 heterostructure. (fh) TEM images of WO3/TiO2-25 heterostructure. (il) HAADF-STEM micrograph of WO3/TiO2-25 heterostructure and the corresponding EDS mapping of Ti, W, and O.
Figure 2. (a) SEM image of TiO2 nanoribbons. (bd) TEM images of TiO2 nanoribbons. (e) SEM image of WO3/TiO2-25 heterostructure. (fh) TEM images of WO3/TiO2-25 heterostructure. (il) HAADF-STEM micrograph of WO3/TiO2-25 heterostructure and the corresponding EDS mapping of Ti, W, and O.
Catalysts 13 00556 g002
Figure 3. XPS spectra: (a) Ti 2p spectra of TiO2 and WO3/TiO2-25 heterojunctions. The orange and blue curves are fitting curves corresponding to the Ti 2p1/2 and Ti 2p3/2, respectively. The black line is the raw data. (b) W 4f spectra of WO3 and WO3/TiO2-25 heterojunctions. The orange, cyan, blue and green curves are fitting curves corresponding to the W 4f7/2 and W 4f5/2, respectively. The black represent the raw data and the red line is the fitting line. (c) O 1s spectra of TiO2 and WO3/TiO2-25 heterojunctions. The magenta, blue and olive curves are fitting curves corresponding to the Ti-O bond. The black represent the raw data and the red line is the fitting line. (d) O 1s spectra of WO3 and WO3/TiO2-25 heterojunctions. The magenta, blue and olive curves are fitting curves corresponding to the W-O bond. The black represent the raw data and the red line is the fitting line.
Figure 3. XPS spectra: (a) Ti 2p spectra of TiO2 and WO3/TiO2-25 heterojunctions. The orange and blue curves are fitting curves corresponding to the Ti 2p1/2 and Ti 2p3/2, respectively. The black line is the raw data. (b) W 4f spectra of WO3 and WO3/TiO2-25 heterojunctions. The orange, cyan, blue and green curves are fitting curves corresponding to the W 4f7/2 and W 4f5/2, respectively. The black represent the raw data and the red line is the fitting line. (c) O 1s spectra of TiO2 and WO3/TiO2-25 heterojunctions. The magenta, blue and olive curves are fitting curves corresponding to the Ti-O bond. The black represent the raw data and the red line is the fitting line. (d) O 1s spectra of WO3 and WO3/TiO2-25 heterojunctions. The magenta, blue and olive curves are fitting curves corresponding to the W-O bond. The black represent the raw data and the red line is the fitting line.
Catalysts 13 00556 g003
Figure 4. (a) Diffuse reflectance spectra for the TiO2 nanoribbons, WO3 nanosheets, and WO3/TiO2-25 heterojunction. (b) Plot of (αhν)1/2 as a function of photon energy (hν).
Figure 4. (a) Diffuse reflectance spectra for the TiO2 nanoribbons, WO3 nanosheets, and WO3/TiO2-25 heterojunction. (b) Plot of (αhν)1/2 as a function of photon energy (hν).
Catalysts 13 00556 g004
Figure 5. (a) Fluorescence spectrum of TiO2 nanoribbons and WO3/TiO2−25 heterojunction. (b) Photocatalytic degradation of RhB over various photocatalysts under sunlight irradiation, where C0 is the initial concentration of the pollutant, and C is the concentration of the pollutant at irradiation time t. (c) First-order reaction kinetic curves for the RhB degradation reaction over various photocatalysts. (d) Curve of the degradation ratio of RhB versus reuse times of WO3/TiO2-25.
Figure 5. (a) Fluorescence spectrum of TiO2 nanoribbons and WO3/TiO2−25 heterojunction. (b) Photocatalytic degradation of RhB over various photocatalysts under sunlight irradiation, where C0 is the initial concentration of the pollutant, and C is the concentration of the pollutant at irradiation time t. (c) First-order reaction kinetic curves for the RhB degradation reaction over various photocatalysts. (d) Curve of the degradation ratio of RhB versus reuse times of WO3/TiO2-25.
Catalysts 13 00556 g005
Figure 6. (a) The effect of WO3/TiO2−25 catalyst loading on RhB photodegradation ratio. (b) Trapping test for active species during RhB photodegradation with the WO3/TiO2−25 photocatalyst. (c) Proposed photocatalytic mechanism of WO3/TiO2 photocatalyst.
Figure 6. (a) The effect of WO3/TiO2−25 catalyst loading on RhB photodegradation ratio. (b) Trapping test for active species during RhB photodegradation with the WO3/TiO2−25 photocatalyst. (c) Proposed photocatalytic mechanism of WO3/TiO2 photocatalyst.
Catalysts 13 00556 g006
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, L.; Xu, K.; Tang, H.; Zhu, L. Vertical Growth of WO3 Nanosheets on TiO2 Nanoribbons as 2D/1D Heterojunction Photocatalysts with Improved Photocatalytic Performance under Visible Light. Catalysts 2023, 13, 556. https://doi.org/10.3390/catal13030556

AMA Style

Wang L, Xu K, Tang H, Zhu L. Vertical Growth of WO3 Nanosheets on TiO2 Nanoribbons as 2D/1D Heterojunction Photocatalysts with Improved Photocatalytic Performance under Visible Light. Catalysts. 2023; 13(3):556. https://doi.org/10.3390/catal13030556

Chicago/Turabian Style

Wang, Ling, Keyi Xu, Hongwang Tang, and Lianwen Zhu. 2023. "Vertical Growth of WO3 Nanosheets on TiO2 Nanoribbons as 2D/1D Heterojunction Photocatalysts with Improved Photocatalytic Performance under Visible Light" Catalysts 13, no. 3: 556. https://doi.org/10.3390/catal13030556

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop