Next Article in Journal
Intensity-Dependent Optical Response of 2D LTMDs Suspensions: From Thermal to Electronic Nonlinearities
Previous Article in Journal
The Effect of Nanobubble Water Containing Cordyceps Extract and Withaferin A on Free Fatty Acid-Induced Lipid Accumulation in HepG2 Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synergistic Effect of Y Doping and Reduction of TiO2 on the Improvement of Photocatalytic Performance

1
College of Materials Science and Technology, Nanjing University of Aeronautics and Astronautics, Nanjing 210016, China
2
State Key Laboratory of Mechanics and Control for Aerospace Structures, Nanjing University of Aeronautics and Astronautics, Nanjing 210016, China
3
Department of Applied Physics, The Hong Kong Polytechnic University, Hong Kong 999077, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(15), 2266; https://doi.org/10.3390/nano13152266
Submission received: 6 July 2023 / Revised: 31 July 2023 / Accepted: 1 August 2023 / Published: 7 August 2023
(This article belongs to the Section Energy and Catalysis)

Abstract

:
Pure TiO2 and 3% Y-doped TiO2 (3% Y-TiO2) were prepared by a one-step hydrothermal method. Reduced TiO2 (TiO2-H2) and 3% Y-TiO2 (3% Y-TiO2-H2) were obtained through the thermal conversion treatment of Ar-H2 atmosphere at 500 °C for 3 h. By systematically comparing the crystalline phase, structure, morphological features, and photocatalytic properties of 3% Y-TiO2-H2 with pure TiO2, 3% Y-TiO2, and TiO2-H2, the synergistic effect of Y doping and reduction of TiO2 was obtained. All samples show the single anatase phase, and no diffraction peak shift is observed. Compared with single-doped TiO2 and single-reduced TiO2, 3% Y-TiO2-H2 exhibits the best photocatalytic performance for the degradation of RhB, which can be totally degraded in 20 min. The improvement of photocatalytic performance was attributed to the synergistic effect of Y doping and reduction treatment. Y doping broadened the range of light absorption and reduced the charge recombination rates, and the reduction treatment caused TiO2 to be enveloped by disordered shells. The remarkable feature of reduced TiO2 by H2 is its disordered shell filled with a limited amount of oxygen vacancies (OVs) or Ti3+, which significantly reduces the Eg of TiO2 and remarkably increases the absorption of visible light. The synergistic effect of Y doping, Ti3+ species, and OVs play an important role in the improvement of photocatalytic performances. The discovery of this work provides a new perspective for the improvement of other photocatalysts by combining doping and reduction to modify traditional photocatalytic materials and further improve their performance.

1. Introduction

Among the various semiconductors, titanium dioxide (TiO2) is considered a useful photocatalyst for the treatment of water pollution and water splitting, owing to its minimal toxicity, strong oxidation capacity, and ready availability [1,2,3,4,5,6]. In recent decades, many research articles have reported the performance, synthesis methods, as well as the reaction mechanisms of TiO2 in photocatalytic systems [7,8,9]. The crystalline phase of TiO2 remarkably affects its photocatalytic activity. Rutile, anatase, and brookite are the three main crystalline phases of TiO2 [10]. The photocatalytic performance of pure anatase TiO2 is better than that of rutile, whereas the brookite phase is unstable and does not show appreciable photoactivity [11]. However, the wide bandgap (Eg; anatase, ~3.20 eV) limits light absorption to the UV region of the solar spectrum (~4% of the total solar irradiance), thus affecting photocatalytic activity [12]. Moreover, the high recombination of the photogenerated electron–hole (e–h+) pair has been proven to reduce photocatalytic efficiency [13]. Many approaches, including metal (Al3+, Zn2+, Fe3+, Ce3+, Mn2+, Ag+) [14,15,16] and nonmetal doping [17], the deposition of noble metal (Au, Pt) [18,19], and construction of heterojunctions with other semiconductors [20,21], have been explored to overcome these limitations.
Among these methods, rare earth (RE) metal doping is a popular technique to reduce the recombination rate of photogenerated carriers and shift the absorption wavelength to the visible region and increase the photoactivity of TiO2 [22,23]. Furthermore, RE metal doping shows other benefits, such as the ability to concentrate pollutants at the TiO2 photocatalyst surface, inhibit the phase transformation from anatase to rutile, and decrease the crystallite size [24,25]. Many studies reported that the concentration of RE metal dopant affects the photoactivity, with the optimum achieved at concentrations below 5 wt.% [25]. Among all the RE ions, yttrium (Y) ion is considered a typical dopant used to modify the electronic structure and optical properties of TiO2 [26]. Specifically, it has been found that the Y-doped TiO2 can reduce the recombination rate of photogenerated electrons/holes pairs, which improves the photocatalytic efficiency of TiO2 [27]. The effect of Y-doped TiO2 on photocatalytic activity has been recently reported in several studies [28,29].
In addition to RE metal doping, reduced TiO2, which is obtained by high-temperature treatments in various reducing atmospheres (e.g., vacuum, Ar, Ar-H2, and pure H2) [30,31,32] or thermite reduction [33], reduction with sodium borohydride [34], and Ti3+ self-doping [35,36], has shown tremendous potential as a photocatalyst in wastewater treatment and water splitting recently. The enhanced photocatalytic performance of reduced TiO2 is related to a significant decrease in Eg and increased light absorption due to the introduction of oxygen vacancies (OV) or the formation of Ti3+ centers in the TiO2 lattice [34]. Current studies on reduced TiO2 focused on the relationship between the Ti3+ or OV concentration and photocatalytic activity of TiO2 and comparison of the reduction degree caused by different reduction methods. However, the synergistic effect of RE doping and reduction treatment for TiO2 photocatalysts has not been reported yet.
In this work, pure anatase TiO2, Y-doped TiO2, and their respective reduced products were prepared to verify the synergistic effect of Y doping and reduction treatment for TiO2 photocatalysts. Pure anatase TiO2 and Y-doped TiO2 were obtained by a one-step hydrothermal method, reduced TiO2 and reduced Y-doped TiO2 were formed by heat treatment in an Ar-H2 atmosphere at 500 °C for 3 h. The crystalline phase, morphological features, and photocatalytic properties of four types of TiO2 were compared systematically. By comparing TiO2 and Y-doped TiO2, the effect of RE doping on the improvement of photocatalytic performance could be obtained. Similarly, by comparing TiO2 with reduced TiO2, the effect of reduction treatment by H2 on the improvement of photocatalytic performance can be obtained. By comparing TiO2, which are simultaneously doped and reduced, with pure TiO2, Y-doped TiO2, and reduced TiO2, it can be verified the existence of a coupling effect of RE doping and reduction treatment on the improvement of photocatalytic performance. Ultimately, a positive effect of the combination of Y doping and reduction of TiO2 on the improvement of photodegradation activity of organic pollutants under simulated sunlight irradiation was found, and their possible photocatalytic mechanism was proposed. The aim of this study was to develop an understanding of the synergistic effect of Y doping and reduction of TiO2 on the improvement of photocatalytic performance.

2. Experimental

2.1. Materials

Titanium isopropoxide (TIP), Y(NO3)3·6H2O, absolute ethanol, acetic acid, and nitric acid (HNO3) were received from Sinopharm Chemical Reagent Co., Ltd., Shanghai, China. All chemicals were not further purified during use.

2.2. Preparation of Pure TiO2 and Y-Doped TiO2 Nanoparticles

A total of 4.2 mL of TIP, 16.8 mL of absolute ethanol, and 1.68 mL of acetic acid were mixed under continuous stirring to form solution A. Solution B was composed of a certain amount of Y(NO3)3·6H2O dissolved in 21 mL of deionized water with a pH value adjusted to 2.5 by dilute HNO3 (1 mol·L−1). The molar ratio of Y/Ti in solution B was set to 0, 1%, 2%, 3%, 4%, and 5%, respectively. Then, solution B was added into solution A to obtain Y-doped TiO2 sol. After stirring for 30 min, the Y-doped TiO2 sol was transferred into a 70 mL Teflon vessel. The Teflon vessel was placed into stainless-steel autoclaves. Then, the autoclaves were put in an oven and heated to 180 °C for 12 h. Then, the autoclaves were cooled to room temperature after hydrothermal treatment. The final products were filtered, washed, and dried at 80 °C for 12 h.

2.3. Preparation of Reduced TiO2 and Y-Doped TiO2 Nanoparticles

TiO2 and Y-doped TiO2 nanoparticles were placed in a tube furnace, respectively, and then evacuated to a base pressure of about 0.5 Pa. The tube was filled with Ar-H2 (95 vol %–5 vol %) atmosphere to normal pressure. The samples were heated at 500 °C with a heating rate of 5 °C /min for 3 h in Ar-H2 flow to obtain the final reduced pure TiO2 and Y-doped TiO2 products by H2.

2.4. Photocatalytic Test

The photocatalytic activity test was conducted using RhB as a simulated pollutant and a 300 W Xe lamp as a simulated sunlight source (PLS-SXE300/300UV, Trusttech Co., Ltd., Beijing, China) to irradiate the pollutant. In a typical photocatalytic process, 10 mg of the catalyst was dispersed into 100 mL of the RhB solution (10 mg L−1) in a 200 mL double-layer reactor cooled by running water to maintain a temperature of 25 °C, and the mixture was magnetically stirred with 300 rpm in the dark for 1 h to achieve an adsorption–desorption balance. Then, the xenon lamp was turned on, and 5 mL of the suspension was collected at every 10 min interval and centrifuged (9000 rpm, 5 min) to remove the catalyst powders. The total irradiation time of all samples was 1 h. The degradation rate of RhB was calculated by converting the absorbance into concentration using the dye standard curve by a UV–vis spectrophotometer.

2.5. Characterization

The phase structure of the products was determined using X-ray diffraction (XRD) (Bruker D8Advance diffractometer) with Cu Kα radiation (λ = 1.5418 Å). The structural information of products was obtained using high-resolution transmission electron microscopy (HRTEM) by a JEM-2100F at 200 kV. X-ray photoelectron spectroscopy (XPS) measurements were carried out using the AXIS-Ultra DLD system. The UV–vis diffuse reflectance absorption spectra (DRS) of samples were used to determine the bandgap by the PE Lambda 950 spectrometer with BaSO4 as reference. The Brunauer–Emmett–Teller (BET) method with a JW-BK200B was used to evaluate the specific surface area of the products.

3. Results and Discussion

3.1. Structure and Morphology Characterization

Pure TiO2 and various Y-doped TiO2 (1–5%) were prepared in advance to verify their phase composition and photocatalytic performance for the degradation of RhB. XRD was used to characterize and compare the crystalline phase of samples. All samples show the single anatase phase (Figure S1a). The catalytic activities were evaluated by the degradation of RhB under simulated sunlight irradiation and the maximum photocatalytic performance for the degradation of RhB when the Y dopant concentration was 3% (Figure S1b). Therefore, 3% Y-doped TiO2 was selected as a typical research object for rare earth doping. Figure 1 shows the crystalline phase of pure TiO2, 3% Y-doped TiO2 (3% Y-TiO2), reduced TiO2 by H2 (TiO2-H2), and reduced 3% Y-doped TiO2 (3% Y-TiO2-H2). All four samples have significant diffraction peaks representing the characteristic of a single anatase phase (PDF No. 211272). Anatase is the only crystalline phase present in the structure of TiO2-H2 and 3% Y-TiO2-H2, indicating that no phase transformation occurs in the annealing process of H2 reduction at 500 °C. After annealing, the colors of TiO2-H2 and 3% Y-TiO2-H2 change from white to gray black. The refraction of Y2O3 is not observed in the XRD patterns of 3% Y-TiO2 or 3% Y-TiO2-H2, indicating that the content of Y2O3 is below the detection limit. No diffraction peak shift is observed for all Y-modified samples, demonstrating that Y3+ species exists at the crystal boundary or surface rather than in the inner crystalline structure of TiO2. The increased diffraction peak intensities of TiO2-H2 and 3% Y-TiO2-H2 after annealing indicate the increased crystallinity of samples in comparison with those of pure TiO2 and 3% Y-TiO2.
The morphology and structure of all four samples are investigated by HRTEM. The low-magnification TEM images and statistical particle size distribution of TiO2, 3% Y-TiO2, TiO2-H2, and 3% Y-TiO2-H2 are shown in Figure 2a–d and their insert, respectively. All samples are primarily composed of dispersed circular, rectangular, and some irregular-shaped nanoparticles with average sizes ranging from 8 nm to 11 nm. The grain sizes of TiO2-H2 and 3% Y-TiO2-H2 increase slightly but not significantly after annealing. High-magnification TEM images show clear lattice fringes, which is indicative of the high crystallinity of TiO2 and 3% Y-TiO2 (Figure 2e,f). However, high-magnification TEM images of TiO2-H2 and 3% Y-TiO2-H2 illustrate a core–shell structure with a ∼1.5 nm-thick disordered surface shell (Figure 2g,h), which is not observed on the surfaces of TiO2 and 3% Y-TiO2 prepared by the single hydrothermal method. Four samples have lattice spacings of 0.34 nm in the same way as in Figure 2e, which is consistent with its (101) planes (Figure 2e–h). The elemental mapping images of 3% Y-TiO2-H2 with individual elements of Ti, O, and Y are shown in Figure 2i. Ti, O, and Y are uniformly distributed throughout the particle space, proving the presence of Y elements in Y-doped TiO2 samples.
The surface areas of all products were estimated by the BET analysis; their N2 adsorption–desorption isotherms are shown in Figure 3. The specific surface areas of pure TiO2 and 3% Y-TiO2 are 148.58 (Figure 3a) and 160.17 m2/g (Figure 3b), respectively, which can be related to changes in the morphology of the TiO2 after doping with Y. Based on the XRD results, no diffraction peak shift is observed for 3% Y-doped TiO2, demonstrating that the Y3+ species exists at the crystal boundary or surface rather than in the inner crystalline structure of TiO2. Therefore, it is speculated that the adhesion of Y3+ species to the surface of particles increases the roughness of the particle surface, resulting in an increase in the specific surface area. The specific surface areas of TiO2-H2 and 3% Y-TiO2-H2 are 85.94 (Figure 3c) and 117.42 m2/g (Figure 3d), respectively, which is lower compared to pure TiO2 and 3% Y-TiO2. This may be attributed to the grain growth and slight aggregation caused by high-temperature annealing treatment, which is consistent with the results of the statistical particle size distribution in TEM images.
Based on the survey scanning the XPS spectra of four samples, all labeled peaks were attributed to Ti 2p and O 1s (Figure S2). Furthermore, a weak Y 3d peak is observed at ~158 eV, indicating the presence of Y3+ species on the surface of 3% Y-TiO2 and 3% Y-TiO2-H2 samples. The high-resolution XPS spectra of Ti 2p for four samples are shown in Figure 4a–d. TiO2 and 3% Y-TiO2 exhibit two peaks at 458.4 and 464.2 eV, which are attributed to 2p3/2 and 2p1/2 of Ti4+, respectively [37]. For TiO2-H2 and 3% Y-TiO2-H2, these peaks shift to low values, broaden, and become unsymmetrical compared with those of pure TiO2 and 3% Y-TiO2, indicating a different bonding environment (Figure S3) [38]. The fitting curves show that the two peaks of Ti 2p are divided into four peaks, found at 458.3 and 464.0 eV, respectively, corresponding to the 2p3/2 and 2p1/2 peaks of Ti4+, and at 457.8 and 463.2 eV, respectively, corresponding to the 2p3/2 and 2p1/2 peaks of Ti3+ [39]. The Ti3+ state indicates that TiO2-H2 and 3% Y-TiO2-H2 are partially reduced through H2 reduction.
The high-resolution XPS spectra of O 1s for four samples are shown in Figure 4e–h. In TiO2 and 3% Y-TiO2, two nonsymmetric fitted peaks signify the presence of two O species. The binding energy at 529.6 eV is ascribed to the characteristic peak of Ti-O in anatase TiO2, namely, lattice oxygen (OL). The binding energy at 531.4 eV is attributed to the O-H, i.e., adsorbed oxygen (OA) [40]. In TiO2-H2 and 3% Y-TiO2-H2, three nonsymmetric fitted peaks are obtained. Except OL (529.8 eV) and OA (531.9 eV), the small peak centered at around 530.9 eV can be assigned to OVs [41]. The estimated ratio of OV to OL (OV/OL) of TiO2-H2 is 0.09 according to the calculated integral areas of the corresponding peaks. For 3% Y-TiO2-H2, the ratio of OV/OL decreases to 0.07, indicating that the Y3+ on the surface or grain boundary of TiO2 by Y doping may slightly inhibit the formation of OVs. Furthermore, the content of OA decreases sharply for TiO2-H2 and 3% Y-TiO2-H2, implying that -OH groups or adsorbed water on the surface of TiO2-H2 and 3% Y-TiO2-H2 are largely scavenged during annealing.
Figure 5a shows the UV–vis DRS of four samples. The light absorption edge of pure TiO2 is approximately 390 nm. The light absorption edges of 3% Y-TiO2, TiO2-H2, and 3% Y-TiO2-H2 gradually blue-shift compared with that of TiO2, and the light absorption capacities in their visible region are gradually enhanced. The light absorption in the region from ~400 nm to the near-infrared region is significant in TiO2-H2 and higher in 3% Y-TiO2-H2.
The Eg can be calculated in accordance with the formula:
( α h v ) 1 / n = A h v E g ,
Among them, α, h, ν, and A are the absorption coefficient, Planck’s constant, frequency of the incident light, and a constant, respectively. For direct and indirect transition semiconductors, n is 1/2 and 2, respectively. The value of n is 2 for the anatate TiO2. In Figure 5b, the Eg values of TiO2, 3% Y-TiO2, TiO2-H2, and 3% Y-TiO2-H2 obtained from the tangent intercept are 2.88, 2.70, 2.24, and 2.17 eV, respectively.

3.2. Photocatalytic Test

The effect on the catalytic activities of all four samples is evaluated by the degradation of RhB under simulated sunlight irradiation, as shown in Figure 6. The UV–vis absorption spectra of TiO2, 3% Y-TiO2, TiO2-H2, 3% Y-TiO2-H2 are separately displayed in Figure 6a–d, and the degradation efficiency curves are summarized in Figure 6e. For pure TiO2, RhB can be totally degraded in approximately 60 min, whereas RhB can be totally degraded by 3% Y-TiO2 in 40 min, which is shorter than that by pure TiO2. This result indicates that a proper amount of RE doping has a positive effect on photocatalytic performance. For TiO2-H2, RhB can be totally degraded only in 30 min, which is beneficial for the reduction role of TiO2. Interestingly, RhB can be totally degraded by 3% Y-TiO2-H2 in 20 min, which is less time than those of other samples. The combined action of 3% Y doping and reduction by H2 is beneficial to the photodegradation activity of TiO2, and the co-modified catalyst exhibits higher photocatalytic activity than any single-modified catalyst.
The kinetics of the photocatalytic activities of all four samples follow the first-order reaction:
−ln(Ct/C0) = k1t,
where k1 is the pseudo-first-order reaction rate constant (min−1) obtained from the slope of −ln(Ct/C0) vs. t, as shown in Figure 6f,g. The k1 values of TiO2, 3% Y-TiO2, TiO2-H2, and 3% Y-TiO2-H2 are 0.0606, 0.0985, 0.1269, and 0.1746, respectively, indicating the efficient photodegradation activity of 3% Y-TiO2-H2. Therefore, the combination of Y doping and H2 reduction is responsible for the high degradation rate.

3.3. Mechanism Analysis

RE metal doping can significantly modify the electrical, physical, and chemical properties of TiO2 photocatalyst and play an effective role for improving photocatalytic performance. Compared to pure TiO2, 3% Y-doped TiO2 has a faster degradation rate of RhB, which is attributed to the influence of Y doping. The ionic radius of Y3+ (93 pm) is larger than that of Ti4+ (68 pm), which is difficult to replace Ti into the TiO2 lattice directly [25]. Based on the XRD results, no diffraction peak shift is observed for 3% Y-doped TiO2, demonstrating that Y3+ species exists at the crystal boundary or surface rather than in the inner crystalline structure of TiO2. Furthermore, the survey scanning the XPS spectra (Figure S2) shows that a weak Y 3d peak is observed at ~158 eV, indicating the presence of Y3+ species on the surface of 3% Y-TiO2. Based on XRD and XPS, the Y3+ species might be deposited on the surface of TiO2. Given that the work function (Φ) of RE metals is lower than that of titanium, RE3+ has more tendencies to attract the e from the sample surface compared with Ti ions, resulting in a reduced e–h+ pair recombination rate. In addition, 3% Y-TiO2 with a higher specific surface area has more reactive active sites compared with pure TiO2, which also helps to improve the photocatalytic performance. Moreover, UV–vis DRS revealed that 3% Y-TiO2 has higher light absorption capacities in their visible region than pure TiO2, resulting in lower Eg values. However, an excessive amount of metal dopants may lead to a decline in photodegradation efficiency towards pollutants by reducing the yield of photoinduced e–h+ pairs (Figure S1b). Therefore, 3% Y doping is the most suitable doping amount and selected as the subsequent research object in our work.
The reduction treatment of TiO2 by H2 is also one of the important methods to improve its photocatalytic performance. Compared to pure TiO2 and 3% Y-TiO2, the specific surface areas of TiO2-H2 significantly decreased, indicating that the reaction active sites decreased. However, TiO2-H2 also had a faster degradation rate of RhB, which is attributed to the changes in microstructure caused by reduction treatment. The remarkable feature of reduced TiO2 by H2 is its disordered shell filled with a limited amount of OVs or Ti3+, which is consistent with the HRTEM and XPS results. Furthemore, the significant decrease in Eg values of TiO2-H2 is due to the significant increase in the sample’s absorption capacity for visible light according to UV–vis DRS results. Compared with the RE metal doping, OV or Ti3+ is a kind of self-doping of the crystal itself without introducing any impurity element, which is considered to reflect the influence of the internal modification of TiO2 on its physicochemical and photocatalytic performance, such as tuning optical absorption, reducing Eg, and increasing carrier concentration [31].
Under the simultaneous action of Y doping and reduction treatment, 3% Y-TiO2-H2 exhibited a higher photocatalytic degradation ability than any single-modified catalyst. Thus, a possible photocatalytic mechanism for 3% Y-TiO2-H2 nanoparticles is proposed (Figure 7). The disordered shell of reduced 3% Y-TiO2-H2 is filled with a limited amount of OVs and Ti3+, which significantly reduces the Eg of TiO2 and remarkably increases the absorption of visible light. The formation of a defective energy level caused by OV and Ti3+ below the conduction band of TiO2 decreases the excitation energy, resulting in highly active photocatalysts [42]. The incorporation of Y3+ on the surface of TiO2 increases the charge separation by acting as an electron trapper, consequently produces more e for reaction on the surface of the catalyst, and reduces the e-h+ recombination rate of photocatalyst. In addition to reducing carrier recombination, as a member of RE elements, Y doping broadens the range of light absorption and increases the photoactivity of TiO2; this phenomenon is consistent with the DRS and photodegradation results. In conclusion, the more reactive active sites, the rapid interfacial charge transfer, and stronger optical absorption of 3% Y-TiO2-H2 caused by Y doping should be the reason for the higher rate of photocatalytic reactions than those of undoped TiO2-H2 [43]. Their photocatalytic process is such that, under simulated sunlight irradiation, e is excited from VB to CB of 3% Y-TiO2-H2, and h+ in VB is abandoned. Generated charge carriers react with oxygen molecules, water molecules, or OH to produce oxidative species for the degradation of organic dye in the aqueous solution, such as hydroxyl radicals (•OH) and (•O2).

4. Conclusions

Y doping and the reduction treatment of TiO2 have been widely proven to be an effective way to improve their photocatalytic performance, respectively. In this work, the coupling treatment of Y doping and reduction of TiO2 can further improve the photocatalytic performance, which is better than any individual method. In summary, pure TiO2 and 3% Y-TiO2 were prepared by a one-step hydrothermal method. Reduced TiO2-H2 and 3% Y-TiO2-H2 were obtained through the thermal conversion treatment of Ar-H2 atmosphere at 500 °C for 3 h. All samples show the single anatase phase, and no diffraction peak shift is observed. The photodegradation efficiency of pure TiO2, 3% Y-TiO2, TiO2-H2, and 3% Y-TiO2-H2 on RhB gradually increased. The 3% Y-TiO2-H2 exhibits the best photocatalytic performance for the degradation of RhB among these four samples, which can be totally degraded in 20 min. The key factors for the improved photocatalytic performance of 3% Y-TiO2-H2 could be attributed to the synergistic effect of Y doping and the reduction of TiO2. Y metal-ion doping broadened the range of light absorption and reduced the charge recombination rates. Reduction treatment can cause TiO2 to be enveloped by disordered shells, and OVs and Ti3+ species efficiently reduced the Eg of TiO2, remarkably increasing the absorption of light. The synergistic effect of rapid interfacial charge transfer and the stronger optical absorption of 3% Y-TiO2-H2 caused by Y doping and reduction treatment should be the reason for the high photocatalytic efficiency. The discovery of this work provides a new perspective for the improvement of other photocatalysts by combining doping and reduction to modify traditional photocatalytic materials and further improve their performance.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano13152266/s1, Figure S1: XRD patterns (a) and degradation curves (b) of pure TiO2 and various Y-doped TiO2 (1~5%) samples; Figure S2. Survey scanning XPS spectra of TiO2, 3% Y-TiO2, TiO2-H2 and 3% Y-TiO2-H2 samples. Figure S3. XPS spectra of Ti 2p for TiO2, 3%Y-TiO2, TiO2-H2 and 3%Y-TiO2-H2 samples.

Author Contributions

Conceptualization, X.L. (Xijuan Li) and H.Z.; experiment, X.L. (Xijuan Li) and H.Z., methodology and writing, X.L. (Xijuan Li) and H.Z.; formal analysis, X.L. (Xia Li), H.Z. and Y.W.; investigation, X.L. (Xijuan Li) and H.Z.; writing—original draft preparation, X.L. (Xijuan Li) and H.Z.; writing—review and editing, H.Z., Y.W. and J.L.; review—comment and editing, J.L., K.Y. and J.W.; supervision, H.Z. and K.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (NSFC No. U1904213 and 52202137), State Key Laboratory of Mechanics and Control of Mechanical Structures (Nanjing University of Aeronautics and Astronautics) (MCMS-E-0521G02), a Project Funded by the Priority Academic Program Development of Jiangsu Higher Education Institutions (PAPD), and Jiangsu Funding Program for Excellent Postdoctoral Talent.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data generated or analyzed during this work are included in this published article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  2. Fujishima, A.; Rao, T.N.; Tryk, D.A. Titanium dioxide photocatalysis. J. Photochem. Photobiol. C Photochem. Rev. 2000, 1, 27–47. [Google Scholar] [CrossRef]
  3. Chen, X.; Mao, S.S. Titanium Dioxide Nanomaterials:  Synthesis, Properties, Modifications, and Applications. Chem. Rev. 2007, 107, 2891–2959. [Google Scholar] [CrossRef] [PubMed]
  4. Guo, Q.; Zhou, C.; Ma, Z.; Yang, X. Fundamentals of TiO2 Photocatalysis: Concepts, Mechanisms, and Challenges. Adv. Mater. 2019, 31, 1901997. [Google Scholar] [CrossRef]
  5. Wu, J.; Hao, J.; Guo, W.; Chen, J.; Min, Y. Regulating photocatalysis by the oxidation state of titanium in TiO2/TiO. J. Colloid Interface Sci. 2022, 613, 616–624. [Google Scholar] [CrossRef]
  6. Liccardo, L.; Bordin, M.; Sheverdyaeva, P.M.; Belli, M.; Moras, P.; Vomiero, A.; Moretti, E. Surface Defect Engineering in Colored TiO2 Hollow Spheres Toward Efficient Photocatalysis. Adv. Funct. Mater. 2023, 33, 2212486. [Google Scholar] [CrossRef]
  7. Noman, M.T.; Ashraf, M.A.; Ali, A. Synthesis and applications of nano-TiO2: A review. Environ. Sci. Pollut. Res. 2019, 26, 3262–3291. [Google Scholar] [CrossRef]
  8. 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]
  9. Nakata, K.; Fujishima, A. TiO2 photocatalysis: Design and applications. J. Photochem. Photobiol. C Photochem. Rev. 2012, 13, 169–189. [Google Scholar] [CrossRef]
  10. Du, X.; Wu, Y.; Kou, Y.; Mu, J.; Yang, Z.; Hu, X.; Teng, F. Amorphous carbon inhibited TiO2 phase transition in aqueous solution and its application in photocatalytic degradation of organic dye. J. Alloys Compd. 2019, 810, 151917. [Google Scholar] [CrossRef]
  11. Bayan, E.M.; Lupeiko, T.G.; Pustovaya, L.E.; Volkova, M.G.; Butova, V.V.; Guda, A.A. Zn–F co-doped TiO2 nanomaterials: Synthesis, structure and photocatalytic activity. J. Alloys Compd. 2020, 822, 153662. [Google Scholar] [CrossRef]
  12. Naldoni, A.; Altomare, M.; Zoppellaro, G.; Liu, N.; Kment, S.; Zboril, R.; Schmuki, P. Photocatalysis with Reduced TiO2: From Black TiO2 to Cocatalyst-Free Hydrogen Production. ACS Catal. 2019, 9, 345–364. [Google Scholar] [CrossRef] [Green Version]
  13. Kusiak-Nejman, E.; Morawski, A.W. TiO2/graphene-based nanocomposites for water treatment: A brief overview of charge carrier transfer, antimicrobial and photocatalytic performance. Appl. Catal. B Environ. 2019, 253, 179–186. [Google Scholar] [CrossRef]
  14. Obata, K.; Kishishita, K.; Okemoto, A.; Taniya, K.; Ichihashi, Y.; Nishiyama, S. Photocatalytic decomposition of NH3 over TiO2 catalysts doped with Fe. Appl. Catal. B Environ. 2014, 160–161, 200–203. [Google Scholar] [CrossRef]
  15. Md Saad, S.K.; Ali Umar, A.; Ali Umar, M.I.; Tomitori, M.; Abd Rahman, M.Y.; Mat Salleh, M.; Oyama, M. Two-Dimensional, Hierarchical Ag-Doped TiO2 Nanocatalysts: Effect of the Metal Oxidation State on the Photocatalytic Properties. ACS Omega 2018, 3, 2579–2587. [Google Scholar] [CrossRef] [Green Version]
  16. Tbessi, I.; Benito, M.; Molins, E.; Liorca, J.; Touati, A.; Sayadi, S.; Najjar, W. Effect of Ce and Mn co-doping on photocatalytic performance of sol-gel TiO2. Solid State Sci. 2019, 88, 20–28. [Google Scholar] [CrossRef]
  17. Asahi, R.; Morikawa, T.; Ohwaki, T.; Aoki, K.; Taga, Y. Visible-Light Photocatalysis in Nitrogen-Doped Titanium Oxides. Science 2001, 293, 269–271. [Google Scholar] [CrossRef]
  18. Weon, S.; Kim, J.; Choi, W. Dual-components modified TiO2 with Pt and fluoride as deactivation-resistant photocatalyst for the degradation of volatile organic compound. Appl. Catal. B Environ. 2018, 220, 1–8. [Google Scholar] [CrossRef]
  19. Al-Azri, Z.H.N.; Chen, W.-T.; Chan, A.; Jovic, V.; Ina, T.; Idriss, H.; Waterhouse, G.I.N. The roles of metal co-catalysts and reaction media in photocatalytic hydrogen production: Performance evaluation of M/TiO2 photocatalysts (M=Pd, Pt, Au) in different alcohol–water mixtures. J. Catal. 2015, 329, 355–367. [Google Scholar] [CrossRef]
  20. Ayekoe, P.Y.; Robert, D.; Goné, D.L. Preparation of effective TiO2/Bi2O3 photocatalysts for water treatment. Environ. Chem. Lett. 2016, 14, 387–393. [Google Scholar] [CrossRef]
  21. Chen, Y.; Huang, W.; He, D.; Situ, Y.; Huang, H. Construction of Heterostructured g-C3N4/Ag/TiO2 Microspheres with Enhanced Photocatalysis Performance under Visible-Light Irradiation. ACS Appl. Mater. Interfaces 2014, 6, 14405–14414. [Google Scholar] [CrossRef] [PubMed]
  22. Tao, X.; Zhu, L.; Wang, X.; Chen, X.; Liu, X. Preparation of Zr/Y co-doped TiO2 photocatalyst and degradation performance of hydroquinone. Environ. Sci. Pollut. Res. 2022, 29, 40854–40864. [Google Scholar] [CrossRef] [PubMed]
  23. Li, J.; Chu, B.; Xie, Z.; Deng, Y.; Zhou, Y.; Dong, L.; Li, B.; Chen, Z. Mechanism and DFT Study of Degradation of Organic Pollutants on Rare Earth Ions Doped TiO2 Photocatalysts Prepared by Sol-Hydrothermal Synthesis. Catal. Lett. 2022, 152, 489–502. [Google Scholar] [CrossRef]
  24. Bingham, S.; Daoud, W.A. Recent advances in making nano-sized TiO2 visible-light active through rare-earth metal doping. J. Mater. Chem. 2011, 21, 2041–2050. [Google Scholar] [CrossRef]
  25. Reszczyńska, J.; Grzyb, T.; Wei, Z.; Klein, M.; Kowalska, E.; Ohtani, B.; Zaleska-Medynska, A. Photocatalytic activity and luminescence properties of RE3+–TiO2 nanocrystals prepared by sol–gel and hydrothermal methods. Appl. Catal. B Environ. 2016, 181, 825–837. [Google Scholar] [CrossRef]
  26. Gao, H.; Liu, J.; Zhang, J.; Zhu, Z.; Zhang, G.; Liu, Q. Influence of carbon and yttrium co-doping on the photocatalytic activity of mixed phase TiO2. Chin. J. Catal. 2017, 38, 1688–1696. [Google Scholar] [CrossRef]
  27. Wang, Y.; Lu, K.; Feng, C. Photocatalytic degradation of methyl orange by polyoxometalates supported on yttrium-doped TiO2. J. Rare Earths 2011, 29, 866–871. [Google Scholar] [CrossRef]
  28. Wu, Y.; Gong, Y.; Liu, J.; Zhang, Z.; Xu, Y.; Ren, H.; Li, C.; Niu, L. B and Y co-doped TiO2 photocatalyst with enhanced photodegradation efficiency. J. Alloys Compd. 2017, 695, 1462–1469. [Google Scholar] [CrossRef]
  29. Tian, X.; Huang, S.; Wang, L.; Li, L.; Lou, Z.; Huang, S.; Zhou, Z. Mitigation of low methane content landfill gas through visible-near-infrared photocatalysis over Y2O3:Er3+/Graphene/TiO2. Appl. Surf. Sci. 2018, 456, 854–860. [Google Scholar] [CrossRef]
  30. Xing, M.; Zhang, J.; Chen, F.; Tian, B. An economic method to prepare vacuum activated photocatalysts with high photo-activities and photosensitivities. Chem. Commun. 2011, 47, 4947–4949. [Google Scholar] [CrossRef]
  31. Eom, J.-Y.; Lim, S.-J.; Lee, S.-M.; Ryu, W.-H.; Kwon, H.-S. Black titanium oxide nanoarray electrodes for high rate Li-ion microbatteries. J. Mater. Chem. A 2015, 3, 11183–11188. [Google Scholar] [CrossRef]
  32. Tian, M.; Mahjouri-Samani, M.; Eres, G.; Sachan, R.; Yoon, M.; Chisholm, M.F.; Wang, K.; Puretzky, A.A.; Rouleau, C.M.; Geohegan, D.B.; et al. Structure and Formation Mechanism of Black TiO2 Nanoparticles. ACS Nano 2015, 9, 10482–10488. [Google Scholar] [CrossRef]
  33. Wang, Z.; Yang, C.; Lin, T.; Yin, H.; Chen, P.; Wan, D.; Xu, F.; Huang, F.; Lin, J.; Xie, X.; et al. Visible-light photocatalytic, solar thermal and photoelectrochemical properties of aluminium-reduced black titania. Energy Environ. Sci. 2013, 6, 3007–3014. [Google Scholar] [CrossRef]
  34. Chen, J.; Ding, Z.; Wang, C.; Hou, H.; Zhang, Y.; Wang, C.; Zou, G.; Ji, X. Black Anatase Titania with Ultrafast Sodium-Storage Performances Stimulated by Oxygen Vacancies. ACS Appl. Mater. Interfaces 2016, 8, 9142–9151. [Google Scholar] [CrossRef]
  35. Zhu, Q.; Peng, Y.; Lin, L.; Fan, C.-M.; Gao, G.-Q.; Wang, R.-X.; Xu, A.-W. Stable blue TiO2−x nanoparticles for efficient visible light photocatalysts. J. Mater. Chem. A 2014, 2, 4429–4437. [Google Scholar] [CrossRef]
  36. Zuo, F.; Wang, L.; Wu, T.; Zhang, Z.; Borchardt, D.; Feng, P. Self-Doped Ti3+ Enhanced Photocatalyst for Hydrogen Production under Visible Light. J. Am. Chem. Soc. 2010, 132, 11856–11857. [Google Scholar] [CrossRef]
  37. Wei, K.; Wang, B.; Hu, J.; Chen, F.; Hao, Q.; He, G.; Wang, Y.; Li, W.; Liu, J.; He, Q. Photocatalytic properties of a new Z-scheme system BaTiO3/In2S3 with a core–shell structure. RSC Adv. 2019, 9, 11377–11384. [Google Scholar] [CrossRef]
  38. Yan, Y.; Hao, B.; Wang, D.; Chen, G.; Markweg, E.; Albrecht, A.; Schaaf, P. Understanding the fast lithium storage performance of hydrogenated TiO2 nanoparticles. J. Mater. Chem. A 2013, 1, 14507–14513. [Google Scholar] [CrossRef]
  39. Awan, I.T.; Lozano, G.; Pereira-da-Silva, M.A.; Romano, R.A.; Rivera, V.A.G.; Ferreira, S.O.; Marega, E. Understanding the electronic properties of BaTiO3 and Er3+ doped BaTiO3 films through confocal scanning microscopy and XPS: The role of oxygen vacancies. Phys. Chem. Chem. Phys. 2020, 22, 15022–15034. [Google Scholar]
  40. Jiang, L.; Huang, Y.; Liu, T. Enhanced visible-light photocatalytic performance of electrospun carbon-doped TiO2/halloysite nanotube hybrid nanofibers. J. Colloid Interface Sci. 2015, 439, 62–68. [Google Scholar] [CrossRef]
  41. Shi, L.; Zhou, W.; Li, Z.; Koul, S.; Kushima, A.; Yang, Y. Periodically Ordered Nanoporous Perovskite Photoelectrode for Efficient Photoelectrochemical Water Splitting. ACS Nano 2018, 12, 6335–6342. [Google Scholar] [CrossRef] [PubMed]
  42. Reszczyńska, J.; Grzyb, T.; Sobczak, J.W.; Lisowski, W.; Gazda, M.; Ohtani, B.; Zaleska, A. Visible light activity of rare earth metal doped (Er3+, Yb3+ or Er3+/Yb3+) titania photocatalysts. Appl. Catal. B Environ. 2015, 163, 40–49. [Google Scholar] [CrossRef]
  43. Reszczyńska, J.; Grzyb, T.; Sobczak, J.W.; Lisowski, W.; Gazda, M.; Ohtani, B.; Zaleska, A. Lanthanide co-doped TiO2: The effect of metal type and amount on surface properties and photocatalytic activity. Appl. Surf. Sci. 2014, 307, 333–345. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of pure TiO2, 3% Y-TiO2, TiO2-H2, and 3% Y-TiO2-H2 samples.
Figure 1. XRD patterns of pure TiO2, 3% Y-TiO2, TiO2-H2, and 3% Y-TiO2-H2 samples.
Nanomaterials 13 02266 g001
Figure 2. HRTEM images of (a,e) TiO2, (b,f) 3% Y-TiO2, (c,g) TiO2-H2, and (d,h) 3% Y-TiO2-H2 (inset is the corresponding statistical particle size distribution). (i) Elemental mapping of 3% Y-TiO2-H2 for Ti (i1), O (i2), and Y (i3).
Figure 2. HRTEM images of (a,e) TiO2, (b,f) 3% Y-TiO2, (c,g) TiO2-H2, and (d,h) 3% Y-TiO2-H2 (inset is the corresponding statistical particle size distribution). (i) Elemental mapping of 3% Y-TiO2-H2 for Ti (i1), O (i2), and Y (i3).
Nanomaterials 13 02266 g002
Figure 3. N2 adsorption–desorption isotherms of TiO2 (a), 3% Y-TiO2 (b), TiO2-H2 (c), and 3% Y-TiO2-H2 (d).
Figure 3. N2 adsorption–desorption isotherms of TiO2 (a), 3% Y-TiO2 (b), TiO2-H2 (c), and 3% Y-TiO2-H2 (d).
Nanomaterials 13 02266 g003
Figure 4. XPS spectra for Ti 2p of TiO2 (a), 3% Y-TiO2 (b), TiO2-H2 (c), 3% Y-TiO2-H2 (d), and for O 1s of TiO2 (e), 3% Y-TiO2 (f), TiO2-H2 (g), 3% Y-TiO2-H2 (h).
Figure 4. XPS spectra for Ti 2p of TiO2 (a), 3% Y-TiO2 (b), TiO2-H2 (c), 3% Y-TiO2-H2 (d), and for O 1s of TiO2 (e), 3% Y-TiO2 (f), TiO2-H2 (g), 3% Y-TiO2-H2 (h).
Nanomaterials 13 02266 g004
Figure 5. (a) UV–vis diffuse reflectance absorption spectra and (b) corresponding bandgap of TiO2, 3% Y-TiO2, TiO2-H2, and 3% Y-TiO2-H2 samples.
Figure 5. (a) UV–vis diffuse reflectance absorption spectra and (b) corresponding bandgap of TiO2, 3% Y-TiO2, TiO2-H2, and 3% Y-TiO2-H2 samples.
Nanomaterials 13 02266 g005
Figure 6. UV–vis absorption spectra of TiO2 (a), 3% Y-TiO2 (b), TiO2-H2 (c), 3% Y-TiO2-H2 (d), degradation efficiency curves (e), kinetic curves (f), and reaction rate constant (g).
Figure 6. UV–vis absorption spectra of TiO2 (a), 3% Y-TiO2 (b), TiO2-H2 (c), 3% Y-TiO2-H2 (d), degradation efficiency curves (e), kinetic curves (f), and reaction rate constant (g).
Nanomaterials 13 02266 g006
Figure 7. Schematic of the photodegradation of RhB in 3% Y-TiO2-H2 nanoparticles.
Figure 7. Schematic of the photodegradation of RhB in 3% Y-TiO2-H2 nanoparticles.
Nanomaterials 13 02266 g007
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

Li, X.; Zheng, H.; Wang, Y.; Li, X.; Liu, J.; Yan, K.; Wang, J.; Zhu, K. Synergistic Effect of Y Doping and Reduction of TiO2 on the Improvement of Photocatalytic Performance. Nanomaterials 2023, 13, 2266. https://doi.org/10.3390/nano13152266

AMA Style

Li X, Zheng H, Wang Y, Li X, Liu J, Yan K, Wang J, Zhu K. Synergistic Effect of Y Doping and Reduction of TiO2 on the Improvement of Photocatalytic Performance. Nanomaterials. 2023; 13(15):2266. https://doi.org/10.3390/nano13152266

Chicago/Turabian Style

Li, Xijuan, Hongjuan Zheng, Yulong Wang, Xia Li, Jinsong Liu, Kang Yan, Jing Wang, and Kongjun Zhu. 2023. "Synergistic Effect of Y Doping and Reduction of TiO2 on the Improvement of Photocatalytic Performance" Nanomaterials 13, no. 15: 2266. https://doi.org/10.3390/nano13152266

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