Next Article in Journal
Device and Circuit Analysis of Double Gate Field Effect Transistor with Mono-Layer WS2-Channel at Sub-2 nm Technology Node
Next Article in Special Issue
Synthesis of Uniform Size Rutile TiO2 Microrods by Simple Molten-Salt Method and Its Photoluminescence Activity
Previous Article in Journal
Significant Modulation of Vortex Resonance Spectra in a Square-Shape Ferromagnetic Dot
Previous Article in Special Issue
Efficient Charge Transfer Channels in Reduced Graphene Oxide/Mesoporous TiO2 Nanotube Heterojunction Assemblies toward Optimized Photocatalytic Hydrogen Evolution
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Construction of Spindle-Shaped Ti3+ Self-Doped TiO2 Photocatalysts Using Triethanolamine-Aqueous as the Medium and Its Photoelectrochemical Properties

1
School of Materials Science and Engineering, Linyi University, Linyi 276005, China
2
School of Chemistry & Chemical Engineering, Linyi University, Linyi 276005, China
3
Yantai North China Microwave Technology Co., Ltd., Yantai 264025, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(13), 2298; https://doi.org/10.3390/nano12132298
Submission received: 7 June 2022 / Revised: 25 June 2022 / Accepted: 2 July 2022 / Published: 4 July 2022
(This article belongs to the Special Issue Synthesis of TiO2 Nanoparticles and Their Catalytic Activity)

Abstract

:
To enhance the utilization efficiency of visible light and reduce the recombination of photogenerated electrons and holes, spindle-shaped TiO2 photocatalysts with different Ti3+ concentrations were fabricated by a simple solvothermal strategy using low-cost, environmentally friendly TiH2 and H2O2 as raw materials and triethanolamine-aqueous as the medium. The photocatalytic activities of the obtained photocatalysts were investigated in the presence of visible light. X-ray diffraction (XRD), Raman spectra, transmission electron microscope (TEM), X-ray photoelectron spectroscopy (XPS), and Fourier transform infrared (FT-IR) spectra were applied to characterize the structure, morphologies, and chemical compositions of as-fabricated Ti3+ self-doped TiO2. The concentration of triethanolamine in the mixed solvent plays a significant role on the crystallinity, morphologies, and photocatalytic activities. The electron–hole separation efficiency was found to increase with the increase in the aspect ratio of as-fabricated Ti3+ self-doped TiO2, which was proved by transient photocurrent response and electrochemical impedance spectroscopy.

1. Introduction

As one of the most promising strategies to deal with the worldwide environmental and energy crises, photocatalytic (PC) technology takes particular attention of being a green technology that can be applied under solar conditions. PC activity is largely determined by the separation efficiency of photogenerated electrons and holes with high chemical energy [1,2]. Due to its earth abundance, nontoxic high-photoelectrical conversion performance [3], and strong chemical stability against photo and chemical corrosion, TiO2 has been extensively applied as photocatalyst for solar hydrogen production [4], photovoltaic power generation [5,6], sewage purification [7,8,9], air purification, and other fields [10,11]. Hence, TiO2-based photocatalysts with high PC activity have received continuous attention. However, two defects, wide band-gap (3.2 eV for pure TiO2) and rapid charge recombination rate, limit the large-scale application of TiO2-based photocatalysts [12,13]. Hence, PC activity may be enhanced by comprehensive studies on band engineering and charge carrier dynamics to solve the above problems of pure TiO2 photocatalyst [14,15].
Many strategies have been developed during the past few decades to enhance the absorption of visible light [1,16]. Heteroatom doping is the most basic way of changing the electronic structure of functional materials; for instance C [17], S [18,19], and N doping [20,21], acting as electronic donation, and receipt have long been actively used to modify wide band-gap TiO2 photocatalysts for visible light absorption. In addition, surface modification strategies also were applied to improve the PC activity of TiO2, such as transition metal doping, dye sensitization, inorganic combination, and nanostructure-tuning [22]. These efforts usually suffer from the leakage of harmful metal ions, thermal instability, and the increase in deleterious recombination centers for photogenerated charges [23,24,25,26]. These vital drawbacks have increased the limited use of TiO2-based materials in developing photocatalysts. Therefore, reasonable modification strategies play a major role in improving PC properties of TiO2-based photocatalysts.
As a typical n-type semiconductor, TiO2 possesses intrinsic oxygen deficiency. The introduction of Ti3+ and oxygen vacancies (OVs) was employed to broaden the visible light absorption range and promote the charge separation [13,27,28]. Hence, Ti3+ self-doped TiO2 nanomaterials have been widely investigated for their extended visible light absorption and high conductivity [29,30]. Inducting Ti3+ and OVs would create new defect states below the conduction band of TiO2, thus increasing its PC activity [31,32,33,34]. Thus far, various strategies for introducing Ti3+ are often used; examples are thermal annealing at a high temperature in various reducing atmospheres, such as H2, CO, H2/Ar, or vacuum [35,36,37]. Nevertheless, the above strategies face the problems of high equipment requirements and high manufacturing costs. Hence, developing a facile strategy to construct Ti3+ self-doped TiO2 is still a potential research topic.
Bulk and surface studies cannot elucidate the shape dependence effect, which is regarded as the origin of the higher performance of smaller particles. Serial organic surfactants, such as PVP [38], PEG [39], CTAB [40], thiophenol [41], and TEA [42] have been employed to tune the size and morphology and to prevent agglomeration. The introduction of surfactant could also stabilize the active adsorption sites that are conducive to the formation of smaller TiO2-based photocatalysts [43,44]. In our previous work, reduced TiO2 photocatalysts with different OVs contents were prepared by hydrothermal method, and the effects of hydrothermal treatment temperature and time consumed on the structure, morphology, and properties of the obtained samples were studied [45,46]. However, the solvent effect has an important impact on the morphology of the product, thus affecting the PC performance of the products. As a soluble, inexpensive, and readily available organic surfactant, triethanolamine (N(CH2CH2OH)3, TEA) plays a vital role in the regulation of crystallinity, morphology, and electronic structure of the TiO2-based photocatalysts [47,48,49,50,51,52,53]. Herein, to optimize the PC performance of TiO2, a facile mixed solvothermal method was developed to fabricate spindle-shaped Ti3+ self-doped TiO2 with TEA as structure-directing and capping agent.

2. Construction of Spindle-Shaped TiO2 Photocatalysts with Ti3+ and OVs

Ti3+ self-doped TiO2 photocatalysts were prepared with TiH2 as Ti source, H2O2 as oxidant, and TEA as structure-directing agent. The entire synthesis process is illustrated in Scheme 1. The sol obtained by oxidizing TiH2 with H2O2 was evenly divided into five parts, and then different amounts of TEA and H2O were added, respectively. The solvothermal reaction subsequently was conducted at 180 °C for 20 h. The products were cleaned with deionized water and absolute ethanol three times, and centrifugate was collected. After drying at 60 °C, light blue photocatalysts were obtained.
The chemicals, detailed preparation process, characterization, PC, and photoelectrochemical measurement experiment of the spindle-shaped TiO2 with Ti3+ self-doped are detailed in the Supplementary Materials.

3. Results and Discussion

Ti3+ self-doped TiO2 nanoparticles were prepared with different concentrations of TEA (0, 10, 20, 30, and 40 mL) via the solvothermal method at 180 °C for 20 h. The corresponding TiO2 nanoparticles were denoted as T00, T10, T20, T30, and T40, respectively. The morphology, structure, and composition of nanomaterials are known to have a direct impact on their PC performance. X-ray diffraction (XRD) was applied to study the structure and crystallization of obtained photocatalysts. As depicted in Figure 1a, those peaks located at 25.3°, 37.9°, 48.0°, 54.1°, 54.9°, 62.7°, and 68.9° represented the (101), (004), (200), (105), (211), (204), and (216) crystal planes of pure TiO2, maintaining the standard card of anatase TiO2 (JCPDS No. 21–1272) [16,54]. The peaks of TiH2 were not found in the XRD results, which proved that TiH2 was converted into anatase TiO2. Furthermore, with the increase in TEA content from 10 to 40 mL, the intensities of diffraction peaks decreased gradually, suggesting a decreased crystallinity for TiO2. It might be attributed to the formation of defects with high content of TEA (Figure 1b). With the increase in TEA amount, the half-width was gradually increasing, indicating that the crystallinity of the sample was deteriorated. The above XRD results show that the content of TEA had a direct effect on the crystallinity of TiO2 photocatalysts.
As a simple, effective, and high-sensitivity detection technology, Raman spectra, originating from the vibration of molecular bonds, were employed to explore the significant structural changes in TiO2. As shown in Figure 2a, these peaks located at 145, 398, 514, and 636 cm−1 could index, respectively, to Eg, B1g, A1g, and Eg lattice vibration modes, indicating that all these nanomaterials were mainly anatase TiO2 [55]. As illustrated in Figure 2b, the strongest Raman bands around 145 cm−1 shifted to a lower wavenumber along with the peak broadening as the content of TEA increased from 10 to 40 mL, which might be attributed to the increase in the aspect ratio of spindle-shaped TiO2 nanoparticles. At the same time, the shift and broadening content reached the highest with 40 mL TEA, indicating that there existed a large number of defects in T40.
Transmission electron microscopy (TEM) and HRTEM were employed to investigate the representative microstructures and crystalline properties of as-prepared nanomaterials. As shown in Figure 3, as-obtained samples were spindle-shaped particles. It was easy to find that the TEA concentration in the medium significantly affected the morphology of TiO2 nanomaterials. In absence of TEA, the obtained nanomaterials were a spherical-like particle (Figure S1a,b). The morphologies of as-obtained TiO2 nanomaterials become spindle-shaped in the presence of TEA. With the increase in TEA amount, the spindle-shaped nanorods became narrower and longer (Figure 3a–c). However, when the amount of TEA reached to 40 mL, the obtained sample become fragmented (Figure 3e), which further indicated that the amount of TEA had an important effect on the morphology of the TiO2 nanomaterials. The HRTEM results demonstrated that the crystallinity of the TiO2 photocatalysts decreased gradually with the increase in TEA (Figure 3d,f and Figure S1c,d). The crystal lattices of T00 were measured at 0.351 nm ), which matched the (101) crystal plane of anatase TiO2. While with increase in TEA, the crystal lattices of T10, T20, T30, and T40 decreased from 0.351 nm to 0.346 nm (Figure S1c,d and Figure 3d,f), indicating that the modification of TEA would affect the crystallinity of TiO2 nanomaterials.
The XPS analysis is applied to further study the surface chemical environment of as-obtained TiO2 photocatalysts. The survey XPS patterns showed that as-fabricated TiO2 photocatalysts all contain O, C, and Ti (Figure S2). With the increase in TEA amounts in the medium, the N element binding energy peak appeared, indicating that the product contains N element on its surface. As illustrated in Figure 4a, the high-resolution XPS spectrum of Ti 2p for T00, T40 showed two peaks, which can be fitted into four peaks at 457.9, 458.6, 463.4, and 464.3 eV, assigning to Ti3+ 2p3/2, Ti4+ 2p3/2, Ti3+ 2p1/2, and Ti4+ 2p1/2, respectively [56,57]. Similarly, the XPS spectrum of O 1s for T00 and T40 also showed two peaks, which can be fitted into three peaks at 529.6, 530.5, and 532.2 eV, assigning to Ti–O bond, O–H, and OVs [58,59]. Compared with T00, the content of Ti3+ and OVs in the sample with TEA was greatly increased, indicating that the addition of TEA could not only form a one-dimensional spindle product but also regulate the content of Ti3+ and OVs in the product. Through comparative analysis of the binding energy peaks of Ti3+ 2p3/2 in all obtained samples, the content of Ti3+ was found to increase with the increase in TEA amount in the mixed solvent, which further demonstrated the influence of solvent effect on the product composition (Supporting Information, Figure S3). On one hand, a one-dimensional structure was conducive to the separation of photogenerated electrons and holes [55]. On the other hand, the introduction of Ti3+/OVs could lower the band-gap of TiO2, which facilitated the application of visible light and the improvement of PC performance. The C 1s spectra of T00, shown in Figure 4c, was deconvolved into one peak located at 284.7 eV, assigning to the C–C bond with sp2 orbital [60]. As comparison, C 1s spectra of T20 were fitted into two peaks at 284.7 and 288.5 eV, the latter of which was assigned as C–N bond. Similarly, the C 1s spectra of T40 were deconvolved into three peaks at 284.7, 286.1, and 288.5 eV, among which the peak at 286.1 eV was assigned to C–H with sp3 hybridization, indicating the introduction of TEA. Moreover, high-resolution N 1s spectra of T20 and T40 showed one more peak at 399.8 eV compared to the N 1s spectra of T00, which was assigned to the C–N–C bond (Figure 4d), further confirming the introduction of TEA on the surface of TiO2. A binding energy peak of Ti–C or Ti–N bond was not in all samples, which indicated that the C or N doping is not formed during the solvothermal treatment, and TEA only attached to the surface of TiO2 by physical adsorption. In addition, the FT-IR analysis results also proved that TEA anchored onto TiO2 photocatalysts (Supporting Information, Figure S4).
The optical performance of as-fabricated TiO2 photocatalysts was investigated by UV-vis diffuse reflectance spectra (DRS). As illustrated in Figure 5a, with TEA as a structure-directing and capping agent, the photo-response performance of TiO2 photocatalysts was greatly improved, extending to visible and infrared light regions. Furthermore, T40 exhibited the highest UV-vis absorption property than the other samples, indicating that the introduction Ti3+/OVs could produce a novel vacancy band located below the conduction band edge of anatase TiO2. As illustrated in Figure 5b, the band-gap energy of T40 was calculated to be 2.93 eV via plotting Kubelka–Munk formula against the photon energy. The band energy of T30, T20, T10, T00, and anatase TiO2 were calculated to be 3.00, 3.03, 3.08, 3.05, and 3.12 eV, which were larger than that of T40.
Photoelectrochemical transient photocurrent response under visible light can feasibly investigate the photogenerated electrons transfer property [61]. The instantaneous photocurrent data of those photocatalysts were investigated under visible light irradiation at an interval of 50 s. As displayed in Figure 6a, the photocurrent density of T10 was measured as 1.3 μA cm−2, while that of T20 and T30 were 2.8 and 3.6 μA cm−2, respectively. The photocurrent disappeared immediately after the visible light was removed, indicating that the photocurrent was completely generated by the photoelectrodes. Charge separation and transfer of photocatalysts were obtained by electrochemical impedance spectroscopy [62]. Figure 6b shows that the radius of the arch for T30 was smallest among those prepared TiO2 photocatalysts, indicating that T30 exhibited the lowest impedance. The above experimental results showed that Ti3+ self-doped and spindle structure were beneficial to the separation and transfer efficiency of photogenerated electron-holes.
The PC performance of as-prepared TiO2 photocatalysts was investigated by the degradation of Rhodamine B (RhB) when exposed to visible light. The time-dependent curve of concentration and spectrum during RhB degradation is shown in Figure 7a. Figure 7a also shows that the photodegradation of RhB dyes was not observed without photocatalyst. This indicates that RhB cannot be degraded under visible light while T30 sample showed the strongest PC activity, which was mainly attributed to its aspect ratio, allowing the generation of photogenerated electrons and holes and electron transfer to the surface. The stability of photocatalyst had an important influence on its practical application. Recycling experiment was employed to investigate the stability of as-prepared photocatalysts. As illustrated in Figure 7b, after six cycles, the PC performance of T30 had almost no loss, indicating that as-obtained possessed high PC stability and high practical application had potential.
Generally, the PC activity is determined directly by the separation efficiency of photogenerated electrons and holes. In addition, it is widely recognized that photocatalysts with a larger specific surface can supply more surface active sites for the adsorption of RhB molecules, resulting in an enhanced PC performance. The BET surface results show that T00 sample has the largest surface area, while T30 sample has the smallest surface (Supporting Information, Figure S5); however, T30 has the best PC performance, which indicates that the specific surface areas are not the main factor affecting the PC performance. In addition, Figure 6 shows that T30 sample has the largest photocurrent and the smallest impedance, indicating that T30 has the higher separation effect of photogenerated electrons and holes. Therefore, the enhanced PC activity of Ti3+ self-doped TiO2 photocatalysts is attributed to the following reasons: (i) The introduction of Ti3+/OVs defects are used to reduce the band-gap of TiO2 into the visible light range and subsequently the new defect states appear below conduction band, and (ii) the spindle-shaped nanoparticles are beneficial to the separation of photogenerated electron holes and electron transport [55]. In addition, the introduction of Ti3+/OVs enhances the charge separation and diffusion while suppressing charge carrier recombination. When exposed to visible light, photogenerated electrons could change from valence band to conduction band, promoting the migration of electrons and holes to the surface TiO2 nanoparticles. As displayed in Figure 8, photogenerated electrons react with dissolved O2 to produce superoxide anion radicals (•O2), while holes oxidize the surface OH- groups and H2O molecules to generate •OH. The highly reductive •O2 and oxidizing •OH will decompose RhB molecules into small micromolecules, such as H2O and CO2. Nevertheless, excessive Ti3+ would lead to the formation of the recombination center for photogenerated electron–hole pairs, which were harmful to the improvement of photocatalytic performance of TiO2.

4. Conclusions

In conclusion, spindle-shaped Ti3+ self-doped TiO2 photocatalysts were fabricated by a simple solvothermal strategy with triethanolamine as a structure-directing agent. The morphology and photochemical properties of TiO2 photocatalysts could be effectively controlled by adjusting the amount of trimethylamine. In addition, the crystallinity, photogenerated electrons, and holes of the TiO2 catalysts could also be effectively controlled by changing the amount of trimethylamine. As-prepared TiO2 photocatalysts were also employed for the photocatalytic degradation of rhodamine B. It was found that T30 exhibited the highest catalytic efficiency, which was mainly due to the large aspect ratio. Hence, the Ti3+ self-doped TiO2 photocatalysts with large aspect ratio will be a feasible alternative strategy for water treatment and organic degradation in the future.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano12132298/s1, Materials and Methods. Figure S1: TEM and HRTEM images of the samples obtained at different triethanolamine concentrations; Figure S2: Survey XPS spectra of the obtained samples at different TEA amount; Figure S3: High-resolution XPS spectrum of Ti3+ 2p3/2 for different samples; Figure S4: FTIP spectra for pure TiO2 and obtained samples with different TEA amount; Figure S5: N2 adsorption-desorption isotherms and the corresponding BET surface area of the different samples.

Author Contributions

Conceptualization, Z.H., X.Z. and S.G.; methodology, Z.H. and J.W.; validation, X.Z. and A.W.; investigation, Z.H., Q.G. and J.W.; resources, X.Z. and S.G.; data curation, Z.H., Q.G., J.W. and A.W.; writing—original draft preparation, Z.H. and S.G.; writing—review and editing, Z.H. and S.G.; visualization, S.G.; supervision, X.Z. and S.G.; project administration, S.G.; funding acquisition, Z.H. and S.G. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Natural Science Foundation of Shandong Province (Grant Number: ZR2019MB019, ZR2020QB170).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data presented in this article are available at request from the corresponding author.

Conflicts of Interest

The authors declare no competing interests.

References

  1. Kang, X.; Song, X.Z.; Han, Y.; Cao, J.; Tan, Z. Defect-engineered TiO2 Hollow Spiny Nanocubes for Phenol Degradation under Visible Light Irradiation. Sci. Rep. 2018, 8, 5904. [Google Scholar] [CrossRef] [PubMed]
  2. Fang, L.; Chen, J.; Zhang, M.; Jiang, X.; Sun, Z. Introduction of Ti3+ ions into heterostructured TiO2 nanotree arrays for enhanced photoelectrochemical performance. Appl. Surf. Sci. 2019, 490, 1–6. [Google Scholar] [CrossRef]
  3. Ma, X.; Wang, C.; Wu, F.; Guan, Y.; Xu, G. TiO2 nanomaterials in photoelectrochemical and electrochemiluminescent biosensing. Top. Curr. Chem. 2020, 378, 28. [Google Scholar] [CrossRef] [PubMed]
  4. Li, J.; Cushing, S.K.; Zheng, P.; Senty, T.; Meng, F.; Bristow, A.D.; Manivannan, A.; Wu, N. Solar hydrogen generation by a CdS-Au-TiO2 sandwich nanorod array enhanced with Au nanoparticle as electron relay and plasmonic photosensitizer. J. Am. Chem. Soc. 2014, 136, 8438–8449. [Google Scholar] [CrossRef]
  5. Pradhan, S.C.; Velore, J.; Hagfeldt, A.; Soman, S. Probing photovoltaic performance in copper electrolyte dye-sensitized solar cells of variable TiO2 particle size using comprehensive interfacial analysis. J. Mater. Chem. C 2022, 10, 3929–3936. [Google Scholar] [CrossRef]
  6. Xiu, Z.; Guo, M.; Zhao, T.; Pan, K.; Xing, Z.; Li, Z.; Zhou, W. Recent advances in Ti3+ self-doped nanostructured TiO2 visible light photocatalysts for environmental and energy applications. Chem. Eng. J. 2020, 382, 123011. [Google Scholar] [CrossRef]
  7. Deng, X.; Zhang, H.; Guo, R.; Ma, Q.; Cui, Y.; Cheng, X.; Xie, M.; Cheng, Q. Effect of Ti3+ on enhancing photocatalytic and photoelectrochemical properties of TiO2 nanorods/nanosheets photoelectrode. Sep. Purif. Technol. 2018, 192, 329–339. [Google Scholar] [CrossRef]
  8. Cheng, D.; Li, Y.; Yang, L.; Luo, S.; Yang, L.; Luo, X.; Luo, Y.; Li, T.; Gao, J.; Dionysiou, D.D. One-step reductive synthesis of Ti3+ self-doped elongated anatase TiO2 nanowires combined with reduced graphene oxide for adsorbing and degrading waste engine oil. J. Hazard. Mater. 2019, 378, 120752. [Google Scholar] [CrossRef]
  9. Hao, W.; Li, X.; Qin, L.; Han, S.; Kang, S.-Z. Facile preparation of Ti3+ self-doped TiO2 nanoparticles and their dramatic visible photocatalytic activity for the fast treatment of highly concentrated Cr(vi) effluent. Catal. Sci. Technol. 2019, 9, 2523–2531. [Google Scholar] [CrossRef]
  10. Wajid Shah, M.; Zhu, Y.; Fan, X.; Zhao, J.; Li, Y.; Asim, S.; Wang, C. Facile Synthesis of Defective TiO2-x Nanocrystals with High Surface Area and Tailoring Bandgap for Visible-light Photocatalysis. Sci. Rep. 2015, 5, 15804. [Google Scholar] [CrossRef] [Green Version]
  11. Liu, G.; Han, K.; Ye, H.; Zhu, C.; Gao, Y.; Liu, Y.; Zhou, Y. Graphene oxide/triethanolamine modified titanate nanowires as photocatalytic membrane for water treatment. Chem. Eng. J. 2017, 320, 74–80. [Google Scholar] [CrossRef]
  12. Cai, J.; Huang, Z.a.; Lv, K.; Sun, J.; Deng, K. Ti powder-assisted synthesis of Ti3+ self-doped TiO2 nanosheets with enhanced visible-light photoactivity. RSC Adv. 2014, 4, 19588–19593. [Google Scholar] [CrossRef]
  13. Xing, H.; Wen, W.; Wu, J.-M. Enhanced UV photoactivity of Ti3+ self-doped anatase TiO2 single crystals hydrothermally synthesized using Ti-H2O2-HF reactants. J. Photochem. Photobiol. A Chem. 2019, 382, 111958. [Google Scholar] [CrossRef]
  14. 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] [PubMed] [Green Version]
  15. Yin, G.; Huang, X.; Chen, T.; Zhao, W.; Bi, Q.; Xu, J.; Han, Y.; Huang, F. Hydrogenated Blue Titania for Efficient Solar to Chemical Conversions: Preparation, Characterization, and Reaction Mechanism of CO2 Reduction. ACS Catal. 2018, 8, 1009–1017. [Google Scholar] [CrossRef]
  16. Hamisu, A.; Gaya, U.I.; Abdullah, A.H. Bi-template assisted sol-gel synthesis of photocatalytically-active mesoporous anatase TiO2 nanoparticles. Appl. Sci. Eng. Prog. 2021, 14, 313–327. [Google Scholar] [CrossRef]
  17. Wang, Y.; Sun, L.; Chudal, L.; Pandey, N.K.; Zhang, M.; Chen, W. Fabrication of Ti3+ Self-doped TiO2 via a Facile Carbothermal Reduction with Enhanced Photodegradation Activities. ChemistrySelect 2019, 4, 14103–14110. [Google Scholar] [CrossRef]
  18. Li, M.; Xing, Z.; Jiang, J.; Li, Z.; Kuang, J.; Yin, J.; Wan, N.; Zhu, Q.; Zhou, W. In-situ Ti3+/S doped high thermostable anatase TiO2 nanorods as efficient visible-light-driven photocatalysts. Mater. Chem. Phys. 2018, 219, 303–310. [Google Scholar] [CrossRef]
  19. Huang, Z.; Gao, Z.; Gao, S.; Wang, Q.; Wang, Z.; Huang, B.; Dai, Y. Facile synthesis of S-doped reduced TiO2-x with enhanced visible-light photocatalytic performance. Chin. J. Catal. 2017, 38, 821–830. [Google Scholar] [CrossRef]
  20. Sun, M.; Yao, Y.; Ding, W.; Anandan, S. N/Ti3+ co-doping biphasic TiO2/Bi2WO6 heterojunctions: Hydrothermal fabrication and sonophotocatalytic degradation of organic pollutants. J. Alloy. Compd. 2020, 820, 153172. [Google Scholar] [CrossRef]
  21. Cao, Y.; Xing, Z.; Shen, Y.; Li, Z.; Wu, X.; Yan, X.; Zou, J.; Yang, S.; Zhou, W. Mesoporous black Ti3+/N-TiO2 spheres for efficient visible-light-driven photocatalytic performance. Chem. Eng. J. 2017, 325, 199–207. [Google Scholar] [CrossRef]
  22. Qiao, M.; Wu, S.; Chen, Q.; Shen, J. Novel triethanolamine assisted sol–gel synthesis of N-doped TiO2 hollow spheres. Mater. Lett. 2010, 64, 1398–1400. [Google Scholar] [CrossRef]
  23. Si, L.; Huang, Z.; Lv, K.; Tang, D.; Yang, C. Facile preparation of Ti3+ self-doped TiO2 nanosheets with dominant {001} facets using zinc powder as reductant. J. Alloy. Compd. 2014, 601, 88–93. [Google Scholar] [CrossRef]
  24. Pan, Y.; Wen, M. Noble metals enhanced catalytic activity of anatase TiO2 for hydrogen evolution reaction. Int. J. Hydrog. Energy 2018, 43, 22055–22063. [Google Scholar] [CrossRef]
  25. Wang, L.; Luo, H.; Zhou, X.; Wei, A.; Zhou, K.; Chen, Z.; Zhang, D. Enhanced permittivity and energy density of P (VDF-HFP)-based capacitor using core-shell structured BaTiO3@TiO2 fillers. Ionics 2018, 24, 3975–3982. [Google Scholar] [CrossRef]
  26. Wang, J.; Wang, X.; Yan, J.; Tan, Q.; Liang, G.; Qu, S.; Zhong, Z. Enhanced Photoelectrochemical Properties of Ti3+ Self-Doped Branched TiO2 Nanorod Arrays with Visible Light Absorption. Materials 2018, 11, 1971. [Google Scholar] [CrossRef] [Green Version]
  27. Li, Y.; Ye, X.; Cao, S.; Yang, C.; Wang, Y.; Ye, J. Oxygen-Deficient Dumbbell-Shaped Anatase TiO2-x Mesocrystals with Nearly 100% Exposed {101} Facets: Synthesis, Growth Mechanism, and Photocatalytic Performance. Chem. A Eur. J. 2019, 25, 3032–3041. [Google Scholar] [CrossRef]
  28. Kuang, J.; Xing, Z.; Yin, J.; Li, Z.; Tan, S.; Li, M.; Jiang, J.; Zhu, Q.; Zhou, W. Ti3+ self-doped rutile/anatase/TiO2(B) mixed-crystal tri-phase heterojunctions as effective visible-light-driven photocatalysts. Arab. J. Chem. 2020, 13, 25682578. [Google Scholar] [CrossRef]
  29. Zhang, X.; Hu, W.; Zhang, K.; Wang, J.; Sun, B.; Li, H.; Qiao, P.; Wang, L.; Zhou, W. Ti3+ Self-Doped Black TiO2 Nanotubes with Mesoporous Nanosheet Architecture as Efficient Solar-Driven Hydrogen Evolution Photocatalysts. ACS Sustain. Chem. Eng. 2017, 5, 6894–6901. [Google Scholar] [CrossRef]
  30. Ren, R.; Wen, Z.; Cui, S.; Hou, Y.; Guo, X.; Chen, J. Controllable Synthesis and Tunable Photocatalytic Properties of Ti3+-doped TiO2. Sci. Rep. 2015, 5, 10714. [Google Scholar] [CrossRef] [Green Version]
  31. Kang, X.; Liu, S.; Dai, Z.; He, Y.; Song, X.; Tan, Z. Titanium dioxide: From engineering to applications. Catalysts 2019, 9, 191. [Google Scholar] [CrossRef] [Green Version]
  32. Nguyen, T.D.; Li, J.; Lizundia, E.; Niederberger, M.; Hamad, W.Y.; MacLachlan, M.J. Black titania with nanoscale helicity. Adv. Funct. Mater. 2019, 29, 1904639. [Google Scholar] [CrossRef]
  33. Zhen, D.; Gao, C.; Yang, D.; Zhu, X.; Grimes, C.A.; Liu, Y.; Cai, Q. Blue Ti3+ self-doped TiO2 nanosheets with rich {001} facets for photocatalytic performance. New J. Chem. 2019, 43, 5759–5765. [Google Scholar] [CrossRef]
  34. Motola, M.; Čaplovičová, M.; Krbal, M.; Sopha, H.; Thirunavukkarasu, G.K.; Gregor, M.; Plesch, G.; Macak, J.M. Ti3+ doped anodic single-wall TiO2 nanotubes as highly efficient photocatalyst. Electrochim. Acta 2020, 331, 135374. [Google Scholar] [CrossRef]
  35. Yu, X.; Fan, X.; An, L.; Li, Z.; Liu, J. Facile synthesis of Ti3+-TiO2 mesocrystals for efficient visible-light photocatalysis. J. Phys. Chem. Solids 2018, 119, 94–99. [Google Scholar] [CrossRef]
  36. Pellegrino, F.; Morra, E.; Mino, L.; Martra, G.; Chiesa, M.; Maurino, V. Surface and Bulk Distribution of Fluorides and Ti3+ Species in TiO2 Nanosheets: Implications on Charge Carrier Dynamics and Photocatalysis. J. Phys. Chem. C 2020, 124, 3141–3149. [Google Scholar] [CrossRef]
  37. Wang, P.; Jia, C.; Li, J.; Yang, P. Ti3+-doped TiO2(B)/anatase spheres prepared using thioglycolic acid towards super photocatalysis performance. J. Alloy. Compd. 2019, 780, 660–670. [Google Scholar] [CrossRef]
  38. Zhao, X.; Guo, K.; Zhang, K.; Duan, S.; Chen, M.; Zhao, N.; Xu, F.J. Orchestrated Yolk-Shell Nanohybrids Regulate Macrophage Polarization and Dendritic Cell Maturation for Oncotherapy with Augmented Antitumor Immunity. Adv. Mater. 2022, 34, 2108263. [Google Scholar] [CrossRef]
  39. Lei, J.; Zhang, M.; Wang, S.; Deng, L.; Li, D.; Mu, C. One-Pot Approach for the Synthesis of Water-Soluble Anatase TiO2 Nanoparticle Cluster with Efficient Visible Light Photocatalytic Activity. J. Phys. Chem. C 2018, 122, 26447–26453. [Google Scholar] [CrossRef]
  40. Mousavi, S.M.; Zarei, M.; Hashemi, S.A.; Ramakrishna, S.; Chiang, W.H.; Lai, C.W.; Gholami, A. Gold nanostars-diagnosis, bioimaging and biomedical applications. Drug Metab. Rev. 2020, 52, 1–20. [Google Scholar] [CrossRef]
  41. Tian, L.; Feng, H.; Dai, Z.; Zhang, R. Resorufin-based responsive probes for fluorescence and colorimetric analysis. J. Mater. Chem. B 2021, 9, 53–79. [Google Scholar] [CrossRef] [PubMed]
  42. Dlamini, N.N.; Rajasekhar Pullabhotla, V.S.R.; Revaprasadu, N. Synthesis of triethanolamine (TEA) capped CdSe nanoparticles. Mater. Lett. 2011, 65, 1283–1286. [Google Scholar] [CrossRef]
  43. Kang, K.-H.; Lee, D.-K. Synthesis of magnesium oxysulfate whiskers using triethanolamine as a morphology control agent. J. Ind. Eng. Chem. 2014, 20, 2580–2583. [Google Scholar] [CrossRef]
  44. Prakash, T.; Navaneethan, M.; Archana, J.; Ponnusamy, S.; Muthamizhchelvan, C.; Hayakawa, Y. Chemical synthesis of highly size-confined triethylamine-capped TiO2 nanoparticles and its dye-sensitized solar cell performance. Bull. Mater. Sci. 2018, 41, 40. [Google Scholar] [CrossRef] [Green Version]
  45. Liu, X.; Gao, S.M.; Xu, H.; Lou, Z.Z.; Wang, W.J.; Huang, B.B.; Dai, Y. Green synthetic approach for Ti3+ self-doped TiO2-x nanoparticles with efficient visible light photocatalytic activity. Nanoscale 2013, 5, 1870–1875. [Google Scholar] [CrossRef]
  46. Wang, X.T.; Li, Y.M.; Liu, X.; Gao, S.M.; Huang, B.B.; Dai, Y. Preparation of Ti3+ self-doped TiO2 nanoparticles and their visible light photocatalytic activity. Chin. J. Catal. 2015, 36, 389–399. [Google Scholar] [CrossRef]
  47. Haque, M.A.; Mahalakshmi, S. Triethanolamine-assisted synthesis of cadmium sulfide nanoclusters. Res. Chem. Intermed. 2014, 41, 5205–5215. [Google Scholar] [CrossRef]
  48. Du, M.; Yin, X.; Gong, H. Effects of triethanolamine on the morphology and phase of chemically deposited tin sulfide. Mater. Lett. 2015, 152, 40–44. [Google Scholar] [CrossRef]
  49. Poortavasoly, H.; Montazer, M.; Harifi, T. Aminolysis of polyethylene terephthalate surface along with in situ synthesis and stabilizing ZnO nanoparticles using triethanolamine optimized with response surface methodology. Mater. Sci. Eng. C Mater. Biol. Appl. 2016, 58, 495–503. [Google Scholar] [CrossRef]
  50. Bai, Q.; Lavenas, M.; Vauriot, L.; Le Trequesser, Q.; Hao, J.; Weill, F.; Delville, J.P.; Delville, M.H. Hydrothermal Transformation of Titanate Scrolled Nanosheets to Anatase over a Wide pH Range and Contribution of Triethanolamine and Oleic Acid to Control the Morphology. Inorg. Chem. 2019, 58, 2588–2598. [Google Scholar] [CrossRef] [Green Version]
  51. Rezaie, A.B.; Montazer, M. Polyester modification through synthesis of copper nanoparticles in presence of triethanolamine optimized with response surface methodology. Fibers Polym. 2017, 18, 434–444. [Google Scholar] [CrossRef]
  52. Liang, J.; Li, L.; Luo, M.; Wang, Y. Fabrication of Fe3O4 octahedra by a triethanolamine-assisted hydrothermal process. Cryst. Res. Technol. 2011, 46, 95–98. [Google Scholar] [CrossRef]
  53. Peng, S.; Dan, M.; Guo, F.; Wang, H.; Li, Y. Template synthesis of ZnIn2S4 for enhanced photocatalytic H2 evolution using triethanolamine as electron donor. Colloids Surf. A Physicochem. Eng. Asp. 2016, 504, 18–25. [Google Scholar] [CrossRef]
  54. Ramanathan, S.; Moorthy, S.; Ramasundaram, S.; Rajan, H.K.; Vishwanath, S.; Selvinsimpson, S.; Durairaj, A.; Kim, B.; Vasanthkumar, S. Grape seed extract assisted synthesis of dual-functional anatase TiO2 decorated reduced graphene oxide composite for supercapacitor electrode material and visible light photocatalytic degradation of bromophenol blue dye. ACS Omega 2021, 6, 14734–14747. [Google Scholar] [CrossRef] [PubMed]
  55. Xiong, H.; Wu, L.; Liu, Y.; Gao, T.; Li, K.; Long, Y.; Zhang, R.; Zhang, L.; Qiao, Z.A.; Huo, Q.; et al. Controllable Synthesis of Mesoporous TiO2 Polymorphs with Tunable Crystal Structure for Enhanced Photocatalytic H2 Production. Adv. Energy Mater. 2019, 9, 1901634. [Google Scholar] [CrossRef]
  56. Yin, G.; Wang, Y.; Yuan, Q. Ti3+-doped TiO2 hollow sphere with mixed phases of anatase and rutile prepared by dual-frequency atmospheric pressure plasma jet. J. Nanoparticle Res. 2018, 20, 1–12. [Google Scholar] [CrossRef]
  57. Abdullah, S.; Sahdan, M.; Nafarizal, N.; Saim, H.; Embong, Z.; Rohaida, C.C.; Adriyanto, F. Influence of substrate annealing on inducing Ti3+ and oxygen vacancy in TiO2 thin films deposited via RF magnetron sputtering. Appl. Surf. Sci. 2018, 462, 575–582. [Google Scholar] [CrossRef]
  58. Duan, Y.; Grah, P.A.; Cai, F.; Yuan, Z. Enhanced photoelectrochemical performance for hydrogen generation via introducing Ti3+ and oxygen vacancies into TiO2 nanorod arrays. J. Mater. Sci. Mater. Electron. 2018, 29, 20236–20246. [Google Scholar]
  59. Pan, J.; Dong, Z.; Wang, B.; Jiang, Z.; Zhao, C.; Wang, J.; Song, C.; Zheng, Y.; Li, C. The enhancement of photocatalytic hydrogen production via Ti3+ self-doping black TiO2/g-C3N4 hollow core-shell nano-heterojunction. Appl. Catal. B Environ. 2019, 242, 92–99. [Google Scholar] [CrossRef]
  60. Ming, H.; Zhang, H.; Ma, Z.; Huang, H.; Lian, S.; Wei, Y.; Liu, Y.; Kang, Z. Scanning transmission X-ray microscopy, X-ray photoelectron spectroscopy, and cyclic voltammetry study on the enhanced visible photocatalytic mechanism of carbon–TiO2 nanohybrids. Appl. Surf. Sci. 2012, 258, 3846–3853. [Google Scholar] [CrossRef]
  61. Yu, J.; Wang, B. Effect of calcination temperature on morphology and photoelectrochemical properties of anodized titanium dioxide nanotube arrays. Appl. Catal. B Environ. 2010, 94, 295–302. [Google Scholar] [CrossRef]
  62. Khan, M.; Ansari, S.; Pradhan, D.; Ansari, M.; Han, D.; Lee, J.; Cho, M. Band gap engineered TiO2 nanoparticles for visible light induced photoelectrochemical and photocatalytic studies. J. Mater. Chem. A 2014, 2, 637–644. [Google Scholar] [CrossRef]
Scheme 1. Schematic diagram of the synthesis route of Ti3+ self-doped TiO2 photocatalysts.
Scheme 1. Schematic diagram of the synthesis route of Ti3+ self-doped TiO2 photocatalysts.
Nanomaterials 12 02298 sch001
Figure 1. (a) XRD spectra of TiO2 photocatalysts obtained by solvothermal treatment of the sol at different triethanolamine concentrations and (b) enlarged XRD results.
Figure 1. (a) XRD spectra of TiO2 photocatalysts obtained by solvothermal treatment of the sol at different triethanolamine concentrations and (b) enlarged XRD results.
Nanomaterials 12 02298 g001
Figure 2. (a) Raman spectra of TiO2 photocatalysts obtained by hydrothermal treatment the sol at different triethanolamine concentrations and (b) enlarged Raman results.
Figure 2. (a) Raman spectra of TiO2 photocatalysts obtained by hydrothermal treatment the sol at different triethanolamine concentrations and (b) enlarged Raman results.
Nanomaterials 12 02298 g002
Figure 3. TEM and HRTEM pictures of TiO2 photocatalysts obtained with different content of TEA: (a) T10, (b) T20, (c,d) T30, and (e,f) T40.
Figure 3. TEM and HRTEM pictures of TiO2 photocatalysts obtained with different content of TEA: (a) T10, (b) T20, (c,d) T30, and (e,f) T40.
Nanomaterials 12 02298 g003
Figure 4. High-resolution XPS spectrum of different samples.
Figure 4. High-resolution XPS spectrum of different samples.
Nanomaterials 12 02298 g004
Figure 5. (a) UV-vis diffuse reflectance spectrum of TiO2 photocatalysts; (b) Plotting curves of Kubelka–Munk against photon energy over pure TiO2, T00, T10, T20, T30 and T40, respectively.
Figure 5. (a) UV-vis diffuse reflectance spectrum of TiO2 photocatalysts; (b) Plotting curves of Kubelka–Munk against photon energy over pure TiO2, T00, T10, T20, T30 and T40, respectively.
Nanomaterials 12 02298 g005
Figure 6. Transient photocurrent response (a) and EIS spectrum (b) of Nyquist plots acquired at open-circuit potential exposed to visible light illumination for different samples.
Figure 6. Transient photocurrent response (a) and EIS spectrum (b) of Nyquist plots acquired at open-circuit potential exposed to visible light illumination for different samples.
Nanomaterials 12 02298 g006
Figure 7. (a) Photocatalytic performance of the different samples; (b) recyclability of T30 in degradation of Rhodamine B.
Figure 7. (a) Photocatalytic performance of the different samples; (b) recyclability of T30 in degradation of Rhodamine B.
Nanomaterials 12 02298 g007
Figure 8. Schematic illustration of the visible-light photocatalytic mechanism of the Ti3+ self-doped TiO2 nanoparticles.
Figure 8. Schematic illustration of the visible-light photocatalytic mechanism of the Ti3+ self-doped TiO2 nanoparticles.
Nanomaterials 12 02298 g008
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hu, Z.; Gong, Q.; Wang, J.; Zheng, X.; Wang, A.; Gao, S. Construction of Spindle-Shaped Ti3+ Self-Doped TiO2 Photocatalysts Using Triethanolamine-Aqueous as the Medium and Its Photoelectrochemical Properties. Nanomaterials 2022, 12, 2298. https://doi.org/10.3390/nano12132298

AMA Style

Hu Z, Gong Q, Wang J, Zheng X, Wang A, Gao S. Construction of Spindle-Shaped Ti3+ Self-Doped TiO2 Photocatalysts Using Triethanolamine-Aqueous as the Medium and Its Photoelectrochemical Properties. Nanomaterials. 2022; 12(13):2298. https://doi.org/10.3390/nano12132298

Chicago/Turabian Style

Hu, Zunfu, Qi Gong, Jiajia Wang, Xiuwen Zheng, Aihua Wang, and Shanmin Gao. 2022. "Construction of Spindle-Shaped Ti3+ Self-Doped TiO2 Photocatalysts Using Triethanolamine-Aqueous as the Medium and Its Photoelectrochemical Properties" Nanomaterials 12, no. 13: 2298. https://doi.org/10.3390/nano12132298

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