Acessibilidade / Reportar erro

TiO2 nanotube arrays with visible light catalytic

Abstract

The TiO2 nanotube arrays were prepared by anodization, and the crystal structure was changed by calcination at different temperatures. The photocatalytic performance of the samples was measured by the degradation of rhodamine B under visible light. The TiO2 nanotubes calcined at 600 °C showed higher photocatalytic activity than other samples. The prepared catalyst is characterized by a variety of techniques, including X-ray diffraction, scanning electron microscopy, ultraviolet-visible diffuse reflectance spectroscopy, Raman, photoluminescence spectroscopy and electrochemical testing. The reasons for improving the catalytic activity were studied from the aspects of crystal structure, surface morphology, and photoelectric properties, and the catalytic mechanism was studied. The results show that the TiO2 nanotubes calcined at 600°C contain two phases of anatase and rutile. Compared with pure phase TiO2, the charge transfer resistance is reduced and the electron-hole reorganization is well suppressed. In addition, it affects the band structure and improves the absorption of visible light. At the same time, studies have found that the main active substances in the catalytic process are h+ and •OH.

Key words
TiO2 nanotube arrays; rhodamine B; anatase; rutile; active substances

INTRODUCTION

The study found that TiO2 semiconductor is an n-type semiconductor, which is widely used in gas sensors (Sun et al. 2019SUN H, GUI Y, WEI H, LONG YK, WANG Q & TANG C. 2019. DFT study of SF6 decomposed products on Pd–TiO2: gas sensing mechanism study. Adsorption 25: 1643-1653. doi: 10.1007/s10450-019-00150-1., Duy et al. 2008DUY NV, HIEU NV, HUY PT, CHIEN ND & THAMILSELVAN M. 2008. Mixed SnO2/TiO2 included with carbon nanotubes for gas-sensing application. Physica E 41: 258-263. doi: 10.1016/j.physe.2008.07.007., Morris et al. 2001MORRIS D, EGDELL RG & MATER J. 2001. Application of V-doped TiO2 as a sensor for detection of SO2. Chem 11: 3207-3210. doi: 10.1039/B104801P.), solar cells (Liu et al. 2018LIU J, LI Y & ARUMUGAM S. 2018. Investigation of low temperature processed titanium dioxide (TiO2) films for printed dye sensitized solar cells (DSSCs)for large area flexible applications. Mater Today Commun 5: 13846-13854. doi: 10.1016/j.matpr.2018.02.026., Fitra et al. 2013FITRA M, DAUT I, IRWANTO M, GOMESH N & IRWAN YM. 2013. Effect of TiO2 thickness dye solar cell on charge generation. Energy Procedia 36: 278-286. doi: 10.1016/j.egypro.2013.07.032., Adawiya et al. 2017ADAWIYA J, HAIDER H & RIYAD. 2017. Exploring potential environmental applications of TiO2 nanoparticles. Energy Procedia 119: 332-345. doi: 10.1016/j.egypro.2017.07.117.) and photocatalysis (Tieng et al. 2011TIENG S, KANAEV A & CHHOR K. 2011. New homogeneously doped Fe(III)-TiO2 photocatalyst for gaseous pollutant degradation. Appl Catal A: Gen 399: 191-197. doi: 10.1016/j.apcata.2011.03.056., Pham et al. 2016PHAM TD, LEE BK, LEE CH & RHEUMATOL BJ. 2016. The advanced removal of benzene from aerosols by photocatalytic oxidation and adsorption of Cu-TiO2/PU under visible light irradiation. Appl Catal B: Environ 36: 1341-1342. doi: 10.1016/j.apcatb.2015.09.023., Gao et al. 2014GAO SW, LAN Z, WU WX, QUE LF & WU JH. 2014. Fabrication and photovoltaic performance of high efficiency front-illuminated dye-sensitized solar cell based on ordered TiO2 nanotube arrays. Acta Phys Chim Sin (3): 446-452. doi: 10.3866/PKU.WHXB201401022.) because of its non-toxicity, low cost, high activity and good stability. However, TiO2 also suffffers from a limited photocatalytic application due to its weak adsorptive property, easy aggregation and recycling problems which would cause the secondary pollution after degradation process. The researchers used different methods to change the morphology, particle size and crystal phase of TiO2 to improve its photocatalytic activity (Chiarello et al. 2005CHIARELLO GL, ROSSETTI I & FORNI L. 2005. Flame-spray pyrolysis preparation of perovskites for methane catalytic combustion. J Catal 236: 251-261. doi: 10.1016/j.jcat.2005.10.003.). At present, the preparation methods of TiO2 mainly include hydrothermal method (Sugimoto et al. 2002SUGIMOTO T, ZHOU X & MURAMATSU A. 2002. Synthesis of Uniform anatase TiO2 nanoparticles by gel–sol method: 1. solution chemistry of Ti(OH)n(4-n)+ complexes. J. Colloid. Inter Sci 252: 339-346. doi: 10.1006/jcis.2002.8454.), anodic oxidation (Liang et al. 2011LIANG YC, WANG CC, KEI CC, HSUEH YC & CHO WH. 2011. Photocatalysis of Ag-loaded TiO2 nanotube arrays formed by atomic layer deposition. J Physica Chem C 115: 9498-9502. doi: 10.1021/cg301230w.), flame spray pyrolysis (Macak et al. 2007MACAK JM, ZLAMAL M, KRYSA J & SCHMUKI P. 2007. Self-organized TiO2 nanotube layers as highly efficient photocatalysts. small 3: 300-304. doi: 10.1002/smll.200600426., Mohamed & Rohani 2011MOHAMED AER & ROHANI S. 2011. Modified TiO2 nanotube arrays (TNATs): progressive strategies towards visible light responsive photoanode, a review. Energ Environ Sci 4: 1065-1086. doi: 10.1039/C0EE00488J.) and sol-gel method (Hwang et al. 2013HWANG KJ, LEE JW, YOO SJ, JEONG SY & JEONG DH. 2013. Fine size-regulation of nanocrystalline anatase-TiO2 via sol-gel synthesis and subsequent phase transformation by calcination. New J Chem 37: 1378-1384. doi: 10.1039/C3NJ41170B.). Among these the TiO2 nanotube arrays (TNTs) prepared by the anodic oxidation method is likely to recover, not easy to agglomerate, and has no secondary pollution. And compared to nanoparticles, vertically oriented structures of TNTs provide especially high energy conversion efficiencies by providing direct electron diffusion pathways at interfaces.

Although these advances, the photocatalytic efficiency of TNTs is still greatly limited due to its wide band gap and high photo-generated electron-hole recombination rate (Zeng et al. 2015ZENG M, LI Y, MAO M, BAI L & REN L. 2015. Synergetic effect between photocatalysis on TiO2 and thermocatalysis on CeO2 for gas-phase oxidation of benzene on TiO2/CeO2 nanocomposites. ACS Catal 5: 3278-3286. doi: 10.1021/acscatal.5b00292.). In order to reduce the band gap and improve the visible light catalytic activity, the predecessors improved the catalytic performance of TiO2 by changing the crystal form (Qian et al. 2018QIAN X, HAN H, CHEN Y & YUAN Y. 2018. Sol-gel solvothermal route to synthesize anatase/brookite /rutile TiO2, nanocomposites with highly photocatalytic activity. J Sol-Gel Sci Technol 85: 394-401. doi: 10.1007/s10971-017-4544-3., Ghayeb & Ghonchegi 2015GHAYEB MMY & GHONCHEGI Z. 2015. Fabrication and characterization of copper doped TiO2, nanotube arrays by in situ electrochemical method as efficient visible-light photocatalyst. Ceram Int 41: 8735-8741. doi: 10.1016/j.ceramint.2015.03.094.), metal ion doping (Tian et al. 2009TIAN H, MA J, LI K & LI JJ. 2009. Hydrothermal synthesis of S-doped TiO2, nanoparticles and their photocatalytic ability for degradation of methyl orange. Ceram Int 35: 1289-1292. doi: 10.1016/j.ceramint.2008.05.003., Choi et al. 2016CHOI Y, KIM H, MOON G, JO SG & CHOI WY. 2016. Boosting up the low catalytic activity of silver for H2 production on Ag/TiO2 photocatalyst: thiocyanate as a selective modifier. ACS Publications 6: 821-828. doi: 10.1021/acscatal.5b02376.), semiconductor composite (Li et al. 2013LI R, LI Q, ZONG L, WANG XD, YANG JJ. 2013. BaTiO3/TiO2 heterostructure nanotube arrays for improved photoelectrochemical and photocatalytic activity. Electrochim Acta 91: 30-35. doi: 10.1016/j.electacta.2012.12.073., Qamar et al. 2015QAMAR M, DRMOSH Q, AHMED MI, QAMARUDDIN M & YAMANI ZH. 2015. Enhanced photoelectrochemical and photocatalytic activity of WO3-surface modified TiO2 thin film. Nanoscale Res Lett 10: 54. doi: 10.1186/s11671-015-0745-2.) and noble metal deposition (Wang et al. 2015WANG Y, FANG HB, ZHENG YZ, YE R & TAO X. 2015. Controllable assembly of well-defined monodisperse Au nanoparticles on hierarchical ZnO microspheres for enhanced visible-light-driven photocatalytic and antibacterial activity. Nanoscale 7: 19118-19128. doi: 10.1039/C5NR06359K.). Among the many methods, it is much simpler to change the crystal phase structure. TiO2 has three crystal forms of anatase, rutile and brookite. The photocatalytic effects of different crystal types are particularly different, and the composite materials of different crystal phases have better photocatalytic activity (Barnard & Curtiss 2005BARNARD AS & CURTISS LA. 2005. Prediction of TiO2 nanoparticle phase and shape transitions controlled by surface chemistry. Nano Lett 5: 1261-1266. doi: doi.org/10.1021/nl050355m.
https://doi.org/10.1021/nl050355m....
, Dicesare & Lakowicz 2001DICESARE N & LAKOWICZ JR. 2001. Spectral properties of fluoro-phores combining the boronic acid group with electron do-nor or withdrawing groups. Implication in the development of fluorescence probed for saccharides. J Phys Chem A 105: 6834-6840., Wiley et al. 2001WILEY JP, HUGHES KA, KAISER RJ, KESICKI EA & LI GS. 2001. Phenylboronic acid salicylhydroxamic acid bioconjugates 2 polyvalent im-mobilization of protein ligands for affinity chromatography. Bioconjugate Chem 12: 240-250. doi: 10.1021/bc0000942.). Thus, predecessors have done a lot of research in this area. Aref Mamakhel et al. (2013)MAMAKHEL A, TYRSTED C, JESEN ED, HALD P & IVERSEN BB. 2013. Direct formation of crystalline phase pure rutile TiO2, nanostructures by a facile hydrothermal method. Cryst Growth Des 13: 4730-4734. doi: 10.1021/cg400858p. used hydrothermal method to synthesize a single rutile nanometer TiO2 with a particle size of less than 10 nm, which provided a new method for obtaining rutile TiO2 at low temperature. Zhang et al. (2009)ZHANG DQ, LI GS & YANG XF. 2009. A micrometer-size TiO2 single-crystal photocatalyst with remarkable 80% level of reactive facets. Chem Commun 29: 4381-4383. doi: 10.1021/cg400858p. has synthesized an anatase TiO2 single crystal photocatalyst with high activity by microwave-assisted hydrothermal method. Li et al. (2017)LI H, WU X, YIN S, KATSUMATA K & WANG YH. 2017. Effect of rutile TiO2 on the photocatalytic performance of g-C3N4/brookite-TiO2-xNyphotocatalyst for NO decomposition. Appl Surf Sci 392: 531-539. doi: 10.1016/j.apsusc.2016.09.075. prepared g-C3N4/rutile TiO2 and g-C3N4/brookite TiO2 composites by a simple solvothermal method, and studied the effect of rutile phase TiO2 and brookite TiO2 on the photocatalytic activity of g-C3N4. Mi et al. (2012)MI JL, CLAUSEN C & BREMHOLM M. 2012. Rapid hydrothermal preparation of rutile TiO2, nanoparticles by simultaneous transformation of primary brookite and anatase: An in situ synchrotron PXRD study. Cryst Growth Des 1: 6092-6097. doi: 10.1016/j.cej.2016.04.092. studied the formation mechanism and crystal growth of TiO2 in high temperature and high pressure fluid. When HCl is used as an additive, pure rutile TiO2 nanoparticles can be formed, while using H2SO4 can obtain pure anatase TiO2. According to previous studies, there are few studies on the preparation of TiO2 with different crystal phases on the titanium mesh by anodization.

In this research, simple anodic oxidation method was used to prepare amorphous TNTs. The TNTs crystal phase was changed by calcination at different temperatures, and the effect of different crystalline TNTs on the photocatalytic activity was investigated. In experiment, titanium mesh is used as the matrix, but titanium does not participate in the catalysis, so it will not interfere with the results of the catalysis. Since the TNTs uniformly cover the substrate and the titanium has not yet entered the TiO2 lattice, it will not form a heterostructure with TNTs.

MATERIALS AND METHODS

Materials

The raw materials used include hydrofluoric acid (HF, 40.0%, CAS: 7664-39-3), nitric acid (HNO3, 65.0-68.0%, CAS: 7697-37-2), isopropanol (C2H6OH, 99.7%, CAS: 67-63-0) and titanium mesh (99%). All raw materials were purchased from Sinopharm Chemical Reagent Network without further purification.

Synthesis

Before anodization, the titanium mesh (2×4 cm) was sequentially sonicated in acetone and pure ethanol for 15 min, followed by etching in a mixture of HF/HNO3/H2O (1:4:5 in volume) for 30 s to remove the oil and oxide layer on the surface, rinsing with deionized water, and drying under flowing N2. The titanium mesh after the above treatment is used as the anode, and the graphite plate is the cathode (2×4 cm). The electrolyte is an aqueous solution containing 1 % HF. Anodization was performed at 15 V for 1 hour, then washed with deionized water and dried. Under a protective atmosphere of argon gas, the obtained sample was calcined for 2 h in a tube furnace, and the calcining temperatures were 300 °C, 400 °C, 500 °C, 600 °C, and 700 °C (heating rates of 5°C/min), respectively.

Catalyst characterizations

The crystal phase structure of the sample was determined using an X’Pert PRO MRD type X-ray diffractometer. The SEM was performed using a JSM-7200 field emission scanning electron microscope. The PL spectrum was measured with a F-7000 type fluorescence spectrophotometer, using a Xe lamp (excitation wavelength is 365 nm) as light source. The DRS spectrum was measured using a U-3900 UV-Vis spectrophotometer and BaSO4 was used as the reference sample.

Photocurrent response (PR) tests were performed using a CS130 electrochemical workstation. A three-electrode system was used, and the corresponding three electrodes are a working electrode, a reference electrode (Ag/AgCl saturated KCl), and a counter electrode (Pt rod). The titanium mesh was cut into 1 cm×1.5 cm, and the anodized area was 1 cm2. An alternating current voltage of 10 mV was applied, and the initial potential was 0 V. The photocurrent response diagram was measured using a 350 W Xe lamp as the light source and an aqueous solution of Na2SO4 (0.3 M) as the electrolyte in the frequency range of 105 Hz to 10-2 Hz.

Photocatalytic reaction

The photocatalytic activities of TNTs were tested by RhB (10 mg/L) degradation under visible-light irradiation. The light source is a 250 W high-pressure mercury lamp (7.0 mW/cm2), and the wavelength below 420 nm is filtered by a filtering device. The photocatalytic performance of TNTs calcined at different temperatures was tested. Place a piece of titanium mesh in 75 ml of the prepared RhB solution for 20 min in the dark, and then transfer the solution to the light source for photodegradation reaction. After centrifuging all the catalyzed solutions twice, the supernatant was taken and the absorbance was measured at 554nm with a 752 spectrophotometer. The sample was catalyzed several times to test its stability.

Mechanism exploration

In order to confirm the mechanism of this catalytic reaction, benzophenone (BQ), triethanolamine (TEOA) and tert-butanol (TBA) were used as scavengers for superoxide radicals (O2 -), holes (h+) and hydroxyl radicals (•OH). These three kinds of scavengers (0.01 mol/L) were added to three same RhB solutions respectively before dark adsorption, and the rest of the process was consistent with the degradation experiment (Hong et al. 2016HONG Y, LI C, ZHANG G, MENG Y & YIN. 2016. Efficient and stable Nb2O5 modified g-C3N4 photocatalyst for removal of antibiotic pollutant. Chem Eng J 299: 74-84. doi: 10.1016/j.jece.2015.03.021.).

RESULTS AND DISCUSSION

XRD analysis

All samples were checked using XRD to determine the crystal phase structure of the TNTs material (Figure 1). The XRD pattern of the titanium mesh is the peak of pure titanium without any other impurities. The titanium mesh after anodization does not show the peak of TiO2, but it shows the peak of TiO2 after calcination under the protection of argon atmosphere, which can indicate that TiO2 can be prepared by anodization, and the TiO2 is amorphous state. The samples after calcination at 300 °C, 400 °C and 500 °C all shows anatase-type TiO2 peaks. The two main diffraction peaks are located at 2θ of 25.28 ° and 48.05 °, which correspond to the characteristic peaks of the anatase phase (101) and (200) crystal planes, respectively. And the peak strength increases slightly with increasing temperature, indicating that anatase TiO2 has increased. The peak strength of the anatase-type TiO2 of the sample calcined at 600 °C weakened, and a peak of rutile-type TiO2 appeared. The diffraction peak of the sample calcined to 700 °C was completely transformed into rutile TiO2. Diffraction peaks at 27.4°, 36.1°, 41.2°, 44.1°, 54.3°, 56.6°, 64.0°, 69.0°and 69.8° were observed in all cases, corresponding to the crystal faces of rutile TiO2, which are (110), (101), (111), (210), (211), (220), (310), (301) and (112), respectively. It can be concluded from the above that TiO2 was prepared after anodization, and the crystalline phase can be changed after roasting at different temperatures. With the increase of temperature, the crystalline phase of TiO2 gradually changes from anatase to rutile.

Figure 1
XRD patterns of TNTs calcined at different temperatures.

Characterization of morphology and structure

SEM was performed to investigate the morphologies of the prepared samples (Figure 2). The TiO2 prepared by anodizing is a tubular structure with a flat surface and uniform diameter, an inner diameter of about 100 nm, a wall thickness of about 8 nm and length is about 350 nm (Figure 2a). After calcining at 300 °C, 400 °C and 500 °C, the sample did not change the nanotube structure and the length has not changed significantly, but the wall thickness increased. The wall thickness of the nanotubes calcined at 500 °C could reach 15 nm (Figure 2b-d). After calcining at 600 °C, the wall thickness of nanotubes increased greatly and cracked severely, and only a small amount of nanotubes still maintained a tubular structure. The ruptured tube wall is spherically stacked (Figure 2e). However, almost all nanotubes were ruptured after calcination at 700 °C and the length also becomes shorter, about 310 nm (Figure 2f).

Figure 2
SEM images of (a) uncalcined TNTs and (b-f) 300 °C, 400 °C, 500 °C, 600 °C and 700 °C calcined TNTs.

Raman spectroscopy

Raman scattering is more sensitive to the detection of the short-range structure of crystals, which makes it easier to research the crystal phase structure and quality of nano-sized crystals. The anatase TiO2 nanocrystals have six Raman active modes (A1g + 2B1g + 3Eg), which are located at 145cm-1 [Eg(1)], 197cm-1 [Eg(2)], 399 cm-1 [B1g(1)], 513 cm-1 [A1g], 519 cm-1 [B1g(2)], and 639 cm-1 [Eg(3)]. Four characteristic peaks were observed in the measured Raman spectrum (Figure 3), at 145 cm-1, 399 cm-1, 513 cm-1, and 639 cm-1. And rutile TiO2 nanocrystals have four Raman active mode (A1g, B1g, B2g, Eg), two characteristic peaks of 443 cm-1 (Eg) and 610 cm-1(A1g) were observed. The broad peak at 232 cm-1 is a composite peak and a characteristic Raman peak of rutile TiO2. The samples calcined at 300°C, 400°C and 500°C are anatase TiO2, and there is no rutile Raman peak. As the temperature rises to 500°C the characteristic peak of anatase increases. The sample calcined at 600°C began to show the characteristic peak of rutile TiO2, but the characteristic peak of corresponding anatase TiO2 still exists, so it is a mixed crystal form of anatase and rutile. After calcination at 700°C, only a small amount of anatase TiO2 is present in the sample.

Figure 3
Raman spectra of TNTs.

Photoluminescence spectroscopy

The recombination ratio of electron-hole pairs is an important factor for measuring the performance of photocatalysts. The photoluminescence spectra (Figure 4) of the samples were tested to compare the recombination ratios of electron-hole pairs of TNTs baked at different temperatures. Generally, the lower the fluorescence intensity, the lower the recombination efficiency of the photoinduced electron-hole pair, and therefore the higher the photocatalytic activity. Figure 4 shows the PL spectral excitation wavelength at 365 nm. The PL spectra of TNTs exhibits a strong emission band centered at about 440 nm, which was attributed to the recombination process of self-trapped excitations (Luo et al. 2015LUO J, ZHOU XS, MA L & XU Y. 2015. Enhancing visible-light photocatalytic activity of g-C3N4 by doping phosphorus and coupling with CeO2 for the degradation of methyl orange under visible light irradiation. RSC Adv 5: 568728-68735. doi: 10.1039/C5RA10848A., Tonda et al. 2015TONDA S, KUMAR S & SHANKER V. 2015. In situ growth strategy for highly efficient Ag2CO3/g-C3N4 hetero/nanojunctions with enhanced photocatalytic activity under sunlight irradiation. J Environ Chem Eng 3: 852-861. doi: 10.1016/j.jece.2015.03.021.). Obviously, the PL strength of uncalcined TNTs is significantly higher than that of TNTs treated at high temperature, indicating amorphous TNTs has the highest electron-hole pair recombination rate, which reduces the photodegradation efficiency. The fluorescence intensity of the calcined TiO2 material is greatly suppressed. The fluorescence quenching degree of TiO2 calcined at 300 °C, 400 °C, 500 °C and 700 °C is almost the same. However, the strength was especially reduced after firing at 600 °C. The results show that the recombination efficiency of electron-hole pairs of TiO2 composites containing both anatase and rutile phases is much lower than that of single crystal phase of TNTs.

Figure 4
PL spectra of TNTs calcined at different temperatures.

Electrochemical test

In order to further verify the conclusions obtained from the PL spectrum, a transient photocurrent response was further performed on the sample (Figure 5). It can be seen that uncalcined TNTs can be excited in visible light, but the photocurrent response is weak, indicating that the inferior photocatalytic activity. After calcination, TNTs can be strongly excited in visible light, and the photocurrent response is particularly enhanced. As the temperature increases, the photocurrent response increases first and then decreases. The TNTs calcined at 600 °C have the strongest response to visible light, indicating that TNTs with a certain proportion of anatase and rutile mixed phase have the highest photocatalytic activity. This is consistent with the results obtained with the PL spectrum.

Figure 5
transient photocurrent responses of TNTs calcined at different temperatures.

Optical performance studies

The UV-vis DRS of TNTs calcined at different temperatures is shown in Figure 6a. Pure TiO2 nanotube arrays showed strong light absorption in the ultraviolet region below 385 nm and showed weak light absorption in the wavelength range of 400-600 nm, which is due to light scattering caused by cracks or pores in the nanotube arrays (Zhang et al. 2015ZHANG M, LU D, ZHANG Z, GUO Y & YANG J. 2015. Enhanced photocurrent and photocatalytic degradation of methyl orange by VN codoped TiO2 nanotube arrays cooperated with H2O2. J Electrochem Soc 162: 557-563. 10.1149/2.0901508jes.). In contrast, the absorption performance of the samples after calcination in the visible light range of 550 nm-800 nm was significantly improved. The samples calcined at 300 °C, 400 °C, and 500 °C were red-shifted compared to the non-calcined samples, which also showed that the absorption of visible light was improved. After sintering at 600 °C, the sample has a strong absorption of visible light, indicating the TNTs of two mixed crystal phases effectively improves the light absorption capacity. Due to the different Fermi energy levels of the two crystalline forms of TiO2 (Bickley et al. 1991BICKLEY IB, GONZALEZ-CARRENO T & LEES JS. 1991. A structural investigation of titanium dioxide photocatalysts. J Solid State Chem 92: 178-190. doi: 10.1016/0022-4596(91)90255-G.), a potential barrier can be created between the two-phase interfaces, which can promote the transfer, separation and migration of electrons and holes to the surface of the catalyst, all of which greatly enhance the visible light absorption performance. However, the samples after calcining at 700 °C did not improve the visible light absorption performance, but greatly reduced the ultraviolet light absorption performance, indicating that the single rutile TiO2 has a poor response to ultraviolet-visible light. In addition, it can be seen from the figure that the absorption edges of the uncalcined TNTs, 500 °C TNTs and 600 °C TNTs are 412 nm, 426 nm and 445 nm, respectively. Substitute the wavelength corresponding to the absorption edge into the formula Eg(eV)=1240/λ (nm)(Feng et al. 2018FENG CY, DENG YC, TANG L, ZENG GM & WANG JJ. 2018. Core-shell Ag2CrO4/N-GQDs@g-C3N4 composites with anti-photocorrosion performance for enhanced full-spectrum-light photocatalytic activities. Appl Catal B: Environ 239: 525-536. doi: 10.1016/j.apcatb.2018.08.049., Wu et al. 2019WU J, FENG YJ, LOGAN BE, DAI CC & HAN XY. 2019. Preparation of Al–O-linked porous-g-C3N4/TiO2-nanotube Z-Scheme composites for efficient photocatalytic CO2 conversion and 2,4-dichlorophenol decomposition and mechanism. ACS Sustainable Chem Eng 7: 15289-15296. doi: 10.1021/acssuschemeng.9b02489.). The calculated band gaps of the samples are 3.06 eV, 2.91 eV and 2.78 eV, respectively. This also explains that the TNTs calcined at 600 °C have better light absorption properties than other samples.

Figure 6
UV-vis spectra of TNTs calcined at different temperatures.

Catalytic performance evaluation

The photocatalytic activity of TNTs calcined at different temperatures was measured by the degradation of RhB under visible light. Figure 7a-g are absorption spectra of RhB at different times during catalytic degradation. In the absorption spectrum of the degraded RhB of the blank sample, it can be seen that RhB is basically not degraded, and its maximum absorption wavelength is 554 nm (Figure 7a). Figure 7b-g shows that all TNTs samples do not have much influence on the absorption wavelength of RhB. Compared with uncalcined TNTs, the degradation rate of RhB by TNTs after heat treatment is greatly increased. With the increase of heat treatment temperature, the degradation rate first increases and then decreases and the sample calcined at 600°C has the best catalytic effect. It can be seen from the Figure 7h that the direct degradation of RhB can be neglected (<5%). Uncalcined TNTs also has degradability, and the final degradation rate can only reach 52.5%. However, the catalytic effect of TNTs calcined at 300 °C and 400 °C was not significantly enhanced, probably because there was only a small amount of anatase TiO2. After 500 °C calcination, the catalytic effect was greatly enhanced, and finally reached 92.7%. And after calcination at 600 °C, the degradation rate is further increased, and can finally reach 96%. The results show that the TNTs mixed with anatase and rutile has the best catalytic effect and the highest photocatalytic activity. It shows that TNTs with a certain proportion of anatase and rutile have the best catalytic effect and the highest photocatalytic activity. The photocatalytic degradation rate of TNTs calcined at 700 °C decreased obviously, indicating that the single rutile TNTs has low catalytic activity and is lower than that of single anatase. The photocatalytic degradation rate of TNTs calcined at 700 °C decreased obviously, indicating that the single rutile TNTs has low catalytic activity and is lower than that of single anatase. From the photocatalytic degradation rate chart in Figure 7i, it can be seen that the degradation rates from high to low are 600 °C > 500 °C > 300 °C > 400 °C> uncalcined > blank. The degradation rate of TNTs calcined at 600 °C can reach 0.0186 min-1, which is four times that of uncalcined TNTs (0.00472 min-1). The degradation rate can also indicate that TNTs with a certain proportion of anatase and rutile have the highest activity. Compared with the TiO2 materials prepared in other articles, although the degradation rate of the TiO2 nanotubes prepared in this article is not the best, it has surpassed some modified TiO2 nanomaterials and far exceeds that of pure TiO2 (P25) as shown in Table I. The standard error of Figure7(i) is calculated as shown in Table II, and all errors are within the acceptable range.

Figure 7
UV-Vis absorption spectra of photocatalytic RhB degradation over (a) blank, (b) uncalcined TNTs, (c) 300 °C TNTs, (d) 400 °C TNTs, (e) 500 °C TNTs, (f) 600 °C TNTs, (g) 700 °C TNTs, (h) photodegradation efficiency toward RhB measured at 554 nm and (i) the photocatalytic rate curves and corresponding fitted kinetics curves.
Table I
Some TiO2 materials and degradation rate table.
Table II
Standard error table of all samples.

Cycle performance and catalytic mechanism

For photocatalytic materials, high catalytic efficiency and high stability are required, which can be reused without losing activity. Therefore, TNTs calcined at 600 °C were recycled and tested for stability. The sample was catalyzed six times and the final degradation rate was recorded each time (Figure 8a and Table III). It can be seen that after six cycles of use, the final catalytic effect of the sample did not obviously decrease. This shows that TNTs are stable and can be reused without losing activity. As shown in Figure 8b and Table IV, the photocatalytic activity did not decrease significantly after the addition of p-benzoquinone (BQ) to the TNTs system. The addition of tert-butanol (TBA) and triethanolamine (TEOA) significantly reduced the degradation rate, indicating that h+ and •OH are the main active substances. The standard error calculation of each catalytic result shows that it is within the allowable range of error.

Figure 8
(a) Photocatalytic cycle performance of TNTs calcined at 600 °C and (b). reactant capture diagram.
Table III
Standard error table of cycle performance test.
Table IV
Standard error table of mechanism test.

In order to further reveal the photocatalytic mechanism of TNTs, the following formulas are used to calculate the valence and conduction band values (Tonda et al. 2015TONDA S, KUMAR S & SHANKER V. 2015. In situ growth strategy for highly efficient Ag2CO3/g-C3N4 hetero/nanojunctions with enhanced photocatalytic activity under sunlight irradiation. J Environ Chem Eng 3: 852-861. doi: 10.1016/j.jece.2015.03.021.).

E C B = X X e E g 2
E V B = X C B E g

X is absolute electronegativity, and the X value of TiO2 is 5.8 Ev (Chen et al. 2014CHEN S, HU Y, MENG S & FU X. 2014. Study on the separation mechanisms of photogenerated electrons and holes for composite photocatalysts g-C3N4-WO3. Appl Catal B-Environ 150-151: 564-573. doi: 10.1016/j.apcatb.2013.12.053.). Ee is a constant relative to the standard H electrode (Ee value is 4.5 eV), Eg is the band gap (Chen et al. 2014CHEN S, HU Y, MENG S & FU X. 2014. Study on the separation mechanisms of photogenerated electrons and holes for composite photocatalysts g-C3N4-WO3. Appl Catal B-Environ 150-151: 564-573. doi: 10.1016/j.apcatb.2013.12.053., Huang et al. 2016HUANG K, LIU C, YAN X, HUANG CY & CHEN JB. 2016. Hydrothermal synthesis of g-C3N4/CdWO4 nanocomposite and enhanced photocatalytic activity for tetracycline degradation under visible light. Crystengcomm 18: 6453-6463. doi: 10.1016/j.seppur.2017.07.026.). EVB values of TNTs were calculated to be 2.69 eV, and ECB values were -0.09 eV. The potential of the valence band edge (2.69 eV) is more positive than the oxidation potentials of OH−/•OH (1.99 eV) and H2O/•OH (2.34 eV). Thus, the photogenerated holes of TNTs can react with OH− or H2O to generate •OH. The potential of O2/•O2 - is -0.33 eV, which is lower than the ECB value of TiO2 nanotubes, so O2 cannot be reduced to •O2 -, and •O2 - is not an active species in the system. Therefore, only h+ and •O2 - can play a role in the catalytic process, and a redox reaction occurs with the pollutants, thereby achieving the effect of degrading the pollutants. TNTs containing both anatase and rutile can effectively improve the separation of photogenerated electron-hole pairs, greatly reduce the possibility of charge recombination, and thus improve the photocatalytic activity of the material. From the above PL, photocurrent and catalytic properties, it can be seen that the recombination of electron-hole pairs is suppressed, and the direct oxidation ability of holes is greatly improved, so that h+ and •OH become active species in TNTs. Figure 9 shows the mechanism of catalytic degradation of RhB. The specific process of catalytic degradation of RhB by TNTs as follows:

Figure 9
Catalytic mechanism diagram of TNTs.

CONCLUSIONS

TiO2 nanotubes are prepared by anodization, and samples with different crystal phases can be obtained by calcination. The TiO2 nanotubes calcined at 600 °C contained two phases of anatase and rutile, and showed the highest RhB degradation rate under visible light. The final catalytic efficiency can reach 96%. This is because the combination of the anatase and rutile phases reduces the charge transfer resistance and effectively inhibits the recombination of electron-hole pairs, thereby improving the photocatalytic activity. At the same time, cyclic tests show that the prepared catalyst has good stability and can be reused many times, so it has broad application prospects in solving environmental pollution.

ACKNOWLEDGMENTS

This work was supported by the Natural Science Foundation of Shandong Province of China (No. ZR2019MEM020, No. ZR2017LEM004) , the Fundamental Research Funds for the Central Universities of China (No. 18CX02091A) and the Open Fund (No. OGE201702-07) of Key Laboratory of Oil & Gas Equipment, Ministry of Education (Southwest Petroleum University).

REFERENCES

  • ADAWIYA J, HAIDER H & RIYAD. 2017. Exploring potential environmental applications of TiO2 nanoparticles. Energy Procedia 119: 332-345. doi: 10.1016/j.egypro.2017.07.117.
  • BARNARD AS & CURTISS LA. 2005. Prediction of TiO2 nanoparticle phase and shape transitions controlled by surface chemistry. Nano Lett 5: 1261-1266. doi: doi.org/10.1021/nl050355m.
    » https://doi.org/10.1021/nl050355m.
  • BICKLEY IB, GONZALEZ-CARRENO T & LEES JS. 1991. A structural investigation of titanium dioxide photocatalysts. J Solid State Chem 92: 178-190. doi: 10.1016/0022-4596(91)90255-G.
  • CAO DD, WANG QY, LIU ZY, ZHANG H, WANG YJ , JIN RC & GAO SM. 2019. Enhanced the photoelectrocatalytic performance of TiO2 nanotube arrays by the synergistic sensitization of AgeAgBr nanospheres. Spectrochim Acta A 227: 1386-1425. doi:10.1016/j.saa.2019.117674.
  • CHEN S, HU Y, MENG S & FU X. 2014. Study on the separation mechanisms of photogenerated electrons and holes for composite photocatalysts g-C3N4-WO3. Appl Catal B-Environ 150-151: 564-573. doi: 10.1016/j.apcatb.2013.12.053.
  • CHIARELLO GL, ROSSETTI I & FORNI L. 2005. Flame-spray pyrolysis preparation of perovskites for methane catalytic combustion. J Catal 236: 251-261. doi: 10.1016/j.jcat.2005.10.003.
  • CHOI Y, KIM H, MOON G, JO SG & CHOI WY. 2016. Boosting up the low catalytic activity of silver for H2 production on Ag/TiO2 photocatalyst: thiocyanate as a selective modifier. ACS Publications 6: 821-828. doi: 10.1021/acscatal.5b02376.
  • DICESARE N & LAKOWICZ JR. 2001. Spectral properties of fluoro-phores combining the boronic acid group with electron do-nor or withdrawing groups. Implication in the development of fluorescence probed for saccharides. J Phys Chem A 105: 6834-6840.
  • DUY NV, HIEU NV, HUY PT, CHIEN ND & THAMILSELVAN M. 2008. Mixed SnO2/TiO2 included with carbon nanotubes for gas-sensing application. Physica E 41: 258-263. doi: 10.1016/j.physe.2008.07.007.
  • FENG CY, DENG YC, TANG L, ZENG GM & WANG JJ. 2018. Core-shell Ag2CrO4/N-GQDs@g-C3N4 composites with anti-photocorrosion performance for enhanced full-spectrum-light photocatalytic activities. Appl Catal B: Environ 239: 525-536. doi: 10.1016/j.apcatb.2018.08.049.
  • FITRA M, DAUT I, IRWANTO M, GOMESH N & IRWAN YM. 2013. Effect of TiO2 thickness dye solar cell on charge generation. Energy Procedia 36: 278-286. doi: 10.1016/j.egypro.2013.07.032.
  • GAO SW, LAN Z, WU WX, QUE LF & WU JH. 2014. Fabrication and photovoltaic performance of high efficiency front-illuminated dye-sensitized solar cell based on ordered TiO2 nanotube arrays. Acta Phys Chim Sin (3): 446-452. doi: 10.3866/PKU.WHXB201401022.
  • GHAYEB MMY & GHONCHEGI Z. 2015. Fabrication and characterization of copper doped TiO2, nanotube arrays by in situ electrochemical method as efficient visible-light photocatalyst. Ceram Int 41: 8735-8741. doi: 10.1016/j.ceramint.2015.03.094.
  • HONG Y, LI C, ZHANG G, MENG Y & YIN. 2016. Efficient and stable Nb2O5 modified g-C3N4 photocatalyst for removal of antibiotic pollutant. Chem Eng J 299: 74-84. doi: 10.1016/j.jece.2015.03.021.
  • HUANG K, LIU C, YAN X, HUANG CY & CHEN JB. 2016. Hydrothermal synthesis of g-C3N4/CdWO4 nanocomposite and enhanced photocatalytic activity for tetracycline degradation under visible light. Crystengcomm 18: 6453-6463. doi: 10.1016/j.seppur.2017.07.026.
  • HWANG KJ, LEE JW, YOO SJ, JEONG SY & JEONG DH. 2013. Fine size-regulation of nanocrystalline anatase-TiO2 via sol-gel synthesis and subsequent phase transformation by calcination. New J Chem 37: 1378-1384. doi: 10.1039/C3NJ41170B.
  • LI H, WU X, YIN S, KATSUMATA K & WANG YH. 2017. Effect of rutile TiO2 on the photocatalytic performance of g-C3N4/brookite-TiO2-xNyphotocatalyst for NO decomposition. Appl Surf Sci 392: 531-539. doi: 10.1016/j.apsusc.2016.09.075.
  • LI R, LI Q, ZONG L, WANG XD, YANG JJ. 2013. BaTiO3/TiO2 heterostructure nanotube arrays for improved photoelectrochemical and photocatalytic activity. Electrochim Acta 91: 30-35. doi: 10.1016/j.electacta.2012.12.073.
  • LI YZ, ZHANG H, GUO ZM, HAN JJ, ZHAO XJ, ZHAO QG & KIM SJ. 2008. Highly efficient visible-light-induced photocatalytic activity of nanostructured AgI/TiO2 photocatalyst. Langmuir 24: 8351-8357. doi:10.1021/la801046u.
  • LIANG YC, WANG CC, KEI CC, HSUEH YC & CHO WH. 2011. Photocatalysis of Ag-loaded TiO2 nanotube arrays formed by atomic layer deposition. J Physica Chem C 115: 9498-9502. doi: 10.1021/cg301230w.
  • LIU J, LI Y & ARUMUGAM S. 2018. Investigation of low temperature processed titanium dioxide (TiO2) films for printed dye sensitized solar cells (DSSCs)for large area flexible applications. Mater Today Commun 5: 13846-13854. doi: 10.1016/j.matpr.2018.02.026.
  • LIU R, WANG P, WANG XF, YU HG & YU JG. 2020. UV- and visible-light photocatalytic activity of simultaneously deposited and doped Ag/Ag(I)-TiO2 photocatalyst. J Phys Chem C 14: 16-22. doi:10.1021/jp305774n.
  • LUO J, ZHOU XS, MA L & XU Y. 2015. Enhancing visible-light photocatalytic activity of g-C3N4 by doping phosphorus and coupling with CeO2 for the degradation of methyl orange under visible light irradiation. RSC Adv 5: 568728-68735. doi: 10.1039/C5RA10848A.
  • MACAK JM, ZLAMAL M, KRYSA J & SCHMUKI P. 2007. Self-organized TiO2 nanotube layers as highly efficient photocatalysts. small 3: 300-304. doi: 10.1002/smll.200600426.
  • MAMAKHEL A, TYRSTED C, JESEN ED, HALD P & IVERSEN BB. 2013. Direct formation of crystalline phase pure rutile TiO2, nanostructures by a facile hydrothermal method. Cryst Growth Des 13: 4730-4734. doi: 10.1021/cg400858p.
  • MI JL, CLAUSEN C & BREMHOLM M. 2012. Rapid hydrothermal preparation of rutile TiO2, nanoparticles by simultaneous transformation of primary brookite and anatase: An in situ synchrotron PXRD study. Cryst Growth Des 1: 6092-6097. doi: 10.1016/j.cej.2016.04.092.
  • MOHAMED AER & ROHANI S. 2011. Modified TiO2 nanotube arrays (TNATs): progressive strategies towards visible light responsive photoanode, a review. Energ Environ Sci 4: 1065-1086. doi: 10.1039/C0EE00488J.
  • MORRIS D, EGDELL RG & MATER J. 2001. Application of V-doped TiO2 as a sensor for detection of SO2. Chem 11: 3207-3210. doi: 10.1039/B104801P.
  • PHAM TD, LEE BK, LEE CH & RHEUMATOL BJ. 2016. The advanced removal of benzene from aerosols by photocatalytic oxidation and adsorption of Cu-TiO2/PU under visible light irradiation. Appl Catal B: Environ 36: 1341-1342. doi: 10.1016/j.apcatb.2015.09.023.
  • QAMAR M, DRMOSH Q, AHMED MI, QAMARUDDIN M & YAMANI ZH. 2015. Enhanced photoelectrochemical and photocatalytic activity of WO3-surface modified TiO2 thin film. Nanoscale Res Lett 10: 54. doi: 10.1186/s11671-015-0745-2.
  • QIAN X, HAN H, CHEN Y & YUAN Y. 2018. Sol-gel solvothermal route to synthesize anatase/brookite /rutile TiO2, nanocomposites with highly photocatalytic activity. J Sol-Gel Sci Technol 85: 394-401. doi: 10.1007/s10971-017-4544-3.
  • SUGIMOTO T, ZHOU X & MURAMATSU A. 2002. Synthesis of Uniform anatase TiO2 nanoparticles by gel–sol method: 1. solution chemistry of Ti(OH)n(4-n)+ complexes. J. Colloid. Inter Sci 252: 339-346. doi: 10.1006/jcis.2002.8454.
  • SUN H, GUI Y, WEI H, LONG YK, WANG Q & TANG C. 2019. DFT study of SF6 decomposed products on Pd–TiO2: gas sensing mechanism study. Adsorption 25: 1643-1653. doi: 10.1007/s10450-019-00150-1.
  • TIAN H, MA J, LI K & LI JJ. 2009. Hydrothermal synthesis of S-doped TiO2, nanoparticles and their photocatalytic ability for degradation of methyl orange. Ceram Int 35: 1289-1292. doi: 10.1016/j.ceramint.2008.05.003.
  • TIENG S, KANAEV A & CHHOR K. 2011. New homogeneously doped Fe(III)-TiO2 photocatalyst for gaseous pollutant degradation. Appl Catal A: Gen 399: 191-197. doi: 10.1016/j.apcata.2011.03.056.
  • TONDA S, KUMAR S & SHANKER V. 2015. In situ growth strategy for highly efficient Ag2CO3/g-C3N4 hetero/nanojunctions with enhanced photocatalytic activity under sunlight irradiation. J Environ Chem Eng 3: 852-861. doi: 10.1016/j.jece.2015.03.021.
  • WANG Y, FANG HB, ZHENG YZ, YE R & TAO X. 2015. Controllable assembly of well-defined monodisperse Au nanoparticles on hierarchical ZnO microspheres for enhanced visible-light-driven photocatalytic and antibacterial activity. Nanoscale 7: 19118-19128. doi: 10.1039/C5NR06359K.
  • WILEY JP, HUGHES KA, KAISER RJ, KESICKI EA & LI GS. 2001. Phenylboronic acid salicylhydroxamic acid bioconjugates 2 polyvalent im-mobilization of protein ligands for affinity chromatography. Bioconjugate Chem 12: 240-250. doi: 10.1021/bc0000942.
  • WU J, FENG YJ, LOGAN BE, DAI CC & HAN XY. 2019. Preparation of Al–O-linked porous-g-C3N4/TiO2-nanotube Z-Scheme composites for efficient photocatalytic CO2 conversion and 2,4-dichlorophenol decomposition and mechanism. ACS Sustainable Chem Eng 7: 15289-15296. doi: 10.1021/acssuschemeng.9b02489.
  • XIAO JQ, LIN KS & YU Y. 2018. Novel Ag@AgCl@AgBr heterostructured nanotubes as high-performance visible-light photocatalysts for decomposition of dyes. Catal today 314: 10-19. doi:10.1016/j.cattod.2018.04.019.
  • ZENG M, LI Y, MAO M, BAI L & REN L. 2015. Synergetic effect between photocatalysis on TiO2 and thermocatalysis on CeO2 for gas-phase oxidation of benzene on TiO2/CeO2 nanocomposites. ACS Catal 5: 3278-3286. doi: 10.1021/acscatal.5b00292.
  • ZHANG DQ, LI GS & YANG XF. 2009. A micrometer-size TiO2 single-crystal photocatalyst with remarkable 80% level of reactive facets. Chem Commun 29: 4381-4383. doi: 10.1021/cg400858p.
  • ZHANG M, LU D, ZHANG Z, GUO Y & YANG J. 2015. Enhanced photocurrent and photocatalytic degradation of methyl orange by VN codoped TiO2 nanotube arrays cooperated with H2O2. J Electrochem Soc 162: 557-563. 10.1149/2.0901508jes.

Publication Dates

  • Publication in this collection
    11 Aug 2023
  • Date of issue
    2023

History

  • Received
    22 July 2020
  • Accepted
    09 Oct 2020
Academia Brasileira de Ciências Rua Anfilófio de Carvalho, 29, 3º andar, 20030-060 Rio de Janeiro RJ Brasil, Tel: +55 21 3907-8100 - Rio de Janeiro - RJ - Brazil
E-mail: aabc@abc.org.br