Next Article in Journal
Tetramethylammonium Fluoride: Fundamental Properties and Applications in C-F Bond-Forming Reactions and as a Base
Previous Article in Journal
The Merit and the Context of Hydrogen Production from Water and Its Effect on Global CO2 Emission
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Microflowery, Microspherical, and Fan-Shaped TiO2 Crystals via Hierarchical Self-Assembly of Nanorods with Exposed Specific Crystal Facets and Enhanced Photocatalytic Performance

1
Department of Chemistry and Chemical Engineering, Jinzhong University, Jinzhong 030619, China
2
Department of Chemistry, Taiyuan Normal University, Jinzhong 030619, China
*
Authors to whom correspondence should be addressed.
Submission received: 27 December 2021 / Revised: 15 February 2022 / Accepted: 15 February 2022 / Published: 18 February 2022

Abstract

:
In this paper, khaki titanium dioxide (TiO2) crystals via hierarchical self-assembly of nanorods with different morphologies and specific exposed crystal facets were prepared for the first time by using a TiCl3 treatment process in the presence and absence of morphology-controlling agents. The crystal structure, morphology, microstructure, specific surface area, and separation efficiency of photogenerated electron-hole pairs of the synthesized TiO2 crystals were characterized. The photocatalytic and recycled performances of the synthesized TiO2 crystals in the presence of shape-controlling agents, such as ammonium sulfate (AS), ammonium carbonate (AC), and urea, and the absence of shape-controlling agents (the obtained TiO2 crystals were expressed as AS-TiO2, AC-TiO2, urea-TiO2, and No-TiO2, respectively) were evaluated and compared with the commercial TiO2 (CM-TiO2) crystals. The AS-TiO2 microspheres with exposed uncertain facets exhibited enhanced photocatalytic activity for the degradation of methylene blue solution, which can be attributed to the combined effect of the anatase phase structure, relatively larger specific surface area, and the effective separation of the photogenerated electron-holes.

1. Introduction

Since Fujishima and Honda first discovered the photocatalytic water-splitting reaction in 1972, TiO2 has become the most suitable and promising semiconductor material in the application of water splitting, dye-sensitized solar cells, Li-ion batteries, gas sensors, etc., because of its advantages of high physical and chemical stability, non-toxic, harmlessness, environmental friendliness, and low price [1,2,3,4]. The design and synthesis of TiO2 with different polymorphs, morphologies, and exposed facets are key to effectively improving their practical application in the field of photocatalysis, as its photocatalytic activities depend critically on the crystal phase, morphology, size, surface area, heterojunction structure, and exposed facets of TiO2 [5,6,7,8]. Among the four polymorphs of TiO2 (i.e., anatase, rutile, brookite, and TiO2 (B)) that mainly exist in nature [9,10,11], anatase usually exhibits the highest photocatalytic activity due to the most increased electron mobility and the lowest photogenerated electron-hole recombination in anatase and the rapid interaction between many organic molecules and anatase surfaces [2]. However, compared with the anatase phase (3.20 eV), brookite phase (3.40 eV), and TiO2 (B) (3.10 eV), the decreased band gap energy of rutile (3.02 eV) leads to part of the photoresponse extending slightly into the visible light region, thus improving the utilization of sunlight [12]. Moreover, rutile is the most thermodynamically stable structure (anatase and brookite phases can be irreversibly transformed into the rutile phase during heating) with the advantages of a high refractive index and good light-reflecting performance [13,14]. Therefore, it is imperative to synthesize rutile and anatase with exposed high-energy surfaces for practical application.
For example, rutile TiO2 nanoflakes with exposed {110} facets were synthesized by using titanium (IV) isopropoxide as the titanium source and hydrochloric acid as the morphology-controlling agent under hydrothermal conditions, which exhibited an enhanced photocatalytic activity (91%) for the degradation of cinnamic acid [15]. Nanotubes/nanowires assembled from anatase TiO2 nanoflakes with exposed {111} facets were prepared using titanium oxysulfate-sulfuric acid hydrate as the titanium source and glacial acetic acid as the morphology-controlling agent under high-temperature conditions, which exhibited an improved photocatalytic activity (17.4%) for CO2 reduction to CH4 [16]. Anatase TiO2 microspheres assembled from ultrathin nanosheets with exposed 100% {001} facets were prepared by using potassium fluorotitanate as the precursor via an on-site precipitation hydrothermal method [17]. Anatase/brookite with tunable ratios were obtained using titanium bis(ammonium lactate) dihydroxide (TALH) as a titanium source and urea as an in situ OH source under hydrothermal conditions [18]. Brookite single-crystal nanosheets with exposed {210}, {101}, and {201} facets were prepared using TiCl4 as a titanium source, urea as an in situ OH source, and sodium lactate as the complexant and surfactant under low-basicity conditions, which exhibited a superior photocatalytic activity toward degradation of methyl orange [19]. Anatase TiO2 nanoparticles exposed to different percentages of {001} facets (5~60%) were synthesized using K-titanate nanowires as a precursor and different amounts of urea as a morphology-controlling agent [20]. Rounded anatase TiO2, rod-like rutile TiO2 growing along the [001] direction, and plate-like brookite TiO2 with exposed {111} facets were prepared by hydrothermal treatment of the mixed solution of TiCl3 and H2O2 with different concentrations under certain pH conditions [21]. Moreover, anatase TiO2 nanocrystals with exposed {010}, {001}, and {111} facets and rutile TiO2 nanorods with exposed {110} facets on the lateral surface were also prepared using the layered titanate as the titanium source, which exhibited good photovoltaic performance (5.14%) or superior photocatalytic activity for the degradation of organic rhodamine B (76.0%) and methyl orange (96.5%) molecules [22,23,24].
Herein, we report on the facile synthesis of rutile/brookite composites (AC-TiO2) containing tufted microflowers self-assembled by tetragonal-shaped rutile TiO2 nanorods with oriented growth along the [001] direction and irregular brookite TiO2 nanoparticles with exposed {001} facets, anatase TiO2 microspheres (AS-TiO2) self-assembled by nanorods with exposed uncertain facets, and fan-shaped particles and microspheres (urea-TiO2 and No-TiO2) self-assembled by tetragonal-shaped rutile TiO2 nanorods with oriented growth along the [001] direction. The preparation of microflowers, microspheres, and fan-shaped TiO2 particles with different crystal forms and various exposed crystal facets using TiCl4, TiOSO4, TALH, etc., as raw material and urea as a morphology-controlling agent has been reported in previous literature [18,19,20,21,25]; however, the previous reports rarely used AC and AS morphology-controlling agents. Furthermore, although there are some reports on the synthesis of TiO2 with TiCl3 and different morphology control agents, there are few reports on the simultaneous use of TiCl3 as a titanium source and urea as a morphology control agent, and few khaki TiO2 samples have been prepared [21,26]. In this study, khaki TiO2 crystals with different crystal forms, different morphologies, and different exposed crystal planes were prepared by simply changing the type of morphology-controlling agent, which is innovative to a certain extent. The anatase microspheres (AS-TiO2) constructed with anatase TiO2 nanorods with exposed uncertain facets exhibited the highest photocatalytic performance to that of the AC-TiO2, urea-TiO2, No-TiO2, and CM-TiO2 crystals, which can be attributed to its anatase phase structure, relatively high specific surface area, and the effective separation of the photogenerated electron-holes.

2. Results and Discussion

2.1. Crystal Structure, Morphology, and Exposed Facets

To identify the crystal phase and estimate the phase composition and crystalline size of the as-prepared khaki crystals, the XRD technique was carried out. The XRD patterns of the TiO2 crystals prepared in the presence and absence of shape-controlling agents are shown in Figure 1. The samples were designated as AC-TiO2, AS-TiO2, urea-TiO2, and No-TiO2 according to the addition of shape-controlling agents. The diffraction peaks of AC-TiO2 at 2θ values of 27.50°, 36.20°, 39.30°, 41.32°, 44.14°, 54.42°, 56.68°, 62.84°, 64.21°, 69.16°, and 69.88° were indexed to the (110), (101), (200), (111), (210), (211), (220), (002), (310), (301), and (112) planes of rutile TiO2 (Joint Committee on Powder Diffraction Standards, JCPDS no. 21-1276), while the diffraction peaks at 2θ values of 25.46°, 25.76°, and 30.88° were indexed to the (120), (111), and (121) planes of brookite TiO2 (JCPDS no. 29-1360) as shown in Figure 1a. The percentage of rutile TiO2 (78.3%) and brookite TiO2 (21.7%) in the AC-TiO2 crystals can be estimated using the following equations [16]:
W R = I R I R + 2.721 I B   W B = 2.721 I B I R + 2.721 I B
where WR and WB represent the weight fraction of rutile and brookite TiO2, respectively [23,27]. IR and IB are the integrated intensity of the rutile TiO2 (110) peak (100.0%) and brookite TiO2 (121) peak (10.2%), respectively [14,16]. For AS-TiO2, the diffraction peaks at 2θ values of 25.34°, 37.08°, 37.88°, 38.64°, 48.10°, 54.00°, 55.14°, 62.86°, 68.88°, 70.30°, and 75.20°, corresponding to the (101), (103), (004), (112), (200), (105), (211), (204), (116), (220), and (215) crystal planes of anatase TiO2 (JCPDS no. 21-1272) as shown in Figure 1b. The diffraction peaks of the urea-TiO2 and No-TiO2 crystals prepared under different conditions corresponded to those of a pure rutile TiO2 (JCPDS no. 21-1276), regardless of the difference in intensity as shown in Figure 1c,d. The diffraction peak intensity of No-TiO2 was stronger than that of urea-TiO2, indicating an improvement in the crystalline sizes and crystallinity.
Figure 2 presents the typical FESEM images of the samples obtained by the solvothermal synthesis method at 160 °C by varying the shape-controlling agents (i.e., AC, AS, and urea) using titanium trichloride (TiCl3) as a precursor and anhydrous and distilled water as solvents. As shown in Figure 2a–c, the AC-TiO2 sample was mainly composed of defined tetragonal-shaped nanorods with a length of 0.15~0.70 μm and width of 15~100 nm and some irregular nanoparticles. Enlarged images revealed that the tetragonal-shaped nanorods were formed by the coalescence of a small number of nanorods. Many tufted microflowers self-assembled by tetragonal-shaped nanorods were observed, indicating that the nanorods began to grow out of the flower clusters [4]. Figure 2d shows the FESEM images of AS-TiO2 microspheres with a diameter of 2.65~5.57 μm. Figure 2e displays an enlarged image of a microsphere, which was self-assembled by nanorods, with a length of several micrometers and width of 20~100 nm along with any directions. As shown in Figure 2f,g, the urea-TiO2 sample was mainly composed of fan-shaped particles, tufted microflowers, and microspheres with an average diameter of 3.50 μm, which were self-assembled by tetragonal-shaped nanorods several micrometers in length and 17~70 nm in width along different directions. Figure 2h,i show the FESEM images of the No-TiO2 sample. It can be seen that the tufted No-TiO2 microflowers were formed by self-assembly of tetragonal-shaped nanorods with a length of 0.30~1.20 μm and a width of 40~180 nm.
The energy-dispersive X-ray spectroscopy (EDS) parameters of prepared TiO2 samples are listed in Table 1. The ready TiO2 samples contained not only titanium and oxygen elements but also a small amount of chlorine (from TiCl3) and other elements from the raw material as shown in Table 1. A tiny amount of carbon in the obtained AC-TiO2 and urea-TiO2 samples came from ammonium carbonate and urea, respectively, and a small amount of sulfur in the obtained AS-TiO2 came from ammonium sulfur.
The facet feature and the crystal growth behavior of the obtained TiO2 samples were further characterized by TEM and HRTEM. Figure 3 shows the TEM and HRTEM images of the AC-TiO2, AS-TiO2, urea-TiO2, and No-TiO2 samples synthesized by the solvothermal process at 160 °C by varying the shape-controlling agents. The TEM image of the AC-TiO2 sample obtained shows a nanorod-shaped morphology with a length of 110~300 nm and a width of 15~45 nm, which is consistent with those observed by FESEM (Figure 3a). For the HRTEM image of AC-TiO2 (Figure 3b), the lattice fringes parallel to the lateral planes were measured to be 0.33 ± 0.01 nm, corresponding to the interplanar distance of the rutile TiO2 (110) planes, indicating that the nanorod are likely to grow along the [001] direction [4]. The fast Fourier transform (FFT) diffraction pattern of the white, dashed lines region (Figure 3b inset) further indicated that the nanorod-shaped TiO2 crystal was single-crystalline. The distances between the lattice fringes, 0.35 ± 0.01 and 0.35 ± 0.01 nm, can be assigned to the (120) and (−120) planes of the brookite TiO2 phase, respectively, and the angle labeled at 80° is identical to the theoretical value for the angle between the (120) and (−120) planes, suggesting that the irregular nanoparticle expose {001} facets on its surface (Figure 3b). The 0.35 ± 0.01 nm of the interplanar spacing of the cuboid-like nanoparticles can be assigned to the (120) planes of the brookite TiO2 (Figure 3c). The above analysis further confirms that the synthesized AC-TiO2 was a mix-phase of rutile and brookite TiO2. Figure 3d–f show TEM and HRTEM images of the AS-TiO2 prepared by the solvothermal treatment of the TiCl3 solution at 160 °C. Dimer particles formed by the aggregation of many nanorods in a specific direction were observed as shown in Figure 3d. The visible lattice fringes with an interplanar spacing of 0.35 ± 0.01 nm matched well with the (101) crystal plane of tetragonal anatase TiO2 (Figure 3e). Viewed along the [010] direction, there were two atomic planes (101) and (002), the lattice spacing was 0.35 ± 0.01 and 0.48 ± 0.01 nm, respectively, and the interface angle was 68.3°, further indicating that the as-prepared AS-TiO2 sample was a single-crystalline anatase TiO2 (Figure 3f) [28,29]. Figure 3g–i present the TEM and HRTEM images of the urea-TiO2 sample obtained by solvothermal reaction of the TiCl3 solution with urea as the shape-controlling agent at 160 °C. Three sets of lattice fringes with intervals of 0.25 ± 0.01, 0.25 ± 0.01, and 0.33 ± 0.01 nm and angles of 45°, 67.5°, and 67.5° can be identified in the HRTEM image, which is in good agreement with the spacing of the (101), (011), and (110) planes of the tetragonal rutile TiO2, respectively, indicating that the as-prepared urea-TiO2 nanorods were single-crystalline. Figure 3i shows a lattice image from the top and side of a tetragonal-shaped nanorod. The distance between two consecutive planes was measured to be 0.33 ± 0.01 nm, which matches well with the distance between the (110) planes of tetragonal rutile TiO2. Furthermore, the lattice fringes were parallel to the lateral planes, indicating that the tetragonal rutile TiO2 grew along the [001] direction. The FFT diffraction pattern of the white, dashed line region (Figure 3i inset) further indicates that the tetragonal rutile TiO2 crystal was a single-crystalline and grew along the [001] direction. Figure 3j shows a TEM image of the tetragonal-shaped No-TiO2 nanorods obtained by solvothermal reaction at 160 °C for 24 h in the absence of a shape-controlling agent. Figure 3k shows a typical HRTEM image of a tetragonal-shaped No-TiO2 nanorod. The growth direction of the No-TiO2 nanorod can be determined as [001] direction by the FFT diffraction pattern (Figure 3k inset). Figure 3l presents a lattice image taken from the middle of a nanorod. The existence of the two atomic planes, (110) and (001), with a lattice spacing of 0.33 ± 0.01 and 0.30 ± 0.01 nm between two consecutive planes and an interfacial angle of 90° between the (110) and (001) crystal planes, proves that the tetragonal-shaped No-TiO2 nanorod should be mainly exposed with {001} facets on its top planes, and grow along the [001] direction. According to the above structural analysis, the tetragonal-shaped rutile TiO2 nanorods (i.e., AC-TiO2, urea-TiO2, and No-TiO2) were mainly exposed to {001} facets on their top planes, and grew along the [001] direction, while the exposed crystal facets of the AS-TiO2 anatase nanorods on the basal surface is uncertain.

2.2. X-ray Photoelectron Spectroscopy Analysis

The surface chemical states of the pure anatase, rutile, and mixed-phase TiO2 were studied with X-ray photoelectron spectroscopy (XPS). Figure 4 shows the surveys of the AC-TiO2, AS-TiO2, urea-TiO2, No-TiO2, and CM-TiO2. Only the correlation peaks of C, Ti, and O are observed in Figure 4a, indicating that the chemical purities of the synthetic and commercial samples were very high. Figure 4b displays the Ti 2p spectra of AC-TiO2, AS-TiO2, urea-TiO2, No-TiO2, and CM-TiO2. Two typical peaks at binding energies of 458.6 and 464.4 eV (or 458.9 and 464.8 eV) observed in the as-prepared and commercial TiO2 samples can be attributed to the Ti4+ states of TiO2 [30]. As for the AC-TiO2, AS-TiO2, urea-TiO2, No-TiO2s, and CM-TiO2 samples, the corresponding binding energies were centered at 458.6 (Ti 2p2/3) and 464.4 eV (Ti 2p1/2) for the AC-TiO2, urea-TiO2, No-TiO2, and CM-TiO2 samples, while the binding energies were centered at 458.9 (Ti 2p2/3) and 464.8 eV (Ti 2p1/2) for the AS-TiO2 sample. This slight discrepancy in binding energies can be attributed to the different surface atomic arrangements and configurations of the as-prepared TiO2 and CM-TiO2 samples with different crystal phases [30,31]. The Ti 2p2/3 peaks shifted from 458.9 of the AS-TiO2 to 458.6 eV for the AC-TiO2, urea-TiO2, No-TiO2, and CM-TiO2 accompanying the negative shift of the Ti 2p1/2 peaks from 464.8 to 464.4 eV, suggesting the partial reduction of TiO2 with the formation of Ti3+ ions on the surface of the as-prepared AC-TiO2, urea-TiO2, No-TiO2, and CM-TiO2 [32]. The high-resolution O 1s and C 1s XPS spectra of the AC-TiO2, AS-TiO2, urea-TiO2, No-TiO2, and CM-TiO2 are shown in Figure 4c,d, respectively. The singe peak for O 1s at 529.8 eV (or 530.2 eV) corresponded to the crystal lattice oxygen in the as-prepared and the commercial TiO2, and the O 1s peaks shifted from 530.2 for the AS-TiO2 to 529.8 eV for the AC-TiO2, urea-TiO2, No-TiO2, and CM-TiO2, suggesting the existence of more oxygen vacancies on the surface of the as-prepared AC-TiO2, urea-TiO2, No-TiO2, and CM-TiO2 [32,33]. Furthermore, the khaki colors of the as-prepared TiO2 samples further confirmed the existence of Ti3+ ions and oxygen vacancies [30] as shown by their digital photograph presented in Figure 8. The two peaks for C 1s at 284.8 and 288.6 eV corresponded to the C–O and C=O bonds in hydrocarbons [34]. In addition, signals related to N were detected at 400.1 eV in the spectra of AC-TiO2, AS-TiO2, and urea-TiO2 (Figure 4e), which was assigned to N 1s from the imino group (=NH) of the ammonium carbonate, ammonium sulfate, and urea, respectively. Signals related to S were detected at 168.9 eV in the spectra of AS-TiO2, which was assigned to S 2p1/2 from the thiol group (–SH) of ammonium sulfate (Figure 4f).

2.3. Photoluminescence Analysis

To understand the origin of the different photocatalytic activities of the prepared AC-TiO2, AS-TiO2, urea-TiO2, and No-TiO2 samples, the recombination of photogenerated electron-hole pairs was studied by PL spectroscopy. Figure 5 exhibits the photoluminescence (PL) spectra of the prepared AC-TiO2, AS-TiO2, urea-TiO2, and No-TiO2 samples excited by 325 nm. As reported, a stronger intensity of PL indicates higher efficiency of electron and hole recombination and, as a consequence, worse charge separation [35]. However, PL emission intensity can be affected by many other factors such as the particle size (PL emission intensity increases with the decrease in particle size), dopant species, semiconductor nanomaterial form, geometry of the particles (which can affect the light diffusion), and number of active defects [36]. The observed broad visible photoluminescence was mainly related to self-trapped excitons and oxygen vacancy-related defect states in the as-prepared TiO2 samples [9]. As shown in Figure 5, the pronounced emission peaks at 558 nm were assigned to oxygen vacancies in the AS-TiO2, AC-TiO2, No-TiO2, and urea-TiO2 samples, which were caused by the rapid recombination of photogenerated electron-hole pairs [37,38]. Except for AS-TiO2, the PL emission spectra of all TiO2 located between 350 and 650 nm, and the emission peak at 394 nm can be attributed to the emission of the band-band PL process of TiO2; the other emission peaks at 394, 436, 467, 481, and 616 nm can be attributed to the excitonic PL process at the band edge of TiO2 [36,39]. In particular, the emission peaks of the as-prepared TiO2 at 467 and 616 nm can be attributed to the shallow trap states originated from oxygen vacancies associated with Ti3+ ions and the intrinsic defect, respectively [40]. The intensity of PL for the TiO2 crystals gradually increased following the order of AS-TiO2, AC-TiO2, No-TiO2, and urea-TiO2, demonstrating that the recombination rate of photogenerated electron-hole pairs increased gradually. The AS-TiO2 crystals had the most robust separation efficiency and the lowest recombination rate, which is conducive to the improvement in the photocatalytic performance.

2.4. Electrochemical Impedance Spectroscopy Analysis

The electron-transfer rate of the as-prepared TiO2 samples was investigated by electrochemical impedance spectroscopy (EIS) as shown in Figure 6. It can be seen that the semicircle radius of the AC-TiO2 sample was smaller than that of AS-TiO2, urea-TiO2, and No-TiO2 in the high-frequency region of the EIS Nyquist plot. In the EIS Nyquist plot, the semicircle radius was related to the electrode resistance, which decreased with electrode resistance. It is universally acknowledged that the separation efficiency of photogenerated electron-hole pairs has an essential impact on the photocatalytic activity. The high conductivity of the AS-TiO2 sample also facilitates electron transfer, resulting in an efficient charge separation [41]. The low conductivity of the AC-TiO2, urea-TiO2, and No-TiO2 samples with the same trend was not conducive to electron transfer, resulting in an inefficient electron-hole separation.

2.5. Photocatalytic Activity

The photocatalytic activity of AC-TiO2, AS-TiO2, urea-TiO2, and No-TiO2 was evaluated by the degradation of organic dye MB under UV irradiation. To better evaluate their photocatalytic efficiency, CM-TiO2 was selected as the photocatalytic benchmark. Figure 7a shows the temporary changes in the adsorption spectra of MB with an irradiation time in the presence of AS-TiO2. The intensity and wavelength of the maximal absorption peak in the visible region gradually decreased and shifted from 664 to 649 nm with increasing UV irradiation time, respectively, indicating the degradation and N-demethylation of MB [42]. Figure 7b displays the variation in the MB photocatalytic degradation efficiency with the irradiation time in the presence and absence of TiO2 samples. In the absence of TiO2, the self-degradation efficiency of MB in the blank test was low, only 6.5% at 120 min. In the presence of TiO2, the degradation efficiency of MB increased on the order of 33.6% (No-TiO2), 35.5% (urea-TiO2), 38.1% (AC-TiO2), 59.7% (CM-TiO2), and 70.4% (AS-TiO2) at 120 min. AS-TiO2 exhibited enhanced photocatalytic activity under UV light irradiation, increasing by a factor of 10.83, 2.09, 1.98, 1.85, and 1.18 compared with that of the blank, No-TiO2, urea-TiO2, AC-TiO2, and CM-TiO2, respectively. Many factors, such as the crystalline phase, heterojunction structure, morphology, crystallinity, crystallite size, specific surface area, exposed facets, the charge carrier’s generation, and photocatalytic performance, have an important influence on photocatalytic activity [36,43]. The specific surface area increased in the order of No-TiO2 (2.58 m2/g), urea-TiO2 (4.71 m2/g), CM-TiO2 (7.27 m2/g), AS-TiO2 (9.93 m2/g), and AC-TiO2 (12.82 m2/g), which is not completely consistent with the order of the photocatalytic activity, indicating that there are other factors affecting the photocatalytic activity. Among the four crystalline phases of TiO2, anatase usually has the highest photocatalytic activity for the degradation of organic dye molecules [44]. Among the five TiO2 samples, AS-TiO2 showed the highest photocatalytic activity, which can be attributed to its anatase phase structure and relatively high specific surface area. The relatively high photocatalytic activity of CM-TiO2 can be attributed to the different energy band structures of anatase (96.8%) and rutile (3.2%), because the photogenerated electrons transfer from rutile to anatase, whereas the photogenerated holes transfer from anatase to rutile, which can inhibit the charge recombination of the photogenerated electron-hole pairs, thereby enhancing the photocatalytic activity [45,46]. No-TiO2 and urea-TiO2 delivered the lower photocatalytic activity with a degradation efficiency of 38.1% and 35.5%, respectively, because these two samples consisted only of single rutile and had smaller specific surface areas (No-TiO2: 2.58 m2/g; urea-TiO2: 4.71 m2/g). Moreover, the rutile had a direct band gap structure, resulting in the rapid recombination of photogenerated electron-hole pairs [47]. Despite the largest surface area for AC-TiO2 (rutile: 78.3%; brookite: 21.7%) compared to the other samples, lower photocatalytic activity is observed, with a degradation efficiency of 38.1%. This was because rutile and brookite TiO2 have direct band gap structures that cannot provide longer electron-hole lifetimes, leading to the fast recombination of photogenerated electron-hole pairs [47,48]. To characterize the surface photocatalytic activity of the as-prepared TiO2 crystals, the photodegradation amount of MB per surface area was estimated according to the formula, degradation amount per surface area (mg(MB)/m2(TiO2)) = (m0(MB) − mt(MB))/(m(TiO2SBET(TiO2)), where m0(MB) and mt(MB) represent the quality of MB in solution before and after illumination for a certain time, respectively; m(TiO2) and SBET(TiO2) are the quality and specific surface area of TiO2, respectively [49]. The photodegradation amount of MB at 120 min was 1.59, 0.88, 0.85, 0.84, and 0.32 mg (MB)/m2(TiO2) for No-TiO2, urea-TiO2, CM-TiO2, AS-TiO2, and AC-TiO2, respectively, as shown in Figure 7c. The maximum amount of photodegradation per specific surface area of No-TiO2 (1.59 mg(MB)/m2(TiO2)) was due to the small differences in the amount of degradation per gram of MB (4.10, 4.13, 6.21, 8.37, and 4.14 mg(MB)/g(TiO2) for No-TiO2, urea-TiO2, CM-TiO2, AS-TiO2, and AC-TiO2, respectively), while its specific surface area was far lower than that of other TiO2 samples (2.58, 4.71, 7.27, 9.93, and 12.82 m2/g for No-TiO2, urea-TiO2, CM-TiO2, AS-TiO2, and AC-TiO2, respectively). The degradation amount of urea-TiO2 (4.13/4.71 = 0.88), CM-TiO2 (6.21/7.27 = 0.85), and AS-TiO2 (8.37/9.93 = 0.84) per specific surface area was almost the same, although the degradation amount per gram (4.13, 6.21, and 8.37 mg(MB)/g(TiO2) for urea-TiO2, CM-TiO2, and AS-TiO2, respectively) and specific surface area (4.71, 7.27, and 9.93 m2/g for urea-TiO2, CM-TiO2, and AS-TiO2, respectively) were different, which may be due to the presence of rutile in CM-TiO2 and specific crystal planes in urea-TiO2 and AS-TiO2. The smallest photodegradation amount of AC-TiO2 (0.32 mg(MB)/m2(TiO2)) per specific surface area was due to the smaller degradation amount per gram (4.14 mg(MB)/g(TiO2)) and the largest specific surface area (12.82 m2/g). The stability and reusability of the AS-TiO2 and CM-TiO2 were evaluated by carrying out cycling experiments three times for the photodegradation of MB as shown in Figure 7d. The degradation rates of the AS-TiO2 and CM-TiO2 samples decreased by only 5.89% and 6.35% for the MB photodegradation, respectively, after recycling three times, indicating the AS-TiO2 and CM-TiO2 samples possessed good stability and reusability.

3. Materials and Methods

3.1. Materials

Titanium trichloride (TiCl3, 15~20%) and ammonium sulfate ((NH4)2SO4, 99.0%) were purchased from Tianjin Beichen District Fangzheng Reagent Factory (Tianjin, China). Ammonium carbonate (NH4NH2CO2 + NH4HCO3), absolute ethyl alcohol (C2H5OH, 99.7%), and urea were purchased from Tianjin Bodi Chemical Co., Ltd., Tianjin Kemiou Chemical Reagent Co., Ltd. (Tianjin, China), and Shanghai Macklin Biochemical Technology Co., Ltd. (Shanghai, China), respectively. All the chemical reagents used in the experiments were not further purified.

3.2. Synthesis of TiO2 Nanocrystals

Firstly, 100 mL of absolute ethyl alcohol and 100 mL deionized water were poured into a round-bottomed flask with a capacity of 500 mL. Then, 50 mL of TiCl3 was added dropwise to the above round-bottomed flask under magnetic stirring. Thirty minutes later, 62.5 mL of the above purple mixed solution was transformed into three Teflon-lined stainless-steel autoclaves with a capacity of 80 mL. Then, 2.5001 g of ammonium sulfate (AS), 2.5013 g of ammonium carbonate (AC), and 2.5014 g of urea were added to the above autoclaves, respectively. For comparison, 62.5 mL of the above purple solution was transformed into the same autoclave without adding any other chemical reagents. After, the above four autoclaves were placed in a constant temperature blast drying oven and heated at 160 °C for 24 h. The khaki powders obtained were denoted as AS-TiO2, AC-TiO2, urea-TiO2, and No-TiO2, respectively (Figure 8).

3.3. Sample Characterization

The crystallite structure of the khaki TiO2 nanocrystals was determined by powder X-ray diffraction (XRD-6100, Shimadzu, Kyoto, Japan) with monochromated Cu Kα radiation (λ = 0.15406 nm) at a scan speed of 8°/min, an accelerating voltage of 40 kV and applied current of 30 mA. The morphology and microstructure of the samples were examined with a field emission scanning electron microscope (FESEM, Hitachi SU8100, Tokyo, Japan) equipped with energy-dispersive X-ray spectroscopy (EDS) at an accelerating voltage of 15 kV and an applied current of 10 μA in the darkfield mode. After the sample was drop-casted on silicon wafers, a transmission electron microscope (TEM) and a high-resolution transmission electron microscope (HRTEM, FEI TALO F200S, Portland, OR, USA) with a 200 kV operating voltage were used after the sample was deposited on a standard copper grid-supported carbon film. The elemental compositions and chemical status were analyzed by X-ray photoelectron spectroscopy (XPS, Thermo Fisher Scientific K-Alpha, New York, NY, USA) fitted with an Al Kα source (1486.6 eV). The specific surface area was determined by nitrogen gas adsorption (micromeritics ASAP 2020, Micromeritics Instrument Corp., Atlanta, GA, USA). The optical property, carrier migration, and recombination were studied by a fluorescence spectrometer (PL, HORIBA Fluoromax-4, HORIBA Instruments Inc., Kyoto, Japan) and electrochemical impedance spectroscopy (EIS, CHI600E, Shanghai Chenhua Instrument Co., Ltd., Shanghai, China), respectively. The photoelectrochemical measurement was carried out on the electrochemical workstation under the irradiation of a 300 W xenon lamp equipped with a cut-off filter (λ > 420 nm). TFO conductive glass (opening area: 1 cm2), platinum, and an Ag/AgCl electrode were used as the working electrode, the counter electrode, and the reference electrode, respectively, and 0.2 mol/L Na2SO4 solution was used as electrolyte. EIS was recorded in the frequency range of 100 kHz to 0.01 Hz under open circuit potential conditions. The absorbance characteristics of the MB solution were determined by a UV-Vis spectrophotometer (TU 1901, Beijing Purkinje General Instrument Co., Ltd., Beijing, China).

3.4. Photocatalytic Experiments

Photocatalytic activity of the synthesized khaki TiO2 samples was evaluated with methylene blue (MB) as the model pollutant under UV irradiation. Typically, 75 mg of khaki TiO2 sample (AS-TiO2, AC-TiO2, urea-TiO2, and No-TiO2) and 150 mL of 10 mg/L MB (2.23 × 10−5 mol/L) sample were mixed and subjected to magnetic stirring in the dark without irradiation for 2 h to establish the adsorption-desorption equilibrium of the MB dye on the surface of TiO2 samples. A low-pressure mercury lamp irradiation (175 W, λmax = 365 nm, Shanghai Mingyao Glass Hardware Tool Factory, Shanghai, China) with a maximum emission at 365 nm was used as the UV resource, and the distance between the mercury lamp and the suspension surface was 25 cm. Within a given irradiation interval, 3 mL of suspension was taken out and analyzed after removing the solid particles by centrifugation. The concentration change of the MB solution was detected by an ultraviolet-visible spectrophotometer. As a comparison, the photocatalytic activity of the commercial TiO2 (CM-TiO2, 96.8% anatase and 3.2% rutile) sample was also measured under the same conditions. As for stability and recyclability, the TiO2 sample was filtered with a sand core filter and thoroughly dried after each cycle, and then a new MB solution was added for further analysis.

4. Conclusions

In summary, mixed-phase AC-TiO2 crystals grew along the [001] direction (rutile nanorods) and were exposed to {001} facets (irregular brookite nanoparticles) on their basal surface; anatase AS-TiO2 microspheres were formed by the self-assembly of nanorods with uncertain facets; rutile urea-TiO2 microspheres and tufted rutile No-TiO2 microflowers were formed by the self-assembly of nanorods with oriented growth along the [001] direction, and exposed {001} facets on their top surface were synthesized via a facile solvothermal route in the presence of TiCl3 and different morphology-controlling agents. The AS-TiO2 microspheres exhibited the highest photocatalytic activity towards decoloration of the MB solution and achieved 70.4% degradation for the MB solution in 120 min, almost 10.83, 2.09, 1.98, 1.85, and 1.18 times as high as that of the blank, No-TiO2, urea-TiO2, AC-TiO2, and CM-TiO2, respectively. Based on the results of XRD, FESEM, HRTEM, specific surface area, PL, and EIS, the high photocatalytic activity of the AS-TiO2 microspheres can be attributed to the combined effect of the anatase phase structure, relatively larger specific surface area, the highest separation efficiency, and the lowest recombination rate. This work provides a simple and economical method to synthesize different TiO2 crystals with exposed specific crystal surfaces for the photodegradation of organic pollutants.

Author Contributions

Conceptualization, Y.-e.D. and X.N.; methodology, Y.-e.D. and X.N.; formal analysis, K.H. and X.H.; writing—original draft preparation, Y.-e.D.; writing—review and editing, Y.-e.D. and C.Z.; funding acquisition, Y.-e.D. and C.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by grants 201901D111303 and 201801D121257 from the Applied Basic Research Project of Shanxi; grant 2019L0881 from the Shanxi Scientific and Technological Innovation Programs of Higher Education Institutions; grant PY201817 from the Shanxi “1331 Project” Key Innovation Team; grant I018038 from the Shanxi “1331 Project” Collaborative Innovation Center Fund Project; grant jzxycxtd2019005 from the Jinzhong University “1331 Project” Key Innovation Team; Research Start-up Fee of Jinzhong University.

Data Availability Statement

Data are contained within the article and are also available from the first corresponding author.

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. Liu, B.; Khare, A.; Aydil, E.S. TiO2-B/anatase core-shell heterojunction nanowires for photocatalysis. ACS Appl. Mater. Interfaces 2011, 3, 4444–4450. [Google Scholar] [CrossRef]
  3. Sarkar, D.; Ghosh, C.K.; Mukherjee, S.; Chattopadhyay, K.K. Three dimensional Ag2O/TiO2 type-II (p-n) nanoheterojunctions for superior photocatalytic activity. ACS Appl. Mater. Interfaces 2013, 5, 331–337. [Google Scholar] [CrossRef] [PubMed]
  4. Xia, H.R.; Li, J.; Peng, C.; Sun, W.T.; Li, L.W.; Peng, L.M. Floating growth of large-scale freestanding TiO2 nanorod films at the gas-liquid interface for additive-free Li-ion battery applications. Appl. Mater. Interfaces 2014, 6, 17376–17383. [Google Scholar] [CrossRef] [PubMed]
  5. Chuangchote, S.; Jitputti, J.; Sagawa, T.; Yoshikawa, S. Photocatalytic activity for hydrogen evolution of electrospun TiO2 nanofibers. ACS Appl. Mater. Interfaces 2009, 1, 1140–1143. [Google Scholar] [CrossRef] [PubMed]
  6. Roy, N.; Park, Y.; Youngku Sohn, Y.; Leung, K.T.; Pradhan, D. Green synthesis of anatase TiO2 nanocrystals with diverse shapes and their exposed facets-dependent photoredox activity. ACS Appl. Mater. Interfaces 2014, 6, 16498–16507. [Google Scholar] [CrossRef] [PubMed]
  7. Liu, X.; Li, Y.X.; Deng, D.Y.; Chen, N.; Xing, X.X.; Wang, Y.D. A one-step nonaqueous sol-gel route to mixed-phase TiO2 with enhanced photocatalytic degradation of Rhodamine B under visible light. Cryst. Eng. Comm. 2016, 18, 1964–1975. [Google Scholar] [CrossRef]
  8. Yamazaki, S.; Kutoh, M.; Yamazaki, Y.; Yamamoto, N.; Fujitsuka, M. One-pot synthesis of long rutile TiO2 nanorods and their photocatalytic activity for O2 evolution: Comparison with near spherical nanoparticles. ACS Omega 2021, 6, 31557–31565. [Google Scholar] [CrossRef]
  9. Wang, C.H.; Zhang, X.T.; Liu, Y.C. Coexistence of an anatase/TiO2(B) heterojunction and an exposed (001) facet in TiO2 nanoribbon photocatalysts synthesized via a fluorine-free route and topotactic transformation. Nanoscale 2014, 6, 5329–5337. [Google Scholar] [CrossRef]
  10. Wang, Q.; Fang, X.X.; Hao, P.; Chi, W.W.; Huang, F.; Shi, X.F.; Cui, G.W.; Liu, Y.; Tang, B. Green preparation of porous hierarchical TiO2(B)/anatase phase junction for effective photocatalytic degradation of antibiotics. Chem. Commun. 2021, 57, 13024–13027. [Google Scholar] [CrossRef]
  11. Shejale, K.P.; Krishnapriya, R.; Patil, H.; Laishram, D.; Rawalf, P.; Sharma, R.K. Recent advances in ultra-low temperature (sub-zero to 100 °C) synthesis, mechanism and applications of titania (TiO2) nanoparticles. Mater. Adv. 2021, 2, 7502–7529. [Google Scholar] [CrossRef]
  12. Serga, V.; Burve, R.; Krumina, A.; Romanova, M.; Kotomin, E.A.; Popov, A.I. Extraction–pyrolytic method for TiO2 polymorphs production. Crystals 2021, 11, 431. [Google Scholar] [CrossRef]
  13. Dima, R.S.; Phuthu, L.; Maluta, N.E.; Kirui, J.K.; Maphanga, R.R. Electronic, structural, and optical properties of mono-doped and co-doped (210) TiO2 brookite surfaces for application in dye-sensitized solar cells–A first principles study. Materials 2021, 14, 3918. [Google Scholar] [CrossRef] [PubMed]
  14. Ye, M.D.; Zheng, D.J.; Wang, M.Y.; Chen, C.; Liao, W.M.; Lin, C.J.; Lin, Z.Q. Hierarchically structured microspheres for high-efficiency rutile TiO2-based dye-sensitized solar cells. ACS Appl. Mater. Interfaces 2014, 6, 2893–2901. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Huyen, T.T.T.; Chi, T.T.K.; Dung, N.D.; Kosslick, H.; Liem, N.Q. Enhanced Photocatalytic Activity of {110}-Faceted TiO2 Rutile Nanorods in the Photodegradation of Hazardous Pharmaceuticals. Nanomaterials 2018, 8, 276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Liu, J.F.; Liu, B.; Ren, Y.; Yuan, Y.K.; Zhao, H.; Yang, H.Q.; Liu, S.Z. Hydrogenated nanotubes/nanowires assembled from TiO2 nanoflakes with exposed {111} facets: Excellent photocatalytic CO2 reduction activity and charge separation mechanism between (111) and (-1-1-1) polar surfaces. J. Mater. Chem. A 2019, 1, 14761–14775. [Google Scholar] [CrossRef]
  17. Di, J.; Yan, H.B.; Liu, Z.Y.; Ding, X. Synthesis and characterization of anatase TiO2 microspheres self-assembled by ultrathin nanosheets. Materials 2021, 14, 2870. [Google Scholar] [CrossRef]
  18. Kandiel, T.A.; Feldhoff, A.; Robben, L.; Dillert, R.; Bahnemann, D.W. Tailored titanium dioxide nanomaterials: Anatase nanoparticles and brookite nanorods as highly active photocatalysts. Chem. Mater. 2010, 22, 2050–2060. [Google Scholar] [CrossRef]
  19. Lin, H.F.; Li, L.P.; Zhao, M.L.; Huang, X.S.; Chen, X.M.; Li, G.S.; Yu, R.C. Synthesis of high-quality brookite TiO2 single-crystalline nanosheets with specific facets exposed: Tuning catalysts from inert to highly reactive. J. Am. Chem. Soc. 2012, 134, 8328–8331. [Google Scholar] [CrossRef]
  20. Han, X.G.; Wang, X.; Xie, S.F.; Kuang, Q.; Ouyang, J.J.; Xie, Z.X.; Zheng, L.S. Carbonate ions-assisted syntheses of anatase TiO2 nanoparticles exposed with high energy (001) facets. RSC Adv. 2012, 2, 3251–3253. [Google Scholar] [CrossRef]
  21. Li, J.G. TiO2 nanocrystals: Phase selection and morphology control. IJMSE 2013, 1, 5–7. [Google Scholar] [CrossRef]
  22. Du, Y.E.; Li, J.; Liu, Y.F.; Niu, X.J.; Guo, F.; Feng, Q. Synthesis of {110}-faceted rutile TiO2 nanocrystals from tetratitanate nanoribbons for improving dye-sensitized solar cell performance. RSC Adv. 2016, 6, 9717–9724. [Google Scholar] [CrossRef]
  23. Du, Y.E.; Niu, X.J.; Li, W.X.; An, J.; Liu, Y.F.; Chen, Y.Q.; Wang, P.F.; Yang, X.J.; Feng, Q. Microwave-assisted synthesis of high-energy faceted TiO2 nanocrystals derived from exfoliated porous metatitanic acid nanosheets with improved photocatalytic and photovoltaic performance. Materials 2019, 12, 3614. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Du, Y.E.; Niu, X.J.; He, J.; Liu, L.; Liu, Y.F.; Chen, C.D.; Yang, X.J.; Feng, Q. Hollow square rodlike microtubes composed of anatase nanocuboids with coexposed {100}, {010}, and {001} facets for improved photocatalytic performance. ACS Omega 2020, 5, 14147–14156. [Google Scholar] [CrossRef]
  25. Liu, G.; Yang, H.G.; Pan, J.; Yang, Y.Q.; Lu, G.Q.; Cheng, H.M. Titanium dioxide crystals with tailored facets. Chem. Rev. 2014, 114, 9559–9612. [Google Scholar] [CrossRef]
  26. Zhao, W.X.; Ma, S.Q.; Xiang, G.L. Direct synthesis of defective ultrathin brookite-phase TiO2 nanosheets showing flexible electronic band states. Chem. Commun. 2021, 57, 500–503. [Google Scholar] [CrossRef]
  27. Zhang, H.Z.; Banfield, F. Understanding polymorphic phase transformation behavior during growth of nanocrystalline aggregates: Insights from TiO2. J. Phys. Chem. B 2000, 104, 3481–3487. [Google Scholar] [CrossRef]
  28. Zhao, J.; Zou, X.X.; Su, J.; Wang, P.P.; Zhou, L.J.; Li, G.D. Synthesis and photocatalytic activity of porous anatase TiO2 microspheres composed of {010}-faceted nanobelts. Dalton Trans. 2013, 42, 4365–4368. [Google Scholar] [CrossRef]
  29. Chen, C.D.; Ikeuchi, Y.; Xu, L.F.; Sewvandi, G.A.; Kusunose, T.; Tanaka, Y.; Nakanishi, S.; Wen, P.H.; Feng, Q. Synthesis of [111]- and {010}-faceted anatase TiO2 nanocrystals from tri-titanate nanosheets and their photocatalytic and DSSC performances. Nanoscale 2015, 7, 7980–7991. [Google Scholar] [CrossRef]
  30. Liu, X.G.; Du, G.R.; Li, M. True photoreactivity origin of Ti3+-doped anatase TiO2 crystals with respectively dominated exposed {001},{101}, and {100} facets. ACS Omega 2019, 4, 14902–14912. [Google Scholar] [CrossRef] [Green Version]
  31. Pan, J.; Liu, G.; Lu, G.Q.; Cheng, H.M. On the true photoreactivity order of {001}, {010}, and {101} facets of anatase TiO2 crystals. Angew. Chem. Int. Ed. 2011, 50, 2133–2137. [Google Scholar] [CrossRef] [PubMed]
  32. Ren, R.; Wen, Z.H.; Cui, S.M.; Hou, Y.; Guo, X.R.; Chen, J.H. Controllable Synthesis and Tunable Photocatalytic Properties of Ti3+-doped TiO2. Sci. Rep. 2015, 5, 10714. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Nandan, S.; Deepak, T.G.; Nair, S.V.; Nair, A.S. TiO2 nanofibers resembling yellow bristle grass in morphology by a soft chemical transformation. Dalton Trans. 2015, 44, 9637–9645. [Google Scholar] [CrossRef]
  34. Xiong, Z.; Luo, Y.; Zhao, Y.C.; Zhang, J.Y.; Zheng, C.G.; Wu, J.C.S. Synthesis, characterization and enhanced photocatalytic CO2 reduction activity of grapheme supported TiO2 nanocrystals with coexposed {001} and {101} facets. Phys. Chem. Chem. Phys. 2016, 18, 13186–13191. [Google Scholar] [CrossRef] [PubMed]
  35. Wei, X.X.; Cui, B.Y.; Wang, X.X.; Cao, Y.Z.; Gao, L.B.; Guo, S.Q.; Chen, C.M. Tuning the physicochemical property of BiOBr via solvent adjustment: Towards an efficient photocatalyst for water treatment. Cryst. Eng. Comm. 2019, 21, 1750–1757. [Google Scholar] [CrossRef]
  36. Jing, L.Q.; Qu, Y.C.; Wang, B.Q.; Li, S.D.; Jiang, B.J.; Yang, L.B.; Fu, W.; Fu, H.G.; Sun, J.Z. Review of photoluminescence performance of nano-sized semiconductor materials and its relationships with photocatalytic activity. Sol. Energy Mater. Sol. Cells 2006, 90, 1773–1787. [Google Scholar]
  37. Du, Y.E.; Niu, X.J.; He, X.R.; Hou, K.; Liu, H.L.; Zhang, C.F. Synthesis and photocatalytic activity of TiO2/CdS nanocomposites with co-exposed anatase highly reactive facets. Molecules 2021, 26, 6031. [Google Scholar] [CrossRef]
  38. Yang, M.J.; Liu, W.; Sun, J.L.; Zhu, J.L. High magnetic field annealing effect on visible photoluminescence enhancement of TiO2 nanotube arrays. Appl. Phys. Lett. 2012, 100, 043106. [Google Scholar] [CrossRef]
  39. Zhang, G.; Zhang, S.; Wang, L.; Liu, R.; Zeng, Y.; Xia, X.; Liu, Y.; Luo, S. Facile synthesis of bird’s nest-like TiO2 microstructure with exposed (001) facets for photocatalytic degradation of methylene blue. Appl. Surf. Sci. 2017, 391, 228–235. [Google Scholar] [CrossRef]
  40. Santara, B.; Giri, P.K.; Imakita, K.; Fujii, M. Evidence of oxygen vacancy induced room temperature ferromagnetism in solvothermally synthesized undoped TiO2 nanoribbons. Nanoscale 2013, 5, 5476–5488. [Google Scholar] [CrossRef]
  41. Xu, J.; Teng, F.; Xu, C.Y.; Yang, Y.; Yang, L.M.; Kan, Y.D. unique anatase TiO2 twinning crystals formed by high-energy {001} facets and the improved photocatalytic activity. J. Phys. Chem. C 2015, 119, 13011–13020. [Google Scholar] [CrossRef]
  42. Zhang, T.Y.; Oyama, T.; Aoshima, A.; Hidaka, H.; Zhao, J.C.; Serpone, N. Photooxidative N-demethylation of methylene blue in aqueous TiO2 dispersions under UV irradiation. J. Photochem. Photobiol. A 2001, 140, 163–172. [Google Scholar] [CrossRef]
  43. Zhao, W.N.; Li, Y.; Shen, W.J. Tuning the shape and crystal phase of TiO2 nanoparticles for catalysis. Chem. Commun. 2021, 57, 6838–6850. [Google Scholar] [CrossRef] [PubMed]
  44. Ong, W.J.; Tian, L.L.; Chai, S.P.; Yong, S.T.; Mohamed, A.R. Highly reactive {001} facets of TiO2-based composites: Synthesis, formation mechanism and characterization. Nanoscale 2014, 6, 1946–2008. [Google Scholar] [CrossRef] [PubMed]
  45. Tiwari, A.; Mondal, I.; Ghosh, S.; Chattopadhyay, N.; Pal, U. Fabrication of mixed phase TiO2 heterojunction nanorods and its enhanced photoactivities. Phys. Chem. Chem. Phys. 2016, 18, 15260–15268. [Google Scholar] [CrossRef]
  46. Zou, F.X.; Hu, J.W.; Miao, W.J.; Shen, Y.J.; Ding, J.D.; Jing, X.H. Synthesis and characterization of enhanced photocatalytic activity with Li+-doping nanosized TiO2 catalyst. ACS Omega 2020, 5, 28510–28516. [Google Scholar] [CrossRef]
  47. Kim, M.G.; Kang, J.M.; Lee, J.E.; Kim, K.S.; Kim, K.H.; Cho, M.; Lee, S.G. Effects of calcination temperature on the phase composition, photocatalytic degradation, and virucidal activities of TiO2 nanoparticles. ACS Omega 2021, 6, 10668–10678. [Google Scholar] [CrossRef]
  48. Liao, Y.L.; Que, W.X.; Jia, Q.Y.; He, Y.C.; Zhang, J.; Zhong, P. Controllable synthesis of brookite/anatase/rutile TiO2 nanocomposites and single-crystalline rutile nanorods array. J. Mater. Chem. 2012, 22, 7937–7944. [Google Scholar] [CrossRef]
  49. Wen, P.H.; Itoh, H.; Tang, W.P.; Feng, Q. Single nanocrystals of anatase-type TiO2 prepared from layered titanate nanosheets: Formation mechanism and characterization of surface properties. Langmuir 2007, 23, 11782–11790. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of as-prepared (a) AC-TiO2; (b) AS-TiO2; (c) urea-TiO2; (d) No-TiO2 microcrystals in the presence and absence of shape-controlling agents.
Figure 1. XRD patterns of as-prepared (a) AC-TiO2; (b) AS-TiO2; (c) urea-TiO2; (d) No-TiO2 microcrystals in the presence and absence of shape-controlling agents.
Catalysts 12 00232 g001
Figure 2. Different magnifications of FESEM images of the (ac) AC-TiO2; (d,e) AS-TiO2; (f,g) urea-TiO2; (h,i) No-TiO2 samples in the presence and absence of shape-controlling agents.
Figure 2. Different magnifications of FESEM images of the (ac) AC-TiO2; (d,e) AS-TiO2; (f,g) urea-TiO2; (h,i) No-TiO2 samples in the presence and absence of shape-controlling agents.
Catalysts 12 00232 g002
Figure 3. TEM and HRTEM images of the prepared (ac) AC-TiO2; (df) AS-TiO2; (gi) urea-TiO2; (jl) No-TiO2 samples in the presence and absence of shape-controlling agents. The insets in (b,i,k) are fast Fourier transform (FFT) diffraction patterns.
Figure 3. TEM and HRTEM images of the prepared (ac) AC-TiO2; (df) AS-TiO2; (gi) urea-TiO2; (jl) No-TiO2 samples in the presence and absence of shape-controlling agents. The insets in (b,i,k) are fast Fourier transform (FFT) diffraction patterns.
Catalysts 12 00232 g003
Figure 4. XPS spectra of AC-TiO2, AS-TiO2, urea-TiO2, No-TiO2, and CM-TiO2: (a) survey spectra; (b) Ti 2p spectra; (c) O 1s spectra; (d) C 1s spectra; (e) N 1s spectra; (f) S 2p spectra.
Figure 4. XPS spectra of AC-TiO2, AS-TiO2, urea-TiO2, No-TiO2, and CM-TiO2: (a) survey spectra; (b) Ti 2p spectra; (c) O 1s spectra; (d) C 1s spectra; (e) N 1s spectra; (f) S 2p spectra.
Catalysts 12 00232 g004
Figure 5. PL emission spectra (excited at 325 nm) of the prepared AC-TiO2, AS-TiO2, urea-TiO2, and No-TiO2 samples.
Figure 5. PL emission spectra (excited at 325 nm) of the prepared AC-TiO2, AS-TiO2, urea-TiO2, and No-TiO2 samples.
Catalysts 12 00232 g005
Figure 6. Electrochemical impedance spectroscopy Nyquist plots of AC-TiO2, AS-TiO2, urea-TiO2, and No-TiO2 electrodes in 0.2 M Na2SO4 aqueous solution.
Figure 6. Electrochemical impedance spectroscopy Nyquist plots of AC-TiO2, AS-TiO2, urea-TiO2, and No-TiO2 electrodes in 0.2 M Na2SO4 aqueous solution.
Catalysts 12 00232 g006
Figure 7. (a) Temporal adsorption spectral changes of MB in aqueous AS-TiO2 suspensions under UV illumination; (b) photocatalysis degradation efficiency of MB with or without TiO2 sample under UV illumination; (c) photocatalysis degradation amount of MB with different TiO2 samples under UV illumination; (d) recycled performances in the presence of AS-TiO2 and CM-TiO2 samples for photodegradation of MB dye.
Figure 7. (a) Temporal adsorption spectral changes of MB in aqueous AS-TiO2 suspensions under UV illumination; (b) photocatalysis degradation efficiency of MB with or without TiO2 sample under UV illumination; (c) photocatalysis degradation amount of MB with different TiO2 samples under UV illumination; (d) recycled performances in the presence of AS-TiO2 and CM-TiO2 samples for photodegradation of MB dye.
Catalysts 12 00232 g007
Figure 8. Photographs of the as-prepared (a) AS-TiO2; (b) AC-TiO2; (c) urea-TiO2; (d) No-TiO2 microcrystals in the presence and absence of shape-controlling agents.
Figure 8. Photographs of the as-prepared (a) AS-TiO2; (b) AC-TiO2; (c) urea-TiO2; (d) No-TiO2 microcrystals in the presence and absence of shape-controlling agents.
Catalysts 12 00232 g008
Table 1. EDS analysis of the as-prepared TiO2 samples.
Table 1. EDS analysis of the as-prepared TiO2 samples.
SampleAC-TiO2Urea-TiO2AS-TiO2No-TiO2
ElementAtom%Mass%Atom%Mass%Atom%Mass%Atom%Mass%
C4.91.94.81.70.00.00.00.0
N0.00.00.00.00.50.20.00.0
O44.622.440.919.838.017.239.618.0
Cl0.20.20.40.40.10.10.20.2
Ti50.375.553.978.159.981.260.281.8
S0.00.00.00.01.51.30.00.0
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Du, Y.-e.; Niu, X.; Hou, K.; He, X.; Zhang, C. Microflowery, Microspherical, and Fan-Shaped TiO2 Crystals via Hierarchical Self-Assembly of Nanorods with Exposed Specific Crystal Facets and Enhanced Photocatalytic Performance. Catalysts 2022, 12, 232. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12020232

AMA Style

Du Y-e, Niu X, Hou K, He X, Zhang C. Microflowery, Microspherical, and Fan-Shaped TiO2 Crystals via Hierarchical Self-Assembly of Nanorods with Exposed Specific Crystal Facets and Enhanced Photocatalytic Performance. Catalysts. 2022; 12(2):232. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12020232

Chicago/Turabian Style

Du, Yi-en, Xianjun Niu, Kai Hou, Xinru He, and Caifeng Zhang. 2022. "Microflowery, Microspherical, and Fan-Shaped TiO2 Crystals via Hierarchical Self-Assembly of Nanorods with Exposed Specific Crystal Facets and Enhanced Photocatalytic Performance" Catalysts 12, no. 2: 232. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12020232

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