Next Article in Journal
Engineering Noble Metal-like Bi onto Hierarchical SrWO4 for the Enhancement of Photocatalytic Activity
Next Article in Special Issue
Reduced Graphene Oxide–Metal Oxide Nanocomposites (ZrO2 and Y2O3): Fabrication and Characterization for the Photocatalytic Degradation of Picric Acid
Previous Article in Journal
Cloning of Cold-Adapted Dextranase and Preparation of High Degree Polymerization Isomaltooligosaccharide
Previous Article in Special Issue
A Kinetic Study of Photocatalytic Degradation of Phenol over Titania–Silica Mixed Oxide Materials under UV Illumination
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Facile Preparation of Highly Active CO2 Reduction (001)TiO2/Ti3C2Tx Photocatalyst from Ti3AlC2 with Less Fluorine

1
College of Resources and Environment, South-Central Minzu University, Wuhan 430074, China
2
Division of Materials and Manufacturing Science, Graduate School of Engineering, Osaka University, Osaka 565-0871, Japan
3
School of Materials Science and Engineering, Zhengzhou University, Zhengzhou 450001, China
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(7), 785; https://doi.org/10.3390/catal12070785
Submission received: 25 June 2022 / Revised: 12 July 2022 / Accepted: 14 July 2022 / Published: 18 July 2022
(This article belongs to the Special Issue New Trends in Photocatalytic Materials for Efficient Performance)

Abstract

:
To date, (001)TiO2/Ti3C2Tx hybridized photocatalyst is usually prepared through the complicated treatment of Ti3AlC2 in the presence of corrosive fluorine with a molar ratio of nF:nTi of more than 20. To reduce the use of corrosive fluorine, herein, exploiting beyond the conventional method, we report a facile synthetic method for (001)TiO2/Ti3C2Tx, elaborately using HF as both an etchant for Al elimination and a morphology control agent for the growth of (001)TiO2 nanosheets, with a sharply diminished use of fluorine (nF:nTi = 4:1) and simplified operation procedures. After optimization, the resulting (001)TiO2/Ti3C2Tx heterojunction exhibited markedly high photocatalytic activity with the CO2 reduction rate of 13.45 μmol g−1 h−1, which even surpasses that of P25 (10.95 μmol g−1 h−1), while the photoelectron selectivity to CH4 is approaching 92.84%. The superior photoactivity is interpreted as the fact that Ti3C2Tx with a lower work function induces photoinduced hole transfer and suppresses the charge recombination, thus facilitating the CO2 multi-electron reduction. This study provides a novel and simple synthesis for (001)TiO2/Ti3C2Tx towards sustainable energy transformations.

Graphical Abstract

1. Introduction

TiO2 is an extremely attractive candidate for photocatalysis, given the multiple advantages of low price, chemical durability, and ecological friendliness [1,2]. However, realistic applications of TiO2 are restricted due to the rapid recombination of photogenerated electron–hole pairs. Therefore, numerous methods have been developed with the focus on the promotion of photoinduced charge transfer, such as loading co-catalyst [3,4], morphology modulation [5], and the construction of heterojunction [6,7,8]. Thereinto, loading co-catalyst on TiO2 has hitherto been a satisfactory choice for improving its photoactivity and stability by accelerating the separation efficiency of photogenerated carriers. Nevertheless, the widespread utilization of noble-metal cocatalysts, such as Au and Pt, remains challenging for both high cost and scarcity. Hence, there is a considerable incentive to exploit cost-effective noble-metal-free cocatalysts for the enhancement of the TiO2 photocatalytic performance.
MXenes, a new family of two-dimensional layered transition metal carbides, nitrides, or carbonitrides, possess a structural formula of Mn+1XnTx, where M refers to the early transition metal, X stands for carbon and/or nitrogen, T represents the surface functional groups (-O, -OH, -F), and n ranges from 1 to 3. Along with a mushrooming in the study of MXenes, up to now, more than 40 MXenes have been synthesized by selectively etching the metal layer from the precursor MAX phases [9]. Moreover, various syntheses for different-dimension MXenes (e.g., quantum nanodots, nanorods, and nanosheets) have been reported [10]. Among all the MXenes, Ti3C2 is one of the most representative and prevalently studied materials in the field of photocatalysis, especially as a co-catalyst, because of the following reasons: (i) Its intrinsically metallic conductivity assures the efficient separation and transfer of photogenerated charge carriers from semiconductors [11,12,13,14]. (ii) The appropriate work function allows it to construct Schottky heterojunction with most semiconductors to alter their photoelectric properties [15,16]. As reported in previous work [17], Ti3C2Tx can be coupled with (Ti, C) co-doped BiOBr to manipulate the carrier flow direction and achieve specific redox reactions via the elaborate Schottky heterojunction. (iii) The enormous surface functional groups favor the enhanced reactant adsorption capability [18].
In recent years, MXene-based materials receive considerable attention in photo/electrocatalysis. Amrillah et al. provided an insightful discussion on a series of syntheses for MXene-based and MXene-derived catalysts, including alcoholysis, hydrothermal treatment, and calcination process [10]. Typically, Ti3C2 is an ideal platform for the in situ growth of TiO2 to prepare TiO2-Ti3C2 composite materials by partial oxidation. TiO2-Ti3C2 heterojunctions prepared by such in situ methods have intimate interfacial contact, which is conducive to the interface transfer of photoinduced electrons. To optimize the photocatalytic performance of TiO2/Ti3C2 catalysts, morphology engineering of TiO2 (originated from partial oxidized Ti3C2) embedded in Ti3C2 has aroused the interest of material chemists [19]. In this way, plenty of TiO2 nanostructures, including nanoparticles, nanoflowers, nanocrystals, and nanosheets, have been derived from Ti3C2 MXene [16,20,21,22]. Among them, (001)TiO2 nanosheets stand out because they offer an excellent photoactivity that renders the TiO2/Ti3C2 heterojunction system attractive for many applications (photocatalytic environmental purification [16,22], photoelectrochemical detection [23], artificial photosynthesis of chemical fuels [24]).
To date, most of the synthesis methods for (001)TiO2/Ti3C2 have been carried out using at least two steps: (i) the raw material Ti3AlC2 was etched for Al layer elimination with HF, followed by washing to remove HF, and the Ti3C2 was finally obtained by drying; (ii) afterward, an additional fluorine source (e.g., NaBF) is added to regulate the in situ growth of (001)TiO2 nanosheets on Ti3C2 by hydrothermal process [16,22]. For step (i), theoretically, to etch the Al layer of 1 mol Ti3AlC2, 3 mol HF is needed according to the reaction equation [25]: Ti3AlC2 + 3HF → Ti3C2 + AlF3 + 3/2H2. However, in those cases, the molar ratio of HF to Ti3AlC2 (nHF:nTi3AlC2) is normally larger than 50:1, which is much more than the actual required amount. In fact, according to the study of Alhabeb et al., fewer HF (nHF:nTi3AlC2 = 10:1) is enough to remove the Al layer of Ti3AlC2 [26]. It is well known that the large use of harmful HF is not environmentally friendly, thereby reducing the use of HF, meanwhile ensuring the elimination of the Al layer is rather significant. For step (ii), NaBF4 is used as the shape-directing agent of (001)TiO2 nanosheets. Additionally, its fluorine ion (F) plays an essential role in preparing (001)TiO2 nanosheets, since F can combine with Ti atom to reduce the surface energy of (001) facet, resulting in the preferential growth of high-energy (001) facets during the growth process of TiO2 crystal [27]. While many efforts have been devoted to fabricating (001)TiO2/Ti3C2, some of the key challenges are still present that need to be addressed, especially the large usage of corrosive HF and the complexity of the operation.
Herein, a facile one-step hydrothermal method of (001)TiO2/Ti3C2Tx composite was developed, as schematically illustrated in Figure 1. In this case, HF was delicately used as both an etchant for Al layer elimination of Ti3AlC2 and a morphology control agent providing F for the controlled growth of (001)TiO2 nanosheets. In sharp contrast to traditional methods, we let the HF used in the etching process continue to participate in the hydrothermal reaction, rather than washing it off, so as to improve the utilization rate of HF and reduce its discharge. As a result, (i) the usage of HF is dramatically reduced; the molar ratio of F to Ti in the pristine materials of HF and Ti3AlC2 (nF:nTi) decreased from 21:1 to 4:1 as compared to the traditional synthesis method [16]. (ii) This work also simplifies the operation procedures, avoiding a series of processes, including washing, drying, and the addition of an extra morphology control agent. For detailed distinction, the experiment conditions and molar ratio (nF:nTi) for synthesizing (001)TiO2/Ti3C2Tx are summarized in Table S1. The optimized (001)TiO2/Ti3C2Tx exhibited superior photocatalytic CO2 reduction activity, which was better than that of commercial P25 TiO2. It also showed remarkable selectivity for CH4 production. The reason for the extraordinary photocatalytic performances may be the fact that Ti3C2Tx with a lower work function induces the photogenerated hole transfer and suppresses the charge recombination, facilitating the CO2 multi-electron reduction.

2. Results and Discussion

The phase structure of the sample was analyzed using X-ray diffraction (XRD) spectroscopy. As shown in Figure 2a, the diffraction peak of Ti3AlC2 MAX is consistent with that reported in the literature [12]. In comparison to Ti3AlC2, the HF etching process causes the (104) diffraction peak at 39° to disappear in Ti3C2Tx, as well as the (002) and (004) diffraction peaks to shift towards lower angles, which are indications of Al layer elimination and increased layer spacing in the Ti3C2Tx structure [28], as likewise evidenced by the observation of SEM images for both Ti3AlC2 and Ti3C2Tx (Figure S2). For conciseness, the TiO2/Ti3C2Tx samples prepared by hydrothermal method are hereafter named TT-x, where x refers to the added volume (unit: mL) of the HF. All of the TiO2/Ti3C2Tx samples are well indexed to the peaks of both Ti3C2Tx and anatase phase TiO2 (JCPDS No. 21-1272), indicating that TiO2/Ti3C2Tx heterojunctions were successfully prepared (Figure 2a). However, there remains a faint (104) diffraction peak of Al at 39° for the TT-2 sample, which is mainly since the amount of HF added during the preparation was too small to etch the Al layer of Ti3AlC2 completely. When the amount of HF added was greater than 3 mL, the peak of Al fully disappeared in the corresponding position of the sample. Notably, the intensity of the diffraction peaks of TiO2 decreased with the increasing HF addition higher than 5mL. This is mainly because the excessive amount of HF will further etch the TiO2 nanosheets and make them less crystalline. Among them, the crystallinity of TiO2 in the TiO2/Ti3C2Tx composites prepared by the HF addition in the range of 3 mL to 5 mL was better.
In addition to XRD, Raman spectra were used to reveal the constituents of the as-prepared samples, yielding similar results. As shown in Figure 2b, Ti3AlC2 contains several characteristic bands that appeared at ca. 169, 268, 411, and 598 cm−1, which are associated with the ω1, ω2, ω3, and ω4 Raman-active phonon vibration modes, respectively, while the two peaks between 1000 and 1800 cm−1 (D and G band) are referred to the stretching vibrations of graphitic carbon [12,22]. For the other two samples, the Ti3C2Tx sample showed the expected Raman vibrational modes [15,29]. Compared to the Ti3C2Tx sample, the TT-5 sample displayed an ascending Raman signal at 150 cm−1 (Eg), together with three other peaks at 393 cm−1 (B1g), 507 cm−1 (A1g), 630 cm−1 (Eg), which are originated from anatase (001)TiO2 [12,30]. Besides, the D- and G-modes peaks of Ti3C2Tx were also observed clearly in TT-5. Reasonably, both XRD and Raman results confirm the transformation of Ti3C2Tx to TiO2 during hydrothermal treatment.
Another direct piece of evidence for the successful preparation of TiO2/Ti3C2Tx is provided by the morphology and microstructure observation via field emission scanning electron microscopy (FESEM) and high-angle annular dark field scanning transmission electron microscope (HAADF-STEM). The most striking feature of the typical heterojunction is the existence of TiO2 nanosheets embedded in Ti3C2Tx that are deliberately created using a fluorine source (HF) under the hydrothermal process. This characteristic is directly observed with FESEM (Figure 2c), showing square TiO2 nanosheets with a thickness of 130 nm embedded in the layered Ti3C2Tx. The TEM image (Figure 2d) of TiO2/Ti3C2Tx displays several square thin nanosheets on a substrate, and its partially enlarged image (Figure 2e) shows the obvious lattice spacing of 0.235 nm, which is identified for the TiO2 (001) plane. Usually, the growth rate of crystals is exponentially related to the surface energy [31]. During the growth of TiO2 nuclei, the (001) planes with high surface energy (0.90 J/m2) decrease rapidly, and the TiO2 nanocrystals evolve spontaneously into a truncated bipyramidal structure dominated by stable (101) facets because of their low surface energy (0.44 J/m2) [32]. It is worth noting that modifying the surface adsorption species of the crystal can change the surface energy, exposing specific highly active facets of the crystal. F is a species with the affinity for Ti atoms so that it can bind Ti atoms and reduce the surface energy of the (001) facets, making them more stable than the (101) facets during the anatase TiO2 crystal growth process [27,33]. Therefore, in our case, the absorbed F atom functioned as a capping agent to modify the surface energy of (001) facets of TiO2, resulting in exceptional stabilization. Furthermore, the elemental distribution in the composite was visualized by STEM-energy-dispersive X-ray spectroscopy (EDS) mapping analysis (Figure 2f). By focusing on an area containing square sheets with lying and standing positions, it is clearly revealed that the distributions of Ti atoms and O atoms are perfectly coincided with the square region, verifying the formation of TiO2 nanosheet structure. Therefore, it can be concluded that the (001)TiO2/Ti3C2Tx has been successfully prepared.
As evidenced above, after HF pretreatment and hydrothermal process for precursor Ti3AlC2, the Al layer has been removed and the TiO2/Ti3C2Tx heterojunction was formed. To further confirm the changes in the surface composition of the prepared samples Ti3AlC2, TT-5, and TT-5-u (the sample pretreated with 5 mL HF but without hydrothermal treatment) at different stages in synthesis, XPS measurements were carried out. As a consequence of the Al layer removal, Al 2s and 2p peaks in Ti3AlC2 disappeared in TT-5 and TT-5-u, which can be observed via the XPS survey spectra (Figure S3). The comparison of the high-resolution Ti 2p spectrum between TT-5 and TT-5-u is shown in Figure 3a,d. For TT-5, the Ti 2p spectrum can be fitted by seven components located at 454.8, 456.3, 456.5, 459.4, 461.3, 462.6, and 465.0 eV, which are attributed to Ti-C (454.8 eV and 461.3 eV), Ti-X (456.3 eV) (substoichiometric TiCx (x < 1) or titanium oxycarbides), Ti3+ (456.5 eV and 462.6 eV), and TiO2 (459.4 eV and 465.0 eV) [12,16,22]. In comparison to TT-5-u, the TT-5 sample shows two more peaks of TiO2 (459.4 eV and 465.0 eV) and the fade of the other Ti species, demonstrating the gradual transformation from Ti3C2Tx to TiO2. Furthermore, the C 1s spectrum (Figure 3b) of TT-5 exhibited several distinct peaks at 288.8, 286.4, 284.8, 282.5, and 281.9 eV, corresponding to C-F, C-O, C-C, C-Ti-O, and Ti-C, respectively [16,22,34]. Apparently, C-F is derived from HF; Ti-C and C-C originate from Ti3C2Tx and adventitious carbon, respectively. It is noteworthy that the TT-5 showed a new C-Ti-O peak at 282.5 eV as compared with TT-5-u (Figure 3e), demonstrating that the TiO2/Ti3C2Tx heterojunction is linked by a C-Ti-O bond. The O 1s spectrum of TT-5-u is fitted by two components located at 532.18 and 530.0 eV (Figure 3f), which are corresponding to Ti-OH, Ti-O [35]. After hydrothermal treatment, the Ti-O intensity of TT-5 significantly increased as compared to Ti-OH (Figure 3c), which suggests the existence of TiO2. The XPS results verify the successful preparation of the TiO2/Ti3C2Tx heterojunction, which is well consistent with the results of XRD, Raman, and EDS analyses.
The photocatalytic activity of the samples was evaluated by photocatalytic CO2 reduction in a gas-solid reaction system under simulated solar light. Mild conditions were kept without using any photosensitizer or organic sacrificial reagents. Methane (CH4) and carbon monoxide (CO) are generated as the predominant products via the two-electron and eight-electron reduction processes, respectively (Figure 4a). Ti3C2Tx with metallic properties is generally not photocatalytically active [28]. However, a weak CO2 photoreduction activity of Ti3C2Tx was detected here, resulting from the TiO2 (partial oxidation Ti3C2Tx) during storage or light exposure. As for the other samples, the photoactivity showed a trend to increase and then decline as the increasing addition of HF. From TT-2 to TT-5, the photoactivity gradually enhanced due to the increase in (001)TiO2 nanosheets on Ti3C2Tx, which is consistent with the trend seen in the XRD and SEM results (Figure 2a and Figure S4). Besides, for nitrogen adsorption-desorption isotherms and corresponding pore-size distribution analysis (Figure S5 and Table S2), TT-5 shows larger surface area and more micropores (<2 nm) as compared to TT-2, which may favor the adsorption of CO2 gas molecules. However, following adding excessive HF, the TiO2 may further be etched (as shown in Figure S4), resulting in a decrease in the photoactivity for TT-6 and TT-8. The results suggest that there is an optimum amount (5 mL) of HF in the synthesis of TiO2/Ti3C2Tx for achieving high photocatalytic CO2 reduction rates. Among all the samples, TT-5 exhibited the best CO2 photoreduction rate of 13.45 μmol g−1 h−1 (10.28 μmol g−1 h−1 for CH4 and 3.17 μmol g−1 h−1 for CO), which was higher than that of P25 TiO2 (10.95 μmol g−1 h−1). Note that methane was not detected for commercial P25, while TT-5 was up to 92.84% methanogenic selectivity (Figure 4b). As reported previously, the existence of Ti3C2 favors the selectivity for photocatalytic CO2–CH4 conversion [36,37]. Combining the reported studies and above experimental results, the reason for the extraordinary selectivity for CH4 production could be interpreted as Ti3C2 with lower work function induces the photogenerated hole transfer and suppresses the charge recombination, facilitating the CO2 multi-electron reduction.
A series of control experiments were conducted to reveal the source of reduction products; the results are shown in Figure S6. The product concentration is below the limit of detection when the experiments were carried out in the absence of catalyst or irradiation or in Ar, proving that the detectable CO and CH4 were derived from photocatalytic CO2 reduction over (001)TiO2/Ti3C2Tx. To further disclose the origin of the products, an isotopically labeled 13CO2 experiment was performed. As shown in Figure 4c,d, the strong signals were observed at m/z = 17 and = 29, which were attributed to 13CH4 and 13CO, respectively. This result further confirms that the reduction products indeed originate from the CO2 photoreduction rather than from the impurity or the sample.
With efforts to acquire an in-depth understanding of the reaction intermediates in CO2 photoreduction for TT-5, in situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) was performed (Figure 5). Obviously, various peaks can be observed during the photocatalytic reaction that corresponds to the formation of carbonate species, in which CO32− and HCO3 are generated from CO2 and -OH [38]. The peaks at 1868, 1844, 1828, 1792, 1698, 1684, 1576, and 1558 cm−1 were attributed to HCOO [12,28,39], while the band for COOH* was located at 1569 cm−1 [40]; they might serve as possible intermediates for CO production. Furthermore, the peaks at 1772, 1750, 1716, 1457, and 1419 cm−1 belong to HCHO [12,28,41], and the peak at 1734 cm−1 originates from CH3O [41], both of which are important intermediate species for the production of CH4. Notably, the signal of CH4 was undetectable on DRIFT spectra, as it is a non-polar molecule. Therefore, the CO2 photoconversion process could be deduced as follows: CO2 molecules are firstly adsorbed on the surface of the TT-5 sample and reacted with H+ dissociated from H2O to form COOH (CO2* + e + H+ → COOH*). Then, COOH* transformed to CO* by a protonation process (COOH* + e + H+ → CO* + H2O) [40,42]. One part of the adsorbed CO* is released to form gaseous CO (CO* → CO); more importantly, the other part of CO* continues to participate in the reduction reaction to form HCHO* and CH3O* (CO* + 2e + 2H+ → HCHO*; HCHO* + e + H+ → CH3O*). Eventually, the CH4 production could be achieved by the further protonation of CH3O* (CH3O* + 3e + 3H+ → CH4↑+ H2O) [43].
Furthermore, considering that the separation and migration of photoinduced electron-hole pairs play an important role in achieving the enhancement of photocatalytic efficiency, the transient photocurrent response test was performed. As shown in Figure 6a, the anode photocurrent increases during irradiation and declines without irradiation. It can clearly be seen that TT-5 exhibited improved photocurrent responses compared with other samples, suggesting the efficient separation of the charge carriers in the composites of TT-5. Notably, the rapid decay characteristic of P25 could be assigned to fast electron–hole recombination kinetics due to the absence of Ti3C2Tx with high conductivity [12,44]. For the others, their photocurrent is relatively stable, because the presence of Ti3C2Tx slows down the recombination rate of photogenerated carriers, which is beneficial for the multi-electron reduction of CO2 (CO2 + 8H+ + 8e → CH4↑ + 2H2O).
In an effort to better understand the pathway of interfacial charge transfer between (001)TiO2 nanosheets (synthesized according to our previous report [45]) and Ti3C2Tx, the work functions (Wsample) of them were studied via Kelvin probe using a gold mesh as the reference. Specifically, the Wsample is calculated by Equation (1) [46]:
Wsample = Wprobe + e • CPDsample
where CPDsample is the contact potential difference (CPD) of the given sample, which was measured by the Kelvin probe; Wprobe represents the work function of the gold mesh (≈4.25 eV); and e is the electronic charge. Based on the probed CPD results (Figure 6b), the W of (001)TiO2 nanosheets and Ti3C2Tx are calculated to be 4.66 and 4.42 eV, respectively (Figure 6c). Accordingly, the Fermi levels (Ef) of the samples are determined via the conversion formulas (Equations (2)–(4)) [28], as shown below:
Ef,vac = Evac − Wsample
Ef(vs.SHE, pH0) = −4.5 V − Ef,vac
Ef(vs.SHE, pH7) = Ef(vs.SHE, pH0) − 0.059 pH
where the Ef,vac represents the Fermi levels of the samples at the vacuum level, and Evac is the energy of a stationary electron at the vacuum level (regarded as 0 eV). Thus, the Ef of Ti3C2Tx and TiO2 are calculated to be −0.49 and −0.25 eV (vs. SHE at pH = 7), respectively. Obviously, the Fermi energy level of Ti3C2Tx is more negative than that of (001)TiO2 nanosheets (i.e., the work function of Ti3C2Tx is lower than that of (001)TiO2), and thus the photogenerated holes rather than electrons of TiO2 nanosheets are trapped by Ti3C2Tx.
Based on the above experimental results, we inferred the mechanism of (001) TiO2/Ti3C2Tx for photocatalytic CO2 reduction (Figure 6d). Under illumination, the electrons on the valence band of (001)TiO2 nanosheets are excited to the conduction band, and these photoelectrons gathered on the conduction band react with the adsorbed CO2 to generate CH4 and CO via the photocatalytic process. Because of the differences in work function for (001)TiO2 nanosheets and Ti3C2Tx with metallic properties, a Schottky barrier could be formed at the interface of (001)TiO2 and Ti3C2Tx [16,22], which would promote the holes rather than electrons transfer to Ti3C2Tx, while enabling the accumulation of electrons on TiO2. Undoubtedly, this is favorable for the efficiently multi-electron CO2 photoreaction.

3. Experimental Procedures

3.1. Chemicals

All the chemicals were of analytic grade and used as received. Deionized (DI) water used in this research was from local sources. Ti3AlC2 power was supplied by 11 Technology Co., Ltd. (Changchun, China). Hydrofluoric acid (HF), sulfuric acid (H2SO4), and sodium bicarbonate (NaHCO3) were obtained from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China).

3.2. Sample Preparation

Preparation of Ti3C2Tx: Ti3C2Tx was prepared by an HF-etching method. In detail, 1 g Ti3AlC2 powder was slowly put into the aqueous HF solution (40 wt%) under magnetic stirring for 24 h to eliminate the Al layer. Then, the black suspension was filtered and washed with distilled water until the pH of the filtrate was about 7. After drying at 60 ℃ under vacuum for 12 h, the Ti3C2Tx powder was obtained.
Preparation of (001) TiO2/Ti3C2Tx: (001) TiO2/Ti3C2Tx composite photocatalysts were synthesized by a one-step hydrothermal method. Typically, 2 g Ti3AlC2 powder was slowly added to HF solution (5 mL, 40 wt%) and kept agitating at room temperature for 24 h in order to remove the Al layer. Then, distilled water was directly added to the above suspensions to a total volume of 120 mL. After sonicating for 10 min, the homogeneous mixture was sealed into a 200 mL Teflon-lined autoclave with a stainless steel jacket for a hydrothermal process at 180 °C for 12 h, in which HF was delicately used as both an etchant for Ti3AlC2 and a morphology control agent for the growth of (001)TiO2 nanosheets on Ti3C2Tx. After natural cooling, the precipitate was collected by filtration, washed thoroughly with distilled water until the pH of the filtrate turned to 7, and dried at 60 °C. Subsequently, the (001)TiO2/Ti3C2Tx was obtained and labeled as TT-5 (5 mL HF) for short.
Similarly, a serial of composites was synthesized by varying the usage of HF under other identical conditions, and the acquired samples were named TT-x (x = 2, 3, 4, 5, 6, and 8), where x represents the usage (unit: mL) of HF. To study the transformation process of Ti3AlC2 to (001)TiO2/Ti3C2Tx, samples pretreated with 5 mL HF but without hydrothermal process were collected and labeled TT-5-u.

3.3. Characterization

The X-ray diffraction (XRD) patterns of the samples were recorded on an X-ray diffractometer (D8 ADVANCE Bruker, Karlsruhe, Germany) with Cu Kα radiation at a scan rate of 0.02° s−1. Raman spectra were collected through a Raman spectroscopy (DXR2 XI, Thermo Fisher, Waltham, MA, USA) with an excitation wavelength of 633 nm. The morphology and microstructure of the samples were analyzed using a field emission scanning electron microscope (FESEM, SU8010 Hitachi, Tokyo, Japan) and high-angle annular dark-field scanning/transmission electron microscopy (HAADF-STEM, Talos F200S, Thermo Scientific, Waltham, MA, USA), respectively. X-ray photoelectron spectroscopy (XPS) measurements were performed using a VG Multilab 2000 XPS system (Thermo VG Scientific, London, UK) to obtain information related to the surface chemical state and composition of the sample. All the spectra were calibrated to the C 1 s peak at 284.8 eV. The UV–vis diffuse reflectance spectra (DRS) were obtained with a UV–vis spectrophotometer (UV2600, Shimadzu, Kyoto, Japan) using BaSO4 as the reference standard. The Brunauer–Emmett–Teller (BET) surface area and pore size distribution data of the samples were analyzed using a nitrogen adsorption apparatus (ASAP-2020, Micromeritics, Norcross, GA, USA). The samples were degassed at 175 °C for 4 h before N2 adsorption measurement. A Kelvin probe apparatus (Instytut Fotonowy, Cracow, Poland) with a sensitivity of 1 mV was utilized to determine the contact potential difference (CPD) of the samples with a gold mesh Kelvin probe as the reference. Surface work function (W) was monitored using an LED lamp as a light source in conjunction with a monochromator. In situ FTIR spectra were measured on a Bruker Tensor II FTIR spectrometer (Bruker, Karlsruhe, Germany).

3.4. Electrochemical Measurements

Electrochemical characterization for transient photocurrent responses (TPR) was conducted on an electrochemical analyzer (CHI 760e, CH Instruments, Shanghai, China). Pt wire, Ag/AgCl (saturated KCl), and 0.4 M Na2SO4 solution were functioned as the counter electrode, reference electrode, and electrolyte, respectively. For the working electrode, 30 mg of catalyst was ground in 1 mL water/absolute ethanol mixed solvent (v/v = 1/1) and 30 μL of Nafion solution to make a slurry, which was coated onto indium tin oxide (ITO) conductive glass with an exposed area of 1 cm2 after ultrasonic dispersion. A 365 nm LED (3 W, Shenzhen Lamplic, Shenzhen, China) was used as the light source.

3.5. Photocatalytic CO2 Reduction and Isotope-Labelling Measurement

The photocatalytic CO2 reduction of the samples was conducted in a glass automatic online gas analysis system (Labsolar-6A, Beijing, China) (Figure S1a). The light source was a 300 W xenon lamp (PLS-SXE300+, Beijing, China). The photocatalytic activity test was carried out in three main steps: coating catalyst, in situ CO2 production, and the occurrence of the CO2 photocatalytic reduction. (1) Coating catalyst: 30 mg of the photocatalyst was dispersed in 15 mL distilled water to form a suspension, which was then transferred to a petri dish (ɸ = 60 mm). After drying, the catalyst film was deposited on the bottom of the dishes. (2) In situ CO2 generation: 2.1 g of NaHCO3 powder was placed on the bottom of a homemade double-neck reactor, and the coated dish in (1) was supported in the reactor using a quartz tripod (Figure S1b). Before illumination, the photocatalytic system was vacuumed to exclude the air. Finally, 7 mL of H2SO4 (2 M) was injected into the reactor to react with NaHCO3 for generating CO2 and H2O vapor within the reactor. (3) CO2 photoreduction (Figure S1c): After adsorption equilibrium, a 300 W xenon lamp (PLS-SXE300+, Perfect Light) was used as the light source to initiate the photocatalytic reaction. The concentration of the CO2 reduction products was analyzed by gas chromatography (GC-2014, Shimazu, Kyoto, Japan) with FID and TCD detectors. For determining the carbon source of the reduction products, 13CO2 isotope labeling experiment was carried out under the same conditions as the above-mentioned CO2 reduction test, except for the substitute of NaH12CO3 using NaH13CO3 (SHANGHAI ZZBIO Co., Ltd., Shanghai, China). The gas products were verified by a mass spectrometer (MS, HPR-20, Hiden Analytical, Warrington, UK). The electron selectivity for CH4 production during the photocatalytic CO2 reduction process is evaluated using the following equation [47]: CH4 selectivity (%) = [(8n(CH4)]/[(8n(CH4) + 2n(CO)] × 100, CO selectivity (%) = [(2n(CO)]/[(8n(CH4) + 2n(CO)] × 100, where n(CH4) and n(CO) denote the number of moles of CH4 and CO, respectively.

4. Conclusions

In summary, a facile one-step hydrothermal approach of (001)TiO2/Ti3C2Tx composite was developed, where HF was delicately used as both an etchant for Ti3AlC2 and a morphology control agent for TiO2 nanosheets. This method simultaneously addressed the issues of the large usage of HF and complex operation procedure. Furthermore, by tailoring the HF amount within a narrow range, the morphology of TiO2 nanosheets could be purposively manipulated. The optimized sample (TT-5) shows a superior photoactivity with a CO2 reduction rate of 13.45 μmol g−1 h−1, as well as a CH4 production selectivity of 92.84%. The excellent photocatalytic performance is mainly since purposeful spatial isolation of photogenerated charge carriers induced by Ti3C2Tx dramatically improves the electron density of TiO2, thereby facilitating the CO2 multi-electron reduction of (001)TiO2/Ti3C2Tx heterojunction. This work may provide an easy and novel approach to fabricating highly efficient MXene-derived composites for sustainable energy conversion.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal12070785/s1, Figure S1: Photocatalytic CO2 reduction device (a), in situ CO2 production (b), CO2 photoreduction reaction (c); Figure S2: SEM image of Ti3AlC2 (a) and Ti3C2 (b), showing that the morphology of bulk Ti3AlC2 changed to accordion-like Ti3C2Tx, as a consequence of the eliminated Al layer via HF etching; Figure S3: X-ray photoelectron spectroscopy (XPS) survey spectra (a) and high-resolution XPS spectra of Al 2p (b) of Ti3AlC2, TT-5-u and TT5; Figure S4: SEM images of TiO2/Ti3C2Tx samples prepared with different usage of HF; Figure S5: Nitrogen adsorption/desorption isotherms and the corresponding pore-size distribution curves (inset) of the TT-2, TT-5 and TT-8 samples; Figure S6: The control experiments of photocatalytic CO2 reduction performance over TT-5 under altered conditions; Table S1: Synthesis conditions of (001)TiO2/Ti3C2Tx composites; Table S2: BET surface area and the corresponding pore volume and pore size of samples [48,49,50,51].

Author Contributions

Conceptualization, K.L. (Kangle Lv); methodology, K.L. (Kangle Lv); formal analysis, J.L., K.L. (Kaining Li), Q.T., Q.L., C.W. and K.L. (Kangle Lv); investigation, J.L. and K.L. (Kaining Li); resources, K.L. (Kangle Lv); data curation, J.L. and K.L. (Kaining Li); writing—original draft preparation, J.L. and K.L. (Kaining Li); Visualization, C.W.; writing—review and editing, Q.L., J.F. and K.L. (Kangle Lv); supervision, project administration and funding acquisition, K.L. (Kangle Lv). All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (51672312, 21972171 and 52073263) and the Fundamental Research Funds for the Central Universities, South-Central Minzu University (CZP22001).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Li, W.; Elzatahry, A.; Aldhayan, D.; Zhao, D. Core-Shell Structured Titanium Dioxide Nanomaterials for Solar Energy Utilization. Chem. Soc. Rev. 2018, 47, 8203–8237. [Google Scholar] [CrossRef] [PubMed]
  2. Gong, H.; Wang, L.; Zhou, K.; Zhang, D.; Zhang, Y.; Adamaki, V.; Sergejevs, A.; Bowen, C. Improved Photocatalytic Performance of Gradient Reduced TiO2 Ceramics with Aligned Pore Channels. Adv. Powder Mater. 2022, 1, 100025. [Google Scholar] [CrossRef]
  3. Xu, F.; Meng, K.; Zhu, B.; Liu, H.; Xu, J.; Yu, J. Graphdiyne: A New Photocatalytic CO2 Reduction Cocatalyst. Adv. Funct. Mater. 2019, 29, 1904256. [Google Scholar] [CrossRef]
  4. Meng, A.; Zhang, L.; Cheng, B.; Yu, J. Dual Cocatalysts in TiO2 Photocatalysis. Adv. Mater. 2019, 31, 1807660. [Google Scholar] [CrossRef] [PubMed]
  5. Sajan, C.P.; Wageh, S.; Al-Ghamdi, A.A.; Yu, J.; Cao, S. TiO2 Nanosheets with Exposed {001} Facets for Photocatalytic Applications. Nano Res. 2016, 9, 3–27. [Google Scholar] [CrossRef]
  6. Xu, F.; Meng, K.; Cheng, B.; Wang, S.; Xu, J.; Yu, J. Unique S-Scheme Heterojunctions in Self-Assembled TiO2/CsPbBr3 Hybrids for CO2 Photoreduction. Nat. Commun. 2020, 11, 4613. [Google Scholar] [CrossRef]
  7. Low, J.; Dai, B.; Tong, T.; Jiang, C.; Yu, J. In Situ Irradiated X-ray Photoelectron Spectroscopy Investigation on a Direct Z-scheme TiO2/CdS Composite Film Photocatalyst. Adv. Mater. 2019, 31, 1802981. [Google Scholar] [CrossRef] [PubMed]
  8. Zhang, Z.; Zhang, Y.; Han, X.; Guo, L.; Wang, D.; Lv, K. Assembly of CaIn2S4 on Defect-Rich BiOCl for Acceleration of Interfacial Charge Separation and Photocatalytic Phenol Degradation via S-Scheme Electron Transfer Mechanism. Catalysts 2021, 11, 1130. [Google Scholar] [CrossRef]
  9. Naguib, M.; Barsoum, M.W.; Gogotsi, Y. Ten Years of Progress in the Synthesis and Development of MXenes. Adv. Mater. 2021, 33, 2103393. [Google Scholar] [CrossRef]
  10. Amrillah, T.; Hermawan, A.; Alviani, V.N.; Seh, Z.W.; Yin, S. MXenes and Their Derivatives as Nitrogen Reduction Reaction Catalysts: Recent Progress and Perspectives. Mater. Today Energy 2021, 22, 100864. [Google Scholar] [CrossRef]
  11. Ran, J.; Gao, G.; Li, F.T.; Ma, T.Y.; Du, A.; Qiao, S.Z. Ti3C2 MXene Co-Catalyst on Metal Sulfide Photo-Absorbers for Enhanced Visible-Light Photocatalytic Hydrogen Production. Nat. Commun. 2017, 8, 13907. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Low, J.; Zhang, L.; Tong, T.; Shen, B.; Yu, J. TiO2/MXene Ti3C2 Composite with Excellent Photocatalytic CO2 Reduction Activity. J. Catal. 2018, 361, 255–266. [Google Scholar] [CrossRef]
  13. Chen, L.; Ye, X.; Chen, S.; Ma, L.; Wang, Z.; Wang, Q.; Hua, N.; Xiao, X.; Cai, S.; Liu, X. Ti3C2 MXene Nanosheet/TiO2 Composites for Efficient Visible Light Photocatalytic Activity. Ceram. Int. 2020, 46, 25895–25904. [Google Scholar] [CrossRef]
  14. Li, W.; Zhuang, C.; Li, Y.; Gao, C.; Jiang, W.; Sun, Z.; Qi, K. Anchoring Ultra-Small TiO2 Quantum Dots onto Ultra-Thin and Large-Sized Mxene Nanosheets for Highly Efficient Photocatalytic Water Splitting. Ceram. Int. 2021, 47, 21769–21776. [Google Scholar] [CrossRef]
  15. Li, Z.; Huang, W.; Liu, J.; Lv, K.; Li, Q. Embedding CdS@Au into Ultrathin Ti3-XC2Ty to Build Dual Schottky Barriers for Photocatalytic H2 production. ACS Catal. 2021, 11, 8510–8520. [Google Scholar] [CrossRef]
  16. Peng, C.; Yang, X.; Li, Y.; Yu, H.; Wang, H.; Peng, F. Hybrids of Two-Dimensional Ti3C2 and TiO2 Exposing {001} Facets toward Enhanced Photocatalytic Activity. ACS Appl. Mater. Interfaces 2016, 8, 6051–6060. [Google Scholar] [CrossRef]
  17. Hermawan, A.; Hasegawa, T.; Asakura, Y.; Yin, S. Enhanced Visible-Light-Induced Photocatalytic NOx Degradation over (Ti,C)-BiOBr/Ti3C2Tx MXene Nanocomposites: Role of Ti and C Doping. Sep. Purif. Technol. 2021, 270, 118815. [Google Scholar] [CrossRef]
  18. Cao, S.; Shen, B.; Tong, T.; Fu, J.; Yu, J. 2D/2D Heterojunction of Ultrathin MXene/Bi2WO6 Nanosheets for Improved Photocatalytic CO2 Reduction. Adv. Funct. Mater. 2018, 28, 1800136. [Google Scholar] [CrossRef]
  19. Sreedhar, A.; Noh, J.S. Recent Advances in Partially and Completely Derived 2D Ti3C2 MXene Based TiO2 Nanocomposites towards Photocatalytic Applications. Sol. Energy 2021, 222, 48–73. [Google Scholar] [CrossRef]
  20. Li, Y.; Deng, X.; Tian, J.; Liang, Z.; Cui, H. Ti3C2 MXene-Derived Ti3C2/TiO2 Nanoflowers for Noble-Metal-Free Photocatalytic Overall Water Splitting. Appl. Mater. Today 2018, 13, 217–227. [Google Scholar] [CrossRef]
  21. Li, Y.; Zhang, D.; Feng, X.; Liao, Y.; Wen, Q.; Xiang, Q. Truncated Octahedral Bipyramidal TiO2/MXene Ti3C2 Hybrids with Enhanced Photocatalytic H2 Production Activity. Nanoscale Adv. 2019, 1, 1812–1818. [Google Scholar] [CrossRef] [Green Version]
  22. Shahzad, A.; Rasool, K.; Nawaz, M.; Miran, W.; Jang, J.; Moztahida, M.; Mahmoud, K.A.; Lee, D.S. Heterostructural TiO2/Ti3C2Tx (MXene) for Photocatalytic Degradation of Antiepileptic Drug Carbamazepine. Chem. Eng. J. 2018, 349, 748–755. [Google Scholar] [CrossRef]
  23. Chen, Y.; Li, X.; Cai, G.; Li, M.; Tang, D. In Situ Formation of (0 0 1) TiO2/Ti3C2 Heterojunctions for Enhanced Photoelectrochemical Detection of Dopamine. Electrochem. Commun. 2021, 125, 106987. [Google Scholar] [CrossRef]
  24. Xu, Y.; Wang, S.; Yang, J.; Han, B.; Nie, R.; Wang, J.; Wang, J.; Jing, H. In-Situ Grown Nanocrystal TiO2 on 2D Ti3C2 Nanosheets for Artificial Photosynthesis of Chemical Fuels. Nano Energy 2018, 51, 442–450. [Google Scholar] [CrossRef]
  25. Hantanasirisakul, K.; Gogotsi, Y. Electronic and Optical Properties of 2D Transition Metal Carbides and Nitrides (MXenes). Adv. Mater. 2018, 30, 1804779. [Google Scholar] [CrossRef] [PubMed]
  26. Alhabeb, M.; Maleski, K.; Anasori, B.; Lelyukh, P.; Clark, L.; Sin, S.; Gogotsi, Y. Guidelines for Synthesis and Processing of Two-Dimensional Titanium Carbide (Ti3C2Tx MXene). Chem. Mater. 2017, 29, 7633–7644. [Google Scholar] [CrossRef]
  27. Yang, H.G.; Sun, C.H.; Qiao, S.Z.; Zou, J.; Liu, G.; Smith, S.C.; Cheng, H.M.; Lu, G.Q. Anatase TiO2 Single Crystals with a Large Percentage of Reactive Facets. Nature 2008, 453, 638–641. [Google Scholar] [CrossRef] [Green Version]
  28. Yang, C.; Tan, Q.; Li, Q.; Zhou, J.; Fan, J.; Li, B.; Sun, J.; Lv, K. 2D/2D Ti3C2 MXene/g-C3N4 Nanosheets Heterojunction for High Efficient CO2 Reduction Photocatalyst: Dual Effects of Urea. Appl. Catal. B Environ. 2020, 268, 118738. [Google Scholar] [CrossRef]
  29. Zhang, C.J.; Pinilla, S.; McEvoy, N.; Cullen, C.P.; Anasori, B.; Long, E.; Park, S.H.; Seral-Ascaso, A.; Shmeliov, A.; Krishnan, D.; et al. Oxidation Stability of Colloidal Two-Dimensional Titanium Carbides (MXenes). Chem. Mater. 2017, 29, 4848–4856. [Google Scholar] [CrossRef]
  30. Naguib, M.; Mashtalir, O.; Lukatskaya, M.R.; Dyatkin, B.; Zhang, C.; Presser, V.; Gogotsi, Y.; Barsoum, M.W. One-Step Synthesis of Nanocrystalline Transition Metal Oxides on Thin Sheets of Disordered Graphitic Carbon by Oxidation of MXenes. Chem. Commun. 2014, 50, 7420–7423. [Google Scholar] [CrossRef] [Green Version]
  31. Liu, S.; Yu, J.; Jaroniec, M. Anatase TiO2 with Dominant High-Energy {001} Facets: Synthesis, Properties, and Applications. Chem. Mater. 2011, 23, 4085–4093. [Google Scholar] [CrossRef]
  32. Diebold, U. The Surface Science of Titanium Dioxide. Surf. Sci. Rep. 2002, 48, 53–229. [Google Scholar] [CrossRef]
  33. Li, X.; Wu, X.; Liu, S.; Li, Y.; Fan, J.; Lv, K. Effects of Fluorine on Photocatalysis. Chin. J. Catal. 2020, 41, 1451–1467. [Google Scholar] [CrossRef]
  34. Rakhi, R.B.; Ahmed, B.; Hedhili, M.N.; Anjum, D.H.; Alshareef, H.N. Effect of Postetch Annealing Gas Composition on the Structural and Electrochemical Properties of Ti2CTx MXene Electrodes for Supercapacitor Applications. Chem. Mater. 2015, 27, 5314–5323. [Google Scholar] [CrossRef] [Green Version]
  35. Li, B.; Zhao, Z.; Gao, F.; Wang, X.; Qiu, J. Mesoporous Microspheres Composed of Carbon-Coated TiO2 Nanocrystals with Exposed {001} Facets for Improved Visible Light Photocatalytic Activity. Appl. Catal. B Environ. 2014, 147, 958–964. [Google Scholar] [CrossRef]
  36. Ye, M.; Wang, X.; Liu, E.; Ye, J.; Wang, D. Boosting the Photocatalytic Activity of P25 for Carbon Dioxide Reduction by Using a Surface-Alkalinized Titanium Carbide MXene as Cocatalyst. ChemSusChem 2018, 11, 1606–1611. [Google Scholar] [CrossRef]
  37. Wang, K.; Li, X.; Wang, N.; Shen, Q.; Liu, M.; Zhou, J.; Li, N. Z-Scheme Core-Shell meso-TiO2@ZnIn2S4/Ti3C2 MXene Enhances Visible Light-Driven CO2-to-CH4 Selectivity. Ind. Eng. Chem. Res. 2021, 60, 8720–8732. [Google Scholar] [CrossRef]
  38. Li, J.; Huang, H.; Xue, W.; Sun, K.; Song, X.; Wu, C.; Nie, L.; Li, Y.; Liu, C.; Pan, Y.; et al. Self-Adaptive Dual-Metal-Site Pairs in Metal-Organic Frameworks for Selective CO2 Photoreduction to CH4. Nat. Catal. 2021, 4, 719–729. [Google Scholar] [CrossRef]
  39. Huo, Y.; Zhang, J.; Dai, K.; Li, Q.; Lv, J.; Zhu, G.; Liang, C. All-Solid-State Artificial Z-Scheme Porous g-C3N4/Sn2S3-DETA Heterostructure Photocatalyst with Enhanced Performance in Photocatalytic CO2 Reduction. Appl. Catal. B Environ. 2019, 241, 528–538. [Google Scholar] [CrossRef]
  40. Di, J.; Chen, C.; Yang, S.Z.; Chen, S.; Duan, M.; Xiong, J.; Zhu, C.; Long, R.; Hao, W.; Chi, Z.; et al. Isolated Single Atom Cobalt in Bi3O4Br Atomic Layers to Trigger Efficient CO2 Photoreduction. Nat. Commun. 2019, 10, 2840. [Google Scholar] [CrossRef] [Green Version]
  41. Xia, P.; Zhu, B.; Yu, J.; Cao, S.; Jaroniec, M. Ultra-Thin Nanosheet Assemblies of Graphitic Carbon Nitride for Enhanced Photocatalytic CO2 Reduction. J. Mater. Chem. A 2017, 5, 3230–3238. [Google Scholar] [CrossRef]
  42. Jiao, X.; Li, X.; Jin, X.; Sun, Y.; Xu, J.; Liang, L.; Ju, H.; Zhu, J.; Pan, Y.; Yan, W.; et al. Partially Oxidized SnS2 Atomic Layers Achieving Efficient Visible-Light-Driven CO2 Reduction. J. Am. Chem. Soc. 2017, 139, 18044–18051. [Google Scholar] [CrossRef] [PubMed]
  43. Li, X.; Sun, Y.; Xu, J.; Shao, Y.; Wu, J.; Xu, X.; Pan, Y.; Ju, H.; Zhu, J.; Xie, Y. Selective Visible-Light-Driven Photocatalytic CO2 Reduction to CH4 Mediated by Atomically Thin CuIn5S8 Layers. Nat. Energy 2019, 4, 690–699. [Google Scholar] [CrossRef]
  44. Xiao, N.; Li, S.; Liu, S.; Xu, B.; Li, Y.; Gao, Y.; Ge, L.; Lu, G. Novel PtPd Alloy Nanoparticle-Decorated g-C3N4 Nanosheets with Enhanced Photocatalytic Activity for H2 Evolution under Visible Light Irradiation. Chin. J. Catal. 2019, 40, 352–361. [Google Scholar] [CrossRef]
  45. Wang, Z.; Lv, K.; Wang, G.; Deng, K.; Tang, D. Study on the Shape Control and Photocatalytic Activity of High-Energy Anatase Titania. Appl. Catal. B Environ. 2010, 100, 378–385. [Google Scholar] [CrossRef]
  46. Xia, Y.; Cheng, B.; Fan, J.; Yu, J.; Liu, G. Unraveling Photoexcited Charge Transfer Pathway and Process of CdS/Graphene Nanoribbon Composites toward Visible-Light Photocatalytic Hydrogen Evolution. Small 2019, 15, 1902459. [Google Scholar] [CrossRef]
  47. Neaţu, Ş.; Maciá-Agulló, J.A.; Concepción, P.; Garcia, H. Gold-Copper Nanoalloys Supported on TiO2 as Photocatalysts for CO2 Reduction by Water. J. Am. Chem. Soc. 2014, 136, 15969–15976. [Google Scholar] [CrossRef]
  48. Li, Y.; Yin, Z.; Ji, G.; Liang, Z.; Xue, Y.; Guo, Y.; Tian, J.; Wang, X.; Cui, H. 2D/2D/2D Heterojunction of Ti3C2 MXene/MoS2 Nanosheets/TiO2 Nanosheets with Exposed (001) Facets toward Enhanced Photocatalytic Hydrogen Production Activity. Appl. Catal. B Environ. 2019, 246, 12–20. [Google Scholar] [CrossRef]
  49. Li, Y.; Ding, L.; Liang, Z.; Xue, Y.; Cui, H.; Tian, J. Synergetic Effect of Defects Rich MoS2 and Ti3C2 MXene as Cocatalysts for Enhanced Photocatalytic H2 Production Activity of TiO2. Chem. Eng. J. 2020, 383, 123178. [Google Scholar] [CrossRef]
  50. Li, Y.; Ding, L.; Yin, S.; Liang, Z.; Xue, Y.; Wang, X.; Cui, H.; Tian, J. Photocatalytic H2 Evolution on TiO2 Assembled with Ti3C2 MXene and Metallic 1T-WS2 as Co-Catalysts. Nano-Micro Lett. 2020, 12, 1–12. [Google Scholar] [CrossRef] [Green Version]
  51. Zhang, X.; Nie, J.; Rao, F.; Liu, H.; Wang, Y.; Qu, D.; Wu, W.; Zhong, P.; Zhu, G. Ti3C2@TiO2/g-C3N4 Heterojunction Photocatalyst with Improved Charge Transfer for Enhancing Visible-Light NO Selective Removal. Ceram. Int. 2021, 47, 31302–31310. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of the synthesis procedure for (001)TiO2/Ti3C2Tx composite photocatalysts.
Figure 1. Schematic illustration of the synthesis procedure for (001)TiO2/Ti3C2Tx composite photocatalysts.
Catalysts 12 00785 g001
Figure 2. Evidence of the formation of (001) TiO2/Ti3C2Tx complex samples. (a) XRD patterns of TiO2/Ti3C2Tx composites with different usage of HF, Ti3C2Tx and Ti3AlC2. (b) Raman spectra of TT-5, Ti3C2Tx and Ti3AlC2. (c) FESEM and (d) TEM image of TT-5, the inset of c shows TiO2 nanosheets embedding in layered Ti3C2Tx substrate. (e) HRTEM image showing the lattice of the TiO2 with the exposed (001) plane in TT-5. (f) HAADF STEM image and the corresponding EDS elemental maps of Ti, O and C of TT-5.
Figure 2. Evidence of the formation of (001) TiO2/Ti3C2Tx complex samples. (a) XRD patterns of TiO2/Ti3C2Tx composites with different usage of HF, Ti3C2Tx and Ti3AlC2. (b) Raman spectra of TT-5, Ti3C2Tx and Ti3AlC2. (c) FESEM and (d) TEM image of TT-5, the inset of c shows TiO2 nanosheets embedding in layered Ti3C2Tx substrate. (e) HRTEM image showing the lattice of the TiO2 with the exposed (001) plane in TT-5. (f) HAADF STEM image and the corresponding EDS elemental maps of Ti, O and C of TT-5.
Catalysts 12 00785 g002
Figure 3. High-resolution XPS spectra for Ti 2p (a,d), C 1s (b,e) and O 1s (c,f) of TT-5 (ac) and TT-5-u (df).
Figure 3. High-resolution XPS spectra for Ti 2p (a,d), C 1s (b,e) and O 1s (c,f) of TT-5 (ac) and TT-5-u (df).
Catalysts 12 00785 g003
Figure 4. (a) Photocatalytic activities of CO2 reduction over the prepared samples. (b) Comparison of product selectivity for TT-5 and P25. (c,d) Mass patterns of produced CH4 (m/z = 17) and CO (m/z = 29) over TT-5 in photocatalytic 13CO2 reduction.
Figure 4. (a) Photocatalytic activities of CO2 reduction over the prepared samples. (b) Comparison of product selectivity for TT-5 and P25. (c,d) Mass patterns of produced CH4 (m/z = 17) and CO (m/z = 29) over TT-5 in photocatalytic 13CO2 reduction.
Catalysts 12 00785 g004
Figure 5. In situ DRIFTs spectra of photocatalytic CO2 reduction on TT-5.
Figure 5. In situ DRIFTs spectra of photocatalytic CO2 reduction on TT-5.
Catalysts 12 00785 g005
Figure 6. (a) CPD values measured by Kelvin probe and (b) the corresponding work functions of TiO2 nanosheets and Ti3C2Tx sample. (c) Photocurrent density of TT-2, TT-8, TT-5 and P25 sample. (d) Schematic for the mechanism of photocatalytic CO2 reduction on (001) TiO2/Ti3C2Tx.
Figure 6. (a) CPD values measured by Kelvin probe and (b) the corresponding work functions of TiO2 nanosheets and Ti3C2Tx sample. (c) Photocurrent density of TT-2, TT-8, TT-5 and P25 sample. (d) Schematic for the mechanism of photocatalytic CO2 reduction on (001) TiO2/Ti3C2Tx.
Catalysts 12 00785 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, J.; Li, K.; Tan, Q.; Li, Q.; Fan, J.; Wu, C.; Lv, K. Facile Preparation of Highly Active CO2 Reduction (001)TiO2/Ti3C2Tx Photocatalyst from Ti3AlC2 with Less Fluorine. Catalysts 2022, 12, 785. https://doi.org/10.3390/catal12070785

AMA Style

Li J, Li K, Tan Q, Li Q, Fan J, Wu C, Lv K. Facile Preparation of Highly Active CO2 Reduction (001)TiO2/Ti3C2Tx Photocatalyst from Ti3AlC2 with Less Fluorine. Catalysts. 2022; 12(7):785. https://doi.org/10.3390/catal12070785

Chicago/Turabian Style

Li, Jibai, Kaining Li, Qiuyan Tan, Qin Li, Jiajie Fan, Chao Wu, and Kangle Lv. 2022. "Facile Preparation of Highly Active CO2 Reduction (001)TiO2/Ti3C2Tx Photocatalyst from Ti3AlC2 with Less Fluorine" Catalysts 12, no. 7: 785. https://doi.org/10.3390/catal12070785

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