Next Article in Journal
A Theoretical and Experimental Study for Enzymatic Biodiesel Production from Babassu Oil (Orbignya sp.) Using Eversa Lipase
Next Article in Special Issue
Recent Progress of Ga-Based Catalysts for Catalytic Conversion of Light Alkanes
Previous Article in Journal
Is Black Titania a Promising Photocatalyst?
Previous Article in Special Issue
Construction of SrTiO3–LaCrO3 Solid Solutions with Consecutive Band Structures for Photocatalytic H2 Evolution under Visible Light Irradiation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Enhanced CO2 Photoreduction over Bi2Te3/TiO2 Nanocomposite via a Seebeck Effect

1
TJU-NIMS International Collaboration Laboratory, Key Laboratory of Advanced Ceramics and Machining Technology (Ministry of Education), Tianjin Key Laboratory of Composite and Functional Materials, School of Material Science and Engineering, Tianjin University, Tianjin 300072, China
2
Collaborative Innovation Centre of Chemical Science and Engineering (Tianjin), Tianjin 300072, China
3
International Center for Materials Nanoarchitectonics (WPI-MANA), National Institute for Materials Science (NIMS), 1-1 Namiki, Tsukuba 305-0044, Japan
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(11), 1323; https://doi.org/10.3390/catal12111323
Submission received: 28 September 2022 / Revised: 15 October 2022 / Accepted: 24 October 2022 / Published: 27 October 2022
(This article belongs to the Special Issue Photocatalysis for Energy Transformation Reactions)

Abstract

:
The activation of carbon dioxide (CO2) molecules and separation/transfer of photoinduced charge carriers are two crucial factors influencing the efficiency of CO2 photoreduction. Herein, we report a p-type Bi2Te3/commercial TiO2 (pBT/P25) nanocomposite for enhanced CO2 photoreduction. Upon light irradiation, a temperature gradient formed in pBT induces the Seebeck effect to build a thermoelectric field, which promotes the charge carriers’ separation/transfer. Additionally, pBT with a strong light absorption capacity generates the photothermal effect favoring the activation of CO2 molecules. In addition, the excellent electric conductivity and large work function render pBT an efficient cocatalyst for further improving the charge carriers’ separation/transfer. Owing to the synergistic enhancement effect of pBT on the activation of CO2 molecules and promotion of charge separation/transfer, we achieved the highest CO evolution rate over pBT(2)/P25 of 19.2 μmol·gcat−1·h−1, which was approximately 5.5 times that of bare P25. This work suggests that a thermoelectric material/semiconductor nanocomposite could be developed as an efficient photo-thermo-electro-chemical conversion system for enhanced CO2 reduction via promoting the charge carriers’ separation/transfer.

Graphical Abstract

1. Introduction

The utilization of fossil fuels leads to the emission of ~1.3 × 104 million tons of carbon dioxide (CO2) in the atmosphere per year, which is one of the reasons for global warming and the greenhouse effect [1,2]. Converting CO2 into high-value-added chemicals has become the best choice for its commercial application. Nonetheless, CO2 conversion reactions usually require a high energy input [3]. Thus, it is of great significance to develop a reliable CO2 conversion path under the condition of low energy consumption. As an ideal green technology for solar-to-chemical conversion, the photoreduction of CO2 into hydrocarbon solar fuels with sunlight as the sole input energy source has been attracting a great deal of research interest in recent years [4,5,6,7,8,9]. In those possible CO2 photoreduction products, CO can be utilized as both a Fischer-Tropsch synthesis reaction feed gas and a reducing agent in metal smelting [10,11]. In this context, the reduction of CO2 to toxic CO for various industrial applications is a promising strategy for solving the two globally concerning issues of the energy crisis and greenhouse effects. Among many photocatalysts, TiO2 has been widely studied for water splitting [12], the degradation of organic pollutants [13], and CO2 photoreduction [14] due to its low price, non-toxic property, and stability. However, it is still a tough challenge to achieve a high photoconversion efficiency because of the extremely thermodynamically stable non-polar CO2 molecules and the fast recombination of photogenerated electron–hole pairs over the TiO2 surface [15,16,17].
On one hand, to promote the adsorption/activation of CO2 molecules, the significant role of the photothermal effect has been paid extensive attention, because the high photothermal temperature is believed to greatly favor the expansion of light absorbance and more importantly the activation of the CO2 molecules, thereby leading to a significantly enhanced CO2 photoreduction performance [18,19,20,21]. On the other hand, to suppress the rapid recombination of photoexcited charge carriers, one effective solution is to introduce co-catalysts, including noble metals [22,23,24], noble-metal-contained alloys, or core-shell structured bimetals [25,26,27,28,29]. Additionally, our lab recently developed the alkalinized Ti3C2 as an excellent noble-metal-free cocatalyst for significantly enhancing the photocatalytic CO2 reduction of commercial TiO2 (P25) [30,31]. In principle, the large work function and excellent electrical conductivity of a cocatalyst favor the separation of photo-generated electrons and holes and thus achieve a high activity of CO2 photoreduction [23,24].
In addition to using cocatalysts, previous studies have shown that electric/magnetic fields could serve as the driving force for effectively promoting the separation/transfer of photoexcited charge carriers and hence increasing the photocatalytic activities. For instance, some researchers found that the oppositely charged ends of tourmaline could drive electrons and holes to transfer in opposite directions to enhance the photocatalytic activity of dye or isopropanol (IPA) degradation [32,33]. Moreover, the non-toxic Fe3O4 was selected as an appropriate supplier of a magnetic field in a TiO2-based photocatalyst with enhanced charge separation, thus resulting in significantly increased photocatalytic activities [34,35].
It is known that most thermoelectric materials possess a strong absorption capacity, which can make full use of ultraviolet–visible–infrared (UV–Vis–IR) light to produce the photothermal effect [36]. In addition to the photothermal effect, thermoelectric materials are able to induce an electric field by a temperature gradient, which is generally known as the Seebeck effect. Previously, the applications of various thermoelectric materials in photocatalysis have been reported, either by the thermoelectric material alone [37,38,39,40] or by combining the thermoelectric material with other wide bandgap semiconductors to form the type-II heterostructure [41,42,43]. To the best of our knowledge, however, the aforementioned Seebeck effect has never been used for enhancing photocatalytic activity. As one of the common thermoelectric materials, the excellent optical, electrical, and thermoelectric properties of the black bismuth telluride (Bi2Te3) would render the formation of the photothermal effect upon light irradiation. Meanwhile, if Bi2Te3 is combined with a suitable semiconductor and the Bi2Te3/semiconductor composite is subjected to light irradiation, a photothermal effect and a temperature gradient could be formed, inducing a thermoelectric field in Bi2Te3 via the Seebeck effect [44]. Moreover, Bi2Te3 could also serve as a cocatalyst for CO2 reduction owing to its excellent electrical conductivity and high work function (~5.3 eV, close to the Fermi-level of Bi2Te3) [45]. In this context, Bi2Te3 is believed to play a critical role in activating the CO2 molecules and promoting the separation/transfer of photoinduced charge carriers, thus significantly enhancing the CO2 photoreduction performance.
Herein, we report a Bi2Te3/TiO2 nanocomposite, showing significantly enhanced CO2 photoreduction activity. With the intention of CO2 reduction, we purposely selected the p-type Bi2Te3 (pBT), from which the thermoelectric field would be favorable for driving the photogenerated electrons from TiO2 to the pBT surface to participate in the CO2 reduction reaction. On the one hand, a temperature gradient formed in pBT induces a thermoelectric field via the Seebeck effect, favoring the separation/transfer of photoinduced electrons and holes. On the other hand, the strong light absorption capacity of pBT offers an efficient photo-thermal conversion to produce a high temperature on the pBT surface, which facilitates the activation of CO2 molecules. Moreover, the interfacial electronic interaction in pBT/P25 endows pBT as a cocatalyst with the electron-rich environment for increasing CO2 photoreduction. Hence, the pBT/P25 nanocomposite exhibited a CO evolution rate of 19.2 μmol·gcat−1·h−1, approximately 5.5 times that of bare P25. Our research demonstrates that the thermoelectric effect could be incorporated into the strategy for developing an efficient photo-thermo-electro-chemical conversion system for CO2 reduction based on a full-spectrum-responsive thermoelectric material/semiconductor nanocomposite.

2. Results and Discussion

2.1. Structural, Morphological, and Optical Characterizations

As shown in Figure 1a, the XRD pattern of pBT(2)/P25 could be indexed well by the joint committee on powder diffraction standards (JCPDS) card of Bi2Te3 [15] and dual-phased P25 [46]. Three main peaks at 25.3, 37.4, and 47.8 were typical for the (101), (004), (200) planes of anatase TiO2 (JCPDS No. 21-1272), and the peaks at 27.4, 36.0, and 41.2 could be assigned to (110), (111), and (220) planes of rutile TiO2 (JCPDS No. 21-1276). The peaks at 27.7°, 37.8°, and 44.7° corresponded to the (015), (1010), and (110) planes of Bi2Te3 (JCPDS No. 15-0863). TEM images showed uniform spherical P25 nanoparticles with a diameter of ~25 nm, and pBT nanoparticles with a diameter of ~10 nm (Figure S1). The HRTEM image, as shown in Figure 1b displayed that the distances of lattice fringes were 0.322, 0.352, and 0.249 nm, and in accordance with the (015) planar space of pBT, (101) planar spaces of anatase, and rutile TiO2, respectively [46]. The high-angle annular dark-field (HAADF) TEM image and corresponding EDS analysis results show the existence of Ti, O, Bi, and Te elements in pBT(2)/P25 (Figure S2). The above results indicate that the pBT nanoparticles were loaded on P25 with an excellent interfacial structure, which was beneficial for the charge-carrier transfer, as will be discussed below.
The SEM images of P25 and pBT(2)/P25 showed that the diameter of P25 in pBT(2)/P25 was maintained at ~25 nm after hydrothermal treatment (Figure S3). Moreover, we studied the Brunauer-Emmett-Teller (BET) surface area of P25 and pBT(2)/P25 using N2 sorption analysis. As shown in Figure S4, the N2 sorption isotherm of pBT(2)/P25 did not exhibit an apparent difference in comparison to that of P25. The specific surface area of P25 and pBT(2)/P25 was 57.6 and 54.9 m2·g−1, respectively. Although the efficiency of surface redox reactions could be dictated by the morphology of photocatalysts or the number of surface reduction/oxidation reactive sites, such similar results of P25 and pBT(2)/P25 in the structure/morphology and specific surface area could not lead to the significantly enhanced CO2 photoreduction activity of pBT(2)/P25 as shown below.
The XPS survey spectrum of pBT(2)/P25 showed the peaks of Ti 2p, O 1s, Bi 4f, and Te 3d (Figure S5). In the high-resolution XPS spectra, we observed that the two peaks of Ti 2p3/2 and Ti 2p1/2 in pBT(2)/P25 were shifted to higher energies (463.90 and 458.20 eV) than those in pure P25 (463.75 and 458.15 eV, Figure S6a) while the peaks of O 1s were almost unchanged (529.50 eV, Figure S6b). The peaks of Bi 4f7/2 and Bi 4f5/2 in pBT(2)/P25 were shifted to lower energies (159.00 and 164.30 eV) than those in pure pBT (159.20 and 164.60 eV, Figure S6c) [47]. Similarly, the peaks of Te 3d5/2 Te 3d5/2 in pBT(2)/P25 were shifted to lower energies (575.70 and 586.10 eV) than those in pure pBT (576.15 and 586.40 eV), and besides, an oxide peak of Te 3d at ~580.30 eV was observed in pBT(2)/P25 (Figure S6d). The shift of binding energies strongly suggested the electronic coupling of pBT and P25 [48,49], which might form an electron-rich environment in pBT favoring CO2 reduction [31].
The UV–Vis–IR DRS spectra in Figure 2a showed that compared with pure P25 only absorbing UV light due to the wide bandgap (~3.2 eV) [30], the absorbance of pBT(2)/P25 was appreciably enhanced, especially in the Vis–IR range because of the strong light absorption ability of loaded pBT nanoparticles. As expected, the strong light absorption ability of pBT would result in quite high photothermal temperature in pBT(x)/P25 depending on the loaded amount of pBT., The in situ monitored photothermal temperatures on the surfaces of P25, pBT, and pBT(x)/P25 (x = 1, 2, 3, 4) upon a 300 W Xe lamp irradiation are shown in Figure 2b. The photothermal temperatures for all the samples increased with the increasing irradiation time and became gradually saturated after approximately 30 min. The photothermal temperatures for P25 and pBT were 120 °C and 350 °C, respectively. As for the pBT(x)/P25 (x = 1, 2, 3, 4) composites, the photothermal temperatures were between 150 and 250 °C. The greater the mass percentage of pBT, the higher photothermal temperature would be. The high photothermal temperature might be one of the important reasons accounting for the enhanced CO2 photoreduction performance of pBT(2)/P25, as discussed below.

2.2. Temperature Gradient and Electric Potential Simulation

A temperature gradient was necessary for the Seebeck effect of pBT to induce a built-in electric field. In practice, it was hard to measure the temperature gradient in pBT/P25. Instead, the temperature gradient in pBT/P25 nanocomposite was theoretically simulated by a commercial FEM solver in COMSOL Multiphysics according to Equation (1) [44,49]. Considering the fact that the thermal conductivities of TiO2 (~8.50 W m−1·K−1) and Bi2Te3 (~1.47 W·m−1·K−1) are different [50,51], and the heat-dissipating from Bi2Te3 of higher temperature to TiO2 of lower temperature was faster than that to a CO2 atmosphere, a temperature gradient could be formed between the exposed surface of Bi2Te3 and the interface of Bi2Te3/TiO2. As shown in Figure 3a, the simulated temperature distribution verified the formation of a temperature gradient in a Bi2Te3/TiO2 nanocomposite under light irradiation. The temperature gradient over Bi2Te3 could be ~40 °C, meaning the Seebeck effect could be excited in Bi2Te3/TiO2 via light illumination [44,52].
In principle, the Seebeck effect suggested that an electric field will be formed by the majority carrier transfer from the high-temperature end to the low-temperature end in a thermoelectric material [53]. For p-type Bi2Te3 (pBT), the majority carrier (holes) would diffuse from the high-temperature end (exposed surface of pBT) to the low-temperature end (interface of pBT/P25), inducing an electric field in the same direction as the directional diffusion of holes. The COMSOL software displayed the electric potential over Bi2Te3 according to the above-simulated temperature gradient. Figure 3b showed that the electric potential was ~7.85 V. The direction of the thermoelectric field was from the high electric potential surface (high-temperature surface of pBT) to the low electric potential surface (interface of pBT/P25). Thus, we expected that the Seebeck effect would undoubtedly influence the CO2 photoreduction process of pBT/P25 by accelerating the transfer of photoinduced electrons.

2.3. CO2 Photoreduction Performances

The GC spectra were obtained to determine the products of CO2 photoreduction over pBT(2)/P25. As shown in Figure S7, we observed that the main reduction product was CO with an evolution rate of 19.2 μmol·gcat−1·h−1. In addition, CH4 and H2 were also detected as the reduction products with much lower evolution rates of 0.4 μmol·gcat−1·h−1 and 2.1 μmol·gcat−1·h−1, respectively. O2 was detected as the oxidation product of H2O with an evolution rate of 9.6 μmol·gcat−1·h−1, indicating that the CO2 photoreduction over pBT/P25 was a complete redox reaction. Considering the aforementioned results, we used the CO evolution rate to evaluate the CO2 photoreduction activities of P25 and pBT(x)/P25 (x = 1, 2, 3, 4).
As displayed in Figure 4a, the pure pBT showed almost no activity, and the pure P25 only exhibited a very low CO evolution rate of 3.5 μmol·gcat−1·h−1. Surprisingly, after loading pBT on P25, the as-obtained pBT(x)/P25 nanocomposites exhibited significantly increased activity compared with either pBT or P25. With an optimal pBT loading of 2 wt%, the pBT(2)/P25 composite showed the highest CO evolution rate of 19.2 μmol·gcat−1·h−1, which was nearly 5.5 times that of P25. Further increasing the pBT loading amount (e.g., pBT(3)/P25, and pBT(4)/P25) decreased the activity, probably due to the light-blocking effect of pBT that had almost no activity for CO2 photoreduction. The O2 evolution rates were shown in Figure 4b. Additionally, pBT(2)/P25 had the highest activity, which was ~5.1 times higher than that of P25 (1.9 μmol·gcat−1·h−1). As mentioned above, the enhanced CO2 photoreduction performance of pBT/P25 could be partially attributed to the high photothermal temperature over the pBT (e.g., ~193 °C for pBT(2)/P25) and thermoelectric field in pBT upon light irradiation.
To verify the important role of the photothermal effect and the Seebeck effect in improving the activity of pBT(2)/P25 for CO2 photoreduction, we carried out several experiments as follows. The thermocatalytic reaction over pBT(2)/P25 was conducted in the dark using a flexible electric heating jacket to heat the reaction cell at ~193 °C. The performance of thermal catalysis (0.2 μmol·gcat−1·h−1) was much lower than that of photocatalysis under irradiation (Figure S8). We also carried out a blank test of CO2 photoreduction on pBT(2)/Al2O3 under the same conditions as on pBT(2)/P25. The photothermal temperature over pBT(2)/Al2O3 was very close to that over pBT(2)/P25 (Figure S9), but Al2O3 did not possess any photocatalytic activity. The CO evolution rate over pBT(2)/Al2O3 (<0.1 μmol·gcat−1·h−1, Figure S8) was negligible in comparison with that over pBT(2)/P25. In addition, the CO2 photoreduction was carried out over pBT(2)/P25 by using a Y44 cutoff filter to completely block UV light (denoted as Xe-Y44). In this case, P25 would not be excited to show any photocatalytic activity judging from the light spectrum of Xe-Y44 (Figure S10) and the absorption spectrum of P25 (Figure 2a). Under the Xe-Y44 irradiation, the photothermal temperature over pBT(2)/P25 reached ~173 °C Figure S11). Surprisingly, only a very low CO evolution rate (~0.2 μmol·gcat−1·h−1, Figure S8) was obtained over pBT(2)/P25 under Xe-Y44 irradiation. The above results indicated that without photoexcitation of P25, only heating the pBT(2)/P25 catalyst to a high temperature could not improve the activity of pBT(2)/P25 for CO2 photoreduction.
In addition, to further confirm the role of the Seebeck effect in enhancing the CO2 photoreduction activity, we used commercial n-type Bi2Te3 (nBT) to prepare the nBT(2)/P25 nanocomposite as a comparison of pBT(2)/P25. It is worth pointing out that nBT and pBT had the same light absorption capacity (Figure 5a), so the in situ monitored photothermal temperature of nBT(2)/P25 (~192 °C) was almost the same as that of pBT(2)/P25 (~193 °C, Figure 5b). Interestingly, the CO evolution rate over nBT(2)/P25 (11.1 μmol·gcat−1·h−1, Figure 5c) was much less than that over pBT(2)/P25 (19.2 μmol·gcat−1·h−1). In contrast to pBT with holes as the majority carrier, electrons are the majority carrier in nBT. Upon light irradiation, the electrons would diffuse from the high-temperature end on the exposed surface of nBT to the low-temperature end at the interface of nBT/P25. Obviously, the thermoelectric field formed in nBT would inhibit the transfer of photoinduced electrons from P25 to the nBT surface. Therefore, the CO evolution rate over nBT(2)/P25 was much less than that over pBT(2)/P25. The difference in CO evolution rates over pBT(2)/P25 and nBT(2)/P25 clearly confirmed that the Seebeck effect enhanced CO2 photoreduction via accelerating photoexcited charge-carrier separation. Actually, several works have proved the promotion of Bi2Te3 on the charge transfer over TiO2 due to the thermoelectric properties, indicating that it is feasible to enhance the photocatalytic activity over the Bi2Te3/TiO2 nanocomposite by inducing a thermoelectric field in Bi2Te3 via the Seebeck effect [54,55].
On the other hand, we performed the CO2 photoreduction experiment by putting the reaction cell in a basin containing cold water, just leaving the top window above the water level for passing through incident light. Under this circumstance, the photothermal temperature of pBT(2)/P25 (denoted pBT(2)/P25-Cool) under the full-arc Xe lamp irradiation was significantly reduced to ~84 °C (Figure S12, Table S1). We found that, as shown in Figure S13, the CO evolution rate over pBT(x)/P25-Cool was remarkably lower than that over pBT(x)/P25, however, the activity of P25-Cool had only a slight decline. The above results confirmed that the influence of the Seebeck effect on promoting charge transfer was weakened by the low-temperature because of the slight temperature gradient. Moreover, in consideration of the high photothermal temperature over the pBT surface, the above results also proved the photothermal effect of pBT on the enhanced activity of pBT(2)/P25 for CO2 photoreduction, in the process of which the activation of CO2 molecules by the high photothermal temperature played a crucial role. In fact, similar results were also reported in Ni-, Rh-, or Au-loaded SiO2 and CeO2 catalysts, in the interface of which the strong local electromagnetic field was formed to greatly favor the activation of CO2 molecules [19,20,56].
Three cycles of reaction for a total of 6 h showed that there was almost no decrease in catalytic activity (Figure S14). The XRD patterns of pBT(2)/P25 before and after the reaction also showed no change (Figure S15). Additionally, the XPS spectra of pBT(2)/P25 showed that no new peak or obvious variation in the peak area ratio of four elements (Ti, O, Bi, and Te) in pBT(2)/P25 was observed after reactions, confirming the stability of pBT(2)/P25 (Figure S16). To clarify whether CO originated from CO2 reduction, a series of blank experiments—as reported in previous studies—were conducted [57,58]. The reaction in the absence of pBT(2)/P25 catalyst (CO2 and H2O were added in the reaction cell) showed no evolution of CO or other hydrocarbons. When CO2 and/or H2O were not added or were replaced by Ar gas while keeping other conditions identical, no CO or other hydrocarbon was detected. These results exclude the possibility that the high yield of CO was from the adsorbed impurities containing carbon (if possible) on pBT(2)/P25.

2.4. Photoelectrochemical and Photoluminescence Properties

Previous results and discussion have proven the essential roles of the photothermal effect and the subsequent Seebeck effect of pBT in the significantly enhanced activity of pBT(2)/P25. Given the excellent electrical conductivity and large work function of pBT [45], we expected that pBT might also serve as a suitable cocatalyst in promoting the separation/transfer of photoexcited charge carriers. From this perspective, we carried out the photoelectrochemical (PEC) measurements of P25 and pBT(x)/P25, in the process of which the photothermal effect and Seebeck effect of pBT could be neglected since the surface temperatures of P25 and pBT(x)/P25 photoanodes immersed in the 0.5 M K2SO4 aqueous solution were in situ monitored to be only ~55.0 °C (Figure S17, Table S2). From Figure 6a, we could see that pBT(2)/P25 showed the highest transient photocurrent (TPC) intensity compared with P25 and other pBT(x)/P25 (x = 1, 3, 4) samples, corresponding to the optimal photocatalytic activity. The pBT(2)/P25 also clearly showed the lowest PL intensity (Figure 6b), indicating a quenching effect of photoinduced electrons by pBT [59]. In no doubt, the photoelectrochemical and photoluminescence properties verified the significant role of pBT as an efficient cocatalyst in enhancing the CO2 photoreduction performance of pBT(2)/P25. Additionally, we also utilized a UV lamp to completely exclude the Vis-NIR light, comparing the photocatalytic performance of P25 and pBT(2)/P25 (Figure S18). Under UV lamp illumination, the photothermal temperature over P25 and pBT(2)/P25 was enough low (~45 °C for P25-UV and 57 °C for pBT(2)/P25-UV, Figure S19) so that the Seebeck effect and photothermal effect were negligible. Additionally, the results show that the CO evolution rate over pBT(2)/P25 was still higher than that of P25, corresponding to the PEC results and confirming the enhancement effect of pBT as a cocatalyst. In fact, the work function of pBT (~5.3 eV) was similar to that of some noble metal cocatalysts (Pt~5.65 eV, Au~5.1 eV) [60]. The enrichment effect of the noble metal cocatalysts with a large work function on the photoexcited electrons from TiO2 will transfer electrons to the cocatalysts and shift the Fermi energy level of the composites to a more negative point (closer to the conduction band of TiO2), significantly improving the photocurrent intensity and reduction ability of the photogenerated electrons [61,62]. In this context, pBT as a cocatalyst not only promoted charge transfer efficiency but also enhanced the reduction ability of photoexcited electrons over pBT/P25.
Based on the above results and discussion, we could clarify that the thermoelectric effect, photothermal effect, and co-catalyzing effect of pBT were incorporated into the pBT/P25 nanocomposite with the promoted separation/transfer of photoinduced charge carriers, thereby realizing enhanced CO2 photoreduction activity. Accordingly, we propose in Figure 7 a mechanism elucidating the enhanced CO2 photoreduction over pBT/P25. Upon light irradiation, a temperature gradient is formed in pBT due to the non-uniform heat dissipation in pBT/P25. Then, the temperature gradient in pBT induces the Seebeck effect to create a built-in thermoelectric field, which drives the electrons excited from P25 to the pBT surface to participate in the CO2 reduction. Meanwhile, the high photothermal temperature over the pBT surface benefits the thermodynamically endothermic CO2 photoreduction by activating the CO2 molecules. In addition, the excellent electric conductivity and large work function render pBT an efficient cocatalyst to improve the transfer and reduction ability of the photoinduced electrons for CO2 reduction. Briefly, pBT plays an essential role in the separation/transfer of photoinduced charge carriers and the activation of CO2 molecules, resulting in highly efficient CO2 photoreduction over pBT/P25.

3. Materials and Methods

3.1. Chemicals and Synthesis

Commercial TiO2 (P25) (Sigma-Aldrich Co., Ltd., St. Louis, MO, USA), p-type Bi2Te3 (Iuding chemistry Co., Ltd., Hong Kong, China), n-type Bi2Te3 (Iuding chemistry Co., Ltd.), and CO2 (99.99%, Tianjin Liufang Gas Co., Ltd., Tianjin, China) were used without further treatment.
The p-type Bi2Te3/P25 (pBT/P25) was synthesized using a hydrothermal method as follows: 500 mg of P25 and different amounts of p-type Bi2Te3 were mixed in 30 mL of Deionized (DI) water. After stirring for 30 min, the mixture was transferred into a 50 mL Teflon-lined stainless-steel autoclave and heated at 170 °C for 24 h. The resulting suspensions were completely washed with DI water and then dried in a vacuum oven at 60 °C for 12 h, obtaining pBT(x)/P25 (x = 1, 2, 3, 4 wt%; x is the weight ratio of pBT to P25) nanocomposites. Following the same procedure as mentioned above, we also prepared n-type Bi2Te3/P25 (nBT(2)/P25) for comparison.
Photoanode samples were fabricated by spin-coating the as-prepared catalysts on fluorine-doped SnO2 (FTO) conductive glass. Typically, 10 mg of catalyst was dispersed in 5 mL of ethanol under ultrasonication. The slurry solution was spin-coated on FTO glass (10 × 10 mm2) for 60 cycles. After heating at 200 °C for 2 h, a photoanode sample was obtained.

3.2. Materials Characterization

The crystal phase was determined by X-ray diffraction (XRD; D8 Advanced, Bruker, Rheinstetten, Germany) using Cu-Kα radiation (λ = 0.154178 nm) at a scanning rate of 0.02°·s−1. The morphology and microstructure of the samples were examined using transmission electron microscopy (TEM; Technai G2 F20, FEI, Eindhoven, The Netherlands). Ultraviolet-visible-infrared (UV-vis-IR) diffuse reflectance spectra (DRS) were recorded at room temperature on a spectrophotometer (UV-3600, Shimadzu, Kyoto, Japan) using BaSO4 as a reference and converted to absorption spectra via the Kubelka-Munk (K-M) method. The chemical state was analyzed by X-ray photoelectron spectroscopy (XPS, Escalab 250Xi, Thermo Scientific, Waltham, MA, USA) calibrated by the C 1s peak (284.6 eV) of the surface adventitious carbon. Photoluminescence (PL) measurement was conducted on a fluorescence spectrometer (Fluorolog-3, Horiba Scientific, Piscataway, NJ, USA) with an excitation light wavelength of λ = 330 nm.

3.3. CO2 Photoreduction Evaluation

The CO2 photoreduction experiment was carried out in a reaction chamber with a quartz lid that secured the adequate transmittance of incident light. The reaction chamber with a total volume of 330 mL (OLPCRS-3, Shanghai Boyi Scientific Instrument Co. Ltd., Shanghai, China) was connected to a closed gas circulation system. Fifty milligrams of catalyst was uniformly dispersed on a porous quartzose slice, which was fixed on the stage inside the reactor. Three milliliters of DI water was injected into the cell as the sacrificial agent and proton source for CO2 photoreduction. After the reactor was sealed and evacuated, CO2 gas was injected until a pressure of 70 kPa. A 300 W full-arc xenon lamp was employed as the light source, and the incident light intensity was measured to be ~1.3 W·cm−2 using an optical power meter (THORLABS, PM 100D, Bergkirchen, Germany). The sample surface temperature was in situ monitored using an infrared thermometer (UT3021, UNI-T Co., Ltd., Dongguan, China), which was pre-calibrated using a reflectance coefficient of 0.36.
During the reaction, a gaseous sample (0.5 mL) was periodically taken from the reaction vessel and analyzed on a gas chromatograph (GC-2014, Shimadzu, Kyoto, Japan) equipped with a flame ionization detector (FID) for analysis of CO and CH4, and a methanizer for an analysis of CO2 using N2 as the carrier gas. The evolved O2 and H2 were analyzed on an online GC (GC-2014C, Shimadzu, Kyoto, Japan) equipped with a thermal conductivity detector (TCD) using Ar as the carrier gas.

3.4. Photoelectrochemical Measurement

A photoelectrochemical (PEC) measurement was conducted on an electrochemical workstation (CHI 660D, Chenhua Co., Ltd., Shanghai, China) in a three-electrode quartz glass cell containing 0.5 M K2SO4 aqueous solution under a 300 W full-arc Xe lamp irradiation. A piece of Pt wire, an Ag/AgCl electrode, and the photoanode were employed as the counter electrode, reference electrode, and working electrode, respectively.

3.5. COMSOL Multiphysics Simulation

A commercial FEM solver in COMSOL Multiphysics was utilized to simulate the heat distribution in pBT/P25. A two-dimensional (2D) geometry with rotation was built to form a three-dimensional (3D) model, in which the dimensions of the pBT and P25 nanoparticles were the same as the actual sizes obtained from SEM and TEM observations. For simplicity, a single-wavelength infrared light of λ = 800 nm was selected as the incident light source to solely excite pBT. The simulation was carried out in the atmosphere of CO2 and H2O. The heat power volume density (Qd) is written as [63]:
Q d = 1 / 2 ε 0 ω ln ( ε r ) | E | 2
where ε0 is the permittivity of the vacuum, ω is the angular frequency of the light, εr is the relative permittivity of Bi2Te3, and E is the light energy absorbed by Bi2Te3, on which the optical properties such as refractive index and absorption coefficient are dependent. All the optical constants of Bi2Te3 were obtained from the material library of the software. The results of in situ monitored photothermal temperatures showed that temperatures over P25, pBT, and pBT(x)/P25 could be steady after 40 min. Thus, to provide the numerical value for a temperature gradient, an initial temperature over Bi2Te3/TiO2 composites was set to 100 °C according to the curves of in situ monitored photothermal temperatures at 20 min. Under this condition, the computational complexity of COMSOL software would be reduced, and the results could be similar to the actual temperatures. According to the simulated temperature gradient over Bi2Te3/TiO2 composite, we further calculated the electric potential via COMSOL software. The Seebeck coefficient and electric conductivity were obtained from the manufacturer of pBT (Iuding chemistry Co., Ltd.). During this simulation process, the electric potential at the bottom of the TiO2 nanoparticle was regarded as zero electric potential.

4. Conclusions

In summary, we used a simple one-step hydrothermal method to successfully fabricate the pBT/P25 nanocomposite exhibiting a remarkably improved activity of CO2 photoreduction. The highest CO evolution of pBT(2)/P25 (19.2 μmol·gcat−1·h−1) was nearly 5.5 times that of P25. The enhancement in the CO2 photoreduction could be attributed to the following reasons: (i) The Seebeck effect induced a thermoelectric field in pBT for promoting the charge-carrier separation and transfer; (ii) The full-spectrum light absorption capacity endowed pBT with a photothermal effect to produce a high photothermal temperature for the activation of CO2 molecules; (iii) The excellent electric conductivity and large work function rendered pBT an efficient cocatalyst to further improve the charge carriers’ separation/transfer and reduction ability. Owing to the synergistic enhancement effect in pBT on the activation of CO2 molecules and the promotion of charge separation/transfer, the CO2 photoreduction efficiency was significantly enhanced. Our research provides a feasible strategy for the development of an efficient photo-thermo-electro-chemical conversion system for CO2 reduction based on constructing a full-spectrum responsive thermoelectric material/semiconductor nanocomposite.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal12111323/s1, Figure S1: TEM image of pBT(2)/P25; Figure S2: HAADF TEM image and EDS analysis of pBT(2)/P25; Figure S3: SEM images of P25 and pBT(2)/P25; Figure S4: N2 sorption isotherms of P25 and pBT(2)/P25; Figure S5: XPS spectra of pBT(2)/P25; Figure S6: HR-XPS spectra of Ti 2p, O 1s in pure P25 and pBT(2)/P25, Bi 4f and Te 3d in pure pBT and pBT(2)/P25; Figure S7: GC spectra of CO, CH4, H2, and O2 in CO2 photoreduction; Figure S8: CO evolution rates over pBT(2)/P25 and pBT(2)/Al2O3; Figure S9: In situ monitoring of the photothermal temperatures over pBT(2)/P25 and pBT(2)/Al2O3; Figure S10: Light spectra of the Xe lamp with or without a Y44 filter; Figure S11: In situ monitoring of the photothermal temperatures over pBT(2)/P25 under Xe lamp irradiation with and without a Y44 filter; Figure S12: In situ monitoring of the photothermal temperatures over P25 and pBT(2)/P25 under photothermal and cooling conditions; Table S1:Temperatures over P25 and pBT(x)/P25 under photothermal and cooling conditions; Figure S13: CO evolution rates over pBT(x)/P25 under photothermal and cooling conditions; Figure S14: Three cycles of CO2 photoreduction over pBT(2)/P25; Figure S15: XRD patterns of pBT(2)/P25 before and after reaction; Figure S16: HR-XPS spectra of Ti 2p, O 1s, Bi 4f, and Te 3d in pBT(2)/P25 before and after reaction; Figure S17: In situ monitoring of photoanodes surface temperatures; Table S2: Surface temperatures of P25 and pBT(x)/P25 photoanodes; Figure S18: CO evolution rates over P25 and pBT(x)/P25 under Xe and UV lamps irradiation; Figure S19: In situ monitoring of the photothermal temperatures over pBT(2)/P25 under Xe and UV lamps; Table S3: Comparison of CO2 photoreduction performance over TiO2-based photocatalysts. References [64,65,66,67,68,69,70,71] are cited in the supplementary materials.

Author Contributions

Conceptualization, Y.L. and D.W.; methodology and investigation, Y.L., Z.J., and H.H.; writing, Y.L. and D.W.; supervision, D.W.; funding acquisition and resources, D.W., L.L., and J.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (51572191, 21633004).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bibi, M.; Ullah, R.; Sadiq, M.; Sadiq, S.; Khan, I.; Saeed, K.; Zia, M.A.; Iqbal, Z.; Ullah, I.; Iqbal, Z.; et al. Catalytic hydrogenation of carbon dioxide over magnetic nanoparticles: Modification in fixed-bed reactor. Catalysts 2021, 11, 592. [Google Scholar] [CrossRef]
  2. Zahedi, G.; Elkamel, A.; Lohi, A. Dynamic optimization strategies of a heterogeneous reactor for CO2 conversion to methanol. Energy Fuels 2007, 21, 2977–2983. [Google Scholar] [CrossRef]
  3. Shi, L.; Liu, H.; Ning, S.; Ye, J. Localized surface plasmon resonance effect enhanced Cu/TiO2 core–shell catalyst for boosting CO2 hydrogenation reaction. Catal. Sci. Technol. 2022, 12, 6155–6162. [Google Scholar] [CrossRef]
  4. Habisreutinger, S.N.; Schmidt-Mende, L.; Stolarczyk, J.K. Photocatalytic reduction of CO2 on TiO2 and other semiconductors. Angew. Chem. Int. Ed. 2013, 52, 7372–7408. [Google Scholar] [CrossRef] [PubMed]
  5. Zeng, F.; Mebrahtu, C.; Xi, X.; Liao, L.; Ren, J.; Xie, J.; Heeres, H.J.; Palkovits, R. Catalysts design for higher alcohols synthesis by CO2 hydrogenation: Trends and future perspectives. Appl. Catal. B Environ. 2021, 291, 120073. [Google Scholar] [CrossRef]
  6. Chang, X.; Wang, T.; Gong, J. CO2 photo-reduction: Insights into CO2 activation and reaction on surfaces of photocatalysts. Energy Environ. Sci. 2016, 9, 2177–2196. [Google Scholar] [CrossRef]
  7. Tan, L.L.; Ong, W.J.; Chai, S.P.; Mohamed, A.R. Photocatalytic reduction of CO2 with H2O over graphene oxide-supported oxygen-rich TiO2 hybrid photocatalyst under visible light irradiation: Process and kinetic studies. Chem. Eng. J. 2017, 308, 248–255. [Google Scholar] [CrossRef]
  8. Roy, S.C.; Varghese, O.K.; Paulose, M.; Grimes, C.A. Toward solar fuels: Photocatalytic conversion of carbon dioxide to hydrocarbons. ACS Nano 2010, 4, 1259–1278. [Google Scholar] [CrossRef]
  9. Zhang, S.; Yin, X.; Zheng, Y. Enhanced photocatalytic reduction of CO2 to methanol by ZnO nanoparticles deposited on ZnSe nanosheet. Chem. Phys. Lett. 2018, 693, 170–175. [Google Scholar] [CrossRef]
  10. Ye, R.P.; Ding, J.; Gong, W.; Argyle, M.D.; Zhong, Q.; Wang, Y.; Russell, C.K.; Xu, Z.; Russell, A.G.; Li, Q.; et al. Double-slit photoelectron interference in strong-field ionization of the neon dimer. Nat. Commun. 2019, 10, 1. [Google Scholar]
  11. Porosoff, M.D.; Yan, B.; Chen, J.G. Catalytic reduction of CO2 by H2 for synthesis of CO, methanol and hydrocarbons: Challenges and opportunities. Energy Environ. Sci. 2016, 9, 62–73. [Google Scholar] [CrossRef]
  12. Shanmugaratnam, S.; Selvaratnam, B.; Baride, A.; Koodali, R.; Ravirajan, P.; Velauthapillai, D.; Shivatharsiny, Y. SnS2/TiO2 nanocomposites for hydrogen production and photodegradation under extended solar irradiation. Catalysts 2021, 11, 589. [Google Scholar] [CrossRef]
  13. Wang, H.; Zhou, D.; Shen, S.; Wan, J.; Zheng, X.; Yu, L.; Phillips, D.L. The photocatalytic activity and degradation mechanism of methylene blue over copper (ii) tetra (4-carboxyphenyl) porphyrin sensitized TiO2 under visible light irradiation. RSC Adv. 2014, 4, 28978–28986. [Google Scholar] [CrossRef]
  14. Dhakshinamoorthy, A.; Navalon, S.; Corma, A.; Garcia, H. Photocatalytic CO2 reduction by TiO2 and related titanium containing solids. Energy Environ. Sci. 2012, 5, 9217. [Google Scholar] [CrossRef]
  15. Hao, L.; Kang, L.; Huang, H.; Ye, L.; Han, K.; Yang, S.; Yu, H.; Batmunkh, M.; Zhang, Y.; Ma, T. Surface-halogenation-induced atomic-site activation and local charge separation for superb CO2 photoreduction. Adv. Mater. 2019, 31, 1900546. [Google Scholar] [CrossRef]
  16. Chen, F.; Ma, Z.; Ye, L.; Ma, T.; Zhang, T.; Zhang, Y.; Huang, H. Macroscopic spontaneous polarization and surface oxygen vacancies collaboratively boosting CO2 photoreduction on BiOIO3 single crystals. Adv. Mater. 2020, 32, 1908350. [Google Scholar] [CrossRef]
  17. Wang, Z.; Zhou, W.; Wang, X.; Zhang, X.; Chen, H.; Hu, H.; Liu, L.; Ye, J.; Wang, D. Enhanced photocatalytic CO2 reduction over TiO2 using metalloporphyrin as the cocatalyst. Catalysts 2020, 10, 654. [Google Scholar] [CrossRef]
  18. Meng, X.; Wang, T.; Liu, L.; Ouyang, S.; Li, P.; Hu, H.; Kako, T.; Iwai, H.; Tanaka, A.; Ye, J. Photothermal conversion of CO2 into CH4 with H2 over group viii nanocatalysts: An alternative approach for solar fuel production. Angew. Chem. Int. Ed. 2014, 53, 11478–11482. [Google Scholar] [CrossRef]
  19. Liu, H.; Dao, T.; Liu, L.; Meng, X.; Nagao, T.; Ye, J. Light assisted CO2 reduction with methane over group VIII metals: Universality of metal localized surface plasmon resonance in reactant activation. Appl. Catal. B: Environ. 2017, 209, 183–189. [Google Scholar] [CrossRef]
  20. Meng, X.; Liu, L.; Ouyang, S.; Xu, H.; Wang, D.; Zhao, N.; Ye, J. Nanometals for solar-to-chemical energy conversion: From semiconductor-based photocatalysis to plasmon-mediated photocatalysis and photo-thermocatalysis. Adv. Mater. 2016, 28, 6781–6803. [Google Scholar] [CrossRef]
  21. Qi, Y.; Song, L.; Ouyang, S.; Liang, X.; Ning, S.; Zhang, Q.; Ye, J. Photoinduced defect engineering: Enhanced photothermal catalytic performance of 2D black In2O3−x nanosheets with bifunctional oxygen vacancies. Adv. Mater. 2020, 32, 1903915. [Google Scholar] [CrossRef] [PubMed]
  22. Zhang, J.; Li, Y.; Sun, J.; Chen, H.; Zhu, Y.; Zhao, X.; Zhang, L.; Wang, S.; Zhang, H.; Duan, X.; et al. Regulation of energetic hot carriers on Pt/TiO2 with thermal energy for photothermal catalysis. Appl. Catal. B Environ. 2022, 133, 121263. [Google Scholar] [CrossRef]
  23. Wang, W.; An, W.J.; Ramalingam, B.; Mukherjee, S.; Niedzwiedzki, D.M.; Gangopadhyay, S.; Biswas, P. Size and structure matter: Enhanced CO2 photoreduction efficiency by size-resolved ultrafine Pt nanoparticles on TiO2 single crystals. J. Am. Chem. Soc. 2012, 134, 11276–11281. [Google Scholar] [CrossRef]
  24. Liu, Z.; Niu, L.; Zong, X.; An, L.; Qu, D.; Wang, X.; Sun, Z. Ambient photothermal catalytic CO oxidation over a carbon-supported palladium catalyst. Appl. Catal. B Environ. 2022, 313, 121439. [Google Scholar] [CrossRef]
  25. Soltani, T.; Zhu, X.; Yamamoto, A.; Singh, S.P.; Fudo, E.; Tanaka, A.; Kominami, H.; Yoshida, H. Effect of transition metal oxide cocatalyst on the photocatalytic activity of Ag loaded CaTiO3 for CO2 reduction with water and water splitting. Appl. Catal. B Environ. 2021, 52, 119899. [Google Scholar] [CrossRef]
  26. Neatu, S.; Macia-Agullo, J.A.; Concepcion, 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] [PubMed]
  27. AlOtaibi, B.; Fan, S.; Wang, D.; Ye, J.; Mi, Z. Wafer-level artificial photosynthesis for CO2 reduction into CH4 and Co using GaN nanowires. ACS Catal. 2015, 5, 5342–5348. [Google Scholar] [CrossRef]
  28. Kang, Q.; Wang, T.; Li, P.; Liu, L.; Chang, K.; Li, M.; Ye, J. Photocatalytic reduction of carbon dioxide by hydrous hydrazine over Au–Cu alloy nanoparticles supported on SrTiO3/TiO2 coaxial nanotube arrays. Angew. Chem. Int. Ed. 2015, 54, 841–845. [Google Scholar] [CrossRef]
  29. Varghese, O.K.; Paulose, M.; LaTempa, T.J.; Grimes, C.A. High-rate solar photocatalytic conversion of CO2 and water vapor to hydrocarbon fuels. Nano Lett. 2009, 9, 731–737. [Google Scholar] [CrossRef]
  30. 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]
  31. Li, J.; Wang, Z.; Chen, H.; Zhang, Q.; Hu, H.; Liu, L.; Ye, J.; Wang, D. A surface-alkalinized Ti3C2 MXene as an efficient cocatalyst for enhanced photocatalytic CO2 reduction over ZnO. Catal. Sci. Technol. 2021, 11, 4953–4961. [Google Scholar] [CrossRef]
  32. Li, K.; Chen, T.; Yan, L.; Dai, Y.; Huang, Z.; Guo, H.; Jiang, L.; Gao, X.; Xiong, J.; Song, D. Synthesis of mesoporous graphene and tourmaline co-doped titania composites and their photocatalytic activity towards organic pollutant degradation and eutrophic water treatment. Catal. Commun. 2012, 28, 196–201. [Google Scholar] [CrossRef]
  33. Yin, L.; Zhao, M.; Hu, H.; Ye, J.; Wang, D. Synthesis of graphene/tourmaline/TiO2 composites with enhanced activity for photocatalytic degradation of 2-propanol. Chin. J. Catal. 2017, 38, 1307–1314. [Google Scholar] [CrossRef]
  34. Pang, F.; Zhang, R.; Lan, D.; Ge, J. Synthesis of magnetite-semiconductor-metal trimer nanoparticles through functional modular assembly: A magnetically separable photocatalyst with photothermic enhancement for water reduction. ACS Appl. Mater. Interfaces 2018, 10, 4929–4936. [Google Scholar] [CrossRef]
  35. Tan, J.; Wang, X.; Hou, W.; Zhang, X.; Liu, L.; Ye, J.; Wang, D. Fabrication of Fe3O4@graphene/TiO2 nanohybrid with enhanced photocatalytic activity for isopropanol degradation. J. Alloy Compd. 2019, 792, 918–927. [Google Scholar] [CrossRef]
  36. Li, F.; Wang, H.; Huang, R.; Chen, W.; Zhang, H. Recent advances in SnSe nanostructures beyond thermoelectricity. Adv. Funct. Mater. 2022, 32, 2200516. [Google Scholar] [CrossRef]
  37. Ji, Y.; Yang, M.; Dong, H.; Wang, L.; Hou, T.; Li, Y. Monolayer group IVA monochalcogenides as potential and efficient catalysts for the oxygen reduction reaction from first-principles calculations. J. Mater. Chem. A 2017, 5, 1734. [Google Scholar] [CrossRef]
  38. Rajamathi, C.R.; Gupta, U.; Pal, K.; Kumar, N.; Yang, H.; Sun, Y.; Shekhar, C.; Yan, B.; Parkin, S.; Waghmare, U.V.; et al. Photochemical water splitting by bismuth chalcogenide topological insulators. ChemPhysChem 2017, 18, 2322–2327. [Google Scholar] [CrossRef]
  39. Li, X.; Zuo, X.; Jiang, X.; Li, D.; Cui, B.; Liu, D. Enhanced photocatalysis for water splitting in layered tin chalcogenides with high carrier mobility. Phys. Chem. Chem. Phys. 2019, 21, 7559. [Google Scholar] [CrossRef]
  40. Wu, Q.; Wei, W.; Lv, X.; Huang, B.; Dai, Y. Computational screening of defective group IVA monochalcogenides as efficient catalysts for hydrogen evolution reaction. J. Phys. Chem. C 2019, 123, 11791–11797. [Google Scholar] [CrossRef]
  41. Chen, P.; Dai, X.; Xing, P.; Zhao, X.; Zhang, Q.; Ge, S.; Si, J.; Zhao, L.; He, Y. Microwave heating assisted synthesis of novel SnSe/g-C3N4 composites for effective photocatalytic H2 production. J. Ind. Eng. Chem. 2019, 80, 74. [Google Scholar] [CrossRef]
  42. Wu, D.; Guo, J.; Ge, Z.; Feng, J. Facile Synthesis Bi2Te3 based nanocomposites: Strategies for enhancing charge carrier separation to improve photocatalytic activity. Nanomaterials 2021, 11, 3390. [Google Scholar] [CrossRef] [PubMed]
  43. Zhang, N.; Zheng, F.; Huang, B.; Ji, Y.; Shao, Q.; Li, Y.; Xiao, X.; Huang, X. Exploring Bi2Te3 nanoplates as versatile catalysts for electrochemical reduction of small molecules. Adv. Mater. 2020, 32, 1906477. [Google Scholar] [CrossRef]
  44. Shim, X.; Zoum, J.; Chenm, Z. Advanced thermoelectric design: From materials and structures to devices. Chem. Rev. 2020, 120, 7399–7515. [Google Scholar]
  45. Ryu, B. Work function of bismuth telluride: First-principles approach. J. Korean Phys. Soc. 2018, 72, 122–128. [Google Scholar] [CrossRef] [Green Version]
  46. Chen, X.; Liu, L.; Huang, F. Black titanium dioxide (TiO2) nanomaterials. Chem. Soc. Rev. 2015, 44, 1861–1885. [Google Scholar] [CrossRef]
  47. Wu, Y.; Lu, G.; Li, S.J. The doping effect of Bi on TiO2 for photocatalytic hydrogen generation and photodecolorization of rhodamine B. J. Phys. Chem. C. 2009, 113, 9950–9955. [Google Scholar] [CrossRef]
  48. Patil, P.; Mali, S.; Kondalkar, V.; Mane, R.; Patil, P.S.; Hong, C.; Bhosale, P. Bismuth telluride quantum dot assisted titanium oxide microflowers for efficient photoelectrochemical performance. Mater. Lett. 2015, 159, 177–181. [Google Scholar] [CrossRef]
  49. Kim, S.; We, J.; Cho, B. A wearable thermoelectric generator fabricated on a glass fabric. Energy Environ. Sci. 2014, 7, 1959–1965. [Google Scholar] [CrossRef]
  50. Yang, L.; Wang, C.; Lin, S.; Chen, T.; Cao, Y.; Zhang, P.; Liu, X. Thermal conductivity of TiO2 nanotube: A molecular dynamics study. J. Phys.: Condens. Matter 2019, 31, 055302. [Google Scholar] [CrossRef]
  51. Muñoz Rojo, M.; Abad, B.; Manzano, C.; Torres, P.; Cartoixà, X.; Alvarezb, F.; Martín Gonzalez, M. Thermal conductivity of Bi2Te3 nanowires: How size affects phonon scattering. Nanoscale 2017, 9, 6741–6747. [Google Scholar] [CrossRef] [PubMed]
  52. Lu, X.; Jiang, P.; Bao, X. Phonon-enhanced photothermoelectric effect in SrTiO3 ultra-broadband photodetector. Nat. Commun. 2019, 10, 138. [Google Scholar] [CrossRef] [PubMed]
  53. Bell, L. Cooling, heating, generating power, and recovering waste heat with thermoelectric systems. Science 2008, 321, 1457–1461. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Chen, T.; Guai, G.H.; Gong, C.; Hu, W.; Zhu, J.; Yang, H.; Yan, Q.; Li, C. Thermoelectric Bi2Te3-improved charge collection for high-performance dye-sensitized solar cells. Energy Environ. Sci. 2012, 5, 6294–6298. [Google Scholar] [CrossRef]
  55. He, H.; Zhang, C.; Liu, T.; Cao, Y.; Wang, N.; Guo, Z. Thermoelectric–photoelectric composite nanocables induced a larger efficiency in dye-sensitized solar cells. J. Mater. Chem. A 2016, 4, 9362–9369. [Google Scholar] [CrossRef]
  56. Jia, Z.; Ning, S.; Tong, Y.; Chen, X.; Hu, H.; Liu, L.; Ye, J.; Wang, D. Selective photothermal reduction of CO2 to CO over Ni-nanoparticle/N-doped CeO2 nanocomposite catalysts. ACS Appl. Nano Mater. 2021, 4, 10485–10494. [Google Scholar] [CrossRef]
  57. Li, Z.; Liu, J.; Shi, R.; Waterhouse, G.I.N.; Wen, X.; Zhang, T. Fe-based catalysts for the direct photohydrogenation of CO2 to value-added hydrocarbons. Adv. Energy Mater. 2021, 11, 2002783. [Google Scholar] [CrossRef]
  58. Han, B.; Wei, W.; Chang, L.; Cheng, P.; Hu, Y. Efficient visible light photocatalytic CO2 reforming of CH4. ACS Catal. 2016, 6, 494–497. [Google Scholar] [CrossRef]
  59. Zhang, X.; Wang, X.; Wang, D.; Ye, J. Conformal BiVO4-layer/WO3-nanoplate-array heterojunction photoanode modified with cobalt phosphate cocatalyst for significantly enhanced photoelectrochemical performances. ACS Appl. Mater. Interfaces 2019, 11, 5623–5631. [Google Scholar] [CrossRef]
  60. Chen, Y.; Chen, H.; Ku, S.; Luo, C.; Lin, J.; Wu, K.; Juang, J.; Kobayashi, T.; Cheng, C.; Tsuei, K.; et al. Manifestation of a second dirac surface state and bulk bands in THz radiation from topological insulators. Sci. Rep. 2015, 5, 14128. [Google Scholar]
  61. Subramanian, V.; Wolf, E.E.; Kamat, P.V. Catalysis with TiO2/Gold nanocomposites. Effect of metal particle size on the fermi level equilibration. J. Am. Chem. Soc. 2004, 126, 4943–4950. [Google Scholar] [CrossRef] [PubMed]
  62. Jakob, M.; Levanon, H.; Kamat, P.V. Charge distribution between UV-Irradiated TiO2 and gold nanoparticles:  Determination of shift in the fermi level. Nano Lett. 2003, 3, 353–358. [Google Scholar] [CrossRef]
  63. Liu, J.; Zhang, Y.; Wang, Z.; Li, M.; Su, W.; Zhao, M.; Huang, S.; Xia, S.; Wang, C. Accurate measurement of Seebeck coefficient. Rev. Sci. Instrum. 2016, 87, 064701. [Google Scholar] [CrossRef] [PubMed]
  64. Li, Y.; Walsh, A.G.; Li, D.; Do, D.; Ma, H.; Wang, C.; Zhang, P.; Zhang, X. W-Doped TiO2 for photothermocatalytic CO2 reduction. Nanoscale 2020, 12, 17245–17252. [Google Scholar] [CrossRef]
  65. Li, G.; Sun, Y.; Zhang, Q.; Gao, Z.; Sun, W.; Zhou, X. Ag quantum dots modified hierarchically porous and defective TiO2 nanoparticles for improved photocatalytic CO2 reduction. Chem. Eng. J. 2021, 410, 128397. [Google Scholar] [CrossRef]
  66. Sun, Y.; Li, G.; Gong, Y.; Sun, Z.; Yao, H.; Zhou, X. Ag and TiO2 nanoparticles co-modified defective zeolite TS-1 for improved photocatalytic CO2 reduction. J. Hazard. Mater. 2021, 403, 124019. [Google Scholar] [CrossRef]
  67. Xing, M.; Zhou, Y.; Dong, C.; Cai, L.; Zeng, L.; Shen, B.; Pan, L.; Dong, C.; Chai, Y.; Zhang, J.; et al. Modulation of the reduction potential of TiO2−x by fluorination for efficient and selective CH4 generation from CO2 photoreduction. Nano Lett. 2018, 18, 3384–3390. [Google Scholar] [CrossRef]
  68. Feng, S.; Zhao, J.; Bai, Y.; Liang, X.; Wang, T.; Wang, C. Facile synthesis of Mo-doped TiO2 for selective photocatalytic CO2 reduction to methane: Promoted H2O dissociation by Mo doping. J. CO2 Util. 2020, 38, 1–9. [Google Scholar] [CrossRef]
  69. Bian, J.; Qu, Y.; Zhang, X.; Sun, N.; Tang, D.; Jing, L. Dimension-matched plasmonic Au/TiO2/BiVO4 nanocomposites as efficient wide-visible-light photocatalysts to convert CO2 and mechanistic insights. J. Mater. Chem. A 2018, 6, 11838–11845. [Google Scholar] [CrossRef]
  70. 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]
  71. He, F.; Zhu, B.; Cheng, B.; Yu, J.; Ho, W.; Macyk, W. 2D/2D/0D TiO2/C3N4/Ti3C2 MXene composite S-scheme photocatalyst with enhanced CO2 reduction activity. Appl. Catal. B 2020, 272, 119006. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns of P25, pBT, and pBT(2)/P25; (b) HRTEM image of pBT(2)/P25.
Figure 1. (a) XRD patterns of P25, pBT, and pBT(2)/P25; (b) HRTEM image of pBT(2)/P25.
Catalysts 12 01323 g001
Figure 2. (a) UV–vis–NIR spectra of pBT, pBT(x)/P25 (x = 1, 2, 3, 4), and P25; (b) In situ monitoring of the photothermal temperatures over P25, pBT(x)/P25, and pBT.
Figure 2. (a) UV–vis–NIR spectra of pBT, pBT(x)/P25 (x = 1, 2, 3, 4), and P25; (b) In situ monitoring of the photothermal temperatures over P25, pBT(x)/P25, and pBT.
Catalysts 12 01323 g002
Figure 3. Simulated (a) temperature distribution and (b) electric potential over pBT(2)/P25 nanocomposite upon a single-wavelength (λ = 800 nm) light irradiation.
Figure 3. Simulated (a) temperature distribution and (b) electric potential over pBT(2)/P25 nanocomposite upon a single-wavelength (λ = 800 nm) light irradiation.
Catalysts 12 01323 g003
Figure 4. (a) CO and (b) O2 evolution rates over P25, pBT(x)/P25, and pBT under a 300 W full-arc Xe lamp illumination.
Figure 4. (a) CO and (b) O2 evolution rates over P25, pBT(x)/P25, and pBT under a 300 W full-arc Xe lamp illumination.
Catalysts 12 01323 g004
Figure 5. (a) UV-vis-IR spectra of pBT, nBT, pBT(2)/P25, and nBT(2)/P25; (b) In situ monitoring of the photothermal temperatures over pBT(2)/P25 and nBT(2)/P25; (c) CO evolution rates over pBT(2)/P25 and nBT(2)/P25 under a 300 W full-arc Xe lamp illumination.
Figure 5. (a) UV-vis-IR spectra of pBT, nBT, pBT(2)/P25, and nBT(2)/P25; (b) In situ monitoring of the photothermal temperatures over pBT(2)/P25 and nBT(2)/P25; (c) CO evolution rates over pBT(2)/P25 and nBT(2)/P25 under a 300 W full-arc Xe lamp illumination.
Catalysts 12 01323 g005
Figure 6. (a) Transient photocurrent responses of P25 and pBT(x)/P25 under a 300 W full-arc Xe lamp irradiation; and (b) photoluminescence spectra of P25 and pBT(x)/P25 with an excitation wavelength of λ = 330 nm.
Figure 6. (a) Transient photocurrent responses of P25 and pBT(x)/P25 under a 300 W full-arc Xe lamp irradiation; and (b) photoluminescence spectra of P25 and pBT(x)/P25 with an excitation wavelength of λ = 330 nm.
Catalysts 12 01323 g006
Figure 7. Enhancement mechanism for CO2 photoreduction over pBT/P25 upon light irradiation.
Figure 7. Enhancement mechanism for CO2 photoreduction over pBT/P25 upon light irradiation.
Catalysts 12 01323 g007
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lei, Y.; Jia, Z.; Hu, H.; Liu, L.; Ye, J.; Wang, D. Enhanced CO2 Photoreduction over Bi2Te3/TiO2 Nanocomposite via a Seebeck Effect. Catalysts 2022, 12, 1323. https://doi.org/10.3390/catal12111323

AMA Style

Lei Y, Jia Z, Hu H, Liu L, Ye J, Wang D. Enhanced CO2 Photoreduction over Bi2Te3/TiO2 Nanocomposite via a Seebeck Effect. Catalysts. 2022; 12(11):1323. https://doi.org/10.3390/catal12111323

Chicago/Turabian Style

Lei, Yiming, Zewei Jia, Huilin Hu, Lequan Liu, Jinhua Ye, and Defa Wang. 2022. "Enhanced CO2 Photoreduction over Bi2Te3/TiO2 Nanocomposite via a Seebeck Effect" Catalysts 12, no. 11: 1323. https://doi.org/10.3390/catal12111323

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