Next Article in Journal
Au@CdS Nanocomposites as a Visible-Light Photocatalyst for Hydrogen Generation from Tap Water
Next Article in Special Issue
Synergetic Adsorption–Photocatalytic Activated Fenton System via Iron-Doped g-C3N4/GO Hybrid for Complex Wastewater
Previous Article in Journal
The Pincer Ligand Supported Ruthenium Catalysts for Acetylene Hydrochlorination: Molecular Mechanisms from Theoretical Insights
Previous Article in Special Issue
Fabrication of a Plasmonic Heterojunction for Degradation of Oxytetracycline Hydrochloride and Removal of Cr(VI) from Water
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Photothermal Catalytic Reduction of CO2 by Cobalt Silicate Heterojunction Constructed from Clay Minerals

1
Jiangsu Key Laboratory of Materials Surface Science and Technology, Changzhou Key Laboratory of Biomass Green, Safe and High Value Utilization, Changzhou University, Changzhou 213164, China
2
Changzhou Railway Higher Vocational and Technical School, Changzhou 213011, China
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(1), 32; https://doi.org/10.3390/catal13010032
Submission received: 14 November 2022 / Revised: 14 December 2022 / Accepted: 20 December 2022 / Published: 24 December 2022
(This article belongs to the Special Issue Synthesis and Application of Composite Photocatalysts)

Abstract

:
The coupled utilization of solar and thermal energy is considered an efficient way to improve the efficiency of CO2 reduction. Herein, palygorskite (Pal) clay is as a silicon source, while Co2+ is introduced to prepare two-dimensional Co2SiO4 nanosheets, and the excess of Co2+ leads to the growth of Co3O4 on the surface of Co2SiO4 to obtain an S-scheme Co2SiO4/Co3O4−x heterojunction, which facilitates the charge transfer and maintains higher redox potentials. Benefiting from black color and a narrow band gap, the cobalt oxide on the surface can increase the light absorption and produce a local photothermal effect. Under proper thermal activation conditions, the photoelectrons captured by the abundant oxygen vacancies can obtain a secondary leap to the semiconductor conduction band (CB), suppressing the recombination of electron-hole pairs, thus favoring the electron transfer on Co2SiO4/Co3O4−x. The composites not only have abundant oxygen vacancies, but also have a large specific surface area for the adsorption and activation of CO2. The yields of CH3OH on Co2SiO4/Co3O4−5% reach as high as 48.9 μmol·g−1·h−1 under simulated sunlight irradiation. In situ DRIFTS is used to explore the photocatalytic reduction CO2 mechanism. It is found that the thermal effect facilitates the generation of the key intermediate COOH* species. This work provides a new strategy for photothermal catalytic CO2 reduction by taking advantage of natural clay and solar energy.

1. Introduction

With the rapid development of the fossil energy economy, CO2 emissions are enormous, which disrupts the balance of nature and causes climate change and serious ecological damage in human society [1]. Therefore, research on CO2 reduction using renewable fuel technology has received widespread attention [2,3].
Inspired by photosynthetic organisms in nature, visible light reduction is used to convert CO2 into chemicals with high added value and solar fuels (e.g., CO, HCOOH, CH3OH, CH4, etc.) [4]. For example, Lin [5] et al. decomposed CO2 into CO with an apparent quantum yield of 0.25% at 420 nm by using graphitic carbon nitride (g-CN) semiconductors. However, the yield of conventional photocatalytic reduction of CO2 is generally low, and the difficulties are mainly related to the limited utilization of solar energy and the low efficiency of photogenerated electron/hole separation [6,7]. Recent studies have shown that the introduction of heat in some photocatalytic reactions can effectively enhance reaction performance [8].
Cobalt-based catalysts with full-spectrum absorption performance and multiple valence states are widely used in CO2 reduction [9]. The modification of Pal clay can significantly change its physical and chemical properties, such as acidification, alkalization, and ion exchange [10]. In our previous work, a unique solar driven catalysis system was developed to convert cellulose into lactic acid using Cu-modified natural palygorskite (Pal) catalyst [11]. Liu [12] et al. reported on one-dimensional Pal treated with acid to remove most of the metal ions from the silica framework, and bismuth ions were incorporated to grow a two-dimensional (2D) bismuth silicate (Bi12SiO20) nanosheet under a microwave-hydrothermal process. Metal silicates can be prepared from silica and are often used in cement and glass processing, as well as in manufacturing [13].
Herein, acid modification of 1D Pal was performed to remove most of the metal ions in the Pal skeleton and retain its unique silicon-oxygen tetrahedral structure. A 2D cobalt silicate Co2SiO4 was synthesized by the microwave hydrothermal method, in which one-dimensional rod silicate was converted to two-dimensional flake silicate. Cobalt oxide particles were grown on the surface of Co2SiO4 by adjusting the amount of Co2+ in Co2SiO4. Co2SiO4/Co3O4 catalysts not only have a large specific surface area, but they also have abundant oxygen vacancies for the adsorption and activation of CO2, which effectively achieves full-spectrum absorption and light-induced heat under solar light.

2. Result and Discussion

2.1. XRD

Figure 1a shows the XRD spectra of the Pal, SiO2 Co3O4, Co2SiO4, and Co2SiO4/Co3O4 composites. The peaks located at 8.5°, 20.1°, 28.0°, and 35.8°correspond to the (110), (040), (400), and (161) crystal planes of the original Pal. Compared with the original Pal, the characteristic peaks of acid-treated Pal basically disappeared, indicating that the metal ions in Pal were largely dissolved after a long treatment with high concentrations of hydrochloric acid, corresponding to the SiO2. The diffraction peaks appearing at 18.8°, 31.0°, and 36.5°, are ascribed to (111), (220), and (311), crystal planes of Co2SiO4 (PDF#29-0508). The characteristic peaks at 31.3°, 36.8°, 44.8°, and 65.2° are consistent with the characteristic peaks of cubic spinel Co3O4 (PDF#42-1467). It is noteworthy that the diffraction peaks of Co3O4 are very close to Co2SiO4. Figure 1b shows the magnified view of Figure 1a from 36.5° to 39.5°. The peaks of Co2SiO4/Co3O4 composites are obviously shifted to a higher angle compared with Co3O4 and Co2SiO4, indicating a stronger interaction between Co2SiO4 and Co3O4 [12,14].

2.2. TEM

Figure 2 shows the TEM picture of the Pal, Co3O4, Co2SiO4, and Co2SiO4/Co3O4 composites. As shown in Figure 2a, the morphology of Pal is a one-dimensional nanofiber structure with a diameter of about 30–50 nm. After a high concentration of acid treatment, the structure of Pal is destroyed, and the rod-like structure tends to break and assemble into sheets. Figure 2b,f indicates TEM and HRTEM of Co2SiO4. The microstructure of Co2SiO4 is a two-dimensional flake, which may be due to the preservation of the silicon-oxygen tetrahedron structure in the skeleton after the destruction of Pal [12]. After the addition of Co2+, the chain-like structure of the silicon-oxygen tetrahedron regularly grows into two-dimensional cobalt silicate flakes under alkaline hydrothermal conditions. The lattice spacing of the Co2SiO4 is 0.09 nm, corresponding to the (400) crystal plane of Co2SiO4. Figure 2c,g displays TEM and HRTEM of Co2SiO4/Co3O4−1%, respectively. It can be seen that, on the surface of Co2SiO4, the Co3O4 particles have a diameter of about 10 nm. Figure 2d,h shows TEM and HRTEM of Co2SiO4/Co3O4−5%; Co3O4 particles on the surface of Co2SiO4 become more stacked. However, when the amount of Co reaches 10%, it is difficult to distinguish Co2SiO4 and Co3O4, as shown in Figure 2e. The HRTEM images (Figure 2g,h) show that the Co3O4 nanoparticles are uniformly loaded on the surface of the Co2SiO4 2D nanosheets. The lattice spacing of Co3O4 on the surface is 0.32 nm and 0.33 nm, corresponding to the (311) and (400) planes of Co3O4, suggesting the existence of a heterogeneous structure between Co2SiO4 and Co3O4.

2.3. UV-Vis-NIR

Figure 3a shows the UV-Vis-NIR absorption spectra of the Pal, Co3O4, Co2SiO4, and Co2SiO4/Co3O4 composites. Pal has a weak UV-Vis-NIR absorption edge of about 380 nm, while Co2SiO4 and Co2SiO4/Co3O4 composites obtained by modifying Pal have full-spectrum absorption to achieve the conversion of UV-responsive natural silicate minerals to visible-responsive silicates. Most of the cobalt-based compounds appear as black or dark green, which can absorb the full range of the spectrum and are ideal photothermal materials [8,15]. The band gaps of Co2SiO4 and Co3O4 in Figure 3b are estimated by the formula of (ahv)n/2 = A (hv − Eg); where a, h, v, A, and Eg represent the optical absorption coefficient, Planck constant, optical frequency, a constant, and band gap. The values n = 1 and n = 4 represent the indirect and direct band gap semiconductors, respectively [16,17]. As shown in Figure 3b, the corresponding band gap energies of Co2SiO4 and Co3O4 are estimated to be 3.39 and 1.53 eV, respectively. The narrow band gap is more favorable for capturing the energy of sunlight, thus promoting the generation of e/h+ pairs [18].

2.4. XPS

Figure 4a shows the XPS survey of Co3O4, Co2SiO4, and Co2SiO4/Co3O4−x, where the presence of Co and O elements can be determined; Si elements are present in the Co2SiO4 and Co2SiO4/Co3O4−x. Figure 4b shows the Co 2p spectra of Co3O4, Co2SiO4, and Co2SiO4/Co3O4−x. Co can be divided into four peaks, the peaks with binding energies at 794.59 eV and 795.96 eV correspond to Co 2p1/2. The peaks with binding energies at 779.77 eV and 781.68 eV correspond to Co 2p3/2, suggesting the presence of +2 valence and +3 valence of Co3O4, and also appear in the Co2SiO4 and Co2SiO4/Co3O4−x [19]. The peaks near the binding energy of 781.06 eV and 797.23 eV in (Figure 4b) correspond to Co 2p1/2 and Co 2p3/2, respectively. Nevertheless, the binding energy of Co2SiO4/Co3O4−x is higher than Co2SiO4, which indicates that Co2SiO4 has a strong interaction with the additional generated Co3O4, increasing the electron cloud density around Co [20]. The studies show that the binding energy of Co 2p in the Co2SiO4 structure is 781.3 eV, while the binding energy of Co 2p in Co3O4 is 779.6 eV. Obviously, the binding energy of cobalt is closer to the former, indicating that the product is Co2SiO4, rather than Co3O4, which is consistent with the results of the XRD analysis. In addition, the energy separation between Co 2p1/2 and Co 2p3/2 of the product is 15.96 eV, which is close to Co2SiO4 (ΔCo 2p = 15.5 eV) [21]. Figure 4c shows the O 1s spectra of Co3O4, Co2SiO4, and Co2SiO4/Co3O4−x. The peaks around the binding energy of 529.85 eV and 532.82 eV correspond to the lattice oxygen (OL) and adsorbed oxygen (OA), respectively. The ratio of OA to OL in Co2SiO4/Co3O4−x is higher than that of Co3O4, indicating the composite Co2SiO4/Co3O4−x has higher content of oxygen vacancy than that of Co3O4 [22].
Figure 4d shows the Si 2p spectra of Co2SiO4 and Co2SiO4/Co3O4−x. The Si 2p of Co2SiO4 is located at the binding energy of 102.75. The Si 2p characteristic peaks of Co2SiO4/Co3O4 composites exhibit lower binding energy and weaker peaks than Co2SiO4, which may be due to the formation of Si-O-Co bonds on the surface [23].

2.5. EIS, Mott–Schottky Plot, and VB-XPS

Figure 5a indicates that the radius of the electrochemical impedance spectroscopy (EIS) spectrum of the Co2SiO4/Co3O4−5% sample is far less than other counterparts, indicating that the value of resistance is lowest, and the transfer of surface charges is fastest [24,25]. Figure 5b–d exhibits the Mott–Schottky patterns of Co2SiO4, Co3O4, and Co2SiO4/Co3O4−x. The flat band potentials (Efb) are 0.79, −0.32, and 0.51 V (vs. Ag/AgCl, pH = 7), respectively, suggesting that their Fermi energy levels are 0.99, −0.12, and 0.71 V (vs. normal hydrogen electrode (NHE), pH = 7) [26]. Generally, valence band (VB) XPS represents the distance between the valence band and the Fermi level [27]. The VB-XPS of Co2SiO4 and Co3O4 are shown in Figure 5e,f, which are 1.34 and 0.43 eV. Therefore, the valence bands (EVB) of Co2SiO4 and Co3O4 are 2.33 and 0.31 V, respectively. According to ECB = EVB − Eg, the conduction bands (CBs) of Co2SiO4 and Co3O4 are calculated to be −1.06 and −1.22 V, respectively.

2.6. In-Situ DRIFTS

The reaction pathway of Co2SiO4/Co3O4−x was investigated by in situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS). The samples are exposed to simulated solar light irradiation (300 W Xe lamp) for 45 minutes, and exporting spectrograms every 5 min. The peak at 2360 cm−1 can be attributed to the symmetric stretching vibration modes of CO2 (Figure 6a) [28]. Figure 6b shows the magnified view of Figure 6a from 1750 cm−1 to 1250 cm−1. The peaks from 1600 to 1750 cm−1 correspond to the stretching vibration of C–O. The peaks at 1560 and 1540 cm−1 can be ascribed to COOH* species. The peaks at 1455 and 1340 cm−1 relate to the O–H bending and C–O stretching vibration of COOH* groups, respectively [29,30]. It is observed that the peak of COOH* species becomes stronger with increasing light-irradiation time, indicating that the COOH* is one of the key intermediate species for CO2-to-CH3OH conversion [31]. Moreover, the signal of CH2* is detected at 1385 cm−1, confirming the hydrogenation of HCHO* [12]. The IR results show that the photocatalytic reduction of CO2 on the Co2SiO4/Co3O4−x surface goes through a sequential process in which COO− and HCHO act as reaction intermediates to generate methanol groups [32]. According to the DRIFTS analysis, a possible CO2 reduction pathway is summarized in the following reactions [33]:
Co2SiO4/Co3O4−x → e + h+
H2O + h+ → H+ + OH
CO2 + e → CO2*
CO2* + H+ + e → COOH*
COOH* + H+ + e → HCOOH* or CO + H2O
HCOOH* + 2H+ + 2 e → HCHO* + H2O
HCHO + 2H+ + 2 e→ CH3OH

2.7. CO2 Reduction Performance

The photothermal catalytic reduction of CO2 performance was performed under solar light. The photothermal catalytic CH3OH generation rate of Co3O4, Co2SiO4, and Co2SiO4/Co3O4−x is shown in Figure 7a. For Co2SiO4/Co3O4−x photocatalysts, the CH3OH generation rate increases with the increase in Co3O4, but excessive Co3O4 may inhibit the photocatalytic reduction activity. Co2SiO4/Co3O4−5% expresses the best CO2 reduction rates of CH3OH and CH4, corresponding to 48.9 and 14.1 μmol·g−1·h−1, which is better than what has been recently reported. After 4 h of sunlight exposure each time, the samples were washed with distilled water for the next test. The samples could maintain considerable photoactivity without significant reduction in activity after five cycling tests (Figure 7b).

2.8. Schematic Mechanism of Photothermal Catalytic Conversion

Figure 8 demonstrates the photothermal catalytic CO2 reduction mechanism of Co2SiO4/Co3O4−x composites. Figure 8a shows the before-contact configuration of the band structures of Co2SiO4 and Co3O4−x. Co2SiO4 has a lower Fermi level than Co3O4−x (Ef = 0.99), and the free electrons of Co3O4−x can be transferred to Co2SiO4 until the Fermi reaches equilibrium (Ef = −0.12). After contact, the band edge of Co2SiO4 with a higher Fermi level bends downward. In comparison, the band edge of Co3O4−x with a higher Fermi level bends upward, thus forming a built-in electric field, facilitating the recombination of electrons in Co2SiO4 and holes in Co3O4−x. The Co2SiO4/Co3O4−x catalyst has an ultrathin two-dimensional structure with abundant sites for physisorption of CO2, and abundant oxygen vacancies for chemisorption and activation of CO2, which facilitates the reduction reaction. Benefiting from a black color and narrow band gap, Co3O4−x can increase light absorption and produce a local photothermal effect. Under proper heat activation, the photoelectrons captured by the abundant oxygen vacancies can obtain a secondary leap to the semiconductor conduction band (CB), suppressing the recombination of electron-hole pairs to a certain extent; thus, the energy-rich environment favors the electron transfer on Co2SiO4/Co3O4−x. Under solar light, Co2SiO4/Co3O4−x has full-spectrum absorption and can be stimulated to produce e and h+ pairs. The photogenerated e in the CB of Co2SiO4 and the photogenerated h+ in the VB of Co3O4−x recombines, facilitated by the built-in electric field, thus establishing an S-scheme process. Since ECB of Co3O4 is more negative than CO2/CH3OH (−0.38 eV), the e in the CB position of Co3O4−x is kept converting CO2 to methanol. Moreover, Co2SiO4 has a more positive VB potential energy than H2O/H+ (1.23 eV). The photogenerated holes in the VB of Co2SiO4 oxidize H2O to O2 and H+ ions (Figure 8b).

3. Experimental Section

3.1. Materials and Chemicals

Raw Pal was obtained from Xuyi, China. Cobalt nitrate (Co(NO3)2·6H2O), ammonium sulfate ((NH4)2SO4), ammonium chloride (NH4Cl), ammonia (NH3·H2O), and hydrochloric acid (HCl) were provided by Shanghai Lingfeng chemical company.

3.2. Preparation of SiO2

An amount of 10 g of Pal was mixed with 30 g of ammonium sulfate and placed in a ceramic crucible. The mixtures were calcined in a muffle furnace at 500 °C for 2 h. An amount of 200 mL 0.5 M HCl was added with refluxing and stirring in a water bath at 80 °C for 6 h. After washing and drying, the SiO2 powders were obtained.

3.3. Preparation of Co2SiO4 and Co2SiO4/Co3O4

An amount of 0.3 g modified Pal was mixed with 10 mmol Co(NO3)2·6H2O in a beaker; 20 mmol NH4Cl was added after 2 h, and then 1 mL ammonia was added drop by drop to adjust the pH. The solution was transferred into a 100 mL microwave hydrothermal reactor for hydrothermal treatment at 160 °C for 90 min. After the reaction, the samples were centrifuged, washed, dried overnight, and the Co2SiO4 was obtained.
The preparation method of Co2SiO4/Co3O4−x composite material was the same as Co2SiO4, except the quantity of Co(NO3)2·6H2O was adjusted to deposit Co3O4. The different mass ratios in the composite material were denoted as Co2SiO4 /Co3O4−x; where x represents the theoretical mass fraction of Co3O4 in the composite from 1% to 10%. The preparation of pure Co3O4 was the same as this method without the addition of SiO2.

3.4. Catalyst Characterizations

The XRD was used to measure the phase structure of materials on a Rigaku D/max 2500PC (Rigaku Corporation, Tokyo, Japan) diffractometer using Cu-Kα radiation (λ = 1.5406 Å) with a scanning range of 2θ from 5° to 80°. The TEM was operated on a JEOL-2100 (Japan Electronics Co., Ltd., Tokyo, Japan) with 200 kV working voltage. The UV-vis-NIR was carried out to measure the optical properties of materials with an integrating sphere (UV-3600, DRS, SHIMADZU, Kyoto, Japan). The XPS was performed on a Quantum 2000 Scanning ESCA Microprobe instrument (Thermo Nicolet Evolution, Thermo Fisher, Waltham, MA, USA). The photocurrent response and Mott-Schottky measured the optical properties of materials using an electrochemical workstation equipped with a 300 W Xe lamp and standard three electrodes (Auto Lab 302N, CH Instruments, Shanghai, China). In situ DRIFTS was conducted by in situ FT-IR spectrometer (Nicolet iS20 FT-IR, ThermoFisher, USA), with a specific designed reaction cell and MCT detector with 120 scans and a resolution of 4 cm−1.

3.5. Photothermal Catalytic CO2 Reduction Experiments

The photothermocatalytic activity of the samples was carried out in a closed photochemical reactor coupled with a condensation system and gas chromatography. First, 0.1 g of catalysts were added into 100 mL of water in the bottle in the reactor. After sealing, the reactor was infiltrated with CO2 gas (99.99%) for 45 min to ensure that the reactor was filled with pure CO2 and to remove the impurity gasses. Then, at room temperature, a xenon lamp was irradiated through the transparent window (diameter 30~40 mm) on the top of the visible reactor. During the reaction process, the temperature in the reactor was monitored, the reaction sampling interval was 30 min, while the mixture product was obtained from the photochemical reactor with a syringe, and the gaseous product was taken out with a gastight needle. The reaction product was detected by a gas chromatograph (GC-7860 Plus, Shanghai Xinuo Instrument Co., Ltd., Shanghai, China) with a flame ionization detector (FID) and a thermal conductivity detector (TCD).

4. Conclusions

In summary, a novel cobalt silicate nanosheet composite (Co2SiO4/Co3O4−x) was prepared using Pal clay as a silicon source. The acid-modified Pal removed most of the metal ions and retained its unique silicon-oxygen tetrahedral structure, then Co2+ were incorporated to acquire Co2SiO4 nanosheet and Co2SiO4/Co3O4−x heterojunction, which realized the conversion of one-dimensional rod silicate to two-dimensional flake silicate. Co2SiO4/Co3O4−x not only have abundant oxygen vacancies, but also have large specific surface areas for the adsorption and activation of CO2. The yields of CH3OH on Co2SiO4/Co3O4−5% were as high as 48.9 μmol·g−1·h−1 under simulated sunlight irradiation. In addition, in situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) was used to explore the photocatalytic reduction CO2 mechanism. Co2SiO4/Co3O4−x effectively and stably achieved full-spectrum absorption and light-induced heat under solar light. The photothermal synergistic effect facilitated the generation of key intermediate COOH* species to enhance methanol production. This work provides a new approach for photothermal catalytic CO2 reduction by taking advantage of natural clay and solar energy.

Author Contributions

Methodology, X.K.; Software, M.F.; Validation, Z.Z.; Investigation, S.Q.; Data curation, C.G.; Writing—original draft, S.Q.; Writing—review & editing, X.L.; Project administration, X.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Jiangsu Province Key Laboratory of Materials Surface Science and Technology and Jiangsu High Institutions Key Basic Research Projects of Natural Science (21KJA430002). We thank the Analysis and Testing Center of Changzhou University for the characterization.

Data Availability Statement

Data can be available on request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zuo, S.; Zhang, H.; Li, X.; Han, C.; Yao, C.; Ni, C. Dual Active Sites Boosting Photocatalytic Nitrogen Fixation over Upconversion Mineral Nanocomposites under the Full Spectrum. ACS Sustain. Chem. Eng. 2022, 10, 1440–1450. [Google Scholar] [CrossRef]
  2. Zhai, Y.; Han, P.; Yun, Q.; Ge, Y.; Zhang, X.; Chen, Y.; Zhang, H. Phase engineering of metal nanocatalysts for electrochemical CO2 reduction. eScience 2022, 2, 467–485. [Google Scholar] [CrossRef]
  3. Navarro-Jaén, S.; Virginie, M.; Bonin, J.; Robert, M.; Wojcieszak, R.; Khodakov, A.Y. Highlights and challenges in the selective reduction of carbon dioxide to methanol. Nat. Rev. Chem. 2021, 5, 564–579. [Google Scholar] [CrossRef]
  4. Zhao, J.; Shi, R.; Waterhouse, G.I.; Zhang, T. Selective photothermal CO2 reduction to CO, CH4, alkanes, alkenes over bimetallic alloy catalysts derived from layered double hydroxide nanosheets. Nano Energy 2022, 102, 107650. [Google Scholar] [CrossRef]
  5. Lin, J.; Pan, Z.; Wang, X. Photochemical Reduction of CO2 by Graphitic Carbon Nitride Polymers. ACS Sustain. Chem. Eng. 2013, 2, 353–358. [Google Scholar] [CrossRef]
  6. Xu, C.; Anusuyadevi, P.R.; Aymonier, C.; Luque, R.; Marre, S. Nanostructured materials for photocatalysis. Chem. Soc. Rev. 2019, 48, 3868–3902. [Google Scholar] [CrossRef]
  7. Liu, Y.; Zhang, C.; Shi, A.; Zuo, S.; Yao, C.; Ni, C.; Li, X. Full solar spectrum driven CO2 conversion over S-Scheme natural mineral nanocomposite enhanced by LSPR effect. Powder Technol. 2021, 396, 615–625. [Google Scholar] [CrossRef]
  8. Gu, Y.; Ding, J.; Tong, X.; Yao, H.; Yang, R.; Zhong, Q. Photothermal catalyzed hydrogenation of carbon dioxide over porous nanosheet Co3O4. J. CO2 Util. 2022, 61, 102003. [Google Scholar] [CrossRef]
  9. Zhang, Q.; Yang, P.; Zhang, H.; Zhao, J.; Shi, H.; Huang, Y.; Yang, H. Oxygen vacancies in Co3O4 promote CO2 photoreduction. Appl. Catal. B Environ. 2021, 300, 120729. [Google Scholar] [CrossRef]
  10. Sun, L.; Ye, X.; Cao, Z.; Zhang, C.; Yao, C.; Ni, C.; Li, X. Upconversion enhanced photocatalytic conversion of lignin biomass into valuable product over CeVO4/palygorskite nanocomposite: Effect of Gd3+ incorporation. Appl. Catal. A Gen. 2022, 648, 118923. [Google Scholar] [CrossRef]
  11. Zhong, M.; Li, X.; Chu, X.; Gui, H.; Zuo, S.; Yao, C.; Li, Z.; Chen, Y. Solar driven catalytic conversion of cellulose biomass into lactic acid over copper reconstructed natural mineral. Appl. Catal. B Environ. 2022, 317, 121718. [Google Scholar] [CrossRef]
  12. Liu, Y.; Chu, X.; Shi, A.; Yao, C.; Ni, C.; Li, X. Construction of 2D Bismuth Silicate Heterojunctions from Natural Mineral toward Cost-Effective Photocatalytic Reduction of CO2. Ind. Eng. Chem. Res. 2022, 61, 12294–12306. [Google Scholar] [CrossRef]
  13. Zhu, J.; Xia, L.; Yang, W.; Yu, R.; Zhang, W.; Luo, W.; Dai, Y.; Wei, W.; Zhou, L.; Zhao, Y.; et al. Activating Inert Sites in Cobalt Silicate Hydroxides for Oxygen Evolution through Atomically Doping. Energy Environ. Mater. 2021, 5, 655–661. [Google Scholar] [CrossRef]
  14. Long, D.; Li, X.; Yin, Z.; Fan, S.; Wang, P.; Xu, F.; Wei, L.; Tadé, M.O.; Liu, S. Novel Co3O4 @ CoFe2O4 double-shelled nanoboxes derived from Metal–Organic Framework for CO2 reduction. J. Alloys Compd. 2020, 854, 156942. [Google Scholar] [CrossRef]
  15. Benchikhi, M.; Hattaf, R.; El Ouatib, R. Sol-Gel-Assisted Molten-Salt Synthesis of Co2SiO4 Pigments for Ceramic Tiles Application. Silicon 2022, 22, 1–8. [Google Scholar] [CrossRef]
  16. Huang, J.; Ye, X.; Li, W.; Shi, A.; Chu, X.; Cao, Z.; Yao, C.; Li, X. Infrared-to-visible upconversion enhanced photothermal catalytic degradation of toluene over Yb3+, Er3+: CeO2/attapulgite nanocomposite: Effect of rare earth doping. J. Ind. Eng. Chem. 2022, 116, 504–514. [Google Scholar] [CrossRef]
  17. Gao, R.; Shi, A.; Cao, Z.; Chu, X.; Wang, A.; Lu, X.; Yao, C.; Li, X. Construction of 2D up-conversion calcium copper silicate nanosheet for efficient photocatalytic nitrogen fixation under full spectrum. J. Alloys Compd. 2022, 910, 164869. [Google Scholar] [CrossRef]
  18. Liu, W.; Li, X.; Chu, X.; Zuo, S.; Gao, B.; Yao, C.; Li, Z.; Chen, Y. Boosting photocatalytic reduction of nitrate to ammonia enabled by perovskite/biochar nanocomposites with oxygen defects and O-containing functional groups. Chemosphere 2022, 294, 133763. [Google Scholar] [CrossRef]
  19. Hassen, D.; Shenashen, M.; El-Safty, A.R.; Elmarakbi, A.; El-Safty, S.A. Anisotropic N-Graphene-diffused Co3O4 nanocrystals with dense upper-zone top-on-plane exposure facets as effective ORR electrocatalysts. Sci. Rep. 2018, 8, 3740. [Google Scholar] [CrossRef]
  20. Bayat, S.; Ghanbari, D.; Salavati-Niasari, M. Pechini synthesis of Co2SiO4 magnetic nanoparticles and its application in photo-degradation of azo dyes. J. Mol. Liq. 2016, 220, 223–231. [Google Scholar] [CrossRef]
  21. Shepit, M.; Paidi, V.K.; Roberts, C.A.; Lierop, V.J. Competing ferro-and antiferromagnetic exchange drives shape-selective nanomagnetism. Sci Rep. 2020, 10, 20990. [Google Scholar] [CrossRef] [PubMed]
  22. Alhogbi, B.G.; Aslam, M.; Hameed, A.; Qamar, M.T. The efficacy of Co3O4 loaded WO3 sheets for the enhanced photocatalytic removal of 2,4,6-trichlorophenol in natural sunlight exposure. J. Hazard. Mater. 2020, 397, 122835. [Google Scholar] [CrossRef] [PubMed]
  23. Bayat, S.; Sobhani, A.; Salavati-Niasari, M. Co2SiO4 nanostructures: New simple synthesis, characterization and investigation of optical property. Mater. Res. Bull. 2017, 88, 248–257. [Google Scholar] [CrossRef] [Green Version]
  24. Zheng, Y.L.; Zhou, L.; Sun, X.C.; Zhang, Q.; Yin, H.J.; Du, P.; Yang, X.F.; Zhang, Y.W. Visible-Light Driven CO2 Reduction to CO by Co3O4 Supported on Tungsten Oxide. J. Phys. Chem. C 2022, 126, 3017–3028. [Google Scholar] [CrossRef]
  25. Wang, Z.; Zheng, Z.; Xue, Y.; He, F.; Li, Y. Acidic Water Oxidation on Quantum Dots of IrOx /Graphdiyne. Adv. Energy Mater. 2021, 11, 2101138. [Google Scholar] [CrossRef]
  26. Li, X.; Wang, Z.; Chu, X.; Gao, B.; Zuo, S.; Liu, W.; Yao, C. Broad spectrum photo-driven selective catalytic reduction of NO over CeO2-CeVO4/palygorskite nanocomposite with enhanced SO2 tolerance. Appl. Clay Sci. 2020, 199, 105871. [Google Scholar] [CrossRef]
  27. He, C.; Li, X.; Qiu, D.; Chen, Y.; Lu, Y.; Cui, X. Nonlinear optical polarization and heterostructure synergistically boosted the built-in electric field of CeF3/LiNbO3 for a higher photocatalytic nitrogen reduction activity. Appl. Surf. Sci. 2021, 556, 149753. [Google Scholar] [CrossRef]
  28. Wu, Y.; Yan, L.; Yu, Y.; Jing, C. Photocatalytic CO2 reduction to CH4 on iron porphyrin supported on atomically thin defective titanium dioxide. Catal. Sci. Technol. 2021, 11, 6103–6111. [Google Scholar] [CrossRef]
  29. Liu, W.; Xiang, W.; Guan, N.; Cui, R.; Cheng, H.; Chen, X.; Song, Z.; Zhang, X.; Zhang, Y. Enhanced catalytic performance for toluene purification over Co3O4/MnO2 catalyst through the construction of different Co3O4-MnO2 interface. Sep. Purif. Technol. 2021, 278, 119590. [Google Scholar] [CrossRef]
  30. Bian, H.; Liu, T.; Li, D.; Xu, Z.; Lian, J.; Chen, M.; Yan, J.; Liu, S.F. Unveiling the effect of interstitial dopants on CO2 activation over CsPbBr3 catalyst for efficient photothermal CO2 reduction. Chem. Eng. J. 2022, 435, 135071. [Google Scholar] [CrossRef]
  31. Cao, G.; Ye, X.; Duan, S.; Cao, Z.; Zhang, C.; Yao, C.; Li, X. Plasmon enhanced Sn:In2O3/attapulgite S-scheme heterojunction for efficient photothermal reduction of CO2. Colloids Surf. A Physicochem. Eng. Asp. 2022, 656, 130398. [Google Scholar] [CrossRef]
  32. Bai, S.; Qiu, H.; Song, M.; He, G.; Wang, F.; Liu, Y.; Guo, L. Porous fixed-bed photoreactor for boosting C–C coupling in photocatalytic CO2 reduction. eScience 2022, 2, 428–437. [Google Scholar] [CrossRef]
  33. Lorber, K.; Djinovic, P. Accelerating photo-thermal CO2 reduction to CO, CH4 or methanol over metal/oxide semiconductor catalysts. iScience 2022, 25, 104107. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) XRD spectra of Pal, SiO2, Co3O4, Co2SiO4, and Co2SiO4/Co3O4−x; (b) enlarged XRD patterns in a 2θ range of 36.5°–39.5°.
Figure 1. (a) XRD spectra of Pal, SiO2, Co3O4, Co2SiO4, and Co2SiO4/Co3O4−x; (b) enlarged XRD patterns in a 2θ range of 36.5°–39.5°.
Catalysts 13 00032 g001
Figure 2. TEM images of (a) Pal, (b) Co2SiO4, (c) Co2SiO4/Co3O4−1%, (d) Co2SiO4/Co3O4−5%, and (e) Co2SiO4/Co3O4−10%. Corresponding HRTEM images of (f) Co2SiO4, (g) Co2SiO4/Co3O4−1%, and (h) Co2SiO4/Co3O4−5%.
Figure 2. TEM images of (a) Pal, (b) Co2SiO4, (c) Co2SiO4/Co3O4−1%, (d) Co2SiO4/Co3O4−5%, and (e) Co2SiO4/Co3O4−10%. Corresponding HRTEM images of (f) Co2SiO4, (g) Co2SiO4/Co3O4−1%, and (h) Co2SiO4/Co3O4−5%.
Catalysts 13 00032 g002
Figure 3. (a) UV-Vis-NIR absorption spectra of Pal, Co3O4, Co2SiO4, and Co2SiO4/Co3O4−x. (b) Tauc curves of Co2SiO4 and Co3O4.
Figure 3. (a) UV-Vis-NIR absorption spectra of Pal, Co3O4, Co2SiO4, and Co2SiO4/Co3O4−x. (b) Tauc curves of Co2SiO4 and Co3O4.
Catalysts 13 00032 g003
Figure 4. XPS spectra of Co3O4, Co2SiO4, and Co2SiO4/Co3O4−x: (a) survey spectrum, (b) Co 2p, (c) O 1s, and (d) Si 2p.
Figure 4. XPS spectra of Co3O4, Co2SiO4, and Co2SiO4/Co3O4−x: (a) survey spectrum, (b) Co 2p, (c) O 1s, and (d) Si 2p.
Catalysts 13 00032 g004
Figure 5. (a) Impedance spectra of Co3O4, Co2SiO4, and Co2SiO4/Co3O4−x; Mott−Schottky plot of (b) Co3O4, (c) Co2SiO4, and (d) Co2SiO4/Co3O4−x; VB-XPS of (e) Co2SiO4 and (f) Co3O4.
Figure 5. (a) Impedance spectra of Co3O4, Co2SiO4, and Co2SiO4/Co3O4−x; Mott−Schottky plot of (b) Co3O4, (c) Co2SiO4, and (d) Co2SiO4/Co3O4−x; VB-XPS of (e) Co2SiO4 and (f) Co3O4.
Catalysts 13 00032 g005
Figure 6. (a) In situ DRIFTS spectra for Co2SiO4/Co3O4−x; (b) enlarged in situ DRIFTS spectra at a wavenumber range of 1750−1350 cm−1.
Figure 6. (a) In situ DRIFTS spectra for Co2SiO4/Co3O4−x; (b) enlarged in situ DRIFTS spectra at a wavenumber range of 1750−1350 cm−1.
Catalysts 13 00032 g006
Figure 7. (a) CO2 reduction performance of Co3O4, Co2SiO4, and Co2SiO4/Co3O4−x under solar light; (b) stability of Co2SiO4/Co3O4−x photothermal CO2 reduction.
Figure 7. (a) CO2 reduction performance of Co3O4, Co2SiO4, and Co2SiO4/Co3O4−x under solar light; (b) stability of Co2SiO4/Co3O4−x photothermal CO2 reduction.
Catalysts 13 00032 g007
Figure 8. (a) Band structure and transfer route of photogenerated carriers before and after contact between Co3O4 and Co2SiO4; (b) schematic mechanism of photocatalytic conversion of CO2 over Co2SiO4/Co3O4−x heterostructure under sunlight.
Figure 8. (a) Band structure and transfer route of photogenerated carriers before and after contact between Co3O4 and Co2SiO4; (b) schematic mechanism of photocatalytic conversion of CO2 over Co2SiO4/Co3O4−x heterostructure under sunlight.
Catalysts 13 00032 g008
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Qin, S.; Ge, C.; Kong, X.; Fu, M.; Zhuang, Z.; Li, X. Photothermal Catalytic Reduction of CO2 by Cobalt Silicate Heterojunction Constructed from Clay Minerals. Catalysts 2023, 13, 32. https://doi.org/10.3390/catal13010032

AMA Style

Qin S, Ge C, Kong X, Fu M, Zhuang Z, Li X. Photothermal Catalytic Reduction of CO2 by Cobalt Silicate Heterojunction Constructed from Clay Minerals. Catalysts. 2023; 13(1):32. https://doi.org/10.3390/catal13010032

Chicago/Turabian Style

Qin, Shan, Chengrong Ge, Xiangming Kong, Meng Fu, Ziheng Zhuang, and Xiazhang Li. 2023. "Photothermal Catalytic Reduction of CO2 by Cobalt Silicate Heterojunction Constructed from Clay Minerals" Catalysts 13, no. 1: 32. https://doi.org/10.3390/catal13010032

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