Next Article in Journal
Direct Patterned Zinc-Tin-Oxide for Solution-Processed Thin-Film Transistors and Complementary Inverter through Electrohydrodynamic Jet Printing
Previous Article in Journal
Soft Actuated Hybrid Hydrogel with Bioinspired Complexity to Control Mechanical Flexure Behavior for Tissue Engineering
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

CuMoxW(1-x)O4 Solid Solution Display Visible Light Photoreduction of CO2 to CH3OH Coupling with Oxidation of Amine to Imine

1
Institute for Advanced Study, Nanchang University, Nanchang 330031, China
2
College of Chemistry, Nanchang University, Nanchang 330031, China
3
School of Information Engineering, Jiangxi Modern Polytechnic College, Nanchang 330095, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2020, 10(7), 1303; https://doi.org/10.3390/nano10071303
Submission received: 16 June 2020 / Revised: 24 June 2020 / Accepted: 28 June 2020 / Published: 3 July 2020
(This article belongs to the Section Nanophotonics Materials and Devices)

Abstract

:
The photoreduction of carbon dioxide (CO2) to valuable fuels is a promising strategy for the prevention of rising atmospheric levels of CO2 and the depletion of fossil fuel reserves. However, most reported photocatalysts are only active in the ultraviolet region, which necessitates co-catalysts and sacrificial agents in the reaction systems, leading to an unsatisfied economy of the process in energy and atoms. In this research, a CuMoxW(1-x)O4 solid solution was synthesized, characterized, and tested for the photocatalytic reduction of CO2 in the presence of amines. The results revealed that the yield of CH3OH from CO2 was 1017.7 μmol/g under 24 h visible light irradiation using CuW0.7Mo0.3O4 (x = 0.7) as the catalyst. This was associated with the maximum conversion (82.1%) of benzylamine to N-benzylidene benzylamine with high selectivity (>99%). These results give new insight into the photocatalytic reduction of CO2 for valuable chemical products in an economic way.

Graphical Abstract

1. Introduction

For energy conservation and environmental protection, direct conversion of CO2 into a source of carbon fuels is an ideal solution [1,2,3,4,5,6,7]. Among the different technologies, solar photocatalytic conversion of CO2 is the most promising because it only needs sunlight at room temperature and ambient pressure, which is clean, safe, and abundant [8,9,10,11,12,13,14]. As such, the development of an efficient photocatalyst under sunlight irradiation is attractive and has become a research hotspot. Over past decades, various photocatalytic materials have been analyzed: TiO2 [15,16,17], CdS [18,19], Bi2O3 [20,21], CeO2 [22,23], and Bi2WO6 [24,25,26]. Their well-designed heterostructures have been examined for reduction of CO2 to fuels, such as CO [27,28,29], CH4 [30,31,32], CH3OH [33,34], and C2H6 [35,36], but the overall photoconversion efficiency and product selectivity is still unsatisfactory and requires further promotion for practical applications.
It has been confirmed that forming a solid solution between semiconductors is an excellent method in the development of visible light-driven photocatalysts for sensitive photoreduction of CO2 [37,38,39,40] and is widely applied in photocatalytic water splitting and pollutant degradation, while exhibiting better performance than single components comprising the solid solution [41,42,43,44,45,46]. Taking this into account, a series of solid solution photocatalysts, such as BiOBrxCl1-x [47], ZnxCd1-xS [48], GaN:ZnO [49], and zinc gallogermanate [39], have been synthesized and examined for CO2 reduction under visible light irradiations. Very recently, Ling and colleagues [50] demonstrated that CuxAgyInzZnkS solid solutions customized with RuO2 or a Rh1.32Cr0.66O3 co-catalyst showed high photocatalytic activity for the reduction of CO2 into CH3OH with a yield up to 118.5 μmol g–1 h–1 under visible light irradiation. Liang’s group [51] reported that a N-doped, graphene-functionalized Zn1.231Ge0.689N1.218O0.782 solid solution exhibited high photocatalytic activity for the evolution of CH4 from the photocatalytic reaction of CO2 and H2O, coupled with oxidation of benzyl alcohol under visible light irradiation.
Since the pioneering work of Benko, CuWO4 has been confirmed as a promising photoanode material for photoelectrochemical (PEC) splitting water, with a narrower band gap of 2.3 eV that allows for the use of visible light and enhanced stability in neutral and moderate to basic pH [52,53]. In this work, a CuMoxW(1-x)O4 solid solution was synthesized by a facile hydrothermal method and then examined for the photocatalytic reduction of CO2 under visible light irradiation without the addition of any co-catalysts and sacrificed reagents in the system. In this process, CO2 was reduced to CH3OH by photo-induced electrons, while amine was selected as the “hydrogen-donor” to capture the photogenerated holes. Then, the amine was converted into imine, resulting in good energy and atomic economies.

2. Experimental

2.1. General

All chemicals and solvents were purchased from commercial suppliers and used as received unless explicitly stated. Using tetramethylsilane (TMS) as an internal standard, 1H nuclear magnetic resonance (NMR) and 13C NMR spectra were measured on a Bruker AVANCE 400 spectrometer (Basel, Switzerland) in CDCl3. X-ray diffraction (XRD) data were collected on a D8 diffractometer from Bruker Instruments (Basel, Switzerland), utilizing Cu Kα radiation at a scan rate of 0.05 s–1. The scanning electron microscopy (SEM) images were obtained by a Quanta 200F FEI (Los Angeles, USA). Transmission electron microscopy (TEM) analysis was carried out on a JEM-2100 (TEM) (Okinawa, Japan), and energy dispersive X-ray spectra (EDX) were obtained at an accelerating voltage of 200 kV. X-ray photoelectron spectroscopy (XPS) data were performed with a Thermo Scientific (New York, USA) ESCALAB250Xi XPS spectrometer with Al Kα at 500 eV. The ultraviolet-visible (UV-Vis) DRS spectra of photocatalysts were conducted using a Cary 60 UV-vis spectrophotometer (London, Europe) with BaSO4 as a reference. Photoluminescence (PL) characterizations were carried out on a FluoroMax-4 spectrometer (London, Europe), using a λex = 320 nm light source. Gas chromatography (GC) product analysis was performed on a flame ionization detector. The composition of liquid products was analyzed by GC-MS (Beijing, China) with a HP-5 capillary column (30 m × 0.25 mm × 0.25 mm).

2.2. The Preparation of Solid Solution Photocatalysts

A series of CuMoxW(1-x)O4 solid solution photocatalysts was prepared by a typical hydrothermal synthesis method, in which X is defined as the molar ratio of Mo/(Mo + W). Setting the preparation of CuMo0.1W0.9O4 (X = 0.1) as an example: 10 mmol Cu(NO3)2·3H2O was dissolved in 250 mL distilled water as precursor solution A. Onethousandth of a mol of NaMoO4·2H2O and 9.0 mmol NaWO4·2H2O were dissolved in 250 mL distilled water as precursor solution B. Precursor solution A was added dropwise to precursor solution B under a strong stirring condition in 30 min, during which the mixtures turned blue-green and more suspended substance appeared. After the addition, the suspended substance (about 80 mL) was then transferred into a 100 mL autoclave and heated at 180 °C for 24 h. After cooling to room temperature, the obtained product was then filtered and washed with distilled water and anhydrous alcohol several times and then dried in a vacuum at 80 °C for 10 h. The other photocatalysts could be easily prepared in the same way by adjusting the molar ratio: Mo/(Mo + W).

2.3. Photocatalytic Reduction of CO2

Photocatalytic reactions were carried out in 40 mL anhydrous acetonitrile with 0.06 g corresponding to the photocatalyst and 1 mmol benzylamine (purified by distilled before using) in an 80 mL self-made quartz reactor at 0.5 MPa CO2 partial pressure. Before the reaction, the reactor was evacuated with high-purity dry CO2 gas and blown through several times to remove lesser quantities of oxygen, and then it was tightly closed. Before illumination, the reaction system was stirred under darkness to prompt CO2 adsorption and desorption equilibrium after blowing CO2 for 30 min. Then, a Xe lamp (300 W) with a 420 nm cut-off filter was turned on as the light source for the photocatalytic reactions. During the processes, the reaction temperature was maintained at room temperature with a circulating cool water bath. After the reactions were complete, the products were immediately analyzed by GC (Figure S1), 1H NMR (Figure S2), and MS (Figure S3). Detailed calculations of methanol yield, conversion rate of benzylamine, and selectivity of N-benzylidenebenzylamine are shown in the Supporting Information.

3. Results and Discussion

3.1. Catalyst Characterization

3.1.1. XRD and EDX Analysis

To reveal the phase structure of various photocatalysts prepared by the hydrothermal method, XRD patterns were recorded and shown in Figure 1. The CuMoxW(1-x)O4 photocatalysts were formed in two different crystalline structures, depending on their chemical composition. For x ≤ 0.7, the XRD patterns exhibit several broad diffraction peaks with low intensity, suggesting that these photocatalysts were in a poor crystalline state. Moreover, the patterns of these materials are in good agreement with the standard spectrum of CuWO4·H2O [54,55] (JCPDS no.: 33-0503). For x ≥ 0.8, several sharp diffraction peaks appeared and intensified, indicating that these photocatalysts are in a good crystalline state. 2θ values of 20.4, 21.4, 24.9, 25.3, 25.4, and 25.9 correspond to the Cu3(MoO4)2(OH)2 of the (021) (101), (121), (130), (040), and (111) faces, respectively, which is in agreement with the reported values [56]. It can be interpreted that the increased Mo content had some effects on the crystal structure of these photocatalysts in the system [56]. Furthermore, the EDX analysis results of the sample are shown in Figure 2 and Table 1. The Mo content in the reaction system gradually increased with increasing x values, and the sample changed from CuWO4·H2O to Cu3(MoO4)2(OH)2. In addition, the XPS spectra of the samples were obtained (Figure S5), which clearly indicate that the Cu and W atoms exist in the form of Cu2+ and W6+, respectively.

3.1.2. SEM and TEM Analysis

Figure 3 shows the SEM and TEM micromorphology of the prepared samples. As seen in Figure 3a, the surface of the selected material prepared by the hydrothermal method was smooth and existed in globular particles. To observe the microstructure of the photocatalyst, the characterization of the photocatalyst (x = 0.7) was carried out using TEM, and the average particle size was shown to be about 30 nm (Figure 3b). The high resolution transmission electron microscopy (HR-TEM) images of the interfacial structure of the photocatalyst in different regions is shown in Figure 3c,d, indicating that the photocatalyst CuMo0.7W0.3O4 (x = 0.7) prepared via hydrothermal method was amorphous because no obvious lattice stripes were observed in the photocatalyst.

3.1.3. UV-Vis Absorption and PL Spectroscopy Analysis

UV-vis DRS spectra of the solid solutions were obtained, and the results are shown in Figure 4. The samples had a wide absorption with a wavelength range from 350 nm to 750 nm. A bathochromic shift absorption band (from 570 to 610 nm) was observed with increasing x values from 0 to 0.7. With a further increasing x value (from 0.8 to 1.0), a blue-shifted absorption band was observed from 620 nm to 580 nm. The different trends of the absorption bands can be verified via XRD spectra. When x ≤ 0.7, these samples were in the form of CuWO4 ·H2O, but, when x > 0.7, they were in the form of Cu 3 ( MoO 4 ) 2 ( OH ) 2 . Generally speaking, the absorption of the lower band is more easily disturbed by the introduction of Mo and W than the higher bands, which indicates that a charge transfer from O 2 to M o 6 + or W 6 + contributes to the formation of the absorption band. This is consistent with a paper by Gaudon et al. [57,58].
As is well known, photoluminescence (PL) emission can be used to discover carrier transfer and recombination efficiency because it is caused by the recombination of photogenic carriers. Lower PL emission intensity suggests a lower recombination rate and exhibits higher photocatalytic activity for a photocatalyst [59,60], as shown in Figure 5, as the value of x increases from 0 to 0.7, the fluorescence intensity of the samples gradually weaken. This is mainly because the Mo content is gradually increasing, which inhibits the recombination of photogenerated charges and carriers, thereby improving the photocatalytic activity. The fluorescence intensity of the samples was gradually strengthened with further increases in x value (from 0.8 to 1.0). It is obvious that the PL emission spectra of samples can be divided into two categories (Figure 5). One type of spectrum was obtained for x values in the range of 0.8 to 1.0, showing strong PL emission with emission intensity increasing with increasing x value. Another emission intensity, lower than the former, was found when the x values were between 0 to 0.7. All optical results revealed that the photocatalytic activity of CuMo0.7W0.3O4 (x = 0.7) is optimal for the system.

3.2. Photocatalytic Reduction of Carbon Dioxide Activity Measurement

To investigate the performance of various photocatalysts prepared hydrothermally, the photocatalytic oxidation of amine to imine coupling with the reduction of CO2 to CH3OH was carried out under visible light (λ > 420 nm) irradiation and a 0.5 MPa CO2 pressure environment. The detailed results are shown in Table 2. These results prove that CH3OH was selectively obtained from CO2, and imine was produced from the oxidation of amine, which were further confirmed with GC, NMR, and MS analysis (Figures S1–S3). We found that benzylamine can be converted to N-benzylidene benzylamine with high selectivity (>99%) and CO2 can be converted to CH3OH with good productivity after 12 h irradiation with photocatalysts. Moreover, the results suggested that CuW0.7Mo0.3O4 (x = 0.7) was the optimum photocatalyst among the examined samples, which showed the highest conversion of benzylamine (57.2%) and high selectivity towards N-benzylidenebenzylamine (>99%) within 14 h irradiations (Table 2, entry 4), and received the highest productivity of CH3OH up to 671.8 μmol/g. These results are almost consistent with optical observations. Therefore, the CuW0.7Mo0.3O4 (x = 0.7) was selected as the photocatalyst in the following experiments. In addition, the results of two blank reactions showed that the conversion of benzylamine was 7.1 and 91.3% in the absence of catalyst and in the presence of catalyst, respectively.
The time plot for the oxidation of benzylamine to the corresponding imine under 0.5 MPa of CO2 on CuW0.7Mo0.3O4 (x = 0.7) is shown in Figure 6a,b. With prolonged irradiation time, the conversion of benzylamine was significantly enhanced, accompanied by an increased yield in CH3OH. As shown, the yield of CH3OH increased slightly within 10 h, and then obviously increased as time continued. With visible irradiation for 24 h, 82.1% of benzylamine was converted and >99% of them were selectively transformed to the corresponding imine, and the CO2 was converted to CH3OH with high productivity up to 1017.7 μmol/g, suggesting a satisfactory photocatalytic efficiency and selectivity under visible light conditions.
Motivated by preceding results, the substrate scope of the oxidation of other amines to corresponding imines coupling with the reduction of CO2 to CH3OH with photocatalytic CuW0.7Mo0.3O4 (x = 0.7) was also carried out. Detailed results, shown in Table 3, show that the oxidative coupling reactions of the amine derivatives to their corresponding imines coupling with the formation of CH3OH from CO2 could proceed well under the same conditions. However, different substrates can present great differences in the conversion of amines and the yield of CH3OH. Methyl-substituted benzylamines at the o-, m-, and p- positions of the benzene ring (Table 3, entries 2–4) can also be converted, but there are slight differences in the conversion process, and it can been seen from the Table 3 that their conversion rates differ vary greatly, which might to be relative to the effect of the steric hindrance. It can be ascribed to the electronic effects associated with electron withdrawing substituents (entry 5–6) and electron donating substituents (entry 4,7) on the benzene ring have an obvious effect on the conversion rate of amines, as well as the yield of formation of CH3OH. Two halo-substituted benzylamines could be oxidized to corresponding imines with a conversion of 30% and yield of 221.4 μmol/g of CH3OH. Moreover, heterocyclic amines were also transformed into the corresponding imines (entries 9–10). However, when aliphatic amines were selected as the substrates (Table 3, entry 11,12), no imines were obtained.
To measure the reusability of the photocatalyst, the CuW0.7Mo0.3O4 (x = 0.7) was selected and re-used with the same reaction conditions three times. Before the proceeding reaction, the used photocatalysts were purified by thorough washing with CH3CN. Figure 7 illustrates cycling runs of CuW0.7Mo0.3O4 (x = 0.7) in the photocatalytic reduction of CO2 coupling with amine selective oxidation to imine under visible light irradiation. It is clear that the high photocatalytic activity of the as-prepared sample minimally decreased after three recycles, indicating good photocatalytic stability of the photocatalyst under visible light irradiation. The decrease in photocatalytic efficiency was mainly due to the loss of crystal water in the photocatalyst after use.
The mechanism for photocatalytic reduction of CO2 to CH3OH coupling with the conversion of amine to corresponding imine is similar to our previous report [61] as shown in Figure 4. Details are as follows:
Catalyst + hv → catalyst + e-cond + p+val,
PhCH2NH2 + 2p+val → PhCH=NH + 2H+cond,
CO2 + 2H+cond + 2e-cond → HCOOH,
HCOOH + 2H+cond + 2e-cond → HCHO + H2O,
HCHO + 2H+cond + 2e-cond → CH3OH,
PhCH=NH + H2O → PhCHO + NH3(S),
PhCHO + PhCH2NH2 → PhCH=NCH2Ph + H2O,
With absorption of visible light (Equations (1) and (2)), the photogenerated holes in the valence band (VB) of the photocatalyst oxidized benzylamine to the corresponding imine and generated H+ (Equation (2)), which shows that CO2 can be reduced by photoelectrons in conduction band (CB) to generate HCOOH, HCHO, and CH3OH (Equations (3)–(5)). The formed imine in Equation (2) was easily decomposed into benzaldehyde and NH3(S) (Equation (6)), and then the obtained benzaldehyde directly reacted with the benzylamine to form the corresponding imine and H2O (Equation (7)).

4. Conclusions

In summary, a series of CuWxMo(1-x)O4 solid solutions were prepared and applied to the reduction of CO2 into CH3OH coupling with the conversion of aromatic amine into its corresponding imine. The results revealed that the yield of CH3OH from CO2 was 1017.7 μmol/g under 24 h visible light irradiations using an optimized photocatalyst, CuW0.7Mo0.3O4 (x = 0.7). The catalyst chosen was associated with the maximum conversion (82.1%) of benzylamine to N-benzylidene benzylamine with high selectivity (>99%). Further experiments revealed that CuW0.7Mo0.3O4 (x = 0.7) exhibited good substrate suitability, photocatalytic activity, and photocatalytic stability. All results should be helpful for the design and application of economical photocatalysts for the reduction of CO2 to CH3OH under visible light.

Supplementary Materials

The following are available online at https://www.mdpi.com/2079-4991/10/7/1303/s1, Figure S1. GC chromatograms of (a) before irradiation and (b) the stand CH3OH and benzylamine in CH3CN and (c) product (CH3OH) after irradiation for 10 h with the partial enlarged drawing. Figure S2. The 1H NMR spectrums of (a) the product (CH3OH) photocatalyzed by CuW0.7Mo0.3O4 (x = 0.7) after reaction for 10 h and (b) the stand CH3OH. Figure S3. the MS of Products photocatalyzed by CuW0.7Mo0.3O4 (x = 0.7) after irradiation for 10 h. Figure S4. GC chromatograms of gas phase products after irradiation for 10 h. Figure S5. XPS spectra of (a) Cu2p and (b) W4f of photocatalysts.

Author Contributions

Writing–original draft, Writing–review & editing, C.L.; Data curation, T.Y., Q.H., and X.L.; Resources, H.L., and Y.Z.; Supervision, G.X., and W.Z.; Formal analysis, H.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Li, D.; Rohani, V.; Fabry, F.; Parakkulam Ramaswamy, A. Direct conversion of CO2 and CH4 into liquid chemicals by plasma-catalysis. Appl. Catal. B Environ. 2020, 261, 118228. [Google Scholar] [CrossRef]
  2. Wang, Y.; Gao, W.; Kazumi, S.; Li, H.; Yang, G.; Tsubaki, N. Direct and Oriented Conversion of CO2 into Value-Added Aromatics. Chemistry 2019, 25, 5149–5153. [Google Scholar] [CrossRef]
  3. Zheng, Y.; Vasileff, A.; Zhou, X.; Jiao, Y.; Jaroniec, M.; Qiao, S.-Z. Understanding the Roadmap for Electrochemical Reduction of CO2 to Multi-Carbon Oxygenates and Hydrocarbons on Copper-Based Catalysts. J. Am. Chem. Soc. 2019, 141, 7646–7659. [Google Scholar] [CrossRef]
  4. Ma, J.; Sun, N.; Zhang, X.; Zhao, N.; Xiao, F.; Wei, W.; Sun, Y. A short review of catalysis for CO2 conversion. Catal. Today 2009, 148, 221–231. [Google Scholar] [CrossRef]
  5. Hu, B.; Guild, C.; Suib, S.L. Thermal, electrochemical, and photochemical conversion of CO2 to fuels and value-added products. J. CO2 Util. 2013, 1, 18–27. [Google Scholar] [CrossRef]
  6. Yang, S.; Rui, P.; Hensley, D.K.; Bonnesen, P.V.; Rondinone, A.J. High-Selectivity Electrochemical Conversion of CO2 to Ethanol using a Copper Nanoparticle/N-doped Graphene Electrode. Chemistryselect 2016, 1, 6055–6061. [Google Scholar]
  7. Benson, E.E.; Kubiak, C.P.; Sathrum, A.J.; Smieja, J.M. Electrocatalytic and homogeneous approaches to conversion of CO2 to liquid fuels. Chem. Soc. Rev. 2008, 38, 89–99. [Google Scholar] [CrossRef]
  8. Tu, W.; Yong, Z.; Zou, Z. ChemInform Abstract: Photocatalytic Conversion of CO2 into Renewable Hydrocarbon Fuels: State-of-the-Art Accomplishment, Challenges, and Prospects. Adv. Mater. 2015, 45, 4607–4626. [Google Scholar]
  9. Li, K.; An, X.; Park, K.H.; Khraisheh, M.; Tang, J. A critical review of CO2 photoconversion: Catalysts and reactors. Catal. Today 2014, 224, 3–12. [Google Scholar] [CrossRef] [Green Version]
  10. Lee, Y.Y.; Jung, H.S.; Kang, Y.T. A review: Effect of nanostructures on photocatalytic CO2 conversion over metal oxides and compound semiconductors. J. CO2 Util. 2017, 20, 163–177. [Google Scholar] [CrossRef]
  11. Yuan, L.; Xu, Y.-J. Photocatalytic conversion of CO2 into value-added and renewable fuels. Appl. Surf. Sci. 2015, 342, 154–167. [Google Scholar] [CrossRef]
  12. Khalil, M.; Gunlazuardi, J.; Ivandini, T.A.; Umar, A. Photocatalytic conversion of CO2 using earth-abundant catalysts: A review on mechanism and catalytic performance. Renew. Sustain. Energy Rev. 2019, 113, 109246. [Google Scholar] [CrossRef]
  13. Das, S.; Daud, W.W. A review on advances in photocatalysts towards CO2 conversion. RSC Adv. 2014, 4, 20856–20893. [Google Scholar] [CrossRef]
  14. 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] [PubMed]
  15. Xiong, Z.; Lei, Z.; Li, Y.; Dong, L.; Zhao, Y.; Zhang, J. A review on modification of facet-engineered TiO2 for photocatalytic CO2 reduction. J. Photochem. Photobiol. C Photochem. Rev. 2018, 36, 24–47. [Google Scholar] [CrossRef]
  16. Zhao, H.; Pan, F.; Ying, L. A review on the effects of TiO2 surface point defects on CO2 photoreduction with H2O. J. Mater. 2016, 3, 17–32. [Google Scholar]
  17. Shehzad, N.; Tahir, M.; Johari, K.; Murugesan, T.; Hussain, M. A critical review on TiO2 based photocatalytic CO2 reduction system: Strategies to improve efficiency. J. CO2 Util. 2018, 26, 98–122. [Google Scholar] [CrossRef]
  18. Cho, K.M.; Kim, K.H.; Park, K.; Kim, C.; Kim, S.; Al-Saggaf, A.; Gereige, I.; Jung, H.-T. Amine-functionalized graphene/CdS composite for photocatalytic reduction of CO2. ACS Catal. 2017, 7, 7064–7069. [Google Scholar] [CrossRef]
  19. Jin, J.; Yu, J.; Guo, D.; Ho, W. A Hierarchical Z-Scheme CdS-WO3 Photocatalyst with Enahnced CO2 Reduction Activity. Mater. Views 2015, 39, 5262–5271. [Google Scholar]
  20. Jin, J.; Tao, H. Facile synthesis of Bi2S3 nanoribbons for photocatalytic reduction of CO2 into CH3OH. Appl. Surf. Sci. 2016, 394, 364–370. [Google Scholar] [CrossRef]
  21. Tran-Phu, T.; Daiyan, R.; Fusco, Z.; Ma, Z.; Amal, R.; Tricoli, A. Nanostructured β-Bi2O3 Fractals on Carbon Fibers for Highly Selective CO2 Electroreduction to Formate. Adv. Funcyional Mater. 2020, 30, 1906478. [Google Scholar] [CrossRef]
  22. Liu, S.-Q.; Zhou, S.-S.; Chen, Z.-G.; Liu, C.-B.; Chen, F. An artificial photosynthesis system based on CeO2 as light harvester and N-doped graphene Cu(II) complex as artificial metalloenzyme for CO2 reduction to methanol fuel. Catal. Commun. 2016, 73, 7–11. [Google Scholar] [CrossRef]
  23. Li, M.; Zhang, L.; Wu, M.; Du, Y.; Fan, X.; Wang, M.; Zhang, L.; Kong, Q.; Shi, J. Mesostructured CeO2/g-C3N4 nanocomposites: Remarkably enhanced photocatalytic activity for CO2 reduction by mutual component activations. Nano Energy 2016, 19, 145–155. [Google Scholar] [CrossRef]
  24. Zhou, Y.; Tian, Z.; Zhao, Z.; Liu, Q.; Kou, J.; Chen, X.; Gao, J.; Yan, S.; Zou, Z. High-Yield Synthesis of Ultrathin and Uniform Bi2WO6 Square Nanoplates Benefitting from Photocatalytic Reduction of CO2 into Renewable Hydrocarbon Fuel under Visible Light. ACS Appl Mater Interfaces 2011, 3, 3594–3601. [Google Scholar] [CrossRef]
  25. Dai, W.; Xu, H.; Yu, J.; Hu, X.; Luo, X.; Tu, X.; Yang, L. Photocatalytic reduction of CO2 into methanol and ethanol over conducting polymers modified Bi2WO6 microspheres under visible light. Appl. Surf. Sci. 2015, 356, 173–180. [Google Scholar] [CrossRef]
  26. Xiao, L.; Lin, R.; Wang, J.; Cui, C.; Wang, J.; Li, Z. A Novel Hollow-Hierarchical Structured Bi2WO6 with Enhanced Photocatalytic Activity for CO2 Photoreduction. J. Colloid Interface Sci. 2018, 523, 151–158. [Google Scholar] [CrossRef]
  27. Wang, J.C.; Zhang, L.; Fang, W.X.; Ren, J.; Li, Z. Enhanced Photoreduction CO2 Activity over Direct Z-Scheme α-Fe2O3/Cu2O Heterostructures Under Visible Light Irradiation. ACS Appl. Mater. Interfaces 2015, 7, 8631–8639. [Google Scholar] [CrossRef]
  28. Guo, Z.; Fei, Y.; Ying, Y.; Leung, C.F.; Ng, S.M.; Ko, C.C.; Cometto, C.; Lau, T.C.; Robert, M. Photocatalytic Conversion of CO2 to CO by a Copper(II) Quaterpyridine Complex. Chemsuschem 2017, 10, 4009–4013. [Google Scholar] [CrossRef]
  29. Niu, K.; Xu, Y.; Wang, H.; Ye, R.; Xin, H.L.; Lin, F.; Tian, C.; Lum, Y.; Bustillo, K.C.; Doeff, M.M. A spongy nickel-organic CO2 reduction photocatalyst for nearly 100% selective CO production. Sci. Adv. 2017, 3, e1700921. [Google Scholar] [CrossRef] [Green Version]
  30. Wu, T.; Zou, L.; Han, D.; Li, F.; Zhang, Q.; Niu, L. A carbon-based photocatalyst efficiently converts CO2 to CH4 and C2H2 under visible light. Green Chem. 2014, 16, 2142. [Google Scholar] [CrossRef]
  31. Xiong, Z.; Lei, Z.; Kuang, C.C.; Chen, X.; Gong, B.; Zhao, Y.; Zhang, J.; Zheng, C.; Wu, J.C.S. Selective photocatalytic reduction of CO2 into CH4 over Pt-Cu2O TiO2 nanocrystals: The interaction between Pt and Cu2O cocatalysts. Appl. Catal. B Environ. 2016, 202, 695–703. [Google Scholar] [CrossRef]
  32. Li, H.; Gao, Y.; Xiong, Z.; Liao, C.; Shih, K. Enhanced selective photocatalytic reduction of CO2 to CH4 over plasmonic Au modified g-C3N4 photocatalyst under UV–vis light irradiation. Appl. Surf. Sci. 2018, 439, 552–559. [Google Scholar] [CrossRef]
  33. Chen, J.; Xin, F.; Niu, H.; MAo, C.-j.; Song, J.-M. Photocatalytic reduction of CO2 with methanol over Bi2S3-ZnIn2S4 nanocomposites. Mater. Lett. 2017, 198, 1–3. [Google Scholar] [CrossRef]
  34. Yu, W.; Xu, D.; Peng, T. Enhanced photocatalytic activity of g-C3N4 for selective CO2 reduction to CH3OH via facile coupling of ZnO: A direct Z-scheme mechanism. J. Mater. Chem. A 2015, 3, 19936. [Google Scholar] [CrossRef]
  35. Phongamwong, T.; Chareonpanich, M.; Limtrakul, J. Role of chlorophyll in Spirulina on photocatalytic activity of CO2 reduction under visible light over modified N-doped TiO2 photocatalysts. Appl. Catal. B Environ. 2014, 168–169, 114–124. [Google Scholar]
  36. Lo, C.C.; Hung, C.-H.; Yuan, C.-S.; Wu, J.-F. Photoreduction of carbon dioxide with H2 and H2O over TiO2 and ZrO2 in a circulated photocatalytic reactor. Sol. Energy Mater. Sol. Cells 2007, 91, 1765–1774. [Google Scholar] [CrossRef]
  37. Wang, J.; Li, G.; Li, Z.; Tang, C.; Feng, Z.; An, H.; Liu, H.; Liu, T.; Li, C. A highly selective and stable ZnO-ZrO2 solid solution catalyst for CO2 hydrogenation to methanol. Sci. Adv. 2017, 3, e1701290. [Google Scholar] [CrossRef] [Green Version]
  38. Maeda, K.; Teramura, L.; Takata, T.; Hara, M.; Saito, N.; Toda, K.; Inoue, Y.; Hisayoshi, K.; Domen, K. Overall Water Splitting on (Ga1-xZnx)(N1-xOx) Solid Solution Photocatalyst: Relationship between Physical Properties and Photocatalytic Activity. J. Phys. Chem. B 2005, 109, 20504–20510. [Google Scholar] [CrossRef] [PubMed]
  39. Yan, S.; Wang, J.; Gao, H.; Wang, N.; Yu, H.; Li, Z.; Zhou, Y.; Zou, Z. Zinc Gallogermanate Solid Solution: A Novel Photocatalyst for Efficiently Converting CO2 into Solar Fuels. Adv. Funct. Mater. 2013, 23, 1839–1845. [Google Scholar] [CrossRef]
  40. Chai, Y.; Li, L.; Lu, J.; Li, D.; Shen, J.; Zhang, Y.; Liang, J.; Wang, X. Germanium-substituted Zn2TiO4 solid solution photocatalyst for conversion of CO2 into fuels. J. Catal. 2019, 371, 144–152. [Google Scholar] [CrossRef]
  41. Ohno, T.; Bai, L.; Hisatomi, T.; Maeda, K.; Domen, K. Photocatalytic Water Splitting Using Modified GaN:ZnO Solid Solution under Visible Light: Long-Time Operation and Regeneration of Activity. J. Am. Chem. Soc. 2012, 134, 8254–8259. [Google Scholar] [CrossRef] [PubMed]
  42. Li, Z.; Zhang, F.; Han, J.; Zhu, J.; Li, M.; Zhang, B.; Fan, W.; Lu, J.; Li, C. Using Pd as a Cocatalyst on GaN–ZnO Solid Solution for Visible-Light-Driven Overall Water Splitting. Catal. Lett. 2018, 148, 933–939. [Google Scholar] [CrossRef]
  43. Maeda, K.; Takata, T.; Hara, M.; Saito, N.; Inoue, Y.; Kobayashi, H.; Domen, K. GaN:ZnO Solid Solution as a Photocatalyst for Visible-Light-Driven Overall Water Splitting. J. Am. Chem. Soc. 2006, 127, 8286–8287. [Google Scholar] [CrossRef] [PubMed]
  44. Yu, X.; Shavel, A.; An, X.; Luo, Z.; Ibá?ez, M.; Cabot, A. Cu2ZnSnS4-Pt and Cu2ZnSnS4-Au Heterostructured Nanoparticles for Photocatalytic Water Splitting and Pollutant Degradation. J. Am. Chem. Soc. 2014, 136, 9236–9239. [Google Scholar] [CrossRef] [PubMed]
  45. Qi, L.; Yang, Y.; Zhang, P.; Le, Y.; Wang, C.; Wu, T. Hierarchical flower-like BiOIxBr(1-x) solid solution spheres with enhanced visible-light photocatalytic activity. Appl. Surf. Sci. 2019, 467–468, 792–801. [Google Scholar] [CrossRef]
  46. Ouyang, S.; Ye, J. β-AgAl(1-x)Ga(x)O2 solid-solution photocatalysts: Continuous modulation of electronic structure toward high-performance visible-light photoactivity. J. Am. Chem. Soc. 2011, 133, 7757. [Google Scholar] [CrossRef] [PubMed]
  47. Gao, M.; Yang, J.; Sun, T.; Zhang, Z.; Zhang, D.; Huang, H.; Lin, H.; Fang, Y.; Wang, X. Persian buttercup-like BiOBrxCl1-x solid solution for photocatalytic overall CO2 reduction to CO and O2. Appl. Catal. B Environ. 2019, 243, 734–740. [Google Scholar] [CrossRef]
  48. Pan, L.; Xuehua, Z.; Chun-Chao, H.; Lin, L.; Chen, Y.; Tao, H. Visible-light driven CO2 photoreduction over ZnxCd1-xS solid solution coupling with tetra(4-carboxyphenyl)porphyrin iron(III) chloride. Phys. Chem. Chem. Phys. 2018, 20, 16985. [Google Scholar]
  49. Zhou, P.; Wang, X.; Yan, S.; Zou, Z. Solid Solution Photocatalyst with Spontaneous Polarization Exhibiting Low Recombination Toward Efficient CO2 Photoreduction. Chemsuschem 2016, 9, 2064–2068. [Google Scholar] [CrossRef] [PubMed]
  50. Liu, J.Y.; Garg, B.; Ling, Y.-C. CuxAgyInzZnkSm solid solutions customized with RuO2 or Rh1.32Cr0.66O3 co-catalyst display visible light-driven catalytic activity for CO2 reduction to CH3OH. Green Chem. 2011, 13, 2029. [Google Scholar] [CrossRef]
  51. Liang, J.; Li, L. Synthesis of N-doped graphene-functionalized Zn1.231Ge0.689N1.218O0.782 solid solution as a photocatalyst for CO2 reduction and oxidation of benzyl alcohol under visible-light irradiation. J. Mater. Chem. A 2017, 5, 10998. [Google Scholar] [CrossRef]
  52. Benko, F.A.; MacLaurin, C.L.; Koffyberg, F.P. CuWO4 and Cu3WO6 as anodes for the photoelectrolysis of water. Mater. Res. Bull. 1982, 17, 133–136. [Google Scholar] [CrossRef]
  53. Lhermitte, C.R.; Bartlett, B.M. Advancing the Chemistry of CuWO4 for Photoelectrochemical Water Oxidation. Acc. Chem. Res. 2016, 49, 1121–1129. [Google Scholar] [CrossRef] [PubMed]
  54. Cheng, J.; Lei, C.; Xiong, E.; Jiang, Y.; Xia, Y. Preparation and characterization of W–Cu nanopowders by a homogeneous precipitation process. J. Alloy. Compd. 2006, 421, 146–150. [Google Scholar] [CrossRef]
  55. Azar, G.T.P.; Rezaie, H.R.; Razavizadeh, H. Synthesis and consolidation of W–Cu composite powders with silver addition. Int. J. Refract. Met. Hard Mater. 2012, 31, 157–163. [Google Scholar] [CrossRef]
  56. Calvert, L.D.; Barnes, W.H. The structure of lindgrenite. Can Miner. 1957, 6, 31–51. [Google Scholar]
  57. Fu, X.; Ji, J.; Tang, W.; Liu, W.; Chen, S. Mo–W based copper oxides: Preparation, characterizations, and photocatalytic reduction of nitrobenzene. Mater. Chem. Phys. 2013, 141, 719–726. [Google Scholar] [CrossRef]
  58. Gaudon, M.; Carbonera, C.; Thiry, A.E.; Demourgues, A.; Deniard, P.; Payen, C.; Létard, J.F.; Jobic, S. Adaptable Thermochromism in the CuMo 1-x W x O 4 Series (0 ≤ x <0.1): A Behavior Related to a First-Order Phase Transition with a Transition Temperature Depending on x. Inorg. Chem. 2007, 46, 10200–10207. [Google Scholar]
  59. Jing, L.; Qu, Y.; Wang, B.; Li, S.; Jiang, B.; Yang, L.; Fu, W.; Fu, H.; Sun, J. Review of photoluminescence performance of nano-sized semiconductor materials and its relationships with photocatalytic activity. Sol. Energy Mater. Sol. Cells 2006, 90, 1773–1787. [Google Scholar]
  60. Yu, J.-G.; Yu, H.-G.; Cheng, B.; Zhao, X.-J.; Yu, J.C.; Ho, W.-K. The Effect of Calcination Temperature on the Surface Microstructure and Photocatalytic Activity of TiO2 Thin Films Prepared by Liquid Phase Deposition. J. Phys. Chem. B 2003, 107, 13871–13879. [Google Scholar] [CrossRef]
  61. Yang, T.; Yu, Q.; Wang, H. Photocatalytic Reduction of CO2 to CH3OH Coupling with the Oxidation of Amine to Imine. Catal. Lett. 2018, 148, 2382–2390. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of nanoparticles prepared by hydrothermal method at 180 °C.
Figure 1. XRD patterns of nanoparticles prepared by hydrothermal method at 180 °C.
Nanomaterials 10 01303 g001
Figure 2. Energy dispersive X-ray spectra (EDX) spectra of (a) CuMo0.5W0.5O4 (x = 0.5) and the plane (A), (b) CuMo0.7W0.3O4 (x = 0.7) and the line (B-B), and (c) CuMo0.8W0.2O4 (x = 0.8) and the plane (C1).
Figure 2. Energy dispersive X-ray spectra (EDX) spectra of (a) CuMo0.5W0.5O4 (x = 0.5) and the plane (A), (b) CuMo0.7W0.3O4 (x = 0.7) and the line (B-B), and (c) CuMo0.8W0.2O4 (x = 0.8) and the plane (C1).
Nanomaterials 10 01303 g002
Figure 3. (a) Scanning electron microscopy (SEM), (b) transmission electron microscopy (TEM), and (c), (d) high resolution transmission electron microscopy (HR-TEM) images of the CuMo0.7W0.3O4 (x = 0.7) photocatalyst.
Figure 3. (a) Scanning electron microscopy (SEM), (b) transmission electron microscopy (TEM), and (c), (d) high resolution transmission electron microscopy (HR-TEM) images of the CuMo0.7W0.3O4 (x = 0.7) photocatalyst.
Nanomaterials 10 01303 g003
Figure 4. Ultraviolet-visible (UV-vis) diffuse reflectance spectra of different samples.
Figure 4. Ultraviolet-visible (UV-vis) diffuse reflectance spectra of different samples.
Nanomaterials 10 01303 g004
Figure 5. Photoluminescence emission spectra of different samples with excitation at 320 nm.
Figure 5. Photoluminescence emission spectra of different samples with excitation at 320 nm.
Nanomaterials 10 01303 g005
Figure 6. (a) Influence of irradiation time on conversion of benzylamine and selectivity of corresponding imine with CuW0.7Mo0.3O4 (x = 0.7); (b) yields of CH3OH production as a function of irradiation time.
Figure 6. (a) Influence of irradiation time on conversion of benzylamine and selectivity of corresponding imine with CuW0.7Mo0.3O4 (x = 0.7); (b) yields of CH3OH production as a function of irradiation time.
Nanomaterials 10 01303 g006aNanomaterials 10 01303 g006b
Figure 7. Photocatalytic activity of the reuse of CuW0.7Mo0.3O4 (x = 0.7).
Figure 7. Photocatalytic activity of the reuse of CuW0.7Mo0.3O4 (x = 0.7).
Nanomaterials 10 01303 g007
Table 1. Molar ratio of Cu:Mo:W.
Table 1. Molar ratio of Cu:Mo:W.
EntryThe Value of xThe Ratio of Cu:Mo:W
10.51:0.12:0.97
20.71:0.23:1.05
30.81:0.76:0.13
Table 2. Photocatalytic activity of different nanoparticles.
Table 2. Photocatalytic activity of different nanoparticles.
EntryCatalystGasCH3OHAmineImine
the Value of x (μmol/g)Conv. (%)Sel. (%)
10CO2296.929.5>99
20.3CO2375.435.9>99
30.5CO2571.248.3>99
40.7CO2671.854.2>99
50.8CO2509.945.7>97
60.9CO2394.634.4>99
71.0CO2249.323.1>99
8 a0.7N2--2.1--
9 b--CO2--1.7--
10 c0.7CO251.79.5>99
11 d0.7CO2------
12 e0.7CO2--91.3--
13 f--CO2--7.1--
a 1 mmol benzylamine; 40 mL CH3CN, 60 mg photocatalyst; b without photocatalyst; c 40 mL benzylamine, without CH3CN; d 40 mL CH3CN, without benzylamine; e O2, f O2, without photocatalyst.
Table 3. a Effect of different amines on photocatalytic activity.
Table 3. a Effect of different amines on photocatalytic activity.
EntryReactantProductYield bConv. (%) cSel. (%) d
1 Nanomaterials 10 01303 i001 Nanomaterials 10 01303 i002671.857.299
2 Nanomaterials 10 01303 i003 Nanomaterials 10 01303 i004127.516.399
3 Nanomaterials 10 01303 i005 Nanomaterials 10 01303 i006227.426.298
4 Nanomaterials 10 01303 i007 Nanomaterials 10 01303 i008251.943.999
5 Nanomaterials 10 01303 i009 Nanomaterials 10 01303 i01033.16.598
6 Nanomaterials 10 01303 i011 Nanomaterials 10 01303 i012137.219.498
7 Nanomaterials 10 01303 i013 Nanomaterials 10 01303 i014214.625.199
8 Nanomaterials 10 01303 i015 Nanomaterials 10 01303 i016221.430.699
9 Nanomaterials 10 01303 i017 Nanomaterials 10 01303 i018126.216.398
10 Nanomaterials 10 01303 i019 Nanomaterials 10 01303 i02029.74.998
11 Nanomaterials 10 01303 i021--------
12 Nanomaterials 10 01303 i022--------
a 1 mmol amine, 0.06 g photocatalysis, 40 mL CH3CN, 0.1 Mpa CO2; b Unit: (µmol g−1); c Conversion of amine; d Selectivity of the corresponding imine.

Share and Cite

MDPI and ACS Style

Luo, C.; Yang, T.; Huang, Q.; Liu, X.; Ling, H.; Zhu, Y.; Xia, G.; Zou, W.; Wang, H. CuMoxW(1-x)O4 Solid Solution Display Visible Light Photoreduction of CO2 to CH3OH Coupling with Oxidation of Amine to Imine. Nanomaterials 2020, 10, 1303. https://doi.org/10.3390/nano10071303

AMA Style

Luo C, Yang T, Huang Q, Liu X, Ling H, Zhu Y, Xia G, Zou W, Wang H. CuMoxW(1-x)O4 Solid Solution Display Visible Light Photoreduction of CO2 to CH3OH Coupling with Oxidation of Amine to Imine. Nanomaterials. 2020; 10(7):1303. https://doi.org/10.3390/nano10071303

Chicago/Turabian Style

Luo, Chao, Tian Yang, Qianfei Huang, Xian Liu, Huan Ling, Yuxin Zhu, Guoming Xia, Wennan Zou, and Hongming Wang. 2020. "CuMoxW(1-x)O4 Solid Solution Display Visible Light Photoreduction of CO2 to CH3OH Coupling with Oxidation of Amine to Imine" Nanomaterials 10, no. 7: 1303. https://doi.org/10.3390/nano10071303

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