Next Article in Journal
Redox Isomerization of Allylic Alcohols Catalyzed by New Water-Soluble Rh(I)-N-Heterocyclic Carbene Complexes
Next Article in Special Issue
Formaldehyde Total Oxidation on Manganese-Doped Hydroxyapatite: The Effect of Mn Content
Previous Article in Journal
Rh Particles Supported on Sulfated g-C3N4: A Highly Efficient and Recyclable Heterogeneous Catalyst for Alkene Hydroformylation
Previous Article in Special Issue
Abatement of Toluene Using a Sequential Adsorption-Catalytic Oxidation Process: Comparative Study of Potential Adsorbent/Catalytic Materials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

A Highly Active Au/In2O3-ZrO2 Catalyst for Selective Hydrogenation of CO2 to Methanol

School of Chemical Engineering and Technology, Tianjin University, Tianjin 300072, China
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(11), 1360; https://doi.org/10.3390/catal10111360
Submission received: 29 October 2020 / Revised: 19 November 2020 / Accepted: 20 November 2020 / Published: 23 November 2020

Abstract

:
A novel gold catalyst supported by In2O3-ZrO2 with a solid solution structure shows a methanol selectivity of 70.1% and a methanol space–time yield (STY) of 0.59 gMeOH h−1 gcat−1 for CO2 hydrogenation to methanol at 573 K and 5 MPa. The ZrO2 stabilizes the structure of In2O3, increases oxygen vacancies, and enhances CO2 adsorption, causing the improved activity.

Graphical Abstract

1. Introduction

With the development of renewable hydrogen, hydrogenation of CO2 to methanol (Equation (1)) has received increasing attention worldwide. Significant efforts have been made towards the preparation of catalysts with high CO2 conversion and high methanol selectivity. The most investigated catalysts are promoted copper catalysts based on the industrial Cu/ZnO catalysts for methanol synthesis from syngas [1,2,3]. Recently, ZnO-ZrO2 [4], GaNi alloy [5], bimetallic PdZn [6], MOF [7,8], and others [1] were reported to have nice activity for CO2 hydrogenation to methanol.
CO2 + 3 H2 → CH3OH + H2O ΔH = −49.7 kJ mol−1
In 2012, our group predicted via density functional theoretical (DFT) studies that In2O3 possesses high activity for CO2 activation [9]. The further DFT [10] and experimental [11] studies confirmed that In2O3 with oxygen vacancies exhibited high activity for CO2 hydrogenation to methanol, with unusually high methanol selectivity. Since then, In2O3-based catalysts have received significant attention, with increasing publications. The reported works include preparation and characterization of In2O3 [12]; combination of In2O3 with other oxides [13,14], especially ZrO2 that forms the solid solution with In2O3 and improves CO2 adsorption with optimum oxygen vacancy [15,16,17] and generates strong electronic interaction between oxides [18]; and the addition of metals, including gold [19,20], copper [21,22], cobalt [23], palladium [24,25,26,27], rhodium [28], nickel [29], and platinum [30,31]. The metal catalysts applied significantly improve the hydrogenation ability of In2O3. It has been found that In2O3 has a strong interaction with metals. This strong interaction causes a total change in catalytic performance. It makes the Ni, Au, Rh, and Pt catalysts become highly selective towards CO2 hydrogenation to methanol with In2O3 support. Especially for the Au catalyst, most of the reported Au catalysts are highly active for oxidation, with almost a thousand papers per year recently. However, only a few papers were published for hydrogenation. The In2O3-supported Au catalyst shows the highest selectivity and best activity ever reported for CO2 hydrogenation to methanol. The methanol selectivity reaches 100% at temperatures below 498 K and is more than 70% at 548 K [19]. It is even 67.8% at 573 K, with a space–time yield (STY) of methanol of 0.47 gMeOH h−1 gcat−1 at 573 K, 5 MPa, and 21,000 cm3 h−1 gcat−1. In this work, we confirm that the In2O3-ZrO2 supported Au catalyst presents even higher activity for CO2 hydrogenation to methanol.

2. Results

In2O3-ZrO2 oxides were prepared via a co-precipitation method, and then Au nanoparticles (NPs) were loaded on the carrier through the deposition–precipitation method. The loading amount of Au was 1%. The samples after reaction at 573 K and 5 MPa were named with “-AR” at the end. As shown in Supplementary Materials Figure S1, the sample with In/Zr molar ratio of 4:1 shows the best catalytic performance. The sample with this In/Zr ratio (labeled as Au/In2O3-ZrO2) was thus selected for the activity test and further characterization. Based on the N2 adsorption–desorption test on 77 K, the In2O3-ZrO2 and Au/In2O3-ZrO2 samples exhibit a similar total specific surface area of 111.5 and 121.6 m2 g−1, respectively. As shown in Figure 1a, the results of X-ray diffraction (XRD) show that the characterized peaks for all samples containing indium belong to cubic-In2O3 (PDF file number 06-0416). Meanwhile, the peaks corresponding to Au cannot be observed, which results from the extremely low loading and ultra-high dispersion of Au NPs. This also means that Au NPs remain stable, preventing deactivation by the agglomeration. The absence of metallic indium after reaction also shows that the structure of the Au/In2O3-ZrO2 catalyst is stable during the reaction. Moreover, the diffraction peak of the In2O3-ZrO2 sample largely differentiates from that of ZrO2, which can be deduced that Zr may be penetrated into In2O3 [32]. In Figure 1b, the high-resolution transmission electron microscopy (HRTEM) image of the Au/In2O3-ZrO2 sample shows the crystalline lattice distances of 0.180, 0.398, 0.275, and 0.291 nm, which are assigned to cubic-In2O3 (440), (211), (321), and (222) facets, respectively. No lattice fringes of ZrO2 or Au can be observed. As shown in Figure 1c, the result of element distribution analysis indicates that the supported Au NPs and Zr are highly dispersed before and after reaction (Supplementary Materials Figure S2). These results further confirm the excellent dispersion of Au NPs and the conclusion that In2O3-ZrO2 is in a solid solution state.
The activity test of CO2 hydrogenation to methanol was conducted in a vertical fixed-bed reactor. The products were analyzed with an online gas chromatograph (Agilent 7890A). For all catalysts, methanol and CO are the main products. Trace methane can be detected at reaction temperatures over 548 K. The carbon balance is better than 99%. As shown in Figure 2a, the Au/In2O3-ZrO2 catalyst exhibits the highest catalytic performance among these three catalysts. The CO2 conversion of the Au/In2O3-ZrO2 catalyst is 14.8%, with a methanol selectivity of 70.1% at 573 K and 5 MPa. The methanol STY reaches 0.59 gMeOH h−1 gcat−1, which is much higher than that over the Au/In2O3 catalyst (0.47 gMeOH h−1 gcat−1) [19] and other Au-based catalysts (Supplementary Materials Table S1). For all catalysts, CO2 conversion increases with the temperature increases, while the methanol selectivity decreases due to the competitive reverse water–gas shift (RWGS) reaction (Equation (2)). Although CO can be detected at 498 K for the Au/In2O3-ZrO2 catalyst, the higher methanol selectivity at 523 K and 548 K is obvious, as compared to the Au/In2O3 catalyst. This result reveals that the Au/In2O3-ZrO2 catalyst maintains high selectivity for methanol synthesis from CO2 hydrogenation. On the other hand, the presence of Zr enhances the CO2 conversion of the Au/In2O3 catalysts. As shown in Figure 2c, the methanol STY on the Au/In2O3-ZrO2 catalyst only decreases slightly and is stabilized at 92% of the initial value after a 6 h reaction at 573 K and 5 MPa. Furthermore, the Au/In2O3-ZrO2 catalyst shows the lowest apparent activation energy (Ea) of 74.1 kJ mol−1 (Figure 2d). These results prove the excellent performance for CO2 hydrogenation to methanol over the Au/In2O3-ZrO2 catalyst.
CO2 + H2 → CO + H2O ΔH = 41.2 kJ mol−1
To further understand the factors responsible for the enhanced catalytic performance, Raman analyses were performed. As shown in Figure 3a, the band at 132 cm−1 is assigned to the In-O vibration (VIn-O) [25]. The scattering features at 306 cm−1 (I1) and 493 cm−1 belong to the bending vibration and stretching vibrations of InO6 octahedrons, respectively [33]. The stretching vibration of In-O-In, occurring at 364 cm−1 (I2), was confirmed to reflect the degree of oxygen defects [19]. For fresh samples, the band of VIn-O visibly shifts towards lower frequencies due to the addition of Zr. Moreover, the characteristic peaks of the In2O3-ZrO2 and Au/In2O3-ZrO2 samples are completely the same as those of In2O3 without the peaks attributed to ZrO2 (Supplementary Materials Figure S3). These results affirm the penetration of Zr atoms into the indium oxides [32]. The intensities of I2 increase with the addition of Zr, suggesting that more oxygen vacancies are generated. However, the loading of Au results in lower I2 intensities, likely owing to the coverage of Au NPs on oxygen vacancies [30]. Subsequently, the variation of related peaks over the samples after H2 reduction for different times under 473 K (labeled with “-HR-1.5h” or “-HR-1h”at the end) was further investigated. As shown in Figure 3b,c, H2 reduction results in the increase of oxygen vacancies for all samples, according to the much larger I2 band [30]. Moreover, prolonging reduction time enables oxygen vacancies to increase sequentially. The area ratio of I2 and I1 (labeled as I2/I1) was calculated to characterize the amount of oxygen vacancies (Figure 3d). The oxygen vacancies of the Au/In2O3 sample are 40% of fresh In2O3, while those of the Au/In2O3-HR-1.5h sample are 46% of the In2O3-HR-1.5h sample. Comparing Au/In2O3-ZrO2 with In2O3-ZrO2 under the same treatment condition, we see that the ratio of their I2/I1 value increases by 21% from fresh samples to the samples after 1.5 h H2 reduction. This result indicates that the well-dispersed Au NPs benefit the formation of oxygen vacancies under the atmosphere including H2. Furthermore, normalized by the I2/I1 value of fresh In2O3, the oxygen vacancies of samples after 1.5 h H2 reduction increase by 2.04, 0.92, 1.00, and 0.90 times in In2O3, In2O3-ZrO2, Au/In2O3, and Au/In2O3-ZrO2, respectively. During the H2 reduction process, the oxygen vacancies of In2O3 increase much more drastically than the other samples. Noticeably, some bulk information of the sample can be obtained under the excitation laser with the wavelength of 532 nm [34]. The rapid increment of oxygen vacancies can be attributed to the over-reduction of pure In2O3 [30]. It results in metallic indium without activity for H2 activation [13]. Thus, it can be deduced that the strong metal-support interaction (SMSI) between Au and carriers could be conducive to maintain the structure of In2O3 [19], which can also be performed through the incorporation of Zr into In2O3. Furthermore, compared to Au/In2O3, the Au/In2O3-ZrO2 sample exhibits the higher I2/I1 values for not only fresh samples but also the samples after reduction, implying the increase of oxygen vacancy density. These results coincide with the better catalytic activity of the Au/In2O3-ZrO2 catalyst.
H2 temperature-programmed reduction (H2-TPR) was employed to investigate the reduction behaviors of the catalysts. The H2-TPR profiles are shown in Figure 4a. For all samples, the reduction peaks located below 500 K can be assigned to the surface reduction [30]. The onset of reduction peak at 530 K belongs to the reduction of bulk In2O3. Due to the fact that In2O3 can be reduced by hydrogen more easily than ZrO2 [17], the enhanced stability of Au/In2O3-ZrO2 under the hydrogen atmosphere is thereby achieved. The XRD patterns of the spent samples were further investigated. As shown in Supplementary Materials Figure S4, there is a noticeable shift towards a lower angle for the In2O3-AR sample compared to fresh In2O3, while the Au/In2O3 and Au/In2O3-ZrO2 samples can remain at the same angle after the reaction at 573 K and 5 MPa. Our previous work reported that the lattice expansion of In2O3 caused by thermal effect and oxygen vacancies could be performed on the shift of XRD diffraction peak [19]. Moreover, it is similar for lattice parameters of In2O3 and Au/In2O3, heated under argon and reactant mixture at 573 K, respectively [19]. As the thermal effect can be excluded, it can be conducted that this shift may be attributed to the increasing oxygen vacancies of bulk In2O3. Thus, it can be deduced that the addition of Zr and Au could effectively hinder the over-reduction of In2O3 under a reactant mixture including H2. In the temperature range from 275 to 530 K, the surface reduction peak shifts towards the lower temperature (below 360 K) with the addition of Au NPs. This result confirms that Au NPs promote the formation of surface oxygen vacancies. In addition, a broader and intense H2 consumption peak at 350 K for the Au/In2O3-ZrO2 sample is obvious. The incorporation of Zr into In2O3 promotes the formation of surface oxygen vacancies on the Au/In2O3-ZrO2 sample. The higher onset temperature of surface reduction and the disappeared peak at 456 K indicate the stabilization function of Zr on In2O3 structure. These results are in line with the results of Raman spectra above. The H2-TPR profiles demonstrate the increasing surface oxygen vacancies over the Au/In2O3-ZrO2 catalyst, as a reason for its prominent catalytic activity.
To further investigate the adsorption behavior of CO2 on the catalyst surface, temperature-programmed desorption of CO2 (CO2-TPD) was conducted. Figure 4b presents the profiles of CO2-TPD, involving several signals in the temperature ranges of 325–425, 485–570, 570–755, and 730–865 K. The single peak around 360 K for all samples belongs to the desorption of physically absorbed CO2 [35]. Other peaks from low to high temperatures can be attributed to chemically absorbed CO2 on the different sites, ascribed to weak, medium, and strong basic sites, respectively. Comparing the Au/In2O3-ZrO2 sample with other samples, we see that the desorption peaks shift to the much higher temperatures of 707 and 798 K, respectively. This result implies that the addition of Zr enhances the strength of CO2 adsorption on these sites, probably owing to the enhanced basicity [36]. It can be proposed that the enhanced CO2 adsorption on the Au/In2O3-ZrO2 sample benefits methanol synthesis. Time-resolved in situ diffuse reflectance infrared Fourier transform spectroscopy (in situ DRIFTS) were also conducted to gain the insights into CO2 adsorption. As shown in Supplementary Materials Figure S5, the CO2 interaction with Au/In2O3-ZrO2 prefers to produce carbonate species rather than bicarbonate species due to the change of intensities [37]. The bicarbonate species disappears much faster after the Ar desorption, especially over the Au/In2O3 sample. Moreover, compared to the Au/In2O3 sample, the blue shift over the Au/In2O3-ZrO2 sample for the bands at 900–1100 cm−1, attributed to CO2 absorbed on vacancies, suggests the stronger adsorption of CO2 [17,38]. These results reveal that the addition of Zr strengthens the interaction between CO2 and the catalyst surface. Based on the discussion above, the SMSI between Au and In2O3-ZrO2 solid solution facilitates H2 dissociation. Meanwhile, the incorporation of Zr into In2O3 leads to the increasing oxygen vacancies, as well as the enhanced CO2 adsorption. The H adatoms, dissociated by the Au NPs strongly interacting with the carrier, can spill over to the oxygen vacancy sites and react with the CO2 molecule activated on the oxygen vacancy sites. This significantly promotes the CO2 conversion into methanol.
In conclusion, we have shown that the integration of Au, In2O3, and ZrO2 enables the high selectivity for methanol synthesis from CO2 hydrogenation. The Au/In2O3-ZrO2 catalyst shows a CO2 conversion of 14.8% with a methanol STY of 0.59 gMeOH h−1 gcat−1 at 573 K and 5 MPa. Such an outstanding activity indicates that the addition of Zr remarkably improves the activity of the Au/In2O3 catalyst. The catalyst characterizations demonstrate that Au NPs are well dispersed on the surface of the In2O3-ZrO2 solid solution. Further investigation through Raman spectra, H2-TPR, and CO2-TPD indicates that the enhanced H2 activation is achieved by the Au NPs strongly interacting with catalysts surface due to the SMSI effect. The presence of Zr mainly exhibits four effects: (1) reducing the apparent activation energy, (2) stabilizing the structure of In2O3, (3) increasing the amount of oxygen vacancies, and (4) enhancing the CO2 adsorption. This work not only extends the utilization of gold catalysts for CO2 hydrogenation but also promotes the fundamental studies for CO2 conversion.

3. Materials and Methods

3.1. Catalyst Preparation

Indium-zirconium oxide carriers were prepared by a co-precipitation approach. Appropriate amounts of In(NO3)3·× H2O (M = 363.3 g mol−1, Beijing HWRK Chem Co., Ltd., Beijing, China, 99.99%) and ZrO(NO3)2·× H2O (M = 257.9 g mol−1, Shanghai Macklin Biochemicals Co., Ltd., Shanghai, China, 99%) with a fixed molar ratio were firstly dissolved in deionized water (0.145 M of In3+). A 0.2 M aqueous solution of Na2CO3 (Tianjin Kemiou Chemical Reagent Co., Ltd., Tianjin, China, 99%) was added dropwise under continuous stirring to reach a pH of 10. The mixture was aged at 353 K for 3 h. Thereafter, the precipitate was filtered and then washed with deionized water several times. Prior to the calcination at 723 K (10 K min−1, 3 h), the sample was dried at 353 K overnight. The obtained carrier was labeled as In2O3-ZrO2. The In2O3 and ZrO2 samples were prepared with the same protocol. The Au/In2O3-ZrO2 sample was prepared via a deposition–precipitation method, which is described in detail elsewhere [19]. The only difference was that the sample after freeze-drying overnight was directly used for the catalytic activity test, without a further reduction step. The samples with other molar ratios of In/Zr were denoted as Au/In2O3-ZrO2 (m:n), where m:n is the molar ratio of In/Zr.

3.2. Characterization

The phase compositions of the samples were characterized by a Bruker D8 Focus diffractometer equipped with Cu Kα radiation source (λ = 1.54056 Å). Diffraction data were collected at a scanning speed of 8° min−1 over the 2θ range of 10°–70°. The N2 adsorption/desorption isotherms, measuring on an AUTOSORB-1-C instrument (Quantachrome) at 77 K, were used to determine the total specific surface area (SBET).
The Raman spectra were recorded by an INVIA Raman analyzer (Renishaw PLC. Ltd., Gloucestershire, UK) with laser irradiation at 532 nm. The SEM images and EDX spectra were used to characterize microstructure and element distribution. All samples were measured on a Hitachi S-4800 SEM at an operating voltage of 25 kV without any conductive coating process. The transmission electron microscopy (TEM) and HRTEM images were obtained using a JEOL JEM-F200 microscope at the accelerating voltage of 200 kV.
H2-TPR was carried out using the Micromeritics Autochem II 2920 chemisorption analyzer. About 100 mg of the as-prepared sample was pretreated under flowing helium for 1 h at 473 K to remove the absorbed impurities. Thereafter, the sample was cooled down to 273 K by liquid nitrogen under flowing helium. Then the sample was heated to 973 K (10 K min−1) after switching to a mixture gas of 10% H2 in N2. The effluent gas was analyzed using a thermal conductivity detector (TCD). The signals were normalized to the weight of the sample for the following analysis.
CO2-TPD was conducted on the same apparatus as H2-TPR. After the same pretreating process as H2-TPR, the sample (about 100 mg) was cooled down to 323 K under flowing helium, followed by CO2 adsorption for 1 h. After being purged with flowing helium for 1 h to remove physically absorbed CO2 at the same temperature, the sample was heated to 1073 K under flowing helium at a rate of 10 K min−1. The effluent gas was analyzed by the TCD. The signals were normalized to the weight of the sample for the following analysis.
Time-resolved in situ DRIFTS was performed using a Tensor 27 spectrometer (Bruker) equipped with a MCT detector by recording 32 scans at a resolution of 2 cm−1. After the pretreatment under flowing argon for 2 h at 323 K, the background spectra were firstly collected and subtracted from the sample spectra. For CO2 adsorption, a 1.0% CO2 in Ar gas with the flow rate of 20 mL min−1 was fed into the system at 323 K. After the collecting of time-resolved spectra for 30 min, the gas was switched to pure Ar. The desorption time-resolved spectra were collected after flushing for 10 min.

3.3. Catalyst Activity Test

The activity test of CO2 hydrogenation to methanol was conducted in a vertical fixed-bed reactor setup. The reactor was loaded with 0.2 g of the catalysts and 1.0 g of SiC, which was purged with N2, at room temperature. After 0.5 h, the gas flow was switched to the reactant mixture (H2/CO2/N2 = 76/19/5, molar ratio) until the pressure reached 5 MPa. The reaction was conducted with a temperature range from 473 to 573 K. The gaseous hourly space velocity (GHSV) was 21,000 cm3 h−1 gcat−1. The reaction conditions were determined by the reported studies [17,39] and our previous study on Au/In2O3 [19] in order to get better catalytic performance. Products were analyzed with an online gas chromatograph (Agilent 7890A) equipped with a flame ionized detector (FID) and a TCD. All the valves and lines between the reactor outlet and GC inlet were heated to 413 K to prevent the condensation of methanol. CO2 conversion (XCO2) and selectivity (Si) and space–time yield (STY, gi gcat−1 h−1) of product i (i = methanol or CO) were calculated as follows:
XCO2 = (FCO2, in − FCO2, out)/FCO2, in × 100%
Si = Fi, out/(FCO2, in − FCO2, out) × 100%
STYi = (FCO2, in × Si × XCO2)/mcat × Mi
where F and M are the molar flow rate and the molar mass of product i, respectively, and mcat is the catalyst sample weight.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/10/11/1360/s1. Table S1: Comparison of Au catalysts for CO2 hydrogenation to methanol. Figure S1: (a) CO2 conversion and methanol selectivity, and (b) methanol space–time yield (STY) respect to the molar In content, which is equal to m/(m + n) for the Au/In2O3-ZrO2 (m:n) sample. Reaction conditions: 5 MPa, H2/CO2/N2 = 76/19/5, GHSV = 21,000 cm3 h−1 gcat−1. Figure S2: TEM images of Au/In2O3-ZrO2 (a), Au/In2O3-ZrO2-AR (b) and Au/In2O3-ZrO2-AR (c); SEM image (d) and the corresponding EDX mapping of Au/In2O3-ZrO2-AR. Figure S3: Visible Raman spectra of fresh ZrO2, Au/In2O3-ZrO2, In2O3-ZrO2, and In2O3. Figure S4: Enlarged XRD patterns of peaks around 32° for fresh samples and the samples after reaction. Figure S5: Time-resolved in situ DRIFTS as a function of CO2 adsorption time and Ar desorption time at 323 K after pretreatment in Ar over Au/In2O3 (a) and Au/In2O3-ZrO2 (b). Symbol definition: bicarbonate (white frame), carbonate (red frame), CO2 absorbed on the vacancies featuring the higher CO2 adsorption energy (yellow frame).

Author Contributions

C.L., conceptualization, supervision, and review and revision of the final version of the paper; Z.L., investigation, methodology, and visualization and writing—original draft preparation; K.S., methodology, and review and editing; J.W., review and editing; Z.Z., visualization. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Key Research and Development Program of China (2016YFB0600902).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jiang, X.; Nie, X.; Guo, X.; Song, C.; Chen, J.G. Recent advances in carbon dioxide hydrogenation to methanol via heterogeneous catalysis. Chem. Rev. 2020, 120, 7984–8034. [Google Scholar]
  2. Wang, W.; Wang, S.; Ma, X.; Gong, J. Recent advances in catalytic hydrogenation of carbon dioxide. Chem. Soc. Rev. 2011, 40, 3703–3727. [Google Scholar]
  3. Su, X.; Xu, J.; Liang, B.; Duan, H.; Hou, B.; Huang, Y. Catalytic carbon dioxide hydrogenation to methane: A review of recent studies. J. Energy Chem. 2016, 25, 553–565. [Google Scholar]
  4. 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]
  5. Studt, F.; Sharafutdinov, I.; Abild-Pedersen, F.; Elkjær, C.F.; Hummelshøj, J.S.; Dahl, S.; Chorkendorff, I.; Nørskov, J.K. Discovery of a Ni-Ga catalyst for carbon dioxide reduction to methanol. Nat. Chem. 2014, 6, 320–324. [Google Scholar]
  6. Liao, F.; Wu, X.-P.; Zheng, J.; Li, M.M.-J.; Kroner, A.; Zeng, Z.; Hong, X.; Yuan, Y.; Gong, X.-Q.; Tsang, S.C.E. A promising low pressure methanol synthesis route from CO2 hydrogenation over Pd@Zn core–shell catalysts. Green Chem. 2017, 19, 270–280. [Google Scholar]
  7. Xiang, W.; Zhang, Y.; Chen, Y.; Liu, C.J.; Tu, X. Synthesis, characterization and application of defective metal–organic frameworks: Current status and perspectives. J. Mater. Chem. A 2020, 8, 21526–21546. [Google Scholar]
  8. Gutterød, E.S.; Lazzarini, A.; Fjermestad, T.; Kaur, G.; Manzoli, M.; Bordiga, S.; Svelle, S.; Lillerud, K.P.; Skúlason, E.; Øien-Ødegaard, S.; et al. Hydrogenation of CO2 to methanol by Pt nanoparticles encapsulated in UiO-67: Deciphering the role of the metal-organic framework. J. Am. Chem. Soc. 2020, 142, 999–1009. [Google Scholar] [PubMed]
  9. Ye, J.; Liu, C.; Ge, Q. DFT study of CO2 adsorption and hydrogenation on the In2O3 surface. J. Phys. Chem. C 2012, 116, 7817–7825. [Google Scholar]
  10. Ye, J.; Liu, C.; Mei, D.; Ge, Q. Active oxygen vacancy site for methanol synthesis from CO2 hydrogenation on In2O3(110): A DFT Study. ACS Catal. 2013, 3, 1296–1306. [Google Scholar]
  11. Sun, K.; Fan, Z.; Ye, J.; Yan, J.; Ge, Q.; Li, Y.; He, W.; Yang, W.; Liu, C.-J. Hydrogenation of CO2 to methanol over In2O3 catalyst. J. CO2 Util. 2015, 12, 1–6. [Google Scholar]
  12. Tsoukalou, A.; Abdala, P.M.; Stoian, D.; Huang, X.; Willinger, M.-G.; Fedorov, A.; Müller, C.R. Structural evolution and dynamics of an In2O3 catalyst for CO2 hydrogenation to methanol: An operando XAS-XRD and in situ TEM study. J. Am. Chem. Soc. 2019, 141, 13497–13505. [Google Scholar] [PubMed]
  13. Chen, T.-Y.; Cao, C.; Chen, T.-B.; Ding, X.; Huang, H.; Shen, L.; Cao, X.; Zhu, M.; Xu, J.; Gao, J.; et al. Unraveling highly tunable selectivity in CO2 hydrogenation over bimetallic In-Zr oxide catalysts. ACS Catal. 2019, 9, 8785–8797. [Google Scholar]
  14. Dou, M.; Zhang, M.; Chen, Y.; Yu, Y. Mechanistic insight into the modification of the surface stability of In2O3 catalyst through metal oxide doping. Catal. Lett. 2018, 148, 3723–3731. [Google Scholar]
  15. Tsoukalou, A.; Abdala, P.M.; Armutlulu, A.; Willinger, E.; Fedorov, A.; Müller, C.R. Operando X-ray absorption spectroscopy identifies a monoclinic ZrO2: In solid solution as the active phase for the hydrogenation of CO2 to methanol. ACS Catal. 2020, 10, 10060–10067. [Google Scholar]
  16. Frei, M.S.; Mondelli, C.; Cesarini, A.; Krumeich, F.; Hauert, R.; Stewart, J.A.; Curulla Ferré, D.; Pérez-Ramírez, J. Role of zirconia in indium oxide-catalyzed CO2 hydrogenation to methanol. ACS Catal. 2020, 10, 1133–1145. [Google Scholar]
  17. Martin, O.; Martín, A.J.; Mondelli, C.; Mitchell, S.; Segawa, T.F.; Hauert, R.; Drouilly, C.; Curulla-Ferré, D.; Pérez-Ramírez, J. Indium oxide as a superior catalyst for methanol synthesis by CO2 hydrogenation. Angew. Chem. Int. Ed. 2016, 55, 6261–6265. [Google Scholar]
  18. Yang, C.; Pei, C.; Luo, R.; Liu, S.; Wang, Y.; Wang, Z.; Zhao, Z.-J.; Gong, J. Strong electronic oxide–support interaction over In2O3 /ZrO2 for highly selective CO2 hydrogenation to methanol. J. Am. Chem. Soc. 2020. [Google Scholar] [CrossRef]
  19. Rui, N.; Zhang, F.; Sun, K.; Liu, Z.; Xu, W.; Stavitski, E.; Senanayake, S.D.; Rodriguez, J.A.; Liu, C.-J. Hydrogenation of CO2 to methanol on a Auδ+–In2O3–x catalyst. ACS Catal. 2020, 10, 11307–11317. [Google Scholar]
  20. Rui, N.; Sun, K.; Shen, C.; Liu, C.-J. Density functional theoretical study of Au4/In2O3 catalyst for CO2 hydrogenation to methanol: The strong metal-support interaction and its effect. J. CO2 Util. 2020, 42, 101313. [Google Scholar]
  21. Shi, Z.; Tan, Q.; Wu, D. A novel core–shell structured CuIn@SiO2 catalyst for CO2 hydrogenation to methanol. AIChE J. 2019, 65, 1047–1058. [Google Scholar]
  22. Shi, Z.; Tan, Q.; Tian, C.; Pan, Y.; Sun, X.; Zhang, J.; Wu, D. CO2 hydrogenation to methanol over Cu-In intermetallic catalysts: Effect of reduction temperature. J. Catal. 2019, 379, 78–89. [Google Scholar]
  23. Bavykina, A.; Yarulina, I.; Al Abdulghani, A.J.; Gevers, L.; Hedhili, M.N.; Miao, X.; Galilea, A.R.; Pustovarenko, A.; Dikhtiarenko, A.; Cadiau, A.; et al. Turning a methanation Co catalyst into an In–Co methanol producer. ACS Catal. 2019, 9, 6910–6918. [Google Scholar]
  24. Ye, J.; Liu, C.-J.; Mei, D.; Ge, Q. Methanol synthesis from CO2 hydrogenation over a Pd4/In2O3 model catalyst: A combined DFT and kinetic study. J. Catal. 2014, 317, 44–53. [Google Scholar]
  25. Rui, N.; Wang, Z.; Sun, K.; Ye, J.; Ge, Q.; Liu, C.-J. CO2 hydrogenation to methanol over Pd/In2O3: Effects of Pd and oxygen vacancy. Appl. Catal. B 2017, 218, 488–497. [Google Scholar]
  26. Frei, M.S.; Mondelli, C.; García-Muelas, R.; Kley, K.S.; Puértolas, B.; López, N.; Safonova, O.V.; Stewart, J.A.; Curulla Ferré, D.; Pérez-Ramírez, J. Atomic-scale engineering of indium oxide promotion by palladium for methanol production via CO2 hydrogenation. Nat. Commun. 2019, 10, 3377. [Google Scholar]
  27. Cai, Z.; Dai, J.; Li, W.; Tan, K.B.; Huang, Z.; Zhan, G.; Huang, J.; Li, Q. Pd supported on MIL-68(In)-derived In2O3 nanotubes as superior catalysts to boost CO2 hydrogenation to methanol. ACS Catal. 2020, 10, 13275–13289. [Google Scholar]
  28. Wang, J.; Sun, K.; Jia, X.; Liu, C.-J. CO2 hydrogenation to methanol over Rh/In2O3 catalyst. Catal. Today 2020. [Google Scholar] [CrossRef]
  29. Jia, X.; Sun, K.; Wang, J.; Shen, C.; Liu, C.-J. Selective hydrogenation of CO2 to methanol over Ni/In2O3 catalyst. J. Energy Chem. 2020, 50, 409–415. [Google Scholar]
  30. Sun, K.; Rui, N.; Zhang, Z.; Sun, Z.; Ge, Q.; Liu, C.-J. A highly active Pt/In2O3 catalyst for CO2 hydrogenation to methanol with enhanced stability. Green Chem. 2020, 22, 5059–5066. [Google Scholar]
  31. Han, Z.; Tang, C.; Wang, J.; Li, L.; Li, C. Atomically dispersed Ptn+ species as highly active sites in Pt/In2O3 catalysts for methanol synthesis from CO2 hydrogenation. J. Catal. 2020. [Google Scholar] [CrossRef]
  32. Su, J.; Wang, D.; Wang, Y.; Zhou, H.; Liu, C.; Liu, S.; Wang, C.; Yang, W.; Xie, Z.; He, M. Direct conversion of syngas into light olefins over zirconium-doped indium(III) oxide and SAPO-34 bifunctional catalysts: Design of oxide component and construction of reaction network. ChemCatChem 2018, 10, 1536–1541. [Google Scholar]
  33. Zhu, H.; Wang, X.; Yang, F.; Yang, X. Template-free, surfactantless route to fabricate In(OH)3 monocrystalline nanoarchitectures and their conversion to In2O3. Cryst. Growth Des. 2008, 8, 950–956. [Google Scholar]
  34. Luo, M.-F.; Yan, Z.-L.; Jin, L.-Y.; He, M. Raman spectroscopic study on the structure in the surface and the bulk shell of CexPr1-xO2-δ mixed oxides. J. Phys. Chem. B 2006, 110, 13068–13071. [Google Scholar] [PubMed]
  35. Jiang, X.; Nie, X.; Gong, Y.; Moran, C.M.; Wang, J.; Zhu, J.; Chang, H.; Guo, X.; Walton, K.S.; Song, C. A combined experimental and DFT study of H2O effect on In2O3/ZrO2 catalyst for CO2 hydrogenation to methanol. J. Catal. 2020, 383, 283–296. [Google Scholar]
  36. Cui, Y.; Xu, L.; Chen, M.; Lian, X.; Wu, C.-e.; Yang, B.; Miao, Z.; Wang, F.; Hu, X. Facilely fabricating mesoporous nanocrystalline Ce–Zr solid solution supported CuO-based catalysts with advanced low-temperature activity toward CO oxidation. Catal. Sci. Technol. 2019, 9, 5605–5625. [Google Scholar]
  37. Gonçalves, A.; Puna, J.F.; Guerra, L.; Campos Rodrigues, J.; Gomes, J.F.; Santos, M.T.; Alves, D. Towards the development of syngas/biomethane electrolytic production, using liquefied biomass and heterogeneous catalyst. Energies 2019, 12, 3787. [Google Scholar] [CrossRef] [Green Version]
  38. Gao, P.; Dang, S.; Li, S.; Bu, X.; Liu, Z.; Qiu, M.; Yang, C.; Wang, H.; Zhong, L.; Sun, Y. Direct production of lower olefins from CO2 conversion via bifunctional catalysis. ACS Catal. 2018, 8, 571–578. [Google Scholar]
  39. Frei, M.; Mondelli, C.; Short, M.; Pérez-Ramírez, J. Methanol as a hydrogen carrier: Kinetic and thermodynamic drivers for its CO2-based synthesis and reforming over heterogeneous catalysts. ChemSusChem 2020, 13, 1–9. [Google Scholar]
Figure 1. (a) XRD patterns of various oxides and Au-supported oxides; (b) HRTEM image of Au/In2O3-ZrO2; (c) scanning electron microscope (SEM) image and corresponding energy-dispersive X-ray spectroscopy (EDX) mapping of Au/In2O3-ZrO2.
Figure 1. (a) XRD patterns of various oxides and Au-supported oxides; (b) HRTEM image of Au/In2O3-ZrO2; (c) scanning electron microscope (SEM) image and corresponding energy-dispersive X-ray spectroscopy (EDX) mapping of Au/In2O3-ZrO2.
Catalysts 10 01360 g001
Figure 2. Comparison of (a) CO2 conversion and methanol selectivity, and (b) methanol space–time yield (STY) over the In2O3, Au/In2O3, and Au/In2O3-ZrO2 catalysts as a function of temperature; (c) CO2 conversion, methanol selectivity, and methanol STY versus time of the Au/In2O3-ZrO2 catalyst on stream (TOS); (d) Arrhenius plots for methanol production over the In2O3, Au/In2O3, and Au/In2O3-ZrO2 catalysts. (Reaction conditions: 5.0 MPa, H2/CO2/N2 = 76/19/5, GHSV = 21,000 cm3 h−1 gcat−1.).
Figure 2. Comparison of (a) CO2 conversion and methanol selectivity, and (b) methanol space–time yield (STY) over the In2O3, Au/In2O3, and Au/In2O3-ZrO2 catalysts as a function of temperature; (c) CO2 conversion, methanol selectivity, and methanol STY versus time of the Au/In2O3-ZrO2 catalyst on stream (TOS); (d) Arrhenius plots for methanol production over the In2O3, Au/In2O3, and Au/In2O3-ZrO2 catalysts. (Reaction conditions: 5.0 MPa, H2/CO2/N2 = 76/19/5, GHSV = 21,000 cm3 h−1 gcat−1.).
Catalysts 10 01360 g002
Figure 3. Visible Raman spectra of fresh samples (a) with the inset showing the enlargement of Raman spectra around 140 cm−1, and the samples after hydrogen reduction for 1 h (b) and 1.5 h (c); the comparison of the relative amount of the oxygen vacancies according to I2/I1 value (d). (H2 reduction conditions: 473 K, 0.1 MPa, 10% H2/Ar.)
Figure 3. Visible Raman spectra of fresh samples (a) with the inset showing the enlargement of Raman spectra around 140 cm−1, and the samples after hydrogen reduction for 1 h (b) and 1.5 h (c); the comparison of the relative amount of the oxygen vacancies according to I2/I1 value (d). (H2 reduction conditions: 473 K, 0.1 MPa, 10% H2/Ar.)
Catalysts 10 01360 g003
Figure 4. H2-TPR profiles (a) and CO2-TPD profiles (b) of In2O3 and Au-supported oxides.
Figure 4. H2-TPR profiles (a) and CO2-TPD profiles (b) of In2O3 and Au-supported oxides.
Catalysts 10 01360 g004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lu, Z.; Sun, K.; Wang, J.; Zhang, Z.; Liu, C. A Highly Active Au/In2O3-ZrO2 Catalyst for Selective Hydrogenation of CO2 to Methanol. Catalysts 2020, 10, 1360. https://doi.org/10.3390/catal10111360

AMA Style

Lu Z, Sun K, Wang J, Zhang Z, Liu C. A Highly Active Au/In2O3-ZrO2 Catalyst for Selective Hydrogenation of CO2 to Methanol. Catalysts. 2020; 10(11):1360. https://doi.org/10.3390/catal10111360

Chicago/Turabian Style

Lu, Zhe, Kaihang Sun, Jing Wang, Zhitao Zhang, and Changjun Liu. 2020. "A Highly Active Au/In2O3-ZrO2 Catalyst for Selective Hydrogenation of CO2 to Methanol" Catalysts 10, no. 11: 1360. https://doi.org/10.3390/catal10111360

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