Open Access Article
Shixian Songa,
Jinyi Weia,
Xuan Hea,
Guangfu Yana,
Mengyan Jiaoa,
Wei Zenga,
Fangfang Daiab and
Midong Shi
*a
aCollege of Chemistry and Environmental Science, Hunan Province Key Laboratory of Rare and Precious Metal Compounds, Xiangnan University, Chenzhou 423000, Hunan, China. E-mail: mdshi032825@xnu.edu.cn
bShaanxi Key Laboratory of Chemical Additives for Industry, Shaanxi University of Science and Technology, Xi'an 710021, China
First published on 2nd November 2021
Oxygen vacancy sites on a catalyst surface have been extensively studied and been proved to promote the adsorption and activation of carbon dioxide. We use Sn-doped ZrO2 to prepare a Zr/Sn catalyst rich in oxygen vacancies (OVs) by co-precipitation. The yield of dimethyl carbonate is 5 times that of ZrO2. Compared with the original ZrO2, Zr/Sn exhibits a higher specific surface area, number of acid–base sites and a lower band gap, which improves the conductivity of electrons and creates more surface. The number of reaction sites greatly enhances the adsorption and activation capacity of CO2 molecules on the catalyst surface. In situ infrared spectroscopy shows that CO2 adsorbs on oxygen vacancies to form monomethyl carbonate, and participates in the reaction as an intermediate species. This work provides new clues for the preparation of ZrO2-based catalysts rich in oxygen vacancies to directly catalyze the synthesis of dimethyl carbonate from methanol and CO2.
Metal oxides have a wide range of applications in the fields of chemistry, physics, and materials science. Their redox properties and acid–base sites are very important for absorption and catalytic applications. In the literature, CeO2, ZrO2, SnO2, TiO2, V2O5, Y2O3, Ga2O3 and other single metal oxides and their composite oxides have been used to catalyze the direct synthesis of DMC from CO2 and methanol.8–12 However, because CO2 has the characteristics of thermodynamic stability and chemical inertness, the route of direct synthesis of DMC still faces limiting factors such as low yield and kinetics/thermodynamics. It is well known that acidity and basicity of the catalytic sites of CO2 adsorption are the key factors that determine the efficiency of CO2 activation.
Oxygen vacancies (OVs) on the surface of the catalyst can not only promote the adsorption and activation of CO2, but also serve as active sites for CO2 conversion in the methanation of CO2.13,14 Oxygen vacancies affect the effect of carbon dioxide catalytic reaction by promoting electron transfer. It has been shown that OVs enhance the catalytic performance of the catalyst by widening the valence band (VB) and narrowing the band gap, and inhibiting photo-induced electron–hole recombination.15,16
Crystal form of zirconia catalyst has a great influence on the acidity and basicity of the catalyst surface and the surface oxygen vacancies. Oxygen vacancies are ubiquitous in metal oxides and have obvious effects on the physical and chemical properties of materials.17,18 Bell et al.19 studied the formation rate of methoxy and methyl carbonate on the tetragonal and monoclinic phases of zirconia, and found that zirconia with a tetragonal crystal phase had a higher density and strength of acid–base sites point, making the formation of methyl groups easier on the tetragonal phase. Methoxy and methyl carbonate are important intermediate species in the direct synthesis of DMC. Compared with pure zirconia, doping other metal ions into zirconia will change the oxygen vacancy concentration and acidity and alkalinity of the catalyst surface.20–22 Tomishige et al.23 found that the catalytic activity of ZrO2 catalyst for the synthesis of DMC is significantly improved compared with pure ZrO2, which is related to the oxygen vacancies generated by the introduction of Ce4+. However, the internal effect of OVs on the catalytic reaction is still elusive, which hinders the rational design of catalyst defects.
Herein, zirconia catalysts with different Sn content were prepared by co-precipitation and their activity in the direct synthesis of dimethyl carbonic from methanol and carbon dioxide was explored. TEM, XRD, N2 adsorption, UV-Vis spectroscopy, Raman spectroscopy, CO2/NH3-TPD, XPS technology and in situ FTIR techniques were employed to identify the structure and composition of the prepared catalysts. The importance of oxygen vacancy for the synthesis of dimethyl carbonate has been studied in detail.
:
1, 1.5
:
1, 2.5
:
1, 4
:
1 and 5.5
:
1 was placed in 250 ml of deionized water, and stirred until completely dissolved. After dissolving, NH3·H2O was added into the above solution to pH = 9–10 and stirring was continued for 1 hour. Then the stirring was stopped, the product was aged and left standing for 12 hours at room temperature. When the aging product was filtered, it was washed with deionized water to neutrality until pH = 7, and washed 3–4 times with ethanol. The product after suction filtration was dried at 100 °C for 12 hours, and then calcined in a muffle furnace at 400 °C for 5 hours to obtain Zrn/Sn, where is the molar ratio of Zr to Sn, and the molar ratio of Zr to Sn is controlled at 1 to 5.5.
![]() | ||
| Fig. 1 Schematic diagram of the synthesis of Sn-doped ZrO2 catalyst and the direct synthesis of dimethyl carbonate from methanol and carbon dioxide. | ||
The phase structure of the catalysts was characterized by XRD (D/max-Ultima IV) using a Cu target Kα-ray (40 kV and 30 mA) as the X-ray source. The scanning range (2θ) was from 10° to 80° with a scanning speed of 8° min−1 and a step function of 0.02.
The specific surface area and pore structure of the catalyst were measured by the Nitrogen adsorption method using the Micromeritics TriStar II Specific Surface Analyzer at the temperature of liquid nitrogen. Before the test, the sample was degassed at a temperature of 180 °C for 6 hours, and the isothermal desorption branch analysis was performed through the Barrett–Joyner–Halenda (BJH) model to calculate the pore volume and pore size distribution, and use the Brunauer–Emmett–Teller (BET) method to estimate the catalyst specific surface area.
The model of the Raman spectrometer used was Horiba Scientific LabRAM HR Evolution installed an Ar laser with a wavelength of 514.5 nm and an output power of 4 mW. The UV-Vis spectrum of the catalyst sample was obtained by Shimadzu UV-2600 UV spectrophotometer.
X-ray photoelectron spectroscopy (XPS) was performed on Thermo Scientific K-Alpha using a 300 W AlK X-ray source, a constant residence time of 100 ms, and a transfer energy of 40 eV. The binding energy refers to the C1s hydrocarbon peak at 284.8 eV.
Temperature programmed desorption (TPD) was performed on the automatic chemical adsorption analyzer PCA-1200. The 0.1 g sample was pretreated at 180 °C for 1 h under He. Then, CO2/NH3 was adsorbed at 25 °C for 60 minutes. After the adsorption was completed, the sample was purified with He for 1 hour to remove physically adsorbed CO2/NH3. At a constant flow rate of He, the test was performed by heating from 50 °C to 600 °C at a heating rate of 10 °C min−1. Use TCD to continuously detect the CO2/NH3 concentration in the exhaust gas.
The in situ infrared spectroscopy experiment was performed on a Thermo Scientific Nicolet IS50 infrared spectrometer with a resolution of 8 cm−1 and 128 scans accumulated. Add 20 mg of catalyst powder to the diffuse reflection cell, heat it for 2 h under a He atmosphere at 180 °C, and measure the background spectrum after it is cooled to room temperature. The flow rates of He and CO2 are 30 ml min−1 and 20 ml min−1, respectively. DMC and methanol are injected into the infrared cell through He gas flow, and the flow rates are 10 ml min−1 and 20 ml min−1. When CO2, methanol and DMC are introduced into the reaction unit, the temperature is maintained at 25 °C.
HRTEM image can determine the crystallization characteristics of the Zrn/Sn catalysts. The lattice fringes of 0.314 nm, 0.261 nm and 0.362 nm can be respectively referred to as the (−111), (002) and (011) planes of monoclinic ZrO2 [PDF 86-1450], while the lattice fringes of 0.284 nm and 0.171 nm are due to (002) and (103) planes of the orthogonal ZrSnO4 [PDF 48-0889]. In the SnO2 low-magnification TEM image, the existence of particles in the dispersed state is clearly observed. The lattice fringe data of 0.331, 0.261 and 0.160 nm obtained by calculating HRTEM images confirmed the existence of (110), (101) and (220) planes of SnO2 [PDF 41-1445], respectively.
| Catalysts | 2θ | Interplanar spacinga (Å) | Lattice parameterb | Average crystallite sizec (Å) | SBET (m2 g−1) | Vpore (cm3 g−1) | ||
|---|---|---|---|---|---|---|---|---|
| a | b | c | ||||||
| a Calculated by (110) plane from XRD.b Calculated for Zr1/Sn and Zr1.5/Sn catalysts by tetragonal ZrO2, for ZrO2, Zr2.5/Sn, Zr4/Sn and Zr5.5/Sn by monoclinic, for SnO2 by tetragonal.c Estimated from XRD by Debye–Scherrer's equation. | ||||||||
| ZrO2 | 24.07 | 3.71 | 5.18 | 5.21 | 5.35 | 93 | 80.07 | 0.147 |
| Zr1/Sn | 23.92 | 3.64 | 4.95 | 5.20 | 5.73 | 127 | 56.91 | 0.134 |
| Zr1.5/Sn | 23.93 | 3.65 | 4.91 | 5.19 | 5.71 | 110 | 87.81 | 0.133 |
| Zr2.5/Sn | 23.96 | 3.67 | 5.17 | 5.19 | 5.34 | 89 | 84.12 | 0.129 |
| Zr4/Sn | 24.01 | 3.68 | 5.16 | 5.17 | 5.31 | 84 | 103.07 | 0.121 |
| Zr5.5/Sn | 24.03 | 3.70 | 5.15 | 5.14 | 5.30 | 79 | 126.03 | 0.109 |
| SnO2 | 26.58 | 3.35 | 4.73 | 4.71 | 3.19 | 56 | — | — |
According to reports by Štefanić et al.,25 metastable ZrSnO4 will only appear in calcination systems at 1000 °C or higher. However, according to our research, in Zr1/Sn and Zr1.5/Sn, 2θ = 31.5°, 51.2° show the (002) and (103) crystal plane growth of orthogonal ZrSnO4 [PDF 48-0889].26 The formation of ZrSnO4 is due to the solubility of Sn4+ ions in the ZrO2 lattice obtained upon crystallization. However, it is not ruled out that other Zrn/Sn samples also contain a small amount of ZrSnO4. This is because the diffraction peaks of ZrSnO4 and that of t-ZrO2 and m-ZrO2 overlap significantly. With the increase of zirconia content, the diffraction peak intensity of SnO2 gradually weakened, which may be caused by the excessive dispersion of SnO2 in the ZrO2 matrix. At the same time, the (111) crystal plane of tetragonal ZrO2 [PDF 88-0287] appeared at 2θ = 29.6°.
In our research, it was observed that pure zirconia particles obtained a mixed tetragonal phase and monoclinic phase at an annealing temperature of 400 °C. The presence of tin makes the samples have stable tetragonal ZrO2 (t-ZrO2) crystal form. These crystalline products may be the result of the surface interaction of ZrO2–SnO2, similarly as in the ZrO2/SO42− system that prevents the diffusion of oxygen from the atmosphere into the ZrO2 lattice and triggers the t-ZrO2 to m-ZrO2 transition.27 The surface interaction seems to be an important factor for stabilizing the t-ZrO2 phase in the crystalline product of the ZrO2–SnO2 system. Therefore, calcination in air leaves enough SnO2 on the surface to delay the phase transition from t-ZrO2 to m-ZrO2 in the ZrO2–SnO2 system.24
| Catalysts | Zr1/Sn | Zr1.5/Sn | Zr2.5/Sn | Zr4/Sn | Zr5.5/Sn |
|---|---|---|---|---|---|
| I190/I177 | 3.66 | 1.72 | 1.69 | 1.66 | 1.38 |
| I190/I75 | 1.16 | 1.47 | 2.71 | 4.35 | 5.71 |
Since the Raman shift of Sn is not observed in Zrn/Sn, it indicates that ZrO2 and SnO2 form a solid solution. For pure SnO2, the vibration modes at 73 cm−1, 633 cm−1 and 770 cm−1 are attributed to the B1g, A1g and B2g vibration modes, respectively, which belong to the natural mode of tetragonal rutile.32 The A1g and B2g modes may be related to the stretching vibration of the Sn–O bond, while B1g is more related to the oxygen atom rotating around the C axis.
With the increase of Sn dopant, the red shift of the interference-free region can be observed, indicating that the band gap of the Zr/Sn catalyst is significantly reduced (from 4.58 eV to 2.70 eV). This may be due to the sp-d spin-exchange interaction between the electron-carrying electron at the cation site and the local d-electron of the Zr ion or the increase in the surface area to volume ratio.33 The first interaction lowers the bottom of the conduction band, and the second interaction increases the top of the valence band, narrowing the band gap. The introduction of impurity bands and the trapping of Zr atoms at the grain boundaries result in the formation of Zr defect states in the forbidden band.34 As Zr doping increases, the density of Zr-induced defect states increases, leading to the observed band gap red shift or decrease.
In order to prove that the introduction of Zr defect states in the band gap region will cause the change of disorder, we introduce a concept 1/D.26 1/D represents the average crystallite size, which is determined by the ratio of the surface area to the volume of the constituent particles of the sample. As the grain size decreases, more and more disorder is introduced into the sample. Therefore, 1/D can be considered as a measure of crystal disorder. Since the radius of Sn ions is smaller than that of Zr ions, as the content of Zr increases, more disorder may be introduced into the particles. Fig. 6g shows that the band gap of the Zr/Sn catalyst increases with increasing particle disorder.
The results show that Sn doping in ZrO2 effectively reduces the band gap energy, the substitution of Zr4+ with Sn species facilitates the formation of Zr3+ions, leading to a rise of oxygen vacancy concentration in Sn-doped ZrO2 through distorting the lattice structure.35,36
The results of using NH3-TPD to study the acid properties of the catalyst surface are shown in Table 3 and Fig. 7b. It can be observed from the figure that ZrO2, Zrn/Sn and SnO2 catalysts have the same type of acid sites. The desorption peaks in the range of 50–200, 200–400, 400–600 °C can be referred to as weak adsorption, medium adsorption and strong adsorption, respectively. The NH3 adsorption capacity of these catalysts is also calculated and summarized in Table 3. Compared with other catalysts, Zr4/Sn catalyst has the highest NH3 adsorption capacity, indicating that Zr4/Sn catalyst has the most acidic sites. The adsorption amount of NH3 in all samples relative to the content of Zr shows a volcanic curve in the following the order: Zr4/Sn > SnO2 > Zr2.5/Sn > Zr5.5/Sn > Zr1.5/Sn > ZrO2 > Zr1/Sn. The total acidity of Zrn/Sn catalyst is higher than that of pure ZrO2 catalyst (except Zr1/Sn), demonstrating ZrO2 and SnO2 have a synergistic effect. At the same time, the total acidity of Zr1/Sn catalyst is lower than that of the ZrO2 catalyst, which may be due to the lower BET surface area (56.91 m2 g−1) and pore volume of Zr1/Sn catalyst than the pure ZrO2 catalyst (80.07 m2 g−1), resulting in a smaller number of exposed Lewis acid sites. Similar to the results of CO2-TPD, with the increase of Zr content, the weakly acidic sites of all Zrn/Sn catalysts first increase and then decrease, and then increase and finally decrease. It can be found that the Zr4/Sn catalyst has the largest number of acidic sites. By doping with oxides, the acid–base properties of ZrO2 surface will be significantly changed, accompanied by differences in performance, similar results are obtained from previous reports.38–40
| Acid sites distribution of NH3-TPD measurement dataa | Basic sites distribution of CO2-TPD measurement datab | |||||||
|---|---|---|---|---|---|---|---|---|
| Weakc | Moderate | Strong | Total | Weak | Moderate | Strong | Total | |
| a Set the amount of acidic sites of Zr4/Sn as 1.00, and compare with other samples.b Set the amount of basicity sites of Zr4/Sn as 1.00, and compare with other samples.c Weak (<200 °C), moderate (200 °C to 400 °C), strong (>400 °C). | ||||||||
| SnO2 | 0.26 | 0.55 | 0.17 | 0.98 | 0.12 | 0.13 | 0.12 | 0.36 |
| Zr1/Sn | 0.07 | 0.26 | 0.07 | 0.40 | 0.11 | 0.46 | 0.14 | 0.71 |
| Zr1.5/Sn | 0.19 | 0.50 | 0.08 | 0.77 | 0.12 | 0.19 | 0.01 | 0.32 |
| Zr2.5/Sn | 0.17 | 0.42 | 0.35 | 0.94 | 0.05 | 1.32 | 0.10 | 1.47 |
| Zr4/Sn | 0.33 | 0.23 | 0.44 | 1.00 | 0.13 | 0.49 | 0.38 | 1.00 |
| Zr5.5/Sn | 0.26 | 0.37 | 0.20 | 0.83 | 0.09 | 0.44 | 0.20 | 0.73 |
| ZrO2 | 0.03 | 0.32 | 0.11 | 0.41 | 0.11 | 0.75 | 0.58 | 1.44 |
Since Sn4+ enters the lattice of zirconia, the binding energy of Zr3d in the Zrn/Sn catalysts (at 182.1 eV, 184.5 eV) is higher than that of zirconia. After deconvolution calculation of the Zr3d spectrum in the ZrO2 catalyst, Zr3d can also be decomposed into 2 groups of peaks: Zr(4−x)+ (∼181.7 eV), ZrO2 (∼182.2 eV).42,43 In Table 4, by calculating the ratio of the area of the Zr(4−x)+ peak to the total area of the Zr(4−x)+ and ZrO2 peaks, the concentration ratio of Zr(4−x)+ can be quantitatively estimated.
| Samples | Zr3d (%) | O1s (%) | Surface atomic ratio Zr/Sna (%) | Relative peak intensity Zr/Ceb (%) | |||
|---|---|---|---|---|---|---|---|
| Zr(4−x)+ | ZrO2 | OL | OV | OC | |||
| a Determined by XPS.b Determined by Raman spectra. The peak intensity at 75 cm−1 in the Raman spectrum of pure SnO2 is recorded as 1, and the peak intensity at 190 cm−1 in the Raman spectrum of pure ZrO2 is recorded as 1. I75/I190 reacts to the relative content change of Zr/Sn. The spectrum has been deconvolved before calculation. | |||||||
| ZrO2 | 4.28 | 95.72 | 75.50 | 9.50 | 15.00 | — | — |
| Zr1/Sn | 7.12 | 92.88 | 77.50 | 18.68 | 3.82 | 1.18 | 1.16 |
| Zr1.5/Sn | 25.41 | 74.59 | 70.40 | 20.20 | 9.40 | 1.64 | 1.79 |
| Zr2.5/Sn | 31.26 | 68.74 | 67.20 | 23.20 | 9.60 | 2.41 | 2.72 |
| Zr4/Sn | 71.12 | 28.88 | 48.41 | 37.6 | 13.83 | 4.44 | 3.87 |
| Zr5.5/Sn | 40.58 | 59.42 | 63.90 | 26.30 | 9.80 | 5.55 | 5.82 |
| SnO2 | — | — | 52.4 | 26.1 | 21.5 | — | — |
The percentages of Zr(4−x)+ to the total Zr (Zrtotal) are 4.28%, 7.12%, 25.41%, 31.26%, 71.12% and 40.58% for the ZrO2, Zr1/Sn, Zr1.5/Sn, Zr2.5/Sn, Zr4/Sn and Zr5.5/Sn (Table 4), respectively. The ratio of Zr(4−x)+ to Zrtotal decreased in the order: Zr4/Sn > Zr5.5/Sn > Zr2.5/Sn > Zr1.5/Sn > Zr1/Sn > ZrO2. Furthermore, the Zr(4−x)+/Zrtotal ratio of Zrx/Sn catalysts is higher than the pure ZrO2 sample. Table 4 shows that in the Zrn/Sn catalysts, Zr(4−x)+ increases with the increase of the Zr content, and then decreases with the further increase of the Zr content. This reduction is believed to be caused by catalyst aggregation caused by excess zirconium.44 The presence of Zr(4−x)+ was confirmed to generate oxygen vacancies on the catalyst surface. Therefore, the concentration of oxygen vacancies increases as the concentration of Zr increases. Compared with other Zrn/Sn catalysts, Zr4/Sn catalysts have the highest concentration of oxygen vacancies because they have the highest Zr(4−x)+/Zrtotal.
Fig. 8b shows the O1s XPS results on Zrn/Sn, SnO2 and ZrO2 catalysts. The first peak is attributed to the lattice oxygen (OL) in the ZrO2 lattice at ∼530.1 eV. The peaks at ∼531.5 and ∼532.7 eV are attributed to the O component associated with the O2− ions in surface oxygen vacancies (OV) and chemisorbed oxygen species (OC), respectively.45,46 Interestingly, the O1s high-resolution XPS spectrum of the Zrn/Sn catalyst (compared to pure zirconia) has a long tail, and hopes to have a higher binding energy. The O1s spectrum of Zrn/Sn is further deconvolved into three peaks. The component at higher binding energy could be due to C–O, O2− peroxides, or OH.47,48 The results of the quantitative calculations of the oxygen vacancy concentration on the catalyst surface are shown in Table 4.
The results show that the addition of Sn significantly reduces the lattice oxygen in the catalyst. The concentration of OV is in good agreement with the results of the Zr(4−x)+ concentration in the Zr3d spectrum. Due to the imbalance of oxidation state between Zr4+ and Zr(4−x)+ sites, oxygen vacancies are generated, resulting in an increase in the concentration of surface oxygen vacancies. Sn doping in the zirconia lattice promotes the formation of surface oxygen vacancies/defects.49
The XPS spectrum of Sn3d is shown in Fig. 8c. The Zrn/Sn catalysts showed two peaks at 495.6–495.7 eV and 487.2–487.5 eV. The peak at higher binding energy belongs to Sn3d3/2, and the peak at lower binding energy belongs to Sn3d5/2, which may be related to the spin–orbit splitting of Sn3d3/2 and Sn3d5/2.50 The binding energy of Sn3d in the Zrn/Sn catalyst is higher than that of tin dioxide, which further confirms that Sn4+ enters the zirconia lattice and forms a solid solution.
:
CO2 = 192
:
200, reaction time was 2 h, and catalyst dosage was 0.5 g). This is not the only case where lower DMC yields are also obtained. In addition, Ikeda et al.8 found that the highest yield of DMC catalyzed by H3PO4/ZrO2 was 0.3 mmol (reaction time was 2 h; reaction temperature was 403 K; CH3OH/CO2 = 192/200 mmol; catalyst weight was 0.5 g). This result is equally unsatisfactory, but it is what it is.
The activity of Zr4/Sn is 3.9 times and 4.7 times that of pure zirconium dioxide and tin dioxide, respectively. The order of catalyst activity from high to low is: Zr4/Sn > Zr2.5/Sn > Zr5.5/Sn > Zr1.5/Sn > Zr1/Sn > ZrO2 > SnO2, which followed the trend of the surface oxygen vacancy concentrations. It can be observed that the methanol conversion rate increases with the increase of the Zr/Sn ratio, reaching the maximum peak value of 0.17% at Zr4/Sn, and then decreases with the further increase of the Zr/Sn ratio.
The stability of heterogeneous catalysts has always been one of the concerns (Fig. 10b). Under the same reaction conditions as above, five reaction cycles were used to test the reusability of the catalyst in DMC synthesis. The activity of all catalysts decreased to varying degrees after five cycles of reaction. However, the catalytic activity of Zr4/Sn is still the highest. The activities of the catalysts were reduced by 57.5, 56.7, 51.3, 31.2, 44.8, 62.9 and 59.9% after five cycle times corresponding to Zr1/Sn, Zr1.5/Sn, Zr2.5/Sn, Zr4/Sn, Zr5.5/Sn, SnO2, ZrO2, respectively.
By comparing the XPS spectra of Zr3d and O1s of fresh and used Zr4/Sn after deconvolution, it is found that the concentration of Zr(4−x)+ and surface oxygen vacancies dropped from 71.12% and 37.6% to 11.03% and 9.1%, respectively. The reason for this reduction is that part of the O atoms used to adsorb carbon dioxide is used to fill the surface oxygen vacancies.51 Compared to fresh, the fourth peak is attributed to adsorbed water (∼535.5 eV), this may be caused by too much water adsorbed in the air.52 A large number of DFT studies and experimental studies have shown that the surface oxygen vacancies are thermodynamically unstable and highly reactive, as well as the interaction between the surface oxygen vacancies and the O atoms that adsorb carbon dioxide.53,54
Among these catalysts, Zr4/Sn has the greatest catalytic activity because it has the highest surface oxygen vacancy concentration. The oxygen vacancies on the surface of the oxygen defect of the metal oxide can adsorb and activate carbon dioxide, and promote the direct reaction of methanol and carbon dioxide to form DMC.55,56 According to previous reports, the activation of carbon dioxide is achieved by filling the oxygen vacancies with free oxygen generated by the interaction at the vacancies, which is easier than on the metal surface.57
The methanation of carbon dioxide proves that surface oxygen vacancies can catalyze the key step that determines the rate at a lower activation temperature.58,59 According to the results obtained in this work and previous reports, tin doping into the zirconia lattice will increase the number of active oxygen vacancies on the surface of zirconia, thereby promoting the interaction between Zrn/Sn and carbon dioxide.25,26,39
![]() | ||
| Fig. 13 DRIFTS spectra of adsorbed species on Zr4/Sn catalyst: (a) adsorbed CO2; (b) adsorbed methanol; (c) comparison of adsorbed DMC and CO2 with adsorbed methanol. | ||
Fig. 13b shows that when methanol is adsorbed on ZrO2 at room temperature, they appear at 1016 cm−1 in addition to the apparent linear (1160 cm−1) and two bridging (1052 and 1031 cm−1) methoxy species.63 Another cone-bridged methoxy species may be related to the abundant surface hydroxyl groups on the surface of ZrO2.63 The two negative peaks at 3755 and 3650 cm−1 may be caused by the reaction of surface hydroxyl groups with methanol. At the same time, the consumption of bridged hydroxyl groups is more than that of linear hydroxyl groups, which indicates that the bridged hydroxyl groups on the surface of ZrO2 are more active than linear hydroxyl groups. Due to the effect of surface lattice oxygen, formate species appeared at 1375 and 1360 cm−1, and obvious carbonate species appeared at 1550 and 1412 cm−1.64
By exposing the catalyst to methanol for 30 minutes, flushing the infrared cell with He, and then introducing the He stream containing CO2, the interaction between CO2 and the adsorbed methanol material was studied. After the introduction of CO2 in the methanol-adsorbed sample, the spectrum is very similar to the methanol adsorption on the ZrO2 surface without pre-adsorption. In Fig. 13c, the bands appearing at 1595, 1497, 1370, and 1201 cm−1 can be attributed to carbonic acid monomethyl ester (m-CH3OCOO–Zr), and the bands at 1157 and 1032 cm−1 are caused by methoxy substances.19,65,66 The adsorption of DMC on the catalyst further verified the existence of monomethyl carbonate, which is an intermediate species in the direct synthesis of DMC from methanol and CO2. Studies have shown that the high selectivity of DMC on the catalyst may be due to the rapid conversion of methoxy species to methoxy carbonate species under high carbon dioxide pressure.67,68
The carbon dioxide is adsorbed on the oxygen vacancies as diacid carbonate.70 Then methanol is cleaved at adjacent acidic sites through OM–H4 bonds to form methyl and methoxy groups.68 Many studies have shown that acid–base sites are critical to the activation of carbon dioxide, which will facilitate the synthesis of DMC.71–73 The activated carbon dioxide is inserted into the oxidation bond of the methoxy species to produce monomethyl carbonate (Zr–OC(O)OCH3), and the methyl group activated by methanol at the acidic site and methyl carbonate produce DMC.19 Oxygen vacancy has the same effect as Lewis acid sites. The interaction with C and O in carbon dioxide and the Lewis acid–base pair (Zr4+O2−) on the catalyst surface together promote this process. The reaction of the methyl carbonate and the methyl group of the methoxy group will regenerate dimethyl carbonate and oxygen vacancy.
The results of XPS and CO2-TPD show that doping Sn in zirconia can promote the generation of surface oxygen vacancies. Carbon dioxide will be adsorbed as bidentate carbonate on the oxygen vacancies.74
| This journal is © The Royal Society of Chemistry 2021 |