Direct Synthesis of Dimethyl Carbonate from CO2 and Methanol by Mg-Doped Ceria Monolithic Catalyst

Yue-Ying YAN Yue LI Jie DENG Xi ZHAO Na TA Yong-Dong CHEN

Citation:  Yue-Ying YAN, Yue LI, Jie DENG, Xi ZHAO, Na TA, Yong-Dong CHEN. Direct Synthesis of Dimethyl Carbonate from CO2 and Methanol by Mg-Doped Ceria Monolithic Catalyst[J]. Chinese Journal of Inorganic Chemistry, 2022, 38(7): 1402-1410. doi: 10.11862/CJIC.2022.139 shu

镁掺杂氧化铈整体式催化剂催化CO2和CH3OH直接合成碳酸二甲酯

    通讯作者: 塔娜, tana@dicp.ac.cn
    陈永东, yongdongchen@swpu.edu.cn
  • 基金项目:

    国家自然科学基金 21773189

    四川省中央引导地方科技发展专项 2021ZYD0044

    四川省科技厅项目 19ZDZX0113

    西南石油大学科研启动项目 2021QHZ023

摘要: 采用共沉淀法成功地合成了不同Mg掺杂量的Ce1-xMgxO2(x=0.05、0.10、0.15、0.20)固溶体催化材料,并运用透射电子显微镜(TEM)、X射线衍射(XRD)、氮气吸附-脱附测试、拉曼光谱、X射线光电子能谱(XPS)、CO2程序升温脱附(CO2-TPD)等技术对这些材料进行了表征。结果发现,通过调控CeO2晶格中Mg的含量,可以调控所制备的Ce1-xMgxO2催化材料的粒径、比表面积、表面缺陷等。其中Ce0.90Mg0.10O2展现了最佳的表面性质,具有最小的平均粒径(约5.8 nm),最大的比表面积(约136 m2·g-1)以及最高的表面氧含量(31.98%)。将Ce1-xMgxO2催化材料涂覆在堇青石蜂窝陶瓷上制成整体催化剂,考察其对CO2和CH3OH直接合成碳酸二甲酯的催化性能。在140℃、2.4 MPa、反应2 h的条件下,Ce0.90Mg0.10O2整体催化剂上碳酸二甲酯的收率高达20.21%,催化效果明显优于CeO2和其余的Ce1-xMgxO2(x=0.05、0.15、0.20)催化材料。

English

  • Dimethyl carbonate (DMC) has been widely applied as a fuel additive, in electrochemistry and organic synthesis due to its environmental-friendly properties[1-3]. Although many methods have been applied for DMC synthesis, such as phosgene method, transesterification method, urea alcoholysis method, epoxy alkane method, methanol, and CO2 direct synthesis method, etc.[4-6]. Direct synthesis of DMC from CO2 and methanol has attracted great attention (Scheme 1)[7]. The utilization of CO2 as the carbon source instead of fossil feedstock may promote the sustainability of the chemical industry and terminate the greenhouse effect caused by excessive CO2 emission. However, there are still some vital challenges such as low yield, deactivation of the catalyst, and thermodynamic limitations for this route[6, 8]. Thus, designing novel catalysts and developing efficient water removing methods from the reaction mixture are crucial to overcoming the thermodynamic equilibrium of the reaction.

    Scheme 1

    Scheme 1.  Direct synthesis of DMC from CO2 and methanol

    Ceria-based nanomaterials have been widely studied in the direct synthesis of DMC from CO2 and methanol. This is mainly due to its fascinating CO2 capture ability which significantly affects the reaction efficiency. Inert CO2 molecular in the gas phase needs to be adsorbed and activated by the surface oxygen vacancy sites and then can react with methanol to generate DMC[9-12]. Doping metal ions while maintaining the fluorite crystalline structure of ceria is one of the effective ways to enhance the concentration of surface oxygen vacancy of CeO2[9, 13-14]. Because the impurity ions can reduce the crystalline size, generate more surface defects and boost the reducibility of surface oxygen[15-16]. On another hand, the surface acid - base property of CeO2 can be mediated by the doping method, which will further favor the formation of DMC to improve the selectivity[9]. According to Scheme 1, the reaction equilibrium can shift toward the right side by water removal[17]. Usually, inorganic dehydrating agents are introduced to physically remove water with limited effect due to the low dehydration capacity at reaction temperatures[18-21]. While organic dehydrating agents are applied to remove water by chemical reactions which may form lots of by-products complicating the entire process[22-25]. Coating the catalyst powder on the surface of cordierite honeycomb ceramics can improve the phase-phase mass transfer performance[26-28]. Therefore, it is reasonable to expect an enhanced efficiency for water removal using a honeycomb structure catalyst, which will improve the DMC yield in return.

    In this contribution, Ce1-xMgxO2 (x=0.05, 0.10, 0.15, 0.20) solid solutions with a variation of magnesium content were prepared by the co-precipitation method to find an optimal ratio. Mg ions doping in CeO2 lattice adjusted the surface acid-base property and the surface oxygen vacancies. Among all the obtained catalytic materials, Ce0.90Mg0.10O2 was found to show the best catalytic activity in the direct synthesis of DMC from methanol and carbon dioxide. Using a unique structure, monolithic catalyst produced by coating powder on cordierite honeycomb ceramics showed high effective and stable catalytic performance. At 140 ℃, 2.4 MPa, and 2 h continuous reaction, the yield of DMC over Ce0.90Mg0.10O2 monolithic catalyst was the highest (20.21%).

    The preparation of Ce0.90Mg0.10O2 by the coprecipitation method is described as an example. We weighed 15.000 0 g (NH4)2Ce(NO3)6, 0.779 5 g Mg(NO3)2 ·6H2O, and 70.000 0 g urea (CH4N2O) and dissolved them completely with 500 mL deionized water under stirring. The mixture was transferred to a 1 000 mL three - neck flask and gradually heated to 90 ℃ under mechanical stirring (600 r·min-1) for 5 h. After the reaction, the product was cooled to room temperature naturally, the precipitate was filtered and washed with water (over 4 000 mL) and absolute ethanol (about 300 mL), dried overnight at 80 ℃, and calcined at 400 ℃ for 4 h in the air to obtain the target product. The obtained Ce1-xMgxO2 powder was ground with the required deionized water to obtain a slurry, which was coated on a cordierite honeycomb ceramics (64 cells per cm2, Φ: 10 mm, L: 25 mm). The load was maintained at 0.5 g, and the excess slurry was blown away. Finally, the coated catalyst was dried overnight at 80 ℃ and calcined at 400 ℃ for 4 h in the air to obtain a Ce0.90Mg0.10O2 monolithic catalyst. The preparation method of Ce0.95Mg0.05O2, Ce0.85Mg0.15O2, and Ce0.80Mg0.20O2 monolithic catalysts were the same as above, only the mass of Mg(NO3)2·6H2O was changed.

    The catalytic activity of the prepared catalyst for the direct synthesis of DMC from CO2 and methanol was evaluated in a continuous fixed-bed reactor. Water was the main disadvantageous factor for the formation of DMC in the synthesis reaction. The flow of the reaction system can remove the water vapor well and detect the reaction products online. A typical procedure was to place the prepared Ce1-xMgxO2 monolith catalyst in a stainless steel reaction tube. The reactor was sealed and purged with a CO2 stream for 30 min to drain the internal air. When the reaction system reached the required temperature, a mixed gas stream of CH3OH (0.145 mL·min-1) and CO2 (40 mL·min-1) (nCH3OHnCO2=2∶1) was introduced. Then the reaction was carried out at 140 ℃, 2.4 MPa, and 2 880 h-1 of gas hourly space velocity (GHSV). The outlet component after the reaction was analyzed online using gas chromatography (Agilent 7890B) equipped with a hydrogen flame ionization detector. The calculation formula for CH3OH conversion and DMC selectivity is as follows:

    $ {\rm{Conversion}} = \frac{{2{c_{{\rm{DMC}}}} + {c_{{\rm{HCHO}}}} + 2{c_{{\rm{DME}}}}}}{{{c_{{\rm{C}}{{\rm{H}}_3}{\rm{OH}}}} + 2{c_{{\rm{DMC}}}} + 2{c_{{\rm{DME}}}} + {c_{{\rm{HCHO}}}}}} \times 100\% $

    (1)

    $ {\rm{Selectivity}} = \frac{{{c_{{\rm{DMC}}}}}}{{{c_{{\rm{DMC}}}} + {c_{{\rm{DME}}}} + {c_{{\rm{HCHO}}}} + {c_{{\rm{CO}}}}}} \times 100\% $

    (2)

    Where ci represents the concentration of a component (i).

    Fig. 1 shows the X - ray diffraction (XRD) patterns of the prepared Ce1-xMgxO2 composite oxides (Detailed characterization conditions can be found in Supporting Information). CeO2 samples showed typical diffraction lines of cubic fluorite structure (PDF No. 43-1002). Besides, it can be seen that the catalyst doped with Mg2+ still maintained the characteristic peak of cubic fluorite ceria after calcination, no diffraction line representing MgO or any other impurities was detected. Compared with pure CeO2, the (111) plane peak shifted to a higher angle with increased Mg concentration (Fig. 1b), indicating a lattice contraction. The calculated lattice constant decreased from 0.541 8 nm for CeO2 to 0.540 6 nm for Ce0.80Mg0.20O2 (Table 1) because the ionic radius of Mg2+ (0.089 nm) is smaller than that of Ce4+ (0.097 nm). The XRD patterns imply that the Mg2+ incorporate into the CeO2 lattice forming no MgO species and part of them substitutes the Ce4+ leading to lattice contraction. These results are in good agreement with previous reports[15, 29-30]. The calculated grain size from (111) for all samples ranges from 5.8 to 6.1 nm and the specific surface area is basically the same, indicating that the addition of Mg has little influence on the micro-textural property.

    Figure 1

    Figure 1.  (a) XRD patterns of Ce1 -xMgxO2 composite oxides and (b) zoomed-in view of the (111) plane peak

    Table 1

    Table 1.  Structural and textural properties of Ce1-xMgxO2 composite oxides
    下载: 导出CSV
    Catalyst (111) plane Lattice parametera / nm Particle sizeb / nm SBET / (m2·g-1) VPore / (m3·g-1)
    2θ / (°) d / nm
    CeO2 28.510 0.312 8 0.541 8 8.6 133 0.131
    Ce0.95Mg0.05O2 28.709 0.311 3 0.540 5 6.7 129 0.142
    Ce0.90Mg0.10O2 28.730 0.310 0 0.540 1 5.8 136 0.188
    Ce0.85Mg0.15O2 28.770 0.310 1 0.539 8 6.1 117 0.129
    Ce0.80Mg0.20O2 28.757 0.310 2 0.540 6 6.3 126 0.168
    a Calculated using Vegard′s law; b Estimated by TEM.

    The N2 adsorption - desorption isotherms and pore size distributions of Ce1 -xMgxO2 catalyst are shown in Fig.S1. As shown in Fig.S1, all catalysts obtained type Ⅳ isotherms with clear H3 hysteresis lines, indicating typical mesoporous materials. In Fig. S2, all catalysts contain mesopore pore size distributions with pore sizes ranging from 2 to 20 nm. The above results show that the Mg2+ content has a significant effect on the pore size distribution. The BET (Brunauer - Emmett - Teller) surface area and pore volume of the synthesized Ce1-xMgxO2 catalyst are summarized in Table 1. It can be observed that Ce0.90Mg0.10O2 composite oxide possesses the highest specific surface area of 136 m2·g-1 and pore volume of 0.188 cm3·g-1.

    Transmission electron microscope (TEM) images (Fig. 2) of as-prepared Ce1-xMgxO2 composite oxides indicated that all samples were in irregular spherical shape exposing no specific facets. The average particle size of as - prepared Ce1-xMgxO2 is consistent with the grain size.

    Figure 2

    Figure 2.  (a-e) TEM images of Ce1-xMgxO2 composite oxides; (f) Size distribution of Ce0.90Mg0.10O2

    There are two bands observed in Raman spectra (Fig. 3). The vibration peak around 461 cm-1 can be attributed to the F2g vibrational mode of Ce—O, which usually shows a sharp and symmetric band at 466 cm-1[9, 31]. Considering the high specific surface area of the prepared material, the peak shifted to low frequency and showed asymmetric character, which are mainly attributed to the small particle size. Compared with asprepared CeO2 nanoparticles, the F2g band gradually blue-shifted with increased Mg2+ content, which demonstrates the decreased average length of Ce—O bond and lattice contraction further. Therefore, it is reasonable to deduce that smaller Mg2+ cations substitute some Ce4+ ions in the fluorite lattice. It is also noted that the intensity of F2g decreased with increased Mg2+ content, revealing structural distortion[32-33]. Another band near 596 cm-1 is related to the oxygen vacancies caused by the Ce3+ ion in the CeO2 lattice (Fig. 3b)[34]. The intensity of this mode increased with an increase of Mg2+ content, pointing at increased intrinsic oxygen vacancy concentration. No Raman shifts of MgO were observed in Ce1-xMgxO2, which further infers Ce1-xMgxO2 prefers a solid solution state.

    Figure 3

    Figure 3.  (a) Raman spectra of Ce1-xMgxO2 composite oxides and (b) zoomed-in view of the peak attributed to oxygen vacancies

    To elaborate on changes in the CeO2 chemical state after Mg doping, X -ray photoelectron spectroscopy (XPS) analysis was carried out. The XPS spectra of Ce3d (Fig. 4a) exhibit complex features with eight peaks. U and V represent spin - orbits of Ce3d3/2 and Ce3d5/2, respectively. Spin-orbit doublet (V‴ ca. 898.3 eV and U‴ ca. 916.8 eV, V″ ca. 888.9 eV and U″ ca. 907.4 eV, V ca. 882.4 eV and U ca. 900.9 eV) are attributed to the Ce4+ species, while (V′ ca. 884.9 eV and U′ ca. 903.4 eV) are assigned to the Ce3+ species[29, 31]. Then the concentration of Ce3+ can be estimated by taking the ratio of the area of the integrated peak corresponding to Ce3+ to the total area of fitted peaks. It is shown that Mg doping has enhanced the concentration of Ce3+ on the surface remarkably, and the maximum ratio (19.42%) has been obtained when 10% Mg doping. The O1s XPS spectra (Fig. 4b) of Ce1-xMgxO2 composite oxides can be deconvoluted into 3 surface oxygen species: lattice oxygen (OL ca. 529.3 eV), surface oxygen vacancies (OV ca. 530.5 eV); and chemisorption oxygen species (OC) at the highest binding energy (ca. 532.2 eV)[35]. The intensity ratio of surface oxygen vacancies to the sum of all oxygen species was summarized in Table 2. It was observed that the incorporation of Mg2+ can effectively increase the number of surface oxygen species (OV+OC). These results confirm that there are enhanced mobility and availability of lattice oxygen species due to the synergistic effect between MgO and CeO2.

    Figure 4

    Figure 4.  (a) Ce3d and (b) O1s XPS spectra of Ce1-xMgxO2 composite oxides

    Table 2

    Table 2.  Relative ratio of Ce3+ species and oxygen vacancies on the surface
    下载: 导出CSV
    Catalyst Surface content of Ce3+ / % Surface content of OV+OC / % Total amount of adsorbed CO2a / (mmolCO2·gcat-1)
    CeO2 1.34 24.97 0.787
    Ce0.95Mg0.05O2 11.57 30.85 0.794
    Ce0.90Mg0.10O2 19.42 31.98 0.845
    Ce0.85Mg0.15O2 15.18 25.06 0.808
    Ce0.80Mg0.20O2 15.40 23.19 0.771
    aDetermined using the CO2 temperature-programmed desorption (CO2-TPD) analysis in Fig.S3.

    Fig. 5 shows the temperature - programmed reduction by hydrogen (H2-TPR) profile of as-prepared Ce1-xMg xO2 composite oxides. The TPR of pure CeO2 showed a broad peak starting at 500 ℃ and one peak at 825 ℃, representing the surface and the bulk reduction process, respectively. The surface reduction initiated around a lower temperature 500 ℃ after Mg2+ ions (less than 20%) were introduced, which means the reducibility of surface oxygen species has been significantly improved. Meanwhile, the area of this broad peak increased gradually with higher Mg concentration as well, indicating the lattice oxygen in bulk can move to the surface and participate in chemical reactions at a relatively lower temperature. Thus not only the reducibility of surface oxygen but also the mobility of lattice have activated due to Mg2+ introduction, resulting in more oxygen vacancies, probably by reducing the interaction between Ce— O with a distorted crystalline structure. This feature will facilitate chemical reactions whose reactants would be activated by oxygen vacancies. According to the related literature, the oxygen vacancy is crucial for activating carbon dioxide in the direct synthesis of DMC from CO2 and methanol[11, 34, 36].

    Figure 5

    Figure 5.  H2-TPR profiles of CeO2 and Ce1-xMgxO2 composite oxides

    Fig. 6a illustrates photographs of as-prepared monolithic catalyst. A scanning electron microscope (SEM) image (Fig. 6) revealed that the average thickness of the catalyst coating was ca. 60 μm. Well uniform coating layers were found, as evidenced in the corner, inner, and frontal channel views from the energy dispersion X-ray spectrum (EDS) mappings of Ce0.90Mg0.10O2 -coated monolithic catalyst. The abnormal distribution of Mg is due to a small amount of Mg in cordierite. It also demonstrates that Ce0.90Mg0.10O2- coated monolithic catalyst can be insufficient contact with the reaction gas stream to promote the conversion and the yield of the product[37]. Catalyst activity of monolithic and particulate (40-60 mesh) Ce0.90Mg0.10O2 catalyst was comparatively studied (Fig. 7). It is easy to conclude this monolithic do have enhanced the DMC yield and methanol conversion even though both were carried out in the same fixed bed reactor. Therefore, it is probable that the unique structure of the monolithic catalyst accelerates the water removal and shifts the reaction equilibrium successfully. Fig. 8 shows the performance of Ce1-xMgxO2 monolithic catalysts on direct DMC synthesis. The optimum temperature and optimum pressure can be obtained from Fig.S4 and S5. The activity of the catalyst was Ce0.90Mg0.10O2 > Ce0.95Mg0.05O2 > CeO2 > Ce0.85Mg0.15O2 > Ce0.80Mg0.20O2. When x=0.10, the yield of DMC reached the maximum of 20.21% and decreased with a higher doping concentration. It is mainly reflected in the decrease of DMC selectivity and the increase of HCHO and DME selectivity.

    Figure 6

    Figure 6.  (a) Photographs, (b) SEM image, and (c-e) EDS element mappings on Ce0.90Mg0.10O2-coated monolithic catalyst

    Figure 7

    Figure 7.  Catalytic activity of monolithic and particulate Ce0.90Mg0.10O2 catalyst

    Figure 8

    Figure 8.  Catalytic performance of Ce1-xMgxO2 monolithic catalysts

    Reaction conditions: catalyst: 500 mg, GHSV: 2 880 mL·gcat-1·h-1, nCH3OHnCO2=2∶1, temperature: 140 ℃, pressure: 2.4 MPa

    According to our previous studies, there are the following reaction processes in this process: (Ⅰ) 2CH3OH → CH3OCH3+H2O; (Ⅱ) 2CH3O+CO2 → HCHO+CO+H2O[35]. It can be seen that the doping of Mg can promote the process of (Ⅰ) and (Ⅱ), which leads to a decrease in the selectivity of DMC.

    Figure 9

    Figure 9.  Durability test of Ce0.90Mg0.10O2 monolithic catalyst

    Reaction conditions: catalyst: 500 mg, GHSV: 2 880 mL·gcat-1·h-1, nCH3OHnCO2=2∶1, temperature: 140 ℃, pressure: 2.4 MPa

    To provide referable information for the industry, we examined the stability of Ce0.90Mg0.10O2 monolithic catalyst at 140 ℃ and 2.4 MPa. There is little deactivation for this catalyst (DMC yield from 20.21% to 19.56%) during the 50 h durability test implies it is a quite promising application for the direct synthesis of DMC from CO2 and methanol.

    In conclusion, doping Mg in CeO2 lattice can enhance the catalytic performance on the direct formation of DMC from methanol and CO2. Since Mg2+ ions play an important role in the activation of oxygen species in CeO2 lattice, which favors the oxygen vacancies formation. At the same time, the honeycomb structure of the monolithic catalyst greatly improves the removal of reaction products, overcoming thermodynamic limitations to some extent. Consequently, the yield of DMC and the stability of the catalyst can be improved.

    Supporting information is available at http://www.wjhxxb.cn


    Acknowledgments: We acknowledge XIAO Yong - Li, JIANG Lan for their aid in this work.
    1. [1]

      Schifter I, Gonzalez U, Gonzalez-Macias C. Effects of Ethanol, Ethyl-tert-butyl Ether and Dimethyl Carbonate Blends with Gasoline on SI engine[J]. Fuel, 2016, 183:  253-261. doi: 10.1016/j.fuel.2016.06.051

    2. [2]

      Tundo P, Musolino M, Aricò F. The Reactions of Dimethyl Carbonate and Its Derivatives[J]. Green Chem., 2018, 20:  28-85. doi: 10.1039/C7GC01764B

    3. [3]

      Selva M, Perosa A, Fiorani G. Dimethyl Carbonate: A Versatile Reagent for a Sustainable Valorization of Renewables[J]. Green Chem., 2018, 20:  288-322. doi: 10.1039/C7GC02118F

    4. [4]

      Keller N, Rebmann G, Keller V. Catalysts, Mechanisms and Industrial Processes for the Dimethyl Carbonate Synthesis[J]. J. Mol. Catal. A: Chem., 2010, 317:  1-18. doi: 10.1016/j.molcata.2009.10.027

    5. [5]

      Saavalainen P, Kabra S, Turpeinen E, Oravisjärvi K, Yadav G D, Keiski R L, Pongrácz E. Sustainability Assessment of Chemical Processes: Evaluation of Three Synthesis Routes of DMC[J]. J. Chem., 2015, :  1-12.

    6. [6]

      Tamboli A H, Chaugule A A, Kim H. Catalytic Developments in the Direct Dimethyl Carbonate Synthesis from Carbon Dioxide and Methanol[J]. Chem. Eng. J., 2017, 323:  530-544. doi: 10.1016/j.cej.2017.04.112

    7. [7]

      Dabral S, Schaub T. The Use of Carbon Dioxide (CO2) as a Building Block in Organic Synthesis from an Industrial Perspective[J]. Adv. Synth. Catal., 2019, 361:  223-246. doi: 10.1002/adsc.201801215

    8. [8]

      Cai Q H, Lu B, Guo L J, Shan Y K. Studies on Synthesis of Dimethyl Carbonate from Methanol and Carbon Dioxide[J]. Catal. Commun., 2009, 10:  605-609. doi: 10.1016/j.catcom.2008.11.002

    9. [9]

      Liu B, Li C M, Zhang G Q, Yan L F, Li Z. Direct Synthesis of Dimethyl Carbonate from CO2 and Methanol over CaO-CeO2 Catalysts: The Role of Acidic-Basic Properties and Surface Oxygen Vacancies[J]. New J. Chem., 2017, 41:  12231-12240. doi: 10.1039/C7NJ02606D

    10. [10]

      Marciniaka A A, Henriqueb F J F S, Limac A F F, Alvesd O C, Moreirae C R, Appele L G, Mota C J A. What are the Preferred CeO2 Exposed Planes for the Synthesis of Dimethyl Carbonate? Answers from Theory and Experiments[J]. Mol. Catal., 2020, 493:  111053. doi: 10.1016/j.mcat.2020.111053

    11. [11]

      Fu Z W, Yu Y H, Li Z, Han D M, Wang S J, Xiao M, Meng Y Z. Surface Reduced CeO2 Nanowires for Direct Conversion of CO2 and Methanol to Dimethyl Carbonate: Catalytic Performance and Role of Oxygen Vacancy[J]. Catalysts, 2018, 8:  164. doi: 10.3390/catal8040164

    12. [12]

      Zhao S Y, Wang S P, Zhao Y J, Ma X B. An In Situ Infrared Study of Dimethyl Carbonate Synthesis from Carbon Dioxide and Methanol over Well-Shaped CeO2[J]. Chin. Chem. Lett., 2017, 28:  65-69. doi: 10.1016/j.cclet.2016.06.003

    13. [13]

      Tamboli A H, Chaugule A A, Gosavi S W, Kim H. CexZr1-xO2 Solid Solutions for Catalytic Synthesis of Dimethyl Carbonate from CO2: Reaction Mechanism and the Effect of Catalyst Morphology on Catalytic Activity[J]. Fuel, 2018, 216:  245-254. doi: 10.1016/j.fuel.2017.12.008

    14. [14]

      Fu Z W, Zhong Y Y, Yu Y H, Long L Z, Xiao M, Han D M, Wang S J, Meng Y Z. TiO2-Doped CeO2 Nanorod Catalyst for Direct Conversion of CO2 and CH3OH to Dimethyl Carbonate: Catalytic Performance and Kinetic Study[J]. ACS Omega, 2018, 3:  198-207. doi: 10.1021/acsomega.7b01475

    15. [15]

      Yu Q, Wu X X, Tang C J, Qi L, Liu B, Gao F, Sun K Q, Dong L, Chen Y. Textural, Structural, and Morphological Characterizations and Catalytic Activity of Nanosized CeO2-MOx (M=Mg2+, Al3+, Si4+) Mixed Oxides for CO Oxidation[J]. J. Colloid Interface Sci., 2011, 354:  341-352. doi: 10.1016/j.jcis.2010.10.043

    16. [16]

      Kang K H, Joe W, Lee C H, Kim M, Kim D B, Jang B, Song I K. Direct Synthesis of Dimethyl Carbonate from Methanol and Carbon Dioxide Over CeO2(X)-ZnO (1-X) Nano-Catalysts[J]. J. Nanosci. Nanotechnol., 2013, 13:  8116-8120. doi: 10.1166/jnn.2013.8177

    17. [17]

      Sakakura T, Choi J, Saito Y, Sako T. Synthesis of Dimethyl Carbonate from Carbon Dioxide: Catalysis and Mechanism[J]. Polyhedron, 2000, 19:  573-576. doi: 10.1016/S0277-5387(99)00411-8

    18. [18]

      Honda M, Tamura M, Nakagawa Y, Nakao K, Suzuki K, Tomishige K. Organic Carbonate Synthesis from CO2 and Alcohol over CeO2 with 2-Cyanopyridine: Scope and Mechanistic Studies[J]. J. Catal., 2014, 318:  95-107. doi: 10.1016/j.jcat.2014.07.022

    19. [19]

      Stoian D, Medina F, Urakawa A. Improving the Stability of CeO2 Catalyst by Rare Earth Metal Promotion and Molecular Insights in the Dimethyl Carbonate Synthesis from CO2 and Methanol with 2-Cyanopyridine[J]. ACS Catal., 2018, 8:  3181-3225. doi: 10.1021/acscatal.7b04198

    20. [20]

      Wang S P, Zhou J J, Zhao S Y, Zhao Y J, Ma X B. Enhancements of Dimethyl Carbonate Synthesis from Methanol and Carbon Dioxide: The In Situ Hydrolysis of 2-Cyanopyridine and Crystal Face Effect of Ceria[J]. Chin. Chem. Lett., 2015, 26:  1096-1100. doi: 10.1016/j.cclet.2015.05.005

    21. [21]

      Sakakura T, Saito Y, Okano M, Choi J C, Sako T. Selective Conversion of Carbon Dioxide to Dimethyl Carbonate by Molecular Catalysis[J]. J. Org. Chem., 1998, 63:  7095-7096. doi: 10.1021/jo980460z

    22. [22]

      Marciniak A A, Alves O C, Appel L G, Mota C J A. Synthesis of Dimethyl Carbonate from CO2 and Methanol over CeO2: Role of Copper as Dopant and the Use of Methyl Trichloroacetate as Dehydrating Agent[J]. J. Catal., 2019, 371:  88-95. doi: 10.1016/j.jcat.2019.01.035

    23. [23]

      Bansode A, Urakawa A. Continuous DMC Synthesis from CO2 and Methanol over a CeO2 Catalyst in a Fixed Bed Reactor in the Presence of a Dehydrating Agent[J]. ACS Catal., 2014, 4:  3877-3880. doi: 10.1021/cs501221q

    24. [24]

      Han D M, Chen Y, Wang S J, Xiao M, Lu Y X, Meng Y Z. Effect of Alkali-Doping on the Performance of Diatomite Supported Cu-Ni Bimetal Catalysts for Direct Synthesis of Dimethyl Carbonate[J]. Catalysts, 2018, 8:  302-312. doi: 10.3390/catal8080302

    25. [25]

      Santos B, Silva V, Loureiro J, Rodrigues A E. Adsorption of H2O and Dimethyl Carbonate at High Pressure over Zeolite 3A in Fixed Bed Column[J]. Ind. Eng. Chem. Res., 2014, 53:  2473-2483. doi: 10.1021/ie403830r

    26. [26]

      Chen Y D, Yang Y, Tian S L, Ye Z B, Li G. Highly Effective Synthesis of Dimethyl Carbonate over CuNi Alloy Nanoparticles@Porous Organic Polymers Composite[J]. Appl. Catal. A, 2019, 587:  117275. doi: 10.1016/j.apcata.2019.117275

    27. [27]

      Vita A, Italiano C, Pino L, Frontera P, Ferraro M, Antonucci V. Activity and Stability of Powder and Monolith-Coated Ni/GDC Catalysts for CO2 Methanation[J]. Appl. Catal. B, 2018, 226:  384-395. doi: 10.1016/j.apcatb.2017.12.078

    28. [28]

      Tsa S B, Ma H. A Research on Preparation and Application of the Monolithic Catalyst with Interconnecting Pore Structure[J]. Sci. Rep., 2018, 8:  16605. doi: 10.1038/s41598-018-35021-2

    29. [29]

      Jin S, Bang G, Liu L, Lee C H. Synthesis of Mesoporous MgO-CeO2 Composites with Enhanced CO2 Capture Rate via Controlled Combustion[J]. Microporous Mesoporous Mater., 2019, 288:  109587. doi: 10.1016/j.micromeso.2019.109587

    30. [30]

      Matović B, Luković J, Stojadinović B, Aškrabić S, Zarubica A, Babić B, Dohčević-Mitrović Z. Influence of Mg doping on Structural, Optical and Photocatalytic Performances of Ceria Nanopowders[J]. Process. Appl. Ceram., 2017, 11:  304-310. doi: 10.2298/PAC1704304M

    31. [31]

      Liu H, Zou W J, Xu X L, Zhang X L, Yang Y Q, Tian G, Feng S H. The Proportion of Ce4+in Surface of CexZr1-xO2 Catalysts: The Key Parameter for Direct Carboxylation of Methanol to Dimethyl Carbonate[J]. J. CO2 Util., 2017, 17:  43-49. doi: 10.1016/j.jcou.2016.11.006

    32. [32]

      Alla S K, Mandal R K, Prasad N K. Optical and Magnetic Properties of Mg2+Doped CeO2 Nanoparticles[J]. RSC Adv., 2016, 6:  103491-103498. doi: 10.1039/C6RA23063F

    33. [33]

      Ma X, Lu P, Wu P. Structural, Optical and Magnetic Properties of CeO2 Nanowires with Nonmagnetic Mg2+Doping[J]. J. Alloys Compd., 2017, 734:  22-28.

    34. [34]

      Liu B, Li C M, Zhang G, Yao X, Chuang S, Li Z. Oxygen Vacancy Promoting Dimethyl Carbonate Synthesis from CO2 and Methanol over Zr-Doped CeO2 Nanorods[J]. ACS Catal., 2018, 8:  10446-10473. doi: 10.1021/acscatal.8b00415

    35. [35]

      Chen Y D, Wang H, Qin Z, Tian S, Li G. TixCe1-xO2 Nanocomposites: A Monolithic Catalyst for Direct Conversion of Carbon Dioxide and Methanol to Dimethyl Carbonate[J]. Green Chem., 2019, 21:  4642-4649. doi: 10.1039/C9GC00811J

    36. [36]

      Pu Y F, Xuan K, Wang F, Li A X, Zhao N, Xiao F K. Synthesis of Dimethyl Carbonate from CO2 and Methanol over a Hydrophobic Ce/SBA-15 Catalyst[J]. RSC Adv., 2018, 8:  27216-27226. doi: 10.1039/C8RA04028A

    37. [37]

      Bustamante F, Orrego A F, Villegas S, Villa A L. Modeling of Chemical Equilibrium and Gas Phase Behavior for the Direct Synthesis of Dimethyl Carbonate from CO2 and Methanol[J]. Ind. Eng. Chem. Res., 2012, 51:  8945-8956. doi: 10.1021/ie300017r

  • Scheme 1  Direct synthesis of DMC from CO2 and methanol

    Figure 1  (a) XRD patterns of Ce1 -xMgxO2 composite oxides and (b) zoomed-in view of the (111) plane peak

    Figure 2  (a-e) TEM images of Ce1-xMgxO2 composite oxides; (f) Size distribution of Ce0.90Mg0.10O2

    Figure 3  (a) Raman spectra of Ce1-xMgxO2 composite oxides and (b) zoomed-in view of the peak attributed to oxygen vacancies

    Figure 4  (a) Ce3d and (b) O1s XPS spectra of Ce1-xMgxO2 composite oxides

    Figure 5  H2-TPR profiles of CeO2 and Ce1-xMgxO2 composite oxides

    Figure 6  (a) Photographs, (b) SEM image, and (c-e) EDS element mappings on Ce0.90Mg0.10O2-coated monolithic catalyst

    Figure 7  Catalytic activity of monolithic and particulate Ce0.90Mg0.10O2 catalyst

    Figure 8  Catalytic performance of Ce1-xMgxO2 monolithic catalysts

    Reaction conditions: catalyst: 500 mg, GHSV: 2 880 mL·gcat-1·h-1, nCH3OHnCO2=2∶1, temperature: 140 ℃, pressure: 2.4 MPa

    Figure 9  Durability test of Ce0.90Mg0.10O2 monolithic catalyst

    Reaction conditions: catalyst: 500 mg, GHSV: 2 880 mL·gcat-1·h-1, nCH3OHnCO2=2∶1, temperature: 140 ℃, pressure: 2.4 MPa

    Table 1.  Structural and textural properties of Ce1-xMgxO2 composite oxides

    Catalyst (111) plane Lattice parametera / nm Particle sizeb / nm SBET / (m2·g-1) VPore / (m3·g-1)
    2θ / (°) d / nm
    CeO2 28.510 0.312 8 0.541 8 8.6 133 0.131
    Ce0.95Mg0.05O2 28.709 0.311 3 0.540 5 6.7 129 0.142
    Ce0.90Mg0.10O2 28.730 0.310 0 0.540 1 5.8 136 0.188
    Ce0.85Mg0.15O2 28.770 0.310 1 0.539 8 6.1 117 0.129
    Ce0.80Mg0.20O2 28.757 0.310 2 0.540 6 6.3 126 0.168
    a Calculated using Vegard′s law; b Estimated by TEM.
    下载: 导出CSV

    Table 2.  Relative ratio of Ce3+ species and oxygen vacancies on the surface

    Catalyst Surface content of Ce3+ / % Surface content of OV+OC / % Total amount of adsorbed CO2a / (mmolCO2·gcat-1)
    CeO2 1.34 24.97 0.787
    Ce0.95Mg0.05O2 11.57 30.85 0.794
    Ce0.90Mg0.10O2 19.42 31.98 0.845
    Ce0.85Mg0.15O2 15.18 25.06 0.808
    Ce0.80Mg0.20O2 15.40 23.19 0.771
    aDetermined using the CO2 temperature-programmed desorption (CO2-TPD) analysis in Fig.S3.
    下载: 导出CSV
  • 加载中
计量
  • PDF下载量:  10
  • 文章访问数:  592
  • HTML全文浏览量:  143
文章相关
  • 发布日期:  2022-07-10
  • 收稿日期:  2022-03-17
  • 修回日期:  2022-05-18
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章