Next Article in Journal
Molecular Catalysis for Precise Olefin Polymerization and ROP 2015
Previous Article in Journal
Hydrogenolysis of Glycerol to 1,2-Propanediol and Ethylene Glycol over Ru-Co/ZrO2 Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of Diethyl Carbonate from Carbon Dioxide, Propylene Oxide and Ethanol over KNO3-CeO2 and KBr-KNO3-CeO2 Catalysts

Key Laboratory of Organic Solar Cells and Photochemical Conversion, School of Chemistry & Chemical Engineering, Tianjin University of Technology, Tianjin 300384, China
*
Author to whom correspondence should be addressed.
Catalysts 2016, 6(4), 52; https://doi.org/10.3390/catal6040052
Submission received: 22 February 2016 / Revised: 18 March 2016 / Accepted: 23 March 2016 / Published: 29 March 2016

Abstract

:
One-pot syntheses of diethyl carbonate (DEC) from CO2, propylene oxide and ethanol were carried out using different solid catalysts. The supercritical CO2 extraction method was used to separate the liquid products and reactants from the catalysts after reaction. The KNO3-CeO2 and KBr-KNO3-CeO2 were found to be active for the reaction after calcinations. The catalyst was also reusable. The thermodynamic properties of the reaction were also evaluated. The effects of various conditions, such as reaction time, amount of catalysts, molar ratio of the reactants, the composition and calcination temperature of the catalysts on the conversion and yields, were investigated, and the yield of DEC was about 13.0% with a selectivity of 38.5% over KBr-KNO3-CeO2. The yield of DEC was improved about 10-fold by using KBr-KNO3-CeO2 catalyst compared to CeO2.

Graphical Abstract

1. Introduction

The synthesis of chemicals using CO2 as a raw material is characterized at present by increasing industrial and academic efforts to use this carbon renewable [1]. The production of carbonates [2,3], carbamates [4], methanol [5], formic acid and its derivatives could be synthesized from CO2 [6,7]. Diethyl carbonate (DEC) is one of the most important green chemicals among carbonate esters. It is an excellent solvent and an intermediate for various pharmaceuticals, such as antibiotics and phenobarbital [8]. DEC has also been proposed as a replacement for MTBE as an attractive oxygen-containing fuel additive for its high oxygen content (40.6 wt. %) compared to MTBE (18.2 wt. %) [9,10].
Since the conventional methodologies for the DEC synthesis, including ethanol phosgenation [11], ethanol oxidative carbonylation [12] and the reaction of ethanol with urea [13], have many problems, such as the toxicity of phosgene, corrosion and low production rates [14], the novel technology for DEC synthesis starting from CO2 and ethanol is a promising route. However, the reaction hardly occurs spontaneously, even under harsh conditions, due to the thermodynamic limitations (yield of less than 0.5%) [8]. To address this issue, chemical dehydration reagent was usually involved to shift the reaction forward to the carbonate side. In the similar reaction of the synthesis dimethyl carbonate (DMC), acetals [15] and orthoesters [16] were used as the organic dehydrants, respectively, and both the DMC yields could be effectively improved above 20-fold. However, the high cost of acetals and orthoesters makes them difficult for industrial production. Acetonitrile [17] and amines [18] were also reported as being used as the dehydrants for the DEC synthesis, but their co-products were complex. Recently, butylene oxide was also used as the dehydrant for the direct synthesis of the DEC over CeO2 catalyst [8,19]. According to the results, the yield of DEC had a nine-fold enhancement compared to that over CeO2 without dehydrant, but it was still not high enough (only 1.5%) in this system and needed to be improved. Besides, CeO2 with 2-cyanopyridine was also used as the carboxylation/hydration cascade catalyst by Tomishige group [20] for the propylene carbonate synthesis from CO2 and 1,2-propanediol, and the yield was much higher (>99%), which might be a landmark in carbonate synthesis using CeO2 catalysis.
In addition, the one-pot synthesis of DEC from carbon dioxide, ethylene oxide (EO) and ethanol on the KI and sodium ethoxide binary homogeneous catalyst was also researched by Wang et al. [21], and the yield of DEC was improved. However, KI and sodium ethoxide are dissolved in ethanol and cannot be separated easily.
On the other hand, the synthesis of cyclic carbonate from epoxide and CO2 was well established in industrial manufacturing. Furthermore, the transesterification of cyclic carbonate with ethanol to produce DEC was also proven to be feasible [22] However, from views of energy consumption, productivity and investment, the one-pot reaction directly from CO2 was undoubtedly superior to the two-step separate reaction. Thus, the development of a more effective one-pot reaction to improve the productivity of DEC directly from CO2 is highly desired.

2. Results and Discussion

The one-pot reaction in DEC synthesis might be composed of two steps, the cycloaddition reaction and subsequent transesterification reaction. The mechanisms of the reaction have been studied and proven by many researchers [21,23]. As analyzed by the GC-MS method in this work, the main products in the one-pot reaction from CO2, ethanol and PO were 1,2-propanediol (PG), DEC and propylene carbonate (PC) with the side-product 1-ethoxy-2-propanol (EP). The EP might be formed from propylene oxide by ethanolysis [23,24] in the basic catalytic environment. The possible equations of these reactions are presented as follows (Scheme 1).

2.1. The One-Pot Synthesis of DEC over KNO3-CeO2 Catalyst

As reviewed in the literature [8], the heterogeneous catalyst CeO2 has catalytic activity for the one-pot reaction in DEC synthesis. In order to improve the catalytic activity in DEC synthesis, the strong base of KOH and several typical alkali and alkaline-earth metal salts, which might be necessary to meet the requirement for catalyzing the transesterification reaction combined with CeO2, were researched. Several metal oxides, such as γ-Al2O3, ZrO2, SiO2 and La2O3, were also tested. In the typical reaction, the molar ratio of CO2, ethanol and PO was fixed as 0.29:0.17:0.14; the reaction temperature was 150 °C; the initial pressure was 5 MPa; and during the reaction, the pressure could reach 9 MPa, which is higher than the critical pressure of CO2. The results are summarized in Table 1, Runs 1–12.
As seen in Table 1, when KNO3-CeO2 was used as the catalyst, the yield of DEC was effectively improved about 10-fold, compared to that using CeO2 catalyst. The KNO3-CeO2 had better catalytic activity than KOH and other alkali and alkaline-earth metal salts loading on CeO2, such as KOH-CeO2, K2CO3-CeO2, NaNO3-CeO2, Ba(NO3)2-CeO2, Mg(NO3)2-CeO2 and Ca(NO3)2-CeO2. Then, KNO3 with other metal oxides, including γ-Al2O3, ZrO2, La2O3 and SiO2, was evaluated. The results indicate that the oxides with acid-base properties, especially ZrO2, γ-Al2O3 and CeO2 combined with KNO3, are more active for DEC synthesis. The basicity of the catalysts KNO3-γ-Al2O3, KNO3-CeO2 and CeO2 were analyzed by the CO2-TPD method (Figure 1). The desorption peaks at about 280 °C–340 °C could be observed for KNO3-γ-Al2O3 and KNO3-CeO2, which had better catalytic activities. It is indicated that the addition of a small amount of a moderate base is more effective for enhancing the activity of the catalyst. Considering the KNO3-CeO2 catalyst obtaining the better yield and selectivity for DEC, it was finally selected for the following reactions. The XRD spectrums of KNO3-CeO2 and CeO2 are shown in Figure 2. By comparison, the characteristic peaks of KNO3 cannot be found, and the diffraction peaks of KNO3-CeO2 are stronger than CeO2. It is indicated that the KNO3 might be well dispersed and caused no considerable distortion in the structure of CeO2 [25].
The preparation conditions for KNO3-CeO2 including the load of KNO3 and the calcination temperature were optimized. The results are shown in Figure 3a,b. It is indicated that the KNO3-CeO2 with n(Ce)/n(K) = 1:0.4 has better catalytic activity (Figure 3a). And the DEC yield reaches higher level (Figure 3b) at the calcination temperature of 500 °C. According to the TG-DTG analysis of KNO3 (Figure 4), the decomposition temperature of KNO3 is 520 °C. When the temperature is higher than 520 °C, the KNO3 will be decomposed to K2O, and the catalytic activity will decrease.
Then, the reaction conditions, including the amount of catalyst, reaction time and volume ratio of ethanol and PO, were studied. The results are shown in Figure 3c–e. As seen in Figure 3c, the DEC yield first increases and then decreases with the increase of catalyst amounts. The DEC yield reaches the peak value when the amount of KNO3-CeO2 is 0.3 g (Figure 4c). Figure 4d shows the dependence of DEC yield on reaction time. The reaction reaches equilibrium in 2 h. Figure 3e shows that the DEC yield reaches the higher level, when both of the volumes of ethanol and PO are 10 mL (0.17 mol and 0.14 mol, respectively) and the CO2 amount is fixed as 0.25 mol.

2.2. The One-Pot Synthesis of DEC over KBr-KNO3-CeO2 Catalyst

As potassium halides (KI, KBr and KCl) were proven to be conducive to the cycloaddition reaction [15]; the KI, KBr and KCl were added in the KNO3-CeO2 catalyst for the one-pot synthesis of DEC. The addition methods for potassium halides, including solid mixed and impregnation methods, were compared.
The results are shown in Table 1, Runs 13–20. As seen in Table 1, when KI-CeO2 is used as the catalyst, the ethanol conversion increases, but the yield of DEC is not improved compared to CeO2 alone. In addition, when KBr or KCl is added to KNO3-CeO2, the ethanol conversion decreases, but the selectivity of DEC is effectively improved. Additionally, the catalysts prepared by the solid mixed method give a higher DEC yield and selectivity. However, when KI-KNO3-CeO2 is used as the catalyst, the DEC yield seriously decreases. This might be because that KI, with a stronger reduction property, is oxidized by KNO3, which causes deactivation.
The molar ratio of KBr and KNO3 was evaluated with fixed n(CeO2)/n(KNO3) = 1:0.4. The results are shown in Figure 5. The XRD patterns of KBr-KNO3-CeO2 are presented in Figure 4, and the characteristic peaks of KBr are labeled. The yield and the selectivity of DEC are improved with the addition of KBr into the catalyst. When the molar ration is n(KBr)/n(KNO3) = 6:4, that is n(CeO2)/n(KNO3)/n(KBr) = 1:0.4:0.6, the yield of DEC reaches 13.0% with a selectivity of 38.5% on the ethanol basis. Both the yield and selectivity of DEC are much higher than reported in Leino’s research [8] by using butylene oxide as the dehydration agent and CeO2 as the catalyst. In their results, the highest obtained yield of DEC was 1.5%, and selectivity to DEC was 10% on the ethanol basis.
The formation kinetics of DEC were also studied, and the dependence of DEC yield on reaction time at 150 °C in 180 min is shown in Figure 6. It is also indicated that the reaction can reach equilibrium in 100 min. The reaction time was selected as 120 min.

2.3. Recycling Experiments

The recycling experiment results of the KBr-KNO3-CeO2 catalyst are listed in Figure 7. After three times reuse, the catalyst still keeps good catalytic activity, and the DEC yield is above 90%, as in the primary reaction. The XRD patterns of fresh KBr-KNO3-CeO2 and the catalyst after using three times were also compared. The calcination temperature KBr-KNO3-CeO2 was also set as 500 °C. Additionally, the KBr was not decomposed according to the TG-DTG analysis (Figure 4). As seen in Figure 7, the characteristic peaks do not change after three runs. This indicates that the active species of KBr and KNO3 do not leach from the catalyst. The reason might be that CO2 was the main reactant in this reaction, and the reactants and products were extracted by Sc-CO2 after the reaction, while the catalyst was not soluble in Sc-CO2 and left in the reactor, which avoided the loss of catalyst. This is one of the advantages of separating the products by Sc-CO2.

2.4. Thermodynamic Evaluation of the One-Pot Synthesis of DEC

In order to perform the thermodynamics evaluations, which are important in seeking novel synthesis ideas, the thermodynamic data of various substances, such as ethanol, CO2, PO, DEC and PG, involved in the reaction are tabulated in Table 2.
The enthalpy and the entropy of the reaction at 298 K estimated from the Δ f H 298 k θ and S 298 k θ values amounted to Δ r H 298 k θ = 101.4   KJ · mol 1 and Δ r S 298 k θ = 357.22   J · mol 1 · K 1 . The Gibbs energy at 298 K and 100 KPa could be calculated by Equation (1) and has a value Δ r G 298 k θ = 5.10   KJ · mol 1 .
Δ r G 298 k θ = Δ r H 298 k θ + T Δ r S 298 k θ
Based on the obtained values, it can be concluded that the reaction is exothermic ( Δ r H 298 k θ = 101.4 KJ · mol 1 < 0 ) and does not occur spontaneously at room temperature Δ r G 298 k θ = 5.10   KJ · mol 1 > 0 .
The relative pattern of the reaction heat with the temperature is expressed by Kirchhoff’s law (Equation (2)), whereas the Gibbs energy of the reaction, at different temperatures, can be given by the Gibbs–Helmholtz equation (Equation (3)) [8]. Gibbs energy values of this reaction at different temperatures were calculated. The results are shown in Figure 8. The value of Δ r G T θ increases with the reaction temperature, and the increase in the temperature is disadvantageous to the formation of DEC.
Δ r H T θ = Δ r H 298 k θ + Δ C p ( T 298 )
d ( Δ r G T θ T ) = ( Δ r H T θ T 2 ) d T
When the temperature is fixed at 423 K, the Gibbs energy of the reaction is a function of pressure and can be written as follows [8]:
Δ r G p = Δ r G θ R T ln ( p p θ )
The equilibrium constant K was determined from the Gibbs energy according to Equation (5):
Δ r G T = R T ln K
K = [ DEC ] [ PG ] [ C 2 H 5 OH ] 2 [ PO ] [ CO 2 ]
In this work, the reaction temperature and pressure are 423 K and 9 MPa; the Δ r G p calculated by Equation (4) is 36.07 kJ·mol−1 > 0; and the equilibrium constant K calculated by Equation (5) is 3.5 × 10−5. When the amounts of CO2, ethanol and PO are 0.29, 0.17 and 0.14 mol, respectively, and assuming that the reaction is in the gas phase, the equilibrium yield of DEC at 423 K and 9 MPa calculated by Equation (6) is about 14.2% based on ethanol. It is indicated that the yield of DEC in the experiment (13.0%) is close to the equilibrium yield calculated.

3. Experimental Section

3.1. Material

Ethanol (99.8%, analytical grade) was dehydrated by adding spherical 3A molecular sieves before the reactions. Propylene oxide (PO), n-propanol, Ce(NO3)3·6H2O, KNO3, KOH, NaNO3, Mg(NO3)2·6H2O, Ba(NO3)2, Ca(NO3)2·4H2O, KBr, KCl, KI, ammonium hydroxide (NH3·H2O, 28%) and spherical 3A molecular sieves were purchased from Jiangtian Chemical Reagent Co., Ltd., Tianjin, China. All of the chemicals were of analytical grade and used without further purification. CO2 (99.99%) was purchased from Lianbo Gas Co., Ltd., Tianjin, China.

3.2. Catalyst Preparation and Characterization

CeO2 was prepared by the precipitation method. The 6.5 g of cerium (III) nitrate hexahydrate (Ce(NO3)3·6H2O) were firstly dissolved in 50 mL water. Then, the precipitating solution of ammonium hydroxide of 50% (v/v) was slowly added into the well-stirred cerium nitrate aqueous solution. The pH of the solution was controlled at 10 throughout the synthesis process. The resulting precipitate was filtrated, washed with deionized water and then dried overnight at 100 °C and calcined at 500 °C for 2 h in air.
KNO3-CeO2 was prepared by the incipient impregnation method. The slurry (n(Ce)/n(K) = 1:0.2, 1:0.4, 1:0.6, 1:0.8 or 1:1) was kept in a static state for 24 h and dried at 110 °C for 5 h. The dried samples were calcined at 500 °C for 3 h. Similar methods were used for the preparation of NaNO3-CeO2, KOH-CeO2, Ca(NO3)2-CeO2, Mg(NO3)2-CeO2 and Ba(NO3)2-CeO2.
KBr-KNO3-CeO2 was prepared by both the impregnation method and solid mixed method. The impregnation method was performed as in the KNO3-CeO2 preparation. For the solid mixed method, the dried solids of KBr and KNO3-CeO2 were mixed evenly and then calcined at 500 °C for 3 h.
The specific surface area was measured by using nitrogen adsorption (Sorptometer 1900, Carlo Erba Instruments, Milan, Italy). The sample was out gassed at 150 °C for 3 h prior to the measurement of the surface area. For calculating the surface area, the BET equation was used.
The structural properties of the catalysts were investigated by X-ray diffraction (XRD) (X’Pert Pro MPD, PANalytical, Lelyweg, The Netherlands) using CuKα (40 kV, 50 mA) radiation with 2θ ranging from 0° to 80° at a scanning speed 0.04°/3 s.
Thermogravimetric (TG) analysis was carried out on a thermogravimetric analyzer (TG209, Netzsch Co., Wittelsbacherstr, Germany) under a N2 stream of 40 mL min−1. The temperature was raised at a heating rate of 10 °C min−1.
Temperature-programmed desorption (TPD) of CO2 was carried out using the Micromeritics Instrument (Auto Chem 2910, Micromeritics Instrument Co., Norcross, GA, USA). Zero-point-one gram of catalyst was placed in a quartz-made U-shaped tube. The catalyst was firstly treated in a flow of helium (50 mL·min−1) at 200 °C for 2 h to remove any physisorbed organic molecules. The sample was then saturated by adsorption of CO2 (50 mL·min−1) at 60 °C for 30 min. Physisorbed CO2 was flushed at 150 °C for 1 h. After the sample was cooled, the furnace temperature was increased from 60 °C to 800 °C at a heating rate of 5 °C·min−1 under a flow of helium (30 mL·min−1). The desorbed CO2 was detected by using a TCD (thermal conductivity detector).

3.3. Catalytic Test

All experiments were carried out in a laboratory-scale stainless steel autoclave (Weihai Chemical Machinery Co., Ltd., Weihai, China) with an inner volume of 100 mL equipped with a stirrer and an electric heater. In a standard procedure, 0.3–0.5 g of catalyst, 170 mmol (10 mL) of ethanol and 140 mmol (10 mL) propylene oxide were introduced into the autoclave. The reactor was purged with CO2 at room temperature. Then, the reaction system was pressurized to the predetermined initial pressure with CO2 by a pump (Model SFC-24, SSI/LabAlliance, State College, PA, USA). The amount of CO2 was calculated from measuring the weight difference of the CO2 cylinder before and after charging the CO2. After the reaction, the reactor was cooled to about 50 °C. The CO2 was released slowly through a receiving flask with methanol as the absorbent. Then, the liquid chemicals in the reactor were extracted in situ by supercritical CO2 at 12 MPa and 50 °C. All of the reactants and products were collected using methanol in a flask. The catalyst was left in the reactor. About 240 g CO2 were required to extract all liquid chemicals. For the recycling experiments, the catalyst was used directly after the extraction process by repeating the procedures above.
Yield   of   product   Y i ( % ) = mole of product i 0.5 × mole of ethanol charged  × 100 %
Selectivity  S i ( % ) = mole of product i 0.5 × mole of ethanol convered × 100 %

3.4. Analysis Method

The products were analyzed by gas chromatography (GC) with an FID detector (6890N, Agilent Co., Santa Clara, CA, USA) and GC–MS (HP 6890 gas chromatograph coupled with mass detector HP 5973N, Agilent Co., Santa Clara, CA, USA) using an HP-5MS (30 m × 0.25 mm × 0.25 μm, Agilent Co., Santa Clara, CA, USA) fused silica capillary column. n-Propanol was used as the internal standard. The chromatography conditions were the same as our previous studies [14].

4. Conclusions

The direct synthesis of DEC from carbon dioxide, ethanol and PO was developed over KNO3-CeO2 and KBr-KNO3-CeO2 catalyst. Supercritical CO2 was used to extract the products and reactants in site, and the catalysts could be reused. In comparison with the direct synthesis of DEC from CO2 and ethanol, the involvement of PO could improve the formation of DEC, simultaneously with the important chemicals of glycol coproduced. The yield of DEC was improved about 10-fold compared to that using single CeO2 as the catalyst. The yield of DEC was about 13.0% with a selectivity of 38.5% under the optimized conditions with KBr-KNO3-CeO2 as the catalyst, which was close to the equilibrium yield calculated in this work.

Acknowledgments

This study is supported by the Youth Fund of the National Natural Science Foundation of China (No. 21106107 and No. 21206077).

Author Contributions

Zhen Zhu synthesized the catalysts and performed most of the catalytic reactions, while Yongyue Sun performed some of the charaterisation. The initial draft of the paper was written by Yanlou Wang; while the final copy and corrections made at various stages of this paper was done by the corresponding author, Dongdong Jia.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Subodh, K.; Suman, L.J. Polyethylene Glycol Enfolded KBr Assisted Base Catalyzed Synthesis of Dimethyl Carbonate from Methanol and Carbon Dioxide. Ind. Eng. Chem. Res. 2014, 53, 15798–15801. [Google Scholar]
  2. Wan Nor Roslam, W.I.; Zatil Amali, C.R.; Mohamed Wahab, M.H.; Mohd, A.Y. The formation of a series of carbonates from carbon dioxide: Capturing and utilization. Renew. Sustain. Energy Rev. 2015, 47, 93–106. [Google Scholar]
  3. Dibenedetto, A.; Angelini, A. Synthesis of Organic Carbonates. Adv. Inorg. Chem. 2014, 66, 25–81. [Google Scholar]
  4. Li, J.; Qi, X.; Wang, L.; He, Y.; Deng, Y. New attempt for CO2 utilization: One-pot catalytic syntheses of methyl, ethyl and n-butyl carbamates. Catal. Commun. 2014, 12, 1224–1227. [Google Scholar] [CrossRef]
  5. Sanguineti, P.B.; Baltanas, M.A.; Bonivardi, A.L. Copper-gallia interaction in Cu-Ga2O3-ZrO2 catalysts for methanol production from carbon oxide(s) hydrogenation. Appl. Catal. A 2014, 504, 476–481. [Google Scholar] [CrossRef]
  6. Ma, J.; Sun, N.; Zhang, X.; Zhao, N.; Xiao, F.; Wei, W.; Sun, Y. A short review of catalysis for CO2 conversion. Catal. Today 2009, 148, 221–231. [Google Scholar] [CrossRef]
  7. Dibenedetto, A.; Angelini, A.; Stufano, P. Use of carbon dioxide as feedstock for chemicals and fuels: Homogeneous and heterogeneous catalysis. J. Chem. Technol. Biotechnol. 2014, 89, 334–353. [Google Scholar] [CrossRef]
  8. Leino, E.; Maki-Arvela, P.; Eranen, K.; Tenho, M.; Murzin, D.Y.; Salmi, T.; Mikkola, J.P. Enhanced yields of diethyl carbonate via one-pot synthesis from ethanol, carbon dioxide and butylene oxide over cerium (IV) oxide. Chem. Eng. J. 2011, 176–177, 124–133. [Google Scholar] [CrossRef]
  9. Dunn, B.C.; Guenneau, C.; Hilton, S.A.; Pahnke, J.; Eyring, E.M.; Dworzanski, J.; Meuzelaar, H.L.C.; Hu, J.Z.; Solum, M.S.; Pugmire, P.J. Production of diethyl carbonate from ethanol and carbon monoxide over a heterogeneous catalyst. Energy Fuel 2002, 16, 177–181. [Google Scholar] [CrossRef]
  10. Ma, X.; Shi, H.; Zhang, Z.; Wang, B.; Wang, S.; Xu, G. Catalytic synthesis of diethyl carbonate from carbon monoxide and ethyl nitrite on supported Pd catalysts. Prepr. Pap. Am. Chem. Soc. Div. Fuel Chem. 2003, 48, 906–907. [Google Scholar]
  11. Ma, X.B.; Fan, M.M.; Zhang, P.B. Study on the catalytic synthesis of diethyl carbonate from CO and ethyl nitrite over supported Pd catalysts. Catal. Commun. 2004, 5, 765–770. [Google Scholar] [CrossRef]
  12. Muskat, I.E.; Strain, F. Preparation of carbonic acid esters. US Patent 2,379,250, June 1941. [Google Scholar]
  13. Wang, D.; Yang, B.; Zhai, X.; Zhou, L. Synthesis of diethyl carbonate by catalytic alcoholisis of urea. Fuel Process. Technol. 2007, 88, 807–812. [Google Scholar] [CrossRef]
  14. Zhang, X.; Jia, D.D.; Zhang, J.; Sun, Y.Y. Direct Synthesis of Diethyl Carbonate from CO2 and Ethanol Catalyzed by ZrO2/Molecular Sieve. Catal. Lett. 2014, 144, 2144–2150. [Google Scholar] [CrossRef]
  15. Choi, J.C.; Kohno, K.; Ohshima, Y.; Yasuda, H.; Sakakura, T. Tin- or titanium-catalyzed dimethyl carbonate synthesis from carbon dioxide and methanol: Large promotion by a small amount of triflate salts. Catal. Commun. 2008, 9, 1630–1633. [Google Scholar] [CrossRef]
  16. Sakakura, T.; Saito, Y.; Okano, M.; Choi, J.C.; Sako, T. Selective Conversion of Carbon Dioxide to Dimethyl Carbonate by Molecular Catalysis. J. Org. Chem. 1998, 63, 7095–7096. [Google Scholar] [CrossRef] [PubMed]
  17. Honda, M.; Kuno, S.; Begum, N.; Fujimoto, K.; Suzuki, K.; Nakagawa, Y.; Tomishige, K. Catalytic synthesis of dialkyl carbonate from low pressure CO2 and alcohols combined with acetonitrile hydration catalyzed by CeO2. Appl. Catal. A 2010, 384, 165–170. [Google Scholar] [CrossRef]
  18. Honda, M.; Sonehara, S.; Yasuda, H.; Nakagawaa, Y.; Tomishige, K. Heterogeneous CeO2 catalyst for the one-pot synthesis of organic carbamates from amines, CO2 and alcohols. Green Chem. 2011, 13, 3406–3413. [Google Scholar] [CrossRef]
  19. Leino, E.; Mäki-Arvela, P.; Eta, V.; Kumar, N.; Demoisson, F.; Samikannu, A.; Leino, A.R.; Shchukarev, A.; Murzin, D.Y.; Mikkola, J.P. The influence of various synthesis methods on the catalytic activity of cerium oxide in one-pot synthesis of diethyl carbonate starting from CO2, ethanol and butylene oxide. Catal. Today 2013, 210, 47–54. [Google Scholar] [CrossRef]
  20. Honda, M.; Tamura, M.; Nakao, K.; Suzuki, K.; Nakagawa, Y.; Tomishige, K. Direct Cyclic Carbonate Synthesis from CO2 and Diol over Carboxylation/Hydration Cascade Catalyst of CeO2 with 2-Cyanopyridine. ACS Catal. 2014, 4, 1893–1896. [Google Scholar] [CrossRef]
  21. Wang, L.; Li, H.; Xin, S.; He, P.; Cao, Y.; Li, F.; Hou, X. Highly efficient synthesis of diethyl carbonate via one-pot reactionfrom carbon dioxide, epoxides and ethanol over KI-based binarycatalyst system. Appl. Catal. A 2014, 471, 19–27. [Google Scholar] [CrossRef]
  22. Qiu, P.; Wang, L.; Jiang, X.; Yang, B. Synthesis of Diethyl Carbonate by the Combined Process of Transesterification with Distillation. Energy Fuel 2012, 26, 1254–1258. [Google Scholar] [CrossRef]
  23. Chang, Y.; Jiang, T.; Han, B.; Liu, Z.; Wu, W.; Gao, L.; Li, J.; Gao, H.; Zhao, G.; Huang, J. One-pot synthesis of dimethyl carbonate and glycols from supercritical CO2, ethylene oxide or propylene oxide, and methanol. Appl. Catal. A 2004, 263, 179–186. [Google Scholar] [CrossRef]
  24. Eta, V.; Maki-Arvela, P.; Leino, A.R.; Kordas, K.; Salmi, T.; Murzin, D.Y.; Mikkola, J.P. Synthesis of Dimethyl Carbonate from Methanol and Carbon Dioxide: Circumventing Thermodynamic Limitations. Ind. Eng. Chem. Res. 2010, 49, 9609–9617. [Google Scholar] [CrossRef]
  25. Wang, Y.; Huang, W.Y.; Wu, Z.; Chun, Y.; Zhu, J.H. Superbase derived from zirconia-supported potassium nitrate. Mater. Lett. 2000, 46, 198–204. [Google Scholar] [CrossRef]
  26. Ma, P.S. Chemical Engineering Thermodynamics; Chemical Industry Press: Beijing, China, 2005. [Google Scholar]
  27. Ma, P.S. Chemical Engineering Data; China Petrochemical Press: Beijing, China, 2003. [Google Scholar]
Scheme 1. The reaction schemes.
Scheme 1. The reaction schemes.
Catalysts 06 00052 g009
Figure 1. CO2-temperature-programmed desorption (TPD) analysis for the catalysts.
Figure 1. CO2-temperature-programmed desorption (TPD) analysis for the catalysts.
Catalysts 06 00052 g001
Figure 2. XRD patterns of CeO2, KNO3-CeO2 and KBr-KNO3-CeO2 (KBr, CeO2).
Figure 2. XRD patterns of CeO2, KNO3-CeO2 and KBr-KNO3-CeO2 (KBr, CeO2).
Catalysts 06 00052 g002
Figure 3. Optimization of the conditions for the catalyst preparation and the DEC synthesis. The ethanol volume was fixed as 10 mL; the reaction temperature was 150 °C; and the initial pressure was 5 MPa.
Figure 3. Optimization of the conditions for the catalyst preparation and the DEC synthesis. The ethanol volume was fixed as 10 mL; the reaction temperature was 150 °C; and the initial pressure was 5 MPa.
Catalysts 06 00052 g003
Figure 4. TG-DTG analysis for KNO3 and KBr.
Figure 4. TG-DTG analysis for KNO3 and KBr.
Catalysts 06 00052 g004
Figure 5. Effects of the molar ratio of KBr and KNO3 on the products (both of the volumes of ethanol and PO were 10 mL; the reaction temperature was 150 °C; the initial pressure was 5 MPa). PC, propylene carbonate.
Figure 5. Effects of the molar ratio of KBr and KNO3 on the products (both of the volumes of ethanol and PO were 10 mL; the reaction temperature was 150 °C; the initial pressure was 5 MPa). PC, propylene carbonate.
Catalysts 06 00052 g005
Figure 6. The relationship between the reaction time and yield of DEC at 150 °C for 180 min. (both of the volumes of ethanol and PO were 10 mL; the initial pressure was 5 MPa).
Figure 6. The relationship between the reaction time and yield of DEC at 150 °C for 180 min. (both of the volumes of ethanol and PO were 10 mL; the initial pressure was 5 MPa).
Catalysts 06 00052 g006
Figure 7. Recycling experiment results of the KBr-KNO3-CeO2 (both of the volumes of ethanol and PO were 10 mL; the reaction temperature was 150 °C; the initial pressure was 5 MPa).
Figure 7. Recycling experiment results of the KBr-KNO3-CeO2 (both of the volumes of ethanol and PO were 10 mL; the reaction temperature was 150 °C; the initial pressure was 5 MPa).
Catalysts 06 00052 g007
Figure 8. Dependence of temperature on the Gibbs energy of DEC synthesis from ethanol, CO2 and PO.
Figure 8. Dependence of temperature on the Gibbs energy of DEC synthesis from ethanol, CO2 and PO.
Catalysts 06 00052 g008
Table 1. Effects of different catalysts on the one-pot reaction. DEC, diethyl carbonate; PG, 1,2-propanediol; EP, 1-ethoxy-2-propanol.
Table 1. Effects of different catalysts on the one-pot reaction. DEC, diethyl carbonate; PG, 1,2-propanediol; EP, 1-ethoxy-2-propanol.
RunCatalystSpecific surface area/m2/gEthanol conversion/%DEC yield/%Selectivity/%
DECPGEP
1CeO2-1.51.384.0--
2NaNO3-CeO2-12.63.628.324.119.3
3Ba(NO3)2-CeO2-8.10.912.610.216.6
4Mg(NO3)2-CeO2-15.92.213.015.917.4
5Ca(NO3)2-CeO2-6.61.322.316.719.8
6KOH-CeO2-13.43.628.020.133.6
7K2CO3-CeO2-9.11.819.017.625.1
8KNO3-CeO230.534.011.233.323.723.5
9KNO3-γ-Al2O310246.811.524.523.116.5
10KNO3-ZrO222.923.76.32724.422.9
11KNO3-SiO278.112.33.125.326.118.8
12KNO3-La2O331.80.60.469.120.710.9
13KI-CeO2 A-18.00.96.36.410.3
14KI-CeO2 B-9.01.314.417.412.8
15KI-KNO3-CeO2 A-17.10.95.84.97.7
16KI-KNO3-CeO2 B-9.60.47.57.97.2
17KCl-KNO3-CeO2 A-14.42.717.414.45.7
18KCl-KNO3-CeO2 B-39.910.325.922.35.9
19KBr-KNO3-CeO2 A-23.19.440.830.57.6
20KBr-KNO3-CeO2 B-33.913.038.527.77.5
Both of the volumes of ethanol and PO were 10 mL; the reaction temperature was 150 °C; the initial pressure was 5 MPa. A Catalysts prepared by the impregnation method; B catalysts prepared by the solid mixed method.
Table 2. Thermodynamic data of various substances in the reaction [26].
Table 2. Thermodynamic data of various substances in the reaction [26].
Substance Δ f H 298 k θ /kJ·mol−1 S 298 k θ /J·mol−1·K−1 C p   298 k i d /J·mol−1·K−1
C2H5OH−235.10282.70112.6
CO2−393.509213.7437.13
DEC−637.9412.94 a211
H2O −241.82188.8233.58
PO−94.68281.1572.55
PG−421.29288 a189.9
a Calculated by the Constantinous-Gani ( CG) group contribution method [27].

Share and Cite

MDPI and ACS Style

Wang, Y.; Jia, D.; Zhu, Z.; Sun, Y. Synthesis of Diethyl Carbonate from Carbon Dioxide, Propylene Oxide and Ethanol over KNO3-CeO2 and KBr-KNO3-CeO2 Catalysts. Catalysts 2016, 6, 52. https://doi.org/10.3390/catal6040052

AMA Style

Wang Y, Jia D, Zhu Z, Sun Y. Synthesis of Diethyl Carbonate from Carbon Dioxide, Propylene Oxide and Ethanol over KNO3-CeO2 and KBr-KNO3-CeO2 Catalysts. Catalysts. 2016; 6(4):52. https://doi.org/10.3390/catal6040052

Chicago/Turabian Style

Wang, Yanlou, Dongdong Jia, Zhen Zhu, and Yongyue Sun. 2016. "Synthesis of Diethyl Carbonate from Carbon Dioxide, Propylene Oxide and Ethanol over KNO3-CeO2 and KBr-KNO3-CeO2 Catalysts" Catalysts 6, no. 4: 52. https://doi.org/10.3390/catal6040052

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