Next Article in Journal
Dynamic Rheological Studies of Poly(p-phenyleneterephthalamide) and Carbon Nanotube Blends in Sulfuric Acid
Previous Article in Journal
Bioavailability of the Polyphenols: Status and Controversies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Gas-Phase Synthesis of Dimethyl Carbonate from Methanol and Carbon Dioxide Over Co1.5PW12O40 Keggin-Type Heteropolyanion

by
Ahmed Aouissi
*,
Zeid Abdullah Al-Othman
and
Amro Al-Amro
Department of Chemistry, King Saud University, P.O. Box 2455, Riyadh-11451, Saudi Arabia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2010, 11(4), 1343-1351; https://doi.org/10.3390/ijms11041343
Submission received: 22 January 2010 / Revised: 9 March 2010 / Accepted: 29 March 2010 / Published: 31 March 2010
(This article belongs to the Section Physical Chemistry, Theoretical and Computational Chemistry)

Abstract

:
The reactivity of Co1.5PW12O40 in the direct synthesis of dimethyl carbonate (DMC) from CO2 and CH3OH was investigated. The synthesized catalyst has been characterized by means of FTIR, XRD, TG, and DTA and tested in gas phase under atmospheric pressure. The effects of the reaction temperature, time on stream, and methanol weight hourly space velocity (MWHSV) on the conversion and DMC selectivity were investigated. The highest conversion (7.6%) and highest DMC selectivity (86.5%) were obtained at the lowest temperature used (200 °C). Increasing the space velocity MWHSV increased the selectivity of DMC, but decreased the conversion. A gain of 18.4% of DMC selectivity was obtained when the MWHSV was increased from 0.65 h−1 to 3.2 h−1.

1. Introduction

Dimethyl carbonate (DMC) has drawn much attention in recent years as an environmentally friendly versatile intermediate. It has been used as a good solvent [1], an alkylation agent [2], and a substitute for highly toxic phosgene and dimethyl sulfate in many chemical processes [3,4]. In addition, it is expected to replace the gasoline oxygenate methyl tert. butyl ether (MTBE), because of its high oxygen content, low toxicity, and rapid biodegradability [1,5,6]. DMC has been produced by the reaction of methanol with phosgene in a concentrated sodium hydroxide solution [7]. However, owing to the high toxicity and the severe corrosivity of phosgene, this process has been abandoned gradually. Currently, DMC is produced mainly by oxidative carbonylation of methanol (non-phosgene route) [8]. The synthesis can be carried out in both liquid- and gas-phases. However, both routes use poisonous gas carbon monoxide and there is the possibility of an explosion. Recently, direct synthesis of DMC from CO2 and CH3OH has been reported as a most attractive route due to the low-cost of CO2 and the environmentally benign process [914]. However, DMC yield in this route is relatively low due to the fact that CO2 is thermodynamically stable and kinetically inert and due to the deactivation of catalysts induced by water formation in the reaction process [10,15]. Therefore, the development of efficient heterogeneous catalytic systems has attracted more attention. Bian et al. [16] studied the reaction over Cu–Ni/graphite nanocomposite catalyst in gaseous phase. They obtained 10.13% of CH3OH conversion and 89.04% of DMC selectivity at 105 °C. Wu et al. [17] studied the synthesis of DMC from gaseous methanol and CO2 over H3PO4 modified V2O5 catalyst with various molar ratios H3PO4/V2O5 (P/V). The best conversion (1.95%) and selectivity of DMC (92.12%) was obtained at 130 °C on the catalyst H3PO4/V2O5 (P/V = 0.20).
Here we report the direct synthesis of DMC from methanol and CO2 in gas phase over Co1.5PW12O40 as a Keggin-type heteropolyanion catalyst. The effect of the reaction temperature, MWHSV, and time on stream on DMC synthesis was investigated.

2. Results and Discussion

2.1. Characterization of Catalysts

The FT-IR spectrum of the Co1.5PW12O40 is shown in Figure 1. The IR spectrum has been assigned according to [18,19]. The main characteristic features of the Keggin structure are observed at 1080–1060 cm−1, 990–960 cm−1, 900–870 cm−1, and 810–760 cm−1, assigned to the stretching vibration νas (P-Oa), νas (M-Od), νas (M-Oc-M), and νas (M-Oc-M), respectively (M = W or Mo). The result of X-ray powder diffraction of the Co1.5PW12O40 salt is shown in Figure 2. In each of the four ranges of 2θ, 7°–10°, 16°–23°, 25°–30°, and 31°–38°, the compound shows a characteristic for well defined Keggin structure of heteropolyanions [2022]. So the presence of the primary Keggin structure in the synthesized phases was confirmed by FTIR and XRD.

Thermogravimetric (TG) and Dfferential Thermal Analysis (DTA)

Thermogravimetric analysis of Co1.5PW12O40 (Figure. 3) showed that the dehydration process starts at low temperature, about 50 °C, and finishes at 250 °C. The differential thermal analysis (DTA) indicates two endothermic peaks below 250 °C, and an exothermic peak at 575 °C. In agreement with the TG results, the endothermic peaks are assigned to the removal of physisorbed or crystallization water molecules (13–14 H2O), whereas the exothermic peak is due to the decomposition of Co1.5PW12O40 into the corresponding oxides (3/2CoO, 12 WO4 and 1/2P2O5). This result is in agreement with published results [23].

2.2. Catalytic Reaction

2.2.1. Effect of Reaction Temperature

The effect of the reaction temperature on the reaction performance was investigated at temperatures ranging from 200–300 °C. The results are illustrated in Figure 4 and 5. It can be seen from Figure 4 that CH3OH conversion and product yields decreased dramatically with increasing temperature. The decrease of both the conversion and product yields with increasing temperature might be due to the decreased CO2 adsorption on the catalyst at high temperatures [10]. Figure 5 shows that the selectivity of DMC decreased with increasing temperature. The decrease of DMC selectivity is probably due to the decomposition of DMC at higher temperatures [10,16, 2426]. It can be seen from these results that the highest conversion of 7.6% and highest selectivity of DMC of 86.5% was obtained at the lowest temperature in this range of temperatures (200 °C).

2.2.2. Effect of Time on Stream

Figure 6 shows CH3OH conversion and products yield as a function of time on stream. It can be seen that both CH3OH conversion and DMC yield increased rapidly during the first three hours, then after increased slowly. DMM and MF were observed as minor products with a rate formation almost constant. As for the product selectivity, it can be seen from the figure 7 that DMC selectivity increased slightly with increasing time on stream to the detriment of DMM and MF.

2.2.3. Effect of Space Velocity

The effect of methanol weight hourly space velocity (MWHSV) on the conversion of methanol and the selectivity of DMC is shown in Figure 8. It can be seen that the conversion dropped sharply to about 6.4%. In fact, when the SV changed from 0.6 h−1 to 1.6 h−1, the conversion decreased from 15.7% to 9.3%. Further increase of MWHSV did not influence the conversion of methanol considerably. The conversion dropped only to about 1.7%. In fact, it has been found that when the MWHSV changed from 1.6 h−1 to 3.2 h−1, the conversion decreased from 9.3% to 7.6%. As for the selectivity of DMC, the result showed that increasing the MWHSV from 0.65 h−1 to 3.2 h−1 increased the selectivity from 68.1% to 86.5% gaining in this way 18.4% of DMC selectivity.

3. Experimental Section

3.1. Catalyst Preparation

The heteropolymolybdate salt namely Co1.5PW12O40 was prepared from 12-tungstophosphoric acid H3PW12O40 as precipitate by adding slowly the required amount of CoCO3. After protons neutralization by CO32−, the cobalt salt, Co1.5PW12O40 was recovered from the solution by filtration

3.2. Physicochemical Techniques

The purity and the Keggin structure of the samples were characterized by means of IR and XRD. IR spectra were recorded with an infrared spectrometer GENESIS II-FTIR (400–4000 cm−1) as KBr pellets. The XRD powder patterns were recorded on a Rigaku diffractometer Ultima IV using CuKα radiation. Thermal analysis was carried out by means of differential thermal analysis (DTA) and thermogravimetric analysis (TGA) in air atmosphere with a 50 Shimadzu thermobalancesis

3.3. Catalytic Tests

The direct synthesis of DMC from methanol and CO2 was carried out at the temperature ranging from 200 °C to 300 °C in a flow-type fixed-bed stainless reactor loaded with 100 mg of catalyst under atmospheric pressure. To supply the reactant, a gas CO2 at a flow rate of 60 mL min−1 was passed through the methanol saturator thermostated at 40 °C. The molar ratio CO2/CH3OH was adjusted by the flux of CO2 controlled by mass flow controller and the temperature of saturator. Prior to the reaction, the catalyst was pretreated at 300 °C with CO2 for 2 h. The reaction products were analyzed with a gas phase chromatograph (Agilent 6890N) equipped with a flame ionization detector and a capillary column (HP-PLOT Q length 30m ID 0.53 mm). The condensed liquid was collected and analyzed. Catalytic activity was indicated by CH3OH conversion and product yields and selectivities. These parameters are calculated according to the following equations:
Yield ( i ) ( % ) = N i N o × 100
Conversion ( % ) = N r N o × 100 = Yield ( i )
S i ( % ) = Yield ( i ) Conversion × 100
  • No: number of moles of methanol introduced (mol/h)
  • Ni: number of moles of the product i formed (mol/h)
  • Nr: number of moles of reacted methanol (mol/h)

4. Conclusion

Synthesis of DMC from CH3OH and CO2 has been studied in gas phase system under atmospheric pressure using Co1.5PW12O40 as a catalyst. It was found that both CH3OH conversion and DMC yield decreased with increasing temperature, owing to the decreased CO2 adsorption on the catalyst at high temperatures. As for the decrease of DMC selectivity, this is probably due to the decomposition of DMC at higher temperatures.
Lower temperature leads to high selectivity of DMC to the detriment of that of DMM and MF. The optimal reaction conditions for the synthesis of DMC is lower temperature and high space velocity.

References

  1. Pacheco, MA; Marshall, CL. Review on Dimethyl carbonate manufacture and its characteristics as a fuel additive. Energy Fuels 1997, 11, 2–29. [Google Scholar]
  2. Guo, XC; Qin, ZF; Wang, GF; Wang, JG. Critical temperatures and pressures of reacting mixture in synthesis of dimethyl carbonate with methanol and carbon dioxide. Chin. Chem. Lett 2008, 19, 249–252. [Google Scholar]
  3. Ono, Y. Catalysis in the production and reactions of dimethyl carbonate, an environmentally benign building block. Appl. Catal. A 1997, 155, 133–166. [Google Scholar]
  4. Tundo, P; Selva, M. The Chemistry of Dimethyl Carbonate. Acc. Chem. Res 2002, 35, 706–716. [Google Scholar]
  5. Richter, M; Fait, MJG; Eckelt, R; Schneider, M; Radnik, J; Heidemann, D; Fricke, R. Gas-phase carbonylation of methanol to dimethyl carbonate on chloride-free Cu-precipitated zeolite Y at normal pressure. J. Catal 2007, 245, 11–24. [Google Scholar]
  6. Delledonne, D; Rivetti, F; Romano, U. Developments in the production and application of dimethylcarbonate. Appl. Catal. A: Gen 2001, 221, 241–251. [Google Scholar]
  7. Shaikh, AG; Sivaram, S. Organic Carbonates. Chem. Rev 1996, 96, 951–976. [Google Scholar]
  8. Romano, U; Tesei, R; Mauri, MM; Rebora, P. Synthesis of Dimethyl Carbonate from Methanol, Carbon Monoxide, and Oxygen Catalyzed by Copper Compounds. Ind. Eng. Chem. Prod. Dev 1980, 19, 396–403. [Google Scholar]
  9. Choi, JC; Sakakura, T; Sako, T. Reaction of Dialkyltin Methoxide with Carbon Dioxide Relevant to the Mechanism of Catalytic Carbonate Synthesis. J. Am. Chem. Soc 1999, 121, 3793–3794. [Google Scholar]
  10. Wu, XL; Xiao, M; Meng, YZ; Lu, YX. Direct synthesis of dimethyl carbonate (DMC) using Cu-Ni/VSO as catalyst. J. Mol. Catal. A: Chem 2006, 249, 93–97. [Google Scholar]
  11. Wu, XL; Xiao, M; Meng, YZ; Lu, YX. Direct synthesis of dimethyl carbonate on H3PO4 modified V2O5. J. Mol. Catal. A: Chem 2005, 238, 158–162. [Google Scholar]
  12. Jessop, PG; Ikariya, T; Noyori, R. Homogeneous Catalysis in Supercritical Fluids. Science 1995, 269, 1065–1069. [Google Scholar]
  13. Tundo, P; Selva, M. The Chemistry of Dimethyl Carbonate. Acc. Chem. Res 2002, 35, 706–716. [Google Scholar]
  14. Omae, I. Aspects of carbon dioxide utilization. Catal. Today 2006, 115, 33–52. [Google Scholar]
  15. Bian, J; Xiao, M; Wang, SJ; Lu, YX; Meng, YZJ. Novel application of thermally expanded graphite as the support of catalysts for direct synthesis of DMC from CH3OH and CO2. J. Colloid. Interface Sci 2009, 334, 50–57. [Google Scholar]
  16. Bian, J; Xiao, M; Wang, S; Wang, X; Lu, Y; Meng, Y. Highly effective synthesis of dimethyl carbonate from methanol and carbon dioxide using a novel copper–nickel/graphite bimetallic nanocomposite catalyst. Chem. Eng. J 2009, 147, 287–296. [Google Scholar]
  17. Wu, XL; Xiao, M; Meng, YZ; Lu, YX. Direct synthesis of dimethyl carbonate on H3PO4 modified V2O5. J. Mol. Catal. A. Chem 2005, 238, 158–162. [Google Scholar]
  18. Rocchiccioli-Deltcheff, C; Fournier, M; Franck, R; Thouvenot, R. Vibrational nvestigations of polyoxometalates. Evidence for anion-anion interactions in molybdenum(VI) and tungsten(VI) compounds related to the Keggin structure. Inorg. Chem 1983, 22, 207–216. [Google Scholar]
  19. Rocchiccioli-Deltcheff, C; Fournier, M. Catalysis by Polyoxometalates. Part 3,–Inflence of Vanadium (V) on the thermal Stability of 12-Metallophosphoric Acids from In Situ Infrared Studies. J. Chem. Soc. Faraday Trans 1991, 87, 3913–3920. [Google Scholar]
  20. Fournier, M; Feumi-Jantou, C; Rabia, C; Herve, G; Launay, S. Polyoxometalates Catalyst Materials: X-Ray Thermal Stability Study of Phosphorus-containing Heteropolyacids H3+xPM12-xVxO40.13–14H2O (M=Mo, W; x=0-1). J. Mater. Chem 1992, 2, 971–978. [Google Scholar]
  21. Gao, R; Chen, H; Le, Y; Dai, W-L; Fan, K. Highly active and selective Cs2.5H0.5PW12O40/SBA-15 composite material in the oxidation of cyclopentane-1,2-diol to glutaric acid by aqueous H2O2. Appl. Catal 2009, 352, 61–65. [Google Scholar]
  22. Fuchs, VM; Pizzio, LR; Blanco, MN. Synthesis and characterization of aluminum or copper tungstophosphate and tungstosilicate immobilized in a polymeric blend. Eur. Polym. J 2008, 44, 801–807. [Google Scholar]
  23. Cabello, CI; Egusquiza, MG; Botto, IL; Minelli, G. Heteropolyoxotungstates containing a wide catalytic target: reducibility and thermal stability. Mater. Chem. Phys 2004, 87, 264–274. [Google Scholar]
  24. Fu, Y; Zhu, H; Shen, J. Thermal decomposition of dimethoxymethane and dimethyl carbonate catalyzed by solid acids and bases. Thermochim. Acta 2005, 434, 88–92. [Google Scholar]
  25. Zhao, T; Han, Y; Sun, Y. Novel reaction route for dimethyl carbonate synthesis from CO2 and methanol. Fuel Proces. Tech 2000, 62, 187–194. [Google Scholar]
  26. Wang, XJ; Xiao, M; Wang, SJ; Lu, YX; Meng, YZ. Direct synthesis of dimethyl carbonate from carbon dioxide and methanol using supported copper (Ni, V, O) catalyst with photo-assistance. J. Mol. Catal. A: Chem 2007, 278, 92–96. [Google Scholar]
Figure 1. IR spectrum of Co1.5PW12O40 Keggin-type heteropolyanion.
Figure 1. IR spectrum of Co1.5PW12O40 Keggin-type heteropolyanion.
Ijms 11 01343f1
Figure 2. XRD pattern of Co1.5PW12O40 Keggin-type heteropolyanion.
Figure 2. XRD pattern of Co1.5PW12O40 Keggin-type heteropolyanion.
Ijms 11 01343f2
Figure 3. Differential thermal analysis (DTA) and thermo-gravimetric analysis (TGA) of Co1.5PW12O40 (heating rate: 5 °C/min).
Figure 3. Differential thermal analysis (DTA) and thermo-gravimetric analysis (TGA) of Co1.5PW12O40 (heating rate: 5 °C/min).
Ijms 11 01343f3
Figure 4. The dependence of the conversion and product yields on reaction temperature over Co1.5PW12O40. Reaction conditions: MWHSV = 3.25 h−1; molar ratio CH3OH/CO2 = 1.9.
Figure 4. The dependence of the conversion and product yields on reaction temperature over Co1.5PW12O40. Reaction conditions: MWHSV = 3.25 h−1; molar ratio CH3OH/CO2 = 1.9.
Ijms 11 01343f4
Figure 5. The dependence of the conversion and product selectivity on reaction temperature over Co1.5PW12O40. Reaction conditions: MWHSV = 3.25 h−1; molar ratio CH3OH/CO2 = 1.9.
Figure 5. The dependence of the conversion and product selectivity on reaction temperature over Co1.5PW12O40. Reaction conditions: MWHSV = 3.25 h−1; molar ratio CH3OH/CO2 = 1.9.
Ijms 11 01343f5
Figure 6. Time on stream effect on the conversion and product yields. Reaction conditions: T = 200 °C; MWHSV = 3.25 h−1; molar ratio CH3OH/CO2 = 1.9.
Figure 6. Time on stream effect on the conversion and product yields. Reaction conditions: T = 200 °C; MWHSV = 3.25 h−1; molar ratio CH3OH/CO2 = 1.9.
Ijms 11 01343f6
Figure 7. The product selectivities as a function of time on stream. Reaction conditions: T = 200 °C; MWHSV = 3.25 h−1; molar ratio CH3OH/CO2 = 1.9.
Figure 7. The product selectivities as a function of time on stream. Reaction conditions: T = 200 °C; MWHSV = 3.25 h−1; molar ratio CH3OH/CO2 = 1.9.
Ijms 11 01343f7
Figure 8. Conversion and DMC selectivity over Co1.5PW12O40 as a function of MWHSV. Reaction conditions: T = 200 °C; molar ratio CH3OH/CO2 = 1.9.
Figure 8. Conversion and DMC selectivity over Co1.5PW12O40 as a function of MWHSV. Reaction conditions: T = 200 °C; molar ratio CH3OH/CO2 = 1.9.
Ijms 11 01343f8

Share and Cite

MDPI and ACS Style

Aouissi, A.; Al-Othman, Z.A.; Al-Amro, A. Gas-Phase Synthesis of Dimethyl Carbonate from Methanol and Carbon Dioxide Over Co1.5PW12O40 Keggin-Type Heteropolyanion. Int. J. Mol. Sci. 2010, 11, 1343-1351. https://doi.org/10.3390/ijms11041343

AMA Style

Aouissi A, Al-Othman ZA, Al-Amro A. Gas-Phase Synthesis of Dimethyl Carbonate from Methanol and Carbon Dioxide Over Co1.5PW12O40 Keggin-Type Heteropolyanion. International Journal of Molecular Sciences. 2010; 11(4):1343-1351. https://doi.org/10.3390/ijms11041343

Chicago/Turabian Style

Aouissi, Ahmed, Zeid Abdullah Al-Othman, and Amro Al-Amro. 2010. "Gas-Phase Synthesis of Dimethyl Carbonate from Methanol and Carbon Dioxide Over Co1.5PW12O40 Keggin-Type Heteropolyanion" International Journal of Molecular Sciences 11, no. 4: 1343-1351. https://doi.org/10.3390/ijms11041343

Article Metrics

Back to TopTop