Next Article in Journal
Wet Peroxide Oxidation of Paracetamol Using Acid Activated and Fe/Co-Pillared Clay Catalysts Prepared from Natural Clays
Previous Article in Journal
Magnetic Nanoparticles for Photocatalytic Ozonation of Organic Pollutants
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effects of Basic Promoters on the Catalytic Performance of Cu/SiO2 in the Hydrogenation of Dimethyl Maleate

State Key Laboratory Breeding Base of Green Chemistry Synthesis Technology, Institute of Industrial Catalysis, Zhejiang University of Technology, Hangzhou 310014, China
*
Author to whom correspondence should be addressed.
Catalysts 2019, 9(9), 704; https://doi.org/10.3390/catal9090704
Submission received: 19 July 2019 / Revised: 17 August 2019 / Accepted: 18 August 2019 / Published: 22 August 2019

Abstract

:
Continuous hydrogenation of dimethyl maleate (DMM) toγ-butyrolactone (GBL), 1,4-butanediol (BDO) and tetrahydrofuran (THF) is a promising process in industry. In this study, Cu-M/SiO2 catalysts modified by basic promoters (M = Mg, Ca, Sr, Ba, La) were prepared, and characterized by physical adsorption of N2, in situ XRD, H2-TPR, CO2-TPD. With the addition of basic promoters, the basicity of Cu-M/SiO2 catalysts was improved. The particle size of CuO on Cu-M/SiO2 catalyst was increased after modified by Mg, Ca, Sr, Ba. However, the CuO particle was decreased on the Cu-La/SiO2 catalyst. The series of Cu-M/SiO2 catalyst was applied to the hydrogenation of DMM. The addition of basic promoters increased the selectivity of GBL during the hydrogenation for the basic promoters improved the dehydrogenation of BDO to GBL in alkaline sites. Furthermore, Cu-La/SiO2 presented a higher activity in the hydrogenation of DMM, due to its higher dispersion of Cu.

1. Introduction

Dimethyl maleate (DMM) as an important chemical intermediate is widely used in pharmaceutical, pesticide, chemical fiber, automobile and other industries [1,2,3]. Moreover, the hydrogenation products of DMM, such as γ-butyrolactone (GBL), 1,4-butanediol (BDO) and tetrahydrofuran (THF) are all high value-added chemical intermediates (the reaction pathways of DMM as shown in Scheme 1) [4,5,6,7]. With the frequent breakthrough of the oxidation technology of n-butane to maleic anhydride and the decrease of production cost, it will be of significance for the development of the downstream products of maleic anhydride. Esterification of maleic anhydride and methanol to DMM, and then continuous hydrogenation to BDO, GBL and THF, are considered to be a very promising technology. The key to this process is the catalyst for hydrogenation of DMM.
The catalysts for the hydrogenation of esters mainly include Ru [8,9,10,11], Pt [12], Pd [13,14], Cu [15,16,17,18], and so on. The noble metal catalysts, such as Ru, Pt and Pd [19,20], have the advantages of high activity, but their price is expensive. A supported Cu catalyst with a strong capacity for the adsorption, activation of C=O bond, and low hydrogenolysis activity of C–C bond, is cheap and easy to prepare—and then, it is widely used in ester hydrogenation reactions [21]. However, unmodified Cu catalyst has the disadvantage of poor activity and difficult to control the selectivity of hydrogenation products. Improving the preparation method of the supported Cu catalyst is the main way to enhance its catalytic performance. The preparation methods of supported Cu catalysts are as follows: Impregnation [22,23], ion exchange [24,25], sol-gel [26,27], deposition-precipitation [28], ammonia evaporation (AE) [29,30,31]. The AE method was invented by the Japanese Ube and used in the hydrogenation of esters [29]. It is a new method, based on ion exchange and deposition–precipitation: Copper ammonia complex ion is formed by ammonia and copper species precursor, followed by ion exchange with the hydroxyl on the surface of SiO2, and the excess copper ammonia complex ions forming Cu(OH)2 deposited on the surface of SiO2 with the evaporation of ammonia. The dispersion of Cu in the Cu/SiO2 catalyst prepared by the AE method is high, and it is difficult to sinter at high temperature. Moreover, there are two kinds of copper species: Cu+ and CuO, both of which have been proved to have a synergistic effect, which can further improve the activity of ester hydrogenation. Therefore, the Cu/SiO2 catalyst prepared by the AE method has the advantages of good dispersion and high activity, and has become a hot spot in recent years. Additives are an important factor affecting the catalyst activity, which changes the electronic structure of catalyst to improve the catalytic activity and selectivity, and also improve the dispersion of the active component and stability [32,33,34,35,36]. Yin et al. introduced NiO to improve the dispersion of metal Cu, and increased the TOF(turn of frequency) value; therefore, the introduction of NiO effectively improved the catalytic performance [32]. He et al. prepared Cu/SiO2 catalysts promoted by different content of B, and found that the addition of an appropriate amount of B could improve the dispersion of Cu species and prevent the agglomeration of copper particles at high temperature [34]. With the increase of B content, the content of Cu+ showed a rising trend, when Cu/B = 6.6, the content of Cu+ was the highest, and the activity and stability of the Cu-B/SiO2 catalyst was much higher than that of the conventional Cu/SiO2 catalyst [34].
In industry, BDO, GBL and THF are co-produced by the hydrogenation of DMM with the yield of THF the highest [31], and the yield of different products was controlled by varying the reaction conditions according to the market. How to further improve the activity of supported Cu catalyst and regulate the selectivity of DMM hydrogenation products controllably is a hot research topic. Guo et al. used Cu–B/γ-Al2O3 catalysts to obtain 100% DMM conversion at 533 K, 5 Mpa, H2/DMM 120 and LHSV (Liquid hourly space velocity) 0.36 h−1 [7]. Wang et al. enhanced the DMM conversion through adsorption precipitation compared to impregnation [37]. What is more, Chen et al. reported that Cu/SBA-15 could reach 100% DMM conversion in dimeathyl maleate hydrogenation [38]. However, the study on the catalytic performance of Cu/SiO2 catalyst modified with alkaline additives in DMM hydrogenation is quite incomplete. In this paper, the basicity of the SiO2 vector was modulated by adding the basic additives (Mg, Ca, Sr, Ba, La), and the effect of different alkali metals on the dispersion of Cu in Cu/SiO2 catalyst prepared by the AE method was explored. Furthermore, the effects of Cu dispersion and the basicity of the supports on the catalytic performance of hydrogenation of DMM were studied.

2. Results and Discussion

2.1. Catalysts Characterization

To begin, 25 wt% Cu/SiO2 and 25 wt% Cu-5 wt% M/SiO2 were characterized by N2 physical adsorption, in situ XRD, H2-TPR and CO2-TPD.
The surface area, pore volume and pore size of SiO2, Cu/SiO2 and Cu-M/SiO2 were characterized by N2 physical adsorption, as shown in Table 1. After loading active component Cu and metal promoter M, the surface area and pore volume of SiO2 were decreased, due to the large loading amount of Cu and M. However, the average pore diameters of Cu/SiO2 and Cu-M/SiO2 were both larger than that of pure supporter SiO2, which should be due to the preparation method (AE) of Cu/SiO2 and Cu-M/SiO2 for the dissolution of partial SiO2 during the preparation under high temperature (363 K) and basic condition. Then, Si(OH)4 formed by the dissolution of SiO2 in the basic solution was further reacted with neutral Cu(OH)2(H2O)4 in solution to form phyllosilicate [31,39].
In order to investigate the effect of adding alkaline additives on the dispersion of Cu species on the catalyst, in situ XRD characterizations of the calcined catalysts were carried out, and the result was shown in Figure 1. The peak at 2θ = 22.5° is attributed to the amorphous SiO2, and the peaks at about 2θ = 43.2° and 50.5° are attributed to Cu. By comparing in situ XRD patterns of Cu-M/SiO2 and Cu/SiO2, it was found that the peaks of CuO of Cu-M/SiO2 (M selected from Mg, Ca, Sr or Ba) at 2θ = 43.2° and 50.5° were much sharper than the peaks of Cu/SiO2 with no alkaline additives, which indicated that adding alkaline additives as Mg, Ca, Sr or Ba led to much larger Cu particles. However, there was no obvious Cu diffraction peaks at 2θ = 43.2° and 50.5° in the in situ XRD patterns of Cu-La/SiO2 modified by La, suggesting the Cu particles were too small to detect by in situ XRD. This conclusion is consistent with the result of Cu-M/SiO2 samples detected by XRD in Figure S1. We think that the effect of the reduction process has little effect on the dispersion of Cu/SiO2 catalyst. Therefore, the Cu dispersion of Cu-La/SiO2 is higher than that of unmodified Cu/SiO2. The addition of La increased the dispersion of Cu species on SiO2, resulting in an increase in the active Cu species per unit area.
H2-TPR tests of Cu-M/SiO2 and Cu/SiO2 were carried out to investigate the changes in the interaction between the active Cu species and the support forces upon the addition of alkaline additives (Figure 2). After the basic additive (except La) was added, the reduction peak of Cu-M/SiO2 was shifted slightly towards the low temperature. According to the results of in situ XRD analysis, the addition of basic additives (except La) decreased the dispersity of Cu species and increased its particle size, which also reduced the interaction between the Cu particles and the SiO2 support, leading to a decrease in the reduction temperature of the catalyst when the alkaline additive (except La) was added. However, for Cu-La/SiO2 catalyst with La additive, the reduction temperature of H2-TPR was slightly higher than that of unmodified Cu/SiO2, which suggested that the addition of La enhanced the interaction of Cu species with the SiO2 support, thus making Cu species more difficult to be reduced. However, for Cu-La/SiO2 with La promoter, the reduction temperature of H2-TPR was slightly higher than that of unmodified Cu/SiO2, which indicated that the addition of La enhanced the interaction between Cu species and carriers and made the Cu species more difficult to be reduced. Similar phenomena have also been found in La modified Cu/SiO2 investigated by Zheng et al., due to the formation of a Cu-O-La bond at the interface between LaOx and Cu species, which improved the stability of the catalyst [40]. In addition, with the characterization of in situ XRD, smaller CuO particles in Cu-La/SiO2 were detected, and this also led to the stronger interaction between Cu and the carrier, thus increasing the reduction temperature of Cu-La/SiO2. What is more, the ratio of different catalysts’ peak area is 1.11:0.98:0.94:1.02:1.07:1 for Cu/SiO2, Cu-La/SiO2, Cu-Sr/SiO2, Cu-Mg/SiO2, Cu-Ca/SiO2 and Cu-Ba/SiO2. It means the H2 consumptions of catalysts are similar.
In order to further explore the acid-base property of the catalyst surface after adding basic promoters, the CO2-TPD of Cu/SiO2 and Cu-M/SiO2 catalysts were studied. As shown in Figure 3, both Cu/SiO2 and Cu-M/SiO2 catalysts have a significant CO2 desorption peak at 400~500 K in the CO2-TPD profiles, and the position and intensity of the desorption peak are very close. However, a CO2 desorption peak of the Cu-M/SiO2 catalyst is observed at 500~580 K, and the intensity of the peak is less than that of the main desorption peak at 400~500 K. In CO2-TPD, the higher the temperature of desorption peak of CO2 means the stronger the basicity of the catalyst; therefore, the addition of basic additives really increased the basicity of the catalyst surface. It is also found that the desorption peak of CO2 from (b) to (f) catalyst shifts slightly towards low temperature, indicating that the basicity of the catalyst surface is Cu-Mg/SiO2 > Cu-Ca/SiO2 > Cu-Sr/SiO2 > Cu-Ba/SiO2 > Cu-La/SiO2. The actual order of basicity of the oxides is: MgO < CaO < SrO < BaO < La2O3 [41], which is contrary to the basicity of the catalyst surface. The reason should be the basic promoters in the calcined Cu-M/SiO2 not only exist in the form of oxides, but also may exist in some form similar to Cu-O-M or M-O-Si [42].

2.2. Hydrogenation of DMM

The hydrogenation of DMM was carried out in a fixed bed reactor over unmodified Cu/SiO2 and basic metal modified Cu-M/SiO2 (M = Mg, Ca, Sr, Ba, La) to evaluate the influence of basic promoters to the catalytic performance of Cu-M/SiO2 catalysts. The conversion of DMM and the selectivity of the products were shown in Table 2.
The conversion rate of DMM reached 100% over the unmodified Cu/SiO2 catalyst showed its good activity, and THF as the only hydrogenated product was observed for the weak alkalinity of the Cu/SiO2 surface. What is more, when Cu/SiO2 exhibits low activity in high LHSV, the THF selectivity still maintains 100%. Therefore, we think high LHSV has little effect on selectivity. After the basic additive (except La) was added to Cu/SiO2 catalyst, the catalytic activity decreased significantly for the conversion of DMM over Cu-M/SiO2 (M = Mg, Ca, Sr, Ba) did not reach 100%, especially the conversion over Cu-Ba/SiO2 just 64.21%. However, the catalytic activity of La modified Cu-La/SiO2 was higher, and the conversion rate of DMM was also 100%.
The activity of Cu-M/SiO2 (M = Mg, Ca, Sr, Ba, La) in the hydrogenation of DMM should be related to the dispersion of Cu on the surface of Cu-M/SiO2. The higher dispersion of Cu on Cu-La/SiO2 led to the higher catalytic activity for the hydrogenation of DMM, and the other Cu-M/SiO2 (M = Mg, Ca, Sr, Ba) with the lower dispersion of Cu than that of unmodified Cu/SiO2 catalyst caused that DMM had not been completely converted under the same reaction condition. The addition of basic promoter to Cu/SiO2 catalyst varied the selectivity of the hydrogenation products. There was just THF as the only product during the hydrogenation of DMM over the unmodified Cu/SiO2. After modified by basic promoter, the selectivity of THF decreased, and more GBL was obtained. In addition, a small amount of BDO was generated. One possible reason for the low selectivity of THF over modified catalysts was the covering of the part of the acid sites on the surface of SiO2 by the basic additives leading to the weakening of the dehydration of BDO to THF under acidic conditions. Moreover, we have tested the Cu-La/SiO2 catalyst for 100 h stability test. As shown in Table 3, We found the catalyst still maintain stable activity and selectivity.

3. Experimental

3.1. Catalysts Preparation

Preparation of Cu/SiO2: 3.78 g Cu(NO3)2·3H2O was dissolved in 40 mL deionized water, and stirred at room temperature for 10 min until it is completely dissolved, then 11.5 mL ammonia (28 wt%) was added to obtain the solution of copper ammonia with stirring for 30 min. Added 4 g SiO2 into the solution and stirred for 240 min. The mixture was quickly moved to the oil bath at 363 K to evaporate ammonia until the pH to 6.5, and then it was cooled to room temperature and filtered. The filter cake was washed by deionized water for three times followed with drying at 393 K for 480 min and calcination at 723 K for 240 min. The final catalyst was denoted as 25 wt% Cu/SiO2.
Preparation of Cu-M/SiO2: A certain amount of M nitrate (M selected from Mg, Ca, Sr, Ba, or La) containing M 0.2 g was dissolved in 40 mL deionized water. 4 g SiO2 was added into the solution. The mixture was filtered after had been stirred for 600 min, followed by drying at 393 K for 480 min and calcination at 723 K for 240 min. The modified carrier was denoted as M/SiO2. 3.78 g Cu(NO3)2·3H2O was dissolved in 40 mL deionized water, and stirred at room temperature for 10 min until it is completely dissolved, then 11.5 mL ammonia (28 wt%) was added to obtain the solution of copper ammonia with stirring for 30 min. Added the above made M/SiO2 into the solution and stirred for 240 min. The mixture was quickly moved to the oil bath at 363 K to evaporate ammonia until the pH to 6.5, and then it was cooled to room temperature and filtered. The filter cake was washed by deionized water for three times followed with drying at 393 K for 480 min and baking at 723 K for 240 min. The final catalyst was denoted as 25wt% Cu-5wt% M/SiO2.

3.2. Catalyst Characterization

N2 physical adsorption: The surface area, pore volume and pore diameter of the catalyst were determined by N2 physical adsorption at 77 K using Micromeritics ASAP 2020 (Micromeritics, Hangzhou, China). Firstly, the sample was heated to 573 K under vacuum condition for 6 h to remove the adsorbed species, and then N2 physical adsorption isotherm was performed. The surface area of the sample was calculated by the BET equation, and the pore volume and pore size distribution of the catalyst were obtained by BJH theory.
In situ XRD: X-ray powder diffraction patterns of the samples were obtained in the scanning angle (2θ) range of 10–80° under H2/Ar atmosphere at 573 K on a Thermo ARL SCINTAG X-TRA (Thermo, Hangzhou, China) using Cu Kα1 radiation (λ = 1.5406 Å) operated at 40 kV and 40 mA. Because of the in situ measurement system, the peak positions had some fluctuations compared with Standard XRD patterns.
The elemental content of the catalysts was detected using XRF (Thermo ARL ADVANT’X, Thermo, Hangzhou, China).
H2 temperature-programed reduction (H2-TPR)—0.1 g calcined catalyst was loaded into a quartz tube, and then a mixture of 5% H2/Ar was passed into the tube with the flow rate at 30 mL/min, while the exhaust gas was detected by a thermal conductivity detector (TCD). After the baseline was stable, the following heating program was carried out: retaining at 303 K for 10 min, and then rising to 1123 K for 82 min.
CO2 temperature-programed desorption (CO2-TPD)—0.1 g calcined catalyst was loaded into a U-tube, and treated at 723 K for 1 h in an Ar atmosphere to remove the adsorbed species. Then the temperature dropped to 373 K, and CO2 was passed into the U-tube. After CO2 adsorbed at 373 K for 30 min, the atmosphere was switched to Ar, and the exhaust gas was detected by a mass spectrum (MS). After the baseline was stable, the following heating program was carried out: rising to 773 K from 373 K for 40 min, and then retaining at 773 K for 10 min.

3.3. Hydrogenation Reaction

The hydrogenation reaction was carried out in a fixed bed reactor (Tianjin Pengxiang Technology Co. Ltd., Tianjin, China) with the diameter of 10 mm and the length of 500 mm and the constant temperature zone length ≥100 mm. The catalyst (20–40 mesh) was loaded into the middle of the constant temperature zone, and activated at 573 K with H2 (30 mL/min) for 240 min. Then the temperature dropped to 513 K. H2 was introduced up to the desired pressure for 5 MPa and the raw material (20 vol% DMM in methanol) was pumped into a vaporizer using a high-pressure pump (Series II), and then reacted with H2. The reaction products were quantitatively analyzed by gas chromatography (Agilent, GC, 7890A, Santa Clara, CA, USA) with DB-1 chromatographic column and FID (flame ionization detector) detector. In the reaction, methane and methol are also generated. However, methane was not detected in the reaction products, and therefore, the methoxy groups of DMM were probably converted to methanol. However, quantification of methanol was not possible because methanol was used as a solvent.

4. Conclusions

By comparing our previous work, Cu-M/SiO2 catalysts were modified by basic promoters (M = Mg, Ca, Sr, Ba, La) to improve the basicity of the surface of catalysts. After modification of the basic promoters, such as Mg, Ca, Sr and Ba, the diameter of CuO particles on the Cu-M/SiO2 catalyst increased, that is, the dispersion of Cu was reduced. On the contrary, adding La promoter reduced the CuO particle and improved the dispersion of CuO on SiO2. The selectivity of the hydrogenation products was modified during the continuous hydrogenation of DMM over Cu-M/SiO2 catalysts, due to the addition of basic promoters, which accelerated the dehydrogenation of BDO to more GBL in the basic active sites.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/9/9/704/s1, Figure S1: X-ray diffraction patterns of the calcined catalysts: (a) Cu/SiO2; (b) Cu-Mg/SiO2; (c) Cu-Ca/SiO2; (d) Cu-Sr/SiO2; (e) Cu-Ba/SiO2; (f) Cu-La/SiO2.

Author Contributions

Conceptualization, J.Y., X.H., Q.Z. and X.L.; Investigation, J.Y., X.H., Q.Z. and X.L.; Methodology, J.Y., X.H., Q.Z. and X.L.; Validation, J.Y., X.H., L.M., C.L., F.F., Q.Z. and X.L.; Visualization, Q.Z. and X.L.; Writing—original draft, J.Y., X.H., Q.Z. and X.L.; Writing—review and editing, L.M., C.L., F.F. and Q.Z. and X.L.

Funding

This research was funded by NSFC (Grant No. 21776258, 21476207, 91534113).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Muller, S.P.; Kucher, M.; Ohlinger, C.; Kraushaar-Czarnetzki, B. Extrusion of Cu/ZnO catalysts for the single-stage gas-phase processing of dimethyl maleate to tetrahydrofuran. J. Catal. 2003, 218, 419–426. [Google Scholar] [CrossRef] [Green Version]
  2. Ohlinger, C.; Kraushaar-Czarnetzki, B. Improved processing stability in the hydrogenation of dimethyl maleate to gamma-butyrolactone, 1,4-butanediol and tetrahydrofuran. Chem. Eng. Sci. 2003, 58, 1453–1461. [Google Scholar] [CrossRef]
  3. Zellmer, D.; Niewa, R.; Kreher, R.P. A functionalized dimethyl maleate (maleic acid dimethyl ester). Acta Cryst. 2014, 55, 823–825. [Google Scholar] [CrossRef]
  4. Emig, G.; Martin, F. Economics of maleic anhydride production from C-4 feedstocks. Catal. Today 1987, 1, 477–498. [Google Scholar] [CrossRef]
  5. Centi, G.; Trifiro, F.; Ebner, J.R.; Franchetti, V.M. Mechanistic aspects of maleic anhydride synthesis from C4 hydrocarbons over phosphorus vanadium oxide. Chem. Rev. 1988, 88, 55–80. [Google Scholar] [CrossRef]
  6. Wang, L.; Abudukelimu, N.; Ma, Y.B.; Qing, S.; Gao, Z.; Wumanjiang, E. Catalytic performance of Fe-modified Ru/Al2O3 in the hydrogenation of dimethyl maleate. J. Fuel Chem. Technol. 2014, 42, 839–844. [Google Scholar]
  7. Guo, P.J.; Chen, L.F.; Yan, S.R.; Dai, W.L.; Qiao, M.H.; Xu, H.L.; Fan, K.N. One-step hydrogenolysis of dimethyl maleate to tetrahydrofuran over chromium-modified Cu–B/γ-Al2O3 catalysts. J. Mol. Catal. A Chem. 2006, 256, 164–170. [Google Scholar] [CrossRef]
  8. Pritchard, J.; Filonenko, G.A.; van Putten, R.; Hensen, E.J.M.; Pidko, E.A. Heterogeneous and homogeneous catalysis for the hydrogenation of carboxylic acid derivatives: History, advances and future directions. Chem. Soc. Rev. 2015, 44, 3808–3833. [Google Scholar] [CrossRef] [PubMed]
  9. Luo, Z.C.; Bing, Q.M.; Kong, J.C.; Liu, J.Y.; Zhao, C. Mechanism of supported Ru3Sn7 nanocluster-catalyzed selective hydrogenation of coconut oil to fatty alcohols. Catal. Sci. Technol. 2018, 8, 1322–1332. [Google Scholar] [CrossRef]
  10. Kuwahara, Y.; Kaburagi, W.; Fujitani, T. Catalytic transfer hydrogenation of levulinate esters to gamma-valerolactone over supported ruthenium hydroxide catalysts. RSC Adv. 2014, 4, 45848–45855. [Google Scholar] [CrossRef]
  11. Fang, X.L.; Zhang, C.Y.; Chen, J.; Zhu, H.P.; Yuan, Y.Z. Synthesis and catalytic performance of ruthenium complexes ligated with rigid o-(diphenylphosphino)aniline for chemoselective hydrogenation of dimethyl oxalate. RSC Adv. 2016, 6, 45512–45518. [Google Scholar] [CrossRef]
  12. Stathis, P.; Stavroulaki, D.; Kaika, N.; Krommyda, K.; Papadogianakis, G. Low trans-isomers formation in the aqueous-phase Pt/TPPTS-catalyzed partial hydrogenation of methyl esters of linseed oil. Appl. Catal. B Environ. 2017, 209, 579–590. [Google Scholar] [CrossRef]
  13. Numwong, N.; Luengnaruemitchai, A.; Chollacoop, N.; Yoshimura, Y. Partial hydrogenation of polyunsaturated fatty acid methyl esters over Pd/activated carbon: Effect of type of reactor. Chem. Eng. J. 2012, 210, 173–181. [Google Scholar] [CrossRef]
  14. Pérez-Cadenas, A.F.; Kapteijn, F.; Zieverink, M.M.P.; Moulijn, J.A. Selective hydrogenation of fatty acid methyl esters over palladium on carbon-based monoliths: Structural control of activity and selectivity. Catal. Today 2007, 128, 13–17. [Google Scholar] [CrossRef]
  15. Li, F.; Lu, C.S.; Li, X.N. The effect of the amount of ammonia on the Cu0/Cu+ ratio of Cu/SiO2 catalyst for the hydrogenation of dimethyl oxalate to ethylene glycol. Chin. Chem. Lett. 2014, 25, 1461–1465. [Google Scholar] [CrossRef]
  16. Huang, H.; Cao, G.; Wang, S. An evaluation of alkylthiols and dialkyl disulfides on deactivation of Cu/Zn catalyst in hydrogenation of dodecyl methyl ester to dodecanol. J. Ind. Eng. Chem. 2014, 20, 988–993. [Google Scholar] [CrossRef]
  17. Gong, J.; Yue, H.; Zhao, Y.; Zhao, S.; Zhao, L.; Lv, J.; Wang, S.; Ma, X. Synthesis of Ethanol via Syngas on Cu/SiO2 Catalysts with Balanced Cu0–Cu+ Sites. J. Am. Chem. Soc. 2012, 134, 13922–13925. [Google Scholar] [CrossRef]
  18. Ma, X.; Yang, Z.; Liu, X.; Tan, X.; Ge, Q. Dynamic redox cycle of Cu0 and Cu+ over Cu/SiO2 catalyst in ester hydrogenation. RSC Adv. 2015, 5, 37581–37584. [Google Scholar] [CrossRef]
  19. Zhang, Q.; Li, K.; Xiang, Y.; Zhou, Y.; Wang, Q.; Guo, L.; Ma, L.; Xu, X.; Lu, C.; Feng, F.; et al. Sulfur-doped porous carbon supported palladium catalyst for high selective o-chloro-nitrobenzene hydrogenation. Appl. Catal. A Gen. 2019, 581, 74–81. [Google Scholar] [CrossRef]
  20. Sadjadi, S. Palladium nanoparticles immobilized on cyclodextrin-decorated halloysite nanotubes: Efficient heterogeneous catalyst for promoting copper- and ligand-free Sonogashira reaction in water–ethanol mixture. Appl. Organomet. Chem. 2018, 32. [Google Scholar] [CrossRef]
  21. Brands, D.S.; Poels, E.K.; Bliek, A. Ester hydrogenolysis over promoted Cu/SiO2 catalysts. Appl. Catal. A Gen. 1999, 184, 279–289. [Google Scholar] [CrossRef]
  22. Toupance, T.; Kermarec, M.; Louis, C. Metal Particle Size in Silica-Supported Copper Catalysts. Influence of the Conditions of Preparation and of Thermal Pretreatments. J. Phys. Chem. B 2000, 104, 965–972. [Google Scholar] [CrossRef]
  23. Munnik, P.; Wolters, M.; Gabrielsson, A.; Pollington, S.D.; Headdock, G.; Bitter, J.H.; de Jongh, P.E.; de Jong, K.P. Copper Nitrate Redispersion To Arrive at Highly Active Silica-Supported Copper Catalysts. J. Phys. Chem. C 2011, 115, 14698–14706. [Google Scholar] [CrossRef] [Green Version]
  24. Shimokawabe, M.; Takezawa, N.; Kobayashi, H. Characterization of copper-silica catalysts prepared by ion exchange. Appl. Catal. 1982, 2, 379–387. [Google Scholar] [CrossRef]
  25. Kohler, M.A.; Curry-Hyde, H.E.; Hughes, A.E.; Sexton, B.A.; Cant, N.W. The structure of CuSiO2 catalysts prepared by the ion-exchange technique. J. Catal. 1987, 108, 323–333. [Google Scholar] [CrossRef]
  26. Shi, L.; Zeng, C.; Jin, Y.; Wang, T.; Tsubaki, N. A sol–gel auto-combustion method to prepare Cu/ZnO catalysts for low-temperature methanol synthesis. Catal. Sci. Technol. 2012, 2, 2569–2577. [Google Scholar] [CrossRef]
  27. Ye, R.P.; Lin, L.; Chen, C.C.; Yang, J.X.; Li, F.; Zhang, X.; Li, D.J.; Qin, Y.Y.; Zhou, Z.; Yao, Y.G. Synthesis of Robust MOF-Derived Cu/SiO2 Catalyst with Low Copper Loading via Sol–Gel Method for the Dimethyl Oxalate Hydrogenation Reaction. ACS Catal. 2018, 8, 3382–3394. [Google Scholar] [CrossRef]
  28. Kasinathan, P.; Hwang, D.W.; Lee, U.H.; Hwang, Y.K.; Chang, J.S. Effect of Cu particle size on hydrogenation of dimethyl succinate over Cu–SiO2 nanocomposite. Catal. Commun. 2013, 41, 17–20. [Google Scholar] [CrossRef]
  29. Miyazaki, H.; Hirai, K.; Uda, T.; Nakamura, Y.; Ikezawa, H.; Tsuchie, T. Process for Producing Ethylene Glycol and/or Glycolic Acid Ester, Catalyst Composition Used Therefor, and Process for the Production of This Composition. Patent EP0064241A1, 4 September 1985. [Google Scholar]
  30. Chen, L.F.; Guo, P.J.; Qiao, M.H.; Yan, S.R.; Li, H.X.; Shen, W.; Xu, H.L.; Fan, K.N. Cu/SiO2 catalysts prepared by the ammonia-evaporation method: Texture, structure, and catalytic performance in hydrogenation of dimethyl oxalate to ethylene glycol. J. Catal. 2008, 257, 172–180. [Google Scholar] [CrossRef]
  31. Han, X.Q.; Zhang, Q.F.; Feng, F.; Lu, C.S.; Ma, L.; Li, X.N. Selective hydrogenation of dimethyl maleate to tetrahydrofuran over Cu/SiO2 catalyst: Effect of Cu+ on the catalytic performance. Chin. Chem. Lett. 2015, 26, 1150–1154. [Google Scholar] [CrossRef]
  32. Yin, A.; Wen, C.; Guo, X.; Dai, W.L.; Fan, K. Influence of Ni species on the structural evolution of Cu/SiO2 catalyst for the chemoselective hydrogenation of dimethyl oxalate. J. Catal. 2011, 280, 77–88. [Google Scholar] [CrossRef]
  33. Ungureanu, A.; Dragoi, B.; Chirieac, A.; Royer, S.; Duprez, D.; Dumitriu, E. Synthesis of highly thermostable copper-nickel nanoparticles confined in the channels of ordered mesoporous SBA-15 silica. J. Mater. Chem. 2011, 21, 12529–12541. [Google Scholar] [CrossRef]
  34. He, Z.; Lin, H.; He, P.; Yuan, Y. Effect of boric oxide doping on the stability and activity of a Cu–SiO2 catalyst for vapor-phase hydrogenation of dimethyl oxalate to ethylene glycol. J. Catal. 2011, 277, 54–63. [Google Scholar] [CrossRef]
  35. Zhao, S.; Yue, H.; Zhao, Y.; Wang, B.; Geng, Y.; Lv, J.; Wang, S.; Gong, J.; Ma, X. Chemoselective synthesis of ethanol via hydrogenation of dimethyl oxalate on Cu/SiO2: Enhanced stability with boron dopant. J. Catal. 2013, 297, 142–150. [Google Scholar] [CrossRef]
  36. Zhu, S.; Gao, X.; Zhu, Y.L.; Zhu, Y.F.; Zheng, H.; Li, Y. Promoting effect of boron oxide on Cu/SiO2 catalyst for glycerol hydrogenolysis to 1,2-propanediol. J. Catal. 2013, 303, 70–79. [Google Scholar] [CrossRef]
  37. Wang, L.; Abudukelimu, N.; Ma, Y.; Qing, S.; Gao, Z.; Eli, W. Enhanced Ru/Alumina catalyst via the adsorption-precipitation (AP) method for the hydrogenation of dimethyl maleate. React. Kinet. Mech. Catal. 2014, 112, 117–129. [Google Scholar] [CrossRef]
  38. Chen, L.; Guo, P.; Zhu, L.; Qiao, M.; Shen, W.; Xu, H.; Fan, K. Preparation of Cu/SBA-15 catalysts by different methods for the hydrogenolysis of dimethyl maleate to 1,4-butanediol. Appl. Catal. A Gen. 2009, 356, 129–136. [Google Scholar] [CrossRef]
  39. Van Der Grift, C.J.G.; Elberse, P.A.; Mulder, A.; Geus, J.W. Preparation of silica-supported copper catalysts by means of deposition-precipitation. Appl. Catal. 1990, 59, 275–289. [Google Scholar] [CrossRef]
  40. Zheng, X.; Lin, H.; Zheng, J.; Duan, X.; Yuan, Y. Lanthanum Oxide-Modified Cu/SiO2 as a High-Performance Catalyst for Chemoselective Hydrogenation of Dimethyl Oxalate to Ethylene Glycol. ACS Catal. 2013, 3, 2738–2749. [Google Scholar] [CrossRef]
  41. Zhang, G.; Hattori, H.; Tanabe, K. Aldol Addition of Acetone, Catalyzed by Solid Base Catalysts: Magnesium Oxide, Calcium Oxide, Strontium Oxide, Barium Oxide, Lanthanum (III) Oxide and Zirconium Oxide. Appl. Catal. 1988, 36, 189–197. [Google Scholar] [CrossRef]
  42. Zhang, B.; Zhu, Y.; Ding, G.; Zheng, H.; Li, Y. Modification of the supported Cu/SiO2 catalyst by alkaline earth metals in the selective conversion of 1,4-butanediol to γ-butyrolactone. Appl. Catal. A Gen. 2012, 443–444, 191–201. [Google Scholar] [CrossRef]
Scheme 1. Reaction pathways of dimethyl maleate (DMM).
Scheme 1. Reaction pathways of dimethyl maleate (DMM).
Catalysts 09 00704 sch001
Figure 1. In situ XRD patterns of the calcined catalysts: (a) Cu-La/SiO2; (b) Cu /SiO2; (c) Cu-Ba/SiO2; (d) Cu-Ca/SiO2; (e) Cu-Sr/SiO2; (f) Cu-Mg/SiO2.
Figure 1. In situ XRD patterns of the calcined catalysts: (a) Cu-La/SiO2; (b) Cu /SiO2; (c) Cu-Ba/SiO2; (d) Cu-Ca/SiO2; (e) Cu-Sr/SiO2; (f) Cu-Mg/SiO2.
Catalysts 09 00704 g001
Figure 2. H2-TPR profiles of the calcined catalysts: (a) Cu/SiO2; (b) Cu-Mg/SiO2; (c) Cu-Ca/SiO2; (d) Cu-Sr/SiO2; (e) Cu-Ba/SiO2; (f) Cu-La/SiO2.
Figure 2. H2-TPR profiles of the calcined catalysts: (a) Cu/SiO2; (b) Cu-Mg/SiO2; (c) Cu-Ca/SiO2; (d) Cu-Sr/SiO2; (e) Cu-Ba/SiO2; (f) Cu-La/SiO2.
Catalysts 09 00704 g002
Figure 3. CO2-TPD profiles of the calcined catalysts: (a) Cu/SiO2; (b) Cu-Mg/SiO2; (c) Cu-Ca/SiO2; (d) Cu-Sr/SiO2; (e) Cu-Ba/SiO2; (f) Cu-La/SiO2.
Figure 3. CO2-TPD profiles of the calcined catalysts: (a) Cu/SiO2; (b) Cu-Mg/SiO2; (c) Cu-Ca/SiO2; (d) Cu-Sr/SiO2; (e) Cu-Ba/SiO2; (f) Cu-La/SiO2.
Catalysts 09 00704 g003
Table 1. Physical properties of the SiO2 support and Cu-M/SiO2 catalysts.
Table 1. Physical properties of the SiO2 support and Cu-M/SiO2 catalysts.
SampleSurface Area (m2/g)Pore Volume (cm3/g)Average Pore Size (nm)Elemental Composition (wt%)
CuM
SiO21670.5011.800
Cu/SiO21170.4214.419.80
Cu-Mg/SiO21080.3914.318.83.2
Cu-Ca/SiO2950.3313.919.03.1
Cu-Sr/SiO21020.3614.118.92.9
Cu-Ba/SiO2990.3514.019.12.8
Cu-La/SiO21120.4014.418.93.0
Table 2. The catalytic performances of Cu/SiO2 and Cu-M/SiO2 catalysts in hydrogenation of DMM.
Table 2. The catalytic performances of Cu/SiO2 and Cu-M/SiO2 catalysts in hydrogenation of DMM.
CatalystsLHSVDMM Conversion/%Selectivity/%
(h−1)GBLBDOTHF
Cu/SiO20.310000100
Cu/SiO20.610000100
Cu/SiO20.983.6500100
Cu-Mg/SiO20.384.2124.434.8270.75
Cu-Ca/SiO20.388.3827.42072.58
Cu-Sr/SiO20.383.1732.173.4364.40
Cu-Ba/SiO20.364.2125.435.0169.56
Cu-La/SiO20.310024.544.2971.17
Reaction conditions: T = 513 K, p = 5 MPa, and H2/ester = 50.
Table 3. The influence of time on the Cu-La/SiO2 catalytic performance.
Table 3. The influence of time on the Cu-La/SiO2 catalytic performance.
Reaction Time (h)DMM Conversion (%)Selectivity (%)
GBLBDOTHF
010024.544.2971.17
510025.124.0470.84
1010024.984.1670.86
1510024.674.2871.05
2010024.384.3571.27
2510024.874.3170.82
3010025.224.1770.61
3510025.164.4270.42
4010024.454.5571
4510024.234.2671.51
5010024.794.1371.08
5510025.724.5969.69
6010025.484.2670.26
6510025.564.3770.07
7010024.944.1470.92
7510024.424.371.28
8010024.14.1771.73
8510023.874.4671.67
9010024.714.5870.71
9510025.664.2770.07
10010025.344.3370.33
Reaction conditions: T = 513 K, p = 5 MPa, LHSV = 0.3 h−1 and H2/ester = 50.

Share and Cite

MDPI and ACS Style

Ying, J.; Han, X.; Ma, L.; Lu, C.; Feng, F.; Zhang, Q.; Li, X. Effects of Basic Promoters on the Catalytic Performance of Cu/SiO2 in the Hydrogenation of Dimethyl Maleate. Catalysts 2019, 9, 704. https://doi.org/10.3390/catal9090704

AMA Style

Ying J, Han X, Ma L, Lu C, Feng F, Zhang Q, Li X. Effects of Basic Promoters on the Catalytic Performance of Cu/SiO2 in the Hydrogenation of Dimethyl Maleate. Catalysts. 2019; 9(9):704. https://doi.org/10.3390/catal9090704

Chicago/Turabian Style

Ying, Juntao, Xueqing Han, Lei Ma, Chunshan Lu, Feng Feng, Qunfeng Zhang, and Xiaonian Li. 2019. "Effects of Basic Promoters on the Catalytic Performance of Cu/SiO2 in the Hydrogenation of Dimethyl Maleate" Catalysts 9, no. 9: 704. https://doi.org/10.3390/catal9090704

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