Anda di halaman 1dari 7

Journal of CO2 Utilization 19 (2017) 33–39

Contents lists available at ScienceDirect

Journal of CO2 Utilization


journal homepage: www.elsevier.com/locate/jcou

Synthesis of dimethyl carbonate from methanol and CO2 over Fe–Zr


mixed oxides
Aixue Lia,b , Yanfeng Pua , Feng Lia , Jing Luoa,b , Ning Zhaoa,c,* , Fukui Xiaoa,c,*
a
State Key Laboratory of Coal Conversion, Institute of Coal Chemistry, Chinese Academy of Sciences, Taiyuan 030001, PR China
b
University of Chinese Academy of Sciences, Beijing 100039, PR China
c
National Engineering Research Center for Coal-based Synthesis, Taiyuan 030001, PR China

A R T I C L E I N F O A B S T R A C T

Article history:
Received 14 September 2016 A series of Fe–Zr mixed oxides with different Fe content were prepared and used for direct synthesis of
Received in revised form 23 February 2017 dimethyl carbonate (DMC) from CO2 and methanol. The best catalytic performance was achieved over the
Accepted 28 February 2017 Fe0.7Zr0.3Oy catalyst, with DMC yield of 0.44 mmolgcat1 and DMC selectivity of 100% under the reaction
Available online 10 March 2017 conditions of 110  C and 12 MPa. Characterization results of N2 physisorption, XRD, XPS, TPR and NH3/
CO2-TPD indicated that the Fe–Zr mixed oxides with coexistence structure of hexagonal Fe2O3 and cubic
Keywords: Fe2O3 favored the formation of moderately acidic and basic sites, which then improved the activation of
Fe–Zr mixed oxides CO2 and methanol. The DMC yield was shown to be linearly related to the amount of moderately acidic
Dimethyl carbonate
and basic sites.
CO2
© 2017 Elsevier Ltd. All rights reserved.
Methanol

1. Introduction system [18–20]. Supported metal catalysts such as Cu–Ni [21–25]


or Cu–Fe [26,27] were also investigated which exhibited better
Dimethyl carbonate (DMC) could be used as green and catalytic performance in fixed bed reactor. Modified metal oxides,
alternative chemical for corrosive and toxic reagents such as especially zirconia-based [28,29], cerium-based [30–33], vanadi-
dimethyl sulfate and phosgene in methylation or carbonylation um-based catalysts [34] had also shown good catalytic activity due
processes [1,2]. Compared with traditional synthetic methods [3– to both the acid and base sites on the catalyst surface. Among them,
10], the direct synthesis of dimethyl carbonate from CO2 and the amphoteric oxide, ZrO2 with vicinal acid-basic active sites
methanol was an economic and environmental route [11,12], which showed better catalytic performance in the direct synthesis of
not only could avoid using of toxic substances such as CO and DMC. Tomishige [35,36] had found that the catalytic activity of
phosgene, but also solved the problem of CO2 mitigation and ZrO2 could be improved after modification of H3PO4. Jiang [37,38]
excess capacity of methanol. However, due to the limit of had reported that H3PW12O40 modified ZrO2 had better catalytic
thermodynamic equilibrium, DMC yield was still very low. activity than ZrO2. Tomishige [39,40] had also prepared the CeO2–
Therefore, the development of high activity and high stability ZrO2 solid solution with better activity under high calcination
catalyst was demanded. temperature.
At present, metal tetra-alkoxides, metal oxides, metal acetates, Iron (III) oxide with acidic and basic sites was low cost, eco-
alkali compound, supported metal catalysts etc. were used. Metal friendly and thus widely investigated in many heterogeneous
tetra-alkoxides [13–17] were considered to be the most active reactions, such as isomerizations, Friedel-Crafts reactions, etc. [41–
catalysts, among which organotin tetra-alkoxides were widely 43], methanol gas-sensing test [44]. Because acid-base bifunction-
studied. However, these homogeneous catalysts were difficult to al catalysts were efficient for the direct synthesis of DMC from
recycling due to the decomposition in the present of H2O that lead methanol and CO2 [28,45,46], and few works about the Fe–Zr
to deactivation. Recently, alkali compound such as potassium mixed oxides as catalyst for the reaction were reported. In the
carbonate were used in which corrosive and expensive methyla- present paper, we proposed to modify ZrO2 with Fe to improve the
tion agent such as methyl iodide need to be added to the reaction acidity and basicity on the catalyst surface and tried to clarify the
relationship between the surface structure, acid-base property and
the catalytic performance.
* Corresponding authors at: State Key Laboratory of Coal Conversion, Institute of
Coal Chemistry, Chinese Academy of Sciences, Taiyuan 030001, PR China.
E-mail addresses: zhaoning@sxicc.ac.cn (N. Zhao), xiaofk@sxicc.ac.cn (F. Xiao).

http://dx.doi.org/10.1016/j.jcou.2017.02.016
2212-9820/© 2017 Elsevier Ltd. All rights reserved.
34 A. Li et al. / Journal of CO2 Utilization 19 (2017) 33–39

2. Experimental were charged into the reactor. Then CO2 was introduced into the
reactor with an initial pressure of 5.0 MPa at room temperature.
2.1. Catalyst preparation The reactor was then heated to 110  C and stirred constantly for 4 h.
After the reaction, the reactor was cooled to room temperature and
Fe-Zr mixed oxides (Fe xZr1-xOy) with different Fe content depressurized. In order to quantitatively analyze composition of
(x = 0.3, 0.5, 0.7, 0.9) were prepared by sol-gel process using Fe the mixture, n-butyl alcohol was added as an internal standard
(NO3)39H2O and Zr(NO3)45H2O as precursors. Typically, ethanol substance. The products were analyzed by a gas chromatograph
solution (200 mL, 0.2 mol L1) with different Fe(NO3)39H2O and Zr (FID-GC920) equipped with a capillary column (DB-210, 25 m
(NO3)44H2O molar ratio was dropped into a 400 mL beaker. A  0.22 mm).
solution of citric acid in ethanol (0.2 mol L1, 100 mL) was then
slowly added with stirring. After continuous stirring for further 3. Results and discussion
24 h at room temperature, the mixed solution was evaporated at
80  C overnight to form the sponge yellow gel. Then, the gel was 3.1. Textural and structural properties of the prepared samples
dried at 80  C for 24 h and milled followed by calcination at 500  C
in air for 4 h. The catalysts with different molar ratio of Fe2O3:ZrO2 Crystal structures of Fe-Zr mixed oxides were investigated by
was denoted as Fe xZr1-xOy (x = 0.3, 0.5, 0.7, 0.9). XRD and the spectra were shown in Fig. 1. For pure ZrO2, the
diffraction peaks at 2u of 24.4 (011), 28.2 (111), 31.5 (111), 40.7
2.2. Catalyst characterization (112) and 45.5 (202) could be assigned to the monoclinic phase
of ZrO2 (JCPDS 37-1484) while the characteristic peaks at 2u of
The specific surface area of the catalysts were determined by N2 30.2 (011), 35.2 (110), 50.3 (112), 60.7 (121), 62.9 (202) and
adsorption-desorption at liquid nitrogen temperature 196  C 74.5 (220) could be ascribed to the tetragonal phase of ZrO2 (JCPDS
using a Micromeritics Tristar II (3020) apparatus. The specific 50-1089). Pure Fe2O3 exhibited a hexagonal phase (JCPDS 33-0664)
surface areas were calculated according to the Brunauer-Emmett- at 2u of 24.1 (012), 33.1 (104), 35.6 (110), 40.8 (113), 49.4 (024),
Teller (BET) equation and the pore volume were calculated from 54.1 (116), 62.4 (214) and 63.9 (300).
the desorption branch of the isotherm using the Barrett, Joyner and After addition of Fe (Fe0.3Zr0.7Oy), the phase of zirconia
Halenda (BJH) equation. disappeared and only a very broad peak attributed to amorphous
The X-ray diffraction (XRD) patterns of catalysts were measured a-Fe2O3 was observed. It was inferred that iron-zirconium solid
on a Bruker D8 Advance (Germany) diffractometer equipped with a solution might be formed because the surface Fe/Zr ratio was
Cu Ka. The scattering intensities were measured over an angular closed to the bulk Fe/Zr ratio (Table 2) [47]. While for Fe0.5Zr0.5Oy
range of 10–80 at a scanning speed of 5 /min. and Fe0.7Zr0.3Oy, cubic phase (2u of 30.1 (220), 43.1 (400), 56.9
The X-ray photoelectron spectroscopy (XPS) spectra were (511), 62.6 (440)) of Fe2O3 was observed which suggested that
obtained using a Thermo ESCALAB 250 spectrometer equipped some Zr4+ ions had incorporated into the Fe2O3 lattice and lead to
with Al Ka radiation (h n = 1486.6 eV) under an ultrahigh vacuum. the crystal phase transformation of some Fe2O3 from hexagonal to
The values of binding energy (eV) were taken relatively to that of C cubic phase [48]. At high Fe content (Fe0.9Zr0.1Oy), the crystal
ls electrons of hydrocarbons on the sample surface (produced by structure was in hexagonal phase of Fe2O3, which was similar to
adventitious carbon), which was accepted to be equal to 284.8 eV. that of the pure Fe2O3.
The reducibility of the catalysts were studied by H2-tempera- The textures of the catalysts were listed in Table 1. The specific
ture-programmed reduction (H2-TPR). 50 mg of freshly calcined surface area of the Fe-Zr solid solutions increased significantly
catalysts was placed on the middle of glass wool in a quartz tube. from 20 m2 g1 to 87 m2 g1 upon increasing of the Fe content,
Then the TPR experiments were carried out in 5% H2/N2 at flow rate suggesting that the introduction of Fe was beneficial to the increase
of 50 mL min1. The temperature of the reactor was increased from of the surface area to some extend. High surface area might cause
room temperature to the final temperature of 1000  C with a the exposure of more active sites and thus lead to a higher catalytic
ramping rate of 10  C/min. Effluent gas was purified by silica gel activity. It could also be seen that the pore volumes increased from
desiccant and 5A molecular sieve. The hydrogen consumed during 0.034 cm3 g1 to 0.088 cm3 g1 with the introduction of Fe and the
TPR was monitored continuously by a thermal conductivity
detector (TCD).
Temperature programmed desorption of NH3 (NH3-TPD) was
carried out on the GAM 200 Mass Spectrometer for the
measurement of the acidity of the catalysts. Each sample (0.1 g)
was placed in the quartz reactor and pretreated in an Ar flow
(25 mL min1) at 500  C for 1 h. The adsorption of NH3 was
performed at 40  C with pure NH3 (50 mL min1) for 30 min,
followed by an Ar purge for 1 h to remove the physisorbed NH3. The
desorption process was performed at a heating rate of 10  Cmin1
from 40  C to 500  C and the evolved NH3 was monitored with a
thermal conductivity detector.
Temperature programmed desorption of CO2 (CO2-TPD) was
carried out on the Micromeritics Auto Chem II Chemisorption
Analyzer for the measurement of the basicity of the catalysts. The
operational process and condition were same with NH3-TPD.

2.3. Catalytic evaluation of catalysts

Direct synthesis of DMC from methanol and carbon dioxide was


carried out in a 50 mL stainless-steel autoclave equipped with a Fig. 1. XRD patterns of the Fe-Zr solid solutions: (^) hexagonal Fe2O3, (^) cubic
magnetic stirrer. Typically, methanol (12 g) and catalyst (1.0 g) Fe2O3, (*) tetragonal ZrO2, (5) monoclinic ZrO2.
A. Li et al. / Journal of CO2 Utilization 19 (2017) 33–39 35

Table 2
XPS results of different catalysts.

Sample Binding Energy (eV) Surface atomic ratios (%)

Zr 3d5/2 Fe 2p3/2 O 1s [Fe/Zr]Bulk [Fe/Zr]Surface O/(Fe + Zr)


ZrO2 182.0 – 529.8 – – –
Fe0.3Zr0.7Oy 182.0 710.7 529.8 0.42 0.43 1.96
Fe0.5Zr0.5Oy 182.1 710.9 529.8 1.00 1.34 1.91
Fe0.7Zr0.3Oy 182.0 711.1 529.8 2.33 2.40 1.97
Fe0.9Zr0.1Oy 182.1 710.8 529.9 9.00 10.54 1.97
Fe2O3 – 710.6 529.5 – – –

Table 1
Physical properties of different catalysts.

Sample Pore volume (cm3 g1) BET area (m2 g1)


ZrO2 0.034 20
Fe0.3Zr0.7Oy 0.021 40
Fe0.5Zr0.5Oy 0.045 66
Fe0.7Zr0.3Oy 0.088 87
Fe0.9Zr0.1Oy 0.035 44
Fe2O3 0.022 5

sample of Fe0.7Zr0.3Oy had the largest specific surface area and pore
volume.

3.2. XPS investigations

The XPS spectra of Zr 3d, O 1s and Fe 2p were shown in Fig. 2


and the peak position were summarized in Table 2. Binding energy
(eV) of Zr 3d5/2 at around 182.0 eV was typically associated with
zirconium in ZrO2 [47]. The spectra of Fe 2p showed the
characteristic Fe 2p3/2 and Fe 2p1/2 peaks at binding energy of
710.6–711.1 eV and 724.5 eV, respectively [49,50]. Two weak
satellite peaks at 718.0–719.0 eV and 733.4 eV were also distin-
guished [51,52], which indicated that the Fe was in +3 oxidation
state.
It was also found that the binding energy of Fe 2p3/2 changed
with the introduction of ZrO2. As Fe3+ was more electronegative
than Zr4+, the electron transferred from the iron atoms to the
zirconium atoms suggested that iron had a higher chemical state in
Fe–Zr solid solutions than that in pure Fe2O3 [53]. The interaction
between Fe2O3 and ZrO2 was in accordance with the XRD results.
The surface atomic ratios of Fe/Zr for Fe–Zr solid solutions were
summarized in Table 2. Each value was higher than that of the bulk,
e.g. although solid solutions were formed for the Fe–Zr solid
solutions, an enrichment of Fe on the surface was detected.
For O1s spectra, a significant peak at 529.5–529.8 eV (denoted
as Ob) was observed for Fe2O3, which could be assigned to the
lattice oxygen. However, a main peak (Ob) and a shoulder peak (Oa)
were detected for ZrO2 and Fe–Zr solid solutions. Generally, the
former corresponded to bulk oxygen, while the later represented
various surface oxygen species, such as O22 and O attributed to Fig. 2. XPS spectra of Fe 2p, Zr 3d and O 1s for Fe–Zr solid solutions.
the defect-oxide or hydroxyl-like group [54]. The surface atomic
ratios of O/(Fe + Zr) for Fe–Zr solid solutions were higher than those
of the stoichiometry (1.85, 1.75, 1.65, 1.55), respectively, suggesting (Fig. 3). The TPR profile of pure iron oxide showed two distinct
the existence of absorbed oxygen on the sample surfaces. reduction peaks, centered at 345  C resulted from the reduction of
Combined with BET results, the increase of absorbed oxygen on Fe2O3 to Fe3O4 and at 628  C assigned to the reduction of Fe3O4 to
the Fe–Zr materials might be associated with higher specific metallic iron [50]. The TPR profile of pure ZrO2 was flat, e.g. no H2-
surface area [55,56] (Table 1). consumption peak were found due to the high stability of the oxide
below 1000  C. For Zr-rich sample (Fe0.3Zr0.7Oy), two reduction
3.3. The reducibility of the catalysts peaks, could be seen which could be ascribed to the process of
Fe2O3 to FeO and FeO to Fe, respectively. The interaction between
H2-TPR experiments were conducted to study the influence of iron oxide and zirconium oxide might lead to stable FeO phase in
Fe content on the reduction property of the Fe–Zr solid solutions the reduction process. As a result, further reduction of the FeO to Fe
36 A. Li et al. / Journal of CO2 Utilization 19 (2017) 33–39

3.4. The surface acidity–basicity of the catalysts

To further explore the surface acid–base properties, NH3-TPD


and CO2-TPD techniques were employed and the results were
shown in Fig. 4 and Fig. 5, respectively.
Fig. 4 displayed the NH3-TPD profiles of Fe-Zr solid solutions.
The amount of acidic sites and peak positions were summarized in
Table 3. All profiles could be deconvoluted into two Gaussian
peaks. For pure ZrO2 and Fe2O3, two small peaks assigned to
weakly (50–150  C) and moderately (150–300  C) acidic sites could
be found. However, for the Fe-Zr solid solutions, two peaks
attributed to moderately and strongly (>300  C) acidic sites could
be observed. The results suggested that for the Fe–Zr solid
solutions, more moderately acidic sites were formed and increased
with the increasing of Fe content. However, the amount of
moderately acidic sites decreased and those of strong acid sites
increased for the Fe0.9Zr0.1Oy catalyst.
Fig. 5 displayed the CO2-TPD profiles of the Fe–Zr solid
Fig. 3. H2-TPR patterns of the Fe–Zr solid solutions.
solutions. The amount of basic sites and peak positions were
listed in Table 4. All profiles in Fig. 5 could be deconvoluted into
two Gaussian peaks. For pure Fe2O3, two weak peaks assigned to
were difficult and the reduction peak shifted to higher temperature weakly (50–150  C) and moderately (150–300  C) basic sites were
[48]. For the Fe-rich catalysts (Fe0.5Zr0.5Oy, Fe0.7Zr0.3Oy), three found. However, for pure ZrO2 and Fe-Zr solid solutions, two peaks
consecutive reduction peaks corresponding to the reduction of attributed to moderately and strongly (>300  C) basic sites were
Fe2O3 to Fe3O4, Fe3O4 to FeO, FeO to Fe were observed. The observed. Small amount of moderately and strongly basic sites of
reduction of the Fe0.9Zr0.1Oy was similar to that of the pure Fe2O3, pure ZrO2 might derive from Zr–O pairs and O2 anions [57–59].
e.g. the two reduction peaks at 360 and 578  C could be assigned to While for the Fe–Zr solid solutions (Fe0.3Zr0.7Oy, Fe0.5Zr0.5Oy,
result from the reduction of Fe2O3 to Fe3O4, Fe3O4 to Fe, Fe0.7Zr0.3Oy) more moderately basic sites were formed and
respectively. increased with the Fe content increasing which could ascribe to

Fig. 4. NH3-TPD profiles of Fe–Zr solid solutions. Fig. 5. CO2-TPD profiles of Fe–Zr solid solutions.
A. Li et al. / Journal of CO2 Utilization 19 (2017) 33–39 37

Table 3
Quantification of the NH3-TPD profiles of Fe-Zr solid solutions.

Sample Weakly Moderately Strongly


1 1

T ( C) Amount (mmol gcat ) T ( C) 
Amount (mmol gcat ) T ( C) Amount (mmol gcat1)
ZrO2 132 0.011 270 0.025 – –
Fe0.3Zr0.7Oy – – 237 0.041 376 0.014
Fe0.5Zr0.5Oy – – 212 0.052 356 0.025
Fe0.7Zr0.3Oy – – 208 0.066 349 0.013
Fe0.9Zr0.1Oy – – 215 0.047 346 0.044
Fe2O3 110 0.010 256 0.010 – –

Table 4
Quantification of the CO2-TPD profiles of Fe–Zr solid solutions.

Sample Weakly Moderately Strongly

T ( C) Amount (mmol gcat1) T ( C) Amount (mmol gcat1) T ( C) Amount (mmol gcat1)


ZrO2 – – 190 0.038 390 0.023
Fe0.3Zr0.7Oy – – 214 0.062 371 0.031
Fe0.5Zr0.5Oy – – 209 0.085 357 0.027
Fe0.7Zr0.3Oy – – 190 0.110 380 0.027
Fe0.9Zr0.1Oy – – 265 0.075 383 0.041
Fe2O3 115 0.019 240 0.012 – –

the following reasons: firstly, the specific surface area increased area might cause the exposure of more active sites and lead to a
upon the increasing of the Fe content which could then result in higher catalytic activity [57,58].
more exposed active sites. Moreover, the formation of Fe–Zr solid In addition, according to the mechanism of the synthesis of
solutions increased the amount of Zr–O, Fe–O pairs on the surface, DMC from CH3OH and CO2, methanol was activated to methyl
and lead to increased amount of moderately basic sites. However, it species and methoxy species on the acidic and basic sites of the
could be observed that the amount of strong basic sites increased catalyst, respectively [45,46]. Methoxy carbonate anion was then
and that of moderately basic sites decreased for Fe0.9Zr0.1Oy. formed by the reaction of methoxy species with carbon dioxide
adsorbed on the basic sites of the catalyst [28,45] followed by
3.5. Catalytic performance further reaction with methyl species to produce DMC. Therefore, in
this study, the low catalytic activity of pure ZrO2 and Fe2O3 might
The effects of the Fe content on the catalytic performance were result from smaller amount of acidic and basic sites on the surface
investigated and the results were shown in Table 5. It could be seen that lead to less activated methanol and CO2. The Fe-Zr solid
that no DMC was detected in the absence of catalyst which solution catalysts exhibited higher catalytic activities than pure
indicated that the catalyst played an important role for the ZrO2 and Fe2O3 which could be attributed to the increased amount
reaction. The pure ZrO2 and Fe2O3 catalyst showed relatively lower of moderately acidic and basic sites with the addition of Fe. Figs. 6
DMC yield than the Fe–Zr solid solutions. The selectivity of DMC on and 7 showed the relationships between catalytic activity and the
Fe–Zr solid solutions nearly approached 100%. The DMC yield amount of moderately acidic and basic sites, respectively. It could
increased with the increasing of Fe content, reaching a maximum be seen that the DMC yield increased linearly upon increasing of
and then decreased with further increase of the Fe amount. Among the amount of moderately acidic and basic sites.
the catalysts examined, the Fe0.7Zr0.3Oy catalyst showed the Combination of the TPD and XRD results might reveal that the
highest DMC yield of 0.44 mmol g cat1. However, the activity on coexistence structure of hexagonal Fe2O3 and cubic Fe2O3 favored
Fe–Zr solid solutions were lower than those of Ce–Zr solid solution the formation of moderately acidic and basic sites, e.g. the
catalysts [39,40,60] or CeO2-based catalysts [33] which showed the decreased catalytic activity of Fe0.9Zr0.1Oy might be due to the
activity of 1.42 mmol gcat1 or 2.3 mmol gcat1.
Reaction conditions: reaction temperature, 110  C; CH3OH
weight, 12 g; initial pressure, 5 MPa; reaction time, 4 h; catalyst
weight, 1.0 g; calcination temperature, 500  C.
In association with the result of the N2 adsorption, the
Fe0.7Zr0.3Oy solid solution with the largest surface area exhibited
higher catalytic performance suggesting that the catalytic activity
was correlated to the surface area of the catalysts, e.g. large surface

Table 5
Influence of Fe content on the yield and selectivity of DMC.

Sample DMC yield (mmol gcat1) DMC selectivity (%)


None 0 0
ZrO2 0.12 100
Fe0.3Zr0.7Oy 0.24 100
Fe0.5Zr0.5Oy 0.35 100
Fe0.7Zr0.3Oy 0.44 100
Fe0.9Zr0.1Oy 0.28 100
Fe2O3 0.04 100 Fig. 6. The relationships between catalytic activity and the amount of moderately
acidic sites.
38 A. Li et al. / Journal of CO2 Utilization 19 (2017) 33–39

[12] Arno Behr, Carbon dioxide as an alternative C1 synthetic unit activation by


transition-metal complexes, Angew. Chem. Int. Ed. 27 (1988) 661–678.
[13] Toshiyasu Sakakura, Jun-Chul Choi, Yuko Saito, Takeshi Sako, Synthesis of
dimethyl carbonate from carbon dioxide: catalysis and mechanism,
Polyhedron 19 (2000) 573–576.
[14] Binbin Fan, Hongyu Li, Weibin Fanb, Jilong Zhang, Ruifeng Li, Organotin
compounds immobilized on mesoporous silicas as heterogeneous catalysts for
direct synthesis of dimethyl carbonate from methanol and carbon dioxide,
Appl. Catal. A 372 (2010) 94–102.
[15] Kazufumi Kohno, Jun-Chul Choi, Yoshihiro Ohshima, Abulimiti Yili, Hiroyuki
Yasuda, Toshiyasu Sakakura, Reaction of dibutyltin oxide with methanol under
CO2 pressure relevant to catalytic dimethyl carbonate synthesis, J. Organomet.
Chem. 693 (2008) 1389–1392.
[16] Jiang Qi, Qihe Lin, Zhongtao Huang, A new catalytic process for the direct
synthesis of dimethyl carbonate from carbon dioxide and methanol, Chin. J.
Catal. 17 (1996) 91–92.
[17] Tiansheng Zhao, Yizhuo Han, Yuhan Sun, Novel reaction route for dimethyl
carbonate synthesis from CO2 and methanol, Fuel Process. Technol. 62 (2000)
187–194.
[18] Fahai Cao, Dingye Fang, Dianhua Liu, Weiyong Ying, Catalytic esterification of
carbon dioxide and methanol for the preparation of dimethyl carbonate, Fuel
Fig. 7. The relationships between catalytic activity and the amount of moderately Chem. Div. Prepr. 74 (2002) 295–297.
basic sites. [19] Shunnong Fang, Kaoru Fujimoto, Direct synthesis of dimethyl carbonate from
carbon dioxide and methanol catalyzed by base, Appl. Catal. A 142 (1996) L1–
L3.
transformation of coexistence structure of hexagonal Fe2O3 and [20] Qinghai Cai, Bin Lu, Lingji Guo, Yongkui Shan, Studies on synthesis of dimethyl
cubic Fe2O3 to hexagonal Fe2O3 that reduced the amount of the carbonate from methanol and carbon dioxide, Catal. Commun. 10 (2009) 605–
moderately acidic and basic sites. 609.
[21] Chuanfeng Li, Shunhe Zhong, Study on application of membrane reactor in
direct synthesis DMC from CO2 and CH3OH over Cu-KF/MgSiO catalyst, Catal.
4. Conclusions Today 82 (2003) 83–90.
[22] X.L. Wua, Y.Z. Meng, M. Xiao, Y.X. Lu, Direct synthesis of dimethyl carbonate
(DMC) using Cu-Ni/VSO as catalyst, J. Mol. Catal. A 249 (2006) 93–97.
Fe–Zr solid solution catalysts with different Fe content were [23] X.J. Wang, M. Xiao, S.J. Wang, Y.X. Lu, Y.Z. Meng, Direct synthesis of dimethyl
synthesized via sol–gel method which showed better catalytic carbonate from carbon dioxide and methanol using supported copper (Ni, V,
performance than pure Fe2O3 and ZrO2 for synthesis of DMC from O) catalyst with photo-assistance, J. Mol. Catal. A 278 (2007) 92–96.
[24] Jun Bian, Min Xiao, Shuanjin Wang, Yixin Lu, Yuezhong Meng, Direct synthesis
CO2 and methanol. The addition of iron to zirconia could influence of DMC from CH3OH and CO2 over V-doped Cu-Ni/AC catalysts, Catal.
the crystal structure of the ZrO2, and thus increased the amount of Commun. 10 (2009) 1142–1145.
the surface moderately acidic and basic sites effectively, which [25] Huiling Chen, Shuanjin Wang, Min Xiao, Dongmei Han, Yixin Lu, Yuezhong
Meng, Direct Synthesis of dimethyl carbonate from CO2 and CH3OH using
then activated the methanol and CO2. The DMC yield linearly
0.4 nm molecular sieve supported Cu-Ni bimetal catalyst, Chin. J. Chem. Eng.
increased with the increasing of the amount of moderately acidic 20 (2012) 906–913.
and basic sites. [26] Yingjie Zhou, Min Xiao, Shuanjin Wang, Dongmei Hana, Yixin Lu, Yuezhong
Meng, Effects of Mo promoters on the Cu-Fe bimetal catalysts for the DMC
formation from CO2 and methanol, Chin. Chem. Lett. 24 (2013) 307–310.
Acknowledgments [27] Yingjie Zhou, Shuanjin Wang, Min Xiao, Dongmei Han, Yixin Lu, Yuezhong
Meng, Novel Cu-Fe bimetal catalyst for the formation of dimethyl carbonate
This work was financially supported by Natural Science from carbon dioxide and methanol, RSC Adv. 17 (2012) 6831–6837.
[28] Keiichi Tomishig, Yoshiki Ikeda, Tomohiro Sakaihori, Kaoru Fujimoto, Catalytic
Foundation of Shanxi Province (No. 201601D102006) and the properties and structure of zirconia catalysts for direct synthesis of dimethyl
Key Science and Technology Program of Shanxi Province, China carbonate from methanol and carbon dioxide, J. Catal. 192 (2000) 355–362.
(MD2014-09, MD2014-10). [29] Keiichi Tomishige, Tomohiro Sakaihori, Yoshiki Ikeda, Kaoru Fujimoto, A novel
method of direct synthesis of dimethyl carbonate from methanol and carbon
dioxide catalyzed by zirconia, Catal. Lett. 58 (1999) 225–229.
References [30] Michele Aresta, Angela Dibenedetto, Carlo Pastore, Antonella Angelini,
Brunella Aresta, Imre Pápai, Influence of Al2O3 on the performance of CeO2
[1] Yoshio Ono, Catalysis in the production and reactions of dimethyl carbonate, used as catalyst in the direct carboxylation of methanol to dimethylcarbonate
an environmentally benign building block, Appl. Catal. A 155 (1997) 133–166. and the elucidation of the reaction mechanism, J. Catal. 269 (2010) 44–52.
[2] A.G. Shaikh, S. Sivaram, Organic carbonates, Chem. Rev. 96 (1996) 951–976. [31] Michele Aresta, Angela Dibenedetto, Carlo Pastore, Corrado Cuocci, Brunella
[3] Happy Babad, Anderew G. Zeiler, The chemistry of phosgene, Chem. Rev. 73 Aresta, Stefania Cometa, Elvira De Giglio, Cerium(IV) oxide modification by
(1973) 75–91. inclusion of a hetero-atom: a strategy for producing efficient and robust nano-
[4] S.T. King, Oxidative carbonylation of methanol to dimethyl carbonate by solid- catalysts for methanol carboxylation, Catal. Today 137 (2008) 125–131.
state ion-exchanged CuY catalysts, Catal. Today 33 (1997) 173–182. [32] Danielle Ballivet-Tkatchenko, Chambrey Stéphane, Riitta Keiski, Rosane
[5] Ugo Romano, Renato Tesel, Marcello Massi Mauri, Pierluigl Rebora, Synthesis Ligabue, Laurent Plasseraud, Philippe Richard, Helka Turunen, Direct
of dimethyl carbonate from methanol, carbon monoxide, and oxygen synthesis of organic carbonates from the reaction of CO2 with methanol and
catalyzed by copper compounds, Ind. Eng. Chem. Prod. Res. Dev. 19 (1980) ethanol over CeO2 catalysts, Catal. Today 115 (2006) 95–101.
396–403. [33] Honda Masayoshi, Tamura Masazumi, Nakagawa Yoshinao, Nakao Kenji,
[6] Tokuo Matsuzaki, Asumaru Nakamura, Dimethyl carbonate synthesis and Suzuki Kimihito, Tomishige Keiichi, Organic carbonate synthesis from CO2 and
other oxidative reactions, Catal. Surv. Asia 1 (1997) 77–88. alcohol over CeO2 with 2-cyanopyridine: scope and mechanistic studies, J.
[7] Hongyou Cui, FujunWang TaoWang, Chaoran Gu, Youyuan Dai PeilinWang, Catal. 318 (2014) 95–107.
Transesterification of ethylene carbonate with methanol in supercritical [34] X.L. Wu, M. Xiao, Y.Z. Meng, Y.X. Lu, Direct synthesis of dimethyl carbonate on
carbon dioxide, J. Supercrit. Fluids 30 (2004) 63–69. H3PO4 modified V2O5, J. Mol. Catal. A 238 (2005) 158–162.
[8] Kyung-Hoon Kim, Dong-Woo Kim, Cheol-Woong Kim, Jae-Cheon Koh, Dae- [35] Yoshiki Ikeda, Tomohiro Sakaihori, Keiichi Tomishige, Kaoru Fujimoto,
Won Park, Synthesis of dimethyl carbonate from transesterification of Promoting effect of phosphoric acid on zirconia catalysts in selective
ethylene carbonate with methanol using immobilized ionic liquid on synthesis of dimethyl carbonate from methanol and carbon dioxide, Catal. Lett.
commercial silica, Korean J. Chem. Eng. 27 (2010) 1441–1445. 66 (2000) 59–62.
[9] Feng Wang, Ning Zhao, Junping Li, Wenbo Zhao, Fukui Xiao, Wei Wei, Yuhan [36] Yoshiki Ikeda, Mohammad Asadullah, Kaoru Fujimoto, Keiichi Tomishige,
Sun, Modeling of the catalytic distillation process for the synthesis of dimethyl, Structure of the active sites on H3PO4/ZrO2 catalysts for dimethyl carbonate
Ind. Eng. Chem. Res. 46 (2007) 8972–8979. synthesis from methanol and carbon dioxide, J. Phys. Chem. B 105 (2001)
[10] Mouhua Wang, Hui Wang, Ning Zhao, Wei Wei, Yuhan Sun, High-yield 10653–10658.
synthesis of dimethyl carbonate from urea and methanol using a catalytic [37] KyungWon La, Ji Chul Jung, Heesoo Kim, Sung-Hyeon Baeck, In Kyu Song, Effect
distillation process, Ind. Eng. Chem. Res. 46 (2007) 2683–2687. of acid-base properties of H3PW12O40/Ce xTi1-xO2 catalysts on the direct
[11] W. Leitner, The coordination chemistry of carbon dioxide and its relevance for synthesis of dimethyl carbonate from methanol and carbon dioxide: a TPD
catalysis: a critical survey, Coord. Chem. Rev. 153 (1996) 257–284. study of H3PW12O40/Ce xTi1-xO2 catalysts, J. Mol. Catal. A 269 (2007) 41–45.
A. Li et al. / Journal of CO2 Utilization 19 (2017) 33–39 39

[38] Chunjie Jiang, Yihang Guo, Chungang Wang, Changwen Hu, Yue Wu, Enbo [49] Catalin F. Petre, Faïçal Larachi, Anoxic alkaline oxidation of bisulfide by Fe/Ce
Wang, Synthesis of dimethyl carbonate from methanol and carbon dioxide in oxides: reaction pathway, AIChE J. 53 (2007) 2170–2187.
the presence of polyoxometalates under mild conditions, Appl. Catal. A 256 [50] Zhaoliang Zhang, Dong Han, Shaojie Wei, Yexin Zhang, Determination of active
(2003) 203–212. site densities and mechanisms for soot combustion with O2 on Fe-doped CeO2
[39] Keiichi Tomishige, Yutaka Furusawa, Yoshiki Ikeda, Mohammad Asadullah, mixed oxides, J. Catal. 276 (2010) 16–23.
Kaoru Fujimoto, CeO2-ZrO2 solid solution catalyst for selective synthesis of [51] Weixin Huang, Wolfgang Ranke, Robert Schlögl, Reduction of an a-Fe2O3
dimethyl carbonate from methanol and carbon dioxide, Catal. Lett. 76 (2001) (0001) film using atomic hydrogen, J. Phys. Chem. C 111 (2007) 2198–2204.
71–74. [52] Toru Yamashita, Peter Hayes, Analysis of XPS spectra of Fe2+ and Fe3+ ions in
[40] Keiichi Tomishige, Kimio Kunimori, Catalytic and direct synthesis of dimethyl oxide materials, Appl. Surf. Sci. 254 (2008) 2441–2449.
carbonate starting from carbon dioxide using CeO2-ZrO2 solid solution [53] Nidhi Garg, Santanu Bera, G. Mangamma, Vinit K. Mittal, R. Krishnan, S.
heterogeneous catalyst: effect of H2O removal from the reaction system, Appl. Velmurugan, Study of Fe2O3-ZrO2 interface of ZrO2 coating grown by
Catal. A 237 (2002) 103–109. hydrothermal process on stainless steel, Surf. Coat. Technol. 258 (2014) 597–
[41] Giovanni Sartori, Raimondo Maggi, Use of solid catalysts in Friedel-Crafts 604.
acylation reaction, Chem. Rev. 106 (2006) 1077–1104. [54] Jean-Charles Dupin, Danielle Gonbeau, Philippe Vinatierb, Alain Levasseur,
[42] Yinyong Sun, Stéphane Walspurger, Jean-Philippe Tessonnier, Benoît Louis, Systematic XPS studies of metal oxides, hydroxides and peroxides, Phys. Chem.
Jean Sommera, Highly dispersed iron oxide nanoclusters supported on ordered Chem. Phys. 6 (2000) 1319–1324.
mesoporous SBA-15: a very active catalyst for Friedel-Crafts alkylations, Appl. [55] Kongzhai Li, Hua Wang, Yonggang Wei, Dongxia Yan, Partial oxidation of
Catal. A 300 (2006) 1–7. methane to syngas with air by lattice oxygen transfer over ZrO2-modified Ce-
[43] Nongyue He, Shulin Bao, Qinhua Xu, Fe-containing mesoporous molecular Fe mixed oxides, Chem. Eng. J. 173 (2011) 574–582.
sieves materials: very active Friedel-Crafts alkylation catalysts, Appl. Catal. A [56] Yi Li, Yuan Wan, Yanping Li, Sihui Zhan, Qingxin Guan, Yang Tian, Low-
169 (1998) 29–36. temperature selective catalytic reduction of NO with NH3 over Mn2O3-doped
[44] G. Neri, A. Bonavita, G. Rizzo, S. Galvagno, S. Capone, P. Siciliano, Methanol gas- Fe2O3 hexagonal microsheets, ACS Appl. Mater. Interfaces 8 (2016) 5224–5233.
sensing properties of CeO2-Fe2O3 thin films, Sens. Actuators B 114 (2006) 687– [57] Hongguang Li, Xi Jiao, Lei Li, Ning Zhao, Fukui Xiao, Wei Wei, Yuhan Sun,
695. Bingsheng Zhang, Synthesis of glycerol carbonate by direct carbonylation of
[45] Kyeong Taek Jung, Alexis T. Bell, An in situ infrared study of dimethyl carbonate glycerol with CO2 over solid catalysts derived from Zn/Al/La and Zn/Al/La/M
synthesis from carbon dioxide and methanol over zirconia, J. Catal. 204 (2001) (M = Li, Mg and Zr) hydrotalcites, Catal. Sci. Technol. 5 (2015) 989–1005.
339. [58] J.I. Di Cosimo, V.K. Diez, M. Xu, E. Iglesia, C.R. Apesteguía, Structure and surface
[46] Nicolas Keller, Guillaume Rebmann, Valérie Keller, Catalysts, mechanisms and and catalytic properties of Mg-Al basic oxides, J. Catal. 178 (1998) 499–510.
industrial processes for the dimethylcarbonate synthesis, J. Mol. Catal. A 317 [59] Hongguang Li, Dengzheng Gao, Peng Gao, Feng Wang, Ning Zhao, Fukui Xiao,
(2010) 1–18. Wei Wei, Yuhan Sun, The synthesis of glycerol carbonate from glycerol and CO2
[47] F. Wyrwalski, J.F. Lamonier, S. Siffert, E.A. Zhilinskaya, L. Gengembre, A. over La2O2CO3-ZnO catalysts, Catal. Sci. Technol. 3 (2013) 2801–2809.
Aboukais, Bulk and surface structures of iron doped zirconium oxide systems: [60] Zhifang Zhang, Zhongwen Liu, Jian Lu, Zhaotie Liu, Synthesis of dimethyl
influence of preparation method, J. Mater. Sci. 40 (2005) 933–942. carbonate from carbon dioxide and methanol over Ce xZr1-xO2 and [EMIM]Br/
[48] Kaidong Chen, Yining Fan, Zheng Hu, Qijie Yan, Study on the structure and Ce0.5Zr0.5O2, Ind. Eng. Chem. Res. 50 (2011) 1981–1988.
reduction behaviour of the iron-zirconium oxide system, J. Mater. Chem. 6
(1996) 1041–1045.

Anda mungkin juga menyukai