You are on page 1of 10

NJC

View Article Online


PAPER View Journal | View Issue
Published on 30 August 2017. Downloaded by National Chemical Laboratory, Pune on 7/25/2023 11:53:42 AM.

Direct synthesis of dimethyl carbonate from CO2


and methanol over CaO–CeO2 catalysts: the role
Cite this: New J. Chem., 2017,
41, 12231 of acid–base properties and surface oxygen
vacancies
Bin Liu, Congming Li,* Guoqiang Zhang, Lifei Yan and Zhong Li *

Herein, the direct synthesis of dimethyl carbonate (DMC) from CO2 and methanol was investigated over
a series of CeO2-based catalysts promoted by variable amounts of CaO. The catalysts were synthesized
by a co-precipitation method and fully characterized by powder X-ray diffraction (XRD), N2 adsorption/
desorption, transmission electron microscopy (TEM), Raman spectroscopy, X-ray photoelectron
spectroscopy (XPS), and NH3-TPD and CO2-TPD techniques. The results demonstrated that the addition
of different amounts of CaO could affect the structure and surface properties of the obtained catalysts.
The acid–base properties and amount of surface oxygen vacancies on the surface of the catalysts were
improved due to the interaction between CaO and CeO2, and the oxygen vacancies enhanced the
Received 18th July 2017, adsorption of CO2. The synthesis of DMC from CO2 and methanol was studied in a slurry bed reactor at
Accepted 30th August 2017 3 MPa and 140 1C. The results revealed that the acid–base properties and oxygen vacancies play a crucial
DOI: 10.1039/c7nj02606d role in promoting the formation of DMC. Among all the bimetallic and pure CeO2 catalysts, the Ca1.5Ce
sample showed the maximum DMC yield (2.47 mmol g1) due to the modification of the acid–base proper-
rsc.li/njc ties of the catalyst and the existence of higher concentration of oxygen vacancies on the CeO2 surface.

1 Introduction As a green chemical substitute, DMC is not only a very


important intermediate in organic synthesis, which is widely
The chemical industry is committed to develop or improve the used in carboxylation and methylation reactions, but also a fuel
hazardous traditional processes, responsible for the environ- additive owing to its non-toxicity.6–8 Before 1980, DMC was
mental problems such as emissions of greenhouse gases, parti- synthesized via the phosgenation of methanol; this method
cles, and toxic compounds, via more sustainable alternatives.1 involved the use of phosgene, a very toxic compound. Then,
Carbon dioxide is regarded as the main greenhouse gas leading several advanced methods such as transesterification route,
to global warming. Therefore, CO2 capture or chemical conver- oxidative carboxylation of methanol, and direct synthesis from
sion of CO2 has been considered as a promising method to methanol and CO2 have been applied to produce DMC. The
reduce CO2 levels.2 It is generally known that CO2 hydrogenation direct synthesis of DMC from methanol and CO2 is considered as
to formic acid or methanol is one of the effective ways to produce a promising method due to its non-toxicity and the abundance of
high-value chemicals. However, this process normally requires reactants as compared to the other routes.9
high temperatures and pressures to achieve a satisfactory CO2 Recently, homogeneous or heterogeneous catalysts have been
conversion.3,4 Another current research focus is to produce applied in this synthesis method. Among the reported catalysts,
organic carbonates from CO2,5 which might be classified into Ni(CH3COO)2, zirconia, CeO2, CeO2–ZrO2, CexTi1xO2, and
two types: cyclic carbonates (such as glycerol carbonate, butylene (H3PW12O40/CexTi1xO2) presented reasonable activity.10–13
carbonate, etc.) and linear carbonates (such as diethyl carbonate, Furthermore, ceria-based catalysts have attracted significant
dimethyl carbonate, etc.). Among them, DMC, the simplest attention due to their high catalytic activity and selectivity.14
organic carbonate, plays an important role in the chemical CeO2 possesses acid–base and redox properties that are favorable
industry due to its remarkable properties. for the activation of CO2 and methanol. Santos et al.15 reported
the kinetics for the direct synthesis of dimethyl carbonate over
CeO2 and developed a kinetic model for this reaction. It is
Key Laboratory of Coal Science and Technology, Ministry of Education and Shanxi
accepted that the addition of some metal oxide promoters to
Province, Taiyuan University of Technology, Taiyuan 030024, Shanxi, China. the CeO2 catalyst can greatly modify the support texture, regulate
E-mail: licongming0523@163.com, lizhong@tyut.edu.cn the particle size, and modulate the redox property to enhance

This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2017 New J. Chem., 2017, 41, 12231--12240 | 12231
View Article Online

Paper NJC

the DMC yield.16,17 Hence, ceria-based catalysts doped with and transformed to O2. The role of CaO in the CaO–CeO2
different metal oxides have been extensively studied. According to catalysts for the DMC synthesis has yet to be investigated in detail.
previous studies conducted by Hye Jin Lee et al.,18 gallium oxide Therefore, it is worthwhile to study the effect of CaO on the surface
supported on CexZr1xO2 catalysts can modify the acidity and properties of CaO–CeO2 catalysts and the catalytic effectiveness of
basicity of the support. The yield of DMC over XGa2O3/Ce0.6Zr0.4O2 these catalysts for DMC synthesis from CO2 and methanol.
catalysts was significantly improved by increasing both acidity In this study, a series of CeO2 catalysts promoted by CaO
Published on 30 August 2017. Downloaded by National Chemical Laboratory, Pune on 7/25/2023 11:53:42 AM.

and basicity of the catalysts. Furthermore, some researchers were synthesized with a mass fraction of Ca between 0 and
prepared Al2O3-modified CeO2 catalysts, which exhibited higher 4.9 wt% using a co-precipitation method, and their catalytic
stability as compared to CeO2 due to the addition of Al2O3 that performance for DMC synthesis from CO2 was evaluated. In
promoted the reduction of Ce(IV) to Ce(III).17 addition, the relationship between the physicochemical proper-
Generally, the addition of alkali and alkaline earth metals to ties and the catalytic activity was investigated in detail. The
catalysts can modify the catalyst structure and improve the focus of this study was to evaluate the surface oxygen vacancies
catalytic activity. Tamao Ishida and co-workers19 found that the of the catalysts as a function of the CaO content, which were
Na promoter led to a decrease in the particle size of Co nano- directly related to the resulting reaction performance.
particles and an increase in the surface basicity of the catalyst,
which was helpful for the catalytic performance in the Fischer–
Tropsch synthesis (FTS). Moreover, it has been reported that the 2 Experimental
introduction of Ca2+ into ZrO2 can effectively improve the thermal 2.1 Materials
stability of the CaO–ZrO2 catalyst.20 Similarly, the addition of Ca to
Cerium(III) nitrate hexahydrate (Ce(NO3)3, 99.9%, AR) and calcium
CeO2-based catalysts was shown to promote the production of
nitrate (Ca(NO3)2, 99.0%, AR) were purchased from Tianjin
biodiesel due to the favorable interactions between calcium and
Guang Fu Fine Chemical Research Institute (Tianjin, China),
cerium species in the solid catalyst.21,22 Kumar et al.23 reported that
and ammonia (NH3H2O, 25%, AR) was obtained from Feng
the CaO–CeO2 catalysts prepared by the surfactant-template method
Chuan Chemical Reagent Co., Ltd (Tianjin, China). All reagents
were used for the direct synthesis of DMC from CO2 and methanol.
were used as received without any further purification.
The author used sieve 3 A as a dehydrating agent, the Ce1–Ca1
catalyst exhibited the best activity, and the DMC yield reached
2.2 Catalyst preparation
2.96 mmol g1. Santos et al.24 also used sieve 3 A as a dehydrating
agent and studied the adsorption of DMC and water on a fixed-bed The CaO/CeO2 catalyst was prepared via a co-precipitation method.
column. The authors have found that to enhance the DMC yield, All mixed oxides were designed to obtain a CaxCe solid solution,
the removal of water from the reaction mixture is a crucial factor where x is the weight percentage of Ca in the catalysts. Typically, a
for the development of novel synthetic methods in this regard. 300 mL aqueous solution containing cerium nitrate and calcium
It is generally believed that oxygen vacancies play a critical nitrate with different Ca/Ce mass ratios and 100 mL ammonia
role by accommodating the oxygen generated from the cleavage solution of 1.0 M were mixed in a dropwise fashion in a 2 L beaker.
of CQO bonds25 and the oxide catalysts; thus, the presence of After the precipitation was completed, the precipitates were aged for
oxygen vacancies on the catalyst surface is expected to have a 2 h at room temperature. Subsequently, the precipitate was filtered
significant impact on the activation and conversion of CO2.26 It and washed several times with deionized water. The precipitate
has been reported that CeO2 and ceria-based catalysts have fine was washed until the conductivity of the filtrate was similar to
redox property and high oxygen storage capacity.27,28 However, that of deionized water. The washed solid was dried at 100 1C
there has been limited studies reporting the effect of oxygen overnight and then calcined at 700 1C for 3 h. Other catalysts
vacancies on the reaction performance. Liu and co-workers29 (Table 1) were prepared following a similar procedure.
have researched the soot oxidation over CeO2 and Ag/CeO2
catalysts and found that a suitable number of surface oxygen 2.3 Characterization
vacancies on the Ag/CeO2 catalyst results in a 10-fold increase 2.3.1 Temperature-programmed desorption of CO2/NH3
in the soot oxidation activity since O can be readily regenerated (CO2-TPD and NH3-TPD). Temperature-programmed desorption

Table 1 Acidity and basicity of the CaxCe catalysts

Basicitya (mmol g1) Acidityb (mmol g1)


Weak Moderate Strong Weak Moderate Strong
(o200 1C) (200–400 1C) (4400 1C) Total (o200 1C) (200–400 1C) (4400 1C) Total
Ca0Ce 0.185 0.134 0.007 0.326 0.188 0.069 0.026 0.283
Ca0.5Ce 0.195 0.144 0.010 0.349 0.194 0.075 0.032 0.301
Ca1.5Ce 0.221 0.160 0.017 0.398 0.219 0.083 0.048 0.350
Ca3.6Ce 0.116 0.028 0.024 0.168 0.072 0.055 0.038 0.165
Ca4.9Ce — 0.009 0.037 0.046 0.059 0.049 0.03 0.138
a b
Calculated by CO2-TPD. Calculated by NH3-TPD.

12232 | New J. Chem., 2017, 41, 12231--12240 This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2017
View Article Online

NJC Paper

(TPD) was carried out using a Micromeritics Autochem II 2920 to the analysis chamber for XPS study. The base pressure
analyzer. Typically, 60 mg sample was pretreated at 500 1C inside the analysis chamber was usually maintained at better
for 1 h under N2. After cooling, CO2 or NH3 adsorption was than 1010 Torr. The binding energies were referenced to the
performed by switching the He flow with a stream of 15 vol% C 1s line at 284.6 eV from adventitious carbon.
CO2 (NH3)–He (40 mL min1) and maintained at this tempera-
ture for 30 min. The sample was purged with He (40 mL min1) 2.4 Catalytic performance evaluation
Published on 30 August 2017. Downloaded by National Chemical Laboratory, Pune on 7/25/2023 11:53:42 AM.

for 2 h to remove the physically adsorbed CO2 (NH3). The TPD To evaluate the catalytic performance of the as-prepared cata-
test was carried out under a constant flow of He (40 mL min1) lysts, the DMC synthesis was conducted on a slurry bed reactor
from 50 1C to 500 1C at a rate of 10 1C min1. The concentration under continuous mechanical stirring; typically, the catalyst
of CO2 (NH3) in the tail gas was continuously measured using a (1 g) and CH3OH (35 mL) were placed inside the batch reactor,
TCD detector. and the mixture was purged with CO2 several times to remove
2.3.2 Powder X-ray diffraction (XRD). The powder X-ray diffrac- the air. Then, the reactor was purged with CO2 to a certain
tion patterns of the investigated catalysts were obtained using a pressure, and the reaction was carried out at 140 1C for 3 h. After
German Bruker D4 (40 kV, 30 mA) X-ray diffractometer equipped the reaction, the reactor was cooled down to room temperature,
with a Cu Ka radiation source (l = 0.15406 nm) and a nickel filter in and then, 1-propanol (CH3CH2CH2OH) was added as an internal
the 2y range of 201–851 at a scanning speed of 81 min1. standard substance. The products were further analyzed by a gas
2.3.3 Atomic absorption spectroscopy (AAS). Atomic absorp- chromatograph (FID-GC, O Hua 9160). It was noted that DMC
tion spectroscopy (AAS) was conducted to determine the actual Ca was the only product, and no other by-products were formed
loading of the as-prepared catalysts by a SpectrAA-220 AAS analyzer. during the reaction process.
Particularly, 25 mg of fresh catalyst was dissolved in 10–15 mL of The mass of DMC was calculated by the GC, and the catalyst
concentrated hydrochloric acid solution, and the mixture was activity was calculated by the following equation:
diluted with distilled water. Then, the solution was transferred to nðDMCÞ=mmol
a 250 mL volumetric flask and again diluted to 250 mL with Catalyst activity ¼
mðcatalystÞ=g
distilled water. After this, 5 mL of the abovementioned solution
was transferred to a 100 mL flask and diluted to 100 mL using
distilled water again (making sure that the Ca solution concen- 3 Results
tration was 1–5 mg L1 for the AAS analysis). Finally, 50 mL of this
3.1 Catalytic activity
solution was obtained to measure the Ca content.
2.3.4 N2 adsorption–desorption measurement. The N2 Fig. 1 shows the DMC yield over the pure CeO2 and the CaxCe
adsorption–desorption measurements were carried out via a samples for the synthesis of DMC from methanol and CO2.
Beishide 3H-2000PS2 apparatus using nitrogen at liquid- Obviously, it was observed that the DMC yield reached a
nitrogen temperature (77 K). Prior to the test, the sample was maximum value of 2.47 mmol g1 when the Ca content was
degassed in situ at 200 1C for 3 h. The specific surface areas of increased from 0 to 1.5 wt%, and thereafter, it decreased with
the samples were calculated by the Brunauer–Emmett–Teller the further increase in the Ca content. It was also noted that
(BET) method, and the pore size distribution was measured the DMC yield obtained over the Ca1.5Ce sample was 1.2 times
using the Barrett–Joyner–Halenda (BJH) model. higher than that obtained with the pure CeO2 sample.
2.3.5 Transmission electron microscopy (TEM). Transmission
3.2 CO2-TPD and NH3-TPD
electron microscopy (TEM) images were obtained using a JEOL
JEM-2100 transmission electron microscope operated at an CO2-TPD and NH3-TPD analysis were conducted to investigate
accelerating voltage of 200 kV. The samples for TEM analyses the effect of Ca content on the acid–base properties of the
were prepared by dispersing the powder products as a slurry in CaxCe catalysts. Fig. 2(a) shows the CO2-TPD profiles of the
ethanol and then depositing on a carbon film coated on a
copper grid.
2.3.6 Raman spectra. Raman spectra were acquired in the
scanning range of 100–1000 cm1 using a Renishaw inVia micro
laser Raman spectrometer (UK) and an Ar+ laser (514.5 nm
wavelength) with an output power of 4 mW.
2.3.7 X-ray photoelectron spectroscopy (XPS). X-ray photo-
electron spectroscopy (XPS) data were obtained via an
ESCALab220i-XL electron spectrometer (VG, UK) using 300 W
Al Ka radiation. Typically, the samples were compressed into
pellets of 2 mm thickness and then mounted on a sample
holder utilizing double-sided adhesive tape for XPS analysis.
The sample holder was then placed in a fast entry air load-lock
chamber without exposure to air and evacuated under vacuum Fig. 1 Catalytic activity of the CaxCe catalysts. Reaction conditions: catalyst
(o106 Torr) overnight. Finally, the sample holder was transferred mass: 1 g; temperature: 140 1C; pressure: 3 MPa; reaction time: 3 h.

This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2017 New J. Chem., 2017, 41, 12231--12240 | 12233
View Article Online

Paper NJC
Published on 30 August 2017. Downloaded by National Chemical Laboratory, Pune on 7/25/2023 11:53:42 AM.

Fig. 2 (a) CO2-TPD and (b) NH3-TPD profiles of the CaxCe catalysts.

CaxCe catalysts from 50 1C to 500 1C, and the total basicity of TCD signals, and this probably represented that the NH3
the CaxCe catalysts calculated from the CO2-TPD peak area is adsorption sites were covered by Ca2+. The acidity of the CaxCe
summarized in Table 1. It revealed that all the CaxCe catalysts, catalysts was calculated based on the NH3-TPD peak area
except the Ca4.9Ce sample, contained the same types of basic percentage, as listed in Table 1. It was obvious that the acidity
sites. The desorption peaks appearing in the lower temperature of the Ca1.5Ce catalyst is higher than that of other CaxCe catalysts
range below 200 1C can be assigned to the interaction between and the adsorption quantity of NH3 is 0.350 mmol g1. The
CO2 and weakly basic sites. The Ca1.5Ce sample exhibits highest total acidity of the catalysts decreased in the following order:
adsorption capacity, and the quantity of CO2 adsorbed is Ca1.5Ce 4 Ca0.5Ce 4 Ca0Ce 4 Ca3.6Ce 4 Ca4.9Ce. This trend is
0.398 mmol g1. Furthermore, it was found that the Ca1.5Ce in good agreement with the catalytic activity results. Moreover,
sample had the largest number of basic sites, and the total the trend for the adsorption strength (weak, moderate, and
basicity of the catalysts decreased in the following order: strong) is the same as the trend for the total acidity.
Ca1.5Ce 4 Ca0.5Ce 4 Ca0Ce 4 Ca3.6Ce 4 Ca4.9Ce; this trend
is in good agreement with the catalytic activity results. It 3.3 Textural properties of the Ca(x)Ce catalysts
was also found that the adsorption of CO2 increased with the The N2 adsorption/desorption isotherms and pore size distri-
increasing Ca content at the strong basic sites, which was butions of the CaxCe catalysts are presented in Fig. 3. As shown
confirmed by the previous literature.30 Furthermore, Ca1.5Ce in Fig. 3(a), a type IV isotherm with a clear H2 hysteresis loop is
showed a higher basicity than bulk CeO2; this may be ascribed obtained for all catalysts except the Ca4.9Ce catalysts, indicating
to the synergic effect between CaO and CeO2. Indeed, some typical mesoporous materials. In Fig. 3(b), the pore size dis-
authors have reported that doping another metal into a metal tribution calculated by the BJH method from the desorption
oxide can affect the textural and surface properties due to their branch of the isotherm revealed that all the catalysts, except the
synergic effect between the metal and the metal oxide.30–32 Ca4.9Ce catalyst, contained mesoporous pore distribution with
However, interestingly, the basicity of the catalysts decreased a pore size of 2–15 nm. In Table 2, it can be seen that the pore
with the increasing Ca content, which was consistent with the size of the Ca4.9Ce catalyst is higher than that of other catalysts;
literature.30 In Fig. 2(a), it was seen that the Ca3.6Ce and Ca4.9Ce however, the surface area of the Ca4.9Ce catalyst is lower
samples almost had no obvious adsorption peak. Thus, the than that of the other catalysts; this may be attributed to the
basicity of these two catalysts was even lower than that of pure formation of piled pores in the aggregated particles. The above-
CeO2. In the previous reports,33,34 CaO-based solid adsorbents mentioned results indicated that the Ca content had a signifi-
have been identified as the most promising candidates, which cant effect on the pore size distribution. The BET surface
are widely used for CO2 capture. The results indicated that the areas and pore volumes of the synthesized CaxCe catalysts are
introduction of an appropriate amount of Ca into CeO2 could summarized in Table 2. It can be seen that the Ca1.5Ce catalyst
increase the number of basic sites in the catalyst, whereas an possesses the highest specific surface area of 36 m2 g1 and pore
excessive amount of Ca could lead to a decrease in the amount volume of 0.14 cm3 g1, whereas Ca4.9Ce presents the lowest
of basic sites. The reduction in the amount of basic sites may be specific surface area of only 1 m2 g1 and pore volume of
due to the aggregation of CaO.35 0.035 cm3 g1. It is noted that there is a correlation between
Fig. 2(b) also shows the NH3-TPD profiles of the CaxCe the total adsorption capacity of CO2 and NH3 and the specific
catalysts. It was found that NH3-TPD and CO2-TPD results surface area. Thus, an increase in the specific surface area (from
exhibited the same trend of variation. This revealed that all 1 to 36 m2 g1) with a suitable content of Ca may provide more
CaxCe catalysts, except Ca4.9Ce, mainly consisted of same acidic available acidic and basic active sites on the catalyst surface. On
sites below 200 1C. The desorption peak for Ca4.9Ce sample can the other hand, an excessive amount of Ca led to a decrease of
hardly be seen in the temperature range below 200 1C; this the specific surface area and pore size, in turn decreasing the
indicates a weak acidity. The Ca4.9Ce catalyst had no obvious CO2 adsorption capacity, as confirmed by the CO2-TPD results.

12234 | New J. Chem., 2017, 41, 12231--12240 This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2017
View Article Online

NJC Paper
Published on 30 August 2017. Downloaded by National Chemical Laboratory, Pune on 7/25/2023 11:53:42 AM.

Fig. 3 (a) N2 adsorption/desorption isotherms and (b) pore size distributions of the CaxCe catalysts.

Table 2 Structure and physical parameters of various samples 3.4 TEM images
Ca loadinga SBET/ Vp Pore Particle Mean Fig. 5 shows the TEM images of the as-prepared catalysts. It was
Sample (wt%) m2 g1 (cm3 g1) size (nm) sizeb (nm) sizec (nm) noted that the particle size was in the range of 10–40 nm in all the
Ca0Ce — 23 0.074 8.0 22 23 cases. This value was in good agreement with the diameter range
Ca0.5Ce 0.5 26 0.11 10.0 20 21 calculated from the XRD data (15–23 nm). In addition, it can be
Ca1.5Ce 1.5 36 0.14 11.0 17 18 seen that with the increase of Ca contents, the particle size of the
Ca3.6Ce 3.6 8 0.049 8.7 16 18
Ca4.9Ce 4.9 1 0.035 25.3 14 15 catalysts decreased, in good line with the XRD results. The above-
a mentioned results indicated that the Ca content had a significant
Determined by AAS. b Calculated for about 100 nanoparticles from
the TEM images. c Estimated by the Scherrer equation, applied to the effect on the particle size of the catalysts, and the Ca1.5Ce catalyst
(111) reflection of fluorite CeO2. possessed the smaller mean size of about 17 nm. In Fig. 5, we can
also see that the Ca4.9Ce catalyst has the smaller mean size of about
14 nm, and the catalyst particles are severely agglomerated.
The XRD patterns of the CaxCe catalysts are presented in
Fig. 4. All the peaks of the samples can be indexed to the face-
centered cubic fluorite structure (JCPDS-ICDD 34-0394). The 3.5 Raman spectra analysis
peaks at 28.51, 33.01, 47.51, 56.51, 59.31, and 69.61 were ascribed The structures of the representative CaxCe samples were
to the (111), (200), (220), (311), (222), and (420) planes of CeO2, studied by Raman spectroscopy (Fig. 6). Ca0Ce exhibits a strong
according to PDF#34-0394.36 After the introduction of Ca F2g vibrational peak at 461 cm1 and a weak oxygen vacancy
species, no new diffraction peaks corresponding to the CaO defect-induced vibrational peak at 602 cm1, ref. 22. The addi-
phase (at around 381 and 541) were observed. This may be tion of Ca2+ species to ceria led to an increase of bandwidth of
attributed to the incorporation of Ca2+ species into the ceria the F2g mode; this suggested that there was a strong interaction
lattice and their high dispersion as a result of their small between CaO and CeO2. The results indicated the creation of
diameter. The mean crystallite size, as calculated by the Scherrer new oxygen vacancies via doping of the CeO2 catalyst with CaO.39
equation, was 23 nm for Ca0Ce, 21 nm for Ca0.5Ce, 18 nm for In addition, this peak exhibited a red shift as the content of CaO
Ca1.5Ce, 18 nm for Ca3.6Ce, and 15 nm for Ca4.9Ce (Table 2). increased; this indicated the incorporation of Ca2+ into the
It indicated that the decline of the crystallite size with the lattice of the resulting catalyst, which was in agreement with
increasing content of CaO could be related to the distortion of that observed via XRD analyses. Furthermore, a weak oxygen
the CeO2 structure due to the incorporation of CaO.37,38 vacancy defect-induced vibrational peak centered at 602 cm1
was observed. Because Ce ion has a valence of 4 or 3, and Ca ion
has only a valence of 2, some vacancies have been created during
the ion substitution process to maintain charge neutrality in
the ionic crystal (Ce4+ + O2 2 Ca2+ + VO, VO: oxygen vacancy).
These vacancies induced lattice distortions to form defects that
were favorable for heterogeneous catalysis. The intensity ratio
between the 602 (ID) and 464 cm1 (If2g) bands was used to
estimate the oxygen vacancy concentration. The ID/If2g values
rapidly increased as the Ca loading increased and followed the
trend Ca4.9Ce 4 Ca1.5Ce 4 Ca0Ce. These results suggested that a
larger amount of oxygen vacancies were created after the addition
of CaO.37,40 Moreover, the weaker band at 255 cm1 was observed
for Ca4.9Ce that belonged to the transverse acoustic (TA) mode.
Fig. 4 XRD patterns of the CaxCe catalysts. It implied that there existed some interaction between CaO and

This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2017 New J. Chem., 2017, 41, 12231--12240 | 12235
View Article Online

Paper NJC
Published on 30 August 2017. Downloaded by National Chemical Laboratory, Pune on 7/25/2023 11:53:42 AM.

Fig. 5 TEM images of the CaxCe catalysts.

CeO2 due to the more vacancies created by the substitution affected the synergistic interactions between the CaO and
of Ca2+ in CeO2.41,42 Therefore, the doped Ca significantly CeO2 nanostructures.

12236 | New J. Chem., 2017, 41, 12231--12240 This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2017
View Article Online

NJC Paper

As listed in Table 3, the percentages of Ce3+ were estimated


to be 19%, 24%, 27%, 17%, and 16% for Ca0Ce, Ca0.5Ce, Ca1.5Ce,
Ca3.6Ce, and Ca4.9Ce, respectively. This trend is exactly the
same as that for the N2 adsorption data. Ca4.9Ce presents the
lowest Ce3+ concentration, whereas Ca1.5Ce presents the highest
concentration. Combining the XPS results with the N2 adsorp-
Published on 30 August 2017. Downloaded by National Chemical Laboratory, Pune on 7/25/2023 11:53:42 AM.

tion data and Raman spectra, the fact that Ca4.9Ce had the
lowest Ce3+ concentration indicated that Ca2+ may be incorpo-
rated into the bulk of the CeO2 lattice due to the smallest BET
specific surface area of the CaO particles. The percentage of Ce3+
of the Ca1.5Ce sample with the highest BET specific surface area
reaches the maximum value of 27%. In general, the chemical
valence of Ce ion is 3 or 4, but Ca ion has only 2 valence; thus,
some vacancies are generated upon substitution to keep charge
neutrality in the ionic crystal, and these vacancies are favorable
for heterogeneous catalysis.21,31,44 Oxygen vacancy-associated
exposed ions on CeO2 are potentially potent surface sites in the
titled activation of CO2.45,46 Herein, combining the N2 adsorp-
tion data, Raman spectra, and XPS results, it was concluded
that the chemical states of the CaxCe samples were clearly
affected by the amount of CaO added, which also affected the
Fig. 6 Raman spectra of Ca0Ce0, Ca1.5Ce, and Ca4.9Ce. structural properties.
The O 1s spectra of the CaxCe samples clearly showed three
states of surface oxygen (Fig. 7, O 1s): lattice oxygen (OL), at a lower
3.6 XPS analysis binding energy (B529.0 eV); weakly bonded oxygen species and/or
Fig. 7 presents the XPS spectra of Ce 3d and O 1s for CaxCe chemisorbed oxygen (OC) (B531.0 eV); and surface oxygen (OS)
samples. The chemical states of the Ce species on the surface of the at the highest binding energy (B532.5 eV).47–49 The calculated
catalysts are obviously related to the concentration of CaO. The Ce OL concentration in the CaxCe catalysts was semi-quantitatively
3d spectra can be deconvoluted into eight peaks:43 V (B882.5 eV), estimated from the integrated peak area using the following
V0 (B884.1 eV), V00 (B888.4 eV), V0 0 0 (B898.0 eV), U (B899.8 eV), equation:
U0 (B901.6 eV), U00 (B907.5 eV), and U0 0 0 (B916.5 eV). The four U AOL
bands represent Ce 3d3/2, and the four V bands represent Ce 3d5/2. ½OL % ¼  100%
AOC þ AOL þ AOS
The 3d104f0 state of the Ce4+ species was labeled as U, U00 , U0 0 0 , V,
V00 , and V0 0 0 , whereas the 3d104f1 state of the Ce3+ was labeled as U0 AOL: photoelectron peaks areas of lattice oxygen; AOC: photo-
and U0 . The calculated Ce3+ concentration of the catalyst CaCe-x electron peaks areas of chemisorbed oxygen; and AOS: photo-
was semi-quantitatively analyzed using the integrated peak area of electron peaks areas of surface oxygen.
the respective valence states via the following equation As shown in Table 3, the ratio of OL to OT (OT = OL + OC + OS)
for all samples was calculated. It was observed that the incor-
 3þ  ACe3þ poration of Ca2+ can effectively increase the amount of lattice
Ce % ¼  100%
ACe3þ þ ACe4þ oxygen (OL) on the surface of the CaxCe catalysts. These results
confirmed that there was an enhanced mobility and availability
ACe3+: photoelectron peaks areas of Ce3+; ACe4+: photoelectron of lattice oxygen species due to the synergistic effect between
peaks areas of Ce4+. CaO and CeO2.50

Fig. 7 XPS spectra of Ce 3d (a) and O 1s (b) for the CaxCe catalysts ((a) Ca0Ce, (b) Ca0.5Ce, (c) Ca1.5Ce, (d) Ca3.6Ce, (e) and Ca4.9Ce).

This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2017 New J. Chem., 2017, 41, 12231--12240 | 12237
View Article Online

Paper NJC

Table 3 Concentrations of Ce3+ (%) and OL estimated by XPS form of CaCO3 and could not effectively react with CO2; this led to
a decline in catalytic activity.35 Consequently, a suitable degree of
Catalyst Ca0Ce Ca0.5Ce Ca1.5Ce Ca3.6Ce Ca4.9Ce
basicity and acidity is necessary for the catalysts to adsorb CO2
OL (%) 0.42 0.50 0.55 0.48 0.37 and CH3OH and ultimately promote the formation of DMC.
Ce3+ (%) 0.19 0.24 0.27 0.18 0.16

4.2 The role of surface oxygen vacancies


Published on 30 August 2017. Downloaded by National Chemical Laboratory, Pune on 7/25/2023 11:53:42 AM.

4 Discussion As seen in Fig. 9, the catalytic activity of the CaxCe catalysts


4.1 The role of basicity–acidity increased with the increasing concentration of Ce3+ on the surface.
Generally, both basic and acidic sites are necessary for the Obviously, the catalytic activity depended on the concentration of
direct synthesis of DMC from methanol and carbon dioxide.13 Ce3+ on the catalyst surface. Among all catalysts, the Ca1.5Ce
Therefore, the total basicity and acidity of the CaxCe catalysts catalyst, with the highest concentration of Ce3+ on the surface,
determined the yield of DMC. As seen in Fig. 8(a), we observed exhibited the best catalytic performance. This indicated that more
that changes from weak to moderate basicity were consistent oxygen vacancies are beneficial for this reaction. According to the
with the activity of the catalysts; however, the changes from previous DFT studies, CO2 could be activated in the oxygen
moderate to strong basicity were not. The improved weak and vacancies on the defective surface of the metal oxide.57 However,
moderate basicity could be ascribed to the synergic effect on perfect metal surfaces, CO2 was very weakly adsorbed and
between CaO and CeO2. The strongly basic sites were caused activated via the Eley–Rideal mechanism.57–60 It was also shown
by the introduction of Ca2+ into CeO2. However, a higher that increasing the content of oxygen vacancies on the CeO2
number of strongly basic sites did not favor this reaction, surface could enhance the adsorption of CO2 and improve the
which was consistent with a previous report.51 As shown in interactions between CO2 and CeO2.61,62 As is well known,
Fig. 8(b), it was observed that more acidic sites were favorable
for increasing the DMC yield. The acidity of the catalysts
increased as the content of Ca increased, whereas the acidity
decreased at an excessive amount of Ca. Calcium oxide is an
alkaline oxide and therefore, an excessive amount of Ca in the
catalysts may result in a decrease of ammonia adsorption.
Ca1.5Ce showed the best DMC yield than other catalysts
due to its higher CO2 and NH3 adsorption capacity. Based
on previous literature reports,52–54 methanol was activated to
methyl and methoxy species at the acidic and basic sites of the
catalyst, and then, the methoxy species reacted with CO2 on the
basic sites of the catalyst to form the methoxy carbonate anion,
which finally reacted with the methyl species to generate DMC.
The adsorption at weakly and moderately basic sites was
enhanced due to the synergistic effect between CaO and
CeO2. However, at strongly basic sites, CO2 could interact with Fig. 9 Correlation between the concentration of Ce3+ (%) and catalytic
CaO to generate calcium carbonate due to the strong basicity of activity. Reaction conditions: catalyst mass of 1 g, temperature of 140 1C,
the catalyst.55,56 Therefore, some active CaO still existed in the pressure of 3 MPa, and reaction time of 3 h.

Fig. 8 Correlation between (a) basicity, (b) acidity, and catalytic activity of catalysts. Reaction conditions: catalyst mass of 1 g, temperature of 140 1C,
pressure of 3 MPa, and reaction time of 3 h.

12238 | New J. Chem., 2017, 41, 12231--12240 This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2017
View Article Online

NJC Paper

vacancies favored CO2 adsorption and improved the catalytic


performance for the synthesis of DMC from CO2 and methanol.
Under the catalytic reaction conditions of 140 1C, 3 MPa, and 2 h
reaction time, the highest DMC yield of 2.47 mmol g1 could be
obtained over the Ca1.5Ce catalyst. Based on the obtained results,
a reaction route for DMC synthesis from CO2 and methanol over
Published on 30 August 2017. Downloaded by National Chemical Laboratory, Pune on 7/25/2023 11:53:42 AM.

the CaxCe catalysts was proposed.

Conflicts of interest
There are no conflicts to declare.

Acknowledgements
Fig. 10 Putative mechanism for the formation of DMC from CO2 and This work was supported by the National Natural Science
methanol at the surface oxygen vacancies over the CaxCe catalysts. Foundation of China (grants U1510203, 21676176), the National
Science Foundation of the Shanxi Province, China (grant
201601D011016), and the fund of State Key Laboratory of Catalysis
oxygen vacancies are Lewis sites and therefore, a higher number
in DICP (grant N-15-05).
of oxygen vacancies favors CO2 adsorption. Thus, a high concen-
tration of Ce3+ helped to produce more oxygen vacancies, which
in turn favored the activation of CO2, thus improving the overall References
reaction performance.
1 B. A. V. Santos, V. M. T. M. Silva, J. M. Loureiro and
4.3 Proposed mechanism for DMC formation over the A. E. Rodrigues, Chem. Biochem. Eng. Rev., 2014, 1, 214–229.
surfaces of CaxCe catalysts 2 S. Wang, S. Fan, Y. Zhao, L. Fan, S. Liu and X. Ma, Ind. Eng.
Chem. Res., 2014, 53, 10457–10464.
On the basis of literature reports and our results, synthesis
3 S. Kattel, B. Yan, Y. Yang, J. G. Chen and P. Liu, J. Am.
mechanism of DMC at the surface oxygen vacancies over the
Chem. Soc., 2016, 138, 12440–12450.
CaxCe catalysts was proposed. As shown in Fig. 10, many oxygen
4 J. Wei, Q. Ge, R. Yao, Z. Wen, C. Fang, L. Guo, H. Xu and
vacancies on the surface of CaxCe samples and metal ions (Ce3+
J. Sun, Nat. Commun., 2017, 8, 15174.
and Ca2+) acted as Lewis acid sites.63 First, CO2 molecules were
5 M. Bersani, K. Gupta, A. K. Mishra, R. Lanza, S. F. R. Taylor,
adsorbed onto oxygen vacancies through the interaction of oxygen
H.-U. Islam, N. Hollingsworth, C. Hardacre, N. H. de Leeuw
atoms of CO2 with the oxygen vacancies. Subsequently, methanol
and J. A. Darr, ACS Catalysis, 2016, 6, 5823–5833.
molecules were adsorbed at a metal ion site adjacent to the oxygen
6 J. Bian, M. Xiao, S. Wang, Y. Lu and Y. Meng, Catal.
vacancy through the interactions of oxygen atoms with the metal
Commun., 2009, 10, 1142–1145.
atom because metal ions constituted Lewis acid sites that could
7 J. Bian, M. Xiao, S. J. Wang, Y. X. Lu and Y. Z. Meng, Chin.
interact with the oxygen atom of the methanol molecules;
Chem. Lett., 2009, 20, 352–355.
afterwards, another methanol molecule was adsorbed at
8 Z.-F. Zhang, Z.-W. Liu, J. Lu and Z.-T. Liu, Ind. Eng. Chem.
another metal ion adjacent to an oxygen vacancy to form an
Res., 2011, 50, 1981–1988.
intermediate. The O atoms in carbon dioxide molecule as Lewis
9 R. Saada, S. Kellici, T. Heil, D. Morgan and B. Saha,
base sites subtract H atom from methanol molecules to form
Appl. Catal., B, 2015, 168-169, 353–362.
water and an intermediate. Finally, after the surface reaction
10 M. H. Y. Kyung Won La and J. S. Chung, Solid State Phenom.,
was completed, the intermediate converted into DMC, whereas
2007, 119, 287–290.
the oxygen vacancy was regenerated.
11 J. C. J. Kyung Won La, H. Kim, S.-H. Baeck and I. K. Song,
J. Mol. Catal. A: Chem., 2007, 41–45.
5 Conclusions 12 Z.-T. L. Zhi-Fang Zhang, Z.-W. Liu and J. Lu, Catal. Lett.,
2008, 129, 428–436.
In this study, a series of CaxCe catalysts with different contents 13 T. S. Keiichi Tomishige, Y. Ikeda and K. Fujimoto, Catal.
of Ca were prepared by a co-precipitation method. The addition Lett., 1999, 58, 225–229.
of CaO to the CeO2 catalyst had a significant impact on the 14 M. Honda, M. Tamura, Y. Nakagawa, K. Nakao, K. Suzuki
acid–base properties and amounts of oxygen vacancies in the and K. Tomishige, J. Catal., 2014, 318, 95–107.
catalyst, as demonstrated by CO2-TPD, NH3-TPD and XPS. 15 C. S. M. P. B. A. V. Santos, V. M. T. M. Silva, J. M. Loureiro
The Raman results also demonstrated that the interactions and A. E. Rodrigues, Appl. Catal., A, 2013, 219–226.
between CaO and CeO2 generated more oxygen vacancies. 16 Y. F. Keiichi Tomishige, Y. Iked, M. Asadullah and K. Fujimoto,
Furthermore, it was observed that a higher number of oxygen Catal. Lett., 2001, 76, 1–2.

This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2017 New J. Chem., 2017, 41, 12231--12240 | 12239
View Article Online

Paper NJC

17 A. D. Michele Aresta and C. Pastore, J. Catal., 2010, 269, 40 K. Gong, H.-L. Wang, D. Fang and Z.-L. Liu, Catal. Commun.,
44–52. 2008, 9, 650–653.
18 H. J. Lee, S. Park, I. K. Song and J. C. Jung, Catal. Lett., 2011, 41 T. Berger, J. Schuh, M. Sterrer, O. Diwald and E. Knozinger,
141, 531–537. J. Catal., 2007, 247, 61–67.
19 T. Ishida, T. Yanagihara, X. Liu, H. Ohashi, A. Hamasaki, 42 Z. Wen, X. Yu, S. T. Tu, J. Yan and E. Dahlquist, Bioresour.
T. Honma, H. Oji, T. Yokoyama and M. Tokunaga, Appl. Technol., 2010, 101, 9570–9576.
Published on 30 August 2017. Downloaded by National Chemical Laboratory, Pune on 7/25/2023 11:53:42 AM.

Catal., A, 2013, 458, 145–154. 43 E. Bêche, P. Charvin, D. Perarnau, S. Abanades and


20 S. Pradhan, V. Sahu and B. G. Mishra, J. Mol. Catal. A: G. Flamant, Surf. Interface Anal., 2008, 40, 264–267.
Chem., 2016, 425, 297–309. 44 Y. C. Wong, Y. P. Tan, Y. H. Taufiq-Yap, I. Ramli and
21 W. Thitsartarn and S. Kawi, Green Chem., 2011, 13, 3423. H. S. Tee, Fuel, 2015, 162, 288–293.
22 X. Yu, Z. Wen, H. Li, S.-T. Tu and J. Yan, Fuel, 2011, 90, 45 S. Sharma, K. B. Sravan Kumar, Y. M. Chandnani, V. S.
1868–1874. Phani Kumar, B. P. Gangwar, A. Singhal and P. A.
23 P. Kumar, V. C. Srivastava, R. Gläser, P. With and I. M. Mishra, Deshpande, J. Phys. Chem. C, 2016, 120, 14101–14112.
Powder Technol., 2017, 309, 13–21. 46 S. Wang, L. Zhao, W. Wang, Y. Zhao, G. Zhang, X. Ma and
24 B. A. V. Santos, V. M. T. M. Silva, J. M. Loureiro and J. Gong, Nanoscale, 2013, 5, 5582–5588.
A. E. Rodrigues, Ind. Eng. Chem. Res., 2014, 53, 2473–2483. 47 Y. Zheng, K. Li, H. Wang, D. Tian, Y. Wang, X. Zhu, Y. Wei,
25 M. D. Porosoff and J. G. Chen, J. Catal., 2013, 301, 30–37. M. Zheng and Y. Luo, Appl. Catal., B, 2017, 202, 51–63.
26 W. Wang, Y. Zhang, Z. Wang, J.-m. Yan, Q. Ge and C.-j. Liu, 48 K. N. R. Benjaram, M. Reddy and P. Bharali, Ind. Eng. Chem.
Catal. Today, 2016, 259, 402–408. Res., 2009, 48, 8478–8486.
27 S. Chang, M. Li, Q. Hua, L. Zhang, Y. Ma, B. Ye and 49 J. wiatowska, V. Lair, C. Pereira-Nabais, G. Cote, P. Marcus
W. Huang, J. Catal., 2012, 293, 195–204. and A. Chagnes, Appl. Surf. Sci., 2011, 257, 9110–9119.
28 S. Carrettin, P. Concepcion, A. Corma, J. M. Lopez Nieto and 50 Y. Wu, Y. Zhang, M. Liu and Z. Ma, Catal. Today, 2010, 153,
V. F. Puntes, Angew. Chem., Int. Ed., 2004, 43, 2538–2540. 170–175.
29 S. Liu, X. Wu, W. Liu, W. Chen, R. Ran, M. Li and D. Weng, 51 O. Arbeláez, A. Orrego, F. Bustamante and A. L. Villa, Catal.
J. Catal., 2016, 337, 188–198. Lett., 2016, 146, 725–733.
30 S. Liu, J. Ma, L. Guan, J. Li, W. Wei and Y. Sun, Microporous 52 T. S. Keiichi Tomishige, Y. Ikeda and K. Fujimoto, Catal.
Mesoporous Mater., 2009, 117, 466–471. Lett., 1999, 58, 225–229.
31 Y. H. Taufiq-Yap, S. H. Teo, U. Rashid, A. Islam, 53 K. Tomishige, Y. Ikeda, T. Sakaihori and K. Fujimoto,
M. Z. Hussien and K. T. Lee, Energy Convers. Manage., J. Catal., 2000, 192, 355–362.
2014, 88, 1290–1296. 54 K. T. Jung and A. T. Bell, J. Catal., 2001, 204, 339–347.
32 S. Yan, M. Kim, S. O. Salley and K. Y. S. Ng, Appl. Catal., A, 55 D. A. J. Carlos Abanades, Energy Fuels, 2003, 17, 308–315.
2009, 360, 163–170. 56 R. Barker, J. Appl. Chem. Biotechnol., 1973, 23, 133–142.
33 S. Wang, H. Shen, S. Fan, Y. Zhao, X. Ma and J. Gong, Chem. 57 J. Ye, C. Liu, D. Mei and Q. Ge, ACS Catalysis, 2013, 3,
Eng. Sci., 2015, 135, 532–539. 1296–1306.
34 H. R. Radfarnia and A. Sayari, Chem. Eng. J., 2015, 262, 58 Y.-F. Zhao, Y. Yang, C. Mims, C. H. F. Peden, J. Li and
913–920. D. Mei, J. Catal., 2011, 281, 199–211.
35 S. F. Shengping Wang, L. Fan, Y. Zhao and X. Ma, Environ. 59 Y. Yang, J. Evans, J. A. Rodriguez, M. G. White and P. Liu,
Sci. Technol., 2015, 49, 5021–5027. Phys. Chem. Chem. Phys., 2010, 12, 9909–9917.
36 W. Huang, J. Yang, C. Wang, B. Zou, X. Meng, Y. Wang, 60 Y. X. Pan, C. J. Liu, D. Mei and Q. Ge, Langmuir, 2010, 26,
X. Cao and Z. Wang, Mater. Res. Bull., 2012, 47, 2349–2356. 5551–5558.
37 J. A. Rodriguez, X. Wang, J. C. Hanson, G. Liu, A. Iglesias-Juez 61 F. Wang, C. Li, X. Zhang, M. Wei, D. G. Evans and X. Duan,
and M. Fernández-Garcı́a, J. Chem. Phys., 2003, 119, 5659. J. Catal., 2015, 329, 177–186.
38 D. K. P. Satish Samantaray, G. Hota and B. G. Mishra, Chem. 62 N. Rui, Z. Wang, K. Sun, J. Ye, Q. Ge and C.-J. Liu, Appl.
Eng. J., 2012, 193–194, 1–9. Catal., B, 2017, 218, 488–497.
39 Z. Wu, M. Li, J. Howe, H. M. Meyer, 3rd and S. H. Overbury, 63 C.-L. Chiang, K.-S. Lin, S.-H. Yu and Y.-G. Lin, Int.
Langmuir, 2010, 26, 16595–16606. J. Hydrogen Energy, 2017, 1–15.

12240 | New J. Chem., 2017, 41, 12231--12240 This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2017

You might also like