You are on page 1of 8

Research Article

Cite This: ACS Catal. 2018, 8, 571−578 pubs.acs.org/acscatalysis

Direct Production of Lower Olefins from CO2 Conversion via


Bifunctional Catalysis
Peng Gao,† Shanshan Dang,†,‡ Shenggang Li,†,§ Xianni Bu,† Ziyu Liu,† Minghuang Qiu,†
Chengguang Yang,† Hui Wang,† Liangshu Zhong,*,† Yong Han,§,∥ Qiang Liu,§,∥ Wei Wei,†,§
and Yuhan Sun*,†,§

CAS Key Laboratory of Low-Carbon Conversion Science and Engineering, Shanghai Advanced Research Institute, Chinese Academy
of Sciences, Shanghai 201203, P. R. China

University of the Chinese Academy of Sciences, Beijing 100049, P. R. China
Downloaded via CHALMERS UNIV OF TECHNOLOGY on September 18, 2020 at 18:54:40 (UTC).
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

§
School of Physical Science and Technology, ShanghaiTech University, Shanghai 201203, P. R. China

State Key Laboratory of Functional Materials for Informatics, Shanghai Institute of Microsystem and Information Technology,
Chinese Academy of Sciences, Shanghai 200050, P. R. China
*
S Supporting Information

ABSTRACT: Direct conversion of carbon dioxide (CO2) into lower olefins


(C2=−C4=), generally referring to ethylene, propylene, and butylene, is highly
attractive as a sustainable production route for its great significance in
greenhouse gas control and fossil fuel substitution, but such a route always
tends to be low in selectivity toward olefins. Here we present a bifunctional
catalysis process that offers C2=−C4= selectivity as high as 80% and C2−C4
selectivity around 93% at more than 35% CO2 conversion. This is achieved by
a bifunctional catalyst composed of indium−zirconium composite oxide and
SAPO-34 zeolite, which is responsible for CO2 activation and selective C−C
coupling, respectively. We demonstrate that both the precise control of oxygen
vacancies on the oxide surface and the integration manner of the components
are crucial in the direct production of lower olefins from CO2 hydrogenation.
No obvious deactivation is observed over 150 h, indicating a promising potential for industrial application.
KEYWORDS: CO2 hydrogenation, lower olefins, bifunctional catalysts, C−C coupling, heterogeneous catalysis

1. INTRODUCTION (ASF) distribution, which is characterized by a maximum of


Carbon dioxide (CO2) is an easily available renewable carbon C2 to C4 hydrocarbon (C2−C4) fraction including olefins
resource with the advantages of being economical, nontoxic, (C2=−C4=) and paraffins (C2o−C4o) of about 56.7% and an
and abundant.1 The utilization of CO2 as a feedstock for undesired methane (CH4) fraction of about 29.2%.16−18
producing various chemicals not only contributes to alleviating Moreover, the low CO2 adsorption rate on the surface due to
global climate changes caused by increasing CO2 emissions but the thermodynamical and chemical stability of CO2 molecules
also offers a sustainable solution to replacing dwindling fossil leads to a high H/C ratio on the catalyst surface during CO2
fuel reserves.2−5 One promising route is the selective hydrogenation.12,19 This favors the hydrogenation of surface-
conversion of CO2 into value-added chemicals, such as lower adsorbed intermediates, resulting in the ready formation of
olefins (C2=−C4=, generally referring to ethylene, propylene, methane with a decrease in chain growth. Iron-based catalyst,
and butylene), which are the key building blocks of the conventionally used for commercial FTS, was also extensively
chemical industry and traditionally produced by thermal studied for the production of olefins from CO2 hydrogenation,
cracking of naphtha.6−8 However, few studies focused on the and a C2=−C4= selectivity of 35−50% was achieved with CH4
selective hydrogenation of CO2 to the product containing two selectivity above 16%.20−22 Thus, it remains a grand challenge
or more carbon atoms (C2+) due to the extreme inertness of to simultaneously achieve high selectivity for lower olefins and
CO2 and a high C−C coupling barrier.9−11 low selectivity for CH4.
Generally, the direct CO2 hydrogenation to hydrocarbons The indirect route of production of lower olefins from CO2
proceeds via a modified Fischer−Tropsch synthesis (FTS) includes conversion of CO2 into methanol (CH3OH) and
process, which consists of two main consecutive reactions: olefins production from CH3OH in a separate stage.13 Different
reverse water gas shift (RWGS) reaction to produce CO
followed by the further conversion of CO to hydrocarbons via Received: August 8, 2017
Fischer−Tropsch reaction.12−15 For FTS, the produced Revised: December 1, 2017
hydrocarbons usually follow the Anderson−Schulz−Flory Published: December 5, 2017

© 2017 American Chemical Society 571 DOI: 10.1021/acscatal.7b02649


ACS Catal. 2018, 8, 571−578
ACS Catalysis Research Article

from industrial methanol synthesis from syngas, the formation with Cu Ka radiation (40 kV, 40 mA). The intensity data were
of water vapor is inevitable for CO2 hydrogenation to collected over a 2θ range of 5−90° and scanning step length of
methanol, which inhibits the reaction strongly and leads to 0.0167°. For in situ XRD measurements, the sample first
serious catalyst deactivation,23,24 and thus an efficient catalyst is remained in pure Ar at a flow rate of 60 mL min−1.
in need to improve catalytic stability. For conversion of Temperature-ramping programs were exhibited from 30 to
methanol to olefins (MTO), the SAPO-34 zeolite is recognized 400 °C at a heating rate of 10 °C min−1 and maintained at 400
as the best catalysts owing to its unique topology, while it °C for 1 h. Then the gas flow was switched to the reactant gas
undergoes quick deactivation as a result of coke deposi- H2/CO2/N2 (73/24/3) mixture, and the temperature was
tion.25−27 Therefore, the selective formation of olefins from maintained at 400 °C.
CH3OH with high stability is still challenging. Moreover, as The textural properties such as surface area (BET),
compared with the indirect route, the direct conversion of CO2 micropore area (t-plot method), pore volume (BJH and HK),
into olefins would be more economic and energy-efficient. and pore size distribution (BJH) of the samples were derived
Here, we report an alternative process for the direct from N2 adsorption−desorption measurements carried out at
production of lower olefins from CO2 hydrogenation via 77 K using a TriStar II 3020 instrument. Prior to the
bifunctional catalysis. The bifunctional catalysis has resulted in measurements, the samples were treated in vacuum at 473 K for
a significant breakthrough in the synthesis of gasoline fuels 10 h.
from CO2 hydrogenation28 and selective conversion of syngas The chemical composition of the zeolites was determined
to lower olefins.7,29 We design a high efficient bifunctional with the use of an X-ray fluorescence (XRF) spectroscopy
catalyst composed of indium−zirconium composite oxides (Rigaku ZSX Primus II, Japan) by SQX calculation.
(In−Zr oxide) and SAPO-34 zeolites to convert CO2 directly The morphology of the samples was observed by
into lower olefins. As a result, a C2=−C4= selectivity reached up SUPRRATM 55 scanning electron microscopy (SEM) with
to 80% in hydrocarbons with only about 4% methane. The an accelerating voltage of 2.0 kV. The nanostructure of the
product distribution was completely different from that catalysts was investigated by a Tecnai G2 20 S-Twin high-
obtained via FTS and deviated greatly from the classical ASF resolution transmission electron microscope (HRTEM) and
distribution. In addition, there was no obvious deactivation over the TEM operated at 200 kV. The energy X-ray dispersive
150 h, suggesting the promising potential for industrial spectroscopic (EDX) analyses have been performed by using a
applications. Li−Si EDS detector with an energy resolution of 0.05 eV.
In situ near ambient pressure X-ray photoelectron spectros-
2. EXPERIMENTAL SECTION copy (NAP−XPS) was performed on a system which is
2.1. Catalyst Preparation. Various oxides were prepared manufactured by SPECS Surface Nano Analysis GmbH. The
by a coprecipitation method. Typically, 3.25 g of In(NO3)3· facility is composed of two chambers, an analysis chamber and a
4.5H2O and 14.63 g of Zr(NO3)3·5H2O were dissolved in a quick ample load-lock chamber. The analysis chamber is
mixture of 48 mL of deionized water and 140 mL of ethanol, equipped with a PHOIBOS NAP hemispherical electron energy
followed by the addition of a mixture of 36 mL of NH4OH analyzer, a microfocus monochromatized Al Kα X-ray source
(28−30 wt % in H2O) and 108 mL of ethanol. The product with beam size of 300 μm, a SPECS IQE-11A ion gun, and an
was aged at 80 °C for 10 min and then filtered and washed with infrared laser heater.
deionized water. The filter cakes were dried overnight at 60 °C CO2 temperature-programmed desorption (CO2−TPD)
and calcined in air at 500 °C for 5 h. Other oxides were experiments were carried out with a Micromeritics Chemi
synthesized by similar procedures. Sorb 2920. First, the catalyst (0.1 g) was pretreated at 400 °C
SAPO-34 zeolites were prepared by a hydrothermal route. for 60 min in a flow of pure Ar of 60 mL min−1 and then cooled
Parent SAPO-34 crystals were prepared by a hydrothermal to 50 °C. After that, the catalyst was saturated in flowing CO2
route. The synthesis gel recipe in molar composition is 1 for 1 h with 30 mL min−1 and followed by flushing in Ar for 1 h
Al2O3:0.44 SiO2: 1.1 P2O5:2.25 triethyl-amine: 35 H2O, and to remove any physisorbed molecules. The CO2−TPD
SAPO-34 crystal seeds with a mass ratio of 1:500 to the gel measurement was then carried out at 50−750 °C and a heating
were mixed in a closed autoclave. Then the mixture was heated rate of 10 °C min−1 under continuous flow of Ar with 40 mL
from room temperature to 165 °C in 7 h and kept at 165 °C for min−1.
33 h before cooling. The solid product was filtered, washed, and In situ diffuse reflectance infrared Fourier transform
dried, followed by calcination at 600 °C for 5 h. spectroscopy (in situ DRIFTS) measurements were performed
For the preparation of the In2O3/SAPO-34 and In−Zr/ in a quartz cell having a cylindrical cavity (5 mm in diameter
SAPO-34 composite catalysts, the In2O3 or In−Zr oxides and and 5 mm vertical length) for the sample placement.
the SAPO-34 were pressed, crushed, and sieved to granules in Approximately, 20 mg of catalyst powder was placed in the
the range of 40−60 mesh (granule sizes of 250−400 μm), cell and pretreated at 400 °C for 1 h under continuous flow of
respectively. Then, the granules of the two samples were mixed He with 40 mL min−1) and then cooled to 30 °C. After that,
together by shaking in a vessel. The samples prepared by this the catalyst was saturated in flowing CO2 for 2 h with 30 mL
method were denoted as In2O3/SAPO-34-G and In−Zr/ min−1 and followed by flushing in He for 30 min to remove any
SAPO-34-G. For comparison, the In2O3 or In−Zr oxides and physisorbed molecules. After that, temperature-ramping pro-
the SAPO-34 were mixed in an agate mortar for 10 min. Then, grams were exhibited from 30 to 100, 150, 200, 250, 300, 350,
the mixed samples were pressed, crushed, and sieved to and 400 °C with the He gas flow of 30 mL min−1. The spectra
particles in the range of 40−60 mesh. The obtained catalysts were collected on a Fourier-transform infrared spectroscopy
were denoted as In2O3/SAPO-34-M and In−Zr/SAPO-34-M. (Thermo Scientific, Nicolet 6700).
2.2. Catalyst Characterization. The crystalline and phase NH3 temperature-programmed desorption (NH3−TPD)
of the samples were investigated by powder X-ray diffraction experiments were performed in a similar manner as CO2−
(XRD) performed on a Rigatku Ultima 4 X-ray diffractometer TPD. After pretreatment of 0.2 g of catalyst in Ar at 400 °C for
572 DOI: 10.1021/acscatal.7b02649
ACS Catal. 2018, 8, 571−578
ACS Catalysis Research Article

1 h, the sample was cooled to 100 °C and brought to saturation where CnHm outlet represents moles C of individual hydro-
with ammonia using NH3 flow at 30 mL min−1. Following the carbon product at the outlet.
ammonia saturation, the system was purged with Ar flow at 100 2.4. DFT Calculations. Periodic density functional theory
°C for 30 min to remove any gas phase ammonia in the system (DFT) calculations were carried out with the Vienna ab initio
and unadsorbed ammonia trapped in the catalyst bed. For simulation package (VASP) using the Perdew−Burke−
desorption analysis, the catalyst bed temperature was raised Ernzerhof (PBE) exchange-correlation functional and projector
from 100 to 600 °C at 10 °C min−1. augmented wave (PAW) potentials.
29
Si, 27Al, and 31P magic-angle spinning (MAS) nuclear
magnetic resonance (NMR) spectra were recorded on a Bruker 3. RESULTS AND DISCUSSION
Avance III 400 WB spectrometer equipped with a 4 mm 3.1. Catalytic Functionality of Oxides and Zeolites.
standard bore CP MAS probehead whose X channel was tuned The bifunctional catalyst contained an In−Zr oxide with an In/
to 79.50, 104.27, and 162 MHz for 29Si, 27Al, and 31P, Zr molar ratio of 1:4 (Table S1 in the Supporting Information)
respectively, using a magnetic field of 9.39T at 297 K. The dried and a SAPO-34 zeolite with Si/(Si+Al+P) molar ratio of 0.062
and finely powdered samples were packed in the ZrO2 rotor and a pore size around 0.4 nm (Table S2 and Figure S1 in the
closed with a Kel-F cap which were spun at 8 or 12 kHz rate. Supporting Information). It was confirmed that the In−Zr
All 29Si, 27Al, and 31P MAS NMR chemical shifts are referenced oxide mainly contained In1−xZrxOy and ZrO2 nanoparticles
to the resonances of 3-(trimethylsilyl)-1-propanesulfonic acid from XRD and TEM characterizations (Figure S2 in the
sodium salt (DSS), [Al(H2O)6]3+ and monoammonium Supporting Information and Figure 1a,b). First, the catalytic
phosphate (NH4H2PO4) standards (d = 0.0), respectively. performance for the sole In−Zr oxide or In2O 3 was
2.3. Catalytic Evaluation. Activity measurements in the investigated. Both In−Zr and In2O3 samples exhibited much
hydrogenation of CO2 were carried out in a continuous-flow, higher activity for CO2 hydrogenation, and CO2 conversion was
high-pressure, fixed-bed reactor. Typically, 1.0 g of composite more than 30% at the reaction temperature of 400 °C, whereas
catalyst (40−60 mesh) was placed in a stainless steel tube usage of ZrO2 alone showed only 7.3% of CO2 conversion
reactor (inner diameter, 12 mm). Prior to reaction, the catalyst (Table S3 in the Supporting Information). In addition, the In−
was pretreated at 400 °C for 1 h in pure Ar (150 mL min−1). Zr oxide displayed an In-based reaction rate (i.e., the number of
Then, the reactant gas mixture with a H2/CO2/N2 ratio of 73/ mol of CO2 converted per gram of indium per second)
24/3 and a pressure of 3.0 MPa was introduced into the approximately five times that of the bulk In2O3, suggesting that
reactor. The catalytic reaction for methanol conversion was the incorporation of Zr enhanced the catalytic activity
performed in the same reactor. After that, 1.0 g of zeolite (40− significantly (Figure 2).
60 mesh) was pretreated in the reactor in pure Ar (150 mL Density functional theory (DFT) calculations revealed that
min−1) at 400 °C for 1 h. In2O3 was a unique catalyst in CO2 activation and hydro-
Methanol was then pumped into the reactor with an Ar or H2 genation to methanol with its surface oxygen vacancy and that
(150 mL min−1) under 0.1−3.0 MPa and 400 °C. The space the reaction followed a mechanism comprising the cyclic
velocity of liquid methanol was set to 0.18 mL gzeolite−1 h−1, creation and annihilation of oxygen vacancies.30,31 Surface Zr
equivalent to the yield of methanol from CO2 hydrogenation dopants were predicted to prefer to substitute the In4a and In4b
over the composite catalyst with In−Zr/SAPO-34 mass ratio of atoms (Figure S3a−d in the Supporting Information). Addi-
2. tionally, based on previous DFT studies, oxygen vacancy at the
The effluents were quantitatively analyzed online with a O3a (D3) or O4a (D4) site may be able to catalyze the CO2
Shimadzu GC-2010C gas chromatograph equipped with hydrogenation reaction.28,30 CO2 chemisorption energies at the
thermal conductivity and flame ionization detectors. The different surface oxygen vacancy sites (D3 and D4) on the pure
catalytic performance after 48 h of reaction was typically used In2O3 and Zr-doped In1−xZrxOy surfaces are shown in Figure
for discussion. S3e,f in the Supporting Information and Figure 3a,b,
CO2 conversion was calculated on a carbon atom basis respectively. For the defect surface, CO2 chemisorption at the
according to the following equation: D3 and D4 sites resulted in the formation of a metal−carbon
CO2inlet − CO2outlet bond. In addition, for the In1−xZrxOy surface, CO2 chem-
CO2 conversion = × 100% isorption at the oxygen vacancy site near the Zr dopant was
CO2inlet much stronger than that at the oxygen vacancy site of pure
In2O3, for example, the presence of the Zr dopant leads to
where CO inlet and CO outlet represent mol of CO at the inlet
stronger CO2 adsorption by ∼1.1 eV at the D4 site (Figure
and outlet, respectively.
3a,b). As shown in Figure 3c, other intermediates were also
CO was formed by the reverse water gas shift reaction and
significantly stabilized during CO2 hydrogenation, especially the
CO selectivity was calculated according to
methoxy (CH3O*) species adsorbed at the defect-free surface
COoutlet (P), which was the least stable among the various intermediates
CO selectivity = × 100% for the reaction on the pure In2O3 surface. Therefore, the
CO2inlet − CO2outlet
incorporation of Zr benefitted the formation of methanol from
where CO2 outlet denotes mol of CO2 at the outlet. CO2 hydrogenation.
The selectivity of individual hydrocarbon product CnHm DFT calculations revealed that oxygen-defective In2O3 can
based on CO-free was obtained according to be created through direct thermal desorption or exposure to
reducing agents.30,31 However, the treatment of fresh In2O3 in
nCnHmoutlet H2 at 300 °C resulted in a significant decrease in surface area.24
CnHm selectivity = Therefore, In2O3 pretreated in Ar can yield more active sites
CO2inlet − CO2outlet − COoutlet
than In2O3 activated in H2. In situ near-ambient pressure X-ray
× 100% photoelectron spectroscopy (NAP-XPS) of the In−Zr oxide
573 DOI: 10.1021/acscatal.7b02649
ACS Catal. 2018, 8, 571−578
ACS Catalysis Research Article

Figure 2. Hydrocarbon distribution and reaction rate (r) over In2O3


and In−Zr oxides and bifunctional catalysts composed of metal oxides
and SAPO-34 zeolites with different mass ratios as shown in
parentheses. Reaction conditions: 400 °C, 3.0 MPa, 9000 mL gcat−1
h−1, H2/CO2/N2 = 73/24/3. CO2 conversions and the formation of
CO by the RWGS reaction are reported in Table S3 in the Supporting
Information.

Figure 1. Characterizations of oxides used in the bifunctional catalyst.


HRTEM images of (a) In−Zr and (b) In2O3 samples. (c) In situ
NAP−XPS O 1s spectra of In−Zr oxide sequentially exposed to 50 Pa Figure 3. CO2 adsorption energy at the O4a surface oxygen vacancy
Ar, CO2, H2, and again to CO2 atmosphere under different conditions. (D4) site on (a) the pure In2O3 and (b) Zr doped In1−xZrxOy surfaces.
In situ NAP-XPS O 1s spectra can be deconvoluted into three distinct (c) Energy profiles of CO2 hydrogenation to form CH3OH on the
peaks: the main oxide peak at 530.2 eV (“Olattice”) and two additional In2O3 and In1−xZrxOy surfaces shown in black and red, respectively. D
peaks at 531.4 and 532.4 eV, which were assigned to O atoms next to a and P stand for defective and perfect surfaces with and without the
defect (“Odefect”) and surface hydroxyls (“OH”), respectively. (d) oxygen vacancy, respectively.
CO2−TPD spectra for the reduced In2O3, ZrO2, and In−Zr oxides
(thermal treatment in Ar at 400 °C for 1 h). (DRIFTS) measurement also confirmed that CO2 could be
activated over such an In2O3 surface with oxygen vacancies.
Several signals in the temperature range of 50−250 and 400−
surface suggested an increase in the amount of surface oxygen 600 °C were observed from CO2−TPD profiles, and a higher
defects detected from oxygen atoms adjacent to the defects temperature peak around 690 °C only appeared for the In−Zr
(Odefect, 22.3%, Table S2 in the Supporting Information; see sample (Figure 1d). The peaks at 400−600 °C originated from
also Figure 1c, 531.4 eV) upon exposure to Ar at 400 °C. The desorption of CO2 that interacted strongly with oxygen
Odefect concentration of In−Zr oxide was higher than the sum of vacancies, and the peak at approximately 690 °C for In−Zr
the individual contributions from pure ZrO2 and In2O3 (Table oxide was attributed to the surface oxygen vacancy site near the
S2 in the Supporting Information). Odefect concentration Zr dopant based on our DFT calculations. In addition, the area
dropped remarkably in CO2 at 400 °C, while H2 benefited of the high temperature peak was larger than the sum of ZrO2
the regeneration of the oxygen vacancies via hydrogenation, and In2O3, similar to the trend of the Odefect signal (Table S2 in
thereby, sustain the catalytic cycle (Figure 1c). The CO2 the Supporting Information). These results indicated that the
temperature-programmed desorption (CO2−TPD) and in incorporation of Zr into In2O3 created new kinds of vacancies
situ diffuse reflectance infrared Fourier transform spectroscopy with high concentration, consistent with our DFT calculations,
574 DOI: 10.1021/acscatal.7b02649
ACS Catal. 2018, 8, 571−578
ACS Catalysis Research Article

which progressively enhanced the reaction rate. As shown in 3.2. Catalytic Performance over Bifunctional Cata-
Figure S4 in the Supporting Information, several bands at 900− lysts. In−Zr/SAPO-34 catalyst exhibited C2=−C4= selectivity of
1100 cm−1 were visible in the DRIFT spectrum collected upon 76.4% in hydrocarbons and CH4 selectivity was only 4.3% with
CO2 adsorption over both In2O3 and In−Zr oxides activated in CO2 conversion of 35.5% at 400 °C. The effects of the H2/CO2
Ar at 400 °C, which were assigned to the adsorbed CO2 ratio and reaction pressure were explored. As shown in Figures
bridging two In atoms around the oxygen vacancy sites and was 5a and S6a in the Supporting Information, CO2 conversion rose
blue-shift with the incorporation of Zr due to stronger CO2 significantly with increasing H2/CO2 ratio (from 22.1% at H2/
adsorption.24,32 In addition, the region between 1200 and 1700 CO2 of 1.0 to 44.0% at H2/CO2 of 5.0) and pressure (from
cm−1 contained a large variety of different bands corresponding 29.5% at 1.0 MPa to 37.9% at 5.0 MPa). CO selectivity was
to carbonate species.33,34 These findings are in agreement with above 80% under different conditions. However, higher H2/
our DFT results. Moreover, the adsorbed species are relatively
stable and cannot be removed completely by He treatment
even up to 350 °C (Figure S4b in the Supporting Information),
while the intensities decreased significantly at 200 °C in the
presence of H2 due to the formation of hydrogenated species
(Figure S4c in the Supporting Information).
As far as the product selectivity was concerned, only about
1% of CH3OH and 1% of hydrocarbons were obtained with
either In−Zr oxide or bulk In2O3 as the sole catalyst for CO2
hydrogenation at 400 °C, and the main product was CO with
much high CH4 selectivity in the hydrocarbon distribution. It is
worth noting that methanol formation from CO2 hydro-
genation is restrained at high temperature due to its exothermic
character,35,36 and the equilibrium selectivity to CH3OH is only
0.5% at 400 °C (Figure S5 in the Supporting Information).
When the metal oxides were combined with SAPO-34 zeolites,
CO selectivity decreased from 97% to 85% and the reaction
rate per gram of In and C2=−C4= selectivity increased
remarkably because of the thermodynamic driving force. The
obtained hydrocarbons distribution from CO2 hydrogenation
over the oxide/zeolite catalyst was similar to that derived from
MTO over the sole SAPO-34 zeolite used in the bifunctional
catalyst (Figure 4). Moreover, the CO2 conversion can be

Figure 4. Influence of the atmosphere and H2 pressure on the product


selectivity for the conversion of methanol over the SAPO-34 catalyst.
Reaction conditions: catalyst, 1.0 g; Ar or H2, 0.1−3.0 MPa; 150 mL
min−1; liquid CH3OH, 0.003 mL min−1; time on stream, 15 h.

tuned by varying the mass ratio of In−Zr oxides to SAPO-34


zeolites (Figure 2). Therefore, it can be concluded that the
Figure 5. Catalytic performance for CO2 hydrogenation. CO2
oxide/zeolite composite catalyst separates CO2 activation and conversion and hydrocarbon distribution at different (a) H2/CO2
C−C coupling onto different sites with complementary and ratios and (b) space velocities. (c) Stability test of composite In−Zr/
compatible properties. During CO2 hydrogenation, CO2 is SAPO-34 catalyst. Reaction conditions: 400 °C, 3.0 MPa, 9000 mL
activated on the oxide surface with oxygen vacancies and gcat−1 h−1, H2/CO2/N2 = 73/24/3, and oxide/zeolite mass ratio = 2.
hydrogenated to intermediate species (methanol), while C−C The formation of CO by the RWGS reaction is reported in Table S4 in
coupling is controlled within the pores of SAPO-34. the Supporting Information.

575 DOI: 10.1021/acscatal.7b02649


ACS Catal. 2018, 8, 571−578
ACS Catalysis Research Article

CO2 ratio and pressure decreased C2=−C4= selectivity and


increased selectivities to CH4 and C2o−C4o gradually, indicating
that the high H2 partial pressure was not favorable for the
formation of C2=−C4=. In order to further reveal the influence
of H2 partial pressure, the catalytic performance of MTO was
investigated in the presence of H2 with different H2 pressures.
Under the Ar atmosphere typically used for the MTO reaction,
the C2=−C4= selectivity reached around 80% over SAPO-34 at
0.1 MPa, and it decreased slightly to 76% when the Ar pressure
rose to 3.0 MPa (Figure 4). The C2−C4 olefin selectivity
maintained 80% when the reaction atmosphere switched from
Ar to H2 at 0.1 MPa, whereas it decreased significantly and C2−
C4 paraffins selectivity increased sharply with increasing H2
pressure due to the second hydrogenation of the olefins over
the acid sites of SAPO-34 (Figure 4).37 Consequently,
decreasing the H2/CO2 ratio and pressure favors the formation
of lower olefin. In addition, the selectivity of lower olefin can be
further increased from 68% to 84% when the space velocity was
changed from 4500 to 15750 mL gcat−1 h−1, and the selectivities
of CH4 and C5+ were largely suppressed (Figure 5b).
The C−C coupling from methanol is thermodynamically
more favorable at high temperature, and 400−450 °C is optimal
for methanol to C2−C4 olefins over SAPO-34.27,29,38 However,
methanol synthesis is thermodynamically restrained at such
high temperature (Figure S5 in the Supporting Information).
Two parallel reactions, methanol synthesis and RWGS, occur
simultaneously during the CO2 hydrogenation process. The
RWGS reaction is favored at higher temperature, since it is an Figure 6. (a) CO2 conversion, CO selectivity, and hydrocarbon
endothermic reaction.12,39,40 Meanwhile, it is kinetically favored distribution over the In−Zr/SAPO-34 catalyst at various reaction
under methanol synthesis conditions for CO2 hydrogenation. temperatures; (b) catalytic performance over In−Zr/SAPO-34 as a
The reaction temperature is usually relative low (200−300 °C) function of CO concentration in the feed at 380 °C. Reaction
over methanol synthesis catalysts. Although the coupling with conditions: 3.0 MPa, 9000 mL gcat−1 h−1, H2/CO2/N2 = 73/24/3, and
the MTO reaction can derive the CO2 conversion, the zeolite is In−Zr/SAPO-34 mass ratio = 0.5.
not active for the C−C coupling reaction at such low
temperature. The mismatch of their reaction temperatures that the intermediates involved in methanol synthesis were
results in the production of a large amount of CO. Therefore, more stable on the oxygen vacancy site near the Zr dopant than
one of the key challenges in the selective formation of lower those on the oxygen vacancy site of pure In2O3, which strongly
olefins from CO2 hydrogenation is how to suppress CO suppressed CO formation. As shown in Table S5 and Figure
formation. S6c in the Supporting Information, CO selectivity for both CO2
Theoretical equilibrium selectivity of CO expected on the hydrogenations to methanol and lower olefins declined with the
basis of thermodynamic analysis of CO2 hydrogenation to incorporation of Zr, which agrees well with our DFT
methanol is as high as 98.5−99.7% with CO2 conversion of calculations. Furthermore, CO selectivity for the bifunctional
31.7−39.9% at 360−420 °C (Figure S5 in the Supporting catalyst can be significantly suppressed from 84% to 70% at 400
Information). Thus, it is necessary to investigate the effect of °C and 64% to 47% at 380 °C when the In−Zr oxide was
reaction conditions on CO selectivity. With the reaction further modified by the Zn promoter (Figure S6c in the
temperature decreasing from 420 to 360 °C, CO selectivity Supporting Information).
decreased sharply from 91% to 51% and these values are far It is noteworthy that the In−Zr/SAPO-34 sample exhibited
from the equilibrium values, and the selectivity of C2−C4 an excellent catalytic stability, and no obvious decline in activity
olefins increased from 68% to 86% (Figure 6a). In addition, (CO2 conversion ∼35%) was observed after a time-on-stream
higher H2/CO2 ratio and pressure also led to significantly of 152 h (Figure 5c). In addition, CH 4 and C 2 =−C 4=
reduced CO selectivity (Table S4 and Figure S6 in the selectivities remained stable around 4.5% and 77%, respectively.
Supporting Information). Furthermore, the CO selectivity can However, In2O3/SAPO-34 encountered severe deactivation
be greatly suppressed by adding CO into the feed gas (Figure after 152 h; for example, the CO2 conversion decreased from
6b). With increasing CO concentration to 14.4%, CO 34.6% to 27.7% and CH4 selectivity increased from 2.0% to
selectivity was only 20.5%, indicating that a larger fraction of 6.0% (Figure S6b in the Supporting Information). As
the available carbon resource in the CO2/H2 mixture has been mentioned above, the partially reduced oxide and SAPO-34
converted into hydrocarbons. The activity for CO2 hydro- zeolite were responsible for the activation of CO2 to methanol
genation to lower olefins and the selectivity of C2−C4 olefins and the selective C−C coupling, respectively. The metal oxide
remained almost unchanged with the increasing CO concen- nanoparticles are susceptible to agglomeration during reaction
tration in the feed gas. at high temperature, which is the primary factor for severe
Aside from the influence of the reaction conditions, the deactivation and low stability.24,41−43 In situ XRD character-
modification of the defective oxide surface can also suppress the ization revealed that the crystal size of sole In2O3 increased
undesired RWGS reaction. Our DFT calculations suggested sharply from 14.8 to 22.5 nm after the exposure to the H2 and
576 DOI: 10.1021/acscatal.7b02649
ACS Catal. 2018, 8, 571−578
ACS Catalysis Research Article

CO2 mixture at 400 °C during the initial stage (8 h) and rose remarkable deactivation of SAPO-34. Especially for In2O3/
gradually in the following reaction time (Figure S7a in the SAPO-34, the CH4 and C2=−C4= selectivities in hydrocarbons
Supporting Information). According to the TEM results of distribution were 98.2% and 0.5%, respectively, and CO2
In2O3/SAPO-34 catalysts in Figure S8 in the Supporting conversion was only 18.2% (Figure 7c). Moreover, In−Zr/
Information, much larger In2O3 nanoparticles with an average SAPO-34 exhibited much higher C2=−C4= selectivity (44.7%)
diameter up to 34.6 nm were present over the spent In2O3/ and CO2 conversion (26.7%) compared with In2O3/SAPO-34,
SAPO-34 after 48 h of reaction compared with the fresh sample though overly close contact of bifunctional active sites
(12.2 nm), consistent well with the XRD results. Nevertheless, decreased its performance for CO2 hydrogenation to lower
the mean particle size grew moderately for the In−Zr sample olefins.
with increasing reaction time (Figure S7b in the Supporting The effect of integration manner on physiochemical
Information), indicating that ZrO2 played a crucial role in properties for the bifunctional catalyst was investigated in
preventing the sintering of the active nanoparticles during detail. It was found that the distance among the two
reaction process. components had no significant effect on textural and structural
3.3. Effect of Integration Manner. We further inves- properties as well as surface acidity for fresh samples. However,
tigated the influence of integration manner of the active the amounts of strongly acidic sites for both spent In2O3/
components on catalytic performance for CO2 hydrogenation SAPO-34 and In−Zr/SAPO-34 catalysts prepared by mixing of
to lower olefins. When SAPO-34 was packed below the oxide two compounds in an agate mortar decreased remarkably after
and separated by a layer of inert quartz sand (dual-bed reaction for 48 h, especially for In2O3/SAPO-34 (Figure S10 in
configuration), the CH4 selectivity was above 55%, and small the Supporting Information). Solid-state magic-angle spinning
amounts of C2=−C4= with the selectivity around 30% were (MAS) nuclear magnetic resonance (NMR) spectroscopy was
formed due to the catalytic function of SAPO-34 in the employed to investigate the chemical environment of frame-
downstream (Figure 7a). Then, we mixed the granules of work atoms (Si, Al, and P) of fresh and spent SAPO-34 used in
the bifunctional catalyst. Nearly no changes were observed for
the 29Si, 27Al, and 31P MAS NMR spectra of spent In2O3/
SAPO-34 and In−Zr/SAPO-34 catalysts prepared by mortar-
mixing configuration compared with the fresh sample,
indicating that the reaction process had little or no effect on
the silicoaluminophosphate frameworks of SAPO-34 (Figure
S11 in the Supporting Information). In addition, the local
elemental compositions of fresh and spent In2O3/SAPO-34 and
In−Zr/SAPO-34 catalysts were analyzed by transmission
electron microscopy and energy-dispersive X-ray spectrometry
(Figures S12 and S13 in the Supporting Information). In the
SAPO-34 zeolite regions, the signal of the In element for the
spent catalysts prepared by mortar-mixing configuration was
much higher than that for the fresh samples, especially for
In2O3/SAPO-34, indicating the presence of In within the
Figure 7. Effect of the integration manner of the active components zeolite crystal or in close proximity to the zeolite crystal after
on catalytic performances. (a) Dual-bed configuration with SAPO-34
the reaction. Nevertheless, traces of In were detected in zeolite
packed below In2O3 or In−Zr oxides, separated by a layer of quartz
sand. (b) Stacking of granules with the In2O3, In−Zr oxides, and regions for spent In2O3/SAPO-34 and In−Zr/SAPO-34
SAPO-34 particle sizes of 250−380 μm. (c) Mixing of the two catalysts prepared by the granule-stacking configuration.
compounds in an agate mortar. All catalysts with oxide/zeolite mass Consequently, overly tight contact of bifunctional active sites
ratio = 2 were evaluated under the same conditions shown in Figure 5. resulted in migration of indium during the reaction followed by
ion exchange of indium ions with zeolite protons, which
oxides and zeolites together by shaking in a vessel to increase decreased the number of strongly acidic sites significantly,
the proximity of the two components. The CH4 selectivity leading to severe deactivation with very low C 2 = −C 4 =
decreased significantly (<5%) and C2=−C4= selectivity increased selectivity.
to above 75% (Figure 7b). Combined with the results that
C2=−C4= selectivity increased gradually with increasing space 4. CONCLUSIONS
velocities (Figure S6c in the Supporting Information), it follows In summary, we have discovered that a bifunctional catalyst
that fast transport of reaction intermediates in the gas phase contained In−Zr oxide and SAPO-34, which was responsible
benefits the selective formation of olefins from CO2 hydro- for the CO2 activation and the selective C−C coupling,
genation. respectively, could realize the direct production of lower olefins
For the granule-stacking configuration, micrometer-sized from CO2 hydrogenation with excellent selectivity and high
particles were stacked together (Figure S9a−d in the activity. The selectivity of C2−C4 olefin reached up to around
Supporting Information). In order to further shorten the 80% with much low CH4 selectivity and CO2 conversion was
distance among the two components, the grinding powders of above 35%. We demonstrated that the incorporation of
oxides and zeolites were mixed in an agate mortar. As shown in zirconium significantly enhanced the activity by creating new
Figures S8 and S9 in the Supporting Information, smaller oxide kinds of vacancies and remarkably improved the catalytic
particles with a mean size of ∼10 nm were observed to be stability by preventing the sintering of the oxide nanoparticles.
attached closely on much bigger SAPO-34 particles of 2 μm in The precise control of the proximity of the two active sites also
size. However, CO2 was converted mainly to CH4 with much played an important role in the direct conversion of CO2 to
lower CO2 conversion, and CH3OH was detected, indicating lower olefins. Additionally, the product distribution depended
577 DOI: 10.1021/acscatal.7b02649
ACS Catal. 2018, 8, 571−578
ACS Catalysis Research Article

on the nature of the oxygen vacancies. The defective In2O3 (13) Centi, G.; Quadrelli, E. A.; Perathoner, S. Energy Environ. Sci.
surface could be modified to further increase the stability of the 2013, 6, 1711−1731.
key intermediates involved in methanol formation, significantly (14) Gnanamani, M. K.; Jacobs, G.; Hamdeh, H. H.; Shafer, W. D.;
suppressing the undesired RWGS reaction. Liu, F.; Hopps, S. D.; Thomas, G. A.; Davis, B. H. ACS Catal. 2016, 6,


913−927.
(15) Dorner, R. W.; Hardy, D. R.; Williams, F. W.; Willauer, H. D.
ASSOCIATED CONTENT Energy & Environmental. Energy Environ. Sci. 2010, 3, 884−890.
*
S Supporting Information (16) Van der Laan, G. P.; Beenackers, A. A. C. M. Catal. Rev.: Sci.
The Supporting Information is available free of charge on the Eng. 1999, 41, 255−318.
ACS Publications website at DOI: 10.1021/acscatal.7b02649. (17) de Smit, E.; Weckhuysen, B. M. Chem. Soc. Rev. 2008, 37, 2758−
2781.
Details of the experiments and Tables S1−S5 and (18) Galvis, H. M. T.; de Jong, K. P. ACS Catal. 2013, 3, 2130−2149.
Figures S1−S13 as described in the text. (PDF) (19) Saeidi, S.; Amin, N. A. S.; Rahimpour, M. R. J. CO2 Util. 2014,


5, 66−81.
(20) Wei, J.; Sun, J.; Wen, Z. Y.; Fang, C. Y.; Ge, Q. J.; Xu, H. Y.
AUTHOR INFORMATION Catal. Sci. Technol. 2016, 6, 4786−4793.
Corresponding Authors (21) Visconti, C. G.; Martinelli, M.; Falbo, L.; Infantes-Molina, A.;
*E-mail: zhongls@sari.ac.cn. Lietti, L.; Forzatti, P.; Iaquaniello, G.; Palo, E.; Picutti, B.; Brignoli, F.
*E-mail: sunyh@sari.ac.cn. Appl. Catal., B 2017, 200, 530−542.
(22) Zhang, J. L.; Lu, S. P.; Su, X. J.; Fan, S. B.; Ma, Q. X.; Zhao, T. S.
ORCID J. CO2 Util. 2015, 12, 95−100.
Shenggang Li: 0000-0002-5173-0025 (23) Li, C.; Yuan, X.; Fujimoto, K. Appl. Catal., A 2014, 469, 306−
Liangshu Zhong: 0000-0002-4167-8630 311.
Notes (24) Martin, O.; Martin, A. J.; Mondelli, C.; Mitchell, S.; Segawa, T.
F.; Hauert, R.; Drouilly, C.; Curulla-Ferre, D.; Perez-Ramirez, J. Angew.
The authors declare no competing financial interest.


Chem., Int. Ed. 2016, 55, 6261−6265.
(25) Ilias, S.; Bhan, A. ACS Catal. 2013, 3, 18−31.
ACKNOWLEDGMENTS (26) Olsbye, U.; Svelle, S.; Bjorgen, M.; Beato, P.; Janssens, T. V. W.;
This work has been supported by the Ministry of Science and Joensen, F.; Bordiga, S.; Lillerud, K. P. Angew. Chem., Int. Ed. 2012, 51,
Technology of China (2016YFA0202802 and 5810−5831.
2017YFB0602202), the National Natural Science Foundation (27) Tian, P.; Wei, Y. X.; Ye, M.; Liu, Z. M. ACS Catal. 2015, 5,
1922−1938.
of China (21773286, 21503260, 91545112, and 11227902), (28) Gao, P.; Li, S. G.; Bu, X. N.; Dang, S. S.; Liu, Z. Y.; Wang, H.;
Shanghai Municipal Science and Technology Commission, Zhong, L. S.; Qiu, M. H.; Yang, C. G.; Cai, J.; Wei, W.; Sun, Y. H. Nat.
China (16DZ1206900 and 15DZ1170500), the Chinese Chem. 2017, 9, 1019−1024.
Academy of Sciences (QYZDB-SSW-SLH035), and Science (29) Cheng, K.; Gu, B.; Liu, X. L.; Kang, J. C.; Zhang, Q. H.; Wang,
and Technology Innovation Fund of Shanghai Advanced Y. Angew. Chem., Int. Ed. 2016, 55, 4725−4728.
Research Institute, CAS (172001). (30) Ye, J. Y.; Liu, C. J.; Mei, D. H.; Ge, Q. F. ACS Catal. 2013, 3,


1296−1306.
REFERENCES (31) Ye, J. Y.; Liu, C. J.; Ge, Q. J. Phys. Chem. C 2012, 116, 7817−
7825.
(1) Olah, G. A.; Goeppert, A.; Prakash, G. K. S. J. Org. Chem. 2009, (32) Bai, S.; Shao, Q.; Wang, P.; Dai, Q.; Wang, X.; Huang, X. J. Am.
74, 487−498. Chem. Soc. 2017, 139, 6827−6830.
(2) Olah, G. A.; Prakash, G. K. S.; Goeppert, A. J. Am. Chem. Soc. (33) Zhu, Y.; Pan, X.; Jiao, F.; Li, J.; Yang, J.; Ding, M.; Han, Y.; Liu,
2011, 133, 12881−12898. Z.; Bao, X. ACS Catal. 2017, 7, 2800−2804.
(3) Banerjee, A.; Dick, G. R.; Yoshino, T.; Kanan, M. W. Nature (34) Larmier, K.; Liao, W. C.; Tada, S.; Lam, E.; Verel, R.; Bansode,
2016, 531, 215−219. A.; Urakawa, A.; Comas-Vives, A.; Coperet, C. Angew. Chem., Int. Ed.
(4) Studt, F.; Sharafutdinov, I.; Abild-Pedersen, F.; Elkjaer, C. F.; 2017, 56, 2318−2323.
Hummelshoj, J. S.; Dahl, S.; Chorkendorff, I.; Norskov, J. K. Nat. (35) Graciani, J.; Mudiyanselage, K.; Xu, F.; Baber, A. E.; Evans, J.;
Chem. 2014, 6, 320−324. Senanayake, S. D.; Stacchiola, D. J.; Liu, P.; Hrbek, J.; Sanz, J. F.;
(5) Moret, S.; Dyson, P. J.; Laurenczy, G. Nat. Commun. 2014, 5, Rodriguez, J. A. Science 2014, 345, 546−550.
4017. (36) Porosoff, M. D.; Yan, B. H.; Chen, J. G. G. Energy Environ. Sci.
(6) Zhong, L. S.; Yu, F.; An, Y. L.; Zhao, Y. H.; Sun, Y. H.; Li, Z. J.; 2016, 9, 62−73.
Lin, T. J.; Lin, Y. J.; Qi, X. Z.; Dai, Y. Y.; Gu, L.; Hu, J. S.; Jin, S. F.; (37) Senger, S.; Radom, L. J. Am. Chem. Soc. 2000, 122, 2613−2620.
Shen, Q.; Wang, H. Nature 2016, 538, 84−87. (38) Lee, Y. J.; Baek, S. C.; Jun, K. W. Appl. Catal., A 2007, 329,
(7) Jiao, F.; Li, J. J.; Pan, X. L.; Xiao, J. P.; Li, H. B.; Ma, H.; Wei, M. 130−136.
M.; Pan, Y.; Zhou, Z. Y.; Li, M. R.; Miao, S.; Li, J.; Zhu, Y. F.; Xiao, D.; (39) Natesakhawat, S.; Lekse, J. W.; Baltrus, J. P.; Ohodnicki, P. R.;
He, T.; Yang, J. H.; Qi, F.; Fu, Q.; Bao, X. H. Science 2016, 351, 1065− Howard, B. H.; Deng, X. Y.; Matranga, C. ACS Catal. 2012, 2, 1667−
1068. 1676.
(8) Galvis, H. M. T.; Bitter, J. H.; Khare, C. B.; Ruitenbeek, M.; (40) Yang, H. Y.; Zhang, C.; Gao, P.; Wang, H.; Li, X. P.; Zhong, L.
Dugulan, A. I.; de Jong, K. P. Science 2012, 335, 835−838. S.; Wei, W.; Sun, Y. H. Catal. Sci. Technol. 2017, 7, 4580−4598.
(9) Sakakura, T.; Choi, J. C.; Yasuda, H. Chem. Rev. 2007, 107, (41) Chaudhuri, R. G.; Paria, S. Chem. Rev. 2012, 112, 2373−2433.
2365−2387. (42) Richard, A. R.; Fan, M. H. ACS Catal. 2017, 7, 5679−5692.
(10) Zhang, X. B.; Zhu, X. B.; Lin, L. L.; Yao, S. Y.; Zhang, M. T.; (43) Karelovic, A.; Ruiz, P. Catal. Sci. Technol. 2015, 5, 869−881.
Liu, X.; Wang, X. P.; Li, Y. W.; Shi, C.; Ma, D. ACS Catal. 2017, 7,
912−918.
(11) Wei, J.; Ge, Q.; Yao, R.; Wen, Z.; Fang, C.; Guo, L.; Xu, H.; Sun,
J. Nat. Commun. 2017, 8, 15174.
(12) Wang, W.; Wang, S. P.; Ma, X. B.; Gong, J. L. Chem. Soc. Rev.
2011, 40, 3703−3727.

578 DOI: 10.1021/acscatal.7b02649


ACS Catal. 2018, 8, 571−578

You might also like