Anda di halaman 1dari 6

ARTICLES

PUBLISHED ONLINE: 12 JUNE 2017 | DOI: 10.1038/NCHEM.2794

Direct conversion of CO2 into liquid fuels with high


selectivity over a bifunctional catalyst
Peng Gao1, Shenggang Li1,2, Xianni Bu1, Shanshan Dang1,3, Ziyu Liu1, Hui Wang1, Liangshu Zhong1*,
Minghuang Qiu1, Chengguang Yang1, Jun Cai2,4, Wei Wei1,2 and Yuhan Sun1,2*

Although considerable progress has been made in carbon dioxide (CO2) hydrogenation to various C1 chemicals, it is still a
great challenge to synthesize value-added products with two or more carbons, such as gasoline, directly from CO2 because
of the extreme inertness of CO2 and a high C–C coupling barrier. Here we present a bifunctional catalyst composed of
reducible indium oxides (In2O3) and zeolites that yields a high selectivity to gasoline-range hydrocarbons (78.6%) with a
very low methane selectivity (1%). The oxygen vacancies on the In2O3 surfaces activate CO2 and hydrogen to form
methanol, and C−C coupling subsequently occurs inside zeolite pores to produce gasoline-range hydrocarbons with a high
octane number. The proximity of these two components plays a crucial role in suppressing the undesired reverse water gas
shift reaction and giving a high selectivity for gasoline-range hydrocarbons. Moreover, the pellet catalyst exhibits a much
better performance during an industry-relevant test, which suggests promising prospects for industrial applications.

O
wing to the growing energy demand, dwindling fossil fuel (HZSM-5) zeolites exhibits an excellent performance for the direct
reserves and increasing atmospheric CO2 concentration, production of gasoline-range hydrocarbons from CO2 hydrogen-
renewable energy sources, such as solar, wind and biomass, ation with a high selectivity. The C5+ selectivity in hydrocarbon dis-
foresee increasing usage. However, the widespread utilization of tribution reached up to 78.6% with only 1% for CH4 selectivity at a
renewable energy sources is currently limited by their intermittent CO2 conversion of 13.1% (Fig. 1a). There was no obvious catalyst
and fluctuating nature. In this context, the chemical conversion of deactivation over 150 hours, and a much better performance for
carbon dioxide (CO2) into value-added products with the assistance CO2 hydrogenation to C5+ hydrocarbons was observed using a
of hydrogen (H2) would represent a promising solution to the pellet catalyst with internal gas recycling. Such results thus suggest
storage of renewable energy1–3. a promising potential for its industrial application.
Thus, much attention has been paid to CO2 hydrogenation to
various C1 feedstocks (for example, methane (CH4), methanol Results and discussions
(CH3OH), carbon monoxide (CO) and formic acid) and considerable The bifunctional catalyst consists of a metal oxide (In2O3) with a
progress has been achieved4–11. However, the extreme inertness of CO2 small particle size of about 10 nm (Fig. 1b and Supplementary Figs
and the high kinetic barriers for the formation of C–C bonds12,13 mean 1 and 2) and a high specific surface area of 120 m2 g–1
it is still a great challenge to synthesize C2+ (hydrocarbons with two or (Supplementary Table 1), and a HZSM-5 zeolite with a hierarchical
more carbons) products directly from CO2 , such as gasoline (C5–C11 micro/mesoporous structure with a mesopore diameter of 4 nm
hydrocarbons), which is a very important transportation fuel widely (Fig. 1c and Supplementary Fig. 3a). The catalytic performance
used around the world. CO2 is well-known to be a very stable molecule and structure information of the sole In2O3 component were
(ΔfG o = –396 kJ mol–1), the end product of any combustion process, investigated first. With In2O3 as the sole catalyst under the same reac-
either biological or chemical, along with water14,15. Another key bottle- tion conditions, CO and CH3OH were the major products. The
neck problem is the assembly of the atoms and formation of chemical hydrocarbon selectivity was rather low (2.4%) and no C5+ hydrocar-
bonds to convert the relatively simple CO2 molecules into the much bons were detected (Fig. 1a). Density functional theory (DFT) calcu-
more complex and energetic C5–C11 hydrocarbons. Obviously, signifi- lations (Supplementary Fig. 4) revealed that In2O3 is reducible, which
cant catalytic advances are required for the large-scale production of results in a surface rich in oxygen vacancies. The oxygen vacancy on
liquid fuels directly from CO2 hydrogenation. Currently, the the In2O3 surface benefits CO2 activation and hydrogenation, and
Fischer–Tropsch synthesis (FTS) route using modified Fe-based can also be generated through thermal desorption18,19. After pre-
catalysts can be utilized to produce hydrocarbons directly from CO2. treatment of In2O3 in argon (Ar) at 400 °C for one hour, the amount
However, the maximum C5–C11 hydrocarbon fraction was limited of surface oxygen defects detected from oxygen atoms adjacent to the
by the Anderson–Schulz–Flory distribution to ∼48% with an undesir- defects by in situ near-ambient pressure X-ray photoelectron spec-
able CH4 fraction of ∼6% (refs 16,17). Furthermore, the heat of adsorp- troscopy (NAP–XPS, O 1s at 531.7 eV (Supplementary Fig. 5a and
tion of CO2 is lower than that of CO because of the thermodynamic Supplementary Table 2)) increased to 24.8%. However, the changes
stability of CO2 , which leads to a much lower coverage of CO2 over in the particle sizes calculated from in situ X-ray diffraction (XRD) pat-
the catalyst, and thus a low CO2 reactivity and high CH4 selectivity. terns (Supplementary Fig. 1) and specific surface areas
In the present work, a bifunctional catalyst that contains partially (Supplementary Table 1) were negligible during the pre-treatment
reducible metal oxides (In2O3) and H-form Zeolite Socony Mobil-5 process. The CO2 and H2 adsorption properties of the In2O3

1
CAS Key Laboratory of Low-Carbon Conversion Science and Engineering, Shanghai Advanced Research Institute, Chinese Academy of Sciences, Shanghai
201210, China. 2 School of Physical Science and Technology, ShanghaiTech University, Shanghai 201203, China. 3 University of the Chinese Academy of
Sciences, Beijing 100049, China. 4 State Key Laboratory of Functional Materials for Informatics, Shanghai Institute of Microsystem and Information
Technology, Chinese Academy of Sciences, Shanghai 200050, China. * e-mail: zhongls@sari.ac.cn; sunyh@sari.ac.cn

NATURE CHEMISTRY | ADVANCE ONLINE PUBLICATION | www.nature.com/naturechemistry 1

© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
ARTICLES NATURE CHEMISTRY DOI: 10.1038/NCHEM.2794

a
CO2 conv. CO sel. CH3OH sel.
100 100

90 90

Conversion and selectivity (%)


Hydrocarbon distribution (%)
80 80

70 70

60 60

50 50
40
40
30
30
20
20
10
10
0
0
) ) ) ) 1) 1) -5
1/2 1/1 2/1 4/1 O3
In 2 5 (2/ (2/ ZSM
5( 5( 5( 5( -5
S M- S M- S M- S M- S M -
S M H
/HZ /HZ /HZ /HZ HZ Zr/HZ
O 3 n 2O 3 n 2O 3 n 2O 3 AI/
In 2 I I I Z n– – Al–
– n
Cu u –Z
C
b c

00
1
1 nm

10
In2O3 (222)

0
(200)

0.290 nm

5 nm 10 nm

Figure 1 | Catalytic performance over various catalysts and the morphology of the In2O3/HZSM-5 bifunctional catalyst. a, CO2 hydrogenation over various
bifunctional catalysts that contained Cu-based catalysts or In2O3 and HZSM-5 with different mass ratios, as shown in the parentheses, and the stand-alone
In2O3 catalyst (reaction conditions, 340 °C, 3.0 MPa, 9,000 ml h–1 gcat–1, H2/CO2/N2 = 73/24/3), as well as the conversion (conv.) of CH3OH over HZSM-5
(reaction conditions, 340 °C, Ar 3.0 MPa, 9,000 ml h–1 gcat–1; liquid CH3OH 0.855 ml h–1 gcat–1). Compared with Cu-based catalysts combined with HZSM-5,
the In2O3/HZSM-5 catalysts show much lower selectivities (sel.) of CO and CH4. The sole In2O3 metal oxide and HZSM-5 zeolite exhibit a typical catalytic
performance of CO2 hydrogenation into CH3OH, and CH3OH into gasoline, respectively. C5+ , red; C2–4 , blue; CH4 , grey. b, HRTEM image of In2O3 with an
exposed facet of (222) after the thermal treatment in Ar at 400 °C for 1 h. c, HRTEM image of HZSM-5 with a lattice spacing of 1.0 nm assigned to the
(200) lattice plane of the ZSM-5 crystals.

surface were investigated further by CO2 and H2 temperature- into hydrocarbons inside the HZSM-5. DFT calculations were
programmed desorption (TPD), respectively. As illustrated in carried out to predict the catalytic cycle of CO2 hydrogenation to
Supplementary Fig. 6a, the observed peak at approximately 417 °C CH3OH at the oxygen-vacancy site of In2O3 , and the potential
originated from the desorption of CO2 that interacted strongly with energy surface is shown in Fig. 2a,c. CO2 first chemisorbed at
the surface. The dissociative adsorption of H2 was detected at 114 °C, the oxygen-vacancy site, and the chemisorbed CO2* species
which formed atomic hydrogen on the In2O3 surface capable of (* represents the surface-adsorbed species) then underwent stepwise
participating in hydrogenation reactions (Supplementary Fig. 6b). hydrogenation to give formate (bi-HCO2*), dioxymethylene
The sole HZSM-5 zeolite used in the bifunctional catalyst was (bi-H2CO2*), methoxy (mono-H3CO*) and finally CH3OH. The
also tested for CH3OH conversion and exhibited a typical catalytic oxygen vacancy was filled during the above process, and was regen-
performance of CH3OH to gasoline with a complete CH3OH erated by subsequent hydrogenation. CH3OH formed over the
conversion (100%) and a high C5+ selectivity of 71.4% (Fig. 1a)20,21. surface of In2O3 passed further through the HZSM-5 zeolite,
We therefore conclude that CO2 was first hydrogenated to where it was transformed into various hydrocarbons at the acidic
CH3OH on the In2O3 surface, and CH3OH was then transformed site of the zeolite via the hydrocarbon-pool mechanism, as shown

2 NATURE CHEMISTRY | ADVANCE ONLINE PUBLICATION | www.nature.com/naturechemistry

© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
NATURE CHEMISTRY DOI: 10.1038/NCHEM.2794 ARTICLES
a b

2 0.24 CnH2n
0.00 1

Cy
*+ P

2 + OH
+ CO +

c
liz
2 D
CO +

5H 3 C
D
H3 H2 3

a
5

tio
0. H
3H −0.66
2 1. Hydrocarbon pool

+ *+

n
Hydrogen
*+

HO
2 D −1.06 CH3OH transfer
−1.13
CO 2+ −1.17
3H *+ H H
+
H2
4
CO P
2
−1.65 CO + D CO + D HZSM-5
H2 H2 −1.79 H 3 2+ H3 O i-CnH2n+2 Aromatic
*+ 2 +
H
H2
O2 +D H*
HC H 2 CO + P −2.32 H2O
5
2. H3 H2
+

c d
CH4 (1%)
(2) (3)
(1) 2H2
CH3OH C2–C4 (20.4%)
CO2
H2O H2
C5+ (78.6%)
(4)

(CH2)n

In2O3 HZSM-5

Figure 2 | Molecular-level mechanism for CO2 hydrogenation into hydrocarbons. a, Energy profile from DFT calculations for CO2 hydrogenation to CH3OH
on the In2O3(110) surface (D and P stand for defective and perfect surfaces with and without the oxygen vacancy, respectively). b, Schematic of the
hydrocarbon-pool mechanism for CH3OH conversion into hydrocarbons inside HZSM-5. c, Schematic for the formation of CH3OH from CO2 at the oxygen-
vacancy site on the In2O3 catalyst surface, which involve four major steps: (1) CO2 adsorption at the oxygen-vacancy site, (2) sequential hydrogenation of
the adsorbed CO2 species to CH3OH, (3) CH3OH desorption, which leads to the surface without the oxygen vacancy, and (4) hydrogenation of the surface
to regenerate the oxygen-vacancy site. d, Schematic for hydrocarbon formation from CH3OH at the acidic site inside the pores of the HZSM-5 catalyst via
the hydrocarbon-pool mechanism.

in Fig. 2b,d22–24. Moreover, it was also found that the C5–C11 iso- integrated with the zeolite (<45%) was much lower than that over
paraffins were mainly obtained from CO2 hydrogenation over the traditional Cu-based catalysts (>90%) at 340 °C. Furthermore,
In2O3/HZSM-5 composite catalyst, and the ratio of isoparaffins to compared with other reducible metal oxides, such as Ga2O3 ,
n-paraffins was 16.8 (Supplementary Table 3 and Supplementary Fe2O3 , ZnO–Cr2O3 and ZnO–ZrO2 combined with HZSM-5, the
Fig. 7). This was very different from that of CO2-based FTS, in In2O3/HZSM-5 composite catalyst exhibited a much better perform-
which normal paraffins and olefins were the main products, and ance for CO2 hydrogenation with a higher activity and a higher
nearly no isoparaffins with a high octane number were formed25–27. selectivity to C5+ hydrocarbons (Supplementary Table 4).
CH3OH synthesis from CO2 hydrogenation is usually The hydrocarbon selectivity and distribution were also affected
accompanied by undesirable CO formation via the reverse water markedly by the integration manner of the active components35,36.
gas shift (RWGS) reaction, which is further enhanced by increasing Using a dual-bed configuration with the HZSM-5 packed above
the reaction temperature because of the endothermicity of the RWGS the In2O3 and separated by a layer of inert quartz sand (Fig. 3a),
reaction. According to the equilibrium calculations (Supplementary the CH4 selectivity was very high (66.3%), whereas the selectivity
Fig. 8), the equilibrium selectivity of CO increases significantly with to C5+ hydrocarbons was only 26.7% and a large amount of
increasing temperature, and reaches 97.1% at 340 °C. When the CH3OH (31.8%) was detected. When HZSM-5 particles were
typical Cu–ZnO–Al2O3 industrial catalyst28 for the synthesis of loaded below the oxides (Fig. 3b), the formed CH3OH can be
CH3OH from syngas and the highly efficient Cu–ZnO–Al2O3– completely converted into hydrocarbons, and the C5+ selectivity
ZrO2 catalyst29 for CO2 hydrogenation to give CH3OH were increased to 70.4%, whereas the CH4 selectivity dropped to 4.5%.
employed for this reaction instead of In2O3 , the main product was Furthermore, by moving the two components into a closer
CO with a much higher CH4 selectivity in the hydrocarbon distri- proximity (from Fig. 3b to Fig. 3d), the CO selectivity decreased
bution under the same reaction conditions (Fig. 1a). Similar results significantly, because of the suppression of the undesired RWGS
were found for other conventional CH3OH synthesis catalysts, such reaction, the C5+ selectivity enhanced and the CH4 selectivity
as Cu–ZnO, Cu–ZnO–ZrO2 and Cu–ZnO–Cr2O3 combined with reduced, although the CO2 conversion only changed slightly. We
various zeolites30–33. Recent DFT calculations demonstrated that observed an even higher C5+ selectivity of 78.6% and a very low
the key intermediates involved in CH3OH synthesis were more CH4 selectivity of 1% over the composite catalyst without quartz
stable on the defective In2O3 surface than those on the Cu surface, sand (Fig. 3d). The effect of the relative amounts of each component
which strongly suppressed the formation of CO (ref. 18). The of In2O3/HZSM-5 on the catalytic performance was also
CH3OH selectivity over In2O3 can be tuned up to 100% when the investigated. With a decreasing relative amount of In2O3 , both
space velocity is above 16,000 h−1, even at the high reaction tempera- CO2 conversion and CO selectivity decreased. The hydrocarbon
ture of 300 °C, whereas much higher space velocities are required for distribution changed slightly when the oxide/zeolite mass ratio
Cu-based catalysts19,34. Consequently, CO selectivity over In2O3 decreased from 2:1 to 1:2, whereas the C5+ selectivity dropped and

NATURE CHEMISTRY | ADVANCE ONLINE PUBLICATION | www.nature.com/naturechemistry 3

© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
ARTICLES NATURE CHEMISTRY DOI: 10.1038/NCHEM.2794

100 100 Fig. 3d). As a result, the number of active sites for the C–C coupling
CH4 C2–C4 C5+
90 90 in the zeolite pore was significantly reduced, which leads to a severe
deactivation with very low C5+ hydrocarbon selectivity.

Conversion and selectivity (%)


80 80
Hydrocarbon distribution (%)

We also investigated the stability of the composite catalyst


70 CO sel. 70 with granule stacking (Fig. 4b). The CO2 conversion and the
60 60 CO selectivity decreased, whereas the C5+ selectivity increased
significantly during the initial 40 hours. However, the CO2
50 50
conversion and C5+ selectivity remained stable at around 12 and
40 40 80%, respectively, after a time-on-stream of 150 hours at 340 °C,
CH3OH sel.
30 30 3.0 MPa and 9,000 ml h–1 gcat–1, which suggests a promising poten-
CO2 conv.
tial for industrial applications. In situ XRD characterization revealed
20 20
that the crystal size of In2O3 increased substantially after the
10 10 exposure to the H2 and CO2 mixture at 340 °C during the initial
0 0 stage (four hours), whereas it remained almost unchanged in the
subsequent reaction period (Supplementary Fig. 1). In addition,
a b c d e the effects of pressure and the H2/CO2 ratio on the catalytic per-
Dual bed Granule stacking Mortar mixing
formance were also studied. A higher pressure and H2/CO2 ratio
can enhance CO2 conversion and decrease CO selectivity
(Supplementary Fig. 9). Moreover, a higher reaction rate with a
Quartz sand

ZSM-5
lower contact time suggested a negative effect of the generated
water (H2O) on the catalyst performance (Supplementary
In2O3 Fig. 10a). The increasing trend of CO2 conversion with decreasing
500–800 μm 0.5–1.0 μm
a 90 90
Figure 3 | Influence of the integration manner of the active components CH4 C2–C4 C5+
(In2O3/HZSM-5 mass ratio = 2:1) on catalytic behaviours under the same 80 80

Conversion and selectivity (%)


conditions. a, Dual-bed configuration with In2O3 packed below HZSM-5 and
Hydrocarbon distribution (%)

70 70
separated by a layer of quartz sand. b, HZSM-5 packed below In2O3 and
separated by quartz sand. c, Stacking of granules with the In2O3 , HZSM-5 and 60 CO sel. 60
quartz sand particle sizes of 250–380 µm. d, In2O3 and HZSM-5 particles 50 50
well mixed without quartz sand. e, In2O3 and HZSM-5 mixed with an agate
mortar. The catalytic performance is improved significantly by moving the two 40 40
components to a closer proximity, whereas the C5+ hydrocarbon selectivity 30 30
decreases remarkably with a further increase in the proximity by grinding the CO2 conv.
powder mixture of the two active components in an agate mortar. 20 20

10 10
CH3OH was detected with a selectivity of 2.5% when the weight
ratio increased from 2:1 to 4:1. In addition, an increase in the 0 0
space velocity can also greatly decrease the CO selectivity and 4,500 6,750 9,000 11,250
enhance the C5+ selectivity (Fig. 4a, 80.3% at 11,250 ml h–1 gcat–1). Space velocity (mL h–1 gcat–1)
Therefore, any measure that improves the transport of the reaction b 90 90
intermediates in the gas phase should favour the shift of the reaction
equilibrium for the selective formation of C5+ hydrocarbons. 80 C5+ 80
The sample illustrated in Fig. 3d was composed of well-mixed

Conversion and selectivity (%)


70 70
Hydrocarbon distribution (%)

micrometre-sized In2O3 and HZSM-5 particles (Supplementary


Fig. 3b). We tried to shorten the distance between the two 60 60
components further by grinding a powder mixture of them both CO sel.
50 50
in an agate mortar to investigate the effect of their intimate
contact (Fig. 3e). For the sample obtained from this integration 40 40
manner, the much smaller In2O3 particles (around 10 nm) were C2–C4
30 30
in much closer contact with the HZSM-5 particles of 500–800 nm
in size (Supplementary Fig. 2c–f ). However, CO2 was converted 20 CO2 conv. 20
mainly into CH4 (94.3% in hydrocarbon distribution) with a
rather low CO2 conversion (8%). In addition, the C5+ hydrocarbon 10 10
CH4
selectivity was only 4.2% and the CH3OH selectivity reached 51.9%, 0 0
which suggests a significant deactivation of HZSM-5. The very close 0 20 40 60 80 100 120 140 160
contact between In2O3 and the zeolite appeared to weaken the Reaction time (h)
synergistic effect. A similar phenomenon was also observed by
de Jong and co-workers, who reported that the closest proximity Figure 4 | Catalytic performance of the composite catalyst presented in
of bifunctional active sites was detrimental to the selective hydrocrack- Fig. 3d. a, CO2 conversion, CO selectivity and hydrocarbon distribution at
ing of hydrocarbons37. Similar textural and structural properties, as different space velocities. CO selectivity decreases and C5+ selectivity
well as surface acidity, were observed for the spent samples pre- increases significantly with increasing space velocities. b, Stability of the
sented in Fig. 3d,e, which indicates that the integration manner composite catalyst with granule stacking at 9,000 ml h–1 gcat–1. Both CO2
has no significant effect on these properties (Supplementary conversion and product selectivity remain stable after the initial 40 h of
Fig. 3). However, the number of strongly acidic sites of the spent reaction. Reaction conditions, 340 °C, 3.0 MPa, H2/CO2/N2 = 73/24/3 and
catalyst presented in Fig. 3e decreased markedly (Supplementary In2O3/HZSM-5 mass ratio = 2:1.

4 NATURE CHEMISTRY | ADVANCE ONLINE PUBLICATION | www.nature.com/naturechemistry

© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
NATURE CHEMISTRY DOI: 10.1038/NCHEM.2794 ARTICLES
a 90 55 similar performance to that of the granule catalyst under the same
C5+ reaction conditions. With an increasing recycle ratio, defined as
80 50
the flow rate of the recycled tail gas over that of the original feed

Conversion and selectivity (%)


Hydrocarbon distribution (%)

45
70 CO sel. gas, both CO2 conversion (from 8.7 to 18.2%) and C5+ hydrocarbon
40
60 selectivity (from 78.0 to 84.1%) increased, whereas CO selectivity
35
50
decreased substantially from 44.5 to 30.0%. The CO concentration
30
in the feed gas would obviously grow with an increasing recycle
40 25
ratio as the recycled tail gas contained a large amount of CO. To
CO2 conv. 20
30 study the effect of CO on the performance of CO2 hydrogenation,
15 CO was added into the feed gas of the CO2/H2 mixture with the
20
10 CO/(CO2 + CO) ratio increased from 0 to 100%. Increasing the
10 CH4 5 CO concentration was found to boost the activity of CO2 hydrogen-
0 0 ation to C5+ hydrocarbons, as shown in Fig. 5b. According to our
0 1 2 3 DFT calculations, the formation of oxygen vacancies on the
Recycle ratio In2O3(110) surface by CO reduction is energetically much more
b
favourable than H2 reduction (Supplementary Fig. 4b), which is
30 4.0
consistent with the experimental observations (Supplementary
3.5 Fig. 6c and Supplementary Table 2). In addition, the adsorption
25 C5+ yield
strength followed the order CO2 > H2 ≫ CO at a relatively low
C5+ yield (mmol h–1 gcat–1)
3.0
CO2 conversion (%)

20 oxygen-vacancy coverage (Supplementary Fig. 4c). Consequently,


2.5
CO can increase the number of active sites (oxygen vacancies),
15 2.0 and then promote the CO2 hydrogenation activity and counterba-
1.5
lance the efficient annihilation of vacancies by CO2. However, we
CO2 conv.
10 did not detect any C5+ hydrocarbon at CO/(CO2 + CO) = 100%
1.0 (CO/H2) within eight hours, because In2O3 was reduced to the met-
5 allic phase in the absence of CO2 at 340 °C (Supplementary Fig. 6d),
0.5
and this metallic phase was unable to catalyse CO2 hydrogenation19.
0 0.0
With the above oxide/zeolite bifunctional catalyst for CO2 hydro-
0 20 40 60 80 100 genation to hydrocarbons, the hydrocarbon distribution can be
CO/(CO + CO2) (%) easily tuned through the shape selectivity of the zeolite. When
SAPO-34 was chosen as the active phase for C−C coupling, the
Figure 5 | Catalytic performance with tail-gas recycling and as a function selectivity to lower olefins (C2=−C4=, generally referring to ethylene,
of CO concentration in the feed. a, Catalytic results over a technical propylene and butylene) reached 76.9% in hydrocarbons with only
catalyst composed of In2O3 and HZSM-5 pellets (size in cylindrical shape of 4.4% of CH4 at a CO2 conversion of 34.1% (Supplementary Fig. 12).
Φ3.0 × 3.5 mm, In2O3/HZSM-5 mass ratio = 1:2) with internal gas recycling In addition, the main hydrocarbon products became liquefied pet-
as functions of the recycle ratio. The recycle ratio represents the flow rate of roleum gas (C3o−C4o, referring to C3 and C4 paraffins) when the
the recycled tail gas over that of the original feed gas. With increasing Beta zeolite was used in combination with In2O3. Additionally, the
recycle ratio, the catalytic performance becomes much better over the catalytic activity depended on the amount of the active vacancies,
well-mixed In2O3 and HZSM-5 pellet catalyst under industrially relevant which could be increased substantially by the introduction of
conditions. b, CO2 conversion and yield of C5+ hydrocarbons over a modifiers and/or the use of a suitable support (such as ZrO2).
In2O3/HZSM-5 composite catalyst (particle size, 250–380 µm) as a
function of CO concentration in the feed at H2/(CO2 + CO)/N2 = 73/24/3. Conclusions
The CO concentration in the feed gas will grow with increasing recycle ratio, In summary, we discovered a bifunctional catalyst composed of par-
and therefore CO was added into the feed gas to investigate the effect of tially reduced In2O3 and HZSM-5 that could convert CO2 directly
CO on the performance without internal gas recycling. The C5+ yield into liquid fuels with an excellent selectivity for value-added pro-
increases substantially with an increasing CO/(CO2 + CO) ratio. Reaction ducts. The C5+ selectivity reached 78.6% with only 1% CH4 at a
conditions, 340 °C, 3.0 MPa, space velocity of the original feed gas CO2 conversion of 13.1%. We demonstrated that the CO2 conver-
9,000 ml h–1 gcat–1 and In2O3/HZSM-5 mass ratio = 1:2. sion could be manipulated by controlling the surface structure of
the oxide and the mass ratio of the oxides/zeolite. The proximity of
space velocity should lead to an increase in the H2O partial pressure the two components played a crucial role to suppress the undesired
across the catalyst bed, which would oxidize In2O3 vacancies (e.g. RWGS reaction and obtain a high C5+ selectivity. Industry-relevant
H2O + D → P + H2) (ref. 38) and lead to a lower catalytic activity. tests using the pellet catalyst were carried out to show that tail-gas
This was proved by the steady drop in the reaction rate when H2O recycling could improve further the catalytic performance for CO2
was co-fed during CO2 hydrogenation (Supplementary Fig. 10b). hydrogenation to C5+ hydrocarbons, which suggests a promising
To investigate further the prospect of the In2O3/HZSM-5 bifunc- prospect for industrial applications.
tional catalyst in industrial applications, the catalytic performance
for CO2 hydrogenation to C5+ hydrocarbons was evaluated using Methods
a pellet catalyst under industrially relevant conditions. The dry Catalyst preparation. Various oxides were prepared by a co-precipitation method.
HZSM-5 zeolites were prepared by the hydrothermal route with ethylamine as the
catalyst powders were compressed by a high-pressure agglomeration template. For the preparation of the composite catalyst presented in Fig. 2d, the In2O3
technique to obtain both In2O3 and HZSM-5 pellets in a and HZSM-5 were pressed, crushed and sieved to granules in the range of 40–60 mesh
cylindrical shape with a diameter of 3.0 mm and a height of (granule sizes of 250–400 µm). Then, the granules of the two samples were mixed
3.5 mm (Φ3.0 × 3.5 mm). The catalytic performance of the together by shaking in a vessel. For the preparation of the composite catalyst presented
well-mixed In2O3 and HZSM-5 pellets (In2O3/HZSM-5 mass in Fig. 2e, the In2O3 and HZSM-5 were mixed in an agate mortar for 20 min. Then, the
mixed sample was pressed, crushed and sieved to particles in the range of 40–60 mesh.
ratio = 1:2) was examined with tail-gas recycling (Supplementary
Fig. 11), commonly used in industry for a more efficient utilization Catalyst characterization. Catalysts were characterized by in situ XRD, nitrogen
of the feed39,40. As shown in Fig. 5a, the pellet catalyst exhibited a physisorption, X-ray fluorescence spectroscopy, scanning electron microscopy, TEM

NATURE CHEMISTRY | ADVANCE ONLINE PUBLICATION | www.nature.com/naturechemistry 5

© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
ARTICLES NATURE CHEMISTRY DOI: 10.1038/NCHEM.2794

and high-resolution TEM (HRTEM), H2 , CO2 and ammonia TPD, as well as H2 and 20. Kim, J., Choi, M. & Ryoo, R. Effect of mesoporosity against the deactivation of
CO temperature-programmed reduction techniques. In situ NAP–XPS was MFI zeolite catalyst during the methanol-to-hydrocarbon conversion process.
performed on a SPECS Surface Nano Analysis. The facility is composed of two J. Catal. 269, 219–228 (2010).
chambers, an analysis chamber and a quick ample load–lock chamber. The analysis 21. Liu, Z. Y. et al. Solvent-free synthesis of c-axis oriented ZSM-5 crystals with enhanced
chamber is equipped with a PHOIBOS NAP hemispherical electron energy analyser, methanol to gasoline catalytic activity. ChemCatChem 8, 3317–3322 (2016).
a microfocus monochromatized Al-Kα X-ray source with a beam size of 300 µm, a 22. Olsbye, U. et al. Conversion of methanol to hydrocarbons: how zeolite cavity and
SPECS IQE-11A ion gun and an infrared laser heater. pore size controls product selectivity. Angew. Chem. Int. Ed. 51, 5810–5831 (2012).
23. Olsbye, U., Bjørgen, M., Svelle, S., Lillerud, K.-P. & Kolboe, S. Mechanistic insight
Catalytic evaluation. Activity measurements in the hydrogenation of CO2 were into the methanol-to-hydrocarbons reaction. Catal. Today 106, 108–111 (2005).
carried out in a continuous-flow, high-pressure, fixed-bed reactor (internal diameter 24. Van Speybroeck, V. et al. Advances in theory and their application within the
of 12 mm). Prior to the reaction, the catalyst was pre-treated in situ at 400 °C for 1 h in field of zeolite chemistry. Chem. Soc. Rev. 44, 7044–7111 (2015).
pure Ar (150 ml min−1). After the reactor was cooled down to 340 °C, the reactant gas 25. Khodakov, A. Y., Chu, W. & Fongarland, P. Advances in the development of
mixture with a H2/CO2/N2 ratio of 73/24/3 and a pressure of 3.0 MPa was introduced novel cobalt Fischer–Tropsch catalysts for synthesis of long-chain hydrocarbons
into the reactor. The catalytic reaction for CH3OH conversion was performed in the and clean fuels. Chem. Rev. 107, 1692–1744 (2007).
same reactor. For the experiments with tail-gas recycling, the catalytic performance of 26. Van der Laan, G. P. & & Beenackers, A. A. C. M. Kinetics and selectivity of the
the pellet catalyst (15 g) was investigated in our pilot-scale fixed-bed reactor (internal Fischer–Tropsch synthesis: a literature review. Catal. Rev. 41, 255–318 (1999).
diameter 19 mm, length 1,180 mm (Supplementary Fig. 11)), which was equipped for 27. Xiang, Y. Z. & Kruse, N. Tuning the catalytic CO hydrogenation to straight- and
operation at industrial working conditions with a recycled tail gas. The effluents were long-chain aldehydes/alcohols and olefins/paraffins. Nat. Commun. 7, 13058 (2016).
analysed quantitatively online with a Shimadzu GC-2010C gas chromatograph 28. Behrens, M. et al. Performance improvement of nanocatalysts by promoter-
equipped with thermal conductivity and flame-ionization detectors. The CO2 induced defects in the support material: methanol synthesis over Cu/ZnO:Al.
conversion was calculated by an internal normalization method. The hydrocarbon J. Am. Chem. Soc. 135, 6061–6068 (2013).
distribution presented in this work was calculated on a molar carbon basis. 29. Gao, P. et al. Influence of Zr on the performance of Cu/Zn/Al/Zr catalysts via
hydrotalcite-like precursors for CO2 hydrogenation to methanol. J. Catal. 298,
DFT calculations. Periodic DFT calculations were carried out with the Vienna ab 51–60 (2013).
initio simulation package using the Perdew–Burke–Ernzerhof exchange-correlation 30. Fujiwara, M., Kieffer, R., Ando, H. & Souma, Y. Development of composite
functional and projector augmented wave potentials. catalysts made of Cu–Zn–Cr oxide zeolite for the hydrogenation of carbon-
dioxide. Appl. Catal. A 121, 113–124 (1995).
Data availability. All the data that support the findings of this study are available 31. Jeon, J.-K., Jeong, K.-E., Park, Y.-K. & Ihm, S.-K. Selective synthesis of C3–C4
within the paper and its Supplementary Information files, or from the corresponding hydrocarbons through carbon-dioxide hydrogenation on hybrid catalysts composed
author on reasonable request. of a methanol synthesis catalyst and SAPO. Appl. Catal. A 124, 91–106 (1995).
32. Fujiwara, M., Satake, T., Shiokawa, K. & Sakurai, H. CO2 hydrogenation for C2+
Received 23 January 2017; accepted 5 May 2017; hydrocarbon synthesis over composite catalyst using surface modified HB
published online 12 June 2017 zeolite. Appl. Catal. B 179, 37–43 (2015).
33. Inui, T., Kitagawa, K., Takeguchi, T., Hagiwara, T. & Makino, Y. Hydrogenation
References of carbon-dioxide to C1–C7 hydrocarbons via methanol on composite catalysts.
1. Kothandaraman, J. et al. Amine-free reversible hydrogen storage in formate salts Appl. Catal. A 94, 31–44 (1993).
catalyzed by ruthenium pincer complex without PH control or solvent change. 34. Martin, O. et al. Zinc-rich copper catalysts promoted by gold for methanol
ChemSusChem 8, 1442–1451 (2015). synthesis. ACS Catal. 5, 5607–5616 (2015).
2. Aresta, M., Dibenedetto, A. & Angelini, A. Catalysis for the valorization of 35. Jiao, F. et al. Selective conversion of syngas to light olefins. Science 351,
exhaust carbon: from CO2 to chemicals, materials, and fuels. Technological use 1065–1068 (2016).
of CO2. Chem. Rev. 114, 1709–1742 (2014). 36. Cheng, K. et al. Direct and highly selective conversion of synthesis gas into lower
3. Li, C. W., Ciston, J. & Kanan, M. W. Electroreduction of carbon monoxide to olefins: design of a bifunctional catalyst combining methanol synthesis and
liquid fuel on oxide-derived nanocrystalline copper. Nature 508, 504–507 (2014). carbon–carbon coupling. Angew. Chem. Int. Ed. 55, 4725–4728 (2016).
4. Li, C.-S. et al. High-performance hybrid oxide catalyst of manganese 37. Zecevic, J., Vanbutsele, G., de Jong, K. P. & Martens, J. A. Nanoscale intimacy in
and cobalt for low-pressure methanol synthesis. Nat. Commun. 6, 6538 (2015). bifunctional catalysts for selective conversion of hydrocarbons. Nature 528,
5. Graciani, J. et al. Highly active copper–ceria and copper–ceria–titania catalysts 245–254 (2015).
for methanol synthesis from CO2. Science 345, 546–550 (2014). 38. Fu, X. Z., Clark, L. A., Zeltner, W. A. & Anderson, M. A. Effects of reaction
6. Scirè, S., Crisafulli, C., Maggiore, R., Minicò, S. & Galvagno, S. Influence of the temperature and water vapor content on the heterogeneous photocatalytic
support on CO2 methanation over Ru catalysts: an FT-IR study. Catal. Lett. 51, oxidation of ethylene. J. Photochem. Photobiol. A 97, 181–186 (1996).
41–45 (1998). 39. Lloyd, L. Handbook of Industrial Catalysts (Springer, 2007).
7. Moret, S., Dyson, P. J. & Laurenczy, G. Direct synthesis of formic acid from 40. Subramani, V. & Gangwal, S. K. A review of recent literature to search for an
carbon dioxide by hydrogenation in acidic media. Nat. Commun. 5, 4017 (2014). efficient catalytic process for the conversion of syngas to ethanol. Energy Fuels
8. Studt, F. et al. Discovery of a Ni–Ga catalyst for carbon dioxide reduction to 22, 814–839 (2008).
methanol. Nat. Chem. 6, 320–324 (2014).
9. Gao, S. et al. Partially oxidized atomic cobalt layers for carbon dioxide Acknowledgements
electroreduction to liquid fuel. Nature 529, 68–71 (2016). This work was supported by the National Natural Science Foundation of China (21503260,
10. Yin, Z. et al. Highly selective palladium–copper bimetallic electrocatalysts for the 21573271, 91545112 and 11227902), the Shanghai Municipal Science and Technology
electrochemical reduction of CO2 to CO. Nano Energy 27, 35–43 (2016). Commission, China (14DZ1207602, 16DZ1206900 and 15DZ1170500), the Ministry of
11. Dixneuf, P. H. Bifunctional catalysis: a bridge from CO2 to methanol. Nat. Chem. Science and Technology of China (2016YFA0202802) and the Chinese Academy of Sciences
3, 578–579 (2011). (QYZDB-SSW-SLH035). We thank Z. Liu and Y. Han for the assistance with in situ NAP–
12. Sakakura, T., Choi, J.-C. & Yasuda, H. Transformation of carbon dioxide. Chem. XPS, located at the State Key Laboratory of Functional Materials for Informatics, Shanghai
Rev. 107, 2365–2387 (2007). Institute of Microsystem and Information Technology, Chinese Academy of Sciences.
13. Montoya, J. H., Peterson, A. A. & Nørskov, J. K. Insights into C–C coupling in
CO2 electroreduction on copper electrodes. ChemCatChem 5, 737–742 (2013). Author contributions
14. Aresta, M., Dibenedetto, A. & Angelini, A. The changing paradigm in CO2 P.G., L.Z. and Y.S. conceived the project, analysed the data and wrote the paper. P.G., S.L.
utilization. J. CO2 Util. 3–4, 65–73 (2013). and W.W. drafted the manuscript. P.G. and S.D. prepared the samples. S.L. performed DFT
15. Benson, E. E., Kubiak, C. P., Sathrum, A. J. & Smieja, J. M. Electrocatalytic and calculations. Z.L. and H.W. studied the effect of the integration manner. X.B., M.Q. and C.Y.
homogeneous approaches to conversion of CO2 to liquid fuels. Chem. Soc. Rev. performed the catalytic evaluation. S.D., Z.L., X.B. and J.C. characterized the samples. All
38, 89–99 (2009). the authors discussed the results and commented on the manuscript.
16. Rodemerck, U. et al. Catalyst development for CO2 hydrogenation to fuels.
ChemCatChem 5, 1948–1955 (2013). Additional information
17. Dorner, R. W., Hardy, D. R., Williams, F. W. & Willauer, H. D. Heterogeneous Supplementary information is available in the online version of the paper. Reprints and
catalytic CO2 conversion to value-added hydrocarbons. Energy Environ. Sci. 3, permissions information is available online at www.nature.com/reprints. Publisher’s note:
884–890 (2010). Springer Nature remains neutral with regard to jurisdictional claims in published maps
18. Ye, J. Y., Liu, C. J., Mei, D. H. & Ge, Q. F. Active oxygen vacancy site for and institutional affiliations. Correspondence and requests for materials should be addressed to
methanol synthesis from CO2 hydrogenation on In2O3(110): a DFT study. ACS L.Z. and Y.S.
Catal. 3, 1296–1306 (2013).
19. Martin, O. et al. Indium oxide as a superior catalyst for methanol synthesis by Competing financial interests
CO2 hydrogenation. Angew. Chem. Int. Ed. 55, 6261–6265 (2016). The authors declare no competing financial interests.

6 NATURE CHEMISTRY | ADVANCE ONLINE PUBLICATION | www.nature.com/naturechemistry

© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.

Anda mungkin juga menyukai