You are on page 1of 19

Catalysis

Science &
Technology
View Article Online
MINIREVIEW View Journal | View Issue

A review of the catalytic hydrogenation of carbon


Published on 06 September 2017. Downloaded by University of Winnipeg on 1/23/2019 4:28:10 AM.

Cite this: Catal. Sci. Technol., 2017,


dioxide into value-added hydrocarbons
7, 4580
Haiyan Yang,†a Chen Zhang,†a Peng Gao, *a Hui Wang,a Xiaopeng Li,a
Liangshu Zhong, a Wei Weiab and Yuhan Sun*ab

Chemical utilization of CO2 to chemicals and fuels is very attractive because it can not only alleviate global
warming caused by increasing atmospheric CO2 concentration but also offer a solution to replace dwin-
dling fossil fuels. Hydrogen is a high-energy material and can be used as the reagent for CO2 transforma-
tion. Moreover, when hydrogen originates directly from renewable energy, CO2 hydrogenation can also
provide an important approach for dealing with the intermittence of renewable sources by storing energy
in chemicals and fuels. Therefore, much attention has been paid to CO2 hydrogenation to various value-
added hydrocarbons, such as lower olefins, liquefied petroleum gas, gasoline, aromatics and so on. The fo-
cus of this perspective article is on the indirect and direct routes for production of hydrocarbons from CO2
Received 12th July 2017, hydrogenation and recent developments in catalyst design, catalytic performance and reaction mechanism.
Accepted 5th September 2017
In addition, a brief overview on CO2 hydrogenation to methanol is given, which is a critical process in the
indirect route involving conversion of CO2 into methanol and subsequent transformation into hydrocar-
DOI: 10.1039/c7cy01403a
bons. We also provide an overview of the challenges in and opportunities for future research associated
rsc.li/catalysis with CO2 hydrogenation to value-added hydrocarbons.

1. Introduction tractive because it can not only alleviate global warming


caused by increasing atmospheric CO2 concentration but also
As a nontoxic, renewable and abundant carbon source, chem- offer a solution to replace dwindling fossil fuels.1–3 However,
ical utilization of CO2 into value-added products is very at- CO2 is well known to be a very stable molecule (ΔfG298K =
−396 kJ mol−1), and the biggest obstacle for establishing in-
a
CAS Key Lab of Low-Carbon Conversion Science and Engineering, Shanghai Ad- dustrial processes based on CO2 as a raw material is its low
vanced Research Institute, Chinese Academy of Sciences, No. 99 Haike Road, energy level.4,5 Thus, a large energy input is required to trans-
Zhangjiang Hi-Tech Park, Shanghai 201210, China. E-mail: gaopeng@sari.ac.cn,
form CO2. Hydrogen is a high-energy material and can be
sunyh@sari.ac.cn; Fax: +86 021 20608066; Tel: +86 021 20608002
b
School of Physical Science and Technology, ShanghaiTech University, Shanghai used as the reagent for CO2 transformation. Moreover, re-
201210, China placement of part of the fossil fuel consumption by renew-
† These authors contributed equally to this work. able energy sources (solar, wind, biomass and so on) is a

Haiyan Yang obtained her mas- Peng Gao was born in May 1987
ter's degree in chemical engineer- in China. Dr Gao received his
ing in 2016 from the CAS Key PhD in chemical engineering and
Laboratory of Low-Carbon Con- technology in 2014 from the Uni-
version Science and Engineering, versity of Chinese Academy of
Shanghai Advanced Research In- Sciences. Currently, he is a pro-
stitute, Chinese Academy of Sci- fessor at the CAS Key Lab of
ences. She is interested in de- Low-Carbon Conversion Science
signing and synthesizing novel and Engineering, Shanghai Ad-
nanocatalysts for catalytic con- vanced Research Institute, Chi-
version of carbon dioxide or syn- nese Academy of Sciences. His
gas into value-added products. research interests include CO2
Haiyan Yang Peng Gao conversion to chemicals and
fuels, Fischer–Tropsch synthesis
and well-defined heterogeneous catalysts.

4580 | Catal. Sci. Technol., 2017, 7, 4580–4598 This journal is © The Royal Society of Chemistry 2017
View Article Online

Catalysis Science & Technology Minireview

central strategy for resource and energy efficiency. When hydrogen


originates directly from renewable energy, CO2 hydrogenation can
also provide an important approach for dealing with the intermit-
tence of renewable sources by storing energy in chemicals and
fuels. Therefore, much attention has been paid to CO2 hydrogena-
tion to various value-added hydrocarbons, such as liquefied petro-
leum gas (LPG, C30–C40), lower olefins (ethylene, propylene and
butylene, C2=–C4=), gasoline (C5–C11), aromatics and so on.
Published on 06 September 2017. Downloaded by University of Winnipeg on 1/23/2019 4:28:10 AM.

There are different possible routes to produce hydrocar-


bons from CO2 hydrogenation (Fig. 1). Two parallel reactions,
Fig. 1 Details of the different possible routes to synthesize value-
the synthesis of methanol (eqn (1)) and reverse water-gas
added hydrocarbons (olefin, LPG, gasoline, aromatics and so on) from
shift (eqn (2), RWGS), are typically present in the CO2 hydro- CO2 hydrogenation.
genation process. The hydrocarbons could be directly pro-
duced from syngas (CO + H2) based on Fischer–Tropsch syn-
thesis (FTS) or indirectly via industrial methanol synthesis and methanol-to-gasoline (MTG) processes. It had also been
and then converting methanol to a range of hydrocarbons reported that aromatic or lower paraffin (LPG) hydrocarbons
using the methanol-to-hydrocarbon (MTH) process, including were synthesized from methanol or dimethyl ether (DME).6–8
the methanol-to-olefin (MTO), methanol-to-propene (MTP), In addition, various hydrocarbons could be indirectly pro-
duced from CO2 hydrogenation via CO2 hydrogenation to
methanol and MTH reactions. Recently, the direct route of
Hui Wang obtained her BS de- converting CO2 into hydrocarbons has been developed based
gree in chemical engineering on FTS (CO2-FTS), via a two-step process with an initial re-
from Hunan Normal University. duction of CO2 to CO via the RWGS reaction followed by the
She obtained her PhD degree in conversion of CO to hydrocarbons via FTS. The catalyst used
physical chemistry from the In- should be active in both RWGS and FTS reactions. It is also
stitute of Coal Chemistry, CAS, possible to combine the catalysts for methanol and the zeo-
in 2006. She has been an assis- lites for MTH to have a direct one-step formation of hydrocar-
tant professor in Shanghai Uni- bons from CO2 hydrogenation.
versity from 2006 to 2013, and
was promoted to professor in CO2 + 3H2 ⇌ CH3OH + H2O ΔrH298K = − 49.5 kJ mol−1 (1)
Shanghai Advanced Research In-
stitute, CAS, in 2013. Her main CO2 + H2 ⇌ CO + H2O ΔrH298K = 41.2 kJ mol−1 (2)
Hui Wang scientific interests are conversion
of carbon dioxide and syngas. Many excellent reviews have discussed the conversion of
She was nominated for the “Outstanding Youth Contribution” CO2 as well as the catalytic hydrogenation of CO2.3,4,9–17 How-
award in the 15th International Congress on Carbon Dioxide Uti- ever, most of the reviews mainly focus on the general aspects
lization (2017, China). of CO2 applications and CO2 hydrogenation to various C1

Wei Wei is a professor, institute Yuhan Sun received his PhD


direct assistant and director of from the Institute of Coal Chem-
the Center for Greenhouse Gas istry of the Chinese Academy of
and Environmental Engineering Sciences in 1989. He is now the
at Shanghai Advanced Research vice president of the preparatory
Institute, CAS. His main re- committee of Shanghai Advanced
search focus includes CO2 ac- Research Institute, CAS. Prof
counting, capture, utilization Sun's main research interests in-
and storage, green chemistry clude C1 chemistry on coal/natu-
and environmental catalysis. ral gas-based synfuels and
Prof Wei received the “Outstand- chemicals, catalysis and engi-
ing Young Scientist” award in neering for CO2 utilization, ap-
Wei Wei the 13th International Congress Yuhan Sun plication of nano-materials in
on Catalysis (2004, France) and green chemistry and optical
was nominated for the “Outstanding Youth Contribution” award films, etc. Prof Sun is also on the board of the Chemistry Society
in the 10th International Congress on Carbon Dioxide Utilization of China and that of the Particle Society of China.
(2009, China).

This journal is © The Royal Society of Chemistry 2017 Catal. Sci. Technol., 2017, 7, 4580–4598 | 4581
View Article Online

Minireview Catalysis Science & Technology

feedstocks (for example, methane, methanol, carbon monox- methanol catalysts for CO hydrogenation, and Cu species
ide, formic acid, etc.). Advances have been made in the past constitute the main active component of the catalytic system;
decade, especially on the conversion of CO2 into C1 (ii) catalysts with noble metals such as Pd and Pt as active
chemicals via heterogeneous catalysts. Due to the extreme in- components; (iii) oxygen-deficient materials used as active
ertness of CO2 and a high C–C coupling barrier, it is still a sites. The reaction mechanism of this kind of catalyst is dif-
challenge to synthesize C2+ hydrocarbons (hydrocarbons with ferent from those of the two kinds of catalytic systems men-
more than two carbons) from CO2. Recently, more and more tioned above, for example, the In2O3/ZrO2 catalytic system.
researchers have focused on this area. Therefore, this crucial 2.1.1. Cu-based catalysts. In view of the significantly lower
Published on 06 September 2017. Downloaded by University of Winnipeg on 1/23/2019 4:28:10 AM.

review tries to discuss the various approaches ultimately lead- selectivity to methanol, catalytic activity and stability, copper
ing to production of hydrocarbons from CO2 hydrogenation alone is not efficient in the synthesis of methanol from
and provide a current understanding of catalyst design, cata- CO2.22,23 An appropriate support not only affects the forma-
lytic performance and reaction mechanism over various types tion and stabilization of the active phase of the catalysts but
of heterogeneous catalysts with an emphasis on practical as- also is capable of tuning the interactions between the major
pects. Additionally, a brief overview on the recent develop- component and the promoter. Whether as a catalyst for CO2
ments in CO2 hydrogenation to methanol is given because hydrogenation to methanol in industry or as a model catalyst
methanol can be subsequently transformed into hydrocar- for theoretical research, Cu/ZnO and further modified cata-
bons over various zeolites using the MTH process in the indi- lysts remain the focus of current research.24,25 The main rea-
rect route. sons for this are that ZnO can improve the dispersion and
stabilization of copper, and the lattice oxygen vacancies and
2. Indirect conversion of CO2 into an electron pair in ZnO are active for methanol synthesis.26
More recently, the formation and properties of Cu–ZnOx
hydrocarbons active sites have been discussed in more depth. Lunkenbein
The two main indirect routes towards CO2 hydrogenation et al. utilized a thermal stimulus using an electron beam to
into hydrocarbons involve (i) conversion of CO2 into metha- transform the industrial Cu/ZnO catalyst structure into Cu
nol and subsequent transformation into hydrocarbons (ole- NPs surrounded by metastable “graphitic-like” ZnO after re-
fins, LPG, gasoline, aromatics and so on) in a separate stage ductive activation.27 The GL ZnO is kinetically stabilized by
and (ii) conversion of CO2 into CO via RWGS; hydrocarbons interacting with the defective and curved Cu surface under-
are then synthesized using the modified FTS process based neath. Defects could play a key role in the stability of the
on two-stage reactors. Generally, the catalysts active in metha- ZnO overlayer and may lead to the presence of some ZnOx
nol synthesis or FTS reactions are also active in RWGS.3,15 species acting as a cocatalyst in methanol synthesis, illustrat-
Thus, this step is integrated in these processes and will be ing that the migration of Zn to the Cu surface is closely re-
discussed in related sections. lated to the formation of a stable Cu–ZnO interface and Cu–
ZnOx active sites in the process of reduction and evaluation.
Similar conclusions are emerging in Tisseraud's latest re-
2.1. CO2 hydrogenation to methanol search, wherein he designed a core–shell Cu@ZnOx catalyst
Beginning in the 1920s, the production of methanol from that allows maximizing the Cu–ZnO contacts and favors a
syngas on an industrial scale was introduced by BASF in Ger- large diffusion of Zn into Cu to cover the metallic Cu surface
many. The Cu/ZnO/Al2O3 catalyst system is employed in the with the oxygen-deficient CuxZn(1−x)Oy active phase for metha-
industrial synthesis of methanol at approximately 6 MPa and nol (Fig. 2).28 Sun et al. also found that a Cu–Zn active nano-
250 °C, and has been widely studied for over 40 years in in- alloy layer was formed on the Cu–ZnO nanoclusters, which
dustry.18 As an alternative feedstock, CO2 replaced by CO is reduced the activation energy of CO adsorption. The surface
considered as an effective way for CO2 utilization in metha- alloy layer and Cu–ZnO sites play a cooperative catalytic role
nol production.19 Different from methanol synthesis from in dimethyl ether synthesis via syngas.29 The experimental
syngas, the formation of water vapor is inevitable in direct and geometrical approach also proved that the methanol for-
CO2 hydrogenation, which inhibits the reaction strongly and mation rate is a linear function of the Zn that has migrated
leads to serious catalyst deactivation.20,21 In addition, other into Cu. In summary, it is widely accepted that the active site
by-products such as CO and hydrocarbons are formed during for methanol synthesis is a Cu–Zn surface alloy and a
the hydrogenation of CO2. Therefore, an efficient catalyst is CuxZn(1−x)Oy active phase formed by atomic diffusion of Zn
needed to improve catalytic stability and avoid the formation from ZnO to the metallic Cu surface.
of undesired by-products for methanol synthesis, which re- To further increase the activity and stability of the Cu/ZnO
quires continuous optimization of the catalyst system or de- catalyst, different modifiers are used as stabilizers and pro-
velopment of a new catalyst system. moters. Toyir et al. found that the promoting effect of Ga2O3
A wide variety of heterogeneous catalysts have been evalu- was strongly associated with Ga2O3 particle size.30 Small
ated in the conversion of CO2 and H2 into methanol. Gener- Ga2O3 particles favor the formation of an intermediate state
ally, methanol catalysts can be classified into three catego- of copper between Cu0 and Cu2+, probably Cu+. Li et al. pro-
ries: (i) Cu-based catalysts, which are mainly modified posed that the introduction of Ga3+ into the Cu/ZnO catalyst

4582 | Catal. Sci. Technol., 2017, 7, 4580–4598 This journal is © The Royal Society of Chemistry 2017
View Article Online

Catalysis Science & Technology Minireview

motion of copper atoms into the ZrO2 lattice and formation


of stable Cu+ (Fig. 3). Additionally, the incorporation of Cu+
or Cu2+ ions into the ZrO2 network compensated for the neg-
ative charge of vacancies and further contributed to the stabi-
lization of the t-ZrO2. The Cu–ZrO2 interface site containing
oxygen vacancies for methanol formation derived from this
enhanced Cu/t-ZrO2 interaction would improve the yield and
selectivity of methanol.
Published on 06 September 2017. Downloaded by University of Winnipeg on 1/23/2019 4:28:10 AM.

2.1.2. Noble metal catalysts. In the noble metal series, Pd-


based catalysts are the most commonly used owing to their
considerable activity and selectivity for hydrogenation of CO2
to methanol. Bahruji et al. investigated structure activity–rela-
tionships for CO2 hydrogenation on Pd/ZnO catalysts pre-
pared by different methods. The results proved that the PdZn
Fig. 2 General scheme for methanol synthesis over Cu/ZnO-based alloy formed during the reaction or high temperature pre-
catalysts.28 reduction greatly reduced CO production by RWGS reaction
and became the active site for methanol synthesis (Fig. 4).36
In addition, the smaller PdZn alloy particle size, less surface
precursor facilitated the deep reduction of ZnO to Zn0 by the metallic Pd content and higher surface area of PdZn alloy
establishment of an electronic heterojunction of ZnO– were beneficial to increase the methanol selectivity. Hartadi
MGa2O4 (M = Zn or Cu).31 The reduced Zn0 when in contact et al. first explored the effect of total pressure and influence
with Cu nanoparticles can form CuZn bimetallic species. of CO on Au/ZnO catalysts.37 Compared with industrial Cu/
Therefore, doping with Ga3+ promoted the formation of a ZnO/Al2O3 catalysts, high temperature and high pressure were
CuZn bimetallic active phase at the interface to enhance the beneficial to the increase in methanol yield, inhibiting the
activity and selectivity to methanol. Martin et al. discovered relative activity of the RWGS reaction and improving the se-
that the introduction of small amounts of a noble metal like lectivity to methanol and conversion of CO2. Increasing the
Au significantly promoted the electronic stabilization of Cu0 CO concentration resulted in the decreasing of methanol for-
and stabilized the interface between Cu and ZnO.32 mation rate, and the rate of CO2 hydrogenation (in CO2/H2
Because of the high stability under reducing or oxidizing mixtures) on Au/ZnO was considerably higher than that of
atmospheres, zirconia has also been deemed as an excellent CO hydrogenation (in CO/H2 mixtures). Therefore, CO2 is ‘di-
promoter or support for the methanol synthesis catalyst.33,34 rectly’ hydrogenated to CH3OH and not converted to CO in
ZrO2 added on Cu-based catalysts could increase the surface the first step, followed by hydrogenation of adsorbed CO.
area and improve Cu particle dispersion. Moreover, the crys- Frauenheim and Xiao investigated CO2 hydrogenation on
tal types of Zr influence the performance of the catalyst. Sam- AgIJ111)/ZnOIJ0001 _) surfaces via first principles calculations.38
son et al. presented that ZrO2 is of great importance in the The calculation results showed that the addition of
catalytic performance of Cu/ZrO2.35 The results showed that ZnOIJ0001 _) substrate to Ag(111) made the Ag's d-band states
only the oxygen vacancy in t-ZrO2 was conducive to the pro- exceed the Fermi level. The binding ability between the

Fig. 3 Methanol formation rate as a function of t-ZrO2 content.35

This journal is © The Royal Society of Chemistry 2017 Catal. Sci. Technol., 2017, 7, 4580–4598 | 4583
View Article Online

Minireview Catalysis Science & Technology


Published on 06 September 2017. Downloaded by University of Winnipeg on 1/23/2019 4:28:10 AM.

Fig. 4 CO2 conversion and CH3OH and CO selectivity as a function of reaction temperature (150–300 °C) on 5% Pd ZnO catalysts prepared by
the sol immobilisation method. The catalyst was reduced at (a) 150 °C to form Pd/ZnO and at (b) 400 °C to form the PdZn/ZnO catalyst.36

supported Ag(111) monolayer and the CO2 molecule signifi- CH4 stems from the presence of the Ni-rich Ni3Ga phase or
cantly improved, and CO2 reduction in the subsequent hydro- small pure Ni particles or it is produced on the adjacent Ni–
genation process was optimized as well. Therefore, the strong Ni sites on the surface of Ni5Ga3 still remains to be clarified.
metal–support interaction between Ag and ZnO is of para- 2.1.3. Other catalytic systems. Apart from the traditional
mount importance to increase the catalytic activity. It was Cu-based catalysts and noble metal catalysts, many CO2 hy-
proved that Ga can also act as an active component for meth- drogenation catalysts based on new reaction mechanisms
anol synthesis from CO2 hydrogenation. Studt et al. and and catalytic structures have been developed in recent years.
Sharafutdinov et al. presented a detailed study of the rela- Ge et al. investigated methanol synthesis from CO2 hydroge-
tionship between active components and product distribu- nation on the defective In2O3IJ110) surface with surface oxy-
tion in intermetallic Ni–Ga catalysts.39,40 Their results show gen vacancies via periodic DFT calculations (Fig. 5).41 The
that the Ni5Ga3 phase is favorable for methane generation surprising calculation results show that the methanol forma-
and increasing methanol selectivity. However, whether the tion mechanism on the D4 surface with the Ov4 oxygen

Fig. 5 Potential energy surfaces for CO2 hydrogenation to methanol on the D4 defective In2O3IJ110) surface.41

4584 | Catal. Sci. Technol., 2017, 7, 4580–4598 This journal is © The Royal Society of Chemistry 2017
View Article Online

Catalysis Science & Technology Minireview

defective site is quite different from that of copper-based cat- C2–C4 hydrocarbons. A number of attempts were devoted to
alysts mentioned earlier, and the Ov4 oxygen defective site is selectively synthesizing light olefins from methanol, not only
found to be more favorable for CO2 activation and hydroge- on medium-pore zeolites but also on small-pore zeolites. A
nation. Typically, the Ov4 oxygen defective site on the range of catalysts are employed, especially with respect to
In2O3IJ110) surface assists CO2 activation and hydrogenation ZSM-5 (MTG directed) and SAPO-34 molecular sieves (MTO
and also stabilizes the key intermediates involved in metha- directed). SAPO-34 molecular sieves exhibit great potential to
nol formation. Martin et al. prepared the In2O3/ZrO2 catalyst obtain high selectivity for light olefins owing to their mild
and achieved the methanol generation mechanism proposed acidity. The preparation conditions, the size and structure of
by Ge et al. successfully.42 ZrO2 was added to make and stabi-
Published on 06 September 2017. Downloaded by University of Winnipeg on 1/23/2019 4:28:10 AM.

the pores, and cation exchange as well as the distribution of


lize the most thermodynamically unstable Ov4 oxygen defec- acid sites are of mounting importance for the catalytic perfor-
tive site cycle. The catalysts show excellent methanol selectiv- mance for synthesizing methanol to lower olefins.47,48 In
ity (∼100%) and stability (1000 h) in the hydrogenation of 2009, a semi-commercial demonstration unit in Feluy, Bel-
CO2 to methanol (Fig. 6). This strongly contrasts with the gium, processing up to 10 tonnes per day of methanol feed
conventional Cu/ZnO/Al2O3 catalyst, which is unselective and was brought on-stream, and in 2011 the construction of a
experiences rapid deactivation. 295 kt per year plant in Nanjing, China, was announced. Re-
searchers at the Dalian Institute of Chemical Physics devel-
oped a similar process using SAPO-34, and a 600 kt per year
2.2. Methanol to hydrocarbons (ethene + propene) plant was started in 2010.49
The MTH process over acid zeolite catalysts, first discovered The MTG process is also among the MTH technologies
by Mobil Research Laboratories in 1976, has been extensively with industrial interest. Catalysts based on H-ZSM-5 or its
investigated in recent years.43 Methanol or its dehydration modified analogs were the main materials given preference
product dimethyl ether (DME) can be used as a feed to pro- by the researchers, with numerous literature studies docu-
duce several different classes of hydrocarbons, including mented. In 1985, the MTG process was commercialized in
lower olefins, gasoline-range hydrocarbons, branched al- New Zealand, where Mobil, in a partnership with the New
kanes, and aromatics. The selectivity to any of these classes Zealand government, built a 14 500 bpd plant based on natu-
of compounds is determined both by the zeolite topology and ral gas converted through synthesis gas into methanol. Issues
the operating conditions used. The key of the MTH process is related to the role of catalyst topology, acidity properties, re-
to control product selectivity and increase the catalytic stabil- action parameters and their effects on catalytic activity, selec-
ity by developing high-performance catalysts and optimizing tivity and catalyst lifetime have been critically covered in the
reaction conditions. Recently, the catalytic materials, mecha- literature review by Galadima and Muraza.50
nism, active intermediates and deactivation and commercial 2.2.2. Reaction mechanism. Mechanistic studies have been
projects of MTH have been reviewed in detail.8,44–46 There- at the core of MTH research. Several reviews that extensively
fore, the catalyst and reaction mechanism of MTH are briefly cover early research on the mechanism can be found in the
introduced in the following. literature.8,46,47,51 Stöcker highlights the milestones of this
2.2.1. Catalysts. The behaviors of zeolite catalysts are development during the past decade before 1990.47 He fo-
based on their shape-selectivity, dimensional structure, sta- cused mainly on the chemistry and mechanism of these reac-
bility and acidity properties with the possibility of modifica- tions, especially the possible mechanistic proposals for the
tion under controlled conditions. Literature data have shown formation of the first C–C bond. He also briefly summarized
that medium-pore zeolite/microporous materials generate the relevant mechanism, such as the oxonium ylide mecha-
C5–C11 hydrocarbons, while small-pore molecular sieves yield nism, carbine mechanism, carbocationic mechanism, free

Fig. 6 (a) Methanol STY and selectivity for CO2 hydrogenation over bulk In2O3, In2O3/ZrO2 (9 wt% In), and the benchmark Cu/ZnO/Al2O3 catalyst
at various temperatures. (b) Evolution of the methanol STY with time on stream (TOS) over In2O3/ZrO2 and Cu/ZnO/Al2O3.42

This journal is © The Royal Society of Chemistry 2017 Catal. Sci. Technol., 2017, 7, 4580–4598 | 4585
View Article Online

Minireview Catalysis Science & Technology

radical mechanism, hydrocarbon pool mechanism and so on.


Among these, the hydrocarbon pool mechanism has been
widely accepted (Fig. 7).8,51–53

2.3. Modified Fischer–Tropsch synthesis


A typical FTS is the catalytic conversion of syngas to a mix-
ture of predominantly linear alkanes and alkenes (eqn (3)
Published on 06 September 2017. Downloaded by University of Winnipeg on 1/23/2019 4:28:10 AM.

and (4)), and a wide distribution of hydrocarbons with differ-


ent chain lengths is produced, which can be described by the
Anderson–Schulz–Flory (ASF) model (Fig. 8).20,21,54,55 Lower
olefins (C2=–C4=) – generally referring to ethylene, propylene
and butylene – could be directly produced from syngas, the
so-called FTO or Fischer–Tropsch to olefin process.56 The ASF
model predicts a maximum C2–C4 hydrocarbon fraction of
about 56.7%, including C2=–C4= olefins and C20–C40 paraffins, Fig. 8 Product distribution according to the Anderson–Schulz–Flory
and an undesired methane fraction of about 29.2%.56,57 (ASF) model.
Therefore, iron and cobalt based catalysts, which are of com-
mercial interest in FTS, should be modified in order to mini-
mize the formation of alkanes (especially CH4) and increase lites for hydrocracking/isomerization reactions turned out to
selectivity to C2–C4 olefins. Recently, it was reported that be promising for the direct production of gasoline-range hy-
iron-based Fischer–Tropsch catalysts promoted by sulfur and drocarbons from syngas.60 The selectivity for C5–C11 hydro-
sodium exhibited excellent selective formation of lower ole- carbons could reach about 70% with a ratio of isoparaffins to
fins (61%) at 340 °C, 2 MPa, and H2/CO = 1.58 Zhong et al. n-paraffins of approximately 2.3 over this catalyst. This C5–
showed that under mild reaction conditions (250 °C, 0.1 C11 selectivity is far beyond the maximum value (around
MPa, H2/CO = 2), cobalt carbide quadrangular nanoprisms 45%) expected from the ASF distribution. Additionally, Plana-
catalyzed the FTO conversion of syngas with high selectivity Pallejà et al. investigated the effect of acidic properties on
for the production of lower olefins (60.8%) while generating ZSM-5 zeolites to evaluate changes in product selectivity dur-
little methane (5%), which deviated markedly from the classi- ing FTS over a physical mixture of the iron-based FTS catalyst
cal ASF distribution.57 They have demonstrated that preferen- and zeolites with different Si/Al ratios.61 Higher acidity (lower
tially exposed {101} and {020} facets play a critical role during Si/Al ratio) in the pristine zeolite was responsible for cracking
syngas conversion in that they favor olefin production and in- of heavy hydrocarbons and the formation of heavier aromatic
hibit methane formation (Fig. 9). products, while a higher Si/Al ratio resulted in a higher selec-
tivity towards gasoline-range products (isoparaffins and
Parrafins: nCO + (2n + 1)H2 → CnH2n+2 + nH2O (3) olefins).

Olefins: nCO + 2nH2 → CnH2n + nH2O (4) 3. Direct conversion of CO2 into
The high-octane gasoline components (C5–C11), which con- hydrocarbons
sist of (predominantly iso-) paraffins, aromatics (predomi- CO2 can be also hydrogenated to hydrocarbons by a direct
nantly methyl-substituted), naphthenes, and olefins, can also route. Compared with the indirect route, the direct route is
be obtained from FTS. Bifunctional FTS catalysts that couple more economical and environmentally benign. The direct
Co nanoparticles for CO hydrogenation and mesoporous zeo- CO2 hydrogenation (CO2-FTS) is often described as the com-
bination of the reduction of CO2 to CO via RWGS reaction
and subsequent hydrogenation of CO to hydrocarbons via
FTS. More recently, the direct and highly selective conversion
of CO2 into value-added hydrocarbons can be achieved by
combining the catalysts for CO2 hydrogenation to methanol
with the zeolites for MTH.

3.1. CO2-based Fischer–Tropsch synthesis via CO


intermediates
A highly attractive route of direct CO2 hydrogenation has
Fig. 7 Mechanism of the catalytic conversion of methanol to been postulated to be a two-step process (RWGS and FTS)
hydrocarbons.8 over the same catalyst.62–65 Thus, the CO2-FTS reaction

4586 | Catal. Sci. Technol., 2017, 7, 4580–4598 This journal is © The Royal Society of Chemistry 2017
View Article Online

Catalysis Science & Technology Minireview


Published on 06 September 2017. Downloaded by University of Winnipeg on 1/23/2019 4:28:10 AM.

Fig. 9 TEM characterization of the CoMn catalysts at the steady stage of reaction. (a and b) Low-resolution TEM images; (c and e) high-resolution
images of Co2C nanoprisms with exposed facets of (101), (−101) and (020); (d) distance (length) of the lattice fringes; (f) the Co2C nanoprism model
with four rectangular faces and two rhomboid faces.59

requires the presence of a bifunctional catalyst that combines Among them, iron-based and cobalt-based catalysts are
RWGS with Fischer–Tropsch chain growth activity. Compared widely used in the CO2-FTS reaction owing to significant frac-
with CO hydrogenation, the degree of hydrogenation of tions of C2+ hydrocarbons produced on them.
surface-adsorbed intermediates in CO2-based FTS is higher 3.1.1. Iron-based catalysts. Since the products derived
due to the slower adsorption rate of CO2, leading to lower from FTS are endowed with a tremendous environmental
CO2 conversion and more ready formation of a fairly large value, CO2-FTS with iron catalysts has been addressed in a
fraction of CH4. Consequently, the main challenge is associ- number of investigations. Iron catalysts have showed great
ated with the development of catalysts not only with high ac- promise in converting CO to value-added hydrocarbons and
tivity but also with high selectivity to value-added hydrocar- proved to be the preferred catalyst candidate for the CO2-FTS
bon products, such as short-chain olefins and long-chain process, attributed to their excellent ability to catalyze both
hydrocarbons. RWGS and FTS processes as well as the highly olefinic nature
Catalysts used with CO2 hydrogenation have been based of the yield products.67 The range of conversion levels and se-
on catalysts for CO hydrogenation so far, and catalyst compo- lectivity of production can be explained by the different prep-
nent is analogous to that for FTS. Different kinds of FTS cata- aration techniques, different promoters, supports and experi-
lysts with different catalytic selectivities have been developed mental conditions.
for the selective production of various hydrocarbons. Cata- Promoters. Despite the fact that iron-based catalysts are
lysts such as Fe, Co, Ru and Ni supported on SiO2, γ-Al2O3, able to transform CO2 with a much higher affinity for long-
TiO2, and carbon nanomaterials have been comprehensively chain hydrocarbons, it should be pointed out that the
investigated in the CO2 conversion to hydrocarbons.62,66 unpromoted catalysts still present a high selectivity to non-

This journal is © The Royal Society of Chemistry 2017 Catal. Sci. Technol., 2017, 7, 4580–4598 | 4587
View Article Online

Minireview Catalysis Science & Technology

desired products such as CH4 and may deactivate rapidly.65,67 reaction conditions of 320 °C, 30 bar, H2/CO2 = 1, and 4000
Therefore, promoters urgently need to fine-tune and optimize ml h−1 gcat−1. The Na–Fe3O4/HZSM-5 multifunctional catalyst
the product distribution and enhance the stability of cata- enables RWGS over Fe3O4 sites, α-olefin synthesis over Fe5C2
lysts. A variety of promoters incorporated onto iron-catalysts sites, and oligomerization/aromatization/isomerization over
for hydrocarbon formation from CO2 hydrogenation have zeolite acid sites (Fig. 10).
been reported.68–74 Copper has also been widely used to modify iron-cata-
It is well known that CO2 conversion is favored by alkaline lysts.70,77,83,84 Herranz et al. indicated that addition of Cu
addition. It has been reported that potassium, as a promoter made reducibility of the iron oxide more facile, leading to a
Published on 06 September 2017. Downloaded by University of Winnipeg on 1/23/2019 4:28:10 AM.

for iron-based catalysts, influenced significantly the catalyst's significant increment in CO2 conversion without altering
overall performance.65,68,75–77 The incorporation of potassium product selectivity.84 The reason for the effects may be that
not only increases the CO2 conversion but also leads to a signif- the introduction of Cu is able to provide active sites for
icant shift in olefin production as well as longer chain hydro- dissociatively adsorbing hydrogen once reduced to its metal-
carbons. Visconti et al. investigated the effect of K-loading on lic form. Recently, Choi et al. reported a new catalyst, which
the activity and the selectivity of the adopted iron catalyst.78 was prepared by reduction of delafossite CuFeO2 followed by
They noticed that high K-loadings made possible the chain in situ carburization to its active phase (Hägg carbide).85 The
growth process and the formation of primary olefins. It is de- catalytic performances of various catalysts derived from
duced that the remarkable catalytic performance properties in Fe2O3, CuFe2O4, and CuFeO2-12 precursors are provided in
the presence of K are attributable to the increase in the ability Fig. 11. It was obvious that the catalyst derived from CuFeO2-
of CO2 chemisorption while impeding H2 adsorption, which in- 12 not only produced C5+ hydrocarbons with a high selectivity
creases the probability of C–C bond formation.68 Wang et al. of 65% but also greatly restrained the formation of CH4 to 2–
investigated the effects of modification by alkali metals like Li, 3%, which reflected a record-breaking selectivity for CO2
Na, K, Cs on the performance of CO2 hydrogenation; they ob- hydrogenation.
served that the K-modified Fe/ZrO2 catalyst with an appropriate Incorporation of manganese into iron-catalysts is able to
K content (0.5–1.0 wt%) outperformed other alkali metals in suppress the formation of CH4 while increasing the ratio of
the synthesis of C2–C4.79 In a word, potassium is an effective olefin/paraffin during CO2 hydrogenation.69,71 It was reported
promoter for the iron catalyst in the enhancement of selectivity that Mn promoted catalyst reduction, dispersion and carburi-
to olefins and long-chain hydrocarbons. zation of the precursor as well as increased the catalyst's sur-
The addition of sodium is essential to tailor the olefin pro- face basicity. However, the positive effect of Mn is valid only
duction, as commented before,73,80,81 since it can obviously in a limited concentration range due to the possible blockage
promote the surface basicity of the catalysts, which is favored of active iron sites at high Mn loadings.71 Therefore, the
for olefin production. In addition, increasing the Na content amount of Mn employed on catalysts should be appropriate.
can greatly improve the carbonization degrees of the iron cat- The positive role of ceria is connected with its good low-
alysts.73,81 Recently, Wei et al. also found that the combina- temperature RWGS activity. The size of CeO2 domains and
tion of a Na–Fe3O4 CO2-FTS catalyst with an HZSM-5 zeolite the order of catalyst impregnation with ceria influence the ac-
into a multifunctional catalyst can shift product distribution tivity and selectivity towards C2–C5 olefins.86,87 For the Fe/
towards high-octane gasoline-range isoparaffins and aro- Mn/K catalysts, promotion by Ce results in an about 22% in-
matics.82 The selectivity for gasoline-range (C5–C11) hydrocar- crease in CO2 conversion and 5% increase in olefin on ac-
bons among all hydrocarbons was up to 78% at a CO2 conver- count of a promotion of RWGS activity compared to the cata-
sion of 22%, while methane selectivity was 4% under the lysts without Ce.86

Fig. 10 Reaction scheme for CO2 hydrogenation to gasoline-range hydrocarbons.82

4588 | Catal. Sci. Technol., 2017, 7, 4580–4598 This journal is © The Royal Society of Chemistry 2017
View Article Online

Catalysis Science & Technology Minireview


Published on 06 September 2017. Downloaded by University of Winnipeg on 1/23/2019 4:28:10 AM.

Fig. 11 (a) Fe time yield (FTY) against time on stream for catalysts derived from Fe2O3, CuFe2O4, and CuFeO2-12 precursors (reaction conditions:
300 °C, 10 bar, 1800 ml h−1 gcat−1, H2/CO2 = 3). (b) Product selectivity on CO-free basis.85

Some other promoters, like Cr and Zn, have also been ap- constitution and structure of the iron catalyst changes from
plied to the iron catalysts. It was reported that promotion of the initial synthesis run, Riedel et al. concluded that the pro-
Fe with Cr and Zn increased the conversion of CO2.76 More- cess of FTS was composed of episodes with distinct kinetic
over, Zn could increase the surface and improve the basicity regimes, and these episodes were closely associated with the
of catalysts as well as offer high selectivity for C2–C4 catalyst transformations in structure and composition, as
alkenes.70,88 shown in Fig. 12.92 The iron phases of the reduced catalyst
Supports. Supporting materials play crucial roles in CO2 before reaction mainly consist of α-Fe and Fe3O4. As time
hydrogenation to hydrocarbons. Iron-based catalysts dis- goes by, the Fe3O4 and Fe2O3 phases are consumed and a
persed on various supports, such as γ-Al2O3, SiO2, TiO2, ZrO2, new amorphous, probably oxidic iron phase emerges, which
and MOFs have been examined extensively.75,76,89 Among the appears to be active for the RWGS reaction. FTS activity starts
supports used in iron catalysts, many researchers considered with the formation of Fe5C2, which they considered is respon-
that γ-Al2O3 performs best (followed by SiO2 and TiO2) be- sible for the formation of olefins and long-chain hydrocar-
cause of the strong metal–support interaction formed with bons. Beyond that, the investigation of the correlation be-
the active component, and it could avoid sintering of active tween the rate of FTS and the percentage of Fe carbide also
particles during reaction.75,76 However, others hold different indicates that iron carbide is the active phase for CO2-FTS.93
views; their studies revealed that ZrO2 provided the highest 3.1.2. Cobalt-based catalysts. Compared to the iron-based
selectivity and yielded lower olefins among a range of sup- catalysts, cobalt catalysts represent the optimal choice for the
ports (SiO2, Al2O3, TiO2, ZrO2, and CNT).79 In addition, a de- synthesis of heavier hydrocarbons from syngas and tend to
tailed study of supports revealed that promoted Fe supported have a much higher chain growth potential owing to their
on TiO2 tended to produce less undesired CH4 than Al2O3- ability to be substantially inactive in the water gas-shift
supported catalysts. Recently, some other supports have also
been introduced, Hu et al. reported that Fe-supported K-
OMS-2 nanocatalysts had the highest selectivity to light
olefins.90
Active sites. In spite of the progress in CO2-based FTS over
iron catalysts, the nature of the active sites of Fe-based cata-
lysts is still in dispute. Either the magnetite phase or the ox-
idic phase of iron is essential for the RWGS reaction, while
the carbide form of iron is responsible for chain propagation.
To assess the nature of iron phases, density functional theory
calculations were applied to Fe-based catalysts by Hakim and
co-workers, and their results have demonstrated that the
Fe3O4IJ111) surface was very capable of activating CO2.91 The
same results were obtained by Wei et al. They observed that
CO2 was initially reduced to CO by H2 via RWGS on Fe3O4
sites followed by a subsequent hydrogenation of CO to
α-olefins via FTS on Fe5C2 sites, as commented before.82
Since iron carbides are visualized as the “true Fischer– Fig. 12 Iron-phase composition as a function of time during
Tropsch catalyst”, much effort has been devoted to identify hydrocarbon synthesis on Fe/Al/Cu catalyst. Reaction conditions: 250
the iron carbides during CO2-FTS.76,79,92,93 By studying the °C, 1 MPa, 1800 ml(NTP) h−1 g(Fe)−1, H2/CO2 = 3.92

This journal is © The Royal Society of Chemistry 2017 Catal. Sci. Technol., 2017, 7, 4580–4598 | 4589
View Article Online

Minireview Catalysis Science & Technology

(WGS) reaction. However, under a CO2/H2 regime, cobalt cata-


lysts, behaving like a methanation catalyst rather than
performing as a FT catalyst, do not follow the usual FT chain
growth mechanism at all, and the product distribution
changes significantly.67,94,95 As shown in Fig. 13, it clearly ap-
pears that the hydrogenation of CO2 over Co/Al2O3 catalysts
does not lead to a typical ASF distribution, differently from
what happens in the case of CO hydrogenation.96 Most at-
tempts yield CH4 as the main product.97,98 Akin et al. ob-
Published on 06 September 2017. Downloaded by University of Winnipeg on 1/23/2019 4:28:10 AM.

served that products of CO2 hydrogenation contained 70


mol% CH4 over Co/Al2O3 catalyst,97 which is analogous to the
literature precedent.99
Therefore, tremendous attempts are employed to reduce
the production of CH4.95,98,99 Donner et al. found that the Fig. 14 ASF distribution over the Co–Pt/Al2O3 catalyst at different H2/
CO2 ratios: (a) 3 : 1, (b) 2 : 1, and (c) 1 : 1. The dashed line is the fitted
shift from CH4 toward longer chain hydrocarbons can be
plot, while the solid line is the actual data obtained from the catalyst,
achieved by reducing the feed gas ratio of H2/CO2, as one can with the data points highlighted by an ×.95
see from Fig. 14.95 They regarded that the behavior was possi-
bly associated with the two different active sites (a distinct
break in the graph of Fig. 14 indicates two different growth is, adding alkali metals on cobalt can increase the selectivity
sites resulting in α1 and α2) for CH4 and C2–C4 products to long-chain hydrocarbons and decrease selectivity to
presenting on the catalyst; the decreasing ratio of H2/CO2 led CH4.73,98,99,101 The variation of alkali metal was investigated
to a lowering of the catalyst's methanation potential, con- by Owen and co-authors.98 They observed that the catalyst be-
versely improving the length of hydrocarbons from (a) 0.41 to ing promoted by sodium or potassium gives greater selectiv-
(c) 0.54. ity to C5+ hydrocarbons and lower selectivity to CH4 with no
Promoters. It has been shown that the addition of noble significant decrease in CO2 conversion. They attributed the
metals to Co-based catalysts would be beneficial in view of re- improved C5+ hydrocarbon selectivity upon employing the al-
ducing the selectivity towards CH4.98,100 It was observed that kali to an enhancement in surface-to-molecule charge trans-
the CH4 selectivity decreased with C5+ hydrocarbons, dou- fer, and the binding strength of CO increased while that of
bling upon the introduction of ruthenium. In the presence of H2 reduced, thus resulting in an increased chain growth
platinum, a large decrease in CH4 selectivity can be achieved, probability.83 As commented before, Gnanamani et al. pro-
as described by Willauer et al.95 Additionally, group VIII posed that Na in association with cobalt carbide plays a piv-
metals have also been used as dopants. Copper present in co- otal role during the RWGS reaction, which is likely a prereq-
balt catalysts has been found to be very active in the WGS uisite for the conversion of CO2 to alcohols, and they
without synthesizing significant quantities of CH4.99 highlighted a role played by Na in stabilizing the carbide
In addition, literature data show that the incorporation of phase.73 Similar phenomena can also be observed in the CO-
alkali metals has a beneficial effect upon the catalysis; that based FTS by Zhong et al.; they suggest that Na enhances the
formation of Co2C, which exhibits high selectivity towards
the formation of lower olefins whereas low selectivity towards
CH4 production.59
Supports. Supports have significant roles in the direct
conversion of CO2 into hydrocarbons over Co-based catalysts.
Most of the research in the field is focused on the use of SiO2
and Al2O3.73,95–98 However, some reports have shown that
other supports could outperform SiO2 and Al2O3.66 Owen
et al. investigated the effect of the supports (including SiO2,
CeO2, TiO2, ZrO2, Al2O3, ZSM-5, MgO) on Co–Na–Mo catalysts
for the conversion of CO2 to hydrocarbons, and the catalytic
performances of Co–Na–Mo deposited on different supports
are given in Table 1.66 It appears that catalysts supported on
SiO2 and ZSM-5 present the highest CO2 conversion values,
and catalysts utilizing ZrO2, Al2O3, TiO2, and CeO2 as support
obtained a similar CO2 conversion. However, the hydrocar-
bon selectivity differs, which decreases in the following order:
Fig. 13 Hydrocarbon experimental ASF plots reported in terms of
ZrO2, Al2O3, TiO2, CeO2. Introducing the MgO as support
selectivity during CO and CO2 hydrogenation over Co/Al2O3 catalysts yields a sole CO without hydrocarbons being formed. These
under the same conditions.96 results demonstrate the importance of the support not only

4590 | Catal. Sci. Technol., 2017, 7, 4580–4598 This journal is © The Royal Society of Chemistry 2017
View Article Online

Catalysis Science & Technology Minireview

Table 1 Catalytic activity of Co–Na–Mo supported on a range of oxides66

CO2 CO HC Hydrocarbon distribution


Catalyst conversion selectivity selectivity
Entry support (%) (%) (%) C1 C2= C2 C3= C3 C4 C5 α
1 SiO2 30.0 21.5 78.5 57.1 0.0 9.3 0.1 11.8 9.3 12.3 0.48
2 SiO2a 15.6 59.0 41.0 42.1 0.0 10.7 1.7 14.1 12.7 18.7 0.52
3 CeO2 15.1 70.2 29.8 22.1 0.92 12.8 9.8 9.1 14.1 31.2 0.59
4 TiO2 13.5 66.7 33.3 23.1 0.77 9.9 9.2 7.0 13.7 24.2 0.60
5 Al2O3 15.4 57.3 42.7 29.3 0.0 12.3 4.1 14.1 13.9 27.1 0.57
Published on 06 September 2017. Downloaded by University of Winnipeg on 1/23/2019 4:28:10 AM.

6 MgO 8.7 100 — 0.0 0.0 0.0 0.0 0.0 0.0 0.0 NA
7 ZrO2 14.4 16.7 83.3 26.8 0.0 14.8 12.9 8.2 14.9 22.4 0.52
8 ZSM-5 29.1 20.6 79.4 95.7 0.0 3.0 0.0 0.9 0.2 0.2 0.17

Reaction conditions: 200 °C, 0.1 MPa, H2/CO2 = 3 : 1 and 8 SCCM total flow, WHSV: 0.35 h−1. a WHSV: 1.73 h−1.

on the cobalt crystallite size but also the critical effect of the 3.1.3. Mixed Fe–Co catalysts. Mixed Fe–Co catalysts are
metal–support interaction on the product distribution. also active for CO2 hydrogenation.80,103,104 Satthawong et al.
Active sites. In spite of the progress in the selectivity to claimed that the addition of a small amount of Co with Fe
heavier hydrocarbons (especially C5+ hydrocarbons) by the led to a dramatic improvement in selectivity to C2+ hydrocar-
promotion of alkali promoters like Na and K, reports on the bons.103 Gnanamani et al. has also shown that Fe–Co bime-
state and nature of the active phase of cobalt that is tallic catalysts outperform monometallic catalysts (Fe or Co)
responsible for the observed increase in C5+ hydrocarbons for CO2 hydrogenation for the production of higher C2+ hy-
are scarce. In order to shed further light on the nature of the drocarbons,80 and the catalyst containing a 50 : 50 Co : Fe
active phase of cobalt during catalysis, Gnanamani et al. composition was found to be the optimum for the production
investigated the influence of pretreatment conditions of of higher hydrocarbons. Although K-promoted or Na-
cobalt on the performance of CO2 hydrogenation over 1% Na/ promoted Fe–Co catalysts exhibit an increased selectivity to
20% Co–SiO2 catalyst.73 Three different activation conditions, higher hydrocarbons, the selectivity to CH4 is still too high.
such as H2, CO/H2 and CO, were employed to control the 3.1.4. Reaction mechanisms (CO-FT vs. CO2-FT). A great
chemical nature of cobalt phases. Fig. 15 shows the product deal of studies have been conducted on catalytic hydrogena-
selectivity described versus the method of activation. As one tion of CO2, for cobalt and iron catalysts; a significant in-
can see, the metallic form of cobalt is obtained and the crease in formation of CH4 (the most thermodynamically fa-
products are primarily CH4 and lower hydrocarbons (C2–C4) vored hydrocarbon product) was noted upon switching the
in the case of H2 pretreated catalysts, while pretreating with reactants from syngas to CO2/H2. A comparative study of FTS
pure H2 or CO/H2 at 250 °C yields a fraction of CoO, which with CO/H2 and CO2/H2 over Co-based catalysts has been car-
suppresses the hydrogenation activity of cobalt to some ried out by many researchers.75,94,96,105 With respect to CO or
degree; as a result, selectivity to CH4 declines. Moreover, after CO2 hydrogenation, Zhang et al. found that the catalytic ac-
CO activation, CoO and cobalt carbide phases form and tivities obtained were similar while the selectivities of the two
produce significantly less CH4, and the selectivity for processes were extremely different.94 For CO hydrogenation,
alcohols increases to 73.2%. Their studies reveal that the normal FTS product distributions were observed with an α of
partially carburized form of cobalt is responsible for the high
alcohol selectivity (73.2%) observed.
However, the above-mentioned Co-based catalysts are pri-
marily prepared by traditional coprecipitation, impregnation,
or physical mixing methods and contain multiple compo-
nents including various promoters as well as supports. Thus,
they exhibit large uncertainties on the spatial arrangements
of active sites. More recently, Xie et al. prepared a well-
defined CeO2–Pt@mSiO2–Co core–shell catalyst (a schematic
of the synthetic process and TEM images of each step are
presented in Fig. 16) for the targeted synthesis of C2–C4 hydro-
carbons from CO2 hydrogenation via a two-step tandem reac-
tion; that is, the RWGS reaction is conducted over a CeO2/Pt
interface and the formed CO subsequently reacts with H2
through the FTS process on the Co/mSiO2 interface.102 A selec-
tivity of 60% toward C2–C4 hydrocarbons was obtained, which Fig. 15 Effect of catalyst activation conditions of 1% Na–20% Co/SiO2
gives impetus to the rational design multifunctional catalysts on CO2 hydrogenation selectivity (reaction conditions: 220 °C, 18.9
with high performance for tandem conversions. bar, 3000 ml h−1 gcat−1, H2/CO2 = 3 : 1).73

This journal is © The Royal Society of Chemistry 2017 Catal. Sci. Technol., 2017, 7, 4580–4598 | 4591
View Article Online

Minireview Catalysis Science & Technology


Published on 06 September 2017. Downloaded by University of Winnipeg on 1/23/2019 4:28:10 AM.

Fig. 16 Synthesis and characterization of the CeO2–Pt@mSiO2–Co tandem catalyst. (a) Schematic of the synthetic process. TEM images of each
step: (b and c) CeO2 nanoparticles, (d and e) overgrowth of Pt nanoparticles on CeO2, (f and g) silica shell coating on CeO2–Pt composite
nanoparticles, (h and i) deposition of Co nanoparticles on CeO2–Pt@mSiO2, and (j) scaled up preparation of CeO2–Pt@mSiO2–Co nanoparticles.
One-pot synthesis can yield 300 mg of catalyst.102

about 0.80; when it comes to the hydrogenation of CO2, with 3.2. Bifunctional catalysis via methanol intermediates
CH4 largely dominating the product distribution, similar re-
sults were reported by Riedel et al.75 As a consequence, it is Direct conversion of CO2 into hydrocarbons can also be
obvious that hydrogenation of CO2 does not lead to a typical achieved via methanol-mediated reactions. In the methanol-
ASF distribution, which differs from what takes place for CO mediated approach, hydrocarbon synthesis was conducted
hydrogenation. Visconti et al. ascribed the change in the yield over a hybrid catalyst composed of a methanol synthesis cata-
products to two main reasons.96,105 On the one hand, the dif- lyst and a zeolite catalyst (MTH) from CO2 hydrogenation,
ferent adsorption strengths of CO2 and CO on the catalyst which is advantageous to produce desirable hydrocarbons. In
surface, that is, the different H/C ratios on the surface of the general, CO2 and H2 react over the Cu-based catalysts to syn-
catalyst within the two processes, cause the decrease in chain thesize methanol, which is subsequently transformed into hy-
propagation ability. They consider that during the hydrogena- drocarbons over a zeolite catalyst. The distribution of hydro-
tion of CO2, a higher H/C ratio is obtained due to the weak carbons over hybrid catalysts greatly depended on the zeolite,
CO2 adsorption on the surface. This favors the hydrogenation which affected both the CO2 hydrogenation rate and the de-
of surface-adsorbed intermediates, leading to the formation sired hydrocarbon yields. Up to now, considerable efforts have
of CH4 and a decrease of chain growth ability. On the other been made to explore effective ways of the direct synthesis of
hand, the lower CO partial pressure may motivate the second- hydrocarbons from CO2 and H2 via methanol intermediates.
ary reactions of olefin in CO2 hydrogenation, thus yielding 3.2.1. Synthesis of olefins. Lower olefins are basic carbon-
more saturated hydrocarbons. based building blocks that are widely used in the chemical in-
To throw light upon the process of CO2 hydrogenation, dustry and are traditionally produced through thermal or cata-
previous work has been reported in the mechanism of cata- lytic cracking of a range of hydrocarbon feed stocks, such as
lytic hydrogenation of CO2 to hydrocarbons.94 Saeidi et al. naphtha, gas oil, condensates and light alkanes.57,58,107 Selec-
presented a reaction mechanism for hydrogenation of CO2 to tive synthesis of lower olefins from a CO2 + H2 mixture by one-
hydrocarbon over iron catalysts as shown in Fig. 17.106 It can pass conversion via methanol synthesis was investigated by
be seen from Fig. 17 that CO2 is initially reduced by ironIJII) Inui et al.108 The authors considered that for the selective syn-
followed by H radical absorption on the catalyst surface. OH, thesis of olefin, a weakly acidic and narrow pore microporous
formic acid and CO are formed when the residual H attacked crystalline catalyst such as SAPO-34 was required. A bifunc-
carbonyl C. The Fe–CH2 radical, considered as a C–C propa- tional catalyst containing In2O3 and SAPO-34 could realize the
gation species, is then formed in the same way.106 direct production of lower olefins from CO2 hydrogenation

4592 | Catal. Sci. Technol., 2017, 7, 4580–4598 This journal is © The Royal Society of Chemistry 2017
View Article Online

Catalysis Science & Technology Minireview


Published on 06 September 2017. Downloaded by University of Winnipeg on 1/23/2019 4:28:10 AM.

Fig. 17 Reaction mechanisms of CO2 to hydrocarbons.106

with excellent selectivity and high activity. The selectivity of C2– also studied by Li et al.111 Their results revealed that the hy-
C4 olefin reached up to around 76.9% with a much lower CH4 brid catalyst was active and stable for LPG synthesis even at
selectivity of 4.4% and CO2 conversion was above 34%.109 low temperature (260 °C), and the conversion of CO2 and the
3.2.2. Synthesis of LPG. LPG is a good substitute for fuel yield of hydrocarbons can reach up to 25.2% and 13.3%, re-
in spark ignition engines and is used as a replacement for spectively, while the selectivity of LPG was as high as 75%
aerosol propellants and refrigerants. Therefore, direct synthe- with limited CH4 formed (only 1%), indicating that the
sis of LPG from CO2 hydrogenation is a desirable route for proper use of a methanol catalyst with zeolite made it a pos-
converting the most abundant carbon resource to highly valu- sibility to obtain the specified hydrocarbons from CO2. Park
able products.110,111 Jeon et al. investigated the catalytic per- et al. proposed that the yield of hydrocarbons was mainly de-
formance of a range of hybrid catalysts consisting of metha- termined by the amount of acid sites of zeolite according to
nol synthesis catalysts (for example, Cu/ZnO/ZrO2 and Cu/ their studies.112 Furthermore, they found that the main prod-
ZnO/Al2O3) and zeolites (for example, SAPO-5, SAPO-44 and uct was C3 or C4 hydrocarbons in the case of the hybrid cata-
HZSM-5) for the direct synthesis of hydrocarbons from CO2 lyst consisting of Cu/ZnO/ZrO2 and SAPO zeolite, and they
hydrogenation via methanol.110 They concluded that propane suggested that the mechanism operating for SAPO was the
was the main product on the hybrid catalysts with SAPO-44, carbon pool mechanism.
while on the hybrid catalysts with SAPO-5 isobutene was the 3.2.3. Synthesis of gasoline. Gasoline is a very important
predominant product due to the different acidity and pore transportation fuel widely used around the world. Systems
size between those zeolites. The selective synthesis of LPG which combine methanol synthesis catalysts and zeolites
over the hybrid catalysts composed of a Zr-modified Cu–Zn– which can produce enhanced yields of C2+ hydrocarbons by
Al methanol catalyst with a Pd-modified zeolite (Pd-β) was combining methanol synthesis and the MTG reaction. Aimed

This journal is © The Royal Society of Chemistry 2017 Catal. Sci. Technol., 2017, 7, 4580–4598 | 4593
View Article Online

Minireview Catalysis Science & Technology


Published on 06 September 2017. Downloaded by University of Winnipeg on 1/23/2019 4:28:10 AM.

Fig. 18 Molecular level mechanism for CO2 hydrogenation into hydrocarbons over an In2O3/HZSM-5 bifunctional catalyst.109

at obtaining a gaseous fuel from CO2 hydrogenation, a cata- two components (mortar mixing) was further shortened,
lyst system involving a methanol synthesis catalyst (Pd–Na- which leads to a severe deactivation with very low C5+ hydro-
modified Cu–Cr–Zn oxides) and a typical MTG catalyst (H- carbon selectivity. Moreover, the catalytic performance for
ZSM-5) was adopted, and the selectivity of C2+ hydrocarbons CO2 hydrogenation was also evaluated using pellet catalysts
among the yielded hydrocarbons was up to 71.8% over the with a cylindrical shape (Φ3.0 mm × 3.5 mm) under industri-
above composite catalysts.113 Fujiwara et al. studied the hy- ally relevant conditions. Tail gas recycling, which was com-
drogenation of CO2 over a Cu–Zn–Cr-oxide/zeolite composite monly used in industry for more efficient utilization of the
catalyst system.114 They found that Cu–Zn–Cr/HY exhibited feed, could improve the catalytic performance for CO2 hydro-
the best performance for CO2 hydrogenation with a CO2 con- genation to C5+ hydrocarbons significantly, suggesting a
version and hydrocarbon selectivity of 39.9% and 12%, re- promising prospect for industrial application.
spectively, and the selectivity of C2+ hydrocarbons was as high To selectively synthesize isoalkane, which is considered as
as 95.8% under reaction conditions of 400 °C, 0.5 MPa, H2/ a fuel additive both for improving the octane value of gasoline
CO2 = 3, and 3000 ml h−1 gcat−1. Moreover, it was also con- and for enhancing automobile exhaust, some authors took
firmed that the MTG reaction and the decomposition of meth- the hybrid catalyst composed of Fe–Zn–Zr and HY zeolite as
anol into CO controlled the production of C2+ hydrocarbons.
However, only low yields of hydrocarbons are obtained,
with CO and CH4 being the predominant products over the
above-mentioned catalysts. Ereña et al. found that the syn-
thetic gasoline fraction was higher over the bifunctional cata-
lysts composed of Cr2O3–ZnO and ZSM-5 than other catalysts,
with a selectivity of 36.1 wt% for total hydrocarbons and C5+
selectivity of 53.9%.115 Very recently, an In2O3/HZSM-5 bifunc-
tional catalyst reported by Gao et al. has exhibited excellent
performance for the direct production of gasoline-range hy-
drocarbons from CO2 hydrogenation with high selectivity and
a remarkable stability.109 The C5–C11 selectivity in hydrocar-
bon distribution reached up to 80% with only 1% CH4 under
reaction conditions of 340 °C, 3 MPa, H2/CO2 = 3, and 9000
ml h−1 gcat−1. The oxygen vacancies on In2O3 surfaces activate
CO2 and H2 to form methanol, and C–C coupling subse-
quently occurs inside zeolite pores to produce gasoline-range
hydrocarbons (Fig. 18). The key intermediates involved in
methanol synthesis are more stable on the defective In2O3
surface than those on the Cu surface, which strongly sup-
presses the formation of CO. Consequently, CO selectivity
over In2O3 integrated with the zeolite (<45%) is much lower Fig. 19 Influence of the integration manner of the active components
than that over traditional Cu-based catalysts (>90%) at 340 (In2O3/HZSM-5 mass ratio = 2 : 1) on catalytic behaviours under the
°C. In addition, by moving the two components into a closer same conditions. (a) Dual-bed configuration with In2O3 packed below
HZSM-5 and separated by a layer of quartz sand. (b) HZSM-5 packed
proximity (from Fig. 19a–d), CO selectivity decreased and the
below In2O3 and separated by quartz sand. (c) Stacking of granules
C5+ selectivity was enhanced. However, the number of with In2O3, HZSM-5 and quartz sand particle sizes of 250–380 μm. (d)
strongly acidic sites of the spent catalyst presented in In2O3 and HZSM-5 particles well mixed without quartz sand. (e) In2O3
Fig. 19e decreased markedly when the distance between the and HZSM-5 mixed with an agate mortar.109

4594 | Catal. Sci. Technol., 2017, 7, 4580–4598 This journal is © The Royal Society of Chemistry 2017
View Article Online

Catalysis Science & Technology Minireview

able energy in the chemical industry value chain while im-


proving resource efficiency and limiting greenhouse gas
emissions. CO2 can be hydrogenated to hydrocarbons by ei-
ther indirect or direct routes, but in all cases there are nu-
merous technical and economic barriers to overcome, which
will further drive the research and development in the pro-
duction processes including the development of catalysts, re-
actors and related process technology from non-petroleum
Published on 06 September 2017. Downloaded by University of Winnipeg on 1/23/2019 4:28:10 AM.

feedstocks.
We have analyzed here the status of research in the indi-
rect routes. The catalytic chemistry of the CO2 hydrogenation
to methanol reaction and the subsequent production of ole-
fins, LPG, gasoline and aromatics via MTH as well as RWGS
Scheme 1 Proposed reaction path of isoalkane formation from CO2 + reaction and value-added hydrogenation are then synthesized
H2 over Fe–Zn–Zr/HY hybrid catalyst.116
via FTS, or cracking of heavy hydrocarbons produced in the
FT process is well established, even if there is still the need
the catalyst for isoalkane synthesis from CO2 hydrogena- for development. Also in terms of process development most
tion.116,117 Ni et al. confirmed that the Fe–Zn–Zr not only acts of the knowledge necessary is available. Minor technological
as a methanol synthesis catalyst but also benefits isoalkane barriers to develop these routes are thus present.
formation.116 A reasonable intact route to isoalkanes is pro- As compared with the indirect route, the direct conversion
posed according to their results, as shown in Scheme 1. First, of CO2 into hydrocarbons would be more economical and en-
the methanol is formed over the Fe–Zn–Zr catalyst and then ergy-efficient. Various attempts have been made to transform
converted to C1 and C4 and further into hydrocarbons includ- carbon dioxide into hydrocarbons directly on FT-based cata-
ing i-C4 and i-C5 over HY zeolite. Moreover, the formed C3 can lysts. The use of CO2 via FT reactions should be focused ei-
further react with methanol to obtain i-C4, while i-C5 can be ther on the selective synthesis of short-chain olefins or on
synthesized from C2 and C3. However, the synergistic effect of the synthesis of hydrocarbons to be used as fuels. It is appar-
metal oxide and zeolite in the mechanical mixture is very ent that catalysts play a key role in driving success. Neverthe-
poor, and the selectivity to isoalkanes is still unsatisfied. To less, the nature of the active sites and interactions among ac-
overcome the above problems, a core–shell catalyst with Fe– tive components, promoter and support as well as reaction
Zn–Zr as the core and a zeolite (HZSM-5, Hbeta and HY) as mechanisms are still in dispute. There is a need for an im-
the shell was prepared using a simple cladding method devel- proved catalyst incorporating both RWGS activity and chain
oped by Wang et al. (Fig. 20).118 Compared with the conven- growth. Nano-structured, porous, and functional materials
tional Fe–Zn–Zr/zeolite prepared by mechanical mixing, the are of paramount importance to these catalytic conversion
Fe–Zn–Zr@zeolite core–shell catalyst exhibited a significantly processes. With a better understanding of the fundamental
improved confinement effect on the promotion of the direct structure–composition–activity relationships of these catalytic
synthesis of isoalkanes from CO2, and the selectivity to iso- systems, the catalysts can be tuned for better catalytic
alkanes in the yielded hydrocarbons reached up to 81.3%. performance.
The CO2-FT reaction is essentially a modification of the
FTS, where CO2 is used instead of CO, and thus the distribu-
4. Conclusions and future tion of hydrocarbons is limited by an ASF model. The meth-
perspectives odology of reaction coupling has resulted in a significant
breakthrough in the direct synthesis of olefin, LPG and gaso-
The utilization of CO2 as the carbon source for the produc- line from CO2 hydrogenation using bifunctional catalysts,
tion of hydrocarbons is an effective way to introduce renew- which provide an important platform for CO2 conversion into

Fig. 20 Illustration of the Fe–Zn–Zr@zeolite core–shell catalyst preparation by a simple cladding method.118

This journal is © The Royal Society of Chemistry 2017 Catal. Sci. Technol., 2017, 7, 4580–4598 | 4595
View Article Online

Minireview Catalysis Science & Technology

chemicals and fuels. As stated above, the bifunctional catalyst 12 S. Saeidi, N. A. S. Amin and M. R. Rahimpour, J. CO2 Util.,
separates CO2 activation and C–C coupling onto two types of 2014, 5, 66–81.
active sites with complementary and compatible properties. 13 M. Aresta, A. Dibenedetto and A. Angelini, Chem. Rev.,
During CO2 hydrogenation, CO2 is activated on Cu-based cat- 2014, 114, 1709–1742.
alysts or an oxide surface with oxygen vacancies and hydroge- 14 E. V. Kondratenko, G. Mul, J. Baltrusaitis, G. O. Larrazabal
nated to methanol intermediate species, while C–C coupling and J. Perez-Ramirez, Energy Environ. Sci., 2013, 6,
is controlled within the pores of zeolite. The activation mech- 3112–3135.
anism of CO2, the rule of the formation of reaction interme- 15 G. Centi, E. A. Quadrelli and S. Perathoner, Energy Environ.
Published on 06 September 2017. Downloaded by University of Winnipeg on 1/23/2019 4:28:10 AM.

diates and C–C precise coupling, the synergy mechanism of Sci., 2013, 6, 1711–1731.
two active sites, as well as the deactivation behavior need to 16 C. S. Song, Catal. Today, 2006, 115, 2–32.
be further clarified, even if relevant research has been made 17 H. Wang, P. Gao, T. J. Zhao, W. Wei and Y. H. Sun, Sci.
recently. With the assistance of numerous in situ characteri- China: Chem., 2015, 58, 79–92.
zation techniques and theories, the design of cost-effective 18 M. Behrens, S. Zander, P. Kurr, N. Jacobsen, J. Senker, G.
catalysts will be promoted a great deal. Last but not least, fu- Koch, T. Ressler, R. W. Fischer and R. Schlogl, J. Am. Chem.
ture research should certainly emphasize the rational design Soc., 2013, 135, 6061–6068.
of highly active catalysts for direct conversion of CO2 into 19 J. Ma, N. N. Sun, X. L. Zhang, N. Zhao, F. K. Mao, W. Wei
value-added hydrocarbons. and Y. H. Sun, Catal. Today, 2009, 148, 221–231.
20 C. Li, X. Yuan and K. Fujimoto, Appl. Catal., A, 2014, 469,
Conflicts of interest 306–311.
21 P. Gao, L. S. Zhong, L. N. Zhang, H. Wang, N. Zhao, W. Wei
There are no conflicts to declare. and Y. H. Sun, Catal. Sci. Technol., 2015, 5, 4365–4377.
22 J. Toyir, P. R. r. de la Piscina, J. L. G. Fierro and N. S.
Acknowledgements Homs, Appl. Catal., B, 2001, 29, 207–215.
This work was financially supported by the National Natural 23 S. Posada-Perez, P. J. Ramirez, R. A. Gutierrez, D. J.
Science Foundation of China (21503260, 21773286), the Minis- Stacchiola, F. Vines, P. Liu, F. Illas and J. A. Rodriguez,
try of Science and Technology of China (2017YFB0602202, Catal. Sci. Technol., 2016, 6, 6766–6777.
2017YFB0602205), Shanghai Municipal Science and Technol- 24 W. Ding, Y. Liu, F. Wang, S. Zhou, A. Chen, Y. Yang and W.
ogy Commission, China (15ZR1444500, 16DZ1206900), the Fang, RSC Adv., 2014, 4, 30677.
“Frontier Science” program of Shell Global Solutions Interna- 25 A. Karelovic and P. Ruiz, Catal. Sci. Technol., 2015, 5,
tional B. V. (PT65197), and Science and Technology Innovation 869–881.
Fund of Shanghai Advanced Research Institute, CAS (172001). 26 N. J. Brown, J. Weiner, K. Hellgardt, M. S. Shaffer and C. K.
Williams, Chem. Commun., 2013, 49, 11074–11076.
References 27 T. Lunkenbein, J. Schumann, M. Behrens, R. Schlogl and
M. G. Willinger, Angew. Chem., Int. Ed., 2015, 54,
1 G. A. Olah, G. K. S. Prakash and A. Goeppert, J. Am. Chem. 4544–4548.
Soc., 2011, 133, 12881–12898. 28 C. Tisseraud, C. Comminges, S. Pronier, Y. Pouilloux and A.
2 A. Banerjee, G. R. Dick, T. Yoshino and M. W. Kanan, Le Valant, J. Catal., 2016, 343, 106–114.
Nature, 2016, 531, 215–219. 29 J. Sun, G. Yang, Q. Ma, I. Ooki, A. Taguchi, T. Abe, Q. Xie,
3 W. Wang, S. P. Wang, X. B. Ma and J. L. Gong, Chem. Soc. Y. Yoneyama and N. Tsubaki, J. Mater. Chem. A, 2014, 2,
Rev., 2011, 40, 3703–3727. 8637–8643.
4 T. Sakakura, J. C. Choi and H. Yasuda, Chem. Rev., 30 J. Toyir, R. Miloua, N. E. Elkadri, M. Nawdali, H. Toufik, F.
2007, 107, 2365–2387. Miloua and M. Saito, Phys. Procedia, 2009, 2, 1075–1079.
5 E. E. Benson, C. P. Kubiak, A. J. Sathrum and J. M. Smieja, 31 M. M.-J. Li, Z. Zeng, F. Liao, X. Hong and S. C. E. Tsang,
Chem. Soc. Rev., 2009, 38, 89–99. J. Catal., 2016, 343, 157–167.
6 K. Asami, Q. W. Zhang, X. H. Li, S. Asaoka and K. Fujimoto, 32 O. Martin, C. Mondelli, D. Curulla-Ferré, C. Drouilly, R.
Catal. Today, 2005, 106, 247–251. Hauert and J. Pérez-Ramírez, ACS Catal., 2015, 5,
7 C. G. Yang, M. H. Qiu, S. W. Hu, X. Q. Chen, G. F. Zeng, 5607–5616.
Z. Y. Liu and Y. H. Sun, Microporous Mesoporous Mater., 33 P. Gao, F. Li, H. J. Zhan, N. Zhao, F. K. Xiao, W. Wei, L. S.
2016, 231, 110–116. Zhong, H. Wang and Y. H. Sun, J. Catal., 2013, 298, 51–60.
8 S. Ilias and A. Bhan, ACS Catal., 2013, 3, 18–31. 34 P. Gao, F. Li, F. K. Xiao, N. Zhao, N. N. Sun, W. Wei, L. S.
9 Q. Q. Chen, M. Lv, Z. Y. Tang, H. Wang, W. Wei and Y. H. Zhong and Y. H. Sun, Catal. Sci. Technol., 2012, 2,
Sun, J. CO2 Util., 2016, 14, 1–9. 1447–1454.
10 W. H. Wang, Y. Himeda, J. T. Muckerman, G. F. Manbeck 35 K. Samson, M. Śliwa, R. P. Socha, K. Góra-Marek, D.
and E. Fujita, Chem. Rev., 2015, 115, 12936–12973. Mucha, D. Rutkowska-Zbik, J. F. Paul, M. Ruggiero-
11 R. M. Cuellar-Franca and A. Azapagic, J. CO2 Util., 2015, 9, Mikołajczyk, R. Grabowski and J. Słoczyński, ACS Catal.,
82–102. 2014, 4, 3730–3741.

4596 | Catal. Sci. Technol., 2017, 7, 4580–4598 This journal is © The Royal Society of Chemistry 2017
View Article Online

Catalysis Science & Technology Minireview

36 H. Bahruji, M. Bowker, G. Hutchings, N. Dimitratos, P. 61 J. Plana-Pallejà, S. Abello, C. Berrueco and D. Montane,


Wells, E. Gibson, W. Jones, C. Brookes, D. Morgan and G. Appl. Catal., A, 2016, 515, 126–135.
Lalev, J. Catal., 2016, 343, 133–146. 62 G. D. Weatherbee and C. H. Bartholomew, J. Catal.,
37 Y. Hartadi, D. Widmann and R. J. Behm, J. Catal., 1984, 87, 352–362.
2016, 333, 238–250. 63 T. Riedel, G. Schaub, K. W. Jun and K. W. Lee, Ind. Eng.
38 J. Xiao and T. Frauenheim, J. Phys. Chem. C, 2013, 117, Chem. Res., 2001, 40, 1355–1363.
1804–1808. 64 D. U. Martin, P. Rohde and Georg Schaub, Ind. Eng. Chem.
39 F. Studt, I. Sharafutdinov, F. Abild-Pedersen, C. F. Elkjaer, Res., 2005, 44, 9653–9658.
Published on 06 September 2017. Downloaded by University of Winnipeg on 1/23/2019 4:28:10 AM.

J. S. Hummelshoj, S. Dahl, I. Chorkendorff and J. K. 65 W. Wang, S. Wang, X. Ma and J. Gong, Chem. Soc. Rev.,
Norskov, Nat. Chem., 2014, 6, 320–324. 2011, 40, 3703–3727.
40 I. Sharafutdinov, C. F. Elkjær, H. W. Pereira de Carvalho, D. 66 R. E. Owen, P. Plucinski, D. Mattia, L. Torrente-Murciano,
Gardini, G. L. Chiarello, C. D. Damsgaard, J. B. Wagner, V. P. Ting and M. D. Jones, J. CO2 Util., 2016, 16, 97–103.
J. D. Grunwaldt, S. Dahl and I. Chorkendorff, J. Catal., 67 R. W. Dorner, D. R. Hardy, F. W. Williams and H. D.
2014, 320, 77–88. Willauer, Energy Environ. Sci., 2010, 3, 884.
41 J. Ye, C. Liu, D. Mei and Q. Ge, ACS Catal., 2013, 3, 68 P. H. Choi, K. W. Jun, S. J. Lee, M. J. Choi and K. W. Lee,
1296–1306. Catal. Lett., 1996, 40, 115–118.
42 O. Martin, A. J. Martin, C. Mondelli, S. Mitchell, T. F. 69 L. Xu, Q. Wang, D. Liang, X. Wang, L. Lin, W. Cui and Y.
Segawa, R. Hauert, C. Drouilly, D. Curulla-Ferre and J. Xu, Appl. Catal., A, 1998, 173, 19–25.
Perez-Ramirez, Angew. Chem., Int. Ed., 2016, 55, 6261–6265. 70 W. Ning, N. Koizumi and M. Yamada, Energy Fuels,
43 C. D. Chang and A. J. Silvestri, J. Catal., 1977, 47, 249–259. 2009, 23, 4696–4700.
44 J. F. Haw, W. Song, D. M. Marcus and J. B. Nicholas, Acc. 71 R. W. Dorner, D. R. Hardy, F. W. Williams and H. D.
Chem. Res., 2003, 36, 317–326. Willauer, Appl. Catal., A, 2010, 373, 112–121.
45 G. Seo, J.-H. Kim and H.-G. Jang, Catal. Surv. Asia, 2013, 17, 72 H. D. Willauer, R. Ananth, M. T. Olsen, D. M. Drab, D. R.
103–118. Hardy and F. W. Williams, J. CO2 Util., 2013, 3–4, 56–64.
46 U. Olsbye, S. Svelle, M. Bjorgen, P. Beato, T. V. Janssens, F. 73 M. K. Gnanamani, G. Jacobs, R. A. Keogh, W. D. Shafer,
Joensen, S. Bordiga and K. P. Lillerud, Angew. Chem., Int. D. E. Sparks, S. D. Hopps, G. A. Thomas and B. H. Davis,
Ed., 2012, 51, 5810–5831. Appl. Catal., A, 2015, 499, 39–46.
47 M. Stöcker, Microporous Mesoporous Mater., 1999, 29, 3–48. 74 N. Fischer, R. Henkel, B. Hettel, M. Iglesias, G. Schaub and
48 S. N. Khadzhiev, N. V. Kolesnichenko and N. N. Ezhova, M. Claeys, Catal. Lett., 2016, 146, 509–517.
Pet. Chem., 2008, 48, 325–334. 75 M. C. Thomas Riedel, H. Schulz, G. Schaub, S.-S. Nam,
49 J. Liang, H. Li, S. Zhao, W. Guo, R. Wang and M. Ying, M. J. C. K.-W. Jun, G. Kishan and K.-W. Lee, Appl. Catal., A,
Appl. Catal., 1990, 64, 31–40. 1999, 186, 201–213.
50 A. Galadima and O. Muraza, J. Nat. Gas Sci. Eng., 2015, 25, 76 P. S. S. Prasad, J. W. Bae, K.-W. Jun and K.-W. Lee, Catal.
303–316. Surv. Asia, 2008, 12, 170–183.
51 U. Olsbye, S. Svelle, K. P. Lillerud, Z. H. Wei, Y. Y. Chen, 77 P. A. Chernavskii, V. O. Kazak, G. V. Pankina, Y. D.
J. F. Li, J. G. Wang and W. B. Fan, Chem. Soc. Rev., Perfiliev, T. Li, M. Virginie and A. Y. Khodakov, Catal. Sci.
2015, 44, 7155–7176. Technol., 2017, 7, 2325–2334.
52 I. M. Dahl and S. Kolboe, J. Catal., 1994, 149, 458–464. 78 C. G. Visconti, M. Martinelli, L. Falbo, A. Infantes-Molina,
53 W. Song, J. F. Haw, J. B. Nicholas and C. S. Heneghan, L. Lietti, P. Forzatti, G. Iaquaniello, E. Palo, B. Picutti and
J. Am. Chem. Soc., 2000, 122, 10726–10727. F. Brignoli, Appl. Catal., B, 2017, 200, 530–542.
54 S. Mousavi, A. Zamaniyan, M. Irani and M. Rashidzadeh, 79 J. Wang, Z. You, Q. Zhang, W. Deng and Y. Wang, Catal.
Appl. Catal., A, 2015, 506, 57–66. Today, 2013, 215, 186–193.
55 R. A. Friedel and R. B. Anderson, J. Am. Chem. Soc., 80 M. K. Gnanamani, G. Jacobs, H. H. Hamdeh, W. D. Shafer,
1950, 72, 1212–1215. F. Liu, S. D. Hopps, G. A. Thomas and B. H. Davis, ACS
56 H. M. T. Galvis and K. P. de Jong, ACS Catal., 2013, 3, Catal., 2016, 6, 913–927.
2130–2149. 81 J. Wei, J. Sun, Z. Wen, C. Fang, Q. Ge and H. Xu, Catal. Sci.
57 L. S. Zhong, F. Yu, Y. L. An, Y. H. Zhao, Y. H. Sun, Z. J. Li, Technol., 2016, 6, 4786–4793.
T. J. Lin, Y. J. Lin, X. Z. Qi, Y. Y. Dai, L. Gu, J. S. Hu, S. F. 82 J. Wei, Q. Ge, R. Yao, Z. Wen, C. Fang, L. Guo, H. Xu and J.
Jin, Q. Shen and H. Wang, Nature, 2016, 538, 84–87. Sun, Nat. Commun., 2017, 8, 15174.
58 H. M. T. Galvis, J. H. Bitter, C. B. Khare, M. Ruitenbeek, A. I. 83 D. B. Bukur, D. Mukesh and S. A. Patel, Ind. Eng. Chem.
Dugulan and K. P. de Jong, Science, 2012, 335, 835–838. Res., 1990, 29, 194–204.
59 L. Zhong, F. Yu, Y. An, Y. Zhao, Y. Sun, Z. Li, T. Lin, Y. Lin, 84 T. Herranz, S. Rojas, F. J. Pérez-Alonso, M. Ojeda, P.
X. Qi, Y. Dai, L. Gu, J. Hu, S. Jin, Q. Shen and H. Wang, Terreros and J. L. G. Fierro, Appl. Catal., A, 2006, 311,
Nature, 2016, 538, 84–87. 66–75.
60 K. Cheng, L. Zhang, J. C. Kang, X. B. Peng, Q. H. Zhang and 85 Y. H. Choi, Y. J. Jang, H. Park, W. Y. Kim, Y. H. Lee, S. H.
Y. Wang, Chem. – Eur. J., 2015, 21, 1928–1937. Choi and J. S. Lee, Appl. Catal., B, 2017, 202, 605–610.

This journal is © The Royal Society of Chemistry 2017 Catal. Sci. Technol., 2017, 7, 4580–4598 | 4597
View Article Online

Minireview Catalysis Science & Technology

86 R. W. Dorner, D. R. Hardy, F. W. Williams and H. D. 102 C. Xie, C. Chen, Y. Yu, J. Su, Y. Li, G. A. Somorjai and P.
Willauer, Catal. Commun., 2011, 15, 88–92. Yang, Nano Lett., 2017, 17, 3798–3802.
87 L. Torrente-Murciano, R. S. L. Chapman, A. Narvaez- 103 R. Satthawong, N. Koizumi, C. Song and P. Prasassarakich,
Dinamarca, D. Mattia and M. D. Jones, Phys. Chem. Chem. Top. Catal., 2014, 57, 588–594.
Phys., 2016, 18, 15496–15500. 104 M. K. Gnanamani, H. H. Hamdeh, G. Jacobs, W. D. Shafer,
88 J. Zhang, S. Lu, X. Su, S. Fan, Q. Ma and T. Zhao, J. CO2 S. D. Hopps, G. A. Thomas and B. H. Davis, ChemCatChem,
Util., 2015, 12, 95–100. 2017, 9, 1303–1312.
89 S. Hu, M. Liu, F. Ding, C. Song, G. Zhang and X. Guo, 105 C. G. Visconti, M. Martinelli, L. Falbo, L. Fratalocchi and L.
Published on 06 September 2017. Downloaded by University of Winnipeg on 1/23/2019 4:28:10 AM.

J. CO2 Util., 2016, 15, 89–95. Lietti, Catal. Today, 2016, 277, 161–170.
90 B. Hu, S. Frueh, H. F. Garces, L. Zhang, M. Aindow, C. 106 S. Saeidi, N. A. S. Amin and M. R. Rahimpour, J. CO2 Util.,
Brooks, E. Kreidler and S. L. Suib, Appl. Catal., B, 2014, 5, 66–81.
2013, 132–133, 54–61. 107 F. Jiao, J. Li, X. Pan, J. Xiao, H. Li, H. Ma, M. Wei, Y. Pan, Z.
91 A. Hakim, T. S. Marliza, N. M. Abu Tahari, R. W. N. Wan Zhou, M. Li, S. Miao, J. Li, Y. Zhu, D. Xiao, T. He, J. Yang, F.
Isahak, R. M. Yusop, W. M. Mohamed Hisham and A. M. Qi, Q. Fu and X. Bao, Science, 2016, 351, 1065–1068.
Yarmo, Ind. Eng. Chem. Res., 2016, 55, 7888–7897. 108 T. Inui, Catal. Today, 1996, 29, 329–337.
92 T. Riedel, H. Schulz, G. Schaub, K.-W. Jun, J.-S. Hwang and 109 P. Gao, S. G. Li, X. N. Bu, S. S. Dang, Z. Y. Liu, H. Wang,
K.-W. Lee, Top. Catal., 2003, 26, 41–54. L. S. Zhong, M. H. Qiu, C. G. Yang, J. Cai, W. Wei and Y. H.
93 M. K. Gnanamani, G. Jacobs, H. H. Hamdeh, W. D. Shafer Sun, Nat. Chem., 2017, DOI: 10.1038/nchem.2794.
and B. H. Davis, Catal. Today, 2013, 207, 50–56. 110 J. K. Jeon, K. E. Jeong, Y. K. Park and S. K. Ihm, Appl.
94 Y. Zhang, G. Jacobs, D. E. Sparks, M. E. Dry and B. H. Catal., A, 1995, 124, 91–106.
Davis, Catal. Today, 2002, 71, 411–418. 111 C. Li, X. Yuan and K. Fujimoto, Appl. Catal., A, 2014, 475,
95 R. W. Dorner, D. R. Hardy, F. W. Williams, B. H. Davis and 155–160.
H. D. Willauer, Energy Fuels, 2009, 23, 4190–4195. 112 Y. K. Park, K. C. Park and S. K. Ihm, Catal. Today, 1998, 44,
96 C. G. Visconti, L. Lietti, E. Tronconi, P. Forzatti, R. Zennaro 165–173.
and E. Finocchio, Appl. Catal., A, 2009, 355, 61–68. 113 T. Inui, K. Kitagawa, T. Takeguchi, T. Hagiwara and Y.
97 A. N. Akin, M. Ataman, A. E. Aksoylu and Z. I. Önsan, React. Makino, Appl. Catal., A, 1993, 94, 31–44.
Kinet. Catal. Lett., 2002, 76, 265–270. 114 M. Fujiwara, R. Kieffer, H. Ando and Y. Souma, Appl.
98 R. E. Owen, J. P. O'Byrne, D. Mattia, P. Plucinski, S. I. Catal., A, 1995, 121, 113–124.
Pascu and M. D. Jones, Chem. Commun., 2013, 49, 115 J. Ereña, J. M. Arandes, R. Garoña, A. G. Gayubo and J.
11683–11685. Bilbao, J. Chem. Technol. Biotechnol., 2003, 78, 161–166.
99 W. W. Russell and G. H. Miller, J. Am. Chem. Soc., 1950, 72, 116 X. Ni, Y. Tan, Y. Han and N. Tsubaki, Catal. Commun.,
2446–2454. 2007, 8, 1711–1714.
100 M. K. Gnanamani, W. D. Shafer, D. E. Sparks and B. H. 117 R. X. Bai, Y. S. Tan and Y. Z. Han, Fuel Process. Technol.,
Davis, Catal. Commun., 2011, 12, 936–939. 2004, 86, 293–301.
101 H. M. Torres Galvis, A. C. J. Koeken, J. H. Bitter, T. 118 X. Wang, G. Yang, J. Zhang, S. Chen, Y. Wu, Q. Zhang, J.
Davidian, M. Ruitenbeek, A. I. Dugulan and K. P. de Jong, Wang, Y. Han and Y. Tan, Chem. Commun., 2016, 52,
J. Catal., 2013, 303, 22–30. 7352–7355.

4598 | Catal. Sci. Technol., 2017, 7, 4580–4598 This journal is © The Royal Society of Chemistry 2017

You might also like