You are on page 1of 19

Results in Engineering 16 (2022) 100711

Contents lists available at ScienceDirect

Results in Engineering
journal homepage: www.sciencedirect.com/journal/results-in-engineering

Methanol synthesis from CO2: A mechanistic overview


Noerma J. Azhari a, 1, Denanti Erika b, 1, St Mardiana b, Thalabul Ilmi c, Melia L. Gunawan a, c,
I.G.B.N. Makertihartha a, c, Grandprix T.M. Kadja b, c, d, *
a
Department of Chemical Engineering, Institut Teknologi Bandung, Jl. Ganesha No. 10, Bandung, 40132, Indonesia
b
Division of Inorganic and Physical Chemistry, Faculty of Mathematics and Natural Sciences, Institut Teknologi Bandung, Jl. Ganesha No. 10, Bandung, 40132,
Indonesia
c
Center for Catalysis and Reaction Engineering, Institut Teknologi Bandung, Jl. Ganesha No. 10, Bandung, 40132, Indonesia
d
Research Center for Nanosciences and Nanotechnology, Institut Teknologi Bandung, Jl. Ganesha No. 10, Bandung, 40132, Indonesia

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

Keywords: Currently, energy demand is growing swiftly with global economic growth. The combustion of fossil fuels is one
Methanol synthesis of the major environmental threats due to the release of CO2 into the atmosphere. Reducing greenhouse gases is
CO2 hydrogenation one of the most critical challenges facing society today. It has been proposed that converting CO2 into fuels and
Reaction pathway
valuable chemicals such as methanol for reducing greenhouse gas emissions can be a favorable solution. This
Cu-based catalyst
Precious metal-based catalyst
review describes the recent development of methanol synthesis from CO2 from a mechanistic point of view.
Oxide-based catalyst Several proposed reaction pathways for methanol formation over various catalysts will be emphasized, i.e., Cu-
Oxygen deficient-based catalyst based catalyst, precious metal-based catalyst, and other catalysts, including oxide-, oxygen deficient-, and
Composite-based catalyst composite-based catalysts such as metal alloys, and supported metal alloys. Additionally, the intermediate
involved, active site, and several aspects affecting the catalytic behavior are also provided. Finally, a remaining
challenge and the perspective for compelling the energy future vision is emphasized.

1. Introduction water using renewable energies, e.g., wind, geothermal, hydro, and solar
[7]. Several important chemical products could be obtained from this
In recent years, many industrial fields used fossil fuels, which process, e.g., methane [8,9], alkanes [10], carbon monoxide [11], ole­
contribute to greenhouse gas (GHG) emissions, especially carbon diox­ fins [12], and alcohol [13]. Among the possible CO2 reduction products,
ide (CO2) [1–3]. Consequently, it raises climate change issues, an urgent methanol is one of the important compounds of C1 chemistry because
global challenge to resolve [4,5]. Compared with the GHG level in 2010, approximately more than 110 Mt/year of methanol is required to fulfill
CO2 reduction of 41–72% by 2050 and 78–118% by 2100 is necessary to its demand in many areas. For instance, methanol acts as a chemical
preserve the temperature change below 2 ◦ C relative to preindustrial intermediate and molecule platform for the synthesis of important
levels [6]. In this case, converting CO2 into valuable products through commodities and products such as formaldehyde, acetic acid, methyl
several processes is one of the promising strategies for reducing its tert-butyl ether (MTBE), and gasoline (MTG) [14,15] (Fig. 1). It also
emission. For instance, electrocatalytic and photocatalytic CO2 reduc­ could be an excellent substitute or additive for fuel and modified diesel
tion is subjected as a potential process for future. However, the recent engines due to their high octane rating [15]. Furthermore, methanol
techniques are still incapable to meet large-scale industrial applications could generate electricity in direct oxidation methanol fuel cells (DMFC)
requirements. At present, thermo-catalytic is still a more feasible [16]. Therefore, methanol production from CO2 benefits from environ­
method for industrial application. mental and economic points of view. Generally, a catalyst is needed to
In the thermo-catalytic reaction process, CO2 hydrogenation has transform the CO2 into a methanol product because CO2 is a highly
gained much attention. Hydrogen (H2) is a suitable reagent to assist the stable molecule. Many types of catalysts can assist the CO2 conversion
CO2 conversion, in which it can be generated from the electrolysis of into methanol, but metal and/or metal oxide-based catalysts are mostly

* Corresponding author. Division of Inorganic and Physical Chemistry, Faculty of Mathematics and Natural Sciences, Institut Teknologi Bandung, Jl. Ganesha No.
10, Bandung, 40132, Indonesia.
E-mail address: grandprix.thomryes@itb.ac.id (G.T.M. Kadja).
1
These authors contributed equally to this work.

https://doi.org/10.1016/j.rineng.2022.100711
Received 9 August 2022; Received in revised form 12 October 2022; Accepted 16 October 2022
Available online 19 October 2022
2590-1230/© 2022 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-
nc-nd/4.0/).
N.J. Azhari et al. Results in Engineering 16 (2022) 100711

2. Methanol synthesis from CO2

Commercially, methanol is synthesized from natural gas through the


syngas route. It contains two processes, i.e., (i) the steam reforming of
methane and (ii) methanol production. The former is the most expensive
step in the production process chain since it proceeds at very high
temperatures (1173 K) and pressure (16–30 bar) [39]. This reaction
follows eqs. (1) and (2), resulting in a mixture of CO, CO2, and H2. Also,
CO2 may form through WGS (Water-Gas Shift) reaction (eq. (4)) [15].
This reaction is usually used to adjust the CO2/H2 ratio in the feed for the
methanol synthesis reaction [40]. The latter process is usually per­
formed at 250–300 ◦ C, 5–10 MPa, in the presence of CuO/ZnO/Al2O3 as
a catalyst [15]. In this case, the methanol product could be obtained
either from CO or CO2 as the reactant (eq. (3) and (5)).
CH4 + H2 O ↔ CO + 3H2 ΔH ◦ = 206 kJ/mol (1)

CH4 + 2H2 O ↔ CO2 + 4H2 ΔH◦ = 165 kJ/mol (2)

Fig. 1. Methanol consumption by industries. CO2 + 3H2 ↔ CH3 OH + H2 O ΔH◦ = − 49.5 kJ/mol (3)

CO + H2 O ↔ CO2 + H2 ΔH◦ = − 41 kJ/mol (4)


used to facilitate the redox reaction, i.e., Cu [17–21], Zn [22–25], Cr
[26,27], and Pd [28–30]. Furthermore, other materials for the promoter CO + 2H2 ↔ CH3 OH ΔH◦ = − 90.5 kJ/mol (5)
of support have been reported.
Many researchers have made significant progress on the CO2 hy­ Producing methanol from CO2 is a promising pathway because it
drogenation to methanol products. The publications related to these benefits both the economics and environmental point of view. It con­
topics gradually grew year by year (Fig. 2). Also, Several reviews are tributes to resolving climate change issues by reducing greenhouse gas
already available to provide extensive insight from various points of emissions. Moreover, it could produce various high-value-added prod­
view. Some reviews only focused on one type of catalyst [31–33]. Others ucts with remarkable economic value although the realization of this
provide a broader range of catalyst types [34,35]. Furthermore, the process still faces many challenges. From the thermodynamic point of
review based on the catalyst design, thermodynamic, kinetic, and the view, CO2 is more stable than CO. The heat formation of CO2 was − 393
development of technical aspects of this reaction has also been provided kJ/mol. Meanwhile, CO is − 110 kJ/mol. Also, CO2 has the high bond
[15,36–38]. Herein, we intend to focus on the mechanistic aspect of the energy of the C– – O double bond (750 kJ/mol). It is larger than that of
methanol synthesis from CO2 hydrogenation over metal-based catalysts, C–C (336 kJ/mol), C–O (327 kJ/mol), or C–H bond (411 kJ/mol) [41].
including Cu-, precious metal-, and other catalysts (composite-, Oxygen Therefore, activating the CO2 molecule needs a much higher energy
deficient- and oxide-based catalyst). input, appropriate reaction conditions, and the presence of the catalyst
Furthermore, the active site, the intermediate involved, and several [34].
factors affecting the reaction pathways will be discussed. An in-depth In addition, the CO2 to methanol reaction is very exothermic and
understanding of this aspect is crucial since it could be used to kinetically limited to 15–25% conversion [42]. CO2 may also convert
improve many aspects of this process. For instance, the build of an into CO product through RWGS reaction, which is endothermic. In that
appropriate synthesis strategy to design an efficient catalyst or the ki­ sense, the reactor model should be designed efficiently. The reactor
netic model for industrial-scale development could be generated. In that should be capable of removing heat released from the methanol syn­
sense, this review benefits from bringing various perspectives regarding thesis reaction to keep the reactor in isothermal condition [43].
the catalytic behavior of several types of catalysts through the recent Furthermore, water is produced as a byproduct of the methanol syn­
development from the mechanistic point of view. Finally, challenges and thesis from CO2. Its presence can cause the deactivation of the catalyst,
insight for future development regarding the sustainable development of especially for the conventional copper-based catalyst [44]. Therefore,
CO2 hydrogenation to methanol will be emphasized. catalyst and reactor design, as well as kinetic study, are crucial for
realizing the methanol production from CO2.

Fig. 2. Numbers of published papers related to CO2 hydrogenation and metal catalyst topics indexed by Scopus on January 25, 2022.

2
N.J. Azhari et al. Results in Engineering 16 (2022) 100711

Various kinetic models were currently reported in the literature on catalyst design has been made, deeper insight into the mechanistic
using either Eley-Rideal (ER) or Langmuir – Hinshelwood (LH) mecha­ aspect of the catalysis system should gain serious attention since it
nism. The conversion of CO2 into methanol products generally could provides essential information to generate rational strategies for the
occur via formate, RWGS, or trans-COOH* pathway (Fig. 3). For the design of the next-generation catalyst.
classical formate pathway, CO2 reacted with adsorbed atomic H and In the Cu-based-catalyst, Cu provides the active sites. Meanwhile,
formed the formate (HCOO*). Then, it gradually transforms into support facilitates the dispersion of active metal Cu. During the catalysis
dioxomethylene (H2COO*), formaldehyde (H2CO*), methoxy (CH3O*), process, Cu should be reduced to generate Cu0 or Cu+ as a dominant
and methanol (CH3OH), respectively. However, others reported that species in the catalyst, which is believed to be an active site for methanol
instead of dioxomethylene (H2COO*), formate species are preferentially synthesis. In that sense, the reaction could occur either on one or two of
hydrogenated into formic acid (HCOOH*). Then, it transforms into those surfaces. Moreover, it has been reported that other active centers
H2COOH*, which continuously splits to generate formaldehyde and OH may generate from the interaction between active metal and its support
species. Finally, methanol was produced from this formaldehyde species [47]. Therefore, it further improved the catalytic performance. For
via methoxy intermediate [45,46]. instance, the formation of the Cu–ZrO2 interface could enhance the
On the other hand, methanol formation via the RWGS pathway began methanol production rate by facilitating CO2 activation and stabilizing
from the transformation of CO2 into CO* species, then it subsequently the intermediates, such as CO and formaldehyde [48]. The synergetic
hydrogenated into methanol through the formation of formyl (HCO*) → effect between metal-support interfaces was also demonstrated by Arena
formaldehyde (H2CO*) → methoxy H3CO* species. This pathway could et al. [49] using Cu–ZnO/ZrO2 as the catalyst. The promotion effect was
also proceed with the hydrocarboxyl species formation (COOH*) fol­ related to the excellent capability of ZrO2 and ZnO to enhance the CO2
lowed by the CO* and the same species with the previous pathway. The affinity and increase the surface availability for generating the reactive
hydrocarboxyl (COOH*) species was generated from the reaction be­ formate intermediate.
tween CO2 with H atoms which is provided by H2O. Therefore, some­ Further, the effect of the zirconia phase in Cu-based catalyst was
times it is called an H2O-mediated mechanism. Some researchers investigated by Witoon and Coworker [50]. In this work, they used three
reported that rather than CO* species production, dihydroxycarbene different phases of ZrO2, i.e., amorphous (a-), tetragonal (t-), and
(COHOH*) was formed as an active species after hydrocarboxyl for­ monoclinic (m-) phases, on Cu/ZrO2 catalyst for converting CO2 to
mation. Then, dihydroxycarbene was dissociated into the hydrox­ methanol. The amorphous Cu/ZrO2 exhibits the highest yield of meth­
ymethylidyne (COH*) species. It further transforms into anol related to its highest Cu surface area. However, the highest turn­
hydroxymethylene (HCOH*) → hydroxymethyl (H2COH*), and finally, over frequency was owned by the Cu/t-ZrO2 catalyst. As demonstrated
methanol product. This pathway was then known as the trans-COOH by Cu K-edges XANES spectra in Fig. 4a, only the monoclinic phase of
mechanism [45,46]. The mechanisms above were affected by various ZrO2 resembles the CuO reference spectra. Contrarily, both amorphous
factors, such as the catalyst properties and reaction conditions. We will and tetragonal phases of ZrO2 exhibit the change of the local catalyst
bring these discussions to the following section. structure toward the octahedral structure (also convinced by the shifting
of the derivative XANES spectra). Therefore, it is further justified that
3. Catalysts for conversion of CO2 into methanol these two phases have a stronger interaction with Cu. It facilitates the
spillover of H2 from Cu on the ZrO2 surface, leading to high adsorption
3.1. Cu-based catalysts of H2 (high ratio of hydrogen to CO2) as revealed by CO2- and H2-TPD
characterization results (Fig. 4b and c).
Cu-based catalyst is the most used catalyst in methanol synthesis. It is Generally, CO2 hydrogenation towards methanol over Cu-based
usually synthesized from its salt through the coprecipitation method in catalyst was believed to occur via the formate mechanism. Grabow
the presence of basic precipitation agents, such as carbonates, bi­ and Mavrikakis [21] performed the DFT calculation of the methanol
carbonates, or hydroxides. This reaction was conducted in an aqueous production on commercial Cu-based catalysts, i.e., Cu/ZnO/Al2O3,
medium. Also, the following process, such as aging, calcination, and especially on the Cu (111) site. This experiment used CO2 and CO as
reduction, may be performed [46]. Currently, Cu-based catalyst with reactants to investigate each component’s role in the reaction. Then, the
various support and promoter for methanol synthesis from CO2 has been microkinetic model was built from the DFT calculation. The result
reported, e.g., Cu/ZnO, Cu/CeO2, Cu-ZO/ZrO2. Although much progress showed that the CO2 hydrogenation occurs via the formate pathway
(CO2* → HCOO* → HCOOH* → CH3O2* → CH2O* → CH3O* →
CH3OH*). Meanwhile, the methanol production from CO follows the
sequence of CO* → HCO* → CH2O* → CH3O* → CH3OH*. For the
CO2-rich feed, the hydrogenation of CH3O* is the rate-determining step.
Meanwhile, the formation of CH3O* is the RDS in CO-rich feed. The
reaction network of methanol synthesis and WGS reaction is shown in
Fig. 5.
Moreover, based on the DFT calculation, CO2 hydrogenation con­
tributes 2/3 of the methanol product. This result was the opposite of the
potential energy surface (PES) analysis (Fig. 6a), which showed that
methanol should be more produced from CO hydrogenation. It is
reasonable because CO hydrogenation is more exothermic than CO2
hydrogenation, and the high energy barrier was owned by CO2 hydro­
genation (CH3O* hydrogenation (black line) and OH* hydrogenation to
form H2O* (blue line after state 3). These phenomena were related to the
promotional effect of CO on CO2 hydrogenation. CO* is more consumed
for the HCO* production, and it could be the H-donor to assists the CO2
hydrogenation [21].
On the other hand, Zhao et al. [52] emphasize the role of water in
methanol synthesis. They demonstrated that in the presence of water,
Fig. 3. Schematic representation of several proposed reaction mechanisms of the formation of hydrocarboxyl (trans-COOH) species on the Cu (111)
methanol formation from CO2. site is kinetically favorable than formate species. It occurs via a hydrogen

3
N.J. Azhari et al. Results in Engineering 16 (2022) 100711

Fig. 4. (a) Cu K-edges XANES Spectra of Cu/ZrO2 catalyst and reference compound. (b) H2-TPD (c) CO2-TPD profiles of Cu/ZrO2 catalysts. Reproduced with
permission from Ref. [50]. Copyright 2016 Elsevier.

Fig. 5. The reaction network of methanol synthesis and WGS reaction over Cu
catalyst. Reproduced with permission from Ref. [21], Copyright 2011 American
Chemical Society.

transfer mechanism. This intermediate is then transformed step by step


Fig. 6. (a) Potential energy surface for methanol synthesis reaction from CO
into dihydroxycarbene (COHOH) → hydroxymethylidyne (COH) →
and CO2. Reproduced with permission from Ref. [21], Copyright 2011 Amer­
hydroxymethylene (HCOH) → hydroxymethyl (H2COH), and finally,
ican Chemical Society. (b) Potential energy surface for the most probable re­
methanol product. This result was confirmed by computational and action pathway in CO2 hydrogenation to methanol over Cu-based catalyst with
experimental methods. The DFT calculation showed that the decompo­ a different promoter. Reproduced with permission from Ref. [51], Copyright
sition of COHOH to COH and OH is the rate-determining step with an 2017 Royal Society of Chemistry.
energy barrier of 0.68 eV. It is much lower than the formate mechanism,
in which the two highest energy barriers are 1.20 and 1.17 eV. Further,
the mechanism of methanol synthesis in other facets of Cu catalyst, i.e.,

4
N.J. Azhari et al. Results in Engineering 16 (2022) 100711

Cu (100) and Cu (110) surfaces, have also been reported by Higham hydrogenation.
et al. [53] using the density functional calculation. Results show that Huš et al. [51] investigated the mechanism of the Cu-based catalyst
those two facets play a role in CO2 activation. Methanol synthesis could with the different promoters, i.e., Zn3O3/Cu, Cr3O3/Cu, Fe3O3/Cu, and
occur via HCOO* and COOH*. Furthermore, the barrier energy of the Mg3O3/Cu, using Ab initio plane-wave density functional theory cal­
initial hydrogenation process on those sites shows a significantly lower culations and the Kinetic Monte Carlo simulations. The result showed
value, resulting in improved catalytic behavior for CO2 dissociation and that all the catalyst follows the formate pathway mechanism, although

Fig. 7. (a) Methanol and CO concentration of outlet on the temporal of the reaction and the phase-resolved DRIFT spectra obtained from a modulated-excitation
experiment for the catalyst which was first pre-treated with CO/H2 (i, iii) and CO2/H2 (iii, iv). Reproduced with permission from Ref. [59], Copyright Wiley
2016. (c) Schematic proposed mechanism of methanol formation over CuO/CeO2/ZrO2. Reproduced with permission from Ref. [60], Copyright Elsevier 2022.

5
N.J. Azhari et al. Results in Engineering 16 (2022) 100711

with the variation in the surface intermediate fraction and adsorption and dissociation of H2. As the hydrogen is dissociated, it will
rate-determining step (Fig. 6b). Also, Zn3O3/Cu catalyst results in high hydrogenate the adsorbed CO2 to form a formate intermediate.
catalytic activity. The understanding of the active site of the Cu–Zn Sequentially, methanol was generated from a series of hydrogenation
catalyst has also been studied by Kattel et al. [54]. In this experiment, processes of CO2 or CO which is produced by RWGS (Fig. 7c). Besides
they used ZnCu and ZnO/Cu as model catalysts to carefully study the ZnO, ZrO2, and CeO2, Au has also been reported to enhance the catalytic
active site during the methanol production from CO2. The result showed activity of Cu-based catalysts on CO2 hydrogenation to methanol.
that methanol formation over both catalysts is via formate intermediate. Adding 1% wt of Au could decrease the 10% mass of CO selectivity. This
Furthermore, the presence of Zn or ZnO in ZnCu and ZnO/Cu catalysts improvement was assigned to the formation of copper-gold interfaces,
assists the system in stabilizing the *HCOOH intermediates via Zn–O which facilitate the hydrogen spillover and enhance the CO and H2
interaction. Also, it enhances the *HCOO activation via the hydroge­ adsorption capacities [63].
nation process. Lam et al. [55] pointed out that the excellent activity in Engineering the catalyst surface properties is one of the crucial fac­
Cu-based catalysts originated from the metal-oxide interface, although tors in obtaining the selective intermediate. Generally, it relates to the
the type of support may affect the selectivity and product distribution. availability of oxygen vacancy and the density of basic site. In the study
For instance, using Al2O3 for Cu support could further convert methanol reported by Marcos et al. [64], preparing CuO–ZnO–ZrO2 catalyst
to dimethyl ether due to the Lewis acid availability. through the coprecipitation method assisted with Pluronic P123 suc­
In commercial methanol synthesis, adding a certain amount of CO2 cessfully improves its catalytic activity. The number of basic sites
to the syngas feed could enhance the methanol synthesis rate [56–58]. increased with the surfactant/Cu–Zn–Zr molar ratio, which is beneficial
Pérez-Ramírez and Co-Worker [59] found that the volcano-type diagram for CO2 adsorption. Based on the investigation by Singh et al. [65], a
was obtained when partially replacing the CO with CO2 feed. The high density of basic site and oxygen vacancies are needed to obtain high
space-time yield of methanol achieved a maximum value when the CO2 selectivity in the hydrogenation reaction of CO2 to methanol. Among
volume content in the feed was 2.4%. Under this optimal mixture, the oxides of CeO2, ZnO, and ZrO2, the combination of Cu with CeO2
sintering between the components of the catalysts could be avoided; resulted in the highest basicity and oxygen vacancy. This basicity plays a
therefore, it enhances the activity of methanol production. Further, they role in species formation during the reaction (Fig. 8) [66]. In low ba­
investigated the effect of this CO2 promotion. In their work, the com­ sicity materials, CO2 adsorption generates bicarbonate species which are
mercial methanol catalyst, Cu–ZnO–Al2O3, was employed. Meanwhile, very easy to desorb into CO2. At materials with medium basicity, CO2
the operando synchrotron x-ray powder diffraction and adsorption produces bidentate carbonate species with hydrogen atoms
modulated-excitation infrared spectroscopy were applied to elucidate adsorbed on Cu, producing CO and H2O through the RWGS reaction
the promotion mechanism. Fig. 7a shows the time-dependant trans­ [67]. While the absorption of CO2 on the materials with higher basicity
formation of CO and methanol during the hydrogenation reaction of CO2 results in unidentate carbonate species where hydrogenation of this
and the optimum mixture of CO + CO2. The catalyst was first pre-treated species will produce methanol through the formate pathway involving
with CO2/H2 (Fig. 7a (i & ii)) and CO/H2 (Fig. 7a (iii & iv)). Under the H2COO, H2COOH, and H3COOH intermediates [68].
CO2/H2, the catalyst size increase and the platelet-like structures of ZnO Very recently, Han et al. [69]. reported that a surface with a large
was formed, in which it contains a high amount of polar facets. It further amount of basic site could be obtained by generating the hollow struc­
enhances the electronic interaction with Cu and preventing the forma­ tured of Cu@ZrO2 from the Cu-loaded Zr-MOF pyrolysis process. This
tion of active sites. Therefore, the RWGS reaction could be suppressed. It structure could improve the adsorption capacity, CO2 activation, and
also confirmed by the Phase-resolved DRIFT spectra obtained from a selective hydrogenation toward methanol products due to the highly
modulated-excitation experiment (Fig. 7a (ii & iv)) which reveal that dispersed Cu nanoparticles coupled with the balanced Cu0/Cu+ sites and
smaller CO adsorption was observed in the CO2/H2-pretreated sample. the abundances of Cu–ZrO2 interfaces. This interface is crucial for the
Christensen and Coworker [61] have also investigated the role of CO intermediate formation, in which it facilitates the CO2 adsorption to
and CO2 at varying conversion levels over Cu/ZnO/Al2O3 catalysts. The form the carbonate (CO3*) and/or bicarbonates (HCO3*) species as
result shows that under very low conversion, CO inhibits Raney Cu observed by in situ FTIR spectroscopy. Furthermore, the appearance
because of the competitive adsorption at the Cu surface. Conversely, the peak associated with formate, methoxy, and methanol species with
beneficial impact of CO was obtained at higher conversion conditions by exposure of time suggests the occurrence of methanol synthesis via
lowering the gas space velocity. It was assigned to the water removal HCOO intermediate.
capability of CO through the water-gas shift reaction because water is a Despite all the factors mentioned above, the size of the Cu cluster was
stronger inhibitor at higher conversion conditions than CO. They also also mentioned to affect the catalytic activity of the Cu-based catalyst in
imply that at the lower conversion condition, the effect of both oxygen methanol synthesis. To investigate this behavior, Yang et al. [70] syn­
vacancy in ZnO and Cu–Zn surface alloy sites from the ZnO is not thesized the Cun/Al2O3 with different clusters of Cu, i.e., n = 3, 4, and
critical. 20. Results showed that the catalytic activity of the catalyst on CO2
On the contrary, Zhu et al. [62] have demonstrated the opposite hydrogenation is strongly related to the Cu cluster with the order of Cu3
effect of CO2 addition in the syngas feed of the methanol synthesis over < Cu20 < Cu4. Furthermore, it was related to the reduced phase in the
Cu–CeO2 catalyst. The presence of CO2 retarded the reaction due to the catalyst. In this case, the temperature needed for each catalyst to
formation of carbonate-like species on the interface site of Cu–CeO2. initially reduce is Cu3 > Cu20 > Cu4. Furthermore, the catalyst activity
Besides, this active site was needed to hydrogenate CO to obtain was also correlated to its charge state. The higher charge state of Cu in
methanol products via a formyl intermediate. The methanol was ob­ Cun/Al2O3 catalyst leads to weaker binding with intermediates, thus
tained from two processes, i.e., the direct hydrogenation of CO2, which resulting in low activity per atom. In the nano-sized scale of the
occurred on the copper site, and the CO intermediate-mediated hydro­ Cu-based catalyst, the methanol selectivity was reduced with the
genation, which involved the Cu–CeO2 interface as the active site. The decrease of the copper clusters. The larger copper dan ZnO particles
latter reaction was much faster. However, it is worth noting that the could generate the new active sites, leading to a higher intrinsic meth­
mechanism should be affected by the reaction condition, as recently anol production rate and simultaneously suppressing the RWGS reaction
reported by Poto et al. [60], in which CuO/CeO2/ZrO2 catalyst was used. [71].
A similar contribution of CO2 and CO (51.5% and 47.4%) was found at From the numerous development and the mechanism proposed,
30 bar, 200 ◦ C, and 3 of the H2: CO2 molar ratio. While at higher tem­ designing the Cu-based catalyst for CO2 hydrogenation to methanol
peratures and H2: CO2 ratio, the contribution of CO is predominant due product is still challenging regarding the low stability of Cu in the
to the Cu0 site saturation. The oxygen vacancy of CeO2–ZrO2 was presence of water byproduct and coke formation. Furthermore, the
responsible for CO2 adsorption, and the metallic copper facilitated the presence of another strong competitive reaction, i.e., RWGS, makes Cu-

6
N.J. Azhari et al. Results in Engineering 16 (2022) 100711

Fig. 8. CO2 adsorption at the surface of materials with different basic streght. Reproduced with permission from Ref. [65], Copyright 2021 Elsevier.

based catalyst difficult to obtain a simultaneously high conversion and uniform nanoparticles (Fig. 9a). Moreover, Cu nanoparticles are highly
selectivity of methanol product. In that sense, combining Cu with suit­ dispersed and surrounded by the Zr–Ce oxide solid solution. The inti­
able support could enhance its ability to activate the very stable CO2 mate contact between these two components produces strong metal
molecule. The strong interaction between Cu and support could enhance support interaction (SMSI) without aggregation. Meanwhile, Cu nano­
the reducibility of the catalyst. Furthermore, the metal-support inter­ particle aggregation was observed for coprecipitation and impregnation
action generates the interface that could stabilize the intermediates methods. Further analysis also demonstrated the presence of defective
involved during the reaction. Therefore, finding other new support Ce3+− Ov–Zr4+ structures (Ov: Oxygen vacancy) and Cu+− O− M (M =
materials and/or alternative preparation method are still needed. Ce, Zr), leading to enhanced methanol formation because of the
Wang et al. [72] introduced a new method called a facile microfluid improved formate intermediate transformation. Meanwhile, the
film-assisted coprecipitation method to improve the Cu dispersion on medium-strength and strong basic sites of Ce3+− O and Cu+− O pairs
the Zr–Ce oxide solution catalyst. The rapid nucleation of ternary were responsible for CO2 adsorption, and Cu0 sites play a role in the
Cu–Zr–Ce is assisted by the H2 bubble obtained from the hydrolysis of activation of hydrogen species (Fig. 9b).
NaBH4 in the micro-liquid film reactor, resulting in the formation of Very recently, the dual mechanism has also been reported for a new

Fig. 9. (a) Schematic description of the microfluid film assisted coprecipitation for synthesis Zr–Ce oxide solid solution surrounded by Cu nanoparticles (b) the
proposed mechanism of methanol for mation from CO2 hydrogenation over Zr–Ce oxide solid solution surrounded by Cu nanoparticles. Reproduced with permission
from Ref. [72] Copyright 2021 American Chemical Society.

7
N.J. Azhari et al. Results in Engineering 16 (2022) 100711

Cu-based catalyst, Cu–ZnO/nanosheet Pr2O3. The RWGS pathway and Cr2O3, SiO2, TiO2, ZnO, and ZrO2. The different activity between these
formate pathway were believed to simultaneously occur based on the in- oxides is associated to their ability to stabilize the Pdn+ (0 < n < 2)
situ DRIFT analysis. Similar to the previous results, the Cu0 was species. Also, it provides an optimal amount of Pdn+. Other researchers
responsible for the hydrogen dissociation at the initial step. Meanwhile, suggested that the enhanced catalytic activity of Pd/Ga2O3 catalyst was
the Cu–Pr2O3 interface facilitates the CO2 adsorption and reacts with the related to the Pd− Ga intermetallic site, which facilitates the methanol
adsorb hydrogen to form formate species through the for­ formation [80], especially the hydrogenation of bidentate formate
mate→methoxy→methanol route (blue route) (Fig. 10). On the other (b-HCOO) intermediates [81]. Moreover, further study revealed that the
hand, the Cu–Pr2O3 interface could also adsorb the CO from the RWGS formation of intermetallic PdxGa species depends on the reduction
reaction in the meantime. Continuously, a series of hydrogenation of this temperature and the different phases of Ga2O3. For instance, on β-Ga2O3,
species generates formyl, formaldehyde, methoxy, and finally, meth­ Pd2Ga was observed at temperature 673 K, and PdGa was detected if the
anol. The occurrence of the CO transformation is associated with the reduction was performed at 773 K. Meanwhile, almost 100% Pd2Ga was
abundance of surface oxygen vacancy and metal-support interfacial site. observed on α-Ga2O3 and γ-Ga2O3 at the 523 K and 673 K temperature
The existence of the oxygen vacancy could be able to generate metal reduction, respectively [82,83]. It is reasonable because the different
atoms with low-coordination numbers and more dangling bonds, phases of Ga2O3 may have different stability under the reduction
resulting in enhanced adsorption energy. The increase of the adsorption condition.
energy of CO and CO2 at these surfaces reveals that those species could García-Trenco et al. [75] developed colloidal Pd2Ga nanoparticle
be easily adsorbed, accelerating the methanol formation [73]. catalysts to hydrogenate CO2 to methanol. This catalyst was prepared
through a two-step procedure, i.e., thermal decomposition reaction of
3.2. Precious metal-based catalysts Pd(OAc)2 (palladium acetate) and Ga(OSt)3 (gallium stearate), which
heating to 190 ◦ C using N2 flow, followed by the reduction of the colloid
Aside from Cu-based catalysts, precious metal-based catalysts (e.g., (Pd/Ga mixture) that utilize a dilute H2 flow (5% H2/N2, 0.5 MPa) at a
Au, Pd, and Pt) have also been tested in the methanol synthesis from CO2 selected temperature (Fig. 11a). As a result, Pd2Ga nanoparticles were
hydrogenation. These precious metals have attracted significant atten­ successfully synthesized with an average size of 5–6 nm. Results show
tion in CO2 to methanol conversion thanks to their high activity in the that the rate of methanol and the TOF increase with the increase of Ga
adsorption and dissociation of H2 [74]. As the active metal, palladium content in the catalyst (maximum at Pd: Ga molar ratio = 1:2) (Fig. 11b
(Pd) exhibits expectant results for methanol synthesis [75,76]. However, and c). Both Pd2Ga and Ga2O3 phase has a crucial role in methanol
it is hard to produce methanol using the pure Pd site. Thus, usually, it formation. However, their work did not provide a detailed mechanism
combines with other metals as promoters or support. In this sense, the and the role of each active phase. The previous study by other re­
catalytic activity and methanol selectivity are influenced by the type of searchers suggests that the Ga2O3 surface facilitates CO2 adsorption to
promoters and support [29]. form bicarbonate species. This species was sequentially transformed into
Several oxides have been used as support for Pd-based catalysts, e.g., methanol through the formates, methylenebisoxy, and methoxy inter­
ZnO [77], CeO2 [78], and Ga2O3 [75,79]. The research on Ga2O3-­ mediate species. This ability was related to its excellence in the disso­
supported palladium (Pd/Ga2O3) for methanol synthesis from CO2 hy­ ciation and chemisorption of H2. Meanwhile, the Pd–Ga promote the
drogenation was pioneered by Fujitani et al. [79]. In their work, the spillover for hydrogen supply and increase the bonded carbonaceous to
catalyst was prepared through the coprecipitation method. The result Ga2O3 phase [84,85].
showed that Pd/Ga2O3 catalyst generates better results (turnover fre­ More recently, Collins et al. [86] reported the Pd-Ga-supported
quency (TOF) 20-fold higher) than Cu/ZnO. This result is also higher mesoporous SiO2 as an efficient catalyst for methanol production from
than those of Pd on other metal oxide support materials such as Al2O3, CO2. Pd–Ga/SiO2 has a lower activation energy (~40 kJ/mol) than

Fig. 10. Proposed mechanism of the CO2 hydrogenation toward methanol over Cu–ZnO/nanosheet Pr2O3. Reproduced with permission form Ref. [73], Copyright
2022 American Chemical Society.

8
N.J. Azhari et al. Results in Engineering 16 (2022) 100711

Fig. 11. (a) Representation of synthesis Pd2Ga nanoparticles. (b) Methanol rate and selectivity and (c) TOF for samples with varying the Pd:Ga molar ratios at
290 ◦ C. Cu–ZnO–Al2O3 data was included as the benchmark. Adapted with permission from Ref. [75], Copyright 2017 American Chemical Society.

Pd/SiO2 (60 kJ/mol), with a 200-fold turnover frequency. Character­ significant increase of the formation methano rate, suggesting the
ization using in situ transmission infrared spectroscopy suggests that thin importance of the intermetallic site in methanol formation. However, at
layer gallium oxide (Ga2O3) facilitates CO2 adsorption to form poly­ the ratio Pd/Ga <1, the exact stoichiometry could not be achieved,
dentate carbonate species, which are further transformed into methanol resulting in a higher amount of Ga2O3 on the surface; thus, the exposed
via hydrogen spillover. Additionally, Pd2Ga bimetallic particles act as intermetallic surface was reduced, and the lower reaction rate was
the catalytic site for H2 dissociation for providing atomic hydrogen to observed.
the Ga2O3 site. Thus, the intimacy between Ga2O3 and Pd2Ga surface is The spillover capability of Pd benefits methanol production; there­
essential for enhancing catalytic performance. fore, Pd has also been used for promoters in Cu-based catalysts, as re­
Although the synergetic effect between Pd and Ga has been reported, ported by Choi et al. [78]. Pd addition on methanol catalyst, Cu/CeO2,
it should be noted that adjusting the molar ratio between the two could enhance the catalytic activity due to the improvement of the Cu
components is also essential. Manrique et al. [81] found that the molar site reducibility through hydrogen spillover property. This electron-rich
ratio 1 of Pd/Ga is the optimum value for obtaining the improved cat­ metal site tends to be more favorable to interacting with CO2 resulting in
alytic activity of the Pd–Ga/SiO2 catalyst. This result was confirmed the formation of the O– – C–O− group and COOH intermediates (RWGS
using XPS and Operando DRIFTS analysis, and the schematic illustration pathway). As studied by Liu and coworkers [42], the formate pathway
of the catalyst surface with various Pd/Ga molar ratios was depicted in was dominant only for Au-doped Cu(111). Meanwhile, for the Pd, Rh Pt,
Fig. 12. At the high Pd/Ga ratio, the Pd phase was detected as the and Ni, the RWGS pathway is faster.
monometallic form along with the Ga2O3 phase covered the Pd sites, Conversely, the formate pathway is more acceptable for Pd–Zn
thus inhibiting the methanol formation since Pd was responsible for catalyst. In the study reported by Bahruji et al. [76], PdZn could stabilize
bidentate formate adsorption. On the other hand, the intermetallic the formate intermediate, leading to high methanol production. PdZn
Pd–Ga was formed only for Pd/Ga ≤ 1. This was followed by a also impedes the CO formations. Meanwhile, Pd metal induces the

Fig. 12. Schematic Illustration of the effect Pd: Ga ratio on the Pd–Ga/SiO2 catalyst surface during CO2 hydrogenation (a) Pd/Ga >1.0 (b) Pd/Ga = 1.0, (c) Pd/Ga
<1. Reproduced with permission from Ref. [81] Copyright 2020 Royal Society of Chemistry.

9
N.J. Azhari et al. Results in Engineering 16 (2022) 100711

decomposition of formate intermediates due to its ability to dissociate Another precious metal, such as Au, is also the catalyst candidate for
H2 but is unlikely to adsorb CO2. Moreover, the controlling of Pd and CO2 hydrogenation to methanol. Nevertheless, it still needs a deeper
PdZn particle size was very crucial for the improvement of their catalytic understanding of its mechanism reaction. Generally, Au-based catalysts
performance. In this case, the choice of a suitable synthesis procedure is are widely used thanks to their high catalytic activity for many reactions
essential. For instance, the sol-immobilized method for Pd/ZnO syn­ with mild conditions, including the selective hydrogenation of various
thesis presents a more eminent way of controlling Pd and PdZn particles organic molecules, the water gas shift (WGS) reaction, and CO oxidation
size than the impregnation method, resulting in a higher CO2 [88]. For CO2 hydrogenation, metal oxide-supported Au catalysts were
conversion. reported as a potential catalyst to form methanol [89]. The Au/ZnO
The formate pathway of the Pd–Zn catalyst was also convinced by catalysts result in the selectivity of methanol product >50%, compara­
Zabilskiy et al. [77] through FTIR, XAS, and XRD characterization. The ble to Cu/ZnO/Al2O3 catalyst (37%). Besides, the results also showed
FTIR results show peaks at 1596 cm− 1 and 1376 cm− 1, corresponding to the catalyst stability with no deactivation throughout >6 h of reaction.
the active formate species attached to Zn. Moreover, the XAS result Furthermore, other oxides have also been reported to support the Au-
suggests the formation of possible intermediates of Pd-formate, in which based catalyst for CO2 hydrogenation. Rezvani et al. [90] reported the
there is no Pd–Zn de-alloying or Pd’s surface oxidation. Notably, the CO2 remarkable activity of Au/CeO2 catalysts. Also, the mechanistic aspect
activation occurs in ZnO site. The Pd–Zn alloy phase did not provide the was investigated through diffuse reflectance infrared fourier transform
active site for direct methanol formation from CO2. Instead, it is spectroscopy (DRIFTS). This catalyst has similar properties to Au/ZnO
responsible for H2 activation phase and allows the subsequent hydro­ catalysts, i.e., a partial reduction of the support as the activation phase as
genation process of formate into methanol on Zinc oxide surfaces. This well as the Auδ− sites formation and the reaction pathways through the
proposed mechanism is similar to that of Cu/ZnO based system. formation with the hydrogenation of surface formate and methoxy
In the case of the combination of PdZn with other metals, such as species. Fig. 13 exhibits the in situ DRIFTS spectra of the Au/CeO2
CdSe, the formate pathway is still reported as the dominant pathway catalyst for observing the temporal evolution of the nature and relative
[87]. The combination of CdSe and PdZn renders the stronger interac­ coverages of the different adsorbed species.
tion of Pd/Zn and more stable HCOO* species as intermediate over As seen, the temporal evolution of DRIFT spectra was correlated to
COOH*. In this sense, it led to the enhancement of methanol production. their catalytic performance (Fig. 13). For the first 15 min, the band at
Furthermore, the XPS result showed that the increment of CdSe con­ ~3853 cm related to the O–H group from adsorbed carboxylic group
centration is associated with the ZnO concentration enhancement. Also, was enhanced, then it decreased after a longer reaction time. It agrees
they found that the methanol selectivity over Pd@Zn is 80%, which was with the methanol formation rate (Fig. 13f and g). It is correlated with
significantly higher than that of the commercial Cu catalyst (10%) in the the band at 2850 cm− 1 (v (C − H) stretch mode of formate-like species),
same reaction condition. Moreover, it reveals that the Pd@Zn catalyst as depicted in Fig. 13b. The weak band at 2711-2720 cm− 1 shows the
inhibits the CO production over RWGS reaction. intensity enhancement through the first 10 min of reaction, which is

Fig. 13. The sequence DRIFT of CO2 hydrogenation over 1 wt % Au/CeO2 on the regions: (a) O–H region, (b) C–H region, (c) C–H region, (d) C– –O region, (e) OCO
bending region with the time scale 5 s, 80 s, 95 s, 110 s, 125 s, 140 s, 4.5 min, 7.5 min, 13 min, 15 min, 30 min, 1 h, 2 h, 5 h, 11 h, and 15 h. This measurement was
conducted at 5 bar. (f) Au-mass-normalized formation rates of methanol on Au/CeO2. (g) Normalized peak intensity of DRIFT spectra and Au-mass-normalized
formation rates of methanol and CO of Au/CeO2 catalyst. Adapted with permission from Ref. [90], Copyright 2020 American Chemical Society.

10
N.J. Azhari et al. Results in Engineering 16 (2022) 100711

assigned to the CH vibration of adsorbed formyl group. Unfortunately, it over 70% at temperatures below 225 ◦ C and 275 ◦ C, respectively. This
is tough to discover if this band disappears due to the strong band at result is almost 20 times higher than the previously reported Au catalyst
2609 cm− 1. The CO stretch and OCO bending regions exhibit an in­ supported by other oxides, such as CeO2 and ZnO, due to the strong
tensity increment (Fig. 13d and e). It is noteworthy that the slow interaction of Au and its support. This work concluded that Au/In2O3
deactivation is caused by the blocking site through adsorbed carbonate could enhance the catalytic performance of Au catalysts.
species. Moreover, the Au/CeO2 catalyst denotes lower selectivity for The different surface properties of the catalyst could significantly
methanol and longer reaction times than Au/ZnO. impact the catalytic activity because it directly interacts with the reac­
Rui et al. [91] reported the Auδ+− In2O3-x catalyst for hydrogenating tant molecules and proceeds the reaction. In that sense, the choice of
CO2 to methanol. The Au electronic structure affects the reactant preparation strategy is essential. Sagar et al. [92] investigate the effect
adsorption and activation regarding the hydrogenation activity. of different preparation methods on the catalyst surface and catalytic
Furthermore, the interfacial sites of Au and its support are also essential performance of Au/ZrO2. Results showed that the best methanol pro­
in the product distribution determination by adjusting the intermediate ductivity was obtained for the catalyst prepared by deposition rather
stabilization. The results show that the methanol selectivity is 100% and than the impregnation method since it produces a large surface area and

Fig. 14. (a) The products acquired by using MIL-101, Pt1@MIL (1 = Pt single atom), Ptn@MIL (n = nanocrystal counterparts), and Cu/ZnO/Al2O3 (as a benchmark)
in CO2:H2 = 1:3 at 150 ◦ C 32 bar after 1 h. (b) Accumulation products obtained by in-situ cycles over Pt1@MIL. (c) The mechanistic insight of Pt1@MIL activity. (d)
Quasi-situ XPS spectra for MIL-101, Pt1@MIL, and Ptn@MIL. (e) and (f) Before/after C and O K-edge XANES spectra for MIL-101, Pt1@MIL, and Ptn@MIL,
respectively. Adapted with permission from Ref [94], Copyright 2019 Nature.

11
N.J. Azhari et al. Results in Engineering 16 (2022) 100711

high-loading Au with controllable particle size. Moreover, the trans­ realization on an industrial scale. Furthermore, even if the atomic scale
formation of amorphous to cubic ZrO2 and the enhanced acid and basic of metal species could be obtained and showed remarkable activity, it
site generated from the deposition method provide the appropriate might change its state during the CO2 hydrogenation process toward the
environment for stabilizing the critical intermediate involved in meth­ larger species due to the reduction or sintering process. Therefore,
anol synthesis, i.e., HCOO*, dioxomethylene, and H3CO*. Whilst the low creating a noble catalyst, especially a single-atom noble metal catalyst
CO2 uptake was observed for the catalyst prepared by the impregnation that stabilizes its phase and state during the reaction process, is chal­
method due to the presence of bulk Au cluster, rendering the lower lenging. In that sense, the search for an appropriate synthesis strategy
methanol formation rate. and/or the improvement of reaction conditions is absolutely needed.
Another precious metal catalyst was reported by Tran et al. [93]
using platinum impregnated onto Nb2O5. The precursor of Pt was pre­ 3.3. Other catalysts
pared from H2PtCl6 dissolution in water and subsequently added into
Nb2O5 synthesized. The characterization exhibit that Pt nanoparticles CO2 hydrogenation to methanol is equilibrium-limited by high
are distributed homogeneously on the surface of Nb2O5 as support. temperature. Generally, a catalyst efficiently operates at high tempera­
Furthermore, the result showed that the acidity controlled by different tures, which requires exceptional thermal stability [54,97–100].
calcination temperatures could increase the metal-support interaction Therefore, if catalysts other than Cu and precious metal-based catalysts
and metal dispersion. Hence, it enhances CO2 conversion and has the are to be used, they must strictly follow the thermal stability require­
highest selectivity of 0.8% at temperatures of 200 ◦ C. ment. In this sense, other catalytic systems with novel catalytic struc­
Furthermore, rather than nanoparticle form, the single atom form of tures, such as composite-based catalysts (metal alloys and supported
Pt was reported to be more reactive for methanol production from CO2. metal-alloys), oxygen-deficient materials, or oxide-based catalysts,
The work reported by Y. Chen et al. [94] reveal that Pt single atoms have been reported as effective catalysts for CO2 hydrogenation to
coordinated with the oxygen atom in MIL-101 (Pt1@MIL) generate methanol.
methanol selectivity up to 90.3% with the formic acid as a byproduct. It In terms of composite-based catalysts, several elements, such as Ni,
is higher than that of Pt as the nanoparticle (Ptn@MIL) and the com­ In, and Mo, have been investigated. Richard and Fan [101] synthesized a
mercial Cu/ZnO/Al2O3 (Fig. 14a). Moreover, the activation energy of composite catalyst, NiaInbAl (NIA)/SiO2, using the
Pt1@MIL (76.4 kJ/mol) was much lower than that of Ptn@MIL(119.7 deposition-precipitation method for the CO2 hydrogenation to methanol
kJ/mol). Pt1@MIL also exhibits the long-term stability shown by pro­ at ambient pressure. Fig. 15a depicts the methanol formation using a
ducing 14.6 mmol of methanol and 1.4 mmol of formic acid After 16 series of NIA catalysts compared to the common CuO/ZnO/Al2O3 (CZA)
in-situ cycles (96 h) (Fig. 14b). catalyst. As seen, NIA/SiO2 catalyst with a Ni: In ratio of 0.7 and 15 wt%
Fig. 14c and d shows that HCOO* and COOH* are not detected in of NIA phase (15 wt% NIA-0.7) exhibits the highest catalytic activity
MIL-101, which are considered intermediates for Pt1@MIL and with a methanol formation rate of 0.33 mol h− 1 mol catalyst metal− 1
Ptn@MIL, respectively. In the reaction using Pt1@MIL as a catalyst, the (Fig. 15a). Combining the Ni with In results in higher methanol synthesis
first H atom adds to the C atom, followed by HCOO* formation as the activity than Ni or In alone and allows it to be reduced at lower tem­
stable intermediates. Meanwhile, in Ptn@MIL, the formation of COOH* peratures. Furthermore, the optimal composition of the Ni and In cata­
is favored, which acquires CO as the major product. The C K-edge lyst was crucial for the catalyst structure formation with a high particle
XANES (Fig. 14e) shows that before treatment with H2 and CO2, dispersion and appropriate particle size distribution after the process of
Pt1@MIL and Ptn@MIL exhibit the same XANES profile. However, after the catalyst reduction.
treatment, it exhibits the more robust feature of C–– O (π*) indicating the The time-dependent infrared spectroscopy was utilized to elucidate
formation of CO2. Moreover, the σ* (O–H) ends up significantly more the reaction mechanism over 15 wt% NIA-0.7 catalysts (Fig. 15b). The
intense for Ptn@MIL but stays the same for Pt1@MIL (Fig. 14f). Thereby, peak located at 1662 cm− 1 is assigned to the monodentate formate (m-
Ptn@MIL with COOH* intermediates could lead to gaseous products’ HCOO) species. Meanwhile, the peaks at 1377 cm− 1, 1576 cm− 1, and
formation. 2857 cm− 1 correspond to the asymmetric- (vas (CO2)), symmetric
Sun et al. [95] reported Pt/In2O3 for catalyzing the methanol for­ stretching vibrations of CO2 (vs (CO2)), and stretching vibration of C–H
mation from CO2. Notably, In2O3 enriched with oxygen vacancies bonding (v(CH)) within the bidentate formate (b-HCOO) species,
showed high activity and selectivity for CO2 activation and hydroge­ respectively. Furthermore, the peaks at 1549 cm− 1 and 1384 cm− 1 are
nation to methanol. Moreover, the support Pt catalyst also exposition a attributed to the asymmetric- (vas (CO3)), symmetric stretching vibra­
promising activity for CO2 hydrogenation. This work exhibited that the tions of carbonate (vs (CO3)) within the monodentate carbonate (m-
Pt/In2O3 catalyst has a remarkable selectivity in the CO2 hydrogenation CO2− 2−
3 ) species. Other carbonate species, bidentate (b-CO3 ) and poly­
reaction with methanol selectivity of ca. 100%, 74%, and 54% at tem­ dentate (p-CO3 ), are indicated by the presence of peaks at 1597 cm− 1
2−

peratures <225 ◦ C, 275 ◦ C, and 300 ◦ C, respectively. Moreover, this (vas (CO3) of b-CO2−3 )), 1478 cm
− 1
(vas (CO3) of p-CO2−
3 )), 1437 cm
− 1
(vs
catalyst shows high stability, in which the activity decreases to 95% (CO3) of p-CO2− 3 ). These results revealed that CO2 hydrogenation over
after 5 h. NIA/SiO2 catalyst occurs via monodentate carbonate, followed by the
Han et al. [96] evaluated the effect of Pt amount on the In2O3 per­ formation of bigger carbonate species, i.e., bidentate and polydentate
formance and its mechanism for methanol synthesis. The result showed carbonates. Moreover, changing the reactant feed from CO2/H2 to CO/
that a small amount of Pt addition into In2O3 through coprecipitation H2 changes the primary hydrocarbon product from methanol to
could enhance both activity and selectivity to methanol products. Dur­ propane.
ing CO2 hydrogenation, a part of the Pt species with a positive charge is Differently, Ren at al [102]. suggested the CO2 → CO3* → HCOO*
reduced and sintered, resulting in the formation of Pt nanoparticles, pathway for the methanol synthesis over unsupported bimetallic Ni–Cu
while the other remains dispersed at the atomic level in a positively (100) catalyst. Detail characterization using Near-Ambient Pressure
charged state (Ptn+). Both species actively assist in more oxygen vacancy X-Ray Photoelectron spectroscopy reveals that the Ni/Cu(100) facili­
formation, on which CO2 is activated. On the other hand, the Ptn + tates the CO2 dissociation into CO and atomic oxygen, which chemisorb
species served as Lewis acid sites in promoting the heterolytic dissoci­ at the surface. During the CO2 activation, the CO3* species will be
ation of H2 and the methanol formation. Contrarily, Pt nanoparticles formed due to the reaction between CO2 and surface oxygen. This spe­
primarily encourage the RWGS reaction, which predominately assists cies was further converted into formate in the presence of hydrogen.
the CO formation and reduces the selectivity to methanol. Furthermore, introducing CO to the CO2 hydrogenation system could
Apart from the excellent performance of precious metal based, pro­ help maintain the catalyst activity because the CO capability reduces the
ducing the Nobel-based catalyst is still expensive, especially for surface oxide that has been proved to retain the CO2 activation.

12
N.J. Azhari et al. Results in Engineering 16 (2022) 100711

Fig. 15. (a) The catalytic performance of NIA-0.5 and NIA-0.7 catalyst at 5–20 wt % loading, and the commercial CuO/ZnO/Al2O3 catalyst in methanol production at
elevated temperature and ambient pressure, (b) Infrared spectra of isothermal CO2 hydrogenation over NIA-0.7 catalyst at 15% loading as a function of time.
Reproduced with permission from Ref. [101]. Copyright 2017 American Chemical Society.

Indium, as the nano-intermetallic form, was also a potential catalyst molybdenum carbide (MoC) at Au, its catalytic performance could be
for methanol production from CO2. Chen et al. [103] synthesized improved, showed by its smaller activation energy (14 kcal/mol) than
InNi-based catalysts through RWGS reaction-oriented pre-design com­ that of its bulk phase (17 kcal/mol) and Cu catalyst system (25 kcal/­
bined with experimental verifications and atomistic-theoretical calcu­ mol). Furthermore, the composition between carbon and metal is very
lations. Based on the DFT calculations, nano-intermetallic InNi3C0.5 can crucial. For the carbon-poor (MoC0.6) carbide nanomaterials supported
reduce the reaction temperature from 400 ◦ C–600 ◦ C–300 ◦ C and below. Au, the CO2 decomposition tends to produce CO and CH4. In contrast,
The transition metal carbides have a good capability for adsorbing CO2 the carbon-rich one ((MoC1.1)/Au(111)) could produce a significant
and breaking the C–O either by themselves or aided by hydrogen. Re­ amount of methanol with no methane. Methanol formation occurs via
sults show that the 1–8% CO2 conversion could be obtained with the formate pathway with the Eley Riedel mechanism. The direction of
methanol selectivity of 60%–98% under 200–300 ◦ C. Initially, CO2 was product selectivity by engineering the catalyst structure properties has
adsorbed on 3Ni–In site and dissociated into CO* and O* on 3Ni–C sites also been documented by Zhou et al. [109]. A two-dimensional multi­
and 3Ni–In sites, respectively. This CO intermediate is further converted layered molybdenum carbide (2D-Mo2C) synthesized from Mo2CTx
into methanol product via HCO* → CH2O* → CH2OH* → CH3OH*. Mxene family is the potential catalyst for producing methanol from CO2.
Molybdenum carbide (Mo2C) is also reported as a good candidate for The terminal Tx group could be tuned to selectively direct the reaction
methanol production from CO2. Adding alkali promoters was expected toward the target, although recent reports suggest that under the H2
to enhance the catalytic performance of Mo2C. Dongil et al. [104] pretreatment, the 2D-Mo2C is still more favorable for RWGS reaction.
evaluated the addition of Cu and Cs to the Mo2C catalyst system. In their However, the broad chance for a methanol catalyst is still opened.
work, Cs is expected to enhance the methanol formation since it has a Besides the metallic composite catalyst, the oxide catalyst, such as
large ionic radius; thus, it can be a better electron donor [105–107]. Indium (In)-based catalyst, also showed promising results. Indium oxide
However, results show the opposite effect due to the occurrence of redox (In2O3) was first tested in 2016 and had an extremely high methanol
surface transformation under the reaction condition. XRD analysis ex­ selectivity [110–112]. Martin et al. [110] reported that In2O3 has a
hibits the presence of the Mo0 phase in the Cs–Mo2C catalyst, which has highly efficient catalytic system and outstanding activity under indus­
poor ability in CO2 hydrogenation. Although, recent reports have found trial conditions for the CO2 to methanol with 100% selectivity at all
that supporting the molybdenum carbide in appropriate materials could temperatures. It is much higher than that of the conventional ternary
improve their catalytic performance in methanol formation. catalyst (Cu/ZnO/Al2O3), which only reached 47% selectivity at 473 K.
Figureas et al. [108] reported that by supporting the nanoparticles of The surface oxygen vacancies of indium oxide promote the selectivity of

13
N.J. Azhari et al. Results in Engineering 16 (2022) 100711

[bi-H2CO + H]* → H2COH* ΔE = +0.83 eV; Ea = 1.33 eV (9)


CO2 hydrogenation to methanol, and it could be increased through CO
co-feeding into the gas feed or using support, e.g., ZrO2. Furthermore, [H3CO + H]* → CH3OH* ΔE = +0.21 eV; Ea = 0.33 eV (10)
this hybrid catalyst is stable even after a 1000 h reaction and could be
upgraded to a technical form with similar catalytic performance. [H2COH + H]* → CH3OH* ΔE = − 0.59 eV; Ea = 2.52 eV (11)
Ye et al. [111] previously predicted the excellent catalytic perfor­ Because of their excellent performance, many studies have been
mance of the In2O3 catalyst in methanol synthesis from CO2 using pe­ conducted to improve their catalyst performance. Perez-Ramirez and
riodic DFT calculations. The initial reaction of CO2 hydrogenation to Coworker [113] evaluate the behavior of In2O3 with different forms, i.e.,
HCOO shows that the calculated reaction energies and activation bar­ bulk-form, In2O3 supported ZrO2, and In2O3 promoted with Pd and Ni.
riers are thermodynamically and kinetically favorable on the D4 Additionally, the effect of CO in the feed stream was evaluated. Results
defective In2O3 (110) surface (Fig. 16b). Afterward, the HCOO hydro­ show that the presence of CO could alter the surface properties of the
genation reaction to H2CO has a slightly endothermic reaction, then the catalyst. The improvement of methanol productivity upon CO addition
H2CO hydrogenation reaction that produced CH3O is more favorable for was obtained for the In2O3 supported on the monoclinic ZrO2 catalyst. It
methanol formation. The oxygen vacancy assists the activation and was attributed to the controlled formation of oxygen vacancies and the
hydrogenation of CO2 as well as stabilizes the critical intermediate re­ ability to restrict sintering during the reaction. Meanwhile, the bulk
action involved in CO2 hydrogenation to methanol (Fig. 16a). The main In2O3-form remains unaffected. For the Ni- and Pd-promoted indium
issue in the CO2 hydrogenation to methanol is the C═O bond cleavage, oxide form, moderate-to-substantial deactivation was observed due to
which can only occur by dissociating intermediates. Furthermore, in the the strong ability of Pd to adsorb CO, the over-reduction, and the sin­
In2O3 system, it can only be achieved by filling the oxygen vacancy with tering between metal promoters and oxide.
one of the O atoms and making this step much easier than on the surface Furthermore, the different phase of In2O3 was also reported to pro­
of the metal catalyst. Detail reaction steps and their calculated energies duce different catalytic behavior in methanol synthesis. Dang et al.
are provided in eq. (6)-(11). [114] evaluate the CO2 hydrogenation pathway over the In2O3 catalyst
[CO2 + H]* → mono-HCOO* ΔE = − 0.21 eV; Ea = 0.15 eV (6) with different phases, i.e., cubic In2O3 (c-In2O3) and corundum-type
hexagonal phase (h-In2O3). Additionally, various exposed surfaces
[HCOO + H]* → bi-H2CO*+Surface O ΔE = +0.05 eV; Ea = 0.57 eV (7) were also investigated, i.e., (110) and (111) surfaces for c-In2O3 and
(104) and (012) surfaces for h-In2O3. The best catalytic performance for
[mono-H2CO + H]* → H3CO* ΔE = − 0.60 eV; Ea = 1.14 eV (8)
methanol synthesis (CO2 conversion >17% with 92.4% methanol
selectivity) was owned by h-In2O3, with the abundance of exposed (104)
surface confirming the DFT calculation results. In situ DRIFT charac­
terization reveals the stronger ability of h-In2O3 (104) to stabilize the
HCOO* and CH3O* species as the key intermediate; thus, higher meth­
anol selectivity could be gained.
Aside from the oxygen vacancy that has been explained previously,
the sulfur vacancies have shown a promising result recently. For
instance, MoS2 is considerably reported as the dominant active center
because of its edge sites in many catalytic reactions [115–119]. Hu et al.
[120] proved that the in-plane sulfur vacancies of MoS2 is the excellent
active centers for CO2 hydrogenation to methanol at the
low-temperature reaction (180 ◦ C). Sulfur vacancies in MoS2 nanosheets
facilitate the dissociation of the CO2 molecule to produce surface-bound
CO and O at ambient temperature, thus enabling the CO2 hydrogenation
reaction to methanol at low temperatures. Notably, methanol selectivity
and CO2 conversion of 94.3% and 12.5%, respectively, were obtained
over the surface of in-plane sulfur vacancy-rich MoS2 nanosheets, which
are specifically higher than previously reported catalysts. Moreover, this
catalyst has high stability, which could be used for over 3000 h without
deactivation.
Molybdenum in the atomic form was also accounted for producing
methanol from CO2 hydrogenation. In the reported work by Len et al.
[121], the atomically dispersed and reducible oxomolybdate species on
TiO2 support results in 35 gMeOH.kg− 1. h− 1 of STY under 275 ◦ C and 3
MPa. Again, the phase of the support affects the catalytic behavior due to
the different oxidation states of Mo. Characterization using
synchrotron-based NAP-XPS shows that the rutile phase of TiO2 support
possesses more reduced Mo than the anatase phase. Under reaction
conditions, TiO2 with a rutile structure has a more labile oxygen envi­
ronment; thus, the methanol productivity is higher.
The recent development of the non-Cu-based and non-precious-
based catalysts that have been enormously studied, i.e., the pure oxide
catalyst, intermetallic, and transition metal carbide catalyst, showed
that these catalysts provide a more promising result regarding lowering
the operation condition. For instance, the sulfur vacancies catalyst could
Fig. 16. (a) Proposed mechanism of methanol formation from CO2 on the be operated for the reaction under 200 ◦ C. Therefore, many explorations
defective In2O3 (110) surface with surface oxygen vacancies, (b) Profiles of of these types of catalysts should be under shed light. One should be paid
potential energy for the generation of oxygen vacancy, 1st and 2nd step of CO2 attention is the catalyst mechanism, both in the reaction process and the
hydrogenation on the D1 (red line) and the D4 surfaces (blue line). Reproduced deactivation mechanism. Also, developing a synthesis strategy to
with permission from Ref. [111]. Copyright 2013 American Chemical Society.

14
N.J. Azhari et al. Results in Engineering 16 (2022) 100711

selectively produce the catalyst with a particular phase or exposed facet direction could be formulated. The catalyst species should be precisely
surface is essential. Therefore, the pre-design-based development will be adjusted to conveniently activate CO2 molecules and stabilize a specific
more visible. intermediate of the reaction. The addition of suitable promoters and
supports may assist the stability of the intermediate species involved in
4. Conclusion and outlook the reaction through the metal-support interaction. Moreover, the
structural evolution of the catalyst and its active sites during the reaction
CO2 conversion to methanol is a promising way to realize a sus­ is also crucial in improving the catalytic performance. Notably, it is
tainable energy future. Although Cu-based catalyst was commonly used essential to adjust the morphology, particle shape, and size, as well as
for this process, this catalyst often suffered from the drawback of low the phase of each component, since it affects the electronic structure and
stability; therefore, other catalysts should be considered. Other than Cu, the activity performance of the catalyst. In that sense, developing the
several types of catalysts have been used for directly producing meth­ synthesis strategy is crucial to be encouraged. For instance, escalating
anol from CO2, i.e., precious metal-based catalysts and other catalysts, the nanocrystal catalyst to single atoms [94] exhibits a much higher
such as composite-based catalysts (metal alloys and supported metal- catalytic activity. However, it is still challenging to engineer the struc­
alloys), oxygen-deficient materials, or oxide-based catalyst. Although tures and accessibility of the high density of single atoms while pre­
all these catalysts show remarkable catalytic performance, the occurring venting aggregation towards a large particle. Moreover, powerful
mechanism remains elusive. characterization techniques, such as X-ray absorption spectroscopy
Lately, researchers have conducted numerous experiments focusing (XAS) and/or high-angle annular dark-field scanning transmission
on the mechanistic aspect of the reaction since it is crucial information electron microscopy (HAADF-STEM), are inevitable. On the other hand,
for developing the next-generation catalyst. At least three plausible developing catalysts that could be operated in milder conditions is also
mechanisms have been proposed, i.e., via formate, RWGS, and trans- still wide-opened.
COOH* pathways. Table 1 summarizes the various catalysts used for Computational studies, such as DFT-based calculations and molec­
methanol synthesis from CO2. As seen, the formate pathway is consid­ ular dynamics simulations, have shed light on the plausible mechanisms
ered the most common route. However, several researchers have sug­ of selective CO2 conversion to methanol. In addition, several in-situ
gested the preference for the trans-COOH* pathway. This is because the characterizations should be involved to characterize the intermediate
hydrogenation of CO2 commonly results in water as the byproduct that species and the behavior of the catalyst during the reaction. Meanwhile,
could react with CO2 to generate hydrocarboxyl species. On the other machine learning (either unsupervised or supervised) could be utilized
hand, the RWGS, usually called RWGS + CO hydro pathway, is also to predict and pre-design the catalyst, as well as determine the appro­
reported for several precious metal-based catalysts, such as Pd–Cu/CeO2 priate reaction condition to produce a more efficient route. From the
and Pt/Nb2O5. However, catalytic behavior is affected by many factors; technological perspective, combining other units, such as the CO2 cap­
therefore, it is difficult to classify catalytic behavior precisely. For ture units [122], the in-situ water removal [123], or the heat-released
instance, the physicochemical properties of the metal and support used, equipment, may improve the thermodynamic equilibrium and other
i.e., state, cluster size, phase, composition, morphology, and basicity, is aspects for compelling the vision of a sustainable energy future.
strongly related to their catalytic performance, in which the discussion Last but not least, the industrialization of the methanol synthesis
of each component effect has been provided in the previous section. process from CO2 is still challenging, although significant progress has
Furthermore, the operation condition, such as the different tempera­ been attained on the laboratory scale. The challenge comes from several
tures and pressure, CO2: H2 ratio, and the presence of CO co-feed in the aspects, especially from the techno-economics point of view. Studies
stream, was also reported to affect the catalytic behavior. that have been conducted by several researchers show that the current
Finally, after the immense development of CO2 conversion to process of methanol production from syngas is still more advantageous
methanol from the mechanistic point of view, several concerns for future related to the techno-economics and life cyle analysis consideration.

Table 1
Various catalyst that have been reported for CO2 hydrogenation to methanol and their correspondingctalytic performance.
Type of catalyst Active site Reaction condition (MPa; Gas ratio (CO2: Conversion Selectivity Ref.

C) H2) (%) (%)

Formate pathway
Cu–ZnO/ZrO2 Cu+/Cu0; oxide-metal interface 1.0; 180 1: 3 ~1.5 ~83 [49]
Cu/tetragonal ZrO2 ZrO2; Cu–ZrO interfaces 3.0; 240 1: 3 ~2 ~80 [50]
hollow structured of Cu0/Cu+ sites; Cu–ZrO2 interfaces 3.0; 150 - 300 1: 3 ~2–18 80–20 [69]
Cu@ZrO2
Cu–CeO2 Cu–CeO2 interfacial site; copper site 3.0; 250 1: 3 ~1.5 ~50 [62]
CuO/CeO2/ZrO2 CeO2–ZrO2; metallic copper 3.0; 200 1: 3 ~5 – [60]
Pd/ZnO PdZn bimetallic phase 2.0; 250 1: 3 11 60 [76]
Pd/ZnO ZnO and palladium-zinc alloy 3.0; 260 1: 3 3.6 49.8 [77]
Au/ZrO2 Zr4+/O2− 4.0; 240 3: 1 5.7 72.5 [92]
Pt1@MIL Pt 3.2; 150 1: 3 – 90.3 [94]
In2O3 (110)* D4 surface (O defective site) – – – – [111]
Bimetallic Ni–Cu (100) Ni–Cu surface 4.0 x 10-4 ; 27–277 ◦ C 1: 3 – – [102]
Cu–Ce–Zr oxide solid Cu0, Cu–O-M (M; Zr, Ce) interfacial sites, and 3.0; 240 1:3 ~6 ~60 [72]
solution Ce3+− Ov–Zr4
Pd2Ga and Ga2O3 phase 5.0; 210 3:1 – 60 [75]
Pd/Ga 3.0; 250 3: 1 – 65 [85]
Au 0.5; 240 1: 3 – 53 [90]
RWGS pathway
Pd–Cu/CeO2 Cu sites 3.0; 190 3: 1 16.1 26.7 [78]
Pt/Nb2O5 Pt 0.5; 280 1: 3 5.7 0.8 [93]
Nano-intermetallic 3Ni–In; 3Ni–C 4.0; 220–310 1: 3 1–8% 60–98 [103]
InNi3C0.5
Trans-COOH* pathway
Cu (111)* Cu (111) – – – – [52]

15
N.J. Azhari et al. Results in Engineering 16 (2022) 100711

Under the current techno-economic condition, methanol production via bifunctional catalysis, ACS Catal. 8 (2018) 571–578, https://doi.org/
10.1021/acscatal.7b02649.
from CO2 did not meet the lower cost compared with the current
[13] F. Studt, I. Sharafutdinov, F. Abild-pedersen, C.F. Elkjær, J.S. Hummelshøj,
methanol market price due to the electricity and hydrogen requirements S. Dahl, I. Chorkendorff, J.K. Nørskov, Discovery of Ni-Ga catalyst for carbon
[124]. However, the environmental benefits will be gained if the elec­ dioxide reduction to methanol, Nat. Chem. (2014) 1, https://doi.org/10.1038/
tricity source comes from low-carbon energy, suggesting the importance nchem.1873. –5.
[14] K.A. Ali, A. Zuhairi, A.R. Mohamed, Recent development in catalytic technologies
of renewable energy to be swiftly developed. for methanol synthesis from renewable sources : a critical review, Renew. Sustain.
Energy Rev. 44 (2015) 508–518, https://doi.org/10.1016/j.rser.2015.01.010.
[15] S.G. Jadhav, P.D. Vaidya, B.M. Bhanage, J.B. Joshi, Catalytic carbon dioxide
Credit author statement
hydrogenation to methanol: a review of recent studies, Chem. Eng. Res. Des. 92
(2014) 2557–2567, https://doi.org/10.1016/j.cherd.2014.03.005.
Noerma J. Azhari: Writing – original draft, Visualization, Denanti [16] G.A. Olah, Towards Oil Independence through Renewable Methanol Chemistry,
2013, pp. 104–107, https://doi.org/10.1002/anie.201204995.
Erika: Writing – original draft, Visualization, St Mardiana: Writing –
[17] F.C.F. Marcos, F.M. Cavalcanti, D.D. Petrolini, L. Lin, L.E. Betancourt, S.
original draft, Visualization, Thalabul Ilmi: Writing – review & editing, D. Senanayake, Effect of Operating Parameters on H2/CO2 Conversion to
Melia L. Gunawan: Writing – review & editing, Supervision, I.G.B.N. Methanol over Cu-Zn Oxide Supported on ZrO2 Polymorph Catalysts :
Makertihartha: Writing – review & editing, Supervision. Grandprix T. Characterization and Kinetics, 2022, p. 427, https://doi.org/10.1016/j.
cej.2021.130947.
M. Kadja: Writing – review & editing, Supervision, Conceptualization, [18] H.W. Lim, M.J. Park, S.H. Kang, H.J. Chae, J.W. Bae, K.W. Jun, Modeling of the
Resources. kinetics for methanol synthesis using Cu/ZnO/Al2O3/ZrO2 catalyst: influence of
carbon dioxide during hydrogenation, Ind. Eng. Chem. Res. 48 (2009)
10448–10455, https://doi.org/10.1021/ie901081f.
Declaration of competing interest [19] K. Chen, H. Fang, S. Wu, X. Liu, J. Zheng, S. Zhou, X. Duan, Y. Zhuang, S. Chi
Edman Tsang, Y. Yuan, CO2 hydrogenation to methanol over Cu catalysts
supported on La-modified SBA-15: the crucial role of Cu–LaOx interfaces, Appl.
The authors declare that they have no known competing financial Catal. B Environ. 251 (2019) 119–129, https://doi.org/10.1016/j.
interests or personal relationships that could have appeared to influence apcatb.2019.03.059.
[20] F. Studt, F. Abild-Pedersen, Q. Wu, A.D. Jensen, B. Temel, J.D. Grunwaldt, J.
the work reported in this paper. K. Norskov, CO hydrogenation to methanol on Cu-Ni catalysts: theory and
experiment, J. Catal. 293 (2012) 51–60, https://doi.org/10.1016/j.
Data availability jcat.2012.06.004.
[21] L.C. Grabow, M. Mavrikakis, Mechanism of methanol synthesis on cu through
CO2 and CO hydrogenation, ACS Catal. 1 (2011) 365–384, https://doi.org/
Data will be made available on request. 10.1021/cs200055d.
[22] V. Deerattrakul, N. Yigit, G. Rupprechter, P. Kongkachuichay, The roles of
nitrogen species on graphene aerogel supported Cu-Zn as efficient catalysts for
Acknowledgments CO2 hydrogenation to methanol, Appl. Catal. Gen. 580 (2019) 46–52, https://doi.
org/10.1016/j.apcata.2019.04.030.
This work is supported by the Ministry of Education, Culture, [23] A. Ye, Z. Li, J. Ding, W. Xiong, W. Huang, Synergistic catalysis of Al and Zn sites
of spinel ZnAl2O4 catalyst for CO hydrogenation to methanol and dimethyl Ether,
Research and Technology of the Republic of Indonesia through PDUPT ACS Catal. 11 (2021) 10014–10019, https://doi.org/10.1021/acscatal.1c02742.
Research Funding (first year) 2022. We dedicate this work to the [24] M. Sadeghinia, A. Nemati Kharat Ghaziani, M. Rezaei, Component ratio
retirement of Prof. Subagjo, who has made an outstanding contribution dependent Cu/Zn/Al structure sensitive catalyst in CO2/CO hydrogenation to
methanol, Mol. Catal. 456 (2018) 38–48, https://doi.org/10.1016/j.
to the development of catalyst and catalysis science in Indonesia.
mcat.2018.06.020.
[25] R. Burch, S.E. Golunski, M.S. Spencer, The role of copper and zinc oxide in
References methanol synthesis catalysts, J. Chem. Soc. Faraday. Trans. 86 (1990)
2683–2691, https://doi.org/10.1039/FT9908602683.
[26] A.V. Kirilin, J.F. Dewilde, V. Santos, A. Chojecki, K. Scieranka, A. Malek,
[1] M. Aresta, A. Dibenedetto, A. Angelini, Catalysis for the valorization of exhaust
Conversion of synthesis gas to light olefins: impact of hydrogenation activity of
carbon : from CO2 to chemicals, materials, and fuels. Technological Use of CO2,
methanol synthesis catalyst on the hybrid process selectivity over Cr-Zn and Cu-
Chem. Rev. 114 (2014) 1709–1742, https://doi.org/10.1021/cr4002758.
Zn with SAPO-34, Ind. Eng. Chem. Res. 56 (2017) 13392–13401, https://doi.org/
[2] T. Herzog, World Greenhouse Gas Emissions in 2005, 2009. Accessed on january
10.1021/acs.iecr.7b02401.
2022, https://www.wri.org/data/world-greenhouse-gas-emissions-2005.
[27] M. Fujiwara, R. Kieffer, H. Ando, Y. Souma, Development of composite catalysts
[3] M. Mikkelsen, M. Jørgensen, F.C. Krebs, The Teraton Challenge . A Review of
made of Cu-Zn-Cr oxide/zeolite for the hydrogenation of carbon dioxide, Appl.
Fixation and Transformation of Carbon Dioxide, 2010, pp. 43–81, https://doi.
Catal. Gen. 121 (1995) 113–124, https://doi.org/10.1016/0926-860X(95)85014-
org/10.1039/b912904a.
7.
[4] M. Ginebra, C. Muñoz, R. Calvelo-pereira, M. Doussoulin, E. Zagal, Science of the
[28] X. Jiang, N. Koizumi, X. Guo, C. Song, Bimetallic Pd-Cu catalysts for selective CO2
Total Environment Biochar Impacts on Soil Chemical Properties , Greenhouse Gas
hydrogenation to methanol, Appl. Catal. B Environ. 170–171 (2015) 173–185,
Emissions and Forage Productivity : A Field Experiment, 2022, p. 806, https://
https://doi.org/10.1016/j.apcatb.2015.01.010.
doi.org/10.1016/j.scitotenv.2021.150465.
[29] N. Iwasa, H. Suzuki, M. Terashita, M. Arai, N. Takezawa, Methanol synthesis from
[5] C.D. Gross, E.W. Bork, C.N. Carlyle, S.X. Chang, Science of the total environment
CO2 under atmospheric pressure over supported Pd catalysts, Catal. Lett. 96
biochar and its manure-based feedstock have divergent effects on soil organic
(2004) 75–78, https://doi.org/10.1023/B:CATL.0000029533.41604.13.
carbon and greenhouse gas emissions in croplands, Sci. Total Environ. 806
[30] J.L. Snider, V. Streibel, M.A. Hubert, T.S. Choksi, E. Valle, D.C. Upham,
(2022), 151337, https://doi.org/10.1016/j.scitotenv.2021.151337.
J. Schumann, M.S. Duyar, A. Gallo, F. Abild-Pedersen, T.F. Jaramillo, Revealing
[6] I.C. Change, Mitigation of Climate Change, Contrib. Working Group III to Fifth
the synergy between oxide and alloy phases on the performance of bimetallic In-
Assessment Report of the Intergovermental Panel on Climate Change, Cambridge
Pd catalysts for CO2 hydrogenation to methanol, ACS Catal. 9 (2019) 3399–3412,
University Press, 2014, p. 1454.
https://doi.org/10.1021/acscatal.8b04848.
[7] J.B. Joshi, Chemical engineering research and design catalytic carbon dioxide
[31] P.S. Murthy, W. Liang, Y. Jiang, J. Huang, Cu-Based nanocatalysts for CO2
hydrogenation to methanol : a review of recent studies, Chem. Eng. Res. (2014)
hydrogenation to Methanol, Energy Fuel. 35 (2021) 8558–8584, https://doi.org/
2557–2567, https://doi.org/10.1016/j.cherd.2014.03.005.
10.1021/acs.energyfuels.1c00625.
[8] A. Durga Devi, S. Pushpavanam, N. Singh, J. Verma, M.P. Kaur, S.C. Roy,
[32] R. Bonetto, F. Crisanti, A. Sartorel, Carbon dioxide reduction mediated by iron
Enhanced methane yield by photoreduction of CO2 at moderate temperature and
catalysts: mechanism and tntermediates that guide selectivity, ACS Omega 5
pressure using Pt coated, graphene oxide wrapped TiO2 nanotubes, Results Eng
(2020) 21309–21319, https://doi.org/10.1021/acsomega.0c02786.
14 (2022), 100441, https://doi.org/10.1016/j.rineng.2022.100441.
[33] J. Wang, G. Zhang, J. Zhu, X. Zhang, F. Ding, A. Zhang, X. Guo, C. Song, CO2
[9] E. Bogel-łukasik, M. Nunes, C.I. Melo, A. Szczepan, Hydrogenation of carbon
hydrogenation to methanol over In2O3-based catalysts: from mechanism to
dioxide to methane by Ruthenium nanoparticles in ionic liquid, ChemSusChem
catalyst development, ACS Catal. 11 (2021) 1406–1423, https://doi.org/
(2016) 1081–1084, https://doi.org/10.1002/cssc.201600203.
10.1021/acscatal.0c03665.
[10] H. Morales, M.R. Goldwasser, M.J. Pérez-zurita, F. González-jiménez, C.U. De N,
[34] X. Jiang, X. Nie, X. Guo, C. Song, J.G. Chen, Recent advances in carbon dioxide
Hydrogenation of carbon oxides over Fe/Al2O3 catalysts, Appl. Catal., A : Gene
hydrogenation to methanol via heterogeneous catalysis, Chem. Rev. 120 (2020)
189 (1999) 87–97, https://doi.org/10.1016/S0926-860X(99)00262-8.
7984–8034, https://doi.org/10.1021/acs.chemrev.9b00723.
[11] H. Sakurai, A. Ueda, T. Kobayashi, M. Haruta, Low-temperature Water – Gas Shift
[35] K. Stangeland, H. Li, Z. Yu, CO2 hydrogenation to methanol: the
Reaction over Gold Deposited on TiO2, 1997, pp. 271–272, https://doi.org/
structure–activity relationships of different catalyst systems, Energy, Ecol.
10.1039/A606192C.
Environ. 5 (2020) 272–285, https://doi.org/10.1007/s40974-020-00156-4.
[12] P. Gao, S. Dang, S. Li, X. Bu, Z. Liu, M. Qiu, C. Yang, H. Wang, L. Zhong, Y. Han,
Q. Liu, W. Wei, Y. Sun, Direct production of lower olefins from CO2 conversion

16
N.J. Azhari et al. Results in Engineering 16 (2022) 100711

[36] G. Bozzano, F. Manenti, Efficient methanol synthesis: perspectives, technologies [60] S. Poto, D. Vico van Berkel, F. Gallucci, M.F.N. d’Angelo, Kinetic modelling of the
and optimization strategies, Prog. Energy Combust. Sci. 56 (2016) 71–105, methanol synthesis from CO2 and H2 over a CuO/CeO2/ZrO2 catalyst: the role of
https://doi.org/10.1016/j.pecs.2016.06.001. CO2 and CO hydrogenation, Chem. Eng. J. 435 (2022), 134946, https://doi.org/
[37] R.P. Ye, J. Ding, W. Gong, M.D. Argyle, Q. Zhong, Y. Wang, C.K. Russell, Z. Xu, A. 10.1016/j.cej.2022.134946.
G. Russell, Q. Li, M. Fan, Y.G. Yao, CO2 hydrogenation to high-value products via [61] N.D. Nielsen, A.D. Jensen, J.M. Christensen, The roles of CO and CO2 in high
heterogeneous catalysis, Nat. Commun. 10 (2019), https://doi.org/10.1038/ pressure methanol synthesis over Cu-based catalysts, J. Catal. 393 (2021)
s41467-019-13638-9. 324–334, https://doi.org/10.1016/j.jcat.2020.11.035.
[38] U.J. Etim, Y. Song, Z. Zhong, Improving the Cu/ZnO-based catalysts for carbon [62] J. Zhu, Y. Su, J. Chai, V. Muravev, N. Kosinov, E.J.M. Hensen, Mechanism and
dioxide hydrogenation to methanol, and the use of methanol as a renewable nature of active sites for methanol synthesis from CO/CO2 on Cu/CeO2, ACS
energy storage media, Front. Earth Sci. 8 (2020) 1–26, https://doi.org/10.3389/ Catal. 10 (2020) 11532–11544, https://doi.org/10.1021/acscatal.0c02909.
fenrg.2020.545431. [63] N. Pasupulety, H. Driss, Y.A. Alhamed, A.A. Alzahrani, M.A. Daous, L. Petrov,
[39] M.J. Da Silva, Synthesis of methanol from methane: challenges and advances on Studies on Au/Cu-Zn-Al catalyst for methanol synthesis from CO2, Appl. Catal.
the multi-step (syngas) and one-step routes (DMTM), Fuel Process, Technol. 145 Gen. 504 (2015) 308–318, https://doi.org/10.1016/j.apcata.2015.01.036.
(2016) 42–61, https://doi.org/10.1016/j.fuproc.2016.01.023. [64] F.C.F. Marcos, L. Lin, L.E. Betancourt, S.D. Senanayake, J.A. Rodriguez, J.
[40] B. Balopi, P. Agachi, Danha, Methanol synthesis chemistry and process M. Assaf, R. Giudici, E.M. Assaf, Insights into the methanol synthesis mechanism
engineering aspects - a review with consequence to Botswana chemical industries, via CO2 hydrogenation over Cu-ZnO-ZrO2 catalysts: effects of surfactant/Cu-Zn-Zr
Procedia Manuf. 35 (2019) 367–376, https://doi.org/10.1016/j. molar ratio, J. CO2 Util. 41 (2020), https://doi.org/10.1016/j.jcou.2020.101215.
promfg.2019.05.054. [65] R. Singh, K. Tripathi, K.K. Pant, Investigating the role of oxygen vacancies and
[41] J. Wu, Y. Huang, W. Ye, Y. Li, CO2 reduction: from the electrochemical to basic site density in tuning methanol selectivity over Cu/CeO2 catalyst during
photochemical approach, Adv. Sci. 4 (2017) 1–29, https://doi.org/10.1002/ CO2 hydrogenation, Fuel 303 (2021), 121289, https://doi.org/10.1016/j.
advs.201700194. fuel.2021.121289.
[42] Y. Yang, M.G. White, P. Liu, Theoretical study of methanol synthesis from CO2 [66] J.I. Di Cosimo, V.K. Díez, M. Xu, E. Iglesia, C.R. Apesteguía, Structure and surface
hydrogenation on metal-doped Cu(111) surfaces, J. Phys. Chem. C 116 (2012) and catalytic properties of Mg-Al basic oxides, J. Catal. 178 (1998) 499–510,
248–256, https://doi.org/10.1021/jp208448c. https://doi.org/10.1006/jcat.1998.2161.
[43] N. Park, M.J. Park, Y.J. Lee, K.S. Ha, K.W. Jun, Kinetic modeling of methanol [67] E.G. Choi, K.H. Song, S.R. An, K.Y. Lee, M.H. Youn, K.T. Park, S.K. Jeong, H.
synthesis over commercial catalysts based on three-site adsorption, Fuel Process. J. Kim, Cu/ZnO/AlOOH catalyst for methanol synthesis through CO2
Technol. 125 (2014) 139–147, https://doi.org/10.1016/j.fuproc.2014.03.041. hydrogenation, Kor. J. Chem. Eng. 35 (2018) 73–81, https://doi.org/10.1007/
[44] X. Cui, S.K. Kær, A comparative study on three reactor types for methanol s11814-017-0230-y.
synthesis from syngas and CO2, Chem. Eng. J. 393 (2020), 124632, https://doi. [68] P. Gao, F. Li, N. Zhao, F. Xiao, W. Wei, L. Zhong, Y. Sun, Influence of modifier
org/10.1016/j.cej.2020.124632. (Mn, La, Ce, Zr and Y) on the performance of Cu/Zn/Al catalysts via hydrotalcite-
[45] S. Zhang, Z. Wu, X. Liu, K. Hua, Z. Shao, B. Wei, C. Huang, H. Wang, Y. Sun, like precursors for CO2 hydrogenation to methanol, Appl. Catal. Gen. 468 (2013)
A Short review of recent advances in direct CO2 hydrogenation to alcohols, Top. 442–452, https://doi.org/10.1016/j.apcata.2013.09.026.
Catal. 64 (2021) 371–394, https://doi.org/10.1007/s11244-020-01405-w. [69] X. Han, M. Li, X. Chang, Z. Hao, J. Chen, Y. Pan, S. Kawi, X. Ma, Hollow
[46] J. Zhong, X. Yang, Z. Wu, B. Liang, Y. Huang, T. Zhang, State of the art and structured Cu@ZrO2 derived from Zr-MOF for selective hydrogenation of CO2 to
perspectives in heterogeneous catalysis of CO2 hydrogenation to methanol, Chem. methanol, J. Energy Chem. 71 (2022) 277–287, https://doi.org/10.1016/j.
Soc. Rev. 49 (2020) 1385–1413, https://doi.org/10.1039/c9cs00614a. jechem.2022.03.034.
[47] J. Niu, H. Liu, Y. Jin, B. Fan, W. Qi, J. Ran, Comprehensive review of Cu-based [70] B. Yang, C. Liu, A. Halder, E.C. Tyo, A.B.F. Martinson, S. Seifert, P. Zapol, L.
CO2 hydrogenation to CH3OH: insights from experimental work and theoretical A. Curtiss, S. Vajda, Copper cluster size effect in methanol synthesis from CO2,
analysis, Int. J. Hydrogen Energy (2022) 1–18, https://doi.org/10.1016/j. J. Phys. Chem. C 121 (2017) 10406–10412, https://doi.org/10.1021/acs.
ijhydene.2022.01.021. jpcc.7b01835.
[48] J. Yu, S. Liu, X. Mu, G. Yang, X. Luo, E. Lester, T. Wu, Cu-ZrO2 catalysts with [71] A. Karelovic, P. Ruiz, The role of copper particle size in low pressure methanol
highly dispersed Cu nanoclusters derived from ZrO2@ HKUST-1 composites for synthesis via CO2 hydrogenation over Cu/ZnO catalysts, Catal. Sci. Technol. 5
the enhanced CO2 hydrogenation to methanol, Chem. Eng. J. 419 (2021), (2015) 869–881, https://doi.org/10.1039/c4cy00848k.
129656, https://doi.org/10.1016/j.cej.2021.129656. [72] H. Wang, G. Zhang, G. Fan, L. Yang, F. Li, Fabrication of Zr-Ce Oxide solid
[49] F. Arena, G. Italiano, K. Barbera, S. Bordiga, G. Bonura, L. Spadaro, F. Frusteri, solution surrounded Cu-based catalyst assisted by a microliquid film reactor for
Solid-state interactions, adsorption sites and functionality of Cu-ZnO/ZrO2 efficient CO2 hydrogenation to produce methanol, Ind. Eng. Chem. Res. 60 (2021)
catalysts in the CO2 hydrogenation to CH3OH, Appl. Catal. Gen. 350 (2008) 16188–16200, https://doi.org/10.1021/acs.iecr.1c03117.
16–23, https://doi.org/10.1016/j.apcata.2008.07.028. [73] G. Zhang, M. Liu, G. Fan, L. Zheng, F. Li, Efficient role of nanosheet-like Pr2O3
[50] T. Witoon, J. Chalorngtham, P. Dumrongbunditkul, M. Chareonpanich, induced surface-interface synergetic structures over Cu-based catalysts for
J. Limtrakul, CO2 hydrogenation to methanol over Cu/ZrO2 catalysts: effects of enhanced methanol production from CO2 hydrogenation, ACS Appl. Mater.
zirconia phases, Chem. Eng. J. 293 (2016) 327–336, https://doi.org/10.1016/j. Interfaces 14 (2022) 2768–2781, https://doi.org/10.1021/acsami.1c20056.
cej.2016.02.069. [74] Z. Ou, J. Ran, J. Niu, C. Qin, W. He, L. Yang, A density functional theory study of
[51] M. Huš, D. Kopač, N.S. Štefančič, D.L. Jurković, V.D.B.C. Dasireddy, B. Likozar, CO2 hydrogenation to methanol over Pd/TiO2 catalyst: the role of interfacial site,
Unravelling the mechanisms of CO2 hydrogenation to methanol on Cu-based Int. J. Hydrogen Energy 45 (2020) 6328–6340, https://doi.org/10.1016/j.
catalysts using first-principles multiscale modelling and experiments, Catal. Sci. ijhydene.2019.12.099.
Technol. 7 (2017) 5900–5913, https://doi.org/10.1039/c7cy01659j. [75] A. García-Trenco, E.R. White, A. Regoutz, D.J. Payne, M.S.P. Shaffer, C.
[52] Y.F. Zhao, Y. Yang, C. Mims, C.H.F. Peden, J. Li, D. Mei, Insight into methanol K. Williams, Pd2Ga-based colloids as highly active catalysts for the hydrogenation
synthesis from CO2 hydrogenation on Cu(1 1 1): complex reaction network and of CO2 to methanol, ACS Catal. 7 (2017) 1186–1196, https://doi.org/10.1021/
the effects of H2O, J. Catal. 281 (2011) 199–211, https://doi.org/10.1016/j. acscatal.6b02928.
jcat.2011.04.012. [76] H. Bahruji, M. Bowker, G. Hutchings, N. Dimitratos, P. Wells, E. Gibson, W. Jones,
[53] M.D. Higham, M.D. Higham, M.G. Quesne, M.G. Quesne, C.R.A. Catlow, C.R. C. Brookes, D. Morgan, G. Lalev, Pd/ZnO catalysts for direct CO2 hydrogenation
A. Catlow, C.R.A. Catlow, Mechanism of CO2 conversion to methanol over Cu to methanol, J. Catal. 343 (2016) 133–146, https://doi.org/10.1016/j.
(110) and Cu(100) surfaces, Dalton Trans. 49 (2020) 8478–8497, https://doi. jcat.2016.03.017.
org/10.1039/d0dt00754d. [77] M. Zabilskiy, V.L. Sushkevich, M.A. Newton, F. Krumeich, M. Nachtegaal, J.
[54] S. Kattel, P. Liu, J.G. Chen, Tuning selectivity of CO2 hydrogenation reactions at A. van Bokhoven, Mechanistic study of carbon dioxide hydrogenation over Pd/
the metal/oxide Interface, J. Am. Chem. Soc. 139 (2017) 9739–9754, https://doi. ZnO-Based Catalysts: the role of Palladium–Zinc alloy in selective methanol
org/10.1021/jacs.7b05362. synthesis, Angew. Chem. Int. Ed. 60 (2021) 17053–17059, https://doi.org/
[55] E. Lam, J.J. Corral-Pérez, K. Larmier, G. Noh, P. Wolf, A. Comas-Vives, 10.1002/anie.202103087.
A. Urakawa, C. Copéret, CO2 Hydrogenation on Cu/Al2O3: role of the metal/ [78] E.J. Choi, Y.H. Lee, D.W. Lee, D.J. Moon, K.Y. Lee, Hydrogenation of CO2 to
support interface in driving activity and selectivity of a bifunctional catalyst, methanol over Pd–Cu/CeO2 catalysts, Mol. Catal. 434 (2017) 146–153, https://
Angew. Chem. Int. Ed. 58 (2019) 13989–13996, https://doi.org/10.1002/ doi.org/10.1016/j.mcat.2017.02.005.
anie.201908060. [79] T. Fujitani, M. Saito, Y. Kanai, T. Watanabe, J. Nakamura, T. Uchijima,
[56] J.S. Lee, K.H. Lee, S.Y. Lee, K.Y. Gul, A comparative study of methanol synthesis Development of an active Ga2O3 supported palladium catalyst for the synthesis of
from CO2/H2 and CO/H2 over a Cu/ZnO/Al2O3 Catalyst, J. Catal. 144 (1993) methanol from carbon dioxide and hydrogen, Applied Cataysis A: Gene 125
414–424, https://doi.org/10.1006/jcat.1993.1342. (1995) L199–L202, https://doi.org/10.1016/0926-860X(95)00049-6.
[57] K. Klier, V. Chatikavanij, R.G. Herman, G.W. Simmons, Catalytic synthesis of [80] L. Li, B. Zhang, E. Kunkes, K. Föttinger, M. Armbrüster, D.S. Su, W. Wei,
methanol from CO H2. IV. The effects of carbon dioxide, J. Catal. 74 (1982) R. Schlögl, M. Behrens, Ga-Pd/Ga2O3 Catalysts: the role of gallia polymorphs,
343–360, https://doi.org/10.1016/0021-9517(82)90040-9. intermetallic compounds, and pretreatment conditions on selectivity and stability
[58] O. Martin, J. Pérez-Ramírez, New (revisited) insights into the promotion of in different reactions, ChemCatChem 4 (2012) 1764–1775, https://doi.org/
methanol synthesis catalysts by CO2, Catal. Sci. Technol. 3 (2013) 3343–3352, 10.1002/cctc.201200268.
https://doi.org/10.1039/C3CY00573A. [81] R. Manrique, J. Rodríguez-Pereira, S.A. Rincón-Ortiz, J.J. Bravo-Suárez, V.
[59] O. Martin, C. Mondelli, A. Cervellino, D. Ferri, D. Curulla-Ferré, J. Pérez-Ramírez, G. Baldovino-Medrano, R. Jiménez, A. Karelovic, The nature of the active sites of
Operando synchrotron X-ray powder diffraction and modulated-excitation Pd-Ga catalysts in the hydrogenation of CO2 to methanol, Catal. Sci. Technol. 10
infrared spectroscopy elucidate the CO2 promotion on a commercial methanol (2020) 6644–6658, https://doi.org/10.1039/d0cy00956c.
synthesis Catalyst, Angew. Chem. Int. Ed. 55 (2016) 11031–11036, https://doi. [82] H. Lorenz, R. Thalinger, E.M. Köck, M. Kogler, L. Mayr, D. Schmidmair, T. Bielz,
org/10.1002/anie.201603204. K. Pfaller, B. Klötzer, S. Penner, Methanol steam reforming: CO2-selective Pd2Ga

17
N.J. Azhari et al. Results in Engineering 16 (2022) 100711

phases supported on α- and γ-Ga2O3, Appl. Catal. Gen. 453 (2013) 34–44, https:// [106] L. Pastor-p, M. Shah, Improving Fe/Al2O3 catalysts for the reverse water-gas shift
doi.org/10.1016/j.apcata.2012.11.010. reaction : on the effect of Cs as activity/selectivity promoter, Catalysts 8 (2018),
[83] A. Haghofer, K. Föttinger, F. Girgsdies, D. Teschner, A. Knop-Gericke, R. Schlögl, https://doi.org/10.3390/catal8120608.
G. Rupprechter, In situ study of the formation and stability of supported Pd2Ga [107] Q. Zhang, L. Pastor-pérez, W. Jin, S. Gu, T.R. Reina, Environmental
methanol steam reforming catalysts, J. Catal. 286 (2012) 13–21, https://doi.org/ Understanding the promoter e ff ect of Cu and Cs over highly e ff ective β - Mo 2 C
10.1016/j.jcat.2011.10.007. catalysts for the reverse water-gas shift reaction, Appl. Catal., B 244 (2019)
[84] A. Aguirre, S.E. Collins, Selective detection of reaction intermediates using 889–898, https://doi.org/10.1016/j.apcatb.2018.12.023.
concentration- modulation excitation DRIFT spectroscopy, Catal. Today 205 [108] M. Figueras, R.A. Gutiérrez, F. Viñes, P.J. Ramírez, J.A. Rodriguez, F. Illas,
(2013) 34–40, https://doi.org/10.1016/j.cattod.2012.08.020. Supported molybdenum carbide nanoparticles as an excellent catalyst for CO2
[85] S.E. Collins, J.J. Delgado, C. Mira, J.J. Calvino, S. Bernal, D.L. Chiavassa, M. hydrogenation, ACS Catal. 11 (2021) 9679–9687, https://doi.org/10.1021/
A. Baltanás, A.L. Bonivardi, The role of Pd-Ga bimetallic particles in the acscatal.1c01738.
bifunctional mechanism of selective methanol synthesis via CO2 hydrogenation [109] H. Zhou, Z. Chen, E. Kountoupi, A. Tsoukalou, P.M. Abdala, P. Florian,
on a Pd/Ga2O3 catalyst, J. Catal. 292 (2012) 90–98, https://doi.org/10.1016/j. A. Fedorov, C.R. Müller, Two-dimensional molybdenum carbide 2D-Mo2C as a
jcat.2012.05.005. superior catalyst for CO2 hydrogenation, Nat. Commun. 12 (2021) 1–10, https://
[86] S.E. Collins, M.A. Baltanás, J.J. Delgado, A. Borgna, A.L. Bonivardi, CO2 doi.org/10.1038/s41467-021-25784-0.
hydrogenation to methanol on Ga2O3-Pd/SiO2 catalysts: dual oxide-metal sites or [110] O. Martin, A.J. Martín, C. Mondelli, S. Mitchell, T.F. Segawa, R. Hauert,
(bi)metallic surface sites? Catal. Today 381 (2021) 154–162, https://doi.org/ C. Drouilly, D. Curulla-Ferré, J. Pérez-Ramírez, Indium oxide as a superior
10.1016/j.cattod.2020.07.048. catalyst for methanol synthesis by CO2 hydrogenation, Angew. Chem. Int. Ed. 55
[87] F. Liao, X.P. Wu, J. Zheng, M.M.J. Li, A. Kroner, Z. Zeng, X. Hong, Y. Yuan, X. (2016) 6261–6265, https://doi.org/10.1002/anie.201600943.
Q. Gong, S.C.E. Tsang, A promising low pressure methanol synthesis route from [111] J. Ye, C. Liu, D. Mei, Q. Ge, Active oxygen vacancy site for methanol synthesis
CO2 hydrogenation over Pd@Zn core-shell catalysts, Green Chem. 19 (2017) from CO2 hydrogenation on In2O3(110): a DFT study, ACS Catal. 3 (2013)
270–280, https://doi.org/10.1039/c6gc02366e. 1296–1306, https://doi.org/10.1021/cs400132a.
[88] Y. Hartadi, D. Widmann, R.J. Behm, CO2 hydrogenation to methanol on [112] J. Ye, C. Liu, Q. Ge, DFT study of CO 2 adsorption and hydrogenation on the in 2
supported Au catalysts under moderate reaction conditions: support and particle O 3 surface, J. Phys. Chem. 116 (2012), https://doi.org/10.1021/jp3004773.
size effects, ChemSusChem 8 (2015) 456–465, https://doi.org/10.1002/ [113] T.P. Araújo, A. Shah, C. Mondelli, J.A. Stewart, D. Curulla Ferré, J. Pérez-
cssc.201402645. Ramírez, Impact of hybrid CO2-CO feeds on methanol synthesis over In2O3-based
[89] Y. Hartadi, D. Widmann, R.J. Behm, Methanol synthesis: via CO2 hydrogenation catalysts, Appl. Catal. B Environ. 285 (2021), https://doi.org/10.1016/j.
over a Au/ZnO catalyst: an isotope labelling study on the role of CO in the apcatb.2021.119878.
reaction process, Phys. Chem. Chem. Phys. 18 (2016) 10781–10791, https://doi. [114] S. Dang, B. Qin, Y. Yang, H. Wang, J. Cai, Y. Han, S. Li, P. Gao, Y. Sun, Rationally
org/10.1039/c5cp06888f. designed indium oxide catalysts for CO2 hydrogenation to methanol with high
[90] A. Rezvani, A.M. Abdel-Mageed, T. Ishida, T. Murayama, M. Parlinska-Wojtan, R. activity and selectivity, Sci. Adv. 6 (2020) 1–12, https://doi.org/10.1126/sciadv.
J. Behm, CO2 reduction to methanol on Au/CeO2 catalysts: mechanistic insights aaz2060.
from activation/deactivation and SSITKA Measurements, ACS Catal. 10 (2020) [115] L.S. Byskov, J.K. Nørskov, B.S. Clausen, H. Topsøe, DFT calculations of
3580–3594, https://doi.org/10.1021/acscatal.9b04655. unpromoted and promoted MoS2-based hydrodesulfurization catalysts, J. Catal.
[91] N. Rui, F. Zhang, K. Sun, Z. Liu, W. Xu, E. Stavitski, S.D. Senanayake, J. 187 (1999) 109–122, https://doi.org/10.1006/jcat.1999.2598.
A. Rodriguez, C.J. Liu, Hydrogenation of CO2 to methanol on a Auδ+-In2O3- x [116] B. Hinnemann, P.G. Moses, J. Bonde, K.P. Jørgensen, J.H. Nielsen, S. Horch,
catalyst, ACS Catal. 10 (2020) 11307–11317, https://doi.org/10.1021/ I. Chorkendorff, J.K. Nørskov, Biomimetic hydrogen evolution : MoS2
acscatal.0c02120. nanoparticles as catalyst for hydrogen evolution, J. Am. Chem. Soc. 127 (2005)
[92] T.V. Sagar, J. Zavašnik, M. Finšgar, N.N. Tušar, A. Pintar, Evaluation of Au/ZrO2 5308–5309, https://doi.org/10.1021/ja0504690.
catalysts prepared via postsynthesis methods in CO2 hydrogenation to methanol, [117] T.F. Jaramillo, K.P. Jørgensen, J. Bonde, J.H. Nielsen, S. Horch, I. Chorkendoff,
Catalysts 12 (2022) 1–25, https://doi.org/10.3390/catal12020218. Identification of active edge sites for electrochemical H2 evolution from MoS2
[93] S.B.T. Tran, H. Choi, S. Oh, J.Y. Park, Influence of support acidity of Pt/Nb2O5 nanocatalysts identification of active edge sites for electrochemical H2 evolution
catalysts on selectivity of CO2 hydrogenation, Catal. Lett. 149 (2019) 2823–2835, from MoS2 nanocatalysts, Science 317 (2007) 100–102, https://doi.org/
https://doi.org/10.1007/s10562-019-02822-7. 10.1126/science.1141483.
[94] Y. Chen, H. Li, W. Zhao, W. Zhang, J. Li, W. Li, X. Zheng, W. Yan, W. Zhang, [118] R. [117] R. Prins, V.H.J. De Beer, G.A. Somorjai, Structure and function of the
J. Zhu, R. Si, J. Zeng, Optimizing reaction paths for methanol synthesis from CO2 catalyst and the promoter in Co—Mo hydrodesulfurization catalysts, Catal. Rev.
hydrogenation via metal-ligand cooperativity, Nat. Commun. 10 (2019) 1–8, Eng. 31 (1989) 1–41, https://doi.org/10.1080/01614948909351347.
https://doi.org/10.1038/s41467-019-09918-z. [119] C. Tsai, F. Abild-pedersen, J.K. Nørskov, Tuning the MoS2 edge-site activity for
[95] K. Sun, N. Rui, Z. Zhang, Z. Sun, Q. Ge, C.J. Liu, A highly active Pt/In2O3 catalyst hydrogen evolution via support interactions, Nano Lett. 13 (2014) 1381–1387,
for CO2 hydrogenation to methanol with enhanced stability, Green Chem. 22 https://doi.org/10.1021/nl404444k.
(2020) 5059–5066, https://doi.org/10.1039/d0gc01597k. [120] J. Hu, L. Yu, J. Deng, Y. Wang, K. Cheng, C. Ma, Q. Zhang, W. Wen, S. Yu, Y. Pan,
[96] Z. Han, C. Tang, J. Wang, L. Li, C. Li, Atomically dispersed Ptn+ species as highly J. Yang, H. Ma, F. Qi, Y. Wang, Y. Zheng, M. Chen, R. Huang, S. Zhang, Z. Zhao,
active sites in Pt/In2O3 catalysts for methanol synthesis from CO2 hydrogenation, J. Mao, X. Meng, Q. Ji, G. Hou, X. Han, X. Bao, Y. Wang, D. Deng, Sulfur vacancy-
J. Catal. 394 (2021) 236–244, https://doi.org/10.1016/j.jcat.2020.06.018. rich MoS2 as a catalyst for the hydrogenation of CO2 to methanol, Nat. Catal. 4
[97] J.A. Rodriguez, P. Liu, D.J. Stacchiola, S.D. Senanayake, M.G. White, J.G. Chen, (2021) 242–250, https://doi.org/10.1038/s41929-021-00584-3.
Hydrogenation of CO2 to methanol: importance of metal-oxide and metal-carbide [121] T. Len, M. Bahri, O. Ersen, Y. Lefkir, L. Cardenas, I.J. Villar-garcia, V.P. Dieste,
interfaces in the activation of CO2, ACS Catal. 5 (2015) 6696–6706, https://doi. J. Llorca, N. Perret, R. Checa, E. Puzenat, P. Afanasiev, F. Morfin, L. Piccolo,
org/10.1021/acscatal.5b01755. Supporting information ultradispersed Mo/TiO2 catalysts for CO2 hydrogenation
[98] M.D. Porosoff, B. Yan, J.G. Chen, Environmental Science CO , methanol and to methanol, Green Chem. 23 (2021) 7259–7268, https://doi.org/10.1039/
hydrocarbons : challenges and opportunities, Energy Environ. Sci. 9 (2015) D1GC01761F.
62–73, https://doi.org/10.1039/C5EE02657A. [122] O.A. Odunlami, D.A. Vershima, T.E. Oladimeji, S. Nkongho, S.K. Ogunlade, B.
[99] K. Li, J.G. Chen, CO2 hydrogenation to methanol over ZrO2-containing catalysts: S. Fakinle, Advanced techniques for the capturing and separation of CO2 – a
insights into ZrO2 induced synergy, ACS Catal. 9 (2019) 7840–7861, https://doi. review, Results Eng 15 (2022), 100512, https://doi.org/10.1016/j.
org/10.1021/acscatal.9b01943. rineng.2022.100512.
[100] S. Dang, H. Yang, P. Gao, H. Wang, X. Li, W. Wei, Y. Sun, A review of research [123] W. Yue, Y. Li, W. Wei, J. Jiang, J. Caro, A. Huang, Highly selective CO2
progress on heterogeneous catalysts for methanol synthesis from carbon dioxide conversion to methanol in a bifunctional zeolite catalytic membrane reactor,
hydrogenation, Catal. Today 330 (2019) 61–75, https://doi.org/10.1016/j. Angew. Chem. Int. Ed. 60 (2021) 18289–18294, https://doi.org/10.1002/
cattod.2018.04.021. anie.202106277.
[101] A.R. Richard, M. Fan, Low-pressure hydrogenation of CO2 to CH3OH using Ni-In- [124] M.A. Adnan, G. Kibria, Comparative techno-economic and life-cycle assessment of
Al/SiO2 catalyst synthesized via a phyllosilicate precursor, ACS Catal. 7 (2017) power-to-methanol synthesis pathways, Appl. Energy 278 (2020), 115614,
5679–5692, https://doi.org/10.1021/acscatal.7b00848. https://doi.org/10.1016/j.apenergy.2020.115614.
[102] Y. Ren, C. Xin, Z. Hao, H. Sun, S.L. Bernasek, W. Chen, G.Q. Xu, Probing the
reaction mechanism in CO2 hydrogenation on bimetallic Ni/Cu(100) with near-
ambient pressure X-ray Photoelectron spectroscopy, ACS Appl. Mater. Interfaces
Further reading
12 (2020) 2548–2554, https://doi.org/10.1021/acsami.9b19523.
[103] P. Chen, G. Zhao, X.R. Shi, J. Zhu, J. Ding, Y. Lu, Nano-Intermetallic InNi3C0.5 [125] M.K. Koh, M.M. Zain, A.R. Mohamed, Exploring transition metal (Cr, Mn, Fe, Co,
compound discovered as a superior catalyst for CO2 reutilization, Science 17 Ni) promoted copper-catalyst for carbon dioxide hydrogenation to methanol, AIP
(2019) 315–324, https://doi.org/10.1016/j.isci.2019.07.006. Conf. Proc. 2124 (2019), https://doi.org/10.1063/1.5117066.
[104] A. Dongil, Q. Zhang, L. Pérez, T. Ramirez Reina, A. Ruiz, I. Rodriguez-Ramos,
Effect of Cu and Cs in the β-Mo2C system for CO2 hydrogenation to methanol,
Catalysts 10 (2020) 1213, https://doi.org/10.3390/catal10101213.
[105] F. Baibars, E. Le Sache, S. Gu, T.R. Reina, CO2 Valorisation via Reverse Water-Gas
Shift Reaction Using Advanced Cs Doped Fe-Cu/Al2O3 Catalysts, vol. 21, 2017,
pp. 423–428, https://doi.org/10.1016/j.jcou.2017.08.009.

18
N.J. Azhari et al. Results in Engineering 16 (2022) 100711

[126] L. Jia, J. Gao, W. Fang, Q. Li, Carbon dioxide hydrogenation to methanol over the [128] W.J. Shen, M. Okumura, Y. Matsumura, M. Haruta, The influence of the support
pre-reduced LaCr0.5Cu0.5O3 catalyst, Catal. Commun. 10 (2009) 2000–2003, on the activity and selectivity of Pd in CO hydrogenation, Appl. Catal. Gen. 213
https://doi.org/10.1016/j.catcom.2009.07.017. (2001) 225–232, https://doi.org/10.1016/S0926-860X(01)00465-3.
[127] J.F. Edwards, G.L. Schrader, Methanol, formaldehyde, and formic acid adsorption
on methanol synthesis catalysts, J. Phys. Chem. 89 (1985) 782–788, https://doi.
org/10.1021/j100251a015.

19

You might also like