Cherevotan 2021

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Letter

http://pubs.acs.org/journal/aelccp

Operando Generated Ordered Heterogeneous


Catalyst for the Selective Conversion of CO2 to
Methanol
Arjun Cherevotan, Jithu Raj,$ Lakshay Dheer,$ Soumyabrata Roy, Shreya Sarkar, Risov Das,
Chathakudath P. Vinod, Shaojun Xu, Peter Wells, Umesh V. Waghmare, and Sebastian C. Peter*
Cite This: ACS Energy Lett. 2021, 6, 509−516 Read Online
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ACCESS *
Downloaded via UNIV OF CONNECTICUT on May 15, 2021 at 15:17:52 (UTC).

Metrics & More Article Recommendations sı Supporting Information

ABSTRACT: The discovery of new materials for efficient trans-


formation of carbon dioxide (CO2) into desired fuel can revolutionize
large-scale renewable energy storage and mitigate environmental
damage due to carbon emissions. In this work, we discovered an
operando generated stable Ni−In kinetic phase that selectively
converts CO2 to methanol (CTM) at low pressure compared to the
state-of-the-art materials. The catalytic nature of a well-known
methanation catalyst, nickel, has been tuned with the introduction
of inactive indium, which enhances the CTM process. The remarkable
change in the mechanistic pathways toward methanol production has
been mapped by operando diffuse reflectance infrared Fourier
transform spectroscopy analysis, corroborated by first-principles calculations. The ordered arrangement and pronounced
electronegativity difference between metals are attributed to the complete shift in mechanism. The approach and findings of
this work provide a unique advance toward the next-generation catalyst discovery for going beyond the state-of-the-art in CO2
reduction technologies.

O ne of the many potential strategies envisioned to


curtail the alarming CO2 levels in the atmosphere is
to use scalable thermocatalytic hydrogenation of
captured CO2 to produce energy dense fuels important in the
material value chain.1−3 This strategy is a “two birds, one stone
have resulted in greater mechanistic insight into the
process.20−22
A relatively new area in CTM catalyst engineering has seen
the advent of transition-metal-based bimetallic or intermetallic
catalysts (IMCs).23,24 Even in the state-of-the-art CZA system,
approach”4 which can potentially shift the linear trend of interfacial Cu−Zn alloy is strongly speculated to be the
carbon utilization to a closed sustainable loop, thus decreasing efficient catalytic site for CTM under the reaction
CO2 emission, mitigating climate change, and reducing the conditions.25−27 Given the role bimetallic alloys have played
dependence on fossil fuels.5,6 On the other hand, methanol in various catalytic processes, such as Cu−Zn in CTM, more
(MeOH), among other catalytic products, has been projected recent studies have explored the possibilities of their ordered
as a future transportation fuel because of its high octane IM variants.28−30 They differ from their bimetallic/alloy
number.7 It can be directly used as additive to gasoline in counterparts having unique long-range crystalline order and
internal combustion engines or its derivatives such as dimethyl exhibit drastic changes from their parent elements in terms of
ether (DME) or oxymethylene ethers.8−10 Also, it is an both electronic and geometric properties.31−33 This inimitable
important starting material for the production of several feature significantly improves the CTM activity of ordered
multicarbon products, including alcohols, ethers, esters, olefins,
IMCs over bimetallic systems that can potentially surpass the
and alkanes.11,12 The catalytic conversion of CO2 to methanol
(CTM) via a thermochemical route of hydrogenation has been
investigated since the early 1990s.13 CuO/ZnO/Al2O3 (CZA) Received: December 17, 2020
is the state-of-the-art catalyst being used for CTM in high Accepted: January 11, 2021
pressure−temperature regimes.14 Since then, many advances in Published: January 14, 2021
catalyst design have been made to facilitate the hydrogenation
of CTM.15−19 Investigations into improved efficiency via
catalyst design in nanodimension and support-based effects

© 2021 American Chemical Society https://dx.doi.org/10.1021/acsenergylett.0c02614


509 ACS Energy Lett. 2021, 6, 509−516
ACS Energy Letters http://pubs.acs.org/journal/aelccp Letter

CZA benchmarks.34,35 One of the major advantages of


transforming bimetallic systems to IMCs is the generation of
the ordered arrangement of the atoms, which indirectly isolates
the active metallic site within the chemical atmosphere made
up of nonactive elements.36,37 In addition to the ordered
arrangement and site isolation, the performance of the catalysts
can be improved further with appropriate selection of support
materials.38−41 The carbon neutrality and the key techno-
economic bottleneck of the CTM process is associated with
the minimization of hydrogen consumption and energy
requirements of high-pressure conditions. Both the issues
have been addressed here by significantly enhancing the
MeOH selectivity at lower pressures.
In this work, we have discovered a novel non-noble metal-
based IMC comprising Ni and In atoms, Ni3In, through in situ
stabilization of kinetic Ni−In phase. This nanoscale-ordered Figure 1. Schematic illustration of Ni−In intermetallic on SBA-15
catalyst dispersed on high surface mesoporous silica SBA-15 by by incipient wetness impregnation.
a facile incipient wetness impregnation exhibits high MeOH
yields in CTM under lower pressures. As a support, SBA-15 is inactive. The experiments at higher pressures for Ni/SBA-15
popular owing to its high surface area (>600 m2/g) and exhibit a notable 12% decline in the conversion, but the
ordered hexagonal cylindrical mesoporous channels well selectivity toward methane remains unchanged at ∼99.5%.
interconnected via micropores.42 These properties help in This confirms that the selectivity of Ni catalyst toward
the uniform distribution of IMC nanoparticles (NPs) during methane is an inherent catalytic property of Ni and
incipient wetness impregnation process. The IMC-NPs nestle independent of operating pressure and WHSVs.
inside well-interconnected channels enhancing the diffusion of On the basis of the optimized reactions conditions for Ni/
reactants and products. Additionally, the SBA-15 surface SBA-15, the catalytic activity of Ni7In3/SBA-15 toward CO2
stabilizes the carbonyl intermediates together with enabling reduction has been tested (Figures S7 and S8). At 573 K, 50
ideal IM distribution.43 Nickel is a well-known and highly bar, and 4.7 h−1 WHSV, Ni7In3/SBA-15 converted ∼17% of
selective CO2 methanation catalyst at ambient pressures,44 and CO2 with product mixture ∼86% MeOH and 14% CO (Figure
also different ratios of Ni to In in a phyllosilicate-based 2a). The dramatic turnaround of the product selectivity from
composite catalyst is reported to enhance CO2 to methanol 100% methane on Ni to 86% methanol (∼1% methane) over
conversion at lower pressure.45 Further, the surface oxygen- IM Ni7In3 can be attributed to the stark structural changes
vacancy-mediated CO2 activation and selective hydrogenation upon alloying Ni with In, which triggers a different mechanistic
to methanol makes In2O3 an attractive catalyst.46,47 Strong pathway in the overall CO2 reduction process. In addition to
metal support interaction between d-block metals like Ni or Au conversion and selectivity, the other important direction in
with In2O3 in a reducing environment enhances the O-vacancy, CTM research aims to reduce the operating pressure to make
hiking the methanol formation.48,49 Recently, the In−Pd and the overall process economically viable. Toward this, Ni7In3/
In−Ni bimetallic systems on silica, reported with enhanced SBA-15 was further tested for CO2 reduction at different
methanol yield, the active phase is generated from synergy pressure ranges, starting from ambient to 50 bar, at a constant
between In−Pd or In−Ni with indium oxide phase.50 Here we temperature of 573 K and 4.7 h−1 WHSV (Figure 2b). As
aimed to understand and sketch the catalytic evolution upon shown, the conversion increases from 0% at ambient pressure
the introduction of In into the active Ni structure. These to more than 16% at 20 bar. A further increase in pressure from
controlled synthesis studies, repeated over several batches, 20 to 50 bar yields less than 2% increment in the CO2
revealed that all high-temperature reaction routes in the Ni−In conversion. A similar trend was obtained for the methanol
system lead majorly to the thermodynamically stable Ni7In3 selectivity, which increases from 0% at ambient pressure to
phase irrespective of the compositions targeted (see Figures 1, 75% methanol at 20 bar. A relatively smaller increment (13%)
S1, and S2 and section 1.2 in the Supporting Information). in the methanol selectivity was achieved between 20 and 50
The TEM, STEM, SEM, and N2 adsorption−desorption bar (Figures S7 and S8 and Table S1). CO was obtained as the
analyses confirm the uniform distribution of Ni7In3 on SBA-15 major product at low pressure, but methane was never found
(Figures S2−S4 and section 1.3 in the Supporting to be more than 1% at any of the operating conditions.
Information). The catalytic performance of Ni/SBA-15 and A sharp change in both conversion and selectivity occurring
Ni7In3/SBA-15 toward CO2 reduction was evaluated in a in the 10−20 bar range can be correlated to a sudden change
continuous flow fixed bed reactor. The products are analyzed in the chemical nature of Ni7In3/SBA-15 during the catalysis
by gas chromatography (Figure S5). Consistent with the earlier test. PXRD studies on the spent Ni7In3/SBA-15 catalysts
works,51,52 Ni/SBA-15 converts CO2 to methane with almost tested at each pressure were deployed to understand this
100% selectivity (Figure S6). Under various pressure regimes, chemical change and to assess the stability of the catalysts
CO2:H2 ratios, and varying WHSV, Ni/SBA-15 favored the (Figure 2c). The studies revealed an interesting phase
conversion of CO2 into methane (Figure S6a−c). transformation of the IM phase from the most unsymmetrical
Furthermore, at a given pressure, better conversion was triclinic Ni7In3 (SG: P1̅) (Figure S9) to the most symmetrical
obtained for a 1:4 ratio of CO2 to H2 (4.9 WHSV) without cubic Ni3In (SG: Pm3̅m) (Figure S10). The Ni3In phase
significantly affecting the methane selectivity. Bare SBA-15 progressively evolved from Ni7In3 with an increase of pressure
(without metal loading) under similar conditions did not favor of the reaction from ambient pressure to 20 bar at 573 K
the CO2 conversion (Table S1), implying that bare support is (Figure 2c). Above 20 bar, the phase transformation of Ni7In3
510 https://dx.doi.org/10.1021/acsenergylett.0c02614
ACS Energy Lett. 2021, 6, 509−516
ACS Energy Letters http://pubs.acs.org/journal/aelccp Letter

Figure 2. Catalytic performance of Ni7In3/SBA-15. (a) CO2 conversion and selectivity for 15 h of time on stream at reaction conditions of
4.7 h−1 (1:3), 50 bar, and 573 K. (b) CO2 conversion and methanol selectivity at different pressures (5−50 bar) with 4.7 h−1 and 573 K. (c)
XRD patterns of spent catalyst after reaction at 5, 10, 15, 20, 40, and 50 bar compared with simulated patterns of Ni (SG:Fm3̅m), Ni3In
(SG:Pm3̅m) and Ni7In3 (SG:P1̅).

to Ni3In was observed along with a negligible amount of In2O3/SBA-15 spent catalyst also confirmed the formation of
metallic Ni formation. The phase transformation is perfectly Ni3In (SG: Pm3̅m) with minor peaks of elemental Ni (SG:
aligned with the enhancement in CO2 conversion and Fm3̅m) (Figure S12d).
methanol selectivity, which ascertains that the operando These controlled studies clearly confirm that the CTM
generated cubic Ni3In is the real active phase for CTM. activity derives from the Ni3In phase and is not due to the
From a fundamental crystallographic perspective, the structural enhancement of O-vacancies on In2O3 by Ni metal. We
transition of Ni7In3 to Ni3In is complex, although earlier conclude that the NiO-In2O3 heterostructure reduces to the
reports have proposed Ni deficient Ni7In3 as the superstructure corresponding Ni and In metals, under the stream of reactant
of Ni3In based on the neutron diffraction and XRD studies.53 hydrogen at high temperature and pressure, followed by their
We speculate that Ni7In3 transformed into a perfectly ordered diffusion to form the Ni3In phase (Figure S13). Furthermore,
and kinetically stable Ni3In structure under the high temper- the spent catalyst after maximum applied pressure (50 bar)
ature and pressure conditions. This brings up a couple of reaction (denoted as Ni3In) when retested at various pressures
critical questions about the mechanistic pathway of metastable (0−20 bar), exhibited CO2 conversion and MeOH selectivity
Ni3In formation: (a) Is it a direct transformation from Ni7In3 similar to that of the first cycle (Figure S14). This proves that
to Ni3In or (b) the does the pathway entail the decomposition the metastable phase of Ni3In formed in the operando
of Ni7In3 to the constituent Ni and In metals followed by their condition is catalytically stable with good CO2 conversion
diffusion into the ordered Ni3In phase under the reaction and excellent MeOH selectivity at lower pressures.
conditions. To confirm this, an alternate synthetic route has The formation of intermetallic Ni7In3 and the operando
been explored. The as synthesized Ni7In3/SBA-15 was oxidized transformation to Ni3In has been proved by XPS. The slight
by calcining in open atmosphere at 773 K for 5 h, which shift of Ni0 2p3/2 toward lower binding energy (851.6 and
resulted in the formation NiO-In2O3/SBA-15 (Figure S11). 852.1 eV for Ni7In3 and Ni3In) compared to the metallic Ni
CO2 hydrogenation on NiO-In2O3/SBA-15 has been per- (852.6 eV) can be attributed to the difference in bonding due
formed at 20 and 50 bar (4.7 h−1 and 573 K) by avoiding the to the formation of IMC (Figure 3a,b).54 This also signifies
catalyst activation step to eliminate any IMC formation. partial electron transfer from In to Ni which confirms the
Although the conversion of CO2 by this catalyst is reduced by formation of relatively electron-rich Ni sites upon IM
7% compared to Ni7In3/SBA-15, the methanol selectivity was formation.54 Additionally, the broad peaks at 855.7, 861.3,
found to be similar to the reactions run over without oxidizing and 873.56 eV identified as Ni2+ 2p3/2, Ni2+ 2p3/2 (satellite),
the catalyst (Figure S12a−c). The O-vacancies of In2O3 have and Ni2+ 2p1/2, respectively, can be assigned to Ni−O because
been reported previously to favor CTM conversion,4,46,47 and of surface oxidation. The peak area ratio of Ni2+ 2p3/2 to Ni0
also recent studies discuss the enhancement of In2O3 O- 2p3/2 of as-synthesized Ni7In3 is two times higher than that of
vacancies by metals (Ni)48 and bimetallic systems (In−Pd and Ni3In (Table S3), which implies that the as-synthesized
In−Ni)50 dispersed on them, improving their CTM activity. catalyst having more surface oxidation gets reduced during
Thus, one can possibly attribute the activity of the catalyst to catalyst activation and then by hydrogen stream in operando
the presence of In2O3. But interestingly, the PXRD of the NiO- conditions. This is well-supported by the O 1s XPS spectra of
511 https://dx.doi.org/10.1021/acsenergylett.0c02614
ACS Energy Lett. 2021, 6, 509−516
ACS Energy Letters http://pubs.acs.org/journal/aelccp Letter

Ni7In3 and Ni3In (Figure 3c,d). The peaks at 531.34 and 528.6
eV correspond to surface hydroxyl and lattice O2− of NiO,
respectively.55 The peak at 528.6 eV gets reduced significantly
in Ni3In as compared to Ni7In3, indicating the reduction in
surface NiO in operando conditions. The charge transfer from
In to Ni upon IM formation is further confirmed by XANES
(Figure 3e,g) with a prominent shift of Ni K absorption edge
by 2 and 3.5 eV for Ni7In3 and Ni3In, respectively.
The white line intensity, which is a characteristic indicator of
charge transfer,56 also drops upon the transformation of Ni7In3
to Ni3In (Figure 3g). Experimental and theoretical studies
were employed to recognize the difference in the CO2
reduction mechanistic pathways between Ni/SBA-15 and
Ni3In/SBA-15. Operando diffuse reflectance infrared Fourier
transform spectroscopy (DRIFTS) has been used to map the
intermediates formed during the CO2 reduction process
(Figures 4 and S15).
Three key intermediates were identified when Ni3In was
used as the CTM catalyst: free COstr (2178 and 2111 cm−1),
linear Ni-COstr (2028 cm−1), bridged Ni3−COstr (1930 cm−1),
and conjugated C−O/C=Ostr of formate (1588 cm−1) (Figures
4a,b and S15).57,58 These peaks started to appear around 9 min
(∼450 K) of temperature ramp (15 K min−1 from 303 K),
which corresponds to the starting temperature for CO2
hydrogenation. The peaks intensify and saturate within an
Figure 3. (a) Ni 2p3/2 XPS spectrum of Ni7In3/SBA-15. (b) Ni hour of reaction at 573 K (Figure 4a,b). In contrast, over Ni,
2p3/2 XPS spectrum of Ni3In/SBA-15. (c) O 1s XPS spectrum of bands corresponding to formate, bridge Ni-CO, and linear Ni-
Ni7In3/SBA-15. (d) O 1s XPS spectrum of Ni3In/SBA-15. (e) XAS
spectra of Ni K-edge of Ni/SBA-15, Ni7In3/SBA-15, and Ni3In/
CO originate at the onset temperature. On further increase of
SBA-15. (f) Enlarged absorption edge region of Ni K-edge of temperature, the linear Ni-COstr (2030 cm−1) band subsides to
XANES spectra and (g) white line intensity region in XANES noise after 12 min of ramp along with a slight intensification of
spectra. bands at 2178 and 2111 cm−1 (free CO) (Figure 4c).
Furthermore, during 1 h on stream there is no enhancement of

Figure 4. Operando DRIFTS measurements. (a) Successive spectra of CO stretching regions over Ni3In/SBA-15 plotted with time of
temperature ramp from 303 to 573 K. (b) Successive spectra of CO stretching regions over Ni3In/SBA-15 for an hour on stream. (c)
Successive spectra of CO stretching regions over Ni/SBA-15 plotted with time of temperature ramp from 303 to 573 K. The ramp rate from
303 to 573 K is 15 K/min. (d) Successive spectra of CH stretching regions over Ni/SBA-15 for an hour on stream.

512 https://dx.doi.org/10.1021/acsenergylett.0c02614
ACS Energy Lett. 2021, 6, 509−516
ACS Energy Letters http://pubs.acs.org/journal/aelccp Letter

Figure 5. Relative activation energy diagrams of CO2 to methanol conversion on Ni3In occurring via (a) cis-COOH and (b) bi-HCOO
pathway. Black and red bars denote the intermediates and transition states along the CO2 hydrogenation pathways. TS3 (transition state 3)
is the kinetic bottleneck along both pathways, while the desorption of CH3OH is the thermodynamically most energy demanding step
(denoted by blue double headed arrows). All values are in electronvolts. (c) Schematic mechanism for the CO2-to-methane formation over
Ni/SBA-15 and CO2 to methanol formation over Ni3In/SBA-15 catalysts.

the peak intensity at 2030 cm−1; instead, other peaks get To conclusively determine the most feasible pathway, we
intensified and saturated (Figure S16). A new band estimated the energy of second elementary step in cis-COOH
corresponding to C−Hstr of methane starts to appear at 3015 and formate pathways, i.e.
cm−1 after 15 min of ramp (Figure S17) and gets intensified *CO2 + 3H 2 → *bi‐HCOO + H + 2H 2
further upon time on stream (Figure 4d).59 This baseline shift
in Figure 4c,d is expected because of the temperature effect and
causing thermal expansion and particle aggregation in the *CO2 + 3H 2 → *trans‐COOH + H + 2H 2
sample. The appearance of bands corresponding to C−Hstr,
free COstr, and subsidized Ni-COstr bands confirms that linear Formation of bi-HCOO from CO2 adsorbed on the (111)
Ni bound CO gets reduced to methane at temperature above surface of Ni3In is quite spontaneous (ΔE = −0.53 eV) and
500 K or the Ni-CO bond gets dissociated to free COstr. The more exothermic than that of its competing intermediate,
absence of the C−Hstr characteristic peak proves that Ni3In trans-COOH, occurring in the cis-COOH pathway (ΔE =
does not favor methane formation (Figure S18). In −0.32 eV). Also, both intermediates show a bidentate
comparison to the pure metal, the reduction ability of Ni has attachment to the (111) surface; however, the chemical
been reduced in the case of Ni3In. The decrease of Ni bonds formed by the adsorbates with Ni3In are distinct: Ni−O
reducibility in Ni3In may be due to the modification of and In−O in the case of bi-HCOO; Ni−C and Ni−O bonds in
hydrogen spilling power of Ni, which was optimized to the case of trans-COOH. The fact that In does not bind with
perform 6e− reduction instead of 8e− to yield methane upon trans-COOH suggests that Ni3In will prefer bi-HCOO.
the IMC formation. To understand the bonding of these intermediates with the
To shed further light on the mechanism of CTM conversion surface, we examined the projected density of states (PDOS)
on Ni3In, we performed first-principles calculations. Eight CO2 of bi-HCOO*Ni3In and trans-COOH*Ni3In (Figure S20),
which further show that trans-COOH interacts only with Ni
hydrogenation pathways (HCOO, CO hydrogenation, trans-
atoms on the surface while bi-HCOO exhibits charge transfer
COOH, and cis-COOH)60 (see Figures 5 and S19−S21) were
from both Ni and In atoms (Table S4). To comment on the
analyzed by determining their reaction energies to pinpoint the
kinetics of the CTM conversion on the present catalyst, we
most feasible pathway for CTM conversion on Ni3In. All
determined the activation energy barriers for each elementary
intermediates occurring on various CO2 reduction pathways, step along the cis-COOH and bi-HCOO pathways using the
with the exceptions of CH3O and bi-HCOO, stabilize in nudged elastic band (NEB) method. These pathways are very
configurations of adsorption at the surface Ni sites (Figure similar and differ only in the initial intermediates. From this
S20) indicating Ni atoms to be the active centers of (111) analysis, we find the following step to be the rate-determining
surface of Ni3In. Energy landscape plots show that the initial step with activation energy barrier of 1.2 eV (Figure 5):
and final steps, i.e., adsorption of CO2 and desorption of
CH3OH, are both nonspontaneous and exhibit an energy cost *HCO + *OH + 2H 2 → *H 2CO + *OH + H + H 2
of ∼0.11 and ∼0.25 eV respectively. The thermodynamically However, the formation of the first intermediate, i.e., trans-
most feasible CO2 reduction pathway has the lowest energy COOH and bi-HCOO along the carboxylic acid and formate
barrier associated with the potential-determining step (PDS). pathways, respectively, have different activation energy barriers
In the context of a thermocatalytic reaction, PDS gives a associated with them. Formation of bi-HCOO from adsorbed
measure of thermodynamic energy barrier. As is evident in CO2 has an activation energy of 0.57 eV, while the trans-
Figure S21, the CO hydrogenation pathway is energetically COOH intermediate requires crossing of energy barrier of 0.83
most expensive, and formation of *CO molecule from eV to form. Hence, the activation energy barrier and the
adsorbed *CO2 species is the PDS (ΔE = 0.9 eV). For the thermodynamic reaction energies of the second elementary
trans-COOH pathway, hydrogenation of *COH to form step in carboxylic acid and formate pathways conclusively show
*trans-HCOH is the PDS with an energy barrier of 0.82 eV. the latter to be more feasible. Therefore, CTM conversion on
Both cis-COOH and formate pathways exhibit similar energy Ni3In is most feasible via the formate pathway because of its
barriers of 0.25 eV, wherein the final step, i.e., desorption of low energy barrier and thermodynamically spontaneous
CH3OH, is the potential-determining step. hydrogenation of CO2 to yield bi-HCOO. These results are
513 https://dx.doi.org/10.1021/acsenergylett.0c02614
ACS Energy Lett. 2021, 6, 509−516
ACS Energy Letters http://pubs.acs.org/journal/aelccp Letter

concurrent with the in situ DRIFT studies and firmly confirm Research, Bangalore 560064, India; orcid.org/0000-
the formate pathway for selective methanol production over 0002-6592-4410
the operando generated Ni3In phase. On the basis of both Risov Das − New Chemistry Unit and School of Advanced
experimental and theoretical studies, the mechanism on Ni and Materials, Jawaharlal Nehru Centre for Advanced Scientific
Ni3In can be schematically represented as in Figure 5c. Research, Bangalore 560064, India
In conclusion, we demonstrate the concept of operando Chathakudath P. Vinod − Catalysis and Inorganic Chemistry
catalyst generation by identifying the metastable kinetic Ni3In Division, CSIR-National Chemical Laboratory, Pune
phase as a new efficient and stable catalyst for CTM 410008, India; orcid.org/0000-0001-9857-4907
conversion. The in situ transformation of the thermodynami- Shaojun Xu − UK Catalysis Hub, Didcot OX11 0FA, U.K.;
cally stable Ni7In3 phase highlights the importance of Cardiff Catalysis Institute, School of Chemistry, Cardiff
identifying and exploring the true active phases of catalysts University, Cardiff CF10 3AT, U.K.
when subjected to drastic thermocatalytic conditions. The Peter Wells − UK Catalysis Hub, Didcot OX11 0FA, U.K.;
catalytic selectivity of the cheap Ni catalyst was completely School of Chemistry, University of Southampton,
modulated toward the formate-mediated methanol pathway by Southampton, U.K.; Diamond Light Source, Didcot, U.K.
ordered alloying with In, as evinced by operando DRIFTS and Umesh V. Waghmare − School of Advanced Materials and
computational studies. This mechanistic tunability can be Theoretical Sciences Unit, Jawaharlal Nehru Centre for
attributed to the ordered arrangement of Ni and In atoms and Advanced Scientific Research, Bangalore 560064, India;
strong difference in their electronegativity, leading to orcid.org/0000-0002-9378-155X
interatomic charge transfer. More importantly, the catalyst’s Complete contact information is available at:
conversion efficiency can be optimized at much lower https://pubs.acs.org/10.1021/acsenergylett.0c02614
pressures, which has huge energy saving implications for the
overall process. This work attempts to advance the state-of-the- Author Contributions
art in CTM catalysis and demonstrate a new avenue for $
J.R. and L.D. contributed equally to this work.
targeting metastable active phases as efficient catalysts for CO2
reduction. Notes


The authors declare no competing financial interest.

*
ASSOCIATED CONTENT
sı Supporting Information
The Supporting Information is available free of charge at
■ ACKNOWLEDGMENTS
We thank Jawaharlal Nehru Centre for Advanced Scientific
https://pubs.acs.org/doi/10.1021/acsenergylett.0c02614. Research and the Department of Science and Technology,
India (DST) (Grant Number: DST/TM/EWO/MI/CCUS/
Experimental details for synthesis, characterizations
13(G)), DST nanomission (Grant Number: SR/NM/NS-
(XRD, EDAX, SEM, TEM, XPS, DRIFTS, and XAFS),
1125/2015I), and Technical Research Centre (TRC) (Grant
catalyst screening details, data calculation details, DFT
Number: JNC/AO/DST-TRC/C.14.10/16-2750) in JNCASR
calculations, additional experiment results, and addi-
for financial support. S.C.P. thanks DST for SwarnaJayanti
tional tables and figures (PDF)
Fellowship (Grant Number: DST/SJF/CSA-02/2017-18).

■ AUTHOR INFORMATION
Corresponding Author
Parts of this research were also carried out at the light source
PETRA III at DESY, a member of the Helmholtz Association
(HGF) and we thank Dr. Wolfgang Drube for assistance in
Sebastian C. Peter − New Chemistry Unit and School of using PETRA III beamline P64 at DESY, Germany. We also
Advanced Materials, Jawaharlal Nehru Centre for Advanced thank DST for financial assistance for the measurement at
Scientific Research, Bangalore 560064, India; orcid.org/ DESY. A.C.H. thanks University Grant Commission for the
0000-0002-5211-446X; Phone: 080-22082998; research fellowship. P.W. and S.X. thank UK Catalysis Hub
Email: sebastiancp@jncasr.ac.in, sebastiancp@gmail.com; Consortium and EPSRC for financial support (Grant numbers
Fax: 080-22082627 EP/K014706/1, EP/K014668/1, EP/K014854/1, EP/
K014714/1, and EP/I019693/1).


Authors
Arjun Cherevotan − New Chemistry Unit and School of REFERENCES
Advanced Materials, Jawaharlal Nehru Centre for Advanced (1) Aresta, M.; Dibenedetto, A.; Angelini, A. Catalysis for the
Scientific Research, Bangalore 560064, India valorization of exhaust carbon: from CO2 to chemicals, materials, and
Jithu Raj − New Chemistry Unit and School of Advanced fuels. technological use of CO2. Chem. Rev. 2014, 114 (3), 1709−
Materials, Jawaharlal Nehru Centre for Advanced Scientific 1742.
Research, Bangalore 560064, India (2) Peter, S. C. Reduction of CO2 to Chemicals and Fuels: A
Lakshay Dheer − School of Advanced Materials, Theoretical Solution to Global Warming and Energy Crisis. ACS Energy Lett.
Sciences Unit, and Chemistry and Physics of Materials Unit, 2018, 3 (7), 1557−1561.
Jawaharlal Nehru Centre for Advanced Scientific Research, (3) Goud, D.; Gupta, R.; Maligal-Ganesh, R.; Peter, S. C. Review of
Bangalore 560064, India; orcid.org/0000-0002-2196- Catalyst Design and Mechanistic Studies for the Production of Olefins
from Anthropogenic CO2. ACS Catal. 2020, 10 (23), 14258−14282.
2589 (4) Roy, S.; Cherevotan, A.; Peter, S. C. Thermochemical CO2
Soumyabrata Roy − New Chemistry Unit and School of hydrogenation to single carbon products: scientific and technological
Advanced Materials, Jawaharlal Nehru Centre for Advanced challenges. ACS Energy Lett. 2018, 3 (8), 1938−1966.
Scientific Research, Bangalore 560064, India (5) Tackett, B. M.; Gomez, E.; Chen, J. G. Net reduction of CO2 via
Shreya Sarkar − New Chemistry Unit and School of Advanced its thermocatalytic and electrocatalytic transformation reactions in
Materials, Jawaharlal Nehru Centre for Advanced Scientific standard and hybrid processes. Nat. Catal. 2019, 2 (5), 381−386.

514 https://dx.doi.org/10.1021/acsenergylett.0c02614
ACS Energy Lett. 2021, 6, 509−516
ACS Energy Letters http://pubs.acs.org/journal/aelccp Letter

(6) Ozin, G. A. You can’t have an energy revolution without (23) Kattel, S.; Liu, P.; Chen, J. G. G. Tuning Selectivity of CO2
transforming advances in materials, chemistry and catalysis into policy Hydrogenation Reactions at the Metal/Oxide Interface. J. Am. Chem.
change and action. Energy Environ. Sci. 2015, 8 (6), 1682−1684. Soc. 2017, 139 (29), 9739−9754.
(7) Goeppert, A.; Czaun, M.; Jones, J. P.; Prakash, G. K. S.; Olah, G. (24) Bai, S.; Shao, Q.; Wang, P.; Dai, Q.; Wang, X.; Huang, X.
A. Recycling of carbon dioxide to methanol and derived products - Highly Active and Selective Hydrogenation of CO2 to Ethanol by
closing the loop. Chem. Soc. Rev. 2014, 43 (23), 7995−8048. Ordered Pd−Cu Nanoparticles. J. Am. Chem. Soc. 2017, 139 (20),
(8) Artz, J.; Müller, T. E.; Thenert, K.; Kleinekorte, J.; Meys, R.; 6827−6830.
Sternberg, A.; Bardow, A.; Leitner, W. Sustainable conversion of (25) Li, M. M. J.; Zeng, Z. Y.; Liao, F. L.; Hong, X. L.; Tsang, S. C.
carbon dioxide: An integrated review of catalysis and life cycle E. Enhanced CO2 hydrogenation to methanol over CuZn nanoalloy in
assessment. Chem. Rev. 2018, 118 (2), 434−504. Ga modified Cu/ZnO catalysts. J. Catal. 2016, 343, 157−167.
(9) Zhong, J.; Yang, X.; Wu, Z.; Liang, B.; Huang, Y.; Zhang, T. (26) Prieto, G.; Zečević, J.; Friedrich, H.; de Jong, K. P.; de Jongh, P.
State of the art and perspectives in heterogeneous catalysis of CO2 E. Towards stable catalysts by controlling collective properties of
hydrogenation to methanol. Chem. Soc. Rev. 2020, 49 (5), 1385− supported metal nanoparticles. Nat. Mater. 2013, 12, 34.
1413. (27) Kattel, S.; Ramírez, P. J.; Chen, J. G.; Rodriguez, J. A.; Liu, P.
(10) Gonzalez-Garay, A.; Frei, M. S.; Al-Qahtani, A.; Mondelli, C.; Active sites for CO2 hydrogenation to methanol on Cu/ZnO catalysts.
Guillen-Gosalbez, G.; Perez-Ramirez, J. Plant-to-planet analysis of Science 2017, 355 (6331), 1296−1299.
CO2-based methanol processes. Energy Environ. Sci. 2019, 12 (12), (28) García-Trenco, A.; White, E. R.; Regoutz, A.; Payne, D. J.;
3425−3436. Shaffer, M. S. P.; Williams, C. K. Pd2Ga-based colloids as highly active
(11) Kondratenko, E. V.; Mul, G.; Baltrusaitis, J.; Larrazabal, G. O.; catalysts for the hydrogenation of CO2 to methanol. ACS Catal. 2017,
Perez-Ramirez, J. Status and perspectives of CO2 conversion into fuels 7 (2), 1186−1196.
and chemicals by catalytic, photocatalytic and electrocatalytic (29) García-Trenco, A.; Regoutz, A.; White, E. R.; Payne, D. J.;
processes. Energy Environ. Sci. 2013, 6 (11), 3112−3135. Shaffer, M. S. P.; Williams, C. K. PdIn intermetallic nanoparticles for
(12) Centi, G.; Quadrelli, E. A.; Perathoner, S. Catalysis for CO2 the hydrogenation of CO2 to methanol. Appl. Catal., B 2018, 220, 9−
conversion: a key technology for rapid introduction of renewable 18.
energy in the value chain of chemical industries. Energy Environ. Sci. (30) Shi, Z.; Tan, Q.; Tian, C.; Pan, Y.; Sun, X.; Zhang, J.; Wu, D.
2013, 6 (6), 1711−1731. CO2 hydrogenation to methanol over Cu-In intermetallic catalysts:
(13) Fujimoto, K.; Yu, Y. Spillover effect on the stabilization of Cu- effect of reduction temperature. J. Catal. 2019, 379, 78−89.
Zn catalyst for CO2 hydrogenation to methanol. Stud. Surf. Sci. Catal. (31) Marakatti, V. S.; Peter, S. C. Synthetically tuned electronic and
1993, 77, 393−396. geometrical properties of intermetallic compounds as effective
(14) Lunkenbein, T.; Girgsdies, F.; Kandemir, T.; Thomas, N.; heterogeneous catalysts. Prog. Solid State Chem. 2018, 52, 1−30.
Behrens, M.; Schlögl, R.; Frei, E. Bridging the time gap: A copper/ (32) Iihama, S.; Furukawa, S.; Komatsu, T. Efficient catalytic system
zinc oxide/aluminum oxide catalyst for methanol synthesis studied for chemoselective hydrogenation of halonitrobenzene to haloaniline
under industrially relevant conditions and time scales. Angew. Chem., using PtZn intermetallic compound. ACS Catal. 2016, 6 (2), 742−
Int. Ed. 2016, 55 (41), 12708−12712. 746.
(15) Behrens, M.; Studt, F.; Kasatkin, I.; Kuhl, S.; Havecker, M.; (33) Gamler, J. T. L.; Ashberry, H. M.; Skrabalak, S. E.; Koczkur, K.
Abild-Pedersen, F.; Zander, S.; Girgsdies, F.; Kurr, P.; Kniep, B. L.; M. Random alloyed versus intermetallic nanoparticles: A comparison
Tovar, M.; Fischer, R. W.; Norskov, J. K.; Schlogl, R. The active site of of electrocatalytic performance. Adv. Mater. 2018, 30 (40), 1801563.
methanol synthesis over Cu/ZnO/Al2O3 industrial catalysts. Science (34) Studt, F.; Sharafutdinov, I.; Abild-Pedersen, F.; Elkjær, C. F.;
2012, 336 (6083), 893−897. Hummelshøj, J. S.; Dahl, S.; Chorkendorff, I.; Nørskov, J. K.
(16) Kuld, S.; Conradsen, C.; Moses, P. G.; Chorkendorff, I.; Discovery of a Ni-Ga catalyst for carbon dioxide reduction to
Sehested, J. Quantification of zinc atoms in a surface alloy on copper methanol. Nat. Chem. 2014, 6, 320.
in an industrial-type methanol synthesis catalyst. Angew. Chem., Int. (35) Fiordaliso, E. M.; Sharafutdinov, I.; Carvalho, H. W. P.;
Ed. 2014, 53 (23), 5941−5945. Grunwaldt, J.-D.; Hansen, T. W.; Chorkendorff, I.; Wagner, J. B.;
(17) van den Berg, R.; Prieto, G.; Korpershoek, G.; van der Wal, L. Damsgaard, C. D. Intermetallic GaPd2 nanoparticles on SiO2 for low-
I.; van Bunningen, A. J.; Lægsgaard-Jørgensen, S.; de Jongh, P. E.; de pressure CO2 hydrogenation to methanol: Catalytic performance and
Jong, K. P. Structure sensitivity of Cu and CuZn catalysts relevant to in situ characterization. ACS Catal. 2015, 5 (10), 5827−5836.
industrial methanol synthesis. Nat. Commun. 2016, 7 (1), 13057. (36) Osswald, J.; Giedigkeit, R.; Jentoft, R. E.; Armbruster, M.;
(18) An, B.; Li, Z.; Song, Y.; Zhang, J.; Zeng, L.; Wang, C.; Lin, W. Girgsdies, F.; Kovnir, K.; Ressler, T.; Grin, Y.; Schlogl, R. Palladium-
Cooperative copper centres in a metal−organic framework for gallium intermetallic compounds for the selective hydrogenation of
selective conversion of CO2 to ethanol. Nat. Catal. 2019, 2 (8), acetylene - Part I: Preparation and structural investigation under
709−717. reaction conditions. J. Catal. 2008, 258 (1), 210−218.
(19) Zhang, X.; Han, S.; Zhu, B.; Zhang, G.; Li, X.; Gao, Y.; Wu, Z.; (37) Osswald, J.; Kovnir, K.; Armbruster, M.; Giedigkeit, R.; Jentoft,
Yang, B.; Liu, Y.; Baaziz, W.; Ersen, O.; Gu, M.; Miller, J. T.; Liu, W. R. E.; Wild, U.; Grin, Y.; Schlogl, R. Palladium-Gallium intermetallic
Reversible loss of core−shell structure for Ni−Au bimetallic compounds for the selective hydrogenation of acetylene - Part II:
nanoparticles during CO2 hydrogenation. Nat. Catal. 2020, 3 (4), Surface characterization and catalytic performance. J. Catal. 2008, 258
411−417. (1), 219−227.
(20) Liu, C.; Yang, B.; Tyo, E.; Seifert, S.; DeBartolo, J.; von (38) Jiao, F.; Frei, H. Nanostructured cobalt oxide clusters in
Issendorff, B.; Zapol, P.; Vajda, S.; Curtiss, L. A. Carbon dioxide mesoporous silica as efficient oxygen-evolving catalysts. Angew. Chem.,
conversion to methanol over size-selected Cu4 clusters at low Int. Ed. 2009, 48 (10), 1841−1844.
pressures. J. Am. Chem. Soc. 2015, 137 (27), 8676−8679. (39) Xu, C.; Chen, G.; Zhao, Y.; Liu, P.; Duan, X.; Gu, L.; Fu, G.;
(21) Parastaev, A.; Muravev, V.; Huertas Osta, E.; van Hoof, A. J. F.; Yuan, Y.; Zheng, N. Interfacing with silica boosts the catalysis of
Kimpel, T. F.; Kosinov, N.; Hensen, E. J. M. Boosting CO2 copper. Nat. Commun. 2018, 9 (1), 3367.
hydrogenation via size-dependent metal−support interactions in (40) Yang, X.; Kattel, S.; Senanayake, S. D.; Boscoboinik, J. A.; Nie,
cobalt/ceria-based catalysts. Nat. Catal. 2020, 3 (6), 526−533. X.; Graciani, J.; Rodriguez, J. A.; Liu, P.; Stacchiola, D. J.; Chen, J. G.
(22) Qian, C.; Sun, W.; Hung, D. L. H.; Qiu, C.; Makaremi, M.; Low Pressure CO2 Hydrogenation to Methanol over Gold Nano-
Hari Kumar, S. G.; Wan, L.; Ghoussoub, M.; Wood, T. E.; Xia, M.; particles Activated on a CeOx/TiO2 Interface. J. Am. Chem. Soc. 2015,
Tountas, A. A.; Li, Y. F.; Wang, L.; Dong, Y.; Gourevich, I.; Singh, C. 137 (32), 10104−10107.
V.; Ozin, G. A. Catalytic CO2 reduction by palladium-decorated (41) Lam, E.; Larmier, K.; Wolf, P.; Tada, S.; Safonova, O. V.;
silicon−hydride nanosheets. Nat. Catal. 2019, 2 (1), 46−54. Copéret, C. Isolated Zr Surface Sites on Silica Promote Hydro-

515 https://dx.doi.org/10.1021/acsenergylett.0c02614
ACS Energy Lett. 2021, 6, 509−516
ACS Energy Letters http://pubs.acs.org/journal/aelccp Letter

genation of CO2 to CH3OH in Supported Cu Catalysts. J. Am. Chem. (60) Tang, Q.; Shen, Z.; Huang, L.; He, T.; Adidharma, H.; Russell,
Soc. 2018, 140 (33), 10530−10535. A. G.; Fan, M. Synthesis of methanol from CO2 hydrogenation
(42) Zhao, D.; Huo, Q.; Feng, J.; Chmelka, B. F.; Stucky, G. D. promoted by dissociative adsorption of hydrogen on a Ga3Ni5(221)
Nonionic triblock and star diblock copolymer and oligomeric surface. Phys. Chem. Chem. Phys. 2017, 19 (28), 18539−18555.
surfactant syntheses of highly ordered, hydrothermally stable,
mesoporous silica structures. J. Am. Chem. Soc. 1998, 120 (24),
6024−6036.
(43) Kattel, S.; Yan, B.; Chen, J. G.; Liu, P. CO2 hydrogenation on
Pt, Pt/SiO2 and Pt/TiO2: Importance of synergy between Pt and
oxide support. J. Catal. 2016, 343, 115−126.
(44) Wu, H. C.; Chang, Y. C.; Wu, J. H.; Lin, J. H.; Lin, I. K.; Chen,
C. S. Methanation of CO2 and reverse water gas shift reactions on Ni/
SiO2 catalysts: the influence of particle size on selectivity and reaction
pathway. Catal. Sci. Technol. 2015, 5 (8), 4154−4163.
(45) Richard, A. R.; Fan, M. Low-Pressure Hydrogenation of CO2 to
CH3OH Using Ni-In-Al/SiO2 Catalyst Synthesized via a Phyllosilicate
Precursor. ACS Catal. 2017, 7 (9), 5679−5692.
(46) Martin, O.; Martin, A. J.; Mondelli, C.; Mitchell, S.; Segawa, T.
F.; Hauert, R.; Drouilly, C.; Curulla-Ferre, D.; Perez-Ramirez, J.
Indium oxide as a superior catalyst for methanol synthesis by CO2
hydrogenation. Angew. Chem., Int. Ed. 2016, 55 (21), 6261−6265.
(47) Ye, J.; Liu, C.; Mei, D.; Ge, Q. Active oxygen vacancy site for
methanol synthesis from CO2 hydrogenation on In2O3(110): A DFT
study. ACS Catal. 2013, 3 (6), 1296−1306.
(48) Jia, X.; Sun, K.; Wang, J.; Shen, C.; Liu, C.-j. Selective
hydrogenation of CO2 to methanol over Ni/In2O3 catalyst. J. Energy
Chem. 2020, 50, 409−415.
(49) Rui, N.; Zhang, F.; Sun, K.; Liu, Z.; Xu, W.; Stavitski, E.;
Senanayake, S. D.; Rodriguez, J. A.; Liu, C.-J. Hydrogenation of CO2
to Methanol on a Auδ+−In2O3−x Catalyst. ACS Catal. 2020, 10 (19),
11307−11317.
(50) Snider, J. L.; Streibel, V.; Hubert, M. A.; Choksi, T. S.; Valle, E.;
Upham, D. C.; Schumann, J.; Duyar, M. S.; Gallo, A.; Abild-Pedersen,
F.; Jaramillo, T. F. Revealing the synergy between oxide and alloy
phases on the performance of bimetallic In−Pd catalysts for CO2
hydrogenation to methanol. ACS Catal. 2019, 9 (4), 3399−3412.
(51) Aziz, M. A. A.; Jalil, A. A.; Triwahyono, S.; Mukti, R. R.; Taufiq-
Yap, Y. H.; Sazegar, M. R. Highly active Ni-promoted mesostructured
silica nanoparticles for CO2 methanation. Appl. Catal., B 2014, 147,
359−368.
(52) Aziz, M. A. A.; Jalil, A. A.; Triwahyono, S.; Sidik, S. M.
Methanation of carbon dioxide on metal-promoted mesostructured
silica nanoparticles. Appl. Catal., A 2014, 486, 115−122.
(53) Noren, L.; Larsson, A. K.; Withers, R. L.; Rundlof, H. A
neutron and X-ray powder diffraction study of B8(2) related
superstructure phases in the Ni-In system. J. Alloys Compd. 2006,
424 (1−2), 247−254.
(54) Marakatti, V. S.; Arora, N.; Rai, S.; Sarma, S. C.; Peter, S. C.
Understanding the role of atomic ordering in the crystal structures of
NixSny toward efficient vapor phase furfural hydrogenation. ACS
Sustainable Chem. Eng. 2018, 6 (6), 7325−7338.
(55) Yan, X.; Tian, L.; Chen, X. Crystalline/amorphous Ni/NiO
core/shell nanosheets as highly active electrocatalysts for hydrogen
evolution reaction. J. Power Sources 2015, 300, 336−343.
(56) Tsoukalou, A.; Abdala, P. M.; Stoian, D.; Huang, X.; Willinger,
M.-G.; Fedorov, A.; Müller, C. R. Structural evolution and dynamics
of an In2O3 catalyst for CO2 hydrogenation to methanol: an operando
XAS-XRD and in Situ TEM study. J. Am. Chem. Soc. 2019, 141 (34),
13497−13505.
(57) Courtois, M.; Teichner, S. J. Infrared studies of CO, O2, and
CO2 gases and their interaction products, chemically adsorbed on
nickel oxide. J. Catal. 1962, 1 (2), 121−135.
(58) Campuzano, J. C.; Greenler, R. G. The adsorption sites of CO
on Ni(111) as determined by infrared reflection-absorption spectros-
copy. Surf. Sci. 1979, 83 (1), 301−312.
(59) Vogt, C.; Groeneveld, E.; Kamsma, G.; Nachtegaal, M.; Lu, L.;
Kiely, C. J.; Berben, P. H.; Meirer, F.; Weckhuysen, B. M. Unravelling
structure sensitivity in CO2 hydrogenation over nickel. Nat. Catal.
2018, 1 (2), 127−134.

516 https://dx.doi.org/10.1021/acsenergylett.0c02614
ACS Energy Lett. 2021, 6, 509−516

You might also like