1 s2.0 S2666386423000693 Main

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

ll

OPEN ACCESS

Review
Atomically dispersed metals as potential
coke-resistant catalysts for dry
reforming of methane
Qian Zhang,1,2 Mohcin Akri,3 Yiwen Yang,1,* and Botao Qiao1,*

SUMMARY
Dry reforming of methane, the reforming of methane with carbon
dioxide to produce syngas, is important within the chemical industry
and has aroused wide interest from both industrial and academic
fields. Many research efforts have been devoted to the develop-
ment of catalysts with both excellent activity and stability. One of
the most common and critical issues for traditional nickel-based cat-
alysts is severe deactivation due to carbon deposition. Recently,
atomically dispersed metals have been found to be promising
coke-resistant catalysts for dry reforming of methane due to their
superior catalytic activity and high resistance to carbon deposition.
This review focuses on the latest developments of atomically
dispersed metal catalysts based on nickel, as well as noble metals,
for dry reforming of methane. Recent understanding of the mecha-
nisms underlying the coke-resistant properties is illustrated, and the
critical challenges and perspectives for the future development of
atomically dispersed catalysts are summarized.

INTRODUCTION
Carbon dioxide (CO2) and methane (CH4) are two of the main greenhouse gases
responsible for global warming, but are also abundant and low-cost carbon sour-
ces.1,2 In particular, CH4 has been considered as a relatively clean energy source
with which to realize a low-carbon economy.3,4 Dry reforming of methane (DRM), a
process of simultaneous conversion of CH4 and CO2 to syngas (H2 + CO), is consid-
ered as one of the most important approaches to address the challenge of miti-
gating both greenhouse gases and producing versatile feedstock (syngas with
desired H2/CO ratio that is typically lower than 2 required for Fischer-Tropsch syn-
thesis) for manufacturing diverse chemicals such as methanol, formic acid and higher
alkanes.1,5–7 The DRM reaction has been studied intensively for more than 30 years
using various catalysts including noble and non-noble metals.8–14 Among them,
nickel (Ni)-based catalysts are the most extensively investigated, and have been re-
garded as promising candidates for DRM due to their high comparable activity to 1CAS Key Laboratory of Science and Technology
noble metals and wide availability (low cost).1,15–17 on Applied Catalysis, Dalian Institute of Chemical
Physics, Chinese Academy of Sciences, Dalian
116023, China
Despite the advances, it is still challenging to develop Ni-based catalyst systems for 2Universityof Chinese Academy of Sciences,
industrial scale application, as the catalysts undergo rapid deactivation due to car- Beijing 100049, China
3Unité de Catalyse et Chimie du Solide(UCCS),
bon deposition. The carbon deposition is mainly derived from the methane decom-
Université de Lille, UMR 8181 CNRS, Centrale
position and CO disproportionation (which dominate at different temperatures, Lille, F-59000 Lille, France
respectively), thermal deterioration of the catalyst, and sintering of metal species *Correspondence: yangyiwen@dicp.ac.cn (Y.Y.),
at high temperature.1,6,18,19 With the aim to avoid or reduce the carbon deposition bqiao@dicp.ac.cn (B.Q.)
on Ni-based catalysts, development and investigations in catalyst design and https://doi.org/10.1016/j.xcrp.2023.101310

Cell Reports Physical Science 4, 101310, March 15, 2023 ª 2023 The Author(s). 1
This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
ll
OPEN ACCESS
Review

Figure 1. Less coke formation on ultra-fine nickel nanoparticles


Less coke formation is observed on ultra-fine nickel nanoparticles (2–3 nm) uniformly distributed on SBA-15 in DRM. Figure reprinted with permission
(Copyright 2020, Elsevier 31 ).

reactor operation technologies via adjusting catalyst structure, surface chemistry,


and reaction conditions have been reported.20–27 It is commonly accepted in litera-
ture that the particle size of the metal is a critical parameter for the control of carbon
deposition in DRM.28 Various studies have concluded that the catalysts with larger
metal particle sizes are more favorable for carbon deposition, while the catalysts
with small metal particle sizes are more resistant to carbon deposition.25,28,29
Lercher et al.30 demonstrated that when nickel particles are below 2 nm, resistance
to coke deposition is significantly improved. And Ning et al.31 found that the small
and dispersed Ni particles performed better in suppressing carbon deposition for
the DRM reaction (Figure 1 31,).

However, small Ni particles generally exhibit poor thermal stability and are prone
to sintering under harsh reaction conditions, leaving a dilemma for the application
of metal catalysts in DRM reaction. There are some reports of nanocatalysts that try
to solve the dilemma. Song et al.32 reported that the aggregation of small molyb-
denum-doped Ni nanoparticles on the surface of single-crystalline MgO support to
step edges to form larger, highly stable nanoparticles also passivated sites for
coking, producing a catalyst with high activity and longevity at 800 C for DRM.
However, the metal atom utilization is rather low for large nanoparticles and the
long induction time for the bimetallic Ni-Mo catalyst may affect its industrial appli-
cation.32,33 Other relevant reports of Ni nanocatalysts wrapped in two-dimensional

2 Cell Reports Physical Science 4, 101310, March 15, 2023


ll
OPEN ACCESS
Review

molecular sieves showed good performance for DRM as a result of highly


dispersed active Ni particles, but a small amount of carbon deposit was still inev-
itable,34–36 and the active sites of these encapsulated Ni-based nanocatalysts may
be partially coated, also causing low metal atom utilization. Thus, efficient coke-
resistant metal catalysts with high metal atom utilization for DRM require further
development.

Considering the high resistance to carbon deposition of small-size metal particles, it


seems isolated metal atoms should be the promising catalysts to suppress carbon
deposition. Recently, supported atomically dispersed metals (ADMs) as a new class
of heterogeneous catalysts that contain atomically dispersed metal active sites and
countable numbers of atoms (only one atom or several atoms forming a cluster)37,38
have attracted considerable interest and given rise to extensive studies,39–42 exhib-
iting outstanding performance in many important industrial thermochemical reac-
tions such as selective oxidation, water gas shift reaction, selective hydrogenation,
hydroformylation and selective CO2 reduction,43–48 as well as in some electrochem-
ical and photochemical conversions.49–53 In this review, both single-atom catalysts
(SACs)54 and atomically dispersed clusters with a few atoms55,56 are classified as
ADMs catalysts,38,57,58 emphasizing the atomic dispersion of the catalytic metal cen-
ters as previously mentioned.59,60 In comparison, ADMs, especially SACs, have
great advantage of high atomic utilization, uniform metal active sites, and elimina-
tion of size effects.

In general, highly dispersed metal atoms are more thermodynamically unstable due
to their high surface energy and tend to aggregate into large metal clusters and
even nanoparticles at high temperature. But recent research in the past few years
has shown that in some cases single atoms can be more stable than nanoparticles.
For example, in 2015, it was reported that isolated gold single atoms dispersed on
iron oxide nano crystallites (Au1/FeOx) were much more sintering-resistant than cor-
responding Au nanoparticles for CO oxidation due to the formation of covalent
metal-support interaction (CMSI) between surface-anchored Au1 atoms with high
valent states and the lattice oxygen atoms of the FeOx support.61 Subsequently,
a strong covalent metal-support interaction was successfully applied to construct
a thermally stable high-metal-loading Pt single-atom catalyst, which could remain
stable after calcination under 800 C.62 Interesting phenomenon was even found
that the supported Pt nanoparticles would disperse spontaneously into more stable
single Pt atom on reducible iron oxide surface under high-temperature calcination
or during the methane combustion reaction. Other studies have described similar
phenomena. Datye et al.63 found that the strong interaction between CeO2 and
Pt could anchor Pt atoms stably on the crystal surface of CeO2 so that the atomically
dispersed Pt could remain stable at high temperature without agglomeration. In
brief, it is expected to achieve high temperature stability for highly dispersed metal
catalysts by constructing strong metal-support interaction (SMSI).

Therefore, stable ADMs with highly dispersed catalytic metal centers show great po-
tential in achieving more efficient and economic DRM processes. The advantages
and superior characteristics of ADMs in DRM, compared with traditional Ni-based
catalysts, are demonstrated by several examples that were reported in recent years.
Next, the most recent understanding on the underlying mechanisms of ADMs to
exhibit high resistance to carbon deposition in DRM are illustrated and discussed.
At the end of the review, the critical challenges and perspectives for the future devel-
opment of ADMs in DRM are also provided.

Cell Reports Physical Science 4, 101310, March 15, 2023 3


ll
OPEN ACCESS
Review

Figure 2. Overview of the coke


formation and coke removal in the
DRM reaction
Different reactions in the process
of coke formation and removal are
shown in the scheme. Figure is
reprinted with permission
(Copyright 2020, Wiley-VCH 67 ).

Potential of ADMs as coke-resistant catalysts for DRM


The first thing to clarify here should be the potential of ADMs with excellent resis-
tance to carbon deposition for DRM. The deactivation mechanisms of catalysts in
DRM have been well investigated before, basically the causes of catalyst deactiva-
tion can be classified into three: mechanical (attrition or crushing), chemical
(poisoning, carbon deposition, and formation of inactive phase), and thermal (sinter-
ing and aggregation).6,64–66 Among them, for conventional Ni-based catalysts in
DRM reaction, the main reasons that result in rapid deactivation are sintering of
the active metal species and coking via Boudouard reaction (2CO / C + CO2)
and/or CH4 dissociation (CH4 / C + 4H), as shown in Figure 2.67 Under the typical
reaction temperature of DRM, which is usually in the range of 750–900 C, the forma-
tion of coke dominantly derived from CH4 dissociation, as the Boudouard reaction
tends to occur at relatively lower temperature (e.g., 557–700 C).1,67–69

For better understanding of the coking issue in DRM, it is necessary to recall the clas-
sical mechanism reported in the literature on metal nanoparticles, especially those
Ni-based catalysts. Several authors suggested that CH4 decomposes totally on
the surface of nanoparticles to form H2 and carbon (Figure 2), at the same time
CO2 converted to CO and O and subsequently the deposited carbon reacted with
dissociated oxygen to generate CO,70,71 while the unreacted carbon becomes
coke. Regarding CH4 transformation, the activation of the first C-H has always
been an extremely hard yet crucial step due to the high stability and good symmetry
of the CH4 molecule. According to the mechanism of carbon deposition, it is ideal for
a DRM catalyst to just activate the first C-H bond of CH4 for the subsequent oxidation
without the generation of C. However, traditional supported metal catalysts often
tend to activate all the C-H bonds in CH4, resulting in the complete dissociation,
which may further lead to carbon formation and the decrease of product selectivity.
Therefore, an ideal catalyst required should efficiently activate methane, and equally
importantly, it should also possess the ability to stabilize the dissociated entities,
such as methyl.72 According to the reported studies on methane dissociation on
metal surfaces, the first C-H bond cleavage of methane decomposition to methyl
needs only one single metal atom, while several metal atoms are requested for
the further three dehydrogenation reactions on the surface.72–76 That means it is crit-
ical for a good DRM catalyst to activate the first C-H bond of methane while avoiding
complete CH4 dehydrogenation.

Relevant results have also been reported on various SACs for direct methane conver-
sion. For instance, Guo et al.76 reported that atomically dispersed iron single atoms
are effective sites for direct conversion of methane to high-value-added products
without any carbon deposition detected. Besides, nanoceria supported Pt SACs

4 Cell Reports Physical Science 4, 101310, March 15, 2023


ll
OPEN ACCESS
Review

Figure 3. Comparison of the reaction pathways between pure and single-atom alloy surfaces
(A and B) (A) Optimized configurations from DFT for the most stable adsorptions of CxHy species (x = 0,1; y = 0, 1, 2, 3, 4) on the Pt/Cu (111) SAA surface
and (B) the transition states (TSs) for C-H bond scissions from CH 4 * to CH 3 * + H* (TS1), CH 3 * to CH 2 * + H* (TS2), CH 2 * to CH* + H* (TS3), and CH* to C* +
H* (TS4) on the Pt/Cu(111) SAA surface.
(C) Energy landscape for the full dehydrogenation of adsorbed CH4* on Cu(111) (black), Pt/Cu(111) SAA (red), and Pt(111) (blue). Reprinted with
permission (Copyright 2018, Nature Publishing Group 79 ).

were designed for the methane conversion reaction and confirmed to be capable of
stabilizing C2 adsorbates, resulting in suppression of carbon deposition.77 It is a
common challenge in direct conversion of methane to achieve selective C-H activa-
tion of CH4. Bai et al.78 reported that cerium dioxide (CeO2) nanowires supported
single Rh atoms and Rh nanoparticles show different behavior in the direct conver-
sion of methane to oxygenates under mild conditions. The experimental data and
density functional theory (DFT) calculation indicated that CH4 was activated
selectively to *CH3 radical to form the oxygenate on Rh SACs directly, while total
oxidation of CH4 to form COx is favored on Rh nanoparticles. As a result, methane
was converted to oxygenates more selectively over Rh1/CeO2 than Rh NPs/CeO2.

By DFT calculations, Marcinkowsi et al.79 investigated the succession of C-H scis-


sions in methane (Figure 379). It was found that compared with the pure metals,
Pt/Cu single-atom alloys (SAAs) have intermediate barriers for C-H activation but
beneficial Cu-like reaction energies, such that the SAA can perform C-H activation

Cell Reports Physical Science 4, 101310, March 15, 2023 5


ll
OPEN ACCESS
Review

without carbon deposition. Meanwhile, Hu et al.72 constructed a model structure


that consists of isolated Pt atoms on a MoO3 (010) surface and illustrated that the
single Pt atom in Pt/MoO3 can easily activate the first C-H bond in methane with a
reaction barrier of 0.32 eV, notably lower than that on a Pt (111) surface, while the
further dehydrogenation reactions would be hindered by significantly increased
energy barrier, which means CH3 dissociation on the Pt/MoO3 substrate is energet-
ically prohibited. This finding was consistent with the earlier findings that reported
carbon formation of alkanes on the surface of platinum-based catalysts was a struc-
turally sensitive reaction, which usually required at least three adjacent Pt atoms to
cause carbon deposition.80 It also agreed with our recent finding that there was no
carbon deposition in the DRM reaction with carbon dioxide over isolated Ni atom
catalyst supported on Ce-doped hydroxyapatite due to its particular ability to
only activate the first C-H bond in CH4.7

Apart from the high resistance of ADMs to carbon deposition derived from CH4
dissociation, certain ADMs also show great potential in the inhibition of coke forma-
tion from Boudouard reaction in low-temperature DRM.81,82 Recently, a highly car-
bon-resistant nanotubular yolk-shell Pt-NiCe@SiO2 SAA catalyst that presents an
excellent stability for 120 h during DRM at 500 C was reported by Kim et al.82 It
was demonstrated that the synergetic combination of the confined yolk-shell
morphology and Pt-Ni SAA structures prevented carbon formation and provided
excellent catalyst stability. The facile CO desorption from the yolk-shell Pt-
NiCe@SiO2 greatly impeded carbon deposition derived from Boudouard reaction.
At the same time, the atomically dispersed Pt on the Ni yolks facilitated Pt-Ni inter-
actions and enhanced the reducibility of the Ni species, which further suppressed
carbon formation during DRM.82

Demonstration of ADM catalysts in the DRM reaction


The above research results raised the possibility that the construction of ADMs with
excellent stability has the potential to solve the carbon deposition problem in DRM
reaction, and there have been some representative reports on the excellent perfor-
mance of ADM catalysts for DRM reactions. Tang et al.83 investigated DRM over Ni1
and Ru1 single-atom supported on CeO2 nanorod, the catalyst performed high
activity and stability at 500–600 C without carbon deposition detected, and a syner-
getic effect between Ni1 and Ru1 single atoms was observed as evidenced by the
lower apparent activation barrier and high activity and stability compared with Ni1
or Ru1 monoatomic catalyst (Figure 4). The excellent performance of the bimetallic
catalyst was attributed to the different roles of Ni1 and Ru1 in DRM process: The acti-
vation of CH4 mainly occurred on Ni1 site, while Ru1 sites drove the activation of CO2
and played a complementary role for the activation of CH4 and formation of H2.

DRM has also been investigated in our group over Ni1 single atoms supported on
hydroxyapatite (HAP) and cerium-substituted hydroxyapatite (HAP-Ce).7 The sup-
ported Ni single atoms initially exhibited high activity and high resistance to carbon
deposition, while the rapid catalyst deactivation was attributed to the easy aggrega-
tion of Ni1 single atoms to large metal particles during DRM over Ni1/HAP (Fig-
ure 5A7). More importantly, we demonstrated that Ce doping of HAP induced
SMSI which prevented Ni single atoms from sintering and resulted in a high activity
and stability for 100 h DRM with negligible carbon deposition (Figure 5B). Zhang
et al.84 also proposed that the SMSI effect could suppress the sintering of metal
Ni sites during the DRM reaction. They reported that atomically dispersed Ni
anchored in ceria-upgraded boron nitride (Ni/CeO2-BN) showed three times in-
crease compared with Ni/BN in terms of specific activity. The introduction of active

6 Cell Reports Physical Science 4, 101310, March 15, 2023


ll
OPEN ACCESS
Review

Figure 4. Scheme for a catalyst consisting of two sets of single-atom sites alongside the catalytic
performance of Ru1/CeOx, Ni1/CeOx, and Ni1-Ru1/CeOx in DRM
(A) Schematic of a catalyst surface consisting of two sets of single-atom sites.
(B)Turnover frequency (TOF) of reforming CH 4 with CO2 in terms of hydrogen production per Ni 1
site of Ce 0.95 Ni 0.05 O2 , per Ru 1 site of Ce 0.95 Ru 0.05 O 2 , and per Ni 1 or Ru 1 site for
Ce 0.95 Ni 0.025 Ru 0.025 O2 at 500–560  C. Reprinted with permission (Copyright 2019, American
Chemical Society 83 ).

metal oxides into inert support could enhance the interaction between metal and
support.

Zhou et al.85 reported a plasmonic photocatalyst that consisted of single-atom Ru


active sites on Cu ‘‘antenna’’ nanoparticle surface, ideal for low-temperature,
light-driven DRM reaction with highly efficient and carbon-resistant performance
(Figure 6). It was proven that photo-excited hot carriers and single atomic active sites
contributed to the excellent performance of the catalytic system. The presence of
hot carriers could accelerate the associative desorption of H2, consequently sup-
pressing the bonding between Hads and Oads and achieving high photocatalytic
selectivity. In addition, this suppression also led to the effective removal of Cads
via oxidation gasification by Oads (Cads + Oads/CO(g)), thus, playing an important
role in eliminating coking and improving the stability of the catalysts. Quantum me-
chanical modeling also showed that single-atom doping of Ru on Cu (111) surface,
combined with excited state activation, resulted in a significant reduction of the bar-
rier for CH4 activation.

Mechanistic understanding of coke-resistance for ADM catalysts


It has been reported in several works that the catalyst resistance of carbon deposi-
tion increases with decreasing of metal particle size,7,28,30,86 and different mecha-
nisms are proposed to explain this phenomenon. As mentioned above, we have
recently reported an atomically dispersed nickel single atoms as carbon-resistant
active sites for DRM at 750 C. Control experiments showed that CH4 decomposition
occurred on both Ni nanoparticle and nickel single atoms, but the products were
completely different. For Ni nanoparticle catalyst (10Ni/HAP-Ce), a high amount
of carbon deposition was detected with much more hydrogen formation than
ethylene. However, on Ni SACs (2Ni1/HAP-Ce), no carbon formation was detected
and the formation of ethylene was much higher than H2, suggesting less degrees
of CH4 decomposing and more C-C coupling product.7 It was concluded that Ni sin-
gle atoms favor only a partial dehydrogenation of CH4, thus enabling more C-C
coupling, intrinsically avoiding the complete dehydrogenation and carbon deposi-
tion. In good agreement with the experiment results, theoretical DFT calculation
demonstrated that the activation of the first C-H bond in CH4 is efficient on single
Ni atoms, while the subsequent dehydrogenation of CH3 is not favorable thermody-
namically. Due to the high thermodynamic barrier of the dehydrogenation of CH3 to

Cell Reports Physical Science 4, 101310, March 15, 2023 7


ll
OPEN ACCESS
Review

Figure 5. DRM performance over Ni/HAP and Ni/HAP-Ce samples


(A) CH 4 conversion during DRM over HAP supported Ni catalysts.
(B) CH 4 conversion during DRM over HAP-Ce supported Ni catalysts. Conditions: T = 750  C, CH 4 /CO 2 /He = 10/10/30, total flow = 50 mL min 1
, GHSV =
60,000 mL h 1 g 1 . Reprinted with permission (Copyright 2019, Nature Publishing Group7 ).

CH2, the formation CH3O is thermodynamically favored as an alternative intermedi-


ate way for the transformation of CH3 to CO instead of complete dehydrogenation
(Figure 77). Liu et al.87 have reported similar mechanism over Ni4 single site identi-
fying the oxidation of CH3 to CH3O as an alternative way for CO generation, which
has also been mentioned on bimetallic single atoms catalysts Ni1-Ru1/CeO2.83

Despite the above-mentioned SACs showing excellent performance in DRM, Liu and
co-workers reported that single Ni atoms were unfavorable to decompose CH4 and
CO2, while a single-site catalyst with Ni4 clusters can achieve the efficient activation
of CH4 and CO2.87 For a better understanding of DRM, with the combination of DFT
calculations, Kinetic Monte Carlo simulations and coordinative experimental
studies, they explored the difference of reaction mechanism between single-atom
Ni1/MgO (001) and single-site Ni4/MgO (001) catalysts and the remarkable promot-
ing effect of site confinement of the single Ni4 site on the catalytic activity.
Compared with MgO alone, the Ni1/MgO was not enough to dissociate CO2 due
to weaker *CO2 adsorption but stronger binding with CO. In addition, the decom-
position of *CH4 to *CH3 over single Ni atom was more difficult than Ni (111) surface.
For Ni4/MgO, Ni4 clusters as single catalytic sites were atomically dispersed on MgO
(Figure 8 87, 88), and the *CO2 dissociation to *CO + *O was accelerated but the
further decomposition of *CO to *C was hindered as the instability of the final state
corresponds to the co-adsorption of *CO + *O by 0.88 eV than their separate ad-
sorptions. Besides, the single Ni4 site facilitated the *C oxidation to suppress coking
formation from CO2 dissociation. For the dehydrogenation of *CH4 to *CH3, in com-
parison with Ni1/MgO (100) (Ea = 1.57 eV) and Ni (111) (Ea = 0.91 or 1.17 eV), it was
also more facile on Ni4/MgO (100) surface (DE = 0.36 eV and Ea = 0.53 eV), while the
formation of *C from *CH dissociation (DE = 0.63 eV and Ea = 1.06 eV) was energet-
ically prohibited.

Except for the nature of ADMs, the properties of the supports also play an important
role in the suppression of carbon deposition during DRM. Wu et al.88 prepared a se-
ries of CeO2 supported Ni1 catalysts with varied oxygen vacancy (OV) concentration
regulated on CeO2 by introducing a smaller-size metal cation M (M = Mg, Co, Zn).
During the DRM process, isolated Ni atoms exhibit high CH4 dissociation activity
but suffered from rapid deactivation caused by the carbon deposition. Interestingly,

8 Cell Reports Physical Science 4, 101310, March 15, 2023


ll
OPEN ACCESS
Review

Figure 6. Effect of ruthenium concentration on the photocatalytic behavior of Cu/Ru surface alloys
(A) Schematic of a Cu single-atom Ru surface alloy catalyst with the dry reforming reactants and products shown on the left. Reaction rate and long-term
stability (B) and selectivity (C) of photocatalytic DRM are shown. Schematics of the compositional dependence of the Cu x Ru y photocatalyst with respect
to carbon resistance includes pure Cu (D), low Ru loading (E), and high Ru loading (F). Reprinted with permission (Copyright 2020, Nature Publishing
Group 85 ).

along with the increase of the OV concentration from 21.9% to 30.8%, the amount of
carbon deposition decreased by 50%. It was proposed that OV activated CO2 to
produce CO and Oad species, which was in favor of carbon removal and activity
retaining (Figure 8C). DFT calculations were carried out to verify the mechanism
of OV formation and the carbon removal on OV. The formation energy of OV (EV)
and adsorption energy (Eads) of CO2 on CeO2 with a second metal was investigated.
It was shown that EV of CeO2 with a second metal follows the order:
CoCe < ZnCe < MgCe < pure CeO2, indicating that Co was the most favorable
for the formation of OV. In addition, the catalyst with more OV exhibited lower
Eads in Figure 8D, suggesting that OV were the favorable active sites for CO2 adsorp-
tion and activation.

Outlook for ADM catalysts


According to the examples introduced above, ADMs show great potential in maxi-
mizing the atom utilization of noble metals as well as prohibiting carbon deposition
during the DRM process. Determined by the intrinsic thermodynamics of DRM, the
dissociation of CH4 that occurs on the active metal sites is much more difficult
compared with the activation of CO2. On the traditional metal nanoparticles, CH4
tends to dissociate completely and generate *C species that are apt to form coke
on the catalyst. Differently, on the ADMs, CH4 undergoes partial dissociation to
form *CH3 that could be converted to CO subsequently without carbon deposition.
Therefore, ADMs possess the ability to resist carbon deposition intrinsically.

Generally, harsh reaction conditions, for example reaction temperatures as high as


900 C, are required for the highly endothermic DRM process to obtain reasonable

Cell Reports Physical Science 4, 101310, March 15, 2023 9


ll
OPEN ACCESS
Review

Figure 7. DFT calculation for CH4 decomposition


(A) Potential energy diagram from DFT calculations of CH 4 dissociation over Ni 1 /CeO 2 .
(B and C) (B) Potential energy diagram from DFT calculations of CH 3 oxidation and CH 3 O
dehydrogenation, and (C) corresponding geometries over Ni 1 /CeO 2 . Optimized structures for
reaction intermediates are shown inset (Ce: yellow, Ni: blue, O: red, C: black, H: white). Reprinted
with permission (Copyright 2019, Nature Publishing Group 7 ).

syngas yield. This poses great challenges to catalyst stability and process engineer-
ing. There are two design directions for DRM catalysts: one direction is to design
ADM catalysts with high thermostability, and the other is to design catalysts that
exhibit high activity for DRM at low temperature. The stability of ADM catalysts
greatly depends on the interaction between the metal and support. As some doped
supports were reported to be able to stabilize ADMs under high temperature,46,89,90
it is essential to get deeper insight into the effect of the specific modifiers that
strengthen the interaction between various metals and supports. The development
of ADMs with high stability is of great importance for the industrial application of
ADMs in DRM processes in the future.

ADM catalysts with two types of metal sites have been recently reported to be
active for low-temperature DRM,81,91,92 providing a new method for the design
and development of catalysts with high activity and selectivity at a relatively low
temperature. The synergetic catalysis of different metal sites requires further
exploration to guide the rational design of ADMs for low-temperature DRM. In
addition, computational methods, especially DFT calculation, have been widely
used to understand the mechanism, including thermodynamics and kinetics of
DRM catalyzed by ADMs. Mechanism of the ADM-catalyzed DRM reaction should
be further explored to establish structure-performance relationships for the
rational design of more efficient catalysts. Importantly, the approaches for the eco-
nomic preparation of ADMs at industrial scale need to be further developed in the
future.

10 Cell Reports Physical Science 4, 101310, March 15, 2023


ll
OPEN ACCESS
Review

Figure 8. Schematics of structures and mechanism for representative ADMs


(A and B) Schematics of structures for the (A) single-atom Ni 1 /MgO (100) catalyst and (B) single-site
Ni4 /MgO (100) catalyst. Colors indicate the following: green, Mg; red, O; and blue, Ni. Reprinted
with permission (Copyright 2018, American Chemical Society 87 ).
(C and D) (C) Schematic of the mechanism of carbon removal on O V and (D) the formation energy of
O V and adsorption energy of CO 2 on modified CeO 2 (110), where the E 0 represents the initial-state
energy of CeO 2 (110) modified by a second metal. Reprinted with permission (Copyright 2022,
Elsevier 88 ).

ACKNOWLEDGMENTS
This work was financially supported by National Key Research and Development Pro-
gram of China (2021YFA1500503), National Natural Science Foundation of China
(21972135, 21961142006), and CAS Project for Young Scientists in Basic Research
(YSBR-022).

AUTHOR CONTRIBUTIONS
Q.Z, M.A., and B.Q. proposed the outline of the review. Q.Z. and M.A. performed
literature survey. Q.Z. wrote the main part of the original draft. Y.Y. wrote the sum-
mary and conclusion of the manuscript. Y.Y. and B.Q. revised the manuscript. All au-
thors contributed to discussion and manuscript review.

DECLARATION OF INTERESTS
The authors declare no competing interests.

REFERENCES
1. Pakhare, D., and Spivey, J. (2014). A review of 3. He, M., Sun, Y., and Han, B. (2013). Green 5. Shi, L., Yang, G., Tao, K., Yoneyama, Y., Tan,
dry (CO2) reforming of methane over noble carbon science: scientific basis for integrating Y., and Tsubaki, N. (2013). An introduction of
metal catalysts. Chem. Soc. Rev. 43, carbon resource processing, utilization, and CO2 conversion by dry reforming with
7813–7837. recycling. Angew. Chem. Int. Ed. Engl. 52, methane and new route of low-temperature
9620–9633. methanol synthesis. Acc. Chem. Res. 46,
2. Sun, Z., Ma, T., Tao, H., Fa N, Q., and Han, 1838–1847.
B. (2017). Fundamentals and challenges 4. Saha, D., Grappe, H.A., Chakraborty, A., and
of electrochemical CO2 reduction Orkoulas, G. (2016). Postextraction separation, 6. Arora, S., and Prasad, R. (2016). An overview on
using two-dimensional Materials. Chem 3, on-board storage, and catalytic conversion of dry reforming of methane: strategies to reduce
560–587. methane in natural gas: a review. Chem. Rev. carbonaceous deactivation of catalysts. RSC
116, 11436–11499. Adv. 6, 108668–108688.

Cell Reports Physical Science 4, 101310, March 15, 2023 11


ll
OPEN ACCESS
Review

7. Akri, M., Zhao, S., Li, X., Zang, K., Lee, A.F., deactivation behaviour of Ni-containing reconstruction of Ni particles on SBA-15 by
Isaacs, M.A., Xi, W., Gangarajula, Y., Luo, J., catalysts. Int. J. Hydrog. Energy 37, 12281– thermal activation for dry reforming of
Ren, Y., et al. (2019). Atomically dispersed 12291. methane with excellent resistant to carbon
nickel as coke-resistant active sites for deposition. J. Energy Inst. 93, 2255–2263.
methane dry reforming. Nat. Commun. 10, 19. Xu, J., Zhou, W., Wang, J., Li, Z., and Ma, J.
5181–5190. (2009). Characterization and analysis of carbon 32. Song, Y., Ozdemir, E., Ramesh, S., Adishev, A.,
deposited during the dry reforming of Subramanian, S., Harale, A., Albuali, M.,
8. Akri, M., Pronier, S., Chafik, T., Achak, O., methane over Ni/La2O3/Al2O3 catalysts. Chin. Fadhel, B.A., Jamal, A., Moon, D., et al. (2020).
Granger, P., Simon, P., Trentesaux, M., and J. Catal. 30, 1076–1084. Dry reforming of methane by stable Ni-Mo
Batiot-Dupeyrat, C. (2017). Development of nanocatalysts on single-crystalline MgO.
nickel supported La and Ce-natural illite 20. Lavoie, J.M. (2014). Review on dry reforming of Science 367, 777–781.
clay for autothermal dry reforming of methane, a potentially more environmentally-
methane: toward a better resistance to friendly approach to the increasing natural gas 33. Hu, Y.H., and Ruckenstein, E. (2020). Comment
deactivation. Appl. Catal. B Environ. 205, exploitation. Front. Chem. 2, 81. on ‘‘Dry reforming of methane by stable Ni–Mo
519–531. nanocatalysts on single-crystalline MgO’’.
21. Aramouni, N.A.K., Touma, J.G., Tarboush, Science 368, eabb5459.
9. Akri, M., Chafik, T., Granger, P., Ayrault, P., and B.A., Zeaiter, J., and Ahmad, M.N. (2018).
Batiot-Dupeyrat, C. (2016). Novel nickel Catalyst design for dry reforming of methane: 34. Dai, C., Zhang, S., Zhang, A., Song, C., Shi, C.,
promoted illite clay based catalyst for analysis review. Renew. Sustain. Energy Rev. 82, and Guo, X. (2015). Hollow zeolite
autothermal dry reforming of methane. Fuel 2570–2585. encapsulated Ni–Pt bimetals for sintering and
178, 139–147. coking resistant dry reforming of methane.
22. Ayodele, B.V., and Cheng, C.K. (2015).
J. Mater. Chem. A 3, 16461–16468.
10. Thakur, R., VahidMohammadi, A., Smith, J., Modelling and optimization of syngas
Hoffman, M., Moncada, J., Beidaghi, M., and production from methane dry reforming over 35. Min, H.K., Kweon, S., Kim, Y.W., An, H., Jo, D.,
Carrero, C.A. (2020). Insights into the genesis of ceria-supported cobalt catalyst using artificial Park, E.D., Shin, C.-H., and Park, M.B. (2021).
a selective and coke-resistant MXene-Based neural networks and Box-Behnken design. Atomically dispersed nickel species in a two-
catalyst for the dry reforming of methane. ACS J. Ind. Eng. Chem. 32, 246–258. dimensional molecular sieve: origin of high
Catal. 10, 5124–5134. 23. Zhao, X., Li, H., Zhang, J., Shi, L., and Zhang, D.
activity and stability in dry reforming of
methane. Appl. Catal. B Environ. 298, 120627–
11. Zhang, F., Liu, Z., Chen, X., Rui, N., Betancourt, (2016). Design and synthesis of NiCe@m-SiO2
120636.
L.E., Lin, L., Xu, W., Sun, C.-j., Abeykoon, yolk-shell framework catalysts with improved
A.M.M., Rodriguez, J.A., et al. (2020). Effects of coke- and sintering-resistance in dry reforming 36. Jin, B., Li, S., and Liang, X. (2021). High-
Zr doping into ceria for the dry reforming of of methane. Int. J. Hydrogen Energy 41, performance catalytic four-channel hollow
methane over Ni/CeZrO2 catalysts: in situ 2447–2456. fibers with highly dispersed nickel
studies with XRD, XAFS, and AP-XPS. ACS 24. Tsyganok, A.I., Inaba, M., Tsunoda, T., Uchida, nanoparticles prepared by atomic layer
Catal. 10, 3274–3284. K., Suzuki, K., Takehira, K., and Hayakawa, T. deposition for dry reforming of methane. Ind.
(2005). Rational design of Mg–Al mixed oxide- Eng. Chem. Res. 61, 10377–10386.
12. Gallego, G.S., Batiot-Dupeyrat, C., Barrault, J.,
supported bimetallic catalysts for dry
Florez, E., and Mondragón, F. (2008). Dry 37. Wang, L., Liu, H., Zhuang, J., and Wang, D.
reforming of methane. Appl. Catal. A Gen. 292,
reforming of methane over LaNi1 yByO3Gd (2022). Small-scale big science: from nano- to
328–343.
(B=Mg, Co) perovskites used as catalyst atomically dispersed catalytic materials. Small
precursor. Appl. Catal. A Gen. 334, 251–258. 25. Liu, W., Zhang, L., Liu, X., Liu, X., Yang, X., Miao, Sci. 2, 2200036.
S., Wang, W., Wang, A., and Zhang, T. (2017).
13. Valderrama, G., Goldwasser, M.R., Navarro, 38. Chen, X., Jia, Z., Huang, F., Diao, J., and Liu, H.
Discriminating catalytically active FeNx species
C.U.d., Tatibouët, J.M., Barrault, J., Batiot- (2021). Atomically dispersed metal catalysts on
of atomically dispersed Fe-N-C catalyst for
Dupeyrat, C., and Martı́nez, F. (2005). Dry nanodiamond and its derivatives: synthesis and
selective oxidation of the C-H bond. J. Am.
reforming of methane over Ni perovskite type catalytic application. Chem. Commun. 57,
Chem. Soc. 139, 10790–10798.
oxides. Catal. Today 107-108, 785–791. 11591–11603.
26. Liu, J., Peng, H., Liu, W., Xu, X., Wang, X., Li, C.,
14. Akri, M., Achak, O., Granger, P., Wang, S., Zhou, W., Yuan, P., Chen, X., Zhang, W., and 39. Yang, X.F., Wang, A., Qiao, B., Li, J., Liu, J., and
Batiot-Dupeyrat, C., and Chafik, T. (2018). Zhan, H. (2014). Tin modification on Ni/Al2O3: Zhang, T. (2013). Single-atom catalysts: a new
Autothermal reforming of model purified designing potent coke-resistant catalysts for Frontier in heterogeneous catalysis. Acc.
biogas using an extruded honeycomb the dry reforming of methane. ChemCatChem Chem. Res. 46, 1740–1748.
monolith: a new catalyst based on nickel 6, 2095–2104.
incorporated illite clay promoted with MgO. 40. Wang, A., Li, J., and Zhang, T. (2018).
J. Clean. Prod. 171, 377–389. 27. Jang, W.J., Shim, J.O., Kim, H.M., Yoo, S.Y., Heterogeneous single-atom catalysis. Nat.
and Roh, H.S. (2019). A review on dry reforming Rev. Chem 2, 65–81.
15. Ferreira-Aparicio, P., Guerrero-Ruiz, A., and of methane in aspect of catalytic properties.
Rodrı́guez-Ramos, I. (1998). Comparative study 41. Qin, R., Liu, P., Fu, G., and Zheng, N. (2018).
Catal. Today 324, 15–26.
at low and medium reaction temperatures of Strategies for stabilizing atomically dispersed
syngas production by methane reforming with 28. Kim, J.H., Suh, D.J., Park, T.J., and Kim, K.L. metal catalysts. Small Methods 2, 1700286–
carbon dioxide over silica and alumina (2000). Effect of metal particle size on coking 1700306.
supported catalysts. Appl. Catal. A Gen. 170, during CO2 reforming of CH4 over Ni–alumina
177–187. aerogel catalysts. Appl. Catal. A Gen. 197, 42. Cui, X., Li, W., Ryabchuk, P., Junge, K., and
191–200. Beller, M. (2018). Bridging homogeneous and

16. Smoláková, L., Kout, M., Capek, L., Rodriguez- heterogeneous catalysis by heterogeneous
Gomez, A., Gonzalez-Delacruz, V.M., 29. Akri, M., El Kasmi, A., Batiot-Dupeyrat, C., and single-metal-site catalysts. Nat. Catal. 1,
Hromádko, L., and Caballero, A. (2016). Nickel Qiao, B. (2020). Highly active and carbon- 385–397.
catalyst with outstanding activity in the DRM resistant nickel single-atom catalysts for
reaction prepared by high temperature methane dry reforming. Catalysts 10, 630. 43. Xiong, Y., Sun, W., Han, Y., Xin, P., Zheng, X.,
calcination treatment. Int. J. Hydrog. Energy Yan, W., Dong, J., Zhang, J., Wang, D., and Li,
41, 8459–8469. 30. Lercher, J.A., Bitter, J.H., Hally, W., Niessen, Y. (2021). Cobalt single atom site catalysts with
W., and Seshan, K. (1996). Design of stable ultrahigh metal loading for enhanced aerobic
17. Muraleedharan Nair, M., and Kaliaguine, S. catalysts for methane-carbon dioxide oxidation of ethylbenzene. Nano Res. 14,
(2016). Structured catalysts for dry reforming of reforming. In Studies in Surface Science and 2418–2423.
methane. New J. Chem. 40, 4049–4060. Catalysis, J.W. Hightower, W. Nicholas
Delgass, E. Iglesia, and A.T. Bell, eds. (Elsevier), 44. Wu, X., Zhang, Q., Li, W., Qiao, B., Ma, D., and
18. De Sousa, H.S.A., Da Silva, A.N., Castro, A.J., pp. 463–472. Wang, S.L. (2021). Atomic-scale Pd on 2D
Campos, A., Filho, J.M., and Oliveira, A.C. titania sheets for selective oxidation of
(2012). Mesoporous catalysts for dry reforming 31. Chen, X., Yin, L., Long, K., Sun, H., Sun, M., methane to methanol. ACS Catal. 11, 14038–
of methane: correlation between structure and Wang, H., Zhang, Q., and Ning, P. (2020). The 14046.

12 Cell Reports Physical Science 4, 101310, March 15, 2023


ll
OPEN ACCESS
Review

45. Lang, R., Li, T., Matsumura, D., Miao, S., Ren, Y., 60. Qin, R., Liu, K., Wu, Q., and Zheng, N. (2020). 75. Michaelides, A., and Hu, P. (2000). Insight into
Cui, Y.T., Tan, Y., Qiao, B., Li, L., Wang, A., et al. Surface coordination chemistry of atomically microscopic reaction pathways in
(2016). Hydroformylation of olefins by a dispersed metal catalysts. Chem. Rev. 120, heterogeneous catalysis. J. Am. Chem. Soc.
rhodium single-atom catalyst with activity 11810–11899. 122, 9866–9867.
comparable to RhCl(PPh3)3. Angew. Chem. Int.
Ed. Engl. 55, 16054–16058. 61. Qiao, B., Liang, J.X., Wang, A., Xu, C.Q., Li, J., 76. Guo, X., Fang, G., Li, G., Ma, H., Fan, H., Yu, L.,
Zhang, T., and Liu, J.J. (2015). Ultrastable Ma, C., Wu, X., Deng, D., Wei, M., et al. (2014).
46. Flytzani-Stephanopoulos, M. (2014). Gold single-atom gold catalysts with strong covalent Direct, nonoxidative conversion of methane to
atoms stabilized on various supports catalyze metal-support interaction (CMSI). Nano Res. 8, ethylene, aromatics, and hydrogen. Science
the water–gas shift reaction. Acc. Chem. Res. 2913–2924. 344, 616–619.
47, 783–792.
62. Lang, R., Xi, W., Liu, J.C., Cui, Y.T., Li, T., Lee, 77. Xie, P., Pu, T., Nie, A., Hwang, S., Purdy, S.C.,
47. Zhu, C., Liang, J.X., Wang, Y.G., and Li, J. A.F., Chen, F., Chen, Y., Li, L., Li, L., et al. (2019). Yu, W., Su, D., Miller, J.T., and Wang, C. (2018).
(2022). Non-noble metal single-atom catalyst Non defect-stabilized thermally stable single- Nanoceria-supported single-atom platinum
with MXene support: Fe1/Ti2CO2 for CO atom catalyst. Nat. Commun. 10, 234–1271. catalysts for direct methane conversion. ACS
oxidation. Chinese J. Catal. 43, 1830–1841. Catal. 8, 4044–4048.
63. Jones, J., Xiong, H., DeLaRiva, A.T., Peterson,
48. Wang, Y., Arandiyan, H., Scott, J., Aguey- 78. Bai, S., Liu, F., Huang, B., Li, F., Lin, H., Wu, T.,
E.J., Pham, H., Challa, S.R., Qi, G., Oh, S.,
Zinsou, K.-F., and Amal, R. (2018). Single atom Sun, M., Wu, J., Shao, Q., Xu, Y., and Huang, X.
Wiebenga, M.H., Pereira Hernández, X.I., et al.
and nanoclustered Pt catalysts for selective (2020). High-efficiency direct methane
(2016). Thermally stable single-atom platinum-
CO2 reduction. ACS Appl. Energy Mater. 1, conversion to oxygenates on a cerium dioxide
on-ceria catalysts via atom trapping. Science
6781–6789. nanowires supported rhodium single-atom
353, 150–154.
catalyst. Nat. Commun. 11, 954–962.
49. Yang, H.B., Hung, S.-F., Liu, S., Yuan, K., Miao, 64. Bartholomew, C.H. (2001). Mechanisms of
S., Zhang, L., Huang, X., Wang, H.-Y., Cai, W., 79. Marcinkowski, M.D., Darby, M.T., Liu, J.,
catalyst deactivation. Appl. Catal. A Gen. Wimble, J.M., Lucci, F.R., Lee, S., Michaelides,
Chen, R., et al. (2018). Atomically dispersed 212, 17–60.
Ni(i) as the active site for electrochemical CO2 A., Flytzani-Stephanopoulos, M., Stamatakis,
reduction. Nat. Energy 3, 140–147. M., and Sykes, E.C.H. (2018). Pt/Cu single-
65. Martı́n, A.J., Mitchell, S., Mondelli, C., Jaydev, atom alloys as coke-resistant catalysts for
S., and Pérez-Ramı́rez, J. (2022). Unifying views efficient C-H activation. Nat. Chem. 10,
50. Tang, T., Wang, Z., and Guan, J. (2022). on catalyst deactivation. Nat. Catal. 5,
Electronic structure regulation of single-site 325–332.
854–866.
M-N-C electrocatalysts for carbon dioxide
80. Yarusov, I.B., Zatolokina, E.V., Shitova, N.V.,
reduction. Acta Phys. Chim. Sin. 39, 10–3866. 66. Mohan, O., Xu, R., and Mushrif, S.H. (2021). Belyi, A.S., and Ostrovskii, N.M. (1992).
Novel nickel-based single-atom alloy catalyst Propane dehydrogenation over Pt-Sn catalysts.
51. Huang, X., Ma, Y., and Zhi, L. (2020). Ultrathin for CO2 conversion reactions: computational
nitrogenated carbon nanosheets with single- Catal. Today 13, 655–658.
screening and reaction mechanism analysis.
atom nickel as an efficient fatalyst for J. Phys. Chem. C 125, 4041–4055. 81. Shen, D., Li, Z., Shan, J., Yu, G., Wang, X.,
electrochemical CO2 reduction. Acta Phys.
Zhang, Y., Liu, C., Lyu, S., Li, J., and Li, L. (2022).
Chim. Sin. 38, 2011050–2011058. 67. Wittich, K., Krämer, M., Bottke, N., and Schunk, Synergistic Pt-CeO2 interface boosting low
S.A. (2020). Catalytic dry reforming of methane: temperature dry reforming of methane. Appl.
52. Zhang, H., Tian, W., Duan, X., Sun, H., Huang,
Y., Fang, Y., and Wang, S. (2022). Single-atom
insights from model systems. ChemCatChem Catal. B Environ. 318, 121809–121819.
12, 2130–2147.
catalysts on metal-based supports for solar 82. Kim, S., Lauterbach, J., and Sasmaz, E. (2021).
photoreduction catalysis. Chinese J. Catal. 43, 68. Bian, Z., and Kawi, S. (2018). Sandwich-like Yolk-shell Pt-NiCe@SiO2 single-atom-alloy
2301–2315. silica@Ni@silica multicore–shell catalyst for the catalysts for low-temperature dry reforming of
low-temperature dry reforming of methane: methane. ACS Catal. 11, 8247–8260.
53. Guo, Y., Huang, Y., Zeng, B., Han, B., Akri, M.,
Shi, M., Zhao, Y., Li, Q., Su, Y., Li, L., et al. (2022). confinement effect against carbon formation.
ChemCatChem 10, 320–328. 83. Tang, Y., Wei, Y., Wang, Z., Zhang, S., Li, Y.,
Photo-thermo semi-hydrogenation of Nguyen, L., Li, Y., Zhou, Y., Shen, W., Tao,
acetylene on Pd1/TiO2 single-atom catalyst. F.F., and Hu, P. (2019). Synergy of single-
69. Wang, S., Lu, G.Q.M., and Millar, G.J. (1996).
Nat. Commun. 13, 2648–2659. atom Ni1 and Ru1 Sites on CeO2 for dry
Carbon dioxide reforming of methane to
produce synthesis gas over metal-supported reforming of CH4. J. Am. Chem. Soc. 141,
54. Zhang, H., Liu, G., Shi, L., and Ye, J. (2018).
catalysts: state of the art. Energy Fuels 10, 7283–7293.
Single-atom catalysts: emerging
multifunctional materials in heterogeneous 896–904.
84. Li, X., Phornphimon, M., Zhang, X., Deng, J.,
catalysis. Adv. Energy Mater. 8, 1701343– and Zhang, D. (2022). Promoting dry reforming
1701366. 70. Zhu, Y.A., Chen, D., Zhou, X.G., and Yuan, W.K.
(2009). DFT studies of dry reforming of of methane catalysed by atomically-dispersed
methane on Ni catalyst. Catal. Today 148, Ni over ceria-upgraded boron nitride. Chem.
55. Thomas, J.M., Raja, R., and Lewis, D.W. (2005). Asian J. 17, e202101428.
Single-site heterogeneous catalysts. Angew. 260–267.
Chem. Int. Ed. 44, 6456–6482. 85. Zhou, L., Martirez, J.M.P., Finzel, J., Zhang, C.,
71. Ferreira-Aparicio, P., Rodrı́guez-Ramos, I.,
Swearer, D.F., Tian, S., Robatjazi, H., Lou, M.,
56. Li, Z., Wang, D., Wu, Y., and Li, Y. (2018). Recent Anderson, J.A., and Guerrero-Ruiz, A. (2000).
Dong, L., Henderson, L., et al. (2020). Light-
advances in the precise control of isolated Mechanistic aspects of the dry reforming of
driven methane dry reforming with single
single-site catalysts by chemical methods. Natl. methane over ruthenium catalysts. Appl. Catal.
atomic site antenna-reactor plasmonic
Sci. Rev. 5, 673–689. A-Gen. 202, 183–196.
photocatalysts. Nat. Energy 5, 61–70.
57. Cai, X., Chen, X., Ying, Z., Wang, S., Chen, Y., 72. Zhang, C.J., and Hu, P. (2002). The possibility 86. Ji, L., Tang, S., Zeng, H.C., Lin, J., and Tan, K.L.
Cai, Y., Long, G., Liu, H., and Wang, N. (2021). of single C-H bond activation in CH4 on a (2001). CO2 reforming of methane to synthesis
Towards a library of atomically dispersed MoO3-supported Pt catalyst: a density gas over sol–gel-made Co/g-Al2O3 catalysts
catalysts. Mater. Des. 210, 110080–110086. functional theory study. J. Chem. Phys. 116, from organometallic precursors. Appl. Catal. A
4281–4285. Gen. 207, 247–255.
58. Flytzani-Stephanopoulos, M., and Gates, B.C.
(2012). Atomically dispersed supported metal 73. Michaelides, A., and Hu, P. (2000). A density 87. Zuo, Z., Liu, S., Wang, Z., Liu, C., Huang, W.,
catalysts. In Annual Review of Chemical and functional theory study of CH2 and H Huang, J., and Liu, P. (2018). Dry reforming of
Biomolecular Engineering, 3, J.M. Prausnitz, adsorption on Ni(111). J. Chem. Phys. 112, methane on single-site Ni/MgO catalysts:
ed., pp. 545–574. 6006–6014. importance of site confinement. ACS Catal. 8,
9821–9835.
59. Gates, B.C. (2019). Atomically dispersed 74. Michaelides, A., and Hu, P. (2000). A first
supported metal catalysts: seeing is believing. principles study of CH3 dehydrogenation on 88. Wu, J., Gao, J., Lian, S., Li, J., Sun, K., Zhao, S.,
Trends Chem. 1, 99–110. Ni(111). J. Chem. Phys. 112, 8120–8125. Kim, Y.D., Ren, Y., Zhang, M., Liu, Q., et al.

Cell Reports Physical Science 4, 101310, March 15, 2023 13


ll
OPEN ACCESS
Review

(2022). Engineering the oxygen vacancies 90. Peng, M., Jia, Z., Gao, Z., Xu, M., Cheng, D., dry reforming of methane at low
enables Ni single-atom catalyst for stable and Wang, M., Li, C., Wang, L., Cai, X., Jiang, Z., temperatures in an electric field. RSC Adv.
efficient C-H activation. Appl. Catal. B Environ. et al. (2022). Antisintering Pd1 catalyst for 12, 28359–28363.
314, 121516–121527. propane direct dehydrogenation with in situ
active sites regeneration ability. ACS Catal. 12, 92. Wu, J., Qiao, L.Y., Zhou, Z.F., Cui, G.J., Zong,
89. Liu, K., Zhao, X., Ren, G., Yang, T., Ren, Y., Lee, 2244–2252. S.S., Xu, D.J., Ye, R.P., Chen, R.P., Si, R., and
A.F., Su, Y., Pan, X., Zhang, J., Chen, Z., et al. Yao, Y.G. (2018). Revealing the synergistic
(2020). Strong metal-support interaction 91. Motomura, A., Nakaya, Y., Sampson, C., effects of Rh and substituted La2B2O7 (B = Zr
promoted scalable production of thermally Higo, T., Torimoto, M., Tsuneki, H., or Ti) for preserving the reactivity of catalyst in
stable single-atom catalysts. Nat. Commun. 11, Furukawa, S., and Sekine, Y. (2022). dry reforming of methane. ACS Catal. 9,
1263–1271. Synergistic effects of Ni-Fe alloy catalysts on 932–945.

14 Cell Reports Physical Science 4, 101310, March 15, 2023

You might also like