You are on page 1of 25

DOI: 10.1002/cnma.

202000407 Minireview

1
2
3 Single-atom Automobile Exhaust Catalysts
4
5 Yubing Lu,[a] Zihao Zhang,[a, b] Fan Lin,[a] Huamin Wang,*[a] and Yong Wang*[a, b]
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
ChemNanoMat 2020, 6, 1 – 25 1 © 2020 Wiley-VCH GmbH

These are not the final page numbers! ��


Minireview

1
Abstract: Single-atom catalysts (SACs) with isolated metal strategies for promoting the reactivity of metal SACs while
2
centers are attracting great research interest because of their maintaining their thermal stability during exhaust-abate-
3
maximized metal utilization rates and unique structures. In ment-related reactions. Highlights are provided on well-
4
the last decade, significant efforts have been made to design performed SACs for CO oxidation, unburnt hydrocarbon (HC)
5
highly efficient and cost-effective SACs with applications in oxidation, and the selective catalytic reduction of NOx. The
6
automobile exhaust aftertreatment. SACs have not only reaction mechanism and structure-performance relationship
7
demonstrated their high potential for improving the catalytic of SAC during these reactions are also discussed. Finally,
8
performance of current automobile exhaust catalysts but also perspectives are given on the trending research fields for
9
provided an ideal platform for understanding the origins of designing more efficient single-atom automobile exhaust
10
catalyst reactivity. In this review, we summarize the general catalysts in the future.
11
12
13
1. Introduction are discussed in Section 2.2. Below are the typical catalysts used
14
for IC engine exhaust emission abatement.
15
Pollutants emitted from automotive internal combustion (IC) To date, most catalytic systems in vehicles become effective at
16
engines have become a global concern because of their adverse ~ 200 °C, but their efficiency at lower cold-start temperatures is
17
impact on human health and the environment. Major automotive limited.[5] Therefore, a large fraction of pollutants are generated
18
emissions include oxides of nitrogen (NOx), carbon monoxide (CO), when the car is starting up at temperatures in the first 30 seconds
19
unburnt hydrocarbon (HC), carbon dioxide (CO2), and particulate when the catalyst is still being warmed up.[6] One major challenge
20
matter (PM).[1] To mitigate air pollutants from automotive exhaust for traditional exhaust emission control catalysts at low temper-
21
emissions, higher fuel economy standards and exhaust emission atures is that the strong adsorption of CO/HC may poison the
22
regulations were established by different countries.[2] In the United metal surface, preventing O2 activation and further reaction.[7]
23
States, with the onset of SULEV and TierIII emissions standards, Another challenge is that the high cost of noble metal catalysts
24
U.S. DRIVE has set the goal of achieving 90% conversion of vehicle (Pt, Rh, etc.) means they must be used efficiently; however,
25
emissions at a low temperature of 150 °C.[3] To meet this 150 °C traditional noble metal catalysts often suffer from low metal
26
challenge and fulfill the more stringent exhaust emission surface dispersion and sintering. Therefore, in the last decade,
27
regulations, more efficient and durable catalytic materials are significant efforts have been made to study extremely small
28
required.[4] Catalytic converters are required for removing the CO, (subnanometer) clusters and even isolated single atoms for
29
unburnt HCs, and NOx from the exhaust of IC engines. Different application in automobile emissions to enhance metal efficiency
30
types of catalysts are involved depending on the air-fuel ratio as and mitigate CO poisoning. Encouragingly, remarkable catalytic
31
well as the engine type. Figure 1 summarizes two typical performances have been reported for catalysts in the subnanom-
32
configurations of automobile aftertreatment systems: a) the diesel eter regime. For example, Pt subnanometer clusters were reported
33
engine aftertreatment system; and b) the three-way converter for to be highly active and resistant to sintering during CO
34
stoichiometric gasoline engines. A typical diesel engine aftertreat- oxidation.[8]
35
ment system (Figure 1a) consists of a diesel oxidation catalyst As the limitations on particle size diminished, single-atom
36
(DOC), a diesel particulate filter (DPF), a selective catalytic catalysts (SACs) not only maximized metal utilization but also
37
reduction catalyst (SCR), and a clean-up catalyst (CUC). The DOC is provided numerous opportunities to alternate the reaction path-
38
designed to convert CO and HC into CO2 and H2O, respectively. ways and therefore attracted increasing research attention in the
39
The DPF is a means of removing the PM or soot. SCR systems are field of exhaust emission abatement.[9] Unlike the supported
40
designed for converting NOx into N2. In a typical SCR system, urea nanoparticle catalysts with mostly metallic surface atoms,[10] the
41
is injected and mixed with the exhaust before entering the SCR supported metal single atoms are not only isolated from each
42
section. CUCs are used to remove the extra ammonia after SCR via other but are also highly tunable as a result of the strong
43
oxidation. The three-way catalyst (TWC) (Figure 1b) consists of a interaction between the metal center and the support.[11] For
44
reduction catalyst and a oxidation catalyst, and detailed reactions example, by adjusting the local environment and adsorption sites
45
of the TiO2-supported Pt single atoms, their oxidation states can
46
vary from Pt II to PtIV; the corresponding CO adsorption energies
47
range from amounts close to those occurring during physisorption
48 [a] Dr. Y. Lu, Dr. Z. Zhang, Dr. F. Lin, Dr. H. Wang, Prof. Y. Wang
Institute for Integrated Catalysis to as high as 295 kJ/mol.[12] In addition, because of the charge
49
Pacific Northwest National Laboratory transfer from or to the support and the capacity for adsorption of
50 Richland, Washington 99354 (USA) multiple ligands, the supported single atoms in many cases are no
51 E-mail: Huamin.Wang@pnnl.gov
Yong.Wang@pnnl.gov longer poisoned by CO.[6a,13] Furthermore, some strongly anchored
52
[b] Dr. Z. Zhang, Prof. Y. Wang metal single atoms also demonstrated their thermal abilities under
53
The Gene and Linda Voiland School of Chemical Engineering and harsh treatment and reaction conditions.[13c,14] As a result of their
54 Bioengineering
unique structures, different SACs have recently (in the past
55 Washington State University
Pullman, Washington 99164 (USA) decade) demonstrated their abilities for not only enhancing
56
This manuscript is part of a special collection on Single-Atom Catalysis. catalytic performance during exhaust emission abatement, but
57
ChemNanoMat 2020, 6, 1 – 25 www.chemnanomat.org 2 © 2020 Wiley-VCH GmbH

These are not the final page numbers! ��


Minireview
also for helping understand the fundamental reaction mecha- 2. Overview of automobile exhaust catalysts
1
nisms. However, a systematic and catalyst-based review about the
2
design and reaction mechanisms of single-atom automobile 2.1. Oxidation catalyst
3
exhaust catalysts is still lacking as of today. In this review, we
4
summarize the recent progress of different types of SACs for Oxidation catalysts are designed for the abatement of HC and CO
5
automobile aftertreatment-related reactions include CO oxidation, only from exhaust emission. Reactions (7) and (8) are the two
6
oxidation of unburnt hydrocarbons, and selective catalytic reduc- major reactions involved. DOCs are typical oxidation catalysts used
7
tion (SCR). We discuss the general strategies for enhancing the as the first treatment in the diesel aftertreatment system. DOCs
8
catalytic performance of the abovementioned reactions over oxidize HCs and CO, as well as NO into NO2, and combust soot
9
different types of SACs and illustrate their reaction mechanisms. into CO2 and H2O. Pt-group metal (PGM) catalysts have been
10
Finally, insights into the future direction of this field will be demonstrated as the most efficient catalysts for the oxidation of
11
provided. CO, HC, and NO.[15] In a typical DOC catalyst, a high-surface-area
12
washcoat is affixed to a ceramic monolith, and the PGM metal is
13
then dispersed onto the washcoat. Al2O3, SiO2, ZrO2, and zeolite
14
are the most commonly applied washcoats,[16] and Pt and Pd are
15
the most efficient noble metals currently available for DOC.[6b,15,16e]
16
17
18
19
20 Yubing Lu received a B.S. in Chemical Huamin Wang is currently a senior
Engineering from Tianjin University and research engineer in Pacific Northwest
21
a M.S. from Carnegie Mellon University. National Laboratory. He received his
22 In May 2019 he obtained his Ph.D. at Ph.D. from Nankai University, China,
23 the Chemical Engineering Department and then did his postdoctoral research
24 of Virginia Polytechnic Institute and in ETH Zurich and UC Berkeley. He has
25 State University in the group of Prof. experience in heterogeneous catalysis,
26 Ayman M. Karim. He is currently a inorganic material synthesis, hydropro-
postdoctoral fellow with Prof. Yong cessing, and biomass conversion. His
27
Wang and Dr. Huamin Wang at Pacific current research involves catalysis in
28 Northwest National Laboratory. His cur- thermochemical conversion of biomass
29 rent research interests are single-atom and fundamental understanding of cat-
30 catalysts with the applications in emis- alytic conversion of oxygenates.
31 sion control and biomass valorisation.
Yong Wang joined Pacific Northwest
32 Zihao Zhang completed his Ph.D. in National Laboratory (PNNL), USA, in
33 September 2019 at Zhejiang University 1994 and was promoted to Laboratory
34 under the guidance of Prof. Xiuyang Lu Fellow in 2005. In 2009, he assumed a
35 and Prof. Jie Fu. From January 2018 to joint position at Washington State
July 2019, he was a visiting Ph.D. University (WSU) and PNNL. In this
36
student in the group of Prof. Sheng Dai unique position, he continues to be a
37 at the University of Tennessee and Oak Laboratory Fellow at PNNL and is the
38 Ridge National Laboratory. He has been Voiland Distinguished Professor in
39 a postdoctoral fellow with Prof. Yong Chemical Engineering at WSU, a full
40 Wang and Dr. Abhijeet Karkamkar at professorship with tenure. His research
41 Washington State University and Pacific interests include the development of
Northwest National Laboratory since novel catalytic materials and reaction
42 October 2019. His current research engineering for the automobile emis-
43 interests are the development of hetero- sion control and conversion of fossil
44 geneous catalysts for biomass valor- and biomass feedstocks to fuels and
45 isation and emission control. chemicals.
46 Fan Lin received his Ph.D. degree in
47 Chemical Engineering and Applied
48 Chemistry in 2017 from University of
49 Toronto, Canada (Prof. Ya-Huei Chin’s
group). After that, he joined Pacific
50
Northwest National Laboratory as a
51 postdoctoral research associate in Jun
52 2017 and then transferred to a chemical
53 engineer in May 2020. His current
54 research interests include catalysis for
55 biomass conversion and automobile
emission control.
56
57
ChemNanoMat 2020, 6, 1 – 25 www.chemnanomat.org 3 © 2020 Wiley-VCH GmbH

These are not the final page numbers! ��


Minireview

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16 Figure 1. Typical configurations of the (a) diesel aftertreatment system and (b) three-way catalytic converter. (a) A typical diesel aftertreatment system consists
17 of a DOC, a DPF, an SCR, and a CUC. (b) A typical three-way catalytic converter consists of reduction catalyst and oxidation catalyst.
18
19
20
2.2. SCR catalyst close to stoichiometric conditions and is typically suitable for
21
gasoline engines. Under high air-fuel ratio conditions, the TWC
22
SCR is currently the most efficient strategy for removing NOx from acts as an oxidation catalyst and the NOx reduction capacity is
23
diesel engine exhaust emissions. Current SCR technologies can be very limited. Diesel engines always run at lean-burn conditions
24
categorized into urea/NH3 SCR,[17] hydrocarbon-SCR,[18] and plas- and are therefore not compatible with the TWC. Pd and Pt are
25
ma-assisted SCR.[19] Among them, urea/NH3 SCR is the most typically applied as the oxidation catalysts, while Pt and Rh are
26
widely commercialized technique for diesel engine aftertreatment. the most common reduction catalysts.[6b,24]
27
During the NH3 SCR process, a catalyst promotes Reactions (1)–(3) Reduction of NOx to nitrogen:
28
at lower temperatures than the thermal process:
29
2 CO þ 2 NO ! 2 CO2 þ N2 (4)
30
4 NO þ 4 NH3 þ O2 ! 4 N2 þ 6 H2 O (1)
31 hydrocarbon þ NO ! CO2 þ H2 O þ N2 (5)
32 6 NO2 þ 8 NH3 ! 7 N2 þ 12 H2 O (2)
33 2 H2 þ 2 NO ! 2 H2 O þ N2 (6)
34 2 NO þ 2 NO2 þ 4 NH3 ! 4 N2 þ 6 H2 O (3)
35
Oxidation of CO to CO2:
36
Reaction (1) is the standard NH3 SCR reaction, and Reaction
37
(3) is the so-called “fast” SCR reaction. Reaction (2) with pure 2 CO þ O2 ! 2 CO2 (7)
38
NO2 is even slower than Reactions (1) and (3). Side reactions
39
that yield NOx products also occur at certain temperatures and Oxidation of unburnt HCs to CO2 and H2O
40
O2 concentrations.[20] Typical SCR catalysts include vanadium-
41
based catalysts,[21] transition metal (e. g., Fe, Cu) ion-exchanged hydrocarbon þ O2 ! CO2 þ H2 O (8)
42
zeolites, and recently mixed metal oxides (e. g., CeO2 TiO2,
43
CeO2 WO3).[22] 3. SACs for CO oxidation
44
In exhaust emission abatement, it is not safe to use gas-
45
phase NH3 as the reductant, so NH3 is generated from solid CO oxidation is an important reaction for diesel oxidation and
46
reductants. Urea, ammonia salts (e. g., ammonia carbamate and three-way catalysts and is a classic probe reaction for hetero-
47
ammonia carbonate) and metal ammine chlorides (e. g., genous catalyst design.[25] Traditional PGM nanoparticle cata-
48
strontium ammine chloride) are the common sources of solid lysts such as Pt, Pd, and Ir require adjacent sites for CO and O2
49
NH3.[23] adsorption, and O2 activation is often poisoned by the strong
50
CO adsorption.[26] SACs, because of their isolated metal center,
51
provide new opportunities to alternate the CO/O2 adsorptions
52
2.3. TWC and reaction pathways of different metal catalysts. Here, we
53
present the recent progress and general strategies of CO
54
The TWC is used to eliminate NOx, CO, and HC at close to oxidation on SACs categorized by their metal centers.
55
stoichiometric conditions. A TWC participates in three functions
56
simultaneously, involving Reactions (4)–(8). The TWC requires
57
ChemNanoMat 2020, 6, 1 – 25 www.chemnanomat.org 4 © 2020 Wiley-VCH GmbH

These are not the final page numbers! ��


Minireview
3.1. Pt SACs
1
2
Pt SACs have been extensively studied for CO oxidation because
3
of their wide applications. Remarkable catalytic performances of
4
CO oxidation were reported on Pt SACs with various supports. For
5
example, as one of the earliest reported SACs for CO oxidation,
6
the Pt1/FeOx catalyst presents a higher specific activity than its
7
nanoparticle counterpart; the turnover frequencies (TOFs) at 27 °C
8
on Pt1/FeOx and Ptn/FeOx were measured at 0.136 and 0.0801 s 1,
9
respectively.[13b] In another study, the specific activity on Pt1/TiO2
10
was measured to be 4–6 times that of the corresponding nano-
11
particle catalysts.[27] However, there were also reports of catalysts
12
such as Pt/CeO2, Pt/θ-Al2O3, and Pt/H ZSM5 where single Pt atoms
13
were claimed to be similarly or less active than their nanoparticle
14
counterparts.[14a,28] Thus, to solve these discrepancies, it is essential
15
to understand the reaction mechanisms and origins of intrinsic
16
activity over SACs. Here, we highlight different thermally stable Pt
17
SACs which are relevant to automobile applications, with a focus
18
on illustrating different strategies for enhancing their CO oxidation
19
performances and summarizing their reaction mechanisms.
20
CeO2 is a common reducible support for anchoring Pt.[29]
21
Doping Pt onto a CeO2 surface could enhance the O2 storage and
22
release capacity of the surface and therefore promote CO
23
oxidation.[30] For Pt/CeO2 nanoparticle catalysts, the Pt CeO2
24
interface was identified as the active site, and the reaction
25
proceeds through a dual-site mechanism in which CO is adsorbed
26
on Pt and O is activated on CeO2.[31] During CO oxidation, a large
27
fraction of non-interface Pt atoms act only as CO reservoirs, and
28
the overall metal utilization rate is low. To address this issue,
29
thermally stable Pt/CeO2 SACs were prepared with an atomic-
30
trapping method in which a Pt atom was mobile at high
31 Figure 2. Illustration of the different strategies for promoting CO oxidation
temperatures before being trapped into the most stable sites on reactivities of Pt SACs without Pt aggregation. (a) Promoting the CO
32
the surface of the CeO2 nanorod.[14a] A density functional theory oxidation reactivity of Pt1/CeO2 by steam treatment at 750 °C or co-feeding
33 water during CO oxidation. Adapted with permission from ref. [13c].
(DFT)-based study indicated that Pt could be anchored on the Pt
34 Copyright 2017 American Association for the Advancement of Science. (b)
(100) surface, forming a square O4 nanopocket structure with high Promoting the CO oxidation reactivity of Pt1/CeO2 by increasing the Pt
35
binding energies (678 kJ/mol) so that the isolated Pt could be weight loading without forming Pt clusters or particles. Adapted with
36 permission from ref. [38]. Copyright 2019 American Chemical Society. (c)
protected from thermal-induced aggregation.[32] A combined
37 Promoting the CO oxidation reactivity of Pt1/TiO2 or Pt/CeO2 by creating a
spectroscopic and theoretical study further demonstrated that the CeOx TiO2 interface to anchor the isolated Pt atoms. Adapted with
38
single Pt atoms could be stabilized on a perturbed CeO2(100) permission from ref. [37]. Copyright 2020 Royal Society of Chemistry.
39
surface with four neighboring O atoms per Pt.[33] The Pt/CeO2 SACs
40
prepared with the atomic-trapping method are active for CO
41
oxidation even after aging under conditions that are harsh for a
42
typical DOC (800 °C in the air).[14a] However, the activity of the lattice O and therefore promoted CO oxidation reactivity. Besides,
43
abovementioned Pt1/CeO2 is still limited when compared with Pereira-Hernández et al. reported an approach to activate the
44
other Pt-group nanoparticle catalysts. Different strategies have thermally stable Pt1/CeO2 in CO at 275 °C, and the high reactivity
45
been applied to improve the low-temperature CO oxidation achieved at low temperatures is related to the improved
46
activity of Pt1/CeO2. One efficient strategy is to promote the reducibility of lattice oxygen on the CeO2 support.[34] Another
47
reducibility of surface lattice O via high-temperature steam strategy to promote CO oxidation on Pt1/CeO2 is to co-feed water.
48
treatment. Nie et al. synthesized a Pt1/CeO2 catalyst using a high- Wang et al. synthesized a Pt1/CeO2 catalyst with an atomic layer
49
temperature atomic-trapping method and further activated the deposition method.[35] The light-off experiment showed that co-
50
Pt1/CeO2 catalyst via steam treatment at 750 °C.[13c] The CO feeding water at 4 kPa significantly improved the CO oxidation
51
oxidation activity of the catalyst was significantly improved after activity and lowered the T50 from 150 °C to 90 °C. DFT calculations
52
the steam treatment, and the T50 of the steam-treated and non- and an isotope labeling experiment using H218O suggest that
53
steam-treated Pt1/CeO2 is 133 °C and 308 °C, respectively (Fig- water could participate in the reaction by dissociating into OH*
54
ure 2a). During high-temperature steam treatment, H was merged and H* at the Pt O Ce interface. The CO oxidation reactivity of Pt/
55
into the top layers of the CeO2 lattice and formed an active Olattice CeO2 with steam treatment can also be further promoted by co-
56
[H] structure. These Olattice[H] sites enhanced the reducibility of the feeding H2O (Figure 2a).[13c] Unlike the steam-treated Pt1/CeO2 with
57
ChemNanoMat 2020, 6, 1 – 25 www.chemnanomat.org 5 © 2020 Wiley-VCH GmbH

These are not the final page numbers! ��


Minireview
a stable Olattice[H] site, the water promotion effect discontinues and harsh reduction, the local environments of single Pt atoms are
1
after depleting water;[13c,35] this effect could potentially be proposed with PtTi6C, PtO2 and Pt(OH), respectively. The local
2
generalized to other SACs. A similar promotion effect by water environment can experience changes, and different environments
3
was reported on a Pt1/Cr1.3Fe0.7O3 catalyst.[36] Adjusting the surface lead to different CO adsorption strengths – for example, harsh
4
properties of CeO2 support by mixing CeO2 with different metal reduction results in a higher CO adsorption strength. The dynamic
5
oxides is also a powerful strategy for promoting the reactivity of evolution of local coordination also leads to different CO oxidation
6
Pt1/CeO2 during CO oxidation. An excellent illustration is a Pt1/ reactivities – harsh reduction, again for example, results in a higher
7
CeOx TiO2 catalyst reported by Yoo et al.[37] that was prepared by CO2 production rate.[12a]
8
adding 1% Ce into the TiO2 before introducing Pt. The CeOx TiO2 Efforts have also been made to synthesize nonreducible
9
interface was identified as the active site for the Pt1/CeOx TiO2 oxide supported or defect-free Pt SACs for CO oxidation
10
catalyst during CO oxidation, and the CeOx TiO2 interface because of their wide industrial applications. Zhang et al.
11
promoted the oxidation of CO via the Mars-van Krevelen (MvK) reported a thermally stable Pt SAC with mesoporous Al2O3 as
12
mechanism.[37] Pt1/CeOx TiO2 presented a similar mass activity to the substrate produced using a modified sol-gel solvent
13
that of steam-treated Pt1/CeO2 at 140 °C but a much higher activity evaporation method.[41] The synthesized Pt1/Al2O3 catalyst was
14
than that of Pt1/TiO2 (15.1 times at 140 °C) (Figure 2c). Overall, the highly stable during CO oxidation, and no deactivation occurred
15
key to promoting the reactivity of Pt1/CeO2 is to enhance the after 60 light-off cycles between 100 and 400 °C. X-ray
16
reducibility of the lattice O next to Pt without reducing the absorption fine structure (XAFS) and Al magic-angle spinning
17
thermal stability of the single Pt atoms. Higher metal loading is nuclear magnetic resonance (MAS NMR) results revealed that
18
also a direct strategy for enhancing the catalytic performance of the Pt was stabilized with the unsaturated pentahedral Al3 +
19
SACs and for practical application to emission control. However, centers via Pt O linkages. FeOx also demonstrated its remark-
20
the higher-loading catalysts often suffered more from metal able ability to anchor to Pt on defect-free sites as a result of the
21
aggregation. Kunwar et al. showed that CeO2(111) step edges covalent interactions between Pt and FeOx, and Pt1/FeOx has
22
could provide abundant sites for adsorption of Pt in a stable been shown to be active for CO oxidation.[13b,42]
23
square planer structure. With the high-temperature trapping The reaction mechanisms of CO oxidation over Pt SACs are
24
method, a high weight loading of 3% Pt/CeO2 SAC was prepared, highly dependent on the type of support, the local coordination
25
which significantly increased the reactivity of low-temperature CO environment of Pt, and the charge density of the Pt metal center
26
oxidation when compared with the 1% Pt/CeO2 SAC (Figure 2b). (Table 1). CO oxidation on reducible oxide-supported SACs that
27
Similar to Pt/CeO2, CO oxidation on a Pt/TiO2 nanoparticle have high O2 storage and release capacities tends to proceed via
28
catalyst also proceeds at the Pt TiO2 interface with CO adsorbed the MvK mechanism. Figure 3a illustrates a typical reaction cycle
29
on Pt and O2 adsorbed on TiO2.[39] By controlling the amount of Pt of CO oxidation on metal oxide-supported Pt SACs with the MvK
30
to ~ 1 Pt atom per anatase TiO2 particle, Pt can be anchored on mechanism. CO is first adsorbed onto Pt (I!II) and then reacts
31
TiO2 as isolated atoms and remain stable during reduction in H2 with an adjacent lattice O to produce CO2 with one surface O
32
(up to 450 °C) and CO oxidation.[27,40] Because of the charge vacancy left (II!III). This step is typically the rate-determining step
33
transfer from Pt to the TiO2 support, the isolated Pt atoms are (RDS). O2 is then adsorbed and activated on the surface O vacancy
34
positively charged and bind very weakly to CO.[27] The adsorbed (III!IV) and further reacts with a CO adsorbed on the Pt (IV!I).
35
CO can be facilely removed by flowing He at room temperature. Pt1/FeOx,[13b,43] Pt1/Ga CeO2 (Ga-doped CeO2),[44] steam-treated Pt1/
36
The TOF of the Pt/TiO2 SAC is 2-fold higher than that of its CeO2,[13c] and Pt1/TiO2[27] were reported to follow the MvK
37
nanoparticle counterparts (average size ~ 1 nm), though similar mechanism. To proceed, the MvK mechanism requires a catalyst
38
reaction orders of CO and O2 were measured on these two with high O2 storage and release properties. Ab initio calculations
39
catalysts.[27] Further study showed that the CO adsorption and showed that a CO oxidation reaction on a CeO2 (111) surface with
40
oxidation performance of Pt1/TiO2 can be fine-tuned by adjusting supported Pt single atoms must overcome a higher energy barrier
41
the local environment of Pt.[12a] During oxidation, mild reduction, than that of the Langmuir-Hinshelwood (LH) pathway to proceed
42
43
44 Table 1. Summary of the structure, adsorption properties, and preferred CO oxidation mechanisms of Pt SACs.
45
Structure CO/O2 adsorption Reaction pathway
46 Catalyst Adsorption site of Pt Q [j e j] Eads_CO [eV] Eads_O2 [eV] υCO [cm 1] Mechanism Rate-limiting step Ref.
47
Pt1/FeOx Replace Fe + 0.33 1.96 1.05 2080 MvK COads + Olattice [13b, 43a]
48 Pt1/TiO2 (PtO2)ads step* + 1.32 0.52 – 2112 MvK COads + Olattice [12, 27]
49 Pt1/CeO2 CeO2(111) + 0.08 3.26 1.42 2095 LH CO2 desorption [44]
50 Pt1/Ga CeO2 Ga CeO2(111) Ov + 0.24 2.22 1.13 2096 MvK COads + Olattice [44]
Pt1/CeO2_S* CeO2(111) – 0.58 0.08 2096 MvK COads + Olattice[H] [13c]
51 Pt1/γ-Al2O3 Al-top (Al-terminated) 0.08 1.27 – – LH COads + O2_ads [46]
52 Pt1/MoS2 3-fold hollow site + 0.04 2.39 0.77 – LH COads + O2_ads [49]
53 Pt1/TiC TiC(100) Ti-defect + 0.39 0.99 0.87 – LH + ER COads + O2_ads [47]
Pt1-N-C N2C + 0.26 1.57 1.31 – TER 2COads + O2(g) [51a]
54 Pt1/graphene C=C site + 0.44 2.05 0.97 – TER 2COads + O2(g) [51b]
55
* Pt/TiO2 SAC with (PtO2)ads step site as an illustration of the Pt1/TiO2 after mild reduction.
56 * S denotes the high-temperature steam treatment.
57
ChemNanoMat 2020, 6, 1 – 25 www.chemnanomat.org 6 © 2020 Wiley-VCH GmbH

These are not the final page numbers! ��


Minireview

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16 Figure 3. Typical reaction pathways of Pt SACs via the MvK (a) or the LH (b) mechanism.[13b,28a,44,46,49]
17
18
19
via the MvK pathway; therefore, the LH pathway is preferred.[13c,44] supported SACs and the molecular oxygen activation has also
20
The high reaction barrier of the MvK pathway on Pt/CeO2(111) is been observed on these cluster supported SACs.[50]
21
due to the high energy required for O vacancy formation, which is A new tri-molecular Eley-Rideal (TER) mechanism was also
22
as high as 2.76 eV.[44] Different strategies have been reported to identified to be favored on SACs with carbon-based supports such
23
lower the O vacancy formation energy enough to favor the MvK as Pt1 N C and Pt1/graphene (Figure 4).[51] The TER mechanism
24
pathway. One strategy is to apply dopant to adjust the redox starts with a gem-dicarbonyl structure because of the strong CO
25
properties of the Pt SACs. For example, doping CeO2(111) with Ga adsorption energy. The O2 then moves toward and then bridges
26
significantly improves its lattice O activation capabilities, and the O between the two adsorbed CO molecules to form an OC OO CO
27
vacancy formation energy on Pt1/Ga CeO2(111) is lowered to only structure (Figure 4, II). The OC OO CO further dissociates into two
28
0.09 eV.[44] The MvK pathway becomes the most favorable path- CO2 molecules that are then desorbed from Pt. For Pt1 N C, the
29
way after adding the Ga dopant, with a rate-limiting step that has reaction barriers of OC OO CO formation and dissociation are
30
a calculated activation barrier of 0.51 eV.[44] Another strategy to 0.06 eV (TS I) and 0.16 eV (TS-II), respectively, compared to the
31
promote the reducibility of the lattice O is to increase the density higher reaction barriers of 0.59 eV and 1.81 eV for the LH and
32
of surface hydroxyl or Olattice[H] via water promotion or steam regular Eley-Rideal (ER) pathways, respectively.[51a] Similar behaviors
33
treatment, as discussed previously.[13c,35] The LH mechanism is were also calculated for Pt1/graphene, where the TER mechanism
34
another common reaction mechanism of CO oxidation. On Pt is preferred.[51b] Although the TER pathway proved to be a highly
35
nanoparticles and single crystals, CO and O2 must be adsorbed by promising pathway for carbon materials supported on Pt SACs,
36
adjacent sites before the reaction can proceed.[25b,26,45] Oxygen experimental evidence is still lacking and waiting to be explored. It
37
activation also requires two neighboring sites; therefore, the metal is worth noting that because the reaction cycle involves the
38
surface is often poisoned by CO.[25b] Unlike how it occurs on Pt formation of two CO2 molecules in one reaction cycle, CO
39
nanoparticles and single crystals, the LH pathway on Pt SACs oxidation on Pt SACs could also involve two different mechanisms.
40
proceeds with CO and O2 adsorbed and activated on the same Pt DFT calculations indicate that Pt1/TiC could initially employ the LH
41
atom. As shown in Figure 3b, after CO and O2 adsorption on the mechanism to form the OCOO intermediate, with the reaction
42
Pt atom (I!III), the reaction can proceed through two possible between CO and the adsorbed O* preferentially proceeding via
43
intermediates: the OCOO intermediate (TS-1 type 1) or the the TER mechanism (Table 1).[47]
44
carbonate intermediate (TS-1 type 2: CO3). The OCOO intermediate A prerequisite for understanding the reaction mechanism of
45
was reported to be the favored on Pt1/γ-Al2O3,[46] Pt1/CeO2(111),[44] Pt SACs is to identify the adsorption site, considering that
46
and Pt1/TiC,[47] while the carbonate intermediate was preferred on different sites may contribute differently during CO oxidation.
47
Pt/θ-Al2O3.[28a] Similar COOO[25b] or carbonate[48] intermediate was The importance of structural uniformity was demonstrated with
48
also identified on supported Pt nanoparticles, however, unlike the Pt1/CeO2.[14c] Low-loading Pt1/CeO2 consists of uniform single Pt
49
Pt single atoms, adjacent sites are always required on Pt atoms anchored on step sites with stable square planar
50
nanoparticles for CO and O2 to adsorb.
51
Currently, DFT-based quantum chemical calculations, X-ray
52
absorption spectroscopy and mass spectroscopy are major
53
techniques to identify the intermediate species during CO
54
oxidation. Identification of the reaction intermediate with the
55
combination of DFT calculation and mass-spectroscopy have
56 Figure 4. Reaction pathway of Pt SACs via the TER mechanism.[51]
been demonstrated on well-controlled metal oxide cluster
57
ChemNanoMat 2020, 6, 1 – 25 www.chemnanomat.org 7 © 2020 Wiley-VCH GmbH

These are not the final page numbers! ��


Minireview
structures. These Pt are stable even after harsh H2 reduction at which makes the identification of Rh single atoms with CO
1
450 °C. However, higher-loading Pt1/CeO2 also consists of Pt chemisorption via infrared spectroscopy an occasionally difficult
2
anchored on other sites with weaker interactions between the task.[57] Other techniques such as high-angle annular dark-field
3
metal and the support. The Pt on those sites agglomerates into scanning transmission electron microscopy (HAADF-STEM) and
4
particles even at a mild reduction temperature of 50 °C. Thus, XAFS are often required to illustrate the surface structure of Rh-
5
the thermal stability of Pt SACs is directly related to the metal- based catalysts.[56c,d]
6
support interactions as well as the metal-CO interactions of the Because of their unique adsorption behaviors, Rh subnanom-
7
Pt anchoring site. A uniform active site is essential for not only eter clusters and single atoms have received recently increasing
8
enhancing the thermal stability of SACs but also understanding research attention. Guan et al. reported a TiO2-supported Rh
9
their origins of the activity, and microscopic and spectroscopic subnanometer cluster catalyst with remarkable catalytic perform-
10
studies have been demonstrated to be powerful tools for ance for CO oxidation at cryogenic temperatures.[58] The average
11
identifying the active anchoring sites of Pt SACs.[52] particle size of the Rh/TiO2 catalyst was measured in transmission
12
electron microscopy (TEM) as 0.62 � 0.16 nm; during CO oxidation,
13
97% conversion was reached at 40 °C with 1% CO and 5% O2 at
14
3.2. Rh and Ir SACs a weight hourly space velocity (WHSV) of 720 L gRh 1 h 1.[58] The Rh/
15
TiO2 catalysts also demonstrate significant size dependence in the
16
Rh-based catalysts are some of the most commonly used catalysts < 2 nm regime; catalysts with particle sizes of 0.4–0.8 nm are
17
for the removal of CO from exhaust streams. Although CO about 5-fold as active as the Rh/TiO2 SAC/pseudo-SACs and one
18
oxidation on Rh single crystals was revealed to be structurally order of magnitude more active than Rh/TiO2 catalysts with
19
insensitive,[53] a significant size dependence has been observed on average particle sizes of ~ 1.8 nm. The high activity on Rh/TiO2
20
Rh nanoparticle catalysts in which the smaller-sized particles and subnanometer cluster catalysts is due to facile O2 activation at the
21
the defected surfaces are more active.[54] Somorjai and co-workers Rh TiO2 interface.[58] Efforts have also been made to engineer the
22
illustrated that the oxide layers of RhOx are responsible for the activity of Rh SACs. Hülsey et al. investigated the CO oxidation
23
high CO oxidation activity of the Rh nanoparticles while the inner reaction mechanism and the dynamic structure of an ammonia
24
metallic Rh0 atoms are inert.[55] The Rh catalyst with a smaller salt of phosphotungstic acid (NPTA, (NH4)xPW12O40) supported Rh
25
particle size consists of a higher RhOx/Rh0 ratio and therefore a SAC with the assistance of operando spectroscopies.[11a] In situ and
26
higher specific activity during CO oxidation.[55] CO oxidation of Rh operando XAFS, XPS, and diffuse reflectance infrared Fourier
27
catalysts is highly related to the configuration of CO and O2 transform spectroscopy (DRIFTS) results were combined to image
28
adsorption. CO adsorption on Rh includes gem-dicarbonyl, linear the CO oxidation cycle, and the MvK mechanism was identified on
29
CO, and bridging CO, which can be denoted as Rh(CO)2, Rh(CO), Rh1/NPTA (Figure 5a). Unlike the Pt or Au SACs, CO oxidation on
30
and Rh2(CO), respectively. Significant efforts have been made by Rh1/NPTA starts with a gem-dicarbonyl (Rhδ + (CO)2) structure. One
31
Gates and co-workers to investigate the adsorption of CO and CO of Rhδ + (CO)2 first reacts with a lattice O on the NPTA support
32
other adsorbates on supported Rh clusters with different numbers to produce one CO2, which also creates an O vacancy cation. In
33
of atoms as well as isolated Rh SACs.[56] Generally, the fraction of the subsequent step, the O vacancy is filled by a gas-phase O2,
34
Rh2(CO) and Rh(CO) among all three species increases with particle and one CO is facilely adsorbed on Rh to restore the Rhδ + (CO)2.
35
size.[68] However, spectroscopic studies showed that well-defined The adsorbed O2 on the O vacancy is then activated at the
36
Rh(CO)2 can be observed with either CO adsorption on isolated Rh Rh NPTA interface and reacts with a CO from Rhδ+ (CO)2 to
37
single atoms or CO adsorbed on the edge atoms of small clusters, produce another CO2.[11a] Systematic mechanistic study of CO
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55 Figure 5. Illustration of CO oxidation reaction pathways on representative Rh and Ir SACs identified with in situ and operando spectroscopies and DFT-based
calculations. (a) Rh1/NPTA. Adapted with permission from ref. [11a]. Copyright 2019 Springer Nature Limited. (b) Ir1/MgAl2O4. Adapted with permission from
56 ref. [13a]. Copyright 2019 Springer Nature Limited.
57
ChemNanoMat 2020, 6, 1 – 25 www.chemnanomat.org 8 © 2020 Wiley-VCH GmbH

These are not the final page numbers! ��


Minireview
oxidation was also conducted on Rh/γ-Al2O3 using DFT-based the lower oxidation potential is more active during CO
1
calculations.[59] The LH pathway was identified as the most oxidation.[11a]
2
favorable pathway. Interestingly, the reaction pathways are Similar to Pt, the metal surface of supported Ir nanoparticles
3
dependent on CO partial pressure. At higher CO partial pressures, and single crystals is poisoned by CO, and the reaction is limited
4
the reaction starts with two CO and one O2 that are co-adsorbed by O2 activation.[26] Therefore, alternating the reaction pathway
5
on one Rh single atom. At lower CO partial pressures, the reaction and preventing CO poisoning is of great importance for supported
6
starts with 1 CO and 1 O2 that are co-adsorbed. For both cases, O2 Ir catalysts. Ir SACs have similar gem-dicarbonyl structures to Rh
7
was activated by bridging between the Rh and the neighboring SACs, which steers their electronic properties away from those of
8
Al.[59] This indicates that the reaction pathway of Rh SACs is their nanoparticle counterparts.[11b,62] Inspired by the high activity
9
dynamic and depends on the reaction conditions; therefore, it is of Pt1/FeOx catalysts,[13b] Liang et al. performed theoretical and
10
important to consider the reaction conditions when comparing experimental studies of catalytic CO oxidation on Ir1/FeOx.[63]
11
different Rh SACs and possibly other metal SACs. Reactivity measurement results showed that Ir1/FeOx is much less
12
Remarkable catalytic performance of CO oxidation over Rh active than Pt1/FeOx, and the TOF of Ir1/FeOx and Pt1/FeOx at 80 °C
13
SACs with various supports was recently reported. A ZnO-nano- was measured as 63 h 1 and 594 h 1, respectively. DFT-based
14
wire-supported Rh SAC (Rh1/ZnO-nw) was observed to be highly calculations indicate that both catalysts proceed through the MvK
15
active for CO oxidation. As shown in Figure 6a, Rh1/ZnO-nw mechanism and that the reaction is limited by CO2 produced by
16
presents the highest CO oxidation performance among three the reaction of an Ir-adsorbed CO and a nearby lattice O. The
17
different ZnO-supported SACs with a T50 of ~ 200 °C, while Pt1/ reaction barrier of this step on Ir1/FeOx (1.41 eV) is much higher
18
ZnO-nw and Au1/ZnO-nw are much less active with T50 values of than that on Pt1/FeOx (0.79 eV), which could partially explain the
19
298 °C and 328 °C, respectively. DFT calculations suggest that CO lower activity of the former. The Ir1/FeOx catalyst was also
20
oxidation on M1/ZnO-nw (M denotes Rh, Pt, or Au) follows the demonstrated to be less active than the FeOx-supported Ir
21
MvK mechanism and that the high activity of Rh1/ZnO could be subnanometer clusters (Figure 7).[64] A higher-loading 2.4% Ir/FeOx
22
attributed to its lower reaction barrier for O2 dissociation at the O2 catalyst consists of single atoms only after synthesis. After 100 °C
23
vacancy site. In another study, a Rh1/CeO2 catalyst also demon- H2 reduction, most Ir is maintained as single atoms – a temper-
24
strated superior reactivity for low-temperature CO oxidation.[60] ature-programmed reduction experiment confirmed the reduction
25
Rh1/CeO2 is slightly more active than the commercial TWC, and its of Ir. After H2 reduction at 200 °C, most of the Ir single atoms
26
T50 is ~ 124 °C (Figure 6b). Rh1/CeO2 is also more active than its aggregate into subnanometer clusters with an average size of
27
corresponding nanoparticles (T50: ~ 187 °C) and much more active about 0.74 nm. A lower-loading 0.22% Ir/FeOx catalyst showed
28
than Pd1/CeO2 (T50: ~ 290 °C) and Pt1/CeO2 (T50: ~ 348 °C). The better thermal stability and remained in the form of Ir single
29
reaction orders of CO and O2 on Rh1/CeO2 are measured as 0.2 atoms after H2 reduction at 200 °C. With the same H2 reduction
30
and 0.03, respectively, which indicates that the reaction follows temperature, the Ir subnanometer cluster catalyst (2.4% IrSubnano/
31
the MvK mechanism.[60] For the same type of Rh SAC, the CO FeOx) is much more active than the Ir SAC (0.22% Ir1/FeOx), and
32
oxidation performance could also be altered by the oxidation and TOFs at 80 °C are 270 h 1 and 65 h 1, respectively (Figure 7).
33
reduction potentials of the support. For example, Rh SACs were However, the activity of the Ir1/FeOx was also affected by H2
34
prepared using various heteropoly acid molecules (HPAs) with reduction. As a result of, the TOF of the 0.22% Ir1/FeOx (64.8 h 1)
35
various oxidation potentials.[11a] Catalytic studies showed that the after 200 °C reduction is the 20-fold equivalent of the TOF of the
36
CO oxidation T20 values of Rh SACs are well correlated to the 2.4% Ir1/FeOx after 100 °C reduction, although both catalysts
37
oxidation potentials of HPA supports for which the catalyst with consist of mostly SAC and the Pd atoms are partially reduced
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54 Figure 6. Catalytic performance of CO oxidation over different ZnO (a) or CeO2 (b) supported SACs and their reference catalysts. (a) Catalytic performance of
55 CO oxidation over ZnO-nanowire-supported Rh, Pt, or Au SACs with ZnO nanowire as reference. Measured with 1% CO and 1% O2. Adapted with permission
from ref. [61]. Copyright 2019 Elsevier B.V. (b) Catalytic performance of CO oxidation over CeO2-supported Rh, Pd, or Pt SACs, CeO2-supported nanoparticle
56 catalysts, commercial TWCs, and CeO2. Measured with 1% CO and 1% O2. Adapted with permission from ref. [60]. Copyright 2020 Royal Society of Chemistry.
57
ChemNanoMat 2020, 6, 1 – 25 www.chemnanomat.org 9 © 2020 Wiley-VCH GmbH

These are not the final page numbers! ��


Minireview
the opposite is true for Ir nanoparticles, where a higher CO partial
1
pressure will further poison the catalyst.[13a,26,65] In addition, the
2
reaction mechanism of the carbon-based materials supported Ir
3
SACs is different from that of other catalysts. DFT-based studies
4
showed that Ir1/graphdiyne preferentially follows the TER mecha-
5
nism in which O2 bridges with two adsorbed CO molecules,
6
forming two CO2 molecules.[66]
7
In summary, the catalytic performance and reaction mecha-
8
nisms of Rh/Ir SACs are highly affected by their supports. MvK,
9
ER, and TER mechanisms were reported on different supports.
10
Because Rh/Ir SACs can bind with more than one CO/O, their
11
reaction mechanisms are also unique from others. Facilitating
12
O2 activation and lattice O* activation are some general
13
strategies to enhance the catalytic performance of Rh/Ir SACs.
14
Figure 7. Catalytic performance of FeOx-supported Ir/Pt SACs and subnan- CO spectator species could be involved during the entire
15
ometer clusters at 80 °C. The H2 reduction temperature of each catalyst is reaction cycle for Rh/Ir SACs. Therefore, it is essential to identify
16 shown on the right y-axis.[13b,63]
both active and resting species for a better understanding of
17
the reaction mechanism.
18
19
(Figure 7). These results suggest that the reduction temperature
20
could affect the local environment of the Ir SACs and therefore 3.3. Pd SACs
21
affect their CO oxidation performance, similar to the case of Pt/
22
TiO2.[12a] Pd nanoparticle catalysts have been well studied for CO
23
Although Ir/FeOx SACs demonstrated inferior reactivities when oxidation. Weakly adsorbed linear and bridge-bonded CO
24
compared with their nanoparticle counterparts, other studies molecules were identified on nanoparticle-based Pd/SiO2,
25
suggest that the structures of the Ir SACs are dynamic during CO Pd Al2O3, and Pd/TiO2.[67] The adsorption of CO and O2 was
26
oxidation and that their reactivity can be enhanced by optimizing mitigated by the steric effect and the epitaxy of the support.[67]
27
the catalyst structure and reaction conditions. Gates and co- Steady-state kinetic experiments revealed that the reaction
28
workers showed that the support acts as a ligand to the Ir SACs mechanism and CO oxidation performance of Pd nanoparticle
29
and has a strong impact on the electronic properties of the metal catalysts are dependent on supports.[68] As a result of the weak
30
center.[11b] For an Ir1/MgO catalyst, for example, charge transfer and comparable adsorption of CO and O2, the Pd particles in Pt/
31
from MgO to Ir single atoms makes the Ir electron rich; however, SiO2, Pd Al2O3 and Pd/TiO2 were not blocked by CO.[68]
32
for an Ir1/zeolite, charge transfer from Ir single atoms to the zeolite However, it was observed that CO adsorption on Pd-based
33
causes the Ir to be electron deficient.[11b] Lu et al. identified the catalysts was strongly size dependent, especially in the
34
active structure of Ir1/MgAl2O4 during CO oxidation with in situ and subnanometer regime. For example, DFT-based calculations
35
operando spectroscopies and ab initio calculations.[13a] Although Ir also showed that Pd7/TiO2 subnanometer cluster catalysts
36
single atoms could bind with one CO very strongly (Eads: adsorb CO strongly, with CO adsorption energies between
37
1.98 eV), the reaction could proceed through a ER mechanism 2.15 and 2.27 eV, which is much higher (more negative)
38
between a second gas-phase CO and an activated O* (reaction than those that occur on nanoparticles.[69] Furthermore, Kaden
39
barrier 1.10 eV). Comparatively, the LH pathway requires a much et al. studied the size dependence of CO oxidation over Pd
40
higher energy (reaction barrier 2.12 and 1.64 eV for TSII and TSIII deposited on TiO2(110) in the size range from Pt1 to Pd25.[70] As
41
in Figure 5b, respectively) as a result of the strong CO adsorption shown in Figure 8, the CO oxidation reactivity of TiO2(110)-
42
(Figure 5b). Also, the Ir MgAl2O4 interface facilitates O2 activation, supported Pd generally increases from Pd1 to Pd20 before
43
and the O2 can be facilely activated with almost no energy barrier. decreasing from Pd20 to Pd25. XPS studies illustrated that the CO
44
What is more important is that during the entire reaction cycle oxidation reactivity of TiO2(110)-supported Pd SACs is directly
45
(Figure 5b), there is always one CO strongly bonded to Ir that correlated to the Pd 3d binding energy. The CO oxidation
46
functions as a spectator species. However, this strongly bonded reactivity is well correlated to the deviation from smooth charge
47
CO facilities O2 activation and lowers the reaction barrier for a scaling, which is an indicator of core-hole screening (Figure 8).
48
second CO to react with the O*. A general implication from this An explanation of the low reactivity of Pd1/TiO2 was given by
49
study is that because Ir (also Rh, Cu etc.) SACs can bind with more another theoretical study in which CO oxidation proceeded on
50
than one adsorbate per metal atom, their reaction mechanisms Pd1/TiO2(100) through the LH mechanism with a relatively high
51
can be dramatically different from those of their nanoparticle energy barrier of 1.13 eV.[71]
52
counterparts; it is important to identify their active states and Stabilizing Pd as isolated atoms is more challenging than
53
resting states for a better optimization of their CO oxidation stabilizing Pt or Rh because of the weak O2 affinity of Pd, and
54
performance.[13a,65] Because the ER mechanism dominates in the the high reducibility also makes Pd SACs unique among all
55
case of Ir/MgAl2O4, the reaction rate is positively dependent on SACs. Peterson et al. stabilized Pd single atoms on γ-Al2O3 by
56
the CO partial pressure but is insensitive to the O2 partial pressure; doping La and studied its CO oxidation performance.[6a] Atomi-
57
ChemNanoMat 2020, 6, 1 – 25 www.chemnanomat.org 10 © 2020 Wiley-VCH GmbH

These are not the final page numbers! ��


Minireview
improved thermal stability after doping metal oxides such as
1
CeO2, La2O3, MnOx and CoOx.[77] Since the active Au nanoparticle
2
catalysts reported by Haruta et al., supported Au catalysts have
3
been widely studied for low-temperature CO oxidation.[78]
4
Although bulk Au is considered inert, highly dispersed Au
5
particles are active for low-temperature CO oxidation.[78–79]
6
Besides particle size, interactions between the metal and the
7
support also contribute to the CO oxidation activity.[80] For
8
example, TiO2- and Al2O3-supported Au nanoparticles were
9
reported to be more active than ZnO- and ZrO2-supported Au
10
nanoparticles with similar sizes.[80] Although significant efforts
11
have been made, the origins of the activity of Au-based
12
catalysts during CO oxidation are still not clear. Because of their
13
highly uniform metal centers, SACs become an ideal platform
14 Figure 8. Correlation between the CO oxidation reactivity of TiO2(110)- for investigating the intrinsic activity of Au-based catalysts.
15 supported Pd (from Pd1 to Pd25) and the Pd 3d binding energy (relative to
expectations from smooth bulk scaling). Reproduced with permission from The reactivities of Au SACs are strongly affected by their
16
ref. [70]. Copyright 2009 American Association for the Advancement of supports and by the charge densities of Au atoms. Single atoms
17 Science. supported on rutile TiO2(110) are almost inactive for CO
18
oxidation, but catalytic activity increases from Au1 (Au1 denotes
19
Au present as single atoms) to Au7.[79e,81] However, Au SACs
20
cally dispersed Pd catalysts (0.5 wt.%) were prepared on both γ- supported by Co3O4,[82] FeOx,[83] and CeO2[84] present remarkable
21
Al2O3 and La-γ-Al2O3 using high-temperature calcination in air. catalytic performance for low-temperature CO oxidation. The
22
Pd1/La-γ-Al2O3 showed better thermal stability and reactivity CO oxidation activity of rutile TiO2(110)-supported single Au
23
than Pd1/Al2O3 during CO oxidation, although both catalysts atoms is limited by their weak O2 adsorption capabilities, but
24
were deactivated as a result of Pd aggregation, especially at the O2 accommodation capability is relatively higher on Au7/
25
higher temperatures. Another advantage of Pd1/La-γ-Al2O3 is TiO2.[85] For Au1/FeOx, O2 was adsorbed on the O vacancy next to
26
that the deactivated catalyst can be regenerated into a Pd SAC the single Au atom, which facilitates O2 activation; the metal-
27
after high-temperature calcination, which will also restore its support interactions and the positively charged metal center
28
original high catalytic performance. An operando XAFS experi- also account for the high activity and stability.[83] Ab initio study
29
ment also confirmed that the isolated Pdδ + is more active than of Au SACs with different supports revealed that the supports
30
the metallic Pd0 for Pd1/La-γ-Al2O3. Besides the differences in with higher reducibilities tend to stabilize single Au atoms
31
reactivities, different reaction kinetics were also identified on better as a result of the charge transfer from the Au 6 s orbital
32
La-γ-Al2O3 supported Pd single atoms and nanoparticles where to the support, which leads to a positively charged Au +.[84] The
33
the Pd SAC is positively ordered in CO while the Pd nanoparticle positively charged Au + not only results in a high thermal
34
catalyst in negatively order in CO. stability but also higher CO and O2 adsorption energies. For
35
Although experimental studies on Pd SACs for CO oxidation Au1/CeO2, single Au atoms prefer to adsorb on the CeO2 step
36
are still at an early stage, theoretical studies were performed to site rather than the terrace site (see Figure 9a for the chemical
37
guide the design of Pd SACs on different types of supports. potentials). The step-site-anchored Au + adsorbs CO strongly,
38
Interestingly, molecular dynamics calculations indicate that Pd1/ and high CO oxidation activity (10 orders of magnitude higher
39
MgO may be able to adsorb more than one CO during CO than that of Au single atom on a CeO2 terrace at 700 K) and
40
oxidation, with Pd(CO)2O2 or Pd(CO)(CO)3 identified as the thermal stability was reported on the Au1 + (CO) structure
41
reaction intermediates.[72] It is worth mentioning that although (Figure 9b).[84] The active structure of Olattice-Au + CO on Au1/
42
both Pd dicarbonyl[72] and Pt dicarbonyl[73] SAC structures have CeO2 was further confirmed by a combined vibrational spectro-
43
been reported theoretically, so far there is no experimental scopic and theoretical study.[86] A combined scanning tunneling
44
evidence. CO oxidation mechanisms on carbon-based or boron- microscopy (STM) and X-ray photoelectron spectroscopy (XPS)
45
nitride-based 2-D materials were also studied theoretically. study on Au1/CuO further illustrated that CO oxidation reactivity
46
First-principle calculations revealed a TER mechanism for CO can be determined by the charge transfer between Au and the
47
oxidation on Pd SACs with defect graphene,[74] a BN support.[87] When a lattice O atom transfers a partial negative
48
nanosheet,[75] or N-doped graphene[76] as support. charge (δ ) to a Au single atom, the lattice O is activated and
49
reacts with a CO (Figure 9c). However, once the O vacancy
50
develops after the CO2 formation, the Au single atoms become
51
3.4. Au SACs neutral and inactive. This is a result of the inability of the
52
neutral Au0 to create an O vacancy; however, the reactivity can
53
Pt, Pd and Rh play a major role in current catalytic converters. be restored by refilling the O vacancy with O2, which again
54
Though the traditional bulk Au catalyst is not as thermally produces the negatively charged Auδ (Figure 9d).[87] Although
55
stable as the abovementioned catalysts, the nano-sized Au Au1/TiO2(110) was shown to be inactive for CO oxidation,[10] DFT
56
catalysts showed promising low-temperature reactivity and calculations revealed that catalytic activity can be significantly
57
ChemNanoMat 2020, 6, 1 – 25 www.chemnanomat.org 11 © 2020 Wiley-VCH GmbH

These are not the final page numbers! ��


Minireview
oxidation reactivity on Au SACs can be significantly promoted
1
by co-feeding water; in comparison, this water promotion effect
2
is much weaker on supported Au nanoparticle catalysts.[93]
3
In summary, the reactivities of Au SACs are highly depend-
4
ent on their local environments and charge states, and the
5
charge transfer between Au and its neighboring atoms could
6
also affect the preferred reaction mechanism of Au SACs.[55]
7
8
9
3.5. Ag and Ru SACs
10
11
CO oxidation on Ag or Ru SACs was also studied theoretically.
12
Interestingly, both Ag1/CeO2 and Ru1/CeO2 were predicted to be
13
actively catalytic for CO oxidation.[94] For a Agn/CeO2 (n = 1–10)
14
catalyst with a CeO2(111) surface„ a single-atom Ag could
15
adsorb CO and migrate onto the CeO2 surface during CO
16
oxidation. The migrated Ag CO may then react with a lattice O
17
via the MvK mechanism. Therefore, the size dependence of CO
18
oxidation reactivity is not obvious in the size range of Ag1 Ag10
19
Figure 9. Illustration of the effect of the adsorption site (a,b) and charge
as a result of the mobility of the Ag atom in the cluster
20
state (a,b) on the reactivities of Au SACs. (a) Chemical potential of Au1/CeO2 catalysts.[94] For Ru1/CeO2, DFT calculations were performed on
21 with different adsorption sites, Au nanoparticle and Au1/MgO. (b) Reactivity the CeO2(111) surface.[95] Both CO and O2 adsorb strongly on
22 of Au1/CeO2 (Au single atom on the CeO2 step site) and Au12/CeO2(111). (c)
STM image of a CuO-monolayer-supported Au single atom after CO
Ru1/CeO2 with adsorption energies of 1.87 and 2.06 eV,
23
adsorption at RT. (d) CO oxidation and the corresponding changes in charge respectively. Between the LH and ER mechanisms, the LH
24 states of Au1/CuO. (a) and (b) are adapted with permission from ref. [84]. mechanism was the preferred pathway when CO and O2
25 Copyright 2017 American Chemical Society. (c) and (d) are adapted with
permission from ref. [87]. Copyright 2018 American Chemical Society.
reacted at the same Ru atom. The rate-limiting step was
26
identified as the reaction between CO* and O* with a low
27
energy barrier of 0.47 eV. The stability of Ru single atoms on
28
TiO2 and Al2O3 was also studied, and the Ru was calculated to
29
improved when doping the TiO2 substrate with Cu or Ni.[88] The be mobile on TiO2(110) and Al2O3(001). Other than the thermally
30
Cu or Ni dopant depletes the d-orbital of the isolated Au atom, stable Ru/CeO2, clusters tended to form with Ru single atoms
31
and the more positively charged Auδ + leads to a lower O2 on TiO2 or Al2O3.
32
activation barrier on Au.
33
It is essential to understand the reaction mechanism of Au
34
SACs for better optimization of their active sites. The LH 3.6. Non-noble metal SACs
35
mechanism was favored on Au/TiO2;[79e,88] the MvK mechanism
36
was preferred on Au/FeOx,[83] Au/CeO2 (step site),[84] and Au/ Although noble metal SACs have been widely adopted for CO
37
CuO;[87] and the ER mechanism was reported on Au1/ oxidation, their high catalytic cost makes them disadvantageous
38
CeO2(110).[89] The preferred reaction mechanism is often for practical application. Research interest has grown in the
39
sensitive to the local structures around Au atoms. For example, study of non-noble metal SACs to enhance their catalytic
40
on single-walled carbon nanotube supported Au SACs, CO performance for CO oxidations. So far, most of these studies are
41
oxidation can proceed through either the LH or ER pathways theoretical with a focus on catalyst screening and reaction
42
depend on the site that the Au atoms are anchored to (e. g., mechanisms. Here, we summarize the recent progress on CO
43
monovacancy, divacancy, or a 5-7-7-5 Stone-Wales defect oxidation over different non-noble, metal-based SACs on metal
44
site).[90] In another study, a new dual-site mechanism is favored oxide material supports.
45
when Au SACs are anchored as adatoms on Th(111); O2 was Metal-oxide-supported, non-noble metal SACs showed a
46
activated on the Th site without the involvement of the Th Au tremendous potential to catalyze low-temperature CO oxida-
47
interface.[91] On the other hand, the MvK mechanism is preferred tion. First-principle calculations revealed that CO oxidation on
48
when a Au single atom was doped on ThO2(111). The CO Ni1/FeOx follows the same MvK mechanism as Pt1/FeOx and Ir1/
49
oxidation reaction mechanism of Au SACs also has an effect on FeOx.[96] Surprisingly, the reaction barrier of Ni1/FeOx (0.75 eV) is
50
the Au nanoparticle catalysts with reducible supports. DFT- lower than that of Pt1/FeOx (0.79 eV) and Ir1/FeOx (1.41 eV),
51
based ab initio molecular dynamics simulations suggest that which suggests that Ni1/FeOx may be an active catalyst for low-
52
isolated Au atoms could migrate away from bulk Au to the temperature CO oxidation. The adsorption energy of CO on Pt1/
53
CeO2 support with the assistance of CO, and these Au single FeOx is calculated to be 0.73 eV, which is much lower than
54
atoms can catalyze CO oxidation with a lower energy barrier. that on Pt1/FeOx (1.08 eV) or Pt1/FeOx (1.77 eV). The high CO
55
The dynamically formed Au single atoms integrate into the Au adsorption energy could hinder the reaction between adsorbed
56
particles again after the reaction.[92] Similar to the Pt SACs, CO CO and the lattice O, which could possibly explain the superior
57
ChemNanoMat 2020, 6, 1 – 25 www.chemnanomat.org 12 © 2020 Wiley-VCH GmbH

These are not the final page numbers! ��


Minireview
activity of Ni1/FeOx.[96] Cu1/CeO2 is another promising candidate the support. Here, we highlight the active trends on several
1
and demonstrated some interesting features for low-temper- commonly used SAC supports.
2
ature CO oxidation. First, the Cu single atom adsorbs strongly Systematic DFT calculations were performed on M1/FeOx,
3
on an O-hollow site of CeO2(111), with an adsorption energy of M1/MgO, and M1/γ-Al2O3 (M denotes metal) to guide the
4
2.89 eV. Second, Cu1/CeO2 adsorbs CO strongly with adsorp- catalyst design. Figure 10 summarizes the reaction barriers of
5
tion energy of 1.52 eV. This indicates that Cu single-atoms the RDS and the CO/O2 adsorption energies reported on the
6
adsorb CO much stronger than Cu nanoparticles, where Cu abovementioned catalysts. DFT studies were performed on
7
(111)/Cu(001) adsorbs CO with an adsorption energy between FeOx-supported Rh, Pd, Co, Cu, Ru, and Ti SACs to guide the
8
0.74 and 0.86 eV.[97] This also suggests that stable CO design of the catalyst.[43a] Apart from the O2-using MvK
9
adsorption bands can be observed in DRIFTS possibly at room mechanism of Pt1/FeOx[13b] and Ir1/FeOx,[63] the LH mechanism
10
temperature where nanoparticles usually requires a much lower was identified as the most favorable mechanism on FeOx-
11
temperature. CO oxidation proceeds on Cu1/CeO2 through the supported Rh, Pd, Co, Cu, Ru, and Ti SACs for both vacancy-free
12
MvK mechanism, and the rate-limiting step was identified as and O-defective FeOx supports.[43a] Among all the calculated O-
13
the CO2 desorption with a low energy barrier of 0.55 eV.[97] defective M1/FeOx, Rh1/FeOx revealed the lowest reaction barrier
14
Clearly, the Cu1/CeO2 may potentially be an active and stable (0.53 eV); Cu1/FeOx and Pd1/FeOx also had lower energy barriers
15
SAC for low-temperature CO oxidation. Besides the metal oxide than Pt1/FeOx (Figure 10a).[43a] For the vacancy-free M1/FeOx,
16
support, recently, a single Cu atom doped CuAl4O7–9 cluster Ru1/FeOx is predicted to be highly active (energy barrier of
17
was also reported to be active for CO oxidation.[98] This study 0.27 eV), and Cu1/FeOx and Ti1/FeOx also demonstrated superior
18
showed that the CO adsorbed CuAl4O9CO is more active than catalytic performance. CO and O2 adsorption energies vary
19
the bare CuAl4O9 structure. This is due to the fact that the among different metals, and a direct relationship has not yet
20
oxidation state is adjusted to around + 1 by the adsorbed CO been determined between the CO/O2 adsorption energies and
21
which facilitates the CO oxidation of another CO. This study the CO oxidation reactivity of M1/FeOx (Figure 10a,b).[43a] The d-
22
demonstrated the correlation between CO oxidation reactivity band centers of high-performance catalysts such as Rh1/FeOx,
23
and the Cu oxidation states at the molecular level. Ru1/FeO, Cu1/FeOx, and Ti1/FeOx are higher than those of Pt1/
24
FeOx, which could partially explain their high CO oxidation
25
reactivity. Interestingly, some of those non-noble M1/FeOx
26
3.7. Theoretically guided SAC designs with different supports catalysts perform better than Pt1/FeOx, and further experimental
27
study into these low-cost catalysts is of great importance.
28
Besides the studies focused on SACs with a specified metal Moreover, first-principle studies on M1/MgO also identified
29
center, tremendous efforts have also been made to compare different active catalysts such as Ag1/MgO, Cu1/MgO, and Ni1/
30
different metal centers with the same support during CO MgO that have better CO oxidation performance than Pt1/MgO
31
oxidation, as well as to understand the general contribution of (Figure 10c; MgO with Fs-defect is selected as an illustration).[99]
32
The LH mechanism is also preferred among all calculated M1/
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
Figure 10. Illustration of the reaction barrier and CO/O2 adsorption energy over (a) M1/FeOx with O vacancy,[43a] (b) M1/FeOx without O vacancy,[43a] (c) M1/Fs-
56 defect MgO,[99] and (d) M1/γ-Al2O3.[100] “ + ” denotes the CO adsorption energy (CO Eads), and “�” denotes the O2 adsorption energy (O2 Eads).
57
ChemNanoMat 2020, 6, 1 – 25 www.chemnanomat.org 13 © 2020 Wiley-VCH GmbH

These are not the final page numbers! ��


Minireview
MgO. In this study, the Sabatier activity is correlated to the CO be less active for HC oxidation compared to their metallic
1
and O2 adsorption energy, and the catalysts with optimized CO cluster counterparts. For example, Jeong et al.[105] compared
2
and O2 adsorption energies were shown to be more active for CeO2-supported noble metal (Pt, Pd, and Rh) clusters and single
3
CO oxidation. The CO oxidation performance of M1/γ-Al2O3 atoms for a TWC reaction, showing that SACs were much less
4
(M=Pd, Fe, Co and Ni) was also investigated theoretically.[100] active than the metallic clusters for HC (C3H6 and C3H8)
5
Although γ-Al2O3 is usually considered an irreducible support, oxidation. Goodman et al.[106] reported that Al2O3-supported Pd
6
the reaction on M1/γ-Al2O3 follows the MvK mechanism. This nanoparticles rapidly lost activity for methane oxidation after
7
illustrates that the lattice O on γ-Al2O3 can be activated by the decomposition into inactive single atoms at high temperatures.
8
metal single atoms. Among all the catalysts for which the A Pd SAC had a methane combustion rate that was more than
9
adsorption energies are calculated (Figure 10d), Ni1/γ-Al2O3 has two orders of magnitude lower than that of Pd particles. The
10
the lowest reaction barrier (0.52 eV) and is predicted to be more low HC oxidation activity of SACs is probably due to the need
11
active than Ni1/γ-Al2O3 (reaction barrier of 0.90 eV). for a critical ensemble metal particle size to adsorb O2 and fully
12
Besides traditional metal oxide materials, perovskite (ABO3 activate the C H bonds.[106]
13
type) materials have also recently been identified as promising There are various kinds of HCs, and their different functional
14
candidates for emission control catalysts.[101] Adding or doping groups can lead to different catalytic requirements for the
15
perovskite materials with highly dispersed metal single atoms oxidation reactions. It is difficult to provide a generalized
16
could be an efficient strategy for enhancing their CO oxidation correlation and a single reaction mechanism for different HCs.
17
performance. First-principle calculations were performed re- However, Palazzolo and Tichenor[107] established the following
18
cently on perovskite-supported SACs to guide their design. For reactivity order regarding the relative ease of oxidation of
19
example, DFT calculations on M1/LaFeO3 (M=Rh, Pd, Pt) revealed different HC types: alcohols < aldehydes < aromatics < ke-
20
that a M1/LaFeO3 catalyst with a La defect demonstrated more tones < acetates < alkanes. Here, we summarize the current
21
optimized CO/O2 adsorption and therefore better catalytic studies on using SACs for catalytic oxidation of lower alkane/
22
performance.[102] Among LaFeO3 Rh, Pd, and Pt single atoms, alkene HCs, aromatics, and oxygenated HCs.
23
Pd1/ LaFeO3 with a La defect presents the highest reactivity. A
24
similar strategy was also applied to guide the design of M1/
25
LaBO3 (M=Rh, Pd, Pt; B=Mn, Fe, Co, Ni).[103] In general, Pd1/LaBO3 4.1. SACs for oxidation of light hydrocarbons
26
exhibits superior catalytic performance to Rh1/ LaBO3 and Pt1/
27
LaBO3. This can be explained by the fact that Pd adsorbs CO Methane is a major hydrocarbon air pollutant generated by
28
much more weakly because of its lower d-band center (more natural gas-fueled vehicles and gas power plants. Methane has
29
filled d state). a much larger detrimental greenhouse effect than carbon
30
In summary, theoretical calculations are powerful tools to dioxide.[108] Other light HCs such as propane and propene are
31
guide the design of metal SACs. The catalytic performance and often used as representative model compounds to study the
32
reaction mechanisms of SACs are strongly affected by the metal catalytic oxidation of alkanes and alkenes.
33
and support selected as well as the anchored site. Different As reported by Asakura et al.,[104] the Pt/MgO SAC showed
34
non-noble metal SACs have been predicted to be highly active, good activity for propane oxidation similar to that of MgO-
35
and those catalysts are excellent candidates for future exper- supported Pt particles. It was proposed that the Pt4 + ion
36
imental studies. CO/O2 adsorption of SACs is directly related to substituted Mg2 + on the top layer of the MgO lattice and
37
the d-orbital density of states. The contribution of CO/O2 distorted the local structure. The Pt O coordination number
38
adsorption to CO oxidation reactivity varies between catalysts, decreased from 5 to 4 after the catalytic combustion reaction,
39
and further study is required to elucidate the underlying trend. suggesting that the O atom coordinated to the Pt atom
40
contributes to the combustion reaction. A vacancy on the Pt
41
site is created through the reaction between the Pt O lattice O
42
4. SACs for HC oxidation and propane and functions as the active site for the steady-
43
state combustion process. It has also been reported that the
44
HCs is one of the major air pollutants in industrial and reactivity of the lattice O atoms in the Pt/MgO SAC was
45
automobile exhausts, and HCs in automotive exhaust are enhanced in the presence of H2O vapor, resulting in an O2-
46
generated by the incomplete combustion of HC fuels. Noble lattice O exchange rate 10 times faster than that without
47
metal SACs, which have been extensively investigated and water.[109] In contrast, such a promoting effect by H2O vapor was
48
proven to have superior activity and stability for CO oxidation, not observed with the Pt particle catalyst.[109] It was proposed
49
attracted much less attention in studies of HC oxidation. As that the H2O vapor generated during HC combustion could
50
early as 1999, Asakura and co-workers[104] reported that a Pt/ promote the activity of the monomeric Pt site.
51
MgO catalyst with atomically dispersed Pt ions was as active as Hoang et al.[110] reported a single-atom Pt catalyst that was
52
metallic Pt particles supported on MgO for propane combus- strongly anchored on a robust nanowire forest of mesoporous
53
tion, which demonstrated the potential of SACs for the catalytic rutile TiO2 grown on the channeled walls of full-size cordierite
54
combustion of HCs. After that study, there have been limited honeycombs. As shown in Figure 11, with over 80% Pt usage
55
studies on using SACs for the catalytic combustion of HCs. reduction compared to that of the commercial benchmark, this
56
Atomically dispersed ionic noble metals were often known to Pt SAC exhibited remarkable oxidation activity not only for CO
57
ChemNanoMat 2020, 6, 1 – 25 www.chemnanomat.org 14 © 2020 Wiley-VCH GmbH

These are not the final page numbers! ��


Minireview
existing surface single-atom Pt species on the support, which
1
dramatically promoted the Pt-support interactions to anchor
2
the Pt single-atoms, making them extremely stable under
3
calcination at 800 °C in humid air for 5 days. This work showed
4
that support reconstruction is a viable approach to solving the
5
problem of instability of noble metal SACs at high temper-
6
atures; however, the methane oxidation activity of this Pt/
7
Mn2O3 catalyst was not as impressive, with a T100 above 600 °C
8
(WHSV of 20 L g 1 h 1). However, while emphasizing the stability
9
of the SACs for methane oxidation, these works did not attempt
10
to reveal the detailed catalytic role of the Pd or Pt single atoms
11
in the oxidation reaction.
12
The highly dispersed but non-isolated ensemble catalysts
13
Figure 11. Diesel oxidation performance of Pt1/TiO2 nano-array (NA) inte- were reported to be alternative of SACs for the improved HCs
14 grated monoliths: Light-off curves for Pt1/TiO2 NA (0.71 gPt L 1) in the CDC
simulated exhaust (12% O2, 6% H2O, 6% CO2, 200 ppm NO, 100 ppm H2,
oxidation reactivity while maintaining 100% metal dispersion.
15
500 ppm CO, 389 ppm C2H4, 233.5 ppm C3H6, 51.9 ppm C3H8, space velocity Jeong et al. reported a CeO2 supported Rh ensemble catalyst
16
30,000 h 1). Reproduced from ref. [110] which is an open access article. with average particle size of 0.9 nm with 100% Rh surface
17 Copyright 2020 Springer Nature Limited.
dispersion under reaction condition of HCs oxidation.[114] The
18
CeO2 supported Rh ensemble catalyst showed much superior
19
reactivity than its SAC and nanoparticle catalyst (NP) counter-
20
(T90 = 160 °C) but also for light HCs (mixture of C2H4, C3H8, and parts during oxidation of both C3H6 (Figure 12a) and C3H8
21
C3H8) (T90 = 170–180 °C) under clean diesel combustion simu- (Figure 12b). The Rh/CeO2 SA is almost inactive up to 400 °C.
22
lated diesel exhaust conditions. The authors ascribed the SACs with the structure of dinuclear cluster cations/anions
23
enhanced low-temperature reactivity of Pt SACs to the distinct such as RhAl3O4 + and RhVO3 also showed promising reactivity
24
electronic structures that modify their interactions with adsor- for methane activation and conversion.[115] The Rh1-doped
25
bate molecules or the involvement of supports that may alter aluminum oxide cluster cations RhAl3O4 + efficiently activate the
26
reaction pathways, as proposed previously for CO oxidation on C H bond and convert methane to syngas and the high C H
27
Pd and Pt SACs.[111] Excellent low-temperature activities were activation is driven by the high electron withdrawn ability of Rh
28
sustained after hydrothermal aging (700 °C for 100 h) and single atoms.[115a] Similarly, the dinuclear RhVO3 species
29
sulfation treatment (5 ppm SO2, 300 °C, 5 h) and the active showed exceptional C H activation performance by converting
30
single-atom Pt species remained stable on the TiO2 nanowire
31
arrays. This excellent stability is a result of a strong electrostatic
32
interaction between Pt single-atom species and the TiO2 nano-
33
wire surface. However, the propane oxidation activity of this
34
SAC is relatively low (Figure 11).
35
Duan et al.[112] synthesized a single-atom 0.23 wt.% Pd/SiO2-
36
ZrO2 catalyst for methane oxidation. The authors employed
37
organo-silane as a surfactant to modify the ZrO2 support by
38
decorating it with a layer of spatial alkyl chain barriers on the
39
surface. Pd acetate was selected as the Pd precursor that
40
existed as a three-dimensional structure in the toluene solution.
41
This approach dispersed Pd as single atoms or sub-nanoscale
42
clusters on the SiO2 ZrO2 support, which proved to be
43
beneficial for catalytic methane combustion. This 0.23 wt.% Pd/
44
SiO2 ZrO2 catalyst presented outstanding activity for methane
45
combustion, with a T100 of 400 °C (WHSV of 30 L g 1 h 1); this is
46
200 °C lower than that of 0.23 wt.% Pd/ZrO2, which contains
47
predominantly Pd nanoparticles. It was found that the hydroxyl
48
groups formed during the reaction were transferred to the
49
silicon atoms, which efficiently reduced the formation of the
50
inactive Pd(OH)x species and lead to the excellent stability of
51
the Pd/SiO2 ZrO2 catalyst. Yan and co-workers[113] reported an
52
approach to synthesize highly stable single-atom 0.8 wt.% Pt/
53 Figure 12. Catalytic activity for (a) C3H6 oxidation (0.2% C3H6 + 1.8% O2,
Mn2O3 catalysts for methane oxidation. They used Mn3O4 as a
54 balanced with He), (b) C3H8 oxidation (0.2% C3H8 + 2% O2, balanced with He)
restructurable support to load single-atom Pt and turn it into a on Rh/CeO2 single-atom catalysts (SAC), highly dispersed ensemble catalyst
55
single-atom Pt-on-Mn2O3 catalyst via high-temperature treat- (ENS) and nanoparticle catalyst (NP). Reproduced from ref. [114]. Copyright
56 2018 American Chemical Society.
ment. The Mn3O4 to Mn2O3 phase transition firmly confined the
57
ChemNanoMat 2020, 6, 1 – 25 www.chemnanomat.org 15 © 2020 Wiley-VCH GmbH

These are not the final page numbers! ��


Minireview
methane and CO2 to CH3OH and CH2O in the temperature range activity for toluene oxidation, with the T50 and T90 being 170 °C
1
of 20 to 327 °C.[115b] and 205 °C, respectively (1000 ppm toluene, 40% O2,
2
40 L g 1 h 1). However, the toluene conversion at 205 °C quickly
3
decreased from 90% to 30% within 2.5 h of the initiation of the
4
4.2. SACs for oxidation of aromatics on-stream reaction. Adding CeO2 to the 0.06 wt.% Ag/Mn2O3
5
catalyst was found to not only improve the catalytic activity but
6
Aromatic compounds such as benzene, toluene, and ethyl- also significantly enhance the stability of the catalyst. Toluene
7
benzene are present in gasoline, diesel, and other fuels conversion on the 0.63 wt.% CeO2–0.06 wt.% Ag/Mn2O3 catalyst
8
produced from crude oil. Consequently, the incomplete (195 °C) decreased by only 10% after 50 h of the on-stream
9
combustion of liquid fuels in automobiles releases significant reaction. Because Ag and CeO2 particles were highly dispersed
10
amounts of aromatic compounds. On the other hand, polycyclic on the Mn2O3 nanowire support, the O species formed at the
11
aromatic hydrocarbons are mainly released by combustion surface O vacancies of CeO2 could efficiently migrate to the
12
processes.[116] active sites (i. e., the Ag Mn2O3 interface) and replenish the
13
Wang and co-workers[117] used the molten salt method to surface reactive lattice O species.
14
synthesize a 0.2 wt.% Pt/TiO2 catalyst containing both single Yang and co-workers[121] synthesized a meso-Fe2O3-sup-
15
atoms and nanoparticles of Pt. This catalyst exhibited good ported Pt SAC (Pt1/meso-Fe2O3) for benzene oxidation. To
16
catalytic performance for toluene oxidation, with a T50 and T90 accomplish this, the authors first synthesized a mesoporous
17
of 173 °C and 183 °C, respectively (1000 ppm toluene, Fe2O3 support (meso-Fe2O3) using the KIT-6 template. Pt1/meso-
18
40 L g 1 h 1). Benzaldehyde was found to be the most stable Fe2O3 SACs were then prepared via a polyvinyl alcohol (PVA)-
19
surface intermediate, and its oxidation could be rate-limiting for protected reduction route with PVA as the protecting agent
20
the entire toluene oxidation reaction. However, it was not clear and NaBH4 as the reducing agent. The 0.15 wt.% Pt1/meso-Fe2O3
21
whether the single atom or the nanoparticle contributed more and 0.25 wt.% Pt1/meso-Fe2O3 catalysts exhibited higher cata-
22
to the catalytic activity. lytic activities than the nanoparticle counterpart (0.25 wt.%
23
Zhao and co-workers[118] synthesized a MgO nanosheets Pt PtNP/meso-Fe2O3). The 0.25 wt.% Pt1/meso-Fe2O3 performed the
24
SAC and tested it for toluene oxidation. This SAC demonstrated best, with a T90 of 198 °C (1000 ppm benzene, 20% O2,
25
optimum activity for toluene oxidation, with a T50 and T90 of 20 L g 1 h 1) and a reaction rate 7 times faster than that of
26
160 °C and 220 °C, respectively (100 ppm toluene, 36 L g 1 h 1). 0.25 wt.% PtNP/meso-Fe2O3. The good catalytic performance of
27
It was proposed that adsorption and activation of molecular O2 0.25 wt.% Pt1/meso-Fe2O3 was associated with the high utiliza-
28
is most critical for toluene oxidation. Pt single atoms on MgO tion of Pt atoms. Furthermore, the 0.15 wt.% Pt1/meso-Fe2O3
29
nanosheets facilitate the formation of O vacancies. Compared and 0.25 wt.% Pt1/meso-Fe2O3 samples were water-resistant,
30
with pure MgO, Pt1/MgO is more facile to generating O2 which was attributed to the formation of active radicals and the
31
vacancies, which facilitates the activation of O2 and thereby decomposition of carbonates. In situ DRIFTS results demon-
32
promotes the formation of corresponding active O2 species. strated that formation of the phenolate and benzoquinone as
33
Introducing H2O (2.5 vol.%) led to an increase in the catalytic well as cyclohexanone and maleate, which were the main
34
activity of Pt1/MgO, with T50 and T90 decreasing to 140 °C and intermediates in the oxidation of benzene. The good stability of
35
170 °C, respectively. In the presence of H2O, the O2 prefers to be the 0.15 wt.% Pt1/meso-Fe2O3 and 0.25 wt.% Pt1/meso-Fe2O3
36 *
activated and dissociated to generate hydroxyl radicals ( OH) samples was associated with the strong interaction between Pt
37
on the O vacancies of the Pt1/MgO surface, which is the and meso-Fe2O3. In addition, the 0.25 wt.% Pt1/meso-Fe2O3
38 *
dominant active O2 species. The critical role of OH in toluene sample possessed a good carbon-dioxide-resistant ability.
39
oxidation was also proposed by Zhang et al.[119] on Pt/MnO2
40
SACs. These researchers synthesized 0.05–0.5 wt.% Pt/MnO2
41
catalysts with atomically dispersed Pt via a one-pot hydro- 4.3. SACs for oxidation of oxygenated HCs
42
thermal process utilizing the richness of Mn defects in
43
birnessite-type MnO2. The doping of single-atom Pt onto MnO2 Currently, there have been limited studies on SACs for catalytic
44
greatly improved its catalytic activity for toluene oxidation. The oxidation of oxygenated HCs. These studies were mainly related
45
0.1 wt.% Pt/MnO2 SAC had a low T50 and T100 of 110 °C and to industrial emission control or indoor air pollutant purification
46
160 °C, respectively (0.1 wt.% Pt, 100 ppm toluene, 48 L g 1 h 1). under much lower temperatures (e. g., < 100 °C) and not directly
47
This catalyst was also capable of toluene degradation at room related to automotive emission control. However, the develop-
48
temperature, achieving 100% conversion of 0.42 ppm toluene ment of biomass-derived oxygenated compounds (including
49
at 28 °C under a high space velocity of 300 L g 1 h 1, making it a alcohols, ketones, and furans) as fuel blendstocks[122] is expected
50
potential catalyst for removal of indoor VOCs. It was found that to increase the amount of oxygenated HCs in engine exhaust;
51
the Mn and O2 defects in MnO2 nanosheets effectively stabilize therefore, more attention will be paid to the application of
52 *
the single-atom Pt, and strong oxidative OH radicals were SACs for catalytic combustion of oxygenated HCs.
53
proposed to contribute to its excellent performance. Jiang et al.[123] prepared a 0.02 wt. % Pt1/Co3O4 SAC that
54
Zhang et al.[120] applied the molten salt method to load a consisted of isolated single Pt atoms anchored to Co3O4 (111)
55
small amount of Ag (0.06 wt.%) on a Mn2O3 nanowire to planes by occupying positions on the Co2 +. This catalyst
56
prepare a Ag/Mn2O3 SAC, which exhibited excellent catalytic exhibited excellent activity for the oxidation of methanol, with
57
ChemNanoMat 2020, 6, 1 – 25 www.chemnanomat.org 16 © 2020 Wiley-VCH GmbH

These are not the final page numbers! ��


Minireview
a low T50 and T90 of 81 °C and 96 °C, respectively (700 ppm the properties of the support. A wide range of metal oxide
1
methanol, 21% O2, GHSV = 30000 h 1). The TOF of methanol supports have been investigated, including reducible metal
2
oxidation (70 °C) on Pt1/Co3O4 was 4 times higher than that on oxides (e. g., TiO2, Fe2O3, Mn2O3, and CeO2), inert oxides (e. g.,
3
the counterpart nanoparticle catalyst 0.05 wt.% Pt/Co3O4. One- SiO2 ZrO2 and Al2O3), and alkali earth metal oxides (e. g., MgO,
4
third of the Co atoms in Pt1/Co3O4 were determined to exist as HMO). On some supports, however, SACs were found to be
5
Co2 + in a tetrahedral coordination to O; the remaining Co relatively inactive for HC oxidation; on other supports, SACs
6
atoms were present as Co3 + in an octahedral O environment. It demonstrate similar or higher activity compared to their nano-
7
was proposed that the Pt sites – which have high occupied particle counterpart for HC oxidation. The strong interactions
8
electronic states – exhibit a strong affinity for the 3d orbital of between the metal single atoms can facilitate HC oxidation in
9
adjacent Co atoms, resulting in significant electron transfer different ways, including by creating O vacancies, generating
10 *
from Pt to Co and ultimately increasing the proportion of O reactive OH radicals, or modifying the electron state of the
11
vacancies present on the catalyst surfaces. Regeneration of support.
12
these O vacancies was determined to promote the co-
13
adsorption of methanol and O2, leading to an increase in the
14
rate of C H bond dissociation. DFT calculations revealed that 5. SACs for SCR
15
electron transfer reduces the activation barriers for methanol
16
oxidation. To meet increasingly stringent NOx (NO and NO2) emission
17
Tang and co-workers[124] conducted a series of works standards, three main technologies have been proposed to
18
investigating Ag1/HMO SACs, with Ag singleatoms supported reduce NOx emissions. These technologies are SCR of NOx, direct
19
on hollandite manganese oxide (HMO), for low-temperature decomposition of NOx, and NOx storage reduction (NSR).[126] SCR
20
formaldehyde (HCHO) oxidation. The AgAOR/HMO SAC was is the most promising technique for the decomposition of NOx
21
synthesized from a traditional supported Ag particle sample (deNOx) and has been commercially applied in automobile
22
(AgNP/HMO) by anti-Ostwald ripening.[124a] At 80 °C, the TOF (ca. vehicles.[127] SCR can be carried out with different reducing
23
0.035 s 1) of AgAOR/HMO is approximately 7 times higher than agents such as NH3/urea (NH3 SCR),[128] HCs (HC SCR),[129] and
24
that of the corresponding AgNP/HMO (ca. 0.005 s 1). The high CO (CO SCR).[130] The removal of NOx with CO or an HC worked
25
activity of the Ag SACs was ascribed to the strong ability of well in stoichiometric conditions over the TWCs when the
26
single-atom Ag to activate both lattice O and molecular O.[124a] reducing pollutants were stoichiometrically matched exactly by
27
Afterward, by comparing AgAOR/HMO and a SAC prepared with NOx and O2 in the gasoline engine; however, the reducing
28
conventional wet impregnation (AgIMP/HMO), these researchers efficiency of TWCs under dynamic lean or rich perturbation
29
found that the different preparation methods could control the conditions will decrease.[131] NH3/urea-SCR technologies were
30
electronic metal-support interactions (EMSI) of single-atom utilized more frequently when coupled with a DOC and a DPF,
31
silver catalysts.[124b] The AgAOR/HMO SAC has a higher depletion especially in lean-condition diesel engines.[132] NH3 can be used
32
of the d electronic state of the Ag active sites, resulting in a directly as the reductant or in situ formed from the decom-
33
stronger EMSI and leading to its stronger activation ability position of urea. In this review, we mainly focus on NH3 SCR
34
toward O2, easier reducibility, and higher catalytic activity for rather than urea-SCR because the goal is to discuss the prospect
35
formaldehyde oxidation. In their subsequent work,[125] these of using SACs for deNOx. We also summarize the current
36
researchers reported a more active Na1/HMO SAC, which application of popular catalysts, including the use of SACs on
37
exhibited higher catalytic activity for HCHO oxidation than the HC SCR[133] and CO SCR.[134]
38
AgAOR/HMO SAC. They demonstrated that the catalytically active The NH3 SCR processes can be classified into different
39
centers include the surface single Na or Ag atoms and the routes, including standard SCR reaction (4 NO + 4 NH3 + O2!
40
vicinal lattice O atoms and that the electronic states of the 4 N2 + 6 H2O), fast SCR reaction (2 NO + 4 NH3 + 2 NO2!4 N2 + 6
41
surface lattice O play a critical role in determining the catalytic H2O), and other SCR reactions with different NO/NO2/NH3/O2
42
performance for HCHO oxidation. The surface lattice O species ratios.[135] During the NH3 SCR process, the type of catalyst used
43
of Na1/HMO have a more negative charge and therefore is undoubtedly the most crucial part of highly efficient and
44
stronger nucleophilic properties, making Na1/HMO more effi- selective NOx abatement. Three types of frequently used
45
cient than AgAOR/HMO at abating HCHO. catalysts include metal ions- (Cu or Fe) exchanged zeolites,
46
In summary, the study of SACs for low-temperature V2O5-based materials, and other V-free metal oxides (Ce, Mn-
47
combustion of HCs has recently emerged as a primary research based materials). These catalysts usually undergo various
48
interest. Depending on their functional groups, different types reaction mechanisms and include respective drawbacks such as
49
of HCs could have different catalytic requirements for oxidation narrow temperature windows, inferior SO2-resistant perform-
50
reactions, which increases the complexity and difficulty of ance, unsatisfactory hydrothermal stability, hydrocarbon coking,
51
catalyst design. Currently, most of the reported works are and alkali or heavy-metal poisoning.[127c,136] For example, V-free
52
dealing with aromatics, while less attention has been paid to metal oxides can operate in a broader temperature window
53
light HCs and oxygenated HCs. While Pt SACs are predominant compared to V-based materials, but they both suffer from issues
54
in current studies, other metals including Pd, Ag, and even Na with low-temperature SO2-tolerance and alkali/heavy-metal
55
have also demonstrated their potential as active SACs for the resistance. Similarly, these catalysts all have acidic and redox
56
oxidation of HCs. The activities of SACs are highly affected by sites that close the SCR reaction cycle, and their redox proper-
57
ChemNanoMat 2020, 6, 1 – 25 www.chemnanomat.org 17 © 2020 Wiley-VCH GmbH

These are not the final page numbers! ��


Minireview
ties affect catalytic activity predominately under low temper- relationship with the Cu density at high Cu densities, which is
1
atures while acidity dominates under relatively high inconsistent with previous reports.[142] A nonlinear trend was
2
temperatures.[127c] To date, numerous reviews about metal- observed at low Cu densities, suggesting two distinct kinetic
3
exchanged zeolites,[132b,137] V2O5-based materials,[138] and V-free regions. The origin of this discrepancy is the non-single site
4
metal oxides[136b,139] have all been summarized. In addition, the behavior in the oxidation half-cycle, where the O-bridged CuII
5
SCR reaction mechanism and some rational explanation about dimer is preferably formed from the CuI species in an O2
6
H2O, SO2, HC or alkali metals-resistant performance have all atmosphere rather than from isolated single CuII species. As
7
been explored to some extent.[127c] Unlike the above works, the seen in Figure 13c, the oxidized Cu dimer does not form as
8
emphasis of this review is to discuss the underlying prospect of easily when the Cu density is low because Cu is spatially
9
using SACs in the SCR reaction, especially because of their restricted by electrostatic tethering to the framework Al center
10
ability to help elucidate complex catalytic mechanisms and and therefore has limited mobility. This causes the redox cycle
11
guide the development of next-generation SCR SACs. (CuI!CuII) to become the kinetically relevant step under low Cu
12
Metal ions-exchanged zeolites (Cu2 + and Fe3 +) have densities and the kinetically irrelevant step under high Cu
13
witnessed a great breakthrough in the SCR process, demon- densities; the CuII-dimer-mediated SCR mechanism can be seen
14
strating a broad reaction window as a result of the presence of in Figure 13d. This density-dependent formation of dinuclear
15
highly isolated active sites.[127c] The categories of the exchanged active sites from relatively inactive isolated single atoms is out
16
cations, the Si/Al ratios, and the topologies of the zeolite hosts of the scope of conventional single-atom heterogeneous
17
are all correlated to catalytic activity, stability, and reaction catalysis, where isolated active sites are immobilized and the
18
routes. It has been generally acknowledged that the small-pore reaction rate of each site is independent of its volumetric
19
SSZ-13 and SAPO-34 (chabazite: CHA) zeolites have better density. Therefore, metal-exchanged zeolites cannot be classi-
20
hydrothermal stabilities than medium-pore (ZSM-5: MFI, ferrier- fied as SACs even though the metal ions are often atomically
21
ite: FER) and large-pore zeolites (mordenite: MOR, beta: BEA) dispersed on supports before the reaction.
22
under harsh hydrothermal conditions.[140] In Cu/CHA, for exam- In contrast, metal-oxide-supported SACs often exhibit
23
ple, two electron-superfluous Al centers tend to be balanced by single-site behavior, which has been widely investigated in the
24
one discrete CuII ion or two [CuII(OH)] + ions. However, it is still CO oxidation field.[28b,143] Before discussing the SACs, we will first
25
not clear whether the real active Cu site is monomeric, dimeric, mention the widely reported metal oxide catalysts used in the
26
or small oligomeric clusters, and the catalytic mechanism is also SCR reaction. The V2O5-based materials typically include
27
widely debated.[137b] In 2017, Paolucci et al. systematically V2O5 TiO2 and V2O5 MOx TiO2 (M: W, Mn, Fe, etc.), and the
28
discussed the Cu active state in a standard SCR reaction by acid-redox dual site proposed by Topsøe et al. is widely
29
regulating the Cu volumetric density from low to high (also accepted as the catalytic active site.[144] However, it is still
30
referred to as the high-to-low average Cu Cu distance in debated whether the SCR reaction proceeds via the mono-
31
Figure 13a,b).[141] It was found that the SCR rate has a linear nuclear V or the dinuclear V2 sites.[145] This disagreement is
32
mainly ascribed to the difficulty of precisely designing uniform
33
mononuclear or dinuclear sites. Other types of metal oxides,
34
including single MnOx,[146] mixed metal (Mn, Fe, Ce, Cu, W, Ti,
35
Sn, Nb) oxides,[146–147] and the TiMOx (Cu, Fe, Co, Mn, Cr), CeWOx,
36
and CeZrOx-based solid solution catalysts,[139f,148] all require acid
37
and redox properties. Among these mixed metal oxide catalysts,
38
Mn-, Fe-, Ce-, and Cu-based oxides are usually considered to be
39
the main catalysts, while other metal oxides are the promoters.
40
Although these metal oxide catalysts exhibited promising SCR
41
reaction performance, unraveling the common structural
42
features of their acid-redox sites is still not easy because of the
43
various structures of the active centers. Recently, Tang and
44
Chen’s group designed a single-site Mo1/Fe2O3 catalyst where a
45
uniform Mo Fe dinuclear site was assembled with one discrete
46
acidic Mo ion and its neighboring redox Fe ion.[145] The
47
characterization and atomic-scale structure of the Mo1/Fe2O3
48
catalyst can be seen in Figure 14. The calculated TOF (NO
49
conversion amount per second per Mo Fe dinuclear site) on
50 Figure 13. (a) The CHA cage (left) and representative high (top right) and the Mo1/Fe2O3 catalyst is ~ 1.7 × 10 3 s 1 at 270 °C, which is
51 low (bottom right) Cu volumetric densities; (b) Standard SCR rates over Cu/
CHA catalysts with different Cu loadings; (c) Snapshots obtained from comparable to the TOFs of 1.34 × 10 3 s 1 and 4.80 × 10 3 s 1
52
simulated initial (time = 0) and final (time!1) CuI geometric distributions (NO conversion amount per second per V site at 277 °C) over
53 with respect to low, medium, and high Cu density; (d) the proposed low- commercial 0.8 wt% V2O5/TiO2 and 1.5 wt% V2O5/TiO2,
54 temperature SCR mechanism over Cu/CHA; the detailed experiment and
simulation conditions can be seen here. Adapted with permission from ref. respectively.[149] Interestingly, SCR rates were found to in-
55
[141]. Copyright 2017 American Association for the Advancement of crcxease linearly as the number of Mo Fe dinuclear sites
56 Science. increased (by increasing the Mo loading), clearly revealing the
57
ChemNanoMat 2020, 6, 1 – 25 www.chemnanomat.org 18 © 2020 Wiley-VCH GmbH

These are not the final page numbers! ��


Minireview

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
Figure 14. (a,b) TEM; (c–f) Energy dispersive X-ray spectroscopy mapping; (g) AC-STEM images of the Mo1/Fe2O3 catalyst; (h) Surface intensity plot and
26 simulated structural model of the Mo1/Fe2O3 catalyst in the chosen area in g. Adapted from ref. [145] which is an open access article. Copyright 2020 Springer
27 Nature Limited.
28
29
30
characteristics of SACs that are different from those of the significantly reduces the cost. However, the reaction temper-
31
abovementioned V2O5/TiO2 catalyst. Furthermore, the SCR rate ature is still too high to be used in diesel engines even though
32
can be adjusted by selectively altering the acid or redox sites, the presence of H2 can increase the activity of low-temperature
33
which provides a more comprehensive identification of the SCR HC SCR.[133] It has been reported that conventionally used Ag/
34
active site. The discrepancy in activity that results from Al2O3 includes various Ag species, including metallic Ag clusters
35
modifying acid-redox properties (Mo Fe, W Fe, and Fe W) in and ionic Ag (Ag + ions and/or Agδ + clusters) species.[154] The
36
SACs[145] seems to explain some previously observed experimen- oxidized Ag species were regarded as the active sites in some
37
tal phenomena, including (1) some dimeric V sites have a higher works, but several studies also suggested the importance of
38
SCR rate than monomeric V sites;[150] (2) the promotional effect Al2O3 supports.[153] Like NH3 SCR, the reaction mechanism of
39
of WO3 or MoO3 in the classical V2O5/TiO2 catalyst is due to the HC SCR on Ag/Al2O3 is also controversial because of the
40
possible partial formation of dinuclear W V or Mo V sites;[151] complex compositions of Ag/Al2O3.[153] In 2020, Wang et al. used
41
and (3) the SCR reaction preferentially proceeds at acid-redox the nanosized γ-Al2O3 with rich terminal hydroxyls to anchor
42
oxide interfaces (Mo Fe, W Fe, W Ce oxides, etc.).[152] The the Ag atoms, and the Ag clusters were instead observed on
43
application of SACs in the SCR reaction is still uncommon, but the microsized γ-Al2O3.[155] The authors used the surface-science
44
the further study of this uniform dinuclear structure by method and DFT calculations to explain the formation of single-
45
selectively tailoring acid-redox sites would help researchers atom Ag on nanosized γ-Al2O3. They also reported that single-
46
understand the SCR reaction mechanism, cast light on the atom Ag demonstrated better catalytic performance than its Ag
47
nature of the SCR active site, and propose reasonable protocols cluster counterpart with the same Ag loading. However, the
48
to design more attractive SACs. Furthermore, the role of a investigation of single-site catalytic behavior and the explora-
49
variety of multiple single sites, such as the ternary tion of the active site were seldom discussed even though the
50
Macid Mredox Macid or Mredox Macid Mredox sites, in the SCR reaction uniform single atom structure was obtained. Therefore, the
51
is worth studying in the future. study of SACs in HC SCR is just at its beginning; more
52
In addition to NH3 being widely accepted as the reducing comprehensive work is urgently required to better understand
53
agent, Al2O3-supported Ag is largely understood as the most the catalytic active center of Ag/Al2O3 for HC SCR.
54
efficient catalyst in the HC SCR reaction even in the presence The application of SACs on CO SCR has also been explored,
55
of H2O vapor and SO2.[153] The major advantage of HC SCR is mainly focusing on Pt- and Rh-based catalysts. Fernandez
56
the use of on-board fuel as the reducing agent, which et al.[156] systematically compared low-temperature CO SCR
57
ChemNanoMat 2020, 6, 1 – 25 www.chemnanomat.org 19 © 2020 Wiley-VCH GmbH

These are not the final page numbers! ��


Minireview
over MCM-22 supported Pt species ranging from single Pt 6. Conclusions and perspectives
1
atoms to Pt clusters (0.2–0.7 nm), and Pt NPs (0.7–3 nm). By
2
coupling operando IR technique and DFT calculation results, In this review, recent developments in automobile exhaust
3
they claimed that single Pt atom is inactive for NO dissociation SACs were summarized and classified according to types of
4
and Pt NPs afford too much strong adsorption towards CO, thus reactions and catalysts. The general principle of catalyst design
5
contributing to a low CO SCR reactivity. In contrast, the and the fundamental reaction mechanism of the automobile
6
outstanding catalytic performance of Pt clusters is attributed to exhaust SACs were also discussed. We also highlight the unique
7
the fact that they allow the low-temperature dissociation of NO features of SACs during CO oxidation, HC oxidation, and SCR.
8
and provide a relatively weak interaction with CO to avoid For CO oxidation, the catalytic performance of SACs can be
9
poisoning. Also, single Pt atoms are not stable under CO SCR promoted with several different strategies. For example, the
10
reaction condition and can be readily transformed into more activity of Pt SACs was reportedly promoted by steam,[13c] CO,[34]
11
active Pt clusters. A similar result was also reported for the high-temperature H2 treatment,[12a] as well as the co-feeding of
12
atomically dispersed Rh(CO)2 on HY zeolites, in which mono- water.[35–36] The activity of Au SACs can be promoted by
13
nuclear rhodium species is not capable of dissociating/ optimizing the charge densities and the local environment.[84,87]
14
recombining of NO molecules.[134] They believed that it is the These strategies can potentially be generalized to other SACs as
15
aggregates of Rh atoms rather than the mononuclear single-site well. Furthermore, the reaction mechanism of CO oxidation
16
Rh(CO)2 species that facilitates the NO reduction with CO. over SACs is very sensitive to their metal center and local
17
Besides, single-atom Rh on inert SiO2 is confirmed by Zhang environment (i. e., metal-support interactions). Therefore, a
18
et al.[157] to be active to reduce the NO with CO via the totally comprehensive understanding of the origin of the activity is
19
different reaction pathways compared to Rh NPs. Very recently, required to optimize the catalyst for better catalytic perform-
20
Christopher and co-workers raised a question if the low- ance. In addition, extensive ab initio quantum chemical
21
temperature NH3 formation from NO reduction in the presence calculations were performed to predict which SACs were active
22
of CO and H2O is due to single atom Rh?[158] They provided the and to investigate the influence of the metal properties.
23
evidence showing that single atom Rh on γ-Al2O3 is selective for Interestingly, some non-noble metal SACs such as Ni, Co, and
24
the low-temperature unselective reduction (USCR) of NO to Cu SACs were predicted to be highly active SACs with similar or
25
NH3, whereas the Rh clusters are responsible for SCR process to even higher TOFs than current noble metal SACs.
26
N2 (Figure 15). As a result, avoiding the formation of single The study of SACs for HC oxidation is a newly emerging
27
atom Rh is recommended for the SCR process. In contrast, research focus. Currently, most of the reported works deal with
28
isolated bimetallic Rh1Co3 site was reported to have a better aromatics, while less attention has been paid to light HCs and
29
performance than closely packed Rh Co alloy NPs towards low- oxygenated HCs. Pt SACs predominant in current studies, while
30
temperature reduction of NO with CO due to its low-barrier NO some other metals, including Pd, Ag, and even Na, also have
31
dissociation route.[159] Overall, the role of single metal atoms on the potential to be active SACs for HC oxidation. The activities
32
CO SCR reaction is still not very clear up to present. A deeper of SACs are highly affected by the properties of the support. A
33
yet fundamental understanding is highly urgent regardless of wide range of metal oxides have been investigated, including
34
whether they have a promotional effect on reactivity and reducible metal oxides (e. g., TiO2, Fe2O3, Mn2O3, and CeO2), inert
35
stability. oxides (e. g., SiO2 ZrO2 and Al2O3), and alkali earth metal oxides
36
(e. g., MgO and HMO). On selected metal oxide supports, SACs
37
show similar or higher activity than their nanoparticle counter-
38
parts for HC oxidation. The strong interaction between the
39
metal single atoms can facilitate HC oxidation via several
40
different pathways, including creating O vacancies, generating
41 *
reactive OH radicals, or modifying the electron state of the
42
support.
43
The use of SACs in SCR, which includes Mo1/Fe2O3 for
44
NH3 SCR, Ag1/γ-Al2O3 for HC SCR, and Pt/ Rh catalysts for
45
CO SCR etc., is still at the beginning stage. The Mo1/Fe2O3 study
46
demonstrated that the SCR rate linearly increase with an
47
increasing quantity of the dual Mo Fe site; the rate can also be
48
adjusted by selectively tuning the acid and/or redox centers
49
(W Fe and Fe W). Although the catalytic activity of these SACs
50
must be further improved, their emergences provides a chance
51
for us to precisely regulate uniform active centers, which is very
52
helpful for understanding catalytic reaction mechanisms. Addi-
53
tionally, Ag1/γ-Al2O3 exhibits better catalytic performance than
54
Figure 15. Illustration of the NO reduction over different Rh structures on its Ag cluster counterparts for HC SCR, but the underlying
55
oxide supports between 150 °C and 300 °C. Reproduced with permission structure-reactivity relationship is not clear. As a result, more
56 from ref. [158], Copyright 2020 American Chemical Society. comprehensive work is urgently needed to better understand
57
ChemNanoMat 2020, 6, 1 – 25 www.chemnanomat.org 20 © 2020 Wiley-VCH GmbH

These are not the final page numbers! ��


Minireview
the catalytic active center of Ag/Al2O3 for HC SCR. The role of including the ternary Macid Mredox Macid or Mredox Macid Mredox
1
single metal atoms on CO SCR is sensitive to the catalyst sites, is worth studying in the future.
2
structure and a more general and fundamental understanding The three-way SAC is also an emerging area of research due
3
is required. Overall, the SACs would provide more opportunities to the urgent need in reduction of precious group metals. A
4
to construct uniform single, dual, or even multiple single-atom recent study showed that defective ceria-supported atomically
5
sites, explore their unique structure-performance relationships, dispersed metal (Pt, Pd and Rh) ensemble catalysts are more
6
and build a bridge between SCR and SACs to further promote active as isolated metal SACs than TWCs.[160] However, the
7
the new industrial SCR technique. reactivity of the isolated SACs may be promoted by tailoring
8
SACs provide numerous opportunities for designing highly their local environments or by using different metal centers.
9
active catalysts for the oxidation of CO and HCs. Although CO Also, different SACs can be applied for both oxidation and
10
oxidation has been extensively studied on SACs, there are still reduction.
11
many fundamental questions that must be answered and
12
problems that need to be solved. First, more systematic studies
13
are required to find more general descriptors for predicting Acknowledgements
14
reaction mechanisms and reactivities. Many strategies have
15
been developed for promoting the reactivities of SACs, but it is This material is based upon work supported by the US Depart-
16
important to understand their underlying mechanisms and ment of Energy (DOE), Energy Efficiency and Renewable Energy,
17
limitations. Second, the dynamic structures of different types of Vehicle Technologies Office. H.W. acknowledges the support
18
SACs under reaction conditions are still not clear; however, from the U.S. Department of Energy (DOE), Office of Basic
19
these structures can be studied experimentally using in situ and Energy Sciences, Division of Chemical Sciences, Geosciences,
20
operando techniques. Third, more thermally stable and higher- and Biosciences (DE-AC05-RL01830), and F.L. would like to
21
loading SACs are the future direction for more practical SACs, thank the financial support by the U.S. Department of Energy
22
especially on non-defect or nonreducible supports. Fourth, non- (DOE) Bioenergy Technologies Office (BETO) and Vehicle
23
noble metal SACs are promising and highly cost-effective Technologies Office (VTO) under the DOE Co-Optimization of
24
materials, and it is important to optimize their structures for Fuels and Engines Initiative.
25
better catalytic performance.
26
Compared to CO oxidation, the catalytic oxidation of HCs
27
are more complex reactions that require the activation of both Conflict of Interest
28
the O2 and HC molecules. Depending on the functional groups
29
present, the specific catalytic requirements for HC oxidation The authors declare no conflict of interest.
30
varies. More mechanistic insights into the activation of HC
31
molecules on the SACs are still needed to guide the design of Keywords: Single-atom catalyst · Vehicle emission control · CO
32
the catalyst. Among the HCs, alkanes (e. g., methane and oxidation · Hydrocarbon oxidation · Selective catalytic reduction
33
propane) are the most difficult to oxidize. So far, there has not
34
been an SAC reported that shows excellent activity for alkane
35
oxidation; this is likely due to the difficulty of activating the
36
C H bond in an alkane. In terms of catalyst design, in addition
37 [1] G. C. Koltsakis, A. M. Stamatelos, Prog. Energy Combust. Sci. 1997, 23, 1–
to developing active single-atom catalytic sites, the combina- 39.
38
tion of metal nanoparticles and single atoms could be a [2] T. V. Johnson, SAE Int. J. Engines 2015, 8, 1152–1167.
39 [3] M. Zammit, C. DiMaggio, C. Kim, C. Lambert, G. Muntean, C. Peden, J.
promising approach for designing catalysts that are compatible
40 Parks, K. Howden, Future Automotive Aftertreatment Solutions: The
with a wider range of HCs. 150 °C Challenge Workshop Report, United States Department of Energy
41
Unlike the oxidation reactions of CO and HCs, the SCR 2017.
42 [4] H. Hirata, Catal. Surv. Asia 2014, 18, 128–133.
technique usually requires a reducing agent and proceeds
43 [5] H.-Y. Chen, S. Mulla, E. Weigert, K. Camm, T. Ballinger, J. Cox, P.
through a completely different mechanism. The current Blakeman, SAE Int. J. Fuels Lubr. 2013, 6, 372–381.
44
challenge in the SCR catalytic system is how to clearly identify [6] a) E. J. Peterson, A. T. De La Riva, S. Lin, R. S. Johnson, H. Guo, J. T.
45 Miller, J. Hun Kwak, C. H. Peden, B. Kiefer, L. F. Allard, F. H. Ribeiro, A. K.
the catalytic active center and guide the development of the
46 Datye, Nat. Commun. 2014, 5, 4885; b) R. M. Heck, R. J. Farrauto, Appl.
next generation of SCR catalysts. The main characteristics of Catal. A 2001, 221, 443–457.
47
next-generation SCR catalysts include low cost, high deNOx [7] a) S. Dey, G. C. Dhal, Mater. Sci. Energy Technol. 2019, 2, 607–623; b) P.-
48 A. Carlsson, M. Skoglundh, P. Thormählen, B. Andersson, Top. Catal.
efficiency, a broad operating temperature window, impressive
49 2004, 30, 375–381.
high-temperature hydrothermal stability, and excellent poison- [8] a) C. R. Yin, F. R. Negreiros, G. Barcaro, A. Beniya, L. Sementa, E. C. Tyo,
50
resistant performance, etc. Taking NH3 SCR as an example, the S. Bartling, K. H. Meiwes-Broer, S. Seifert, H. Hirata, N. Isomura, S.
51 Nigam, C. Majumder, Y. Watanabe, A. Fortunelli, S. Vajda, J. Mater.
acid-redox properties and local electronic structure of the
52 Chem. A 2017, 5, 4923–4931; b) H. Wang, J.-X. Liu, L. F. Allard, S. Lee, J.
current dinuclear Moacid Feredox catalyst can be adjusted by Liu, H. Li, J. Wang, J. Wang, S. H. Oh, W. Li, M. Flytzani-Stephanopoulos,
53
substituting the Mo and/or Fe sites with other metals or non- M. Shen, B. R. Goldsmith, M. Yang, Nat. Commun. 2019, 10, 3808.
54 [9] a) X. F. Yang, A. Q. Wang, B. T. Qiao, J. Li, J. Y. Liu, T. Zhang, Acc. Chem.
metals, thereby possibly providing a more promising SAC
55 Res. 2013, 46, 1740–1748; b) L. C. Liu, A. Corma, Chem. Rev. 2018, 118,
candidate. In addition, a variety of multiple single sites, 4981–5079.
56
[10] B. Roldan Cuenya, F. Behafarid, Surf. Sci. Rep. 2015, 70, 135–187.
57
ChemNanoMat 2020, 6, 1 – 25 www.chemnanomat.org 21 © 2020 Wiley-VCH GmbH

These are not the final page numbers! ��


Minireview
[11] a) M. J. Hulsey, B. Zhang, Z. Ma, H. Asakura, D. A. Do, W. Chen, T. [29] T. Montini, M. Melchionna, M. Monai, P. Fornasiero, Chem. Rev. 2016,
1 Tanaka, P. Zhang, Z. Wu, N. Yan, Nat. Commun. 2019, 10, 1330; b) J. Lu, 116, 5987–6041.
2 P. Serna, C. Aydin, N. D. Browning, B. C. Gates, J. Am. Chem. Soc. 2011, [30] a) J. W. Qin, J. F. Lu, M. H. Cao, C. W. Hu, Nanoscale 2010, 2, 2739–2743;
3 133, 16186–16195; c) J. Liu, ACS Catal. 2017, 7, 34–59. b) Z. P. Hu, H. Metiu, J. Phys. Chem. C 2011, 115, 17898–17909.
[12] a) L. De Rita, J. Resasco, S. Dai, A. Boubnov, H. V. Thang, A. S. Hoffman, [31] a) Y. B. Lu, C. Thompson, D. Kunwar, A. K. Datye, A. M. Karim,
4 I. Ro, G. W. Graham, S. R. Bare, G. Pacchioni, X. Pan, P. Christopher, Nat. ChemCatChem 2020, 12, 1726–1733; b) A. M. Ganzler, M. Casapu, F.
5 Mater. 2019, 18, 746–751; b) H. V. Thang, G. Pacchioni, L. De Rita, P. Maurer, H. Stormer, D. Gerthsen, G. Ferre, P. Vernaux, B. Bornmann, R.
6 Christopher, J. Catal. 2018, 367, 104–114. Frahm, V. Murzin, M. Nachtegaal, M. Votsmeier, J. D. Grunwaldt, ACS
[13] a) Y. Lu, J. Wang, L. Yu, L. Kovarik, X. Zhang, A. S. Hoffman, A. Gallo, Catal. 2018, 8, 4800–4811; c) A. M. Ganzler, M. Casapu, P. Vernoux, S.
7 S. R. Bare, D. Sokaras, T. Kroll, V. Dagle, H. Xin, A. M. Karim, Nat. Can. Loridant, F. J. C. S. Aires, T. Epicier, B. Betz, R. Hoyer, J. D. Grunwaldt,
8 2019, 2, 149–156; b) B. T. Qiao, A. Q. Wang, X. F. Yang, L. F. Allard, Z. Angew. Chem. Int. Ed. 2017, 56, 13078–13082; Angew. Chem. 2017, 129,
9 Jiang, Y. T. Cui, J. Y. Liu, J. Li, T. Zhang, Nat. Chem. 2011, 3, 634–641; 13258–13262.
c) L. Nie, D. H. Mei, H. F. Xiong, B. Peng, Z. B. Ken, X. I. Pereira- [32] A. Bruix, Y. Lykhach, I. Matolinova, A. Neitzel, T. Skala, N. Tsud, M.
10 Hernández, A. De La Riva, M. Wang, M. H. Engelhard, L. Kovarik, A. K. Vorokhta, V. Stetsovych, K. Sevcikova, J. Myslivecek, R. Fiala, M. Vaclavu,
11 Datye, Y. Wang, Science 2017, 358, 1419–1423. K. C. Prince, S. Bruyere, V. Potin, F. Illas, V. Matolin, J. Libuda, K. M.
12 [14] a) J. Jones, H. F. Xiong, A. T. De La Riva, E. J. Peterson, H. Pham, S. R. Neyman, Angew. Chem. Int. Ed. Engl. 2014, 53, 10525–10530.
Challa, G. S. Qi, S. Oh, M. H. Wiebenga, X. I. Pereira-Hernández, Y. [33] M. Kottwitz, Y. Li, R. M. Palomino, Z. Liu, G. Wang, Q. Wu, J. Huang, J.
13 Wang, A. K. Datye, Science 2016, 353, 150–154; b) W. G. Liu, Y. J. Chen, Timoshenko, S. D. Senanayake, M. Balasubramanian, D. Lu, R. G. Nuzzo,
14 H. F. Qi, L. L. Zhang, W. S. Yan, X. Y. Liu, X. F. Yang, S. Miao, W. T. Wang, A. I. Frenkel, ACS Catal. 2019.
15 C. G. Liu, A. Q. Wang, J. Li, T. Zhang, Angew. Chem. Int. Ed. 2018, 57, [34] X. I. Pereira-Hernández, A. De La Riva, V. Muravev, D. Kunwar, H. Xiong,
7071–7075; Angew. Chem. 2018, 130, 7189–7193; c) J. Resasco, L. B. Sudduth, M. Engelhard, L. Kovarik, E. J. M. Hensen, Y. Wang, A. K.
16 De Rita, S. Dai, J. P. Chada, M. J. Xu, X. X. Yan, J. Finzel, S. Hanukovich, Datye, Nat. Commun. 2019, 10, 1358.
17 A. S. Hoffman, G. W. Graham, S. R. Bare, X. Q. Pan, P. Christopher, J. Am. [35] C. L. Wang, X. K. Gu, H. Yan, Y. Lin, J. J. Li, D. D. Liu, W. X. Li, J. L. Lu, ACS
18 Chem. Soc. 2020, 142, 169–184. Catal. 2017, 7, 887–891.
[15] A. Russell, W. S. Epling, Catal. Rev. 2011, 53, 337–423. [36] T. Wang, J. Y. Xing, A. P. Jia, C. Tang, Y. J. Wang, M. F. Luo, J. Q. Lu, J.
19
[16] a) T. Maunula, A. Suopanki, K. Torkkell, M. Härkönen, SAE Trans. 2004, Catal. 2020, 382, 192–203.
20 113, 1963–1974; b) A. Winkler, D. Ferri, M. Aguirre, Appl. Catal. B 2009, [37] M. Yoo, Y. S. Yu, H. Ha, S. Lee, J. S. Choi, S. Oh, E. Kang, H. Choi, H. An,
21 93, 177–184; c) A. Valiheikki, M. Karkkainen, M. Honkanen, O. K. S. Lee, J. Y. Park, R. Celestre, M. A. Marcus, K. Nowrouzi, D. Taube,
Heikkinen, T. Kolli, K. Kallinen, M. Huuhtanen, M. Vippola, J. Lahtinen, D. A. Shapiro, W. Jung, C. Kim, H. Y. Kim, Energy Environ. Sci. 2020, 13,
22
R. L. Keiski, Appl. Catal. B 2017, 218, 409–419; d) M. R. Ward, T. Hyde, 1231–1239.
23 E. D. Boyes, P. L. Gai, ChemCatChem 2012, 4, 1622–1631; e) M. [38] D. Kunwar, S. Zhou, A. De La Riva, E. Peterson, H. Xiong, X. I. Pereira-
24 Skoglundh, L. O. Lowendahl, J. E. Ottersted, Appl. Catal. 1991, 77, 9–20. Hernández, S. C. Purdy, R. ter Veen, H. H. Brongersma, J. T. Miller, H.
[17] a) S. Mohan, P. Dinesha, S. Kumar, Chem. Eng. J. 2020, 384; b) S. S. R. Hashiguchi, L. Kovarik, S. Lin, H. Guo, Y. Wang, A. Datye, ACS Catal.
25
Putluru, A. Riisager, R. Fehrmann, Appl. Catal. B 2010, 97, 333–339; 2019.
26 c) W. Weisweiler, F. Buchholz, Chem. Ing. Tech. 2001, 73, 882–887. [39] N. Li, Q. Y. Chen, L. F. Luo, W. X. Huang, M. F. Luo, G. S. Hu, J. Q. Lu,
27 [18] a) R. Brosius, J. A. Martens, Top. Catal. 2004, 28, 119–130; b) V. Houel, P. Appl. Catal. B 2013, 142, 523–532.
28 Millington, R. Rajaram, A. Tsolakis, Appl. Catal. B 2007, 73, 203–207; [40] C. T. Kuo, Y. B. Lu, L. Kovarik, M. Engelhard, A. M. Karim, ACS Catal.
c) B. Sawatmongkhon, A. Tsolakis, S. Sitshebo, J. Rodriguez-Fernandez, 2019, 9, 11030–11041.
29 M. Ahmadinejad, J. Collier, R. R. Rajaram, Appl. Catal. B 2010, 97, 373– [41] Z. Zhang, Y. Zhu, H. Asakura, B. Zhang, J. Zhang, M. Zhou, Y. Han, T.
30 380. Tanaka, A. Wang, T. Zhang, N. Yan, Nat. Commun. 2017, 8, 16100.
31 [19] a) B. K. Cho, J. H. Lee, C. C. Crellin, K. L. Olson, D. L. Hilden, M. K. Kim, [42] R. Lang, W. Xi, J. C. Liu, Y. T. Cui, T. B. Li, A. F. Lee, F. Chen, Y. Chen, L.
P. S. Kim, I. Heo, S. H. Oh, I. S. Nam, Catal. Today 2012, 191, 20–24; Li, L. Li, J. Lin, S. Miao, X. Y. Liu, A. Q. Wang, X. D. Wang, J. Luo, B. T.
32 b) D. H. Lee, J. O. Lee, K. T. Kim, Y. H. Song, E. Kim, H. S. Han, Int. J. Qiao, J. Li, T. Zhang, Nat. Commun. 2019, 10.
33 Hydrogen Energy 2011, 36, 11718–11726; c) Q. F. Su, H. Pan, J. Chen, [43] a) F. Y. Li, Y. F. Li, X. C. Zeng, Z. F. Chen, ACS Catal. 2015, 5, 544–552;
34 Abstr. Pap. Am. Chem. Soc. 2009, 237. b) G. Liu, A. G. Walsh, P. Zhang, J. Phys. Chem. Lett. 2020, 11, 2219–
[20] I. Nova, D. Bounechada, R. Maestri, E. Tronconi, A. K. Heibel, T. A. 2229.
35 Collins, T. Boger, Ind. Eng. Chem. Res. 2011, 50, 299–309. [44] Y. X. Feng, Q. Wang, H. F. Xiong, S. L. Zhou, X. Chen, X. I. Pereira-
36 [21] a) T. H. Vuong, J. Radnik, E. Kondratenko, M. Schneider, U. Armbruster, Hernández, Y. Wang, S. Lin, A. K. Datye, H. Guo, J. Phys. Chem. C 2018,
37 A. Brückner, Appl. Catal. B 2016, 197, 159–167; b) X. Zhang, J. Wang, Z. 122, 22460–22468.
Song, H. Zhao, Y. Xing, M. Zhao, J. Zhao, Z. a Ma, P. Zhang, N. Tsubaki, [45] a) M. S. Chen, Y. Cai, Z. Yan, K. K. Gath, S. Axnanda, D. W. Goodman,
38 Mol. Cancer 2019, 463, 1–7. Surf. Sci. 2007, 601, 5326–5331; b) Y. Tang, Z. Yang, X. Dai, Phys. Chem.
39 [22] a) X. Wang, Y. Liu, Q. Ying, W. Yao, Z. Wu, Appl. Catal. A 2018, 562, 19– Chem. Phys. 2012, 14, 16566–16572.
40 27; b) H. Chang, J. Li, J. Yuan, L. Chen, Y. Dai, H. Arandiyan, J. Xu, J. [46] H. W. Gao, Appl. Surf. Sci. 2016, 379, 347–357.
Hao, Catal. Today 2013, 201, 139–144. [47] Y. Wang, X. L. Zhang, C. Cheng, Z. X. Yang, Appl. Surf. Sci. 2018, 453,
41 [23] a) G. Fulks, G. B. Fisher, K. Rahmoeller, M.-C. Wu, E. D’Herde, J. Tan, SAE 159–165.
42 Int. 2009; b) C. Gerhart, H.-P. Krimmer, B. Hammer, B. Schulz, O. [48] M. A. Newton, D. Ferri, G. Smolentsev, V. Marchionni, M. Nachtegaal,
43 Kröcher, D. Peitz, T. Sattelmayer, P. Toshev, G. Wachtmeister, A. Nat. Commun. 2015, 6, 8675.
Heubuch, E. Jacob, SAE. Int. J. Engines 2012, 5, 938–946. [49] C. M. Du, H. P. Lin, B. Lin, Z. Y. Ma, T. J. Hou, J. X. Tang, Y. Y. Li, J. Mater.
44
[24] a) P. Brijesh, S. Sreedhara, Int. J. Auto. Tech.-Kor. 2013, 14, 195–206; Chem. A 2015, 3, 23113–23119.
45 b) R. J. Farrauto, M. Deeba, S. Alerasool, Nat. Can. 2019, 2, 603–613. [50] a) X. N. Li, L. N. Wang, L. H. Mou, S. G. He, J. Phys. Chem. A 2019, 123,
46 [25] a) A. Boubnov, S. Dahl, E. Johnson, A. P. Molina, S. B. Simonsen, F. M. 9257–9267; b) X. N. Li, Z. Yuan, J. H. Meng, Z. Y. Li, S. G. He, J. Phys.
Cano, S. Helveg, L. J. Lemus-Yegres, J. D. Grunwaldt, Appl. Catal. B Chem. C 2015, 119, 15414–15420; c) Z. Y. Li, Z. Yuan, X. N. Li, Y. X. Zhao,
47
2012, 126, 315–325; b) A. D. Allian, K. Takanabe, K. L. Fujdala, X. Hao, S. G. He, J. Am. Chem. Soc. 2014, 136, 14307–14313.
48 T. J. Truex, J. Cai, C. Buda, M. Neurock, E. Iglesia, J. Am. Chem. Soc. [51] a) X. L. Zhang, Z. S. Lu, G. L. Xu, T. X. Wang, D. W. Ma, Z. X. Yang, L.
49 2011, 133, 4498–4517; c) M. Cargnello, V. V. T. Doan-Nguyen, T. R. Yang, Phys. Chem. Chem. Phys. 2015, 17, 20006–20013; b) R. Krishnan,
Gordon, R. E. Diaz, E. A. Stach, R. J. Gorte, P. Fornasiero, C. B. Murray, S. Y. Wu, H. T. Chen, Phys. Chem. Chem. Phys. 2019, 21, 12201–12208.
50
Science 2013, 341, 771–773. [52] J. D. Kistler, N. Chotigkrai, P. Xu, B. Enderle, P. Praserthdam, C. Y. Chen,
51 [26] P. J. Berlowitz, C. H. F. Peden, D. W. Goodman, J. Phys. Chem. 1988, 92, N. D. Browning, B. C. Gates, Angew. Chem. Int. Ed. Engl. 2014, 53, 8904–
52 5213–5221. 8907.
53 [27] L. De Rita, S. Dai, K. Lopez-Zepeda, N. Pham, G. W. Graham, X. Q. Pan, [53] a) R. F. Hicks, H. Qi, M. L. Young, R. G. Lee, J. Catal. 1990, 122, 280–294;
P. Christopher, J. Am. Chem. Soc. 2017, 139, 14150–14165. b) I. E. Wachs, R. J. Madix, Surf. Sci. 1978, 76, 531–558.
54 [28] a) M. Moses-DeBusk, M. Yoon, L. F. Allard, D. R. Mullins, Z. L. Wu, X. F. [54] V. Nehasil, I. Stara, V. Matolin, Surf. Sci. 1996, 352, 305–309.
55 Yang, G. Veith, G. M. Stocks, C. K. Narula, J. Am. Chem. Soc. 2013, 135, [55] M. E. Grass, Y. W. Zhang, D. R. Butcher, J. Y. Park, Y. M. Li, H. Bluhm,
56 12634–12645; b) K. Ding, A. Gulec, A. M. Johnson, N. M. Schweitzer, K. M. Bratlie, T. F. Zhang, G. A. Somorjai, Angew. Chem. Int. Ed. 2008, 47,
G. D. Stucky, L. D. Marks, P. C. Stair, Science 2015, 350, 189–192. 8893–8896; Angew. Chem. 2008, 120, 9025–9028.
57
ChemNanoMat 2020, 6, 1 – 25 www.chemnanomat.org 22 © 2020 Wiley-VCH GmbH

These are not the final page numbers! ��


Minireview
[56] a) D. Yang, P. H. Xu, E. J. Guan, N. D. Browning, B. C. Gates, J. Catal. [87] X. Zhou, Q. Shen, K. D. Yuan, W. S. Yang, Q. W. Chen, Z. H. Geng, J. L.
1 2016, 338, 12–20; b) V. Bhirud, J. F. Goellner, A. M. Argo, B. C. Gates, J. Zhang, X. Shao, W. Chen, G. Q. Xu, X. M. Yang, K. Wu, J. Am. Chem. Soc.
2 Phys. Chem. B 2004, 108, 9752–9763; c) A. J. Liang, V. A. Bhirud, J. O. 2018, 140, 554–557.
3 Ehresmann, P. W. Kletnieks, J. F. Haw, B. C. Gates, J. Phys. Chem. B 2005, [88] J. L. Shi, X. J. Zhao, L. Y. Zhang, X. L. Xue, Z. X. Guo, Y. F. Gao, S. F. Li, J.
109, 24236–24243; d) J. O. Ehresmann, P. W. Kletnieks, A. Liang, V. A. Mater. Chem. A 2017, 5, 19316–19322.
4 Bhirud, O. P. Bagatchenko, E. J. Lee, M. Klaric, B. C. Gates, J. F. Haw, [89] W. Y. Song, E. J. M. Hensen, J. Phys. Chem. C 2013, 117, 7721–7726.
5 Angew. Chem. Int. Ed. 2006, 45, 574–576; Angew. Chem. 2006, 118, [90] S. Ali, T. F. Liu, Z. Lian, B. Li, D. S. Su, Phys. Chem. Chem. Phys. 2017, 19,
6 588–590; e) P. Serna, B. C. Gates, J. Catal. 2013, 308, 201–212; f) D. 22344–22354.
Yardimci, P. Serna, B. C. Gates, Chem. Eur. J. 2013, 19, 1235–1245. [91] B. Long, Y. Tang, J. Li, Nano Res. 2016, 9, 3868–3880.
7 [57] a) R. R. Cavanagh, J. T. Yates, J. Chem. Phys. 1981, 74, 4150–4155; b) A. [92] Y. G. Wang, D. Mei, V. A. Glezakou, J. Li, R. Rousseau, Nat. Commun.
8 Cervantes Uribe, G. A. Del Angel Montes, G. Torres-Torres, A. Vázquez- 2015, 6, 6511.
9 Zavala, F. González-García, A. Cordero-García, R. Ojeda-López, J. [93] S. Zhao, F. Chen, S. Duan, B. Shao, T. Li, H. Tang, Q. Lin, J. Zhang, L. Li,
Compos. Sci. 2019, 3. J. Huang, N. Bion, W. Liu, H. Sun, A. Q. Wang, M. Haruta, B. Qiao, J. Li, J.
10 [58] H. Guan, J. Lin, B. Qiao, X. Yang, L. Li, S. Miao, J. Liu, A. Wang, X. Wang, Liu, T. Zhang, Nat. Commun. 2019, 10, 3824.
11 T. Zhang, Angew. Chem. Int. Ed. 2016, 55, 2820–2824; Angew. Chem. [94] Y. L. Shen, K. J. Yin, Z. H. Xiao, Phys. Chem. Chem. Phys. 2019, 21,
12 2016, 128, 2870–2874. 20346–20353.
[59] T. K. Ghosh, N. N. Nair, ChemCatChem 2013, 5, 1811–1821. [95] F. Y. Li, L. Li, X. Y. Liu, X. C. Zeng, Z. F. Chen, ChemPhysChem 2016, 17,
13 [60] B. Han, T. Li, J. Zhang, C. Zeng, H. Matsumoto, Y. Su, B. Qiao, T. Zhang, 3170–3175.
14 Chem. Commun. 2020. [96] J. X. Liang, X. F. Yang, A. Q. Wang, T. Zhang, J. Li, Catal. Sci. Technol.
15 [61] B. Han, R. Lang, H. L. Tang, J. Xu, X. K. Gu, B. T. Qiao, J. Liu, Chin. J. 2016, 6, 6886–6892.
Catal. 2019, 40, 1847–1853. [97] M. Gajdos, J. Hafner, Surf. Sci. 2005, 590, 117–126.
16 [62] a) C. Aydin, J. Lu, N. D. Browning, B. C. Gates, Angew. Chem. Int. Ed. [98] X. P. Zou, L. N. Wang, X. N. Li, Q. Y. Liu, Y. X. Zhao, T. M. Ma, S. G. He,
17 2012, 51, 5929–5934; Angew. Chem. 2012, 124, 6031–6036; b) A. S. Angew. Chem. Int. Ed. Engl. 2018, 57, 10989–10993.
18 Hoffman, C.-Y. Fang, B. C. Gates, J. Phys. Chem. Lett. 2016, 3854–3860. [99] H. X. Xu, C. Q. Xu, D. J. Cheng, J. Li, Catal. Sci. Technol. 2017, 7, 5860–
[63] J. X. Liang, J. Lin, X. F. Yang, A. Q. Wang, B. T. Qiao, J. Y. Liu, T. Zhang, J. 5871.
19
Li, J. Phys. Chem. C 2014, 118, 21945–21951. [100] T. Yang, R. Fukuda, S. Hosokawa, T. Tanaka, S. Sakaki, M. Ehara,
20 [64] J. Lin, Y. Chen, Y. L. Zhou, L. Li, B. T. Qiao, A. Q. Wang, J. Y. Liu, X. D. ChemCatChem 2017, 9, 1222–1229.
21 Wang, T. Zhang, AIChE J. 2017, 63, 4003–4012. [101] a) C. H. Kim, G. S. Qi, K. Dahlberg, W. Li, Science 2010, 327, 1624–1627;
[65] Y. B. Lu, C. T. Kuo, L. Kovarik, A. S. Hoffman, A. Boubnov, D. M. Driscoll, b) K. Simmance, D. Thompsett, W. L. Wang, B. Thiebaut, Catal. Today
22
J. R. Morris, S. R. Bare, A. M. Karim, J. Catal. 2019, 378, 121–130. 2019, 320, 40–50; c) D. Blanck, A. Schon, A. S. Mamede, C. Dujardin,
23 [66] G. L. Xu, R. Wang, Y. C. Ding, Z. S. Lu, D. W. Ma, Z. X. Yang, J. Phys. J. P. Dacquin, P. Granger, J. F. Paul, E. Berrier, Catal. Today 2017, 283,
24 Chem. C 2018, 122, 23481–23492.
151–157.
[67] S. N. Pavlova, V. A. Sadykov, V. A. Razdobarov, E. A. Paukshtis, J. Catal.
25 [102] L. Zhang, I. A. W. Filot, Y. Q. Su, J. X. Liu, E. J. M. Hensen, J. Phys. Chem.
1996, 161, 507–516.
26 C 2019, 123, 7290–7298.
[68] S. N. Pavlova, V. A. Sadykov, N. N. Bulgakov, M. N. Bredikhin, J. Catal.
[103] L. Zhang, Y. Q. Su, M. W. Chang, I. A. W. Filot, E. J. M. Hensen, J. Phys.
27 1996, 161, 517–523.
Chem. C 2019, 123, 31130–31141.
28 [69] W. An, P. Liu, Phys. Chem. Chem. Phys. 2016, 18, 30899–30902.
[104] K. Asakura, H. Nagahiro, N. Ichikuni, Y. Iwasawa, Appl. Catal. A 1999,
[70] W. E. Kaden, T. P. Wu, W. A. Kunkel, S. L. Anderson, Science 2009, 326,
29 188, 313–324.
826–829.
[105] H. Jeong, O. Kwon, B.-S. Kim, J. Bae, S. Shin, H.-E. Kim, J. Kim, H. Lee,
30 [71] S. F. Li, X. J. Zhao, J. L. Shi, Y. Jia, Z. X. Guo, J. H. Cho, Y. F. Gao, Z. Y.
Nat. Can. 2020.
31 Zhang, Phys. Chem. Chem. Phys. 2016, 18, 24872–24879.
[106] E. D. Goodman, A. C. Johnston-Peck, E. M. Dietze, C. J. Wrasman, A. S.
[72] S. Abbet, U. Heiz, H. Hakkinen, U. Landman, Phys. Rev. Lett. 2001, 86,
32 Hoffman, F. Abild-Pedersen, S. R. Bare, P. N. Plessow, M. Cargnello, Nat.
5950–5953.
Can. 2019, 2, 748–755.
33 [73] S. C. Ammal, A. Heyden, ACS Catal. 2017, 7, 301–309.
[74] G. L. Xu, R. Wang, F. Yang, D. W. Ma, Z. X. Yang, Z. S. Lu, Carbon 2017, [107] B. A. Tichenor, M. A. Palazzolo, Environ. Prog. 1987, 6, 172–176.
34 [108] T. V. Choudhary, S. Banerjee, V. R. Choudhary, Appl. Catal. A 2002, 234,
118, 35–42.
35 [75] M. D. Esrafili, S. Asadollahi, ChemistrySelect 2018, 3, 9181–9188. 1–23.
36 [76] M. D. Esrafili, S. Asadollahi, Appl. Surf. Sci. 2019, 463, 526–534. [109] H. Liu, T. Hirota, K. Asakura, Y. Iwasawa, J. Chem. Soc. Faraday Trans.
[77] Y. L. Zhang, R. W. Cattrall, I. D. McKelvie, S. D. Kolev, Gold Bull. 2011, 44, 1998, 94, 2639–2646.
37 [110] S. Hoang, Y. Guo, A. J. Binder, W. Tang, S. Wang, J. Liu, H. Tran, X. Lu, Y.
145–153.
38 [78] a) M. Haruta, T. Kobayashi, H. Sano, N. Yamada, Chem. Lett. 1987, 405– Wang, Y. Ding, E. A. Kyriakidou, J. Yang, T. J. Toops, T. R. Pauly, R.
39 408; b) M. Haruta, N. Yamada, T. Kobayashi, S. Iijima, J. Catal. 1989, Ramprasad, P.-X. Gao, Nat. Commun. 2020, 11, 1062.
115, 301–309; c) M. Haruta, S. Tsubota, T. Kobayashi, H. Kageyama, [111] a) B. Qiao, A. Wang, X. Yang, L. F. Allard, Z. Jiang, Y. Cui, J. Liu, J. Li, T.
40
M. J. Genet, B. Delmon, J. Catal. 1993, 144, 175–192. Zhang, Nat. Chem. 2011, 3, 634–641; b) E. J. Peterson, A. T. De La Riva,
41 [79] a) S. Tsubota, D. Cunningham, Y. Bando, M. Haruta, Stud. Surf. Sci. S. Lin, R. S. Johnson, H. Guo, J. T. Miller, J. Hun Kwak, C. H. F. Peden, B.
42 Catal. 1993, 77, 325–328; b) J. X. Liu, I. A. W. Filot, Y. G. Su, B. Zijlstra, Kiefer, L. F. Allard, F. H. Ribeiro, A. K. Datye, Nat. Commun. 2014, 5,
E. J. M. Hensen, J. Phys. Chem. C 2018, 122, 8327–8340; c) Y. He, J. C. 4885.
43
Liu, L. L. Luo, Y. G. Wang, J. F. Zhu, Y. G. Du, J. Li, S. X. Mao, C. M. Wang, [112] Q. Duan, C. Zhang, S. Sun, Y. Pan, X. Zhou, Y. Liu, K. Chen, C. Li, X.
44 Wang, W. Li, J. Mater. Chem. A 2020, 8, 7395–7404.
Proc. Natl. Acad. Sci. USA 2018, 115, 7700–7705; d) Y. Gao, N. Shao, Y.
45 Pei, Z. F. Chen, X. C. Zeng, ACS Nano 2011, 5, 7818–7829; e) L. Li, Y. [113] D. Yan, J. Chen, H. Jia, Angew. Chem. Int. Ed., 2020. 59, doi: 10.1002/
46 Gao, H. Li, Y. Zhao, Y. Pei, Z. Chen, X. C. Zeng, J. Am. Chem. Soc. 2013, anie.202004929.
135, 19336–19346. [114] H. Jeong, G. Lee, B. S. Kim, J. Bae, J. W. Han, H. Lee, J. Am. Chem. Soc.
47 2018, 140, 9558–9565.
[80] M. Comotti, W.-C. Li, B. Spliethoff, F. Schüth, J. Am. Chem. Soc. 2006,
48 128, 917–924. [115] a) Y. K. Li, Z. Yuan, Y. X. Zhao, C. Y. Zhao, Q. Y. Liu, H. Chen, S. G. He, J.
49 [81] S. S. Lee, C. Y. Fan, T. P. Wu, S. L. Anderson, J. Am. Chem. Soc. 2004, Am. Chem. Soc. 2016, 138, 12854–12860; b) Y. Yang, B. Yang, Y. X.
126, 5682–5683. Zhao, L. X. Jiang, Z. Y. Li, Y. Ren, H. G. Xu, W. J. Zheng, S. G. He, Angew.
50 Chem. Int. Ed. 2019, 58, 17287–17292.
[82] B. T. Qiao, J. Lin, A. Q. Wang, Y. Chen, T. Zhang, J. Y. Liu, Chin. J. Catal.
51 2015, 36, 1505–1511. [116] M. S. Kamal, S. A. Razzak, M. M. Hossain, Atmos. Environ. 2016, 140,
52 [83] B. T. Qiao, J. X. Liang, A. Q. Wang, C. Q. Xu, J. Li, T. Zhang, J. Y. Liu, 117–134.
Nano Res. 2015, 8, 2913–2924. [117] Z. Wang, H. Yang, R. Liu, S. Xie, Y. Liu, H. Dai, H. Huang, J. Deng, J.
53
[84] J. C. Liu, Y. G. Wang, J. Li, J. Am. Chem. Soc. 2017, 139, 6190–6199. Hazard. Mater. 2020, 392, 122258.
54 [85] L. Li, Y. Gao, H. Li, Y. Zhao, Y. Pei, Z. F. Chen, X. C. Zeng, J. Am. Chem. [118] S. Zhao, Y. Wen, X. Liu, X. Pen, F. Lü, F. Gao, X. Xie, C. Du, H. Yi, D.
55 Soc. 2013, 135, 19336–19346. Kang, X. Tang, Nano Res. 2020.
56 [86] C. Schilling, M. Ziemba, C. Hess, M. V. Ganduglia-Pirovano, J. Catal. [119] H. Zhang, S. Sui, X. Zheng, R. Cao, P. Zhang, Appl. Catal. B 2019, 257,
2020, 383, 264–272. 117878.
57
ChemNanoMat 2020, 6, 1 – 25 www.chemnanomat.org 23 © 2020 Wiley-VCH GmbH

These are not the final page numbers! ��


Minireview
[120] Y. Zhang, Y. Liu, S. Xie, H. Huang, G. Guo, H. Dai, J. Deng, Environ. Int. [144] a) X. Zhao, L. Huang, H. Li, H. Hu, J. Han, L. Shi, D. Zhang, Chin. J. Catal.
1 2019, 128, 335–342. 2015, 36, 1886–1899; b) G. Ramis, G. Busca, F. Bregani, P. Forzatti, Appl.
2 [121] K. Yang, Y. Liu, J. Deng, X. Zhao, J. Yang, Z. Han, Z. Hou, H. Dai, Appl. Catal. 1990, 64, 259–278; c) L. Lietti, J. Svachula, P. Forzatti, G. Busca, G.
Catal. B 2019, 244, 650–659. Ramis, P. J. C. t Bregani, Catal. Today 1993, 17, 131–139; d) A.
3
[122] S. Sinha Majumdar, J. A. Pihl, T. J. Toops, Appl. Energy 2019, 255, Marberger, D. Ferri, M. Elsener, O. Kröcher, Angew. Chem. Int. Ed. Engl.
4 113640. 2016, 55, 11989–11994; e) N. Topsoe, H. Topsoe, J. Dumesic, J. Catal.
5 [123] Z. Jiang, X. Feng, J. Deng, C. He, M. Douthwaite, Y. Yu, J. Liu, Z. Hao, Z. 1995, 151, 226–240; f) N. Topsoe, J. Dumesic, H. Topsoe, J. Catal. 1995,
Zhao, Adv. Funct. Mater. 2019, 29, 1902041. 151, 241–252; g) Z. Liu, Y. Li, T. Zhu, H. Su, J. Zhu, Ind. Eng. Chem. Res.
6
[124] a) Z. Huang, X. Gu, Q. Cao, P. Hu, J. Hao, J. Li, X. Tang, Angew. Chem. 2014, 53, 12964–12970.
7 Int. Ed. 2012, 51, 4198–4203; Angew. Chem. 2012, 124, 4274–4279; b) P. [145] W. Qu, X. Liu, J. Chen, Y. Dong, X. Tang, Y. Chen, Nat. Commun. 2020,
8 Hu, Z. Huang, Z. Amghouz, M. Makkee, F. Xu, F. Kapteijn, A. 11, 1–7.
Dikhtiarenko, Y. Chen, X. Gu, X. Tang, Angew. Chem. Int. Ed. 2014, 53, [146] Z.-b. Xiong, J. Liu, F. Zhou, D.-y. Liu, W. Lu, J. Jin, S.-f. Ding, Appl. Surf.
9
3418–3421; Angew. Chem. 2014, 126, 3486–3489. Sci. 2017, 406, 218–225.
10 [125] Y. Chen, J. Gao, Z. Huang, M. Zhou, J. Chen, C. Li, Z. Ma, J. Chen, X. [147] a) L. Chen, J. Li, M. Ge, Environ. Sci. Technol. 2010, 44, 9590–9596; b) Y.
11 Tang, Environ. Sci. Technol. 2017, 51, 7084–7090. Zhang, W. Xiaolei, S. Kai, X. Haitao, S. Keqin, Z. Changcheng, Chin. J.
[126] G. Liu, P.-X. Gao, Catal. Sci. Technol. 2011, 1, 552–568. Catal. 2012, 33, 1523–1531; c) Y. Shu, H. Sun, X. Quan, S. Chen, J. Phys.
12
[127] a) H. Sjövall, L. Olsson, E. Fridell, R. J. Blint, Appl. Catal. B 2006, 64, 180– Chem. C 2012, 116, 25319–25327; d) H. Chang, X. Chen, J. Li, L. Ma, C.
13 188; b) J. H. Kwak, R. G. Tonkyn, D. H. Kim, J. Szanyi, C. H. Peden, J. Wang, C. Liu, J. W. Schwank, J. Hao, Environ. Sci. Technol. 2013, 47,
14 Catal. 2010, 275, 187–190; c) L. Han, S. Cai, M. Gao, J.-y. Hasegawa, P. 5294–5301; e) L. Yan, Y. Liu, K. Zha, H. Li, L. Shi, D. Zhang, Catal. Sci.
Wang, J. Zhang, L. Shi, D. Zhang, Chem. Rev. 2019, 119, 10916–10976. Technol. 2017, 7, 502–514; f) X. Yao, L. Zhang, L. Li, L. Liu, Y. Cao, X.
15
[128] J. A. Sullivan, J. A. Doherty, Appl. Catal. B 2005, 55, 185–194. Dong, F. Gao, Y. Deng, C. Tang, Z. Chen, Appl. Catal. B 2014, 150, 315–
16 [129] C. M. Kalamaras, G. G. Olympiou, V. I. Pârvulescu, B. Cojocaru, A. M. 329.
17 Efstathiou, Appl. Catal. B 2017, 206, 308–318. [148] a) J. Liu, H. Cheng, J. Tan, B. Liu, Z. Zhang, H. Xu, M. Zhao, W. Zhu, J.
[130] A. Patel, T. E. Rufford, V. Rudolph, Z. Zhu, Catal. Today 2011, 166, 188– Liu, Z. Zhao, J. Mater. Chem. A 2020, 8, 6717–6731; b) S. Roy, B.
18
193. Viswanath, M. Hegde, G. Madras, J. Phys. Chem. C 2008, 112, 6002–
19 [131] M. Haneda, Y. Tomida, T. Takahashi, Y. Azuma, T. Fujimoto, Catal. 6012.
20 Commun. 2017, 90, 1–4. [149] L. Lietti, I. Nova, G. Ramis, L. Dall’Acqua, G. Busca, E. Giamello, P.
[132] a) M. V. Twigg, Appl. Catal. B 2007, 70, 2–15; b) S. Mohan, P. Dinesha, S. Forzatti, F. Bregani, J. Catal. 1999, 187, 419–435.
21
Kumar, Chem. Eng. J. 2020, 384, 123253. [150] a) G. He, Z. Lian, Y. Yu, Y. Yang, K. Liu, X. Shi, Z. Yan, W. Shan, H. He, Sci.
22 [133] F. Liu, Y. Yu, H. J. C. C. He, Chem. Commun. 2014, 50, 8445–8463. Adv. 2018, 4, eaau4637; b) G. T. Went, L.-J. Leu, R. R. Rosin, A. T. Bell, J.
23 [134] A. D. Vityuk, S. G. Ma, O. S. Alexeev, M. D. Amiridis, React. Chem. Eng. Catal. 1992, 134, 492–505.
2019, 4, 418–426. [151] a) N. R. Jaegers, J. K. Lai, Y. He, E. Walter, D. A. Dixon, M. Vasiliu, Y.
24
[135] V. I. Pârvulescu, P. Grange, B. Delmon, Catal. Today 1998, 46, 233–316. Chen, C. Wang, M. Y. Hu, K. T. Mueller, Angew. Chem. Int. Ed. Engl.
25 [136] a) F. Gao, X. Tang, H. Yi, S. Zhao, C. Li, J. Li, Y. Shi, X. Meng, J. Catal. 2019, 131, 12739–12746; b) L. Lietti, I. Nova, P. Forzatti, Top. Catal.
26 2017, 7, 199; b) C. Tang, H. Zhang, L. Dong, Catal. Sci. Technol. 2016, 6, 2000, 11, 111–122.
1248–1264; c) R. Q. Long, R. T. Yang, J. Am. Chem. Soc. 1999, 121, [152] a) Y. Xin, N. Zhang, Q. Li, Z. Zhang, X. Cao, L. Zheng, Y. Zeng, J. A.
27
5595–5596; d) K. Rahkamaa-Tolonen, T. Maunula, M. Lomma, M. Anderson, ACS Catal. 2018, 8, 1399–1404; b) Z. Liu, H. Su, B. Chen, J. Li,
28 Huuhtanen, R. L. Keiski, Catal. Today 2005, 100, 217–222. S. I. Woo, Chem. Eng. J. 2016, 299, 255–262; c) J. Chen, Y. Chen, M.
29 [137] a) S. Brandenberger, O. Kröcher, A. Tissler, R. Althoff, Catal. Rev. 2008, Zhou, Z. Huang, J. Gao, Z. Ma, J. Chen, X. Tang, Environ. Sci. Technol.
50, 492–531; b) A. M. Beale, F. Gao, I. Lezcano-Gonzalez, C. H. Peden, J. 2017, 51, 473–478.
30
Szanyi, Chem. Soc. Rev. 2015, 44, 7371–7405. [153] H. Deng, Y. Yu, F. Liu, J. Ma, Y. Zhang, H. He, ACS Catal. 2014, 4, 2776–
31 [138] a) J. Xu, G. Chen, F. Guo, J. Xie, Chem. Eng. J. 2018, 353, 507–518; b) G. 2784.
32 Busca, L. Lietti, G. Ramis, F. Berti, Appl. Catal. B 1998, 18, 1–36. [154] a) M. K. Kim, P. S. Kim, J. H. Baik, I.-S. Nam, B. K. Cho, S. H. Oh, Appl.
[139] a) J. Li, H. Chang, L. Ma, J. Hao, R. T. Yang, Catal. Today 2011, 175, 147– Catal. B 2011, 105, 1–14; b) S. T. Korhonen, A. M. Beale, M. A. Newton,
33
156; b) M. Fu, C. Li, P. Lu, L. Qu, M. Zhang, Y. Zhou, M. Yu, Y. Fang, B. M. Weckhuysen, J. Phys. Chem. C 2011, 115, 885–896.
34 Catal. Sci. Technol. 2014, 4, 14–25; c) C. Gao, J.-W. Shi, Z. Fan, G. Gao, C. [155] F. Wang, J. Ma, S. Xin, Q. Wang, J. Xu, C. Zhang, H. He, X. C. Zeng, Nat.
35 Niu, Catalysts 2018, 8, 11; d) C. Liu, J.-W. Shi, C. Gao, C. Niu, Appl. Catal. Commun. 2020, 11, 1–9.
A 2016, 522, 54–69; e) T. Boningari, P. G. Smirniotis, Curr. Opin. Chem. [156] E. Fernandez, L. C. Liu, M. Boronat, R. Arenal, P. Concepcion, A. Corma,
36
Eng. 2016, 13, 133–141; f) J. Liu, Z. Zhao, C. Xu, J. Liu, Chin. J. Catal. ACS Catal. 2019, 9, 11530–11541.
37 2019, 40, 1438–1487. [157] S. R. Zhang, Y. Tang, L. Nguyen, Y. F. Zhao, Z. L. Wu, T. W. Goh, J. J. Y.
38 [140] a) A. Sultana, T. Nanba, M. Sasaki, M. Haneda, K. Suzuki, H. Hamada, Liu, Y. Y. Li, T. Zhu, W. Y. Huang, A. I. Frenkel, J. Li, F. F. Tao, ACS Catal.
Catal. Today 2011, 164, 495–499; b) B. Palella, M. Cadoni, A. Frache, H. 2018, 8, 110–121.
39
Pastore, R. Pirone, G. Russo, S. Coluccia, L. Marchese, J. Catal. 2003, [158] C. Asokan, Y. Yang, A. Dang, A. Getsoian, P. Christopher, ACS Catal.
40 217, 100–106; c) Q. Ye, L. Wang, R. T. J. A. C. A. G. Yang, Appl. Catal. A 2020, 10, 5217–5222.
41 2012, 427, 24–34; d) J. H. Kwak, D. Tran, S. D. Burton, J. Szanyi, J. H. Lee, [159] S. R. Zhang, L. Nguyen, J. X. Liang, J. J. Shan, J. Y. Liu, A. I. Frenkel, A.
C. H. Peden, J. Catal. 2012, 287, 203–209. Patlolla, W. X. Huang, J. Li, F. Tao, Nat. Commun. 2015, 6.
42
[141] C. Paolucci, I. Khurana, A. A. Parekh, S. Li, A. J. Shih, H. Li, J. R. Di Iorio, [160] H. Jeong, O. Kwon, B. S. Kim, J. Bae, S. Shin, H. E. Kim, J. Kim, H. J. Lee,
43 J. D. Albarracin-Caballero, A. Yezerets, J. T. Miller, Science 2017, 357, Nat. Can. 2020, 3, 368–375.
44 898–903.
[142] a) S. A. Bates, A. A. Verma, C. Paolucci, A. A. Parekh, T. Anggara, A.
45
Yezerets, W. F. Schneider, J. T. Miller, W. N. Delgass, F. H. Ribeiro, J.
46 Catal. 2014, 312, 87–97; b) F. Gao, E. D. Walter, M. Kollar, Y. Wang, J.
47 Szanyi, C. H. Peden, J. Catal. 2014, 319, 1–14. Manuscript received: July 15, 2020
48 [143] J. Jones, H. Xiong, A. T. De La Riva, E. J. Peterson, H. Pham, S. R. Challa, Revised manuscript received: August 27, 2020
G. Qi, S. Oh, M. H. Wiebenga, X. I. P. Hernández, Y. Wang, A. K. Datye, Accepted manuscript online: September 1, 2020
49 Science 2016, 353, 150–154. Version of record online: ■■■, ■■■■
50
51
52
53
54
55
56
57
ChemNanoMat 2020, 6, 1 – 25 www.chemnanomat.org 24 © 2020 Wiley-VCH GmbH

These are not the final page numbers! ��


Minireview

1 MINIREVIEW
2
3
Less can do more: Single-atom
4
catalysts (SACs) with isolated metal
5 Dr. Y. Lu, Dr. Z. Zhang, Dr. F. Lin, Dr. H.
centers are attracting great research
6 Wang*, Prof. Y. Wang*
7
interest. This critical review summa-
8 rizes the recent advances of SACs 1 – 25
9 with applications in automobile
Single-atom Automobile Exhaust
10 exhaust aftertreatment. Strategies for
Catalysts
11 promoting the reactivity and reaction
12 mechanisms of different single-atom
13 catalysts during CO oxidation, hydro-
14 carbon oxidation and selective
15 catalytic reduction of NOx are
16 discussed.
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57

You might also like