Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

pubs.acs.

org/acscatalysis Research Article

Unraveling the Intermediate Reaction Complexes and Critical Role


of Support-Derived Oxygen Atoms in CO Oxidation on Single-Atom
Pt/CeO2
Yubing Lu,∇ Shulan Zhou,∇ Chun-Te Kuo, Deepak Kunwar, Coogan Thompson, Adam S. Hoffman,
Alexey Boubnov, Sen Lin, Abhaya K. Datye, Hua Guo,* and Ayman M. Karim*
Cite This: ACS Catal. 2021, 11, 8701−8715 Read Online
Downloaded via VIRGINIA POLYTECH INST STATE UNIV on August 31, 2021 at 20:22:58 (UTC).
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ACCESS Metrics & More Article Recommendations *


sı Supporting Information

ABSTRACT: CeO2-supported Pt single-atom catalysts have been


extensively studied due to their relevance in automobile emission
control and for the fundamental understanding of CeO2-based
catalysts. Though CeO2-supported Pt nanoparticles are often more
active than their single-atom counterparts, the former could easily
redisperse to Pt single atom under oxidizing diesel conditions.
Therefore, to maximize the reactivity of every Pt atom, it is
important to fully understand the reaction mechanism of CeO2-
supported Pt single atoms. Here, we report a CO oxidation study on
a Pt/CeO2 single-atom catalyst, where we can account for all of the
neighbors using in situ and operando spectroscopy techniques and
microcalorimetric measurements. Coupled with density functional
theory calculations, we present a comprehensive picture of the dynamics of the surface species, the role of surface intermediates, and
explain the observed reaction kinetics. We started with a catalyst containing exclusively single atoms and used in situ/operando
spectroscopy to provide evidence for their stability during the reaction and to identify the Pt1 complexes before and during the
reaction and their binding to CO. The results reveal that in the precatalyst, Pt is present as Pt(O)4 on the CeO2(111) step edge sites,
but during CO oxidation, we find that two Pt1 complexes coexist, representing two states of the same active site in the reaction cycle.
The dominant state/complex remains Pt(O)4, which adsorbs CO very weakly as shown by CO microcalorimetry. The second,
minority state/complex, Pt(CO)(O)3 is generated through the reaction of Pt(O)4 with CO, and CO is bound strongly to Pt1. Labile
oxygen adatoms from the CeO2 surface play a major role in the regeneration of Pt(O)4 either directly from Pt(O)3 or by reaction
with the strongly adsorbed CO in Pt(CO)(O)3. We show that the formation of an oxygen vacancy and generation of a labile O* are
not barrierless, which explains the long lifetime of Pt(CO)(O)3 and its detectability despite being a minority complex. The results
help to develop a comprehensive view of the dynamic evolution of Pt1 complexes along the reaction cycle and provide mechanistic
insights to guide the design of Pt-based single-atom catalysts.
KEYWORDS: single-atom catalysts, reaction mechanism, CO oxidation, operando characterization, density functional theory

■ INTRODUCTION
Atomically dispersed noble metals on different supports have
activity relationship and detailed mechanistic steps for these
novel classes of catalysts.
gained much attention due to their maximized metal utilization Due to the importance of Pt for many applications,
and promise for improving catalytic efficiency.1−4 Additionally, supported Pt single atoms (SA) have been investigated
their fundamentally different electronic properties compared extensively.4,13,16,26,30−36 A prototype single Pt atom catalyst
with their nanoparticle counterparts2,4−11 hold potential to is Pt1/CeO2, which has attracted intensive attention for CO
perform difficult selective chemical transformations,12−19 even oxidation because of its importance in exhaust emission
those traditionally limited to homogeneous catalysts.20,21 On a abatement, especially for its stability and resistance to
fundamental level, supported metal single sites with uniform
atomic-scale structure serve as model systems for bridging the Received: April 27, 2021
material gap and studying reaction mechanisms by relating the Revised: June 21, 2021
theory and experiment. Therefore, tremendous efforts have Published: July 1, 2021
recently been devoted to the synthesis of uniform, atomically
dispersed catalysts with much success.4,22−29 However, there is
still a substantial lack of detailed knowledge on the structure−

© 2021 American Chemical Society https://doi.org/10.1021/acscatal.1c01900


8701 ACS Catal. 2021, 11, 8701−8715
ACS Catalysis pubs.acs.org/acscatalysis Research Article

Figure 1. Surface structure of Pt1/CeO2 before CO oxidation. (a) AC-HAADF-STEM image of 1 wt % Pt1/CeO2. (b) Enlarged AC-HAADF-
STEM image of the region encased by the square in (a). (c) Side and top views of the DFT-optimized structure of the Pt single atom on the CeO2
step edge site with coordination to four O atoms. (d) Side and top views of DFT-optimized structure of weakly adsorbed O2 on Pt1/CeO2
(structure in d). (e) Pt L3 XANES and (f) magnitude of the Fourier-transformed k2-weighted χ(k) data (Δk = 3−12.5 Å−1) in R-space of 1 wt %
Pt1/CeO2 as-synthesized (calcined) and after pretreatment. (g) Wavelet transforms for the k2-weighted χ(k) signal of 1 wt % Pt1/CeO2 after
pretreatment and Pt foil. Scale bar of wavelet contour: 0−0.064 for calcined Pt1/CeO2; 0−0.08 for Pt foil. Na2Pt(OH)6 with +4 oxidation state and
Pt(acac)2 with +2 oxidation state are shown as standards in (e) and (f). Color of DFT structures in (c) and (d): Ce, light yellow; O, red; weakly
adsorbed O2 on Pt, orange; Pt, blue.

oxidation at high temperatures, unlike Pt nanoparticles that be much less active than the Pt/CeO2 nanoparticle (NP)
deactivate when oxidized.25,36−38 Several experimental and catalyst.36,42−44 Turnover frequencies (TOFs) of representa-
theoretical studies on Pt1/CeO2 have been reported. Stable Pt tive PtSA/CeO2 and PtNP/CeO2 at similar reaction conditions
single atoms on CeO2 support can be prepared via a high- are summarized in Table S1. Higher-activity Pt/CeO2 single-
temperature atom-trapping method, where Pt atoms diffuse to atom catalysts were recently reported, though still less active
stable adsorption sites on the CeO2 support.25 Up to 3 wt % Pt than Pt nanoparticles. For example, Pt single atoms on twisted
loading can be stabilized as single atoms through this atom- CeO2 surfaces are more active than PtSA/CeO2 with rod-type
trapping method to form a stable Pt square-planar structure on CeO2 support for low-temperature CO oxidation.45 Besides,
CeO2(111) step sites.24 Pt1/CeO2 catalysts can also be the Ga-doped Pt1/CeO246 showed better activity than the
synthesized using wet impregnation at lower calcination traditional Pt1/CeO2. It is clear that Pt1/CeO2 prepared with
temperatures.39,40 Resasco et al.40 showed that at low loadings, different methods performed differently. Hence, in search of a
Pt single atoms tend to bond strongly on step sites of CeO2 general guiding principle to the design of Pt1/CeO2, it is
with good thermal stability, but the stability decreases with an important to understand the origin of its activity, the rate-
increase in metal loading. Density functional theory (DFT) limiting step, and the reaction mechanism. More importantly,
calculations showed that while Pt single atoms on CeO2(111), though in many cases Pt nanoparticles are more active than Pt
(110), and (100) are all thermally stable,41 the activity and single atoms, multiple studies clearly showed that the Pt
stability during CO oxidation are site-dependent. Pt on nanoparticles on CeO2 easily redisperse into less active Pt
CeO2(110) was calculated to be the most thermally stable, single atoms under lean mixtures and oxidizing conditions in a
and the activity is in the order Pt/CeO2(111)/(100) > Pt/ practical diesel engine. Indeed, the redispersion of Pt particles
CeO2(110). The site-specific activity and stability indicate that is known to occur after oxygen treatment at 400 °C47 or in lean
the support stability and the characteristics of Pt binding site diesel exhaust gas conditions at ∼140 °C and above.43 Since
are important for the reactivity and stability of the Pt1 species. forming single atoms is unavoidable on CeO2-supported Pt
The Pt/CeO2 single-atom catalysts are in general not the most nanoparticle/cluster catalysts, it is highly desirable to under-
active for CO oxidation. Indeed, many Pt/CeO2 single-atom stand the catalytic mechanism of Pt single atoms to improve
catalysts prepared with different methods25,36 were reported to their activity.
8702 https://doi.org/10.1021/acscatal.1c01900
ACS Catal. 2021, 11, 8701−8715
ACS Catalysis pubs.acs.org/acscatalysis Research Article

Table 1. Summary of EXAFS Modeling Results at the Pt L3-Edge for the Pt References, 1 wt % Pt1/CeO2 Catalyst As-
Synthesized (Calcined), and Pt1/CeO2 Catalyst after Pretreatment (O2 and He Treatments at 400 °C)
Na2Pt(OH)6 Pt(acac)2 Pt1/CeO2 as-synthesized Pt1/CeO2 after pretreatmenta
NPt−O 6.1 ± 0.7 4.0 ± 0.4 6.1 ± 0.4 3.5 ± 0.4
RPt−O (Å) 2.00 ± 0.01 1.99 ± 0.01 2.02 ± 0.01 2.02 ± 0.01
σPt−O2 (Å2) 0.002 ± 0.001 0.002 ± 0.001 0.002 ± 0.001 0.000 ± 0.001
ΔE0 Pt−O (eV) 10 ± 2 9±1 12 ± 1 12 ± 1
reduced χ2 487 717 15 67
R-factor 0.003 0.002 0.0007 0.002
a
Spectrum collected in He at 150 °C. All of the spectra for the other samples were collected in air at room temperature.

Therefore, connecting the intrinsic CO oxidation activity (O)3, our DFT calculations show that CO can be oxidized with
with the local structure of Pt1 is crucial to fully understand the the involvement of labile surface oxygen (O*) species derived
reaction mechanism and the factors controlling the reactivity of from the CeO2 surface.
Pt. While the CO oxidation mechanism on Pt nanoparticles
supported on CeO2 is well understood,42,47,48 many questions
remain unanswered for the mechanism of Pt single atoms.
■ RESULTS
Structure of the Pt1 /CeO2 Catalyst before CO
Moreover, there is disagreement in the literature regarding the Oxidation (Precatalyst). The atomically dispersed Pt
spectroscopic observation of CO adsorption on Pt1/CeO2. For supported on CeO2 was prepared by a high-temperature
example, Pt1/CeO2 prepared by atom trapping was reported to atom-trapping method as previously reported.24,53 Figure 1a,b
adsorb CO strongly during CO oxidation. The vibrational CO shows the representative aberration-corrected high-angle
band (νCO) in DRIFTS (∼2095 cm−1) was stable in He and annular dark-field scanning transmission electron microscopy
O2 at 180 °C.24,25 On the other hand, Pt1/CeO2 prepared by (AC-HAADF-STEM) images showing isolated Pt atoms. We
impregnation methods showed either no CO adsorption at note that no clusters were detected before CO oxidation, and
room temperature40 or a similar CO peak (∼2090 cm−1)39 to their presence as a minority species can be ruled out since any
that reported24,25 on the Pt1/CeO2 catalyst prepared by atom weakly bound Pt not trapped as single atoms will agglomerate
trapping. It remains elusive how a very weakly (or strongly) into large Pt nanoparticles at the synthesis temperature of 800
adsorbed CO on Pt1 gets oxidized and what role does oxygen °C in air, which can be easily detected by techniques such as X-
from the CeO2 surface play in the reaction mechanism. ray diffraction and X-ray absorption spectroscopy.25 To probe
Previous studies have demonstrated that the surface oxygen the local environment of the Pt species and their oxidation
plays an important role in low-temperature CO oxidation on state, X-ray absorption spectroscopy (XAS) was performed at
Pt-supported nanoparticles.47,49−52 For example, recent work the Pt L3-edge. Spectra were collected at room temperature on
by our group and Ganzler et al. showed that the Pt-CeO2 the as-synthesized catalyst and after in situ pretreatment in O2
interfacial sites significantly promote the CO oxidation by at 400 °C followed by flushing and cooling down in He. The
supplying reactive/labile O* to overcome the CO poisoning of X-ray absorption near-edge structure (XANES) spectra are
Pt clusters,42,49 and this promotion effect can be further shown in Figure 1e along with reference samples, Pt4+ in
enhanced by CO treatment.47 However, it is still unclear how Na2Pt(OH)6 and Pt2+ in Pt(acac)2. The as-synthesized catalyst
the surface/labile oxygen is involved in CO oxidation on single showed white line intensity slightly lower than that of
Pt atoms on CeO2. Hence, to fully realize the advantages of Pt Na2Pt(OH)6, indicating that the oxidation state of the Pt
single-atom catalysts, it is essential to investigate and single atoms is close to Pt4+. After the in situ pretreatment in
fundamentally understand the reaction mechanism to help O2 at 400 °C and cooling down in He, the white line intensity
identify design rules. Specifically, we need to study the local of the single atoms is very similar to that of Pt(acac)2,
coordination of Pt single atoms and the adsorption behavior of indicating that the Pt single atoms are present as Pt2+. The
CO on the as-prepared catalyst and under reaction conditions decrease in oxidation state after 400 °C in O2 and cooling in
to identify the reaction intermediates and the rate-limiting He treatment suggests that O2 molecules are weakly bonded to
step(s). Pt single atoms during exposure to air and can be removed
In this work, we combined in situ/operando spectroscopy, during high-temperature treatment in He. Given the facile
microcalorimetry, and reaction kinetic measurements with removal of this adsorbed O2 in He, it is unlikely that the Pt in
first-principles calculations to study the catalyst structure and the precatalyst is located as Pt4+ in substitutional Ce4+ sites.
CO oxidation reaction mechanism on a uniform and stable Further details on the local coordination of the Pt single
Pt1/CeO2 catalyst. We show that in the catalyst before the atoms are revealed by modeling the extended X-ray absorption
reaction (precatalyst), Pt1 is adsorbed on CeO2(111) step edge fine structure (EXAFS) spectra of the Pt1/CeO2 and Pt
sites as a square-planar Pt(O)4 complex. However, during CO standards (Figures S1−S5; the modeling results are summar-
oxidation, a combination of infrared and X-ray absorption ized in Table 1). We note that no Pt−Pt backscattering feature
spectroscopy techniques reveals that two Pt1 complexes was observed or could be modeled in the Pt1/CeO2 EXAFS
coexist, while we can rule out the presence of Pt clusters spectra (Figure 1f and Table 1). Consistent with this picture,
based on the operando experimental evidence. The dominant the wavelet transform 54,55 EXAFS of Pt 1 /CeO 2 after
state/complex remains Pt(O)4, while the second complex is calcination (Figure 1g) shows only one intensity maximum
identified as Pt(CO)(O)3 and is generated through the between 5 and 6 Å−1 in k-space and 1.5 Å in R-space, which
reaction of Pt(O)4 with CO. We show that the CO adsorbs corresponds to Pt−O scattering, while no Pt−Pt scattering
weakly on Pt(O)4, while CO is bound very strongly to Pt in could be seen as compared to the wavelet transform EXAFS of
Pt(CO)(O)3. Despite this strong Pt−CO bond in Pt(CO)- Pt foil. The as-synthesized catalyst showed a Pt−O
8703 https://doi.org/10.1021/acscatal.1c01900
ACS Catal. 2021, 11, 8701−8715
ACS Catalysis pubs.acs.org/acscatalysis Research Article

coordination number of 6 similar to the Na2Pt(OH)6 standard, atoms with Pt−O distances between 1.89 and 2.03 Å, four Ce
confirming that Pt1 is present as Pt4+ as also shown by XANES atoms at 3.2−3.3 Å, three oxygen atoms at 3.5 Å, and one
(Figure 1e). The Pt−O coordination number after pretreat- oxygen atom at 3.61 Å (see Figure S6 for the optimized DFT
ment in O2 at 400 °C followed by flushing and cooling in He is structure and Table S2 for the detailed bond distances). The
3.5, which is similar to 4.0 for Pt in Pt(acac)2, confirming the coordination numbers and bond distances are in very good
XANES results that Pt is present as Pt2+. To provide more agreement with our EXAFS modeling results (Table 2 and
details on the local coordination of the Pt single atoms and Figure S7). Therefore, we conclude that in the pretreated
their bonding with CeO2, we fit the EXAFS spectra of the catalyst Pt1 is sitting at a CeO2(111) step edge site and is
pretreated Pt1/CeO2 catalyst using a model containing Pt−O bound to four oxygen atoms, Pt(O)4, in a square-planar
(both short and long) and Pt−Ce single scattering paths. The geometry. This conclusion is also consistent with our previous
results in Table 2 show that, on average, Pt is surrounded by aberration-corrected STEM results on the same catalyst
showing Pt single atoms at step edges of CeO2(111).24
Table 2. Comparison between EXAFS Modeling Results at O2 and CO Adsorption on the Pt1/CeO2 Precatalyst.
the Pt L3-Edge for the Pt1/CeO2 Catalyst after O2 and He Having established that our DFT model closely matches the
Treatments at 400 °C and the DFT-Optimized Model of the experimental results, we investigated the adsorption of O2 and
Pt Single Atom on CeO2(111) Step Edgea CO on Pt1/CeO2 by DFT calculations and microcalorimetry.
The calculated O2 adsorption energy on Pt1 in the optimized
EXAFS fit DFT modelb
structure shown in Figure 1c is −24 kJ/mol (see Figure 1d for
NPt−O 3.4 ± 0.3 4 the structure with O2 adsorbed on Pt1/CeO2 with the square-
NPt−Ce 4.1 ± 1.9 4 planar Pt(O)4 structure). Calorimetry measurements of O2
NPt−O_long 4.5 ± 1.4 4 adsorption on Pt1/CeO2 showed heats consistent with weakly
RPt−O (Å) 2.02 ± 0.01 1.9−2.0 adsorbed O2. The initial adsorption energy is around −20 to
RPt−Ce (Å) 3.13 ± 0.01 3.2−3.3 −40 kJ/mol (Figure 2a), which is in reasonably good
RPt−O_long (Å) 3.58 ± 0.04 3.5−3.6 agreement with the DFT calculation results. The results are
σPt−O2 (Å2) 0.000 ± 0.001 consistent with the fact that weakly adsorbed O2 on Pt1 can be
σPt−Ce2 (Å2) 0.008 ± 0.003 easily removed by treatment in He as observed by EXAFS,
σPt−O_long2 (Å2) 0.004c where the Pt−O coordination decreased from 6 to 4 (Table 1).
ΔE0 (eV) 12 ± 1 Additionally, our DFT calculations show that CO adsorption
reduced χ2 53
on Pt1/CeO2 with the square-planar Pt(O)4 structure is weak,
R-factor 0.0032
a
−42 kJ/mol (see Figure S8 for the structure). We measured
The Pt−Olong denotes the coordination between Pt and a further the CO adsorption energy using microcalorimetry at −75 °C
away O on CeO2 surface as shown in the DFT model (Figure 1c). on the Pt1/CeO2 catalyst after in situ treatment in O2 and He
b
DFT model is shown in Figure 1c. cNotation: N, coordination at 400 °C. As shown in Figure 2a, at low coverage (CO/Pt <
number of the absorber−backscatterer pair; R, radial absorber−
backscatterer distance; σ2, mean square displacement of the half-path
0.2), the initial differential heat of CO adsorption is −44 kJ/
length and it represents the stiffness of the bond for a single scattering mol, which is in excellent agreement with the DFT calculation
path; ΔE0, correction to the threshold energy. results. This indicates that CO adsorbs weakly on Pt(O)4.
State of Pt1 under Reaction Conditions. To further
characterize the Pt−CO binding and local coordination of Pt1/
∼4 oxygen atoms at 2.01 Å, ∼4 Ce atoms at 3.13 Å, and ∼4 CeO2 under reaction conditions, in situ/operando DRIFTS and
oxygen atoms at 3.56 Å. Pt single atoms on CeO2(111) XANES/EXAFS experiments were performed. For both
terraces were reported to adopt sixfold coordination with experiments, Pt1/CeO2 was pretreated in O2 and then He at
oxygen atoms at 2.12−2.18 Å, which is not consistent with our 400 °C. After pretreatment, CO oxidation was performed
results.24,40,41 On the other hand, it has been reported that Pt under 3 kPa CO and 2 kPa O2 for 20 min at 150 °C and CO2
single atoms adsorbed on CeO2(100) or CeO2(111) step edge was detected in the reactor effluent in the DRIFTS and XAFS
sites are coordinated to four oxygen atoms in a square-planar experiments (Figures 2b and S9). During the in situ/operando
geometry. Although Pt adopts a square-planar geometry on DRIFTS experiment, one major peak at 2096 cm−1 is observed
both surfaces, the distance between Pt and Ce is different. in the νCO region during CO oxidation (Figure 2b) and the
Kottwitz et al.39 reported a Pt−Ce distance of 3.34 Å for Pt peak remained in the same position after flushing the gas-phase
single atoms on CeO2(100), while our results show a much CO and O2 with N2/O2 (Figure 2d,e). We note that the CO
shorter distance of 3.13 Å. Therefore, in our Pt1/CeO2, Pt is peak position (2096 cm−1) and shape are identical on a lower
likely adsorbed on CeO2(111) step edges rather than on Pt weight loading catalyst (0.2 vs 1 wt %), showing that
CeO2(100). loading does not affect the anchoring site for Pt1 (Figures 2c
To provide further insights, we performed DFT calculations and S10). This is also reflected in their similar kinetics as
for Pt single atoms adsorbed on CeO2(111) step edge sites, discussed below. We assign the 2096 cm−1 peak to CO
and the optimized model is shown in Figure 1c. The adsorbed on Pt single atoms as previously reported.25,31,56
adsorption energy of Pt at this site of CeO2 is −705 kJ/mol, Notably, the adsorbed CO on Pt1 is stable in N2 (Figure 2d)
which is more negative than the cohesive energy of Pt (−534 and O2 (Figure 2e) with no significant change in peak intensity
kJ/mol), underscoring its stability and inclination to exist in after 5 min in either gas environment at 150 °C, indicating that
the atomic form. The results are consistent with the well- CO is bound very strongly to Pt1. Since Pt1 on CeO2(111) step
established ability of CeO2(111) step edges to trap Pt single edges is present as Pt(O)4 and is shown to adsorb CO very
atoms and stabilize them against agglomeration at high weakly (see above), we hypothesize that the strong CO
temperatures during the synthesis.24,25 Furthermore, the adsorption is due to the reaction of CO with Pt(O)4 to form
DFT model shows that Pt1 is surrounded by four oxygen Pt(CO)(O)3 during CO oxidation, i.e., Pt(O)4 + 2CO →
8704 https://doi.org/10.1021/acscatal.1c01900
ACS Catal. 2021, 11, 8701−8715
ACS Catalysis pubs.acs.org/acscatalysis Research Article

Figure 2. CO/O2 adsorption and resting state of Pt1/CeO2. (a) CO/O2 calorimetry on 1 wt % Pt1/CeO2 after pretreatment in 20 kPa O2 at 400 °C
for 30 min followed by evacuation under vacuum. The catalyst was cooled to at −75 °C under high vacuum and CO/O2 was pulsed onto the
catalyst surface. CO/O (1/2 O2) coverage was normalized to mol of Pt in the catalyst (nominal loading). (b) CO chemisorption DRIFTS of 1 wt
% Pt1/CeO2 under CO oxidation at 150 °C. (c) CO chemisorption DRIFTS of 1 wt % Pt1/CeO2 and 0.2 wt % Pt1/CeO2. Spectra were recorded
after 20 min CO oxidation at 150 °C and 2 min N2 purging to remove the gas-phase CO. CO chemisorption DRIFTS characterizing the 1 wt %
Pt1/CeO2 catalyst under the N2 flow (d) or O2 flow (e) after CO oxidation. Spectra were recorded after CO oxidation at 150 °C for 20 min and
then short N2/O2 (1 kPa O2) purging to remove the gas-phase CO. (f) DFT-optimized structure of the Pt single atoms on the CeO2 (111) step
edge site with three O coordination and one CO coordination. Conditions for (b−e): CO oxidation with 3 kPa CO + 2 kPa O2 and total flow rate
of 50 sccm. The flow rate of the N2 or O2 step is 80 sccm. Color: Ce, light yellow; O, red; Pt, blue; C, gray.

Pt(CO)(O)3 + CO2. In fact, our DFT calculations show that three oxygens instead of four) and can adsorb CO to saturate
this reaction has a relatively low barrier (62 kJ/mol) and CO is their coordination. Therefore, the high 2096 cm−1 peak
bound strongly to Pt1 in Pt(CO)(O)3 (see Figure 2f for the intensity during CO oxidation shows that the initial Pt(O)4
structure) with a binding energy of −260 kJ/mol. Furthermore, can react with CO to form Pt(CO)(O)3.
the stability of this Pt(CO)(O)3 complex indicates that it is an XANES and EXAFS spectra after the O2/He pretreatment
intermediate complex during CO oxidation. We note that we and during CO oxidation are shown in Figure 3. The XANES
performed the in situ CO chemisorption DRIFTS experiment spectra (Figure 3a) are almost identical, indicating that Pt is
at −75 °C on Pt1/CeO2 before CO oxidation to capture the maintained as Pt2+ under reaction conditions. The EXAFS
relatively weak CO adsorption on Pt(O)4. However, the results modeling results (Table 3 and Figures S3−S4) show a slight
in Figure S11 show a dominant peak for CO adsorbed on decrease in Pt−O coordination from 3.5 in He to 3.2 in CO +
CeO2 (2174 and 2156 cm−1), and a very small peak at 2093 O2, but the difference was within the uncertainty of the fit. A fit
cm−1 is also detected (20× smaller than the 2096 cm−1 peak that includes adsorbed CO on Pt resulted in Pt−C
under reaction conditions). Additionally, a small peak at 2123 coordination of ∼0.2 at 1.87 Å, but the fit was not stable
cm−1 was detected during the flow of CO but was not stable (e.g., after small changes to the range in k- or R-space or the
and disappeared once CO was flushed with N2, indicating initial parameter guesses) and therefore was not shown. The
weakly adsorbed CO. The 2123 cm−1 peak could be assigned wavelet transform plot of Pt1/CeO2 EXAFS during CO
to CO adsorbed on Pt(O)4 similar to a recent report by oxidation (Figure 3d) shows a slightly different magnitude
Maurer et al.36 (2129 cm−1). The results are consistent with than the pretreated Pt1/CeO2 (Figure 3c) at R-space ∼1.5 Å
the low CO−Pt(O)4 binding energy calculated by DFT and and k-space between 6 and 12 Å−1. The more focused
measured by microcalorimetry and indicate that the majority of maximum at R-space 2.5−3 Å and k-space ∼4 Å−1 could
Pt atoms are present as Pt(O)4 and adsorb CO weakly. correspond to the contribution of the Pt−CO paths (single
However, the presence of the 2093 cm−1 band at −75 °C along and multiple scattering). We note that the differences are
with the slightly higher initial CO (and O2) heat of adsorption small, which is consistent with the inability to accurately fit
from microcalorimetry (also at −75 °C) suggests that a Pt−CO scattering paths in the EXAFS model and indicates
minority of Pt sites are already partially reduced (i.e., bound to that only a small percentage of Pt atoms have CO adsorbed
8705 https://doi.org/10.1021/acscatal.1c01900
ACS Catal. 2021, 11, 8701−8715
ACS Catalysis pubs.acs.org/acscatalysis Research Article

Figure 3. Operando XAFS characterization of Pt1/CeO2. (a) Pt L3 XANES and (b) magnitude of the Fourier-transformed k2-weighted χ(k) data
(Δk = 3−12.5 Å−1) in R-space of the 1 wt % Pt1/CeO2 catalyst under He flow at 150 °C and during CO oxidation at 150 °C, 3 kPa CO + 2 kPa O2,
and a total flow rate of 50 sccm. Corresponding wavelet transforms of the k2-weighted χ(k) signal (c) under He flow at 150 °C and (d) during CO
oxidation at 150 °C.

Table 3. Summary of EXAFS Modeling Results at the Pt L3- atom intermediate complexes along the reaction pathway. In
Edge for the Pt1/CeO2 in He at 150 °C and during CO other words, the EXAFS results show that Pt exists in more
Oxidation at 150 °C under 3 kPa CO + 2 kPa O2 than one state under reaction conditions, Pt(O)4 and
Pt(CO)(O)3, and only the state bound to CO is detected by
Pt1/CeO2 in He at Pt1/CeO2 in CO + O2 at
150 °C 150 °C DRIFTS (see Figure S12−S13 and the associated text for
NPt−O 3.5 ± 0.4a 3.2 ± 0.3 further evidence that the majority of Pt1 is not bound to CO in
RPt−O (Å) 2.02 ± 0.01 2.02 ± 0.01 DRIFTS). Moreover, the EXAFS results suggest that the most
σPt−O2 (Å2) 0.000 ± 0.001 0.000 ± 0.001 abundant state under reaction conditions is Pt(O)4 and not the
ΔE0 Pt−O (eV) 12 ± 1 13 ± 1 state detected by DRIFTS, i.e., Pt(CO)(O)3. We note that
reduced χ2 67 89 these two complexes are two states of the same active site
R-factor 0.002 0.002 during the reaction, i.e., two states along the reaction cycle
a
Notation: N, coordination number of the absorber−backscatterer (vide infra). To our knowledge, the possibility that metal single
pair; R, radial absorber−backscatterer distance; σ2, mean square atoms can be present in more than one intermediate stable
displacement of the half-path length and it represents the stiffness of state along the reaction coordinate/pathway has only
the bond for a single scattering path; ΔE0, correction to the threshold previously been reported by computational studies,57,58 and
energy. our work provides the first experimental evidence.
CO Oxidation Kinetics on Pt1. CO oxidation kinetics on
during the reaction. In summary, the operando XANES and the 1 wt % Pt1/CeO2 catalyst were measured at two different
EXAFS results confirm that Pt remains as Pt1 in Pt2+ oxidation temperatures: 100 and 150 °C (Figure 4). The CO order was
state during the reaction. 0.7 ± 0.1 and 0.5 ± 0.1 at 100 and 150 °C, respectively, and
It may appear that operando DRIFTS and EXAFS are at the O2 order was ∼0 at both temperatures. The apparent
odds with what they tell us about the working catalyst, so it is activation energy (Eapp) of 69 ± 3 kJ/mol between 80 and 100
important to note the differences between what they actually °C was slightly lower than the Eapp measured between 145 and
measure. Specifically, the DRIFTS experiment is sensitive to 165 °C. To study the effect of loading, which can affect the
only the Pt1 complexes that are bound to CO, i.e., uniformity of the Pt1 anchoring sites on CeO2,40 the reaction
Pt(CO)(O)3 and not Pt(O)4. On the other hand, EXAFS kinetics were also measured on a lower loading of 0.2 wt %
captures an average of all of the Pt in the catalyst. Therefore, Pt1/CeO2. A similar positive order in CO (0.8) and a ∼0th
the combined results indicate that the Pt−CO peak detected in order in O2 were also measured on 0.2 wt % Pt1/CeO2 at 150
DRIFTS represents only a small fraction of Pt, which is °C (Figure S14). The Eapp of 80 ± 1 kJ/mol measured on 0.2
strongly bound to CO. This is very intriguing and indicates wt % Pt1/CeO2 is also similar to that on 1 wt % Pt1/CeO2 (77
that the DRIFTS spectra only capture one of the Pt single- ± 1 kJ/mol). The similar kinetics on 0.2 and 1% Pt1/CeO2 are
8706 https://doi.org/10.1021/acscatal.1c01900
ACS Catal. 2021, 11, 8701−8715
ACS Catalysis pubs.acs.org/acscatalysis Research Article

Figure 4. CO oxidation kinetic measurement on 1 wt % Pt1/CeO2. Effect of CO (a) and O2 (b) partial pressure on the TOF on 1 wt % Pt1/CeO2
measured at both 150 and 100 °C. PO2 = 10 kPa, PCO = 0.3−1 kPa in (a); PCO = 1 kPa, PO2 = 2−14 kPa in (b). (c) Eapp of 1 wt % Pt1/CeO2
measured at two different temperature regions. Region 1, between 145 and 163 °C; Region 2, between 80 and 100 °C; measured with 2 kPa CO
and 1.5 kPa O2 and 50 sccm total flow rate.

Figure 5. Reaction pathways of CO oxidation on Pt1 at the CeO2(111) step edge site proposed by DFT calculations. Color: Ce, light yellow; O,
red; Pt, blue; C, gray; O in O2 and labile O, orange. Details on the generation of oxygen vacancies and labile oxygen reaction steps can be found in
the Supporting Information (SI). The energy is given in kJ/mol.

not surprising given their identical DRIFTS spectra under metallic Pt or Pt clusters were observed on the postreaction
reaction conditions (Figure 2c) and confirm the uniformity of catalyst which confirms the excellent thermal stability of Pt1/
the Pt single atoms. CeO2 during low-temperature CO oxidation. Furthermore, the
We note that other groups have proposed that trans- reaction kinetics (orders and Eapp) are a signature of the active
formation of Pt1 to form metallic clusters may be an essential site and are sensitive to the presence of different types of
step to obtain measurable CO oxidation activity below 200
sites,11,57 which in the case of Pt/CeO2 would be Pt single
°C.36 However, our operando EXAFS and XANES results
atoms and Pt clusters. The reaction kinetics on Pt1/CeO2
(Figure 3a) show that Pt remains isolated (Pt1) as Pt2+ similar
to Pt1 in the precatalyst at 150 °C and shows significant CO shown in Figure 2 are substantially different from those
oxidation reactivity. To further confirm the thermal stability of measured on Pt clusters (see the previous work42 and Figures
Pt1/CeO2 during CO oxidation, we also characterized the S12−S13 and the associated text in the Supporting
postreaction Pt1/CeO2 catalyst using in situ DRIFTS at 35 °C Information (SI) for the effect of cluster formation on
(Figure S15) as well as AC-HAADF-STEM (Figure S16). No kinetics), which combined with our operando and postreaction
8707 https://doi.org/10.1021/acscatal.1c01900
ACS Catal. 2021, 11, 8701−8715
ACS Catalysis pubs.acs.org/acscatalysis Research Article

Scheme 1. Key Reaction Steps and Intermediate Complexes for CO Oxidation on Pt1/CeO2a

a
(i) denotes Pt(O)4, (ii) denotes Pt(CO)(O)4, (iii) denotes Pt(O)3, and (iv) denotes Pt(CO)(O)3. The Pt, O, and C at the active site are
represented by blue, red, and black colors, while the CeO2 substrate in gray. O* denotes labile/reactive oxygen provided by CeO2. The Pt(O)4 (i)
and Pt(CO)(O)3 (iv) intermediate complexes were observed experimentally during CO oxidation. The CO adsorption on Pt(O)4 (i, ii) is
equilibrated, but the Pt(CO)(O)4 complex is not detected experimentally during CO oxidation due to its very low concentration (very weak CO
binding, −42 kJ/mol). Pt(CO)(O)4 reaction to Pt(O)3 (ii, iii) is slow because of the reaction barrier and low Pt(CO)(O)4 concentration. CO
adsorption on Pt(O)3 (iii, iv) is facile with a high CO adsorption energy (−260 kJ/mol). The CO in Pt(CO)(O)3 (iv) reacts with labile O* with a
relatively low barrier (green pathway). Alternatively, Pt(O)3 can react directly with labile O* to regenerate Pt(O)4 (blue pathway). However, both
of these pathways are limited by the generation of labile O*, which is initiated by the gas-phase CO reaction with CeO2.

characterizations clearly show that Pt single atoms (Pt1) are kJ/mol energy change). In the next step, the adsorbed CO2 on
responsible for the measured activity and kinetics. Pt (iv) is formed with an endoergicity of 22 kJ/mol.
We now discuss the mechanistic implications of the reaction Subsequently, the formed CO2 desorbs with a 40 kJ/mol
orders measured on Pt1 (Figure 4). Overall, the positive barrier, leaving behind the Pt(O)3 (v) structure (Figure S17).
reaction order in CO indicates that the Pt single atoms are not From Pt(O)4 (i) to Pt(O)3 (v), the overall reaction barrier is
poisoned (kinetically) by CO. Additionally, the ∼0th order in 62 kJ/mol. The Pt(O)3 (iv) structure leaves an oxygen vacancy
O2 indicates that O2 activation is facile. The reaction orders next to Pt. This vacancy can be filled by either CO or O2.
suggest that the reaction follows an Eley−Rideal/Mars−van Energetic calculations show that the adsorption energy of CO
Krevelen (MvK) mechanism. However, to have a positive (v to vi) is −260 kJ/mol, which is much higher than that of O2
reaction order in CO, the rate-limiting step would likely (−96 kJ/mol, v to vi′, see Figure S18 for the detailed
involve a reaction of CO with an O bound to Pt1 in one of the structure). Given the strong adsorption of CO, the results
resting states, either Pt(O)4 that adsorbs CO weakly or suggest that Pt(CO)(O)3 is likely the intermediate complex
Pt(CO)(O)3 in which CO is adsorbed strongly. Additionally, observed in our in situ/operando DRIFTS experiments. The
Pt(CO)(O)3 was not affected by O2 (Figure 2e), suggesting oxidation of such strongly adsorbed CO (vi) by the
that an oxygen vacancy needs to be created by CO before O2 coadsorbed O2 (vii) to recover Pt(O)4 and release a CO2
can be adsorbed and activated. Dosing CO on the catalyst at (ix) is difficult with a high barrier of 156 kJ/mol. Interestingly,
150 °C (after flushing CO and O2 with N2) to simulate the this Pt(CO)(O)3 intermediate can be rescued by labile
half-reactions involved in an Eley−Rideal/Mars−van Krevelen oxygens (O*), which are presumably generated via O2
mechanism leads to over-reduction and agglomeration of Pt dissociation at the available oxygen vacancies over ceria
(Figures S12−S13). However, despite the over-reduction of Pt (Figures S19−S20 and Tables S3−S4 and the associated text
in pure CO, the experimental results suggest that CO is needed for further discussion on the formation of labile O*s). As
to create a vacancy site where O2 gets activated, but it is not shown in Figure 5, the DFT results suggest that the CO
clear whether this vacancy is adjacent to one of the Pt1 adsorbed on Pt(O)3 (vi) could easily combine with a nearby
complexes or on a nearby site on CeO2 and which reaction labile O* (vii′) adsorbed on the lower surface of the step to
step is rate-limiting. To answer these questions about the form a bent CO2 (ix′) with a low barrier of 31 kJ/mol. This
possible reaction mechanism, we performed DFT calculations step is highly exothermic (−181 kJ/mol). Subsequently, a
starting from Pt1 on the CeO2(111) step edge site. second labile O* (x′) adsorbed nearby recovers Pt(O)4 (i) at
Reaction Mechanism from DFT. The proposed reaction the CeO2(111) step edge by overcoming a small barrier of 22
pathways from DFT calculations are shown in Figure 5. The kJ/mol and, in the meantime, facilitates the desorption of CO2
reaction cycle starts with Pt(O)4 (i), where four lattice oxygens (xii′). Alternatively, the Pt(O)3 species (v) can also be
are bonded with one Pt single atom. In the next step, CO converted directly to Pt(O)4 (i) by the same labile O*, which
weakly adsorbs to form Pt(O)4(CO) (ii) and the adsorption is an exothermic (−342 kJ/mol) process with a minor barrier
energy of CO is −42 kJ/mol. The adsorbed CO facilely reacts of 11 kJ/mol. This can be the reason for the low concentration
with one lattice O to form a bent CO2 intermediate (iii) (−174 of Pt(CO)(O)3 detected experimentally. We note in passing
8708 https://doi.org/10.1021/acscatal.1c01900
ACS Catal. 2021, 11, 8701−8715
ACS Catalysis pubs.acs.org/acscatalysis Research Article

that no entropic contribution is considered in Figure 5, which Furthermore, under our experimental conditions (150 °C, 1
could significantly influence the rate constants via entropic kPa CO, and 10 kPa O2), Pt(O)4 is the dominant intermediate
contributions. Pt1 complex, while Pt(CO)(O)3 is a minority species. The
The existence of labile O* on ceria-based material is well results indicate that [O*] is somewhat limited, resulting in a
known,42,59−61 but its formation and diffusion are not well combination of the blue and green pathways in Figure 5/
understood. Although a complete elucidation of these steps is Scheme 1. Moreover, the stability of the Pt(CO)(O)3 peak in
beyond the scope of this work, we show in the SI that the O2 observed in DRIFTS (Figure 2e) further confirms that CO
presence of Pt1 significantly enhances the formation of O* is required to create a vacancy to activate O2 and create labile
species on ceria. In CO oxidation, the formation is likely O*. The results suggest that the apparent activation energy
initiated by the reaction between the impinging CO and lattice measured experimentally involves the barriers for reduction of
oxygen, leading to a gaseous CO2 molecule and an oxygen Pt(O)4 as well as creating a vacancy near Pt for O2 activation
vacancy on the surface. While this reaction is exothermic, it is and generation of labile O*. As discussed above, the blue
not without a barrier. In fact, the barrier is ∼72−78 kJ/mol, as pathway in Scheme 1 is the major pathway under the reaction
shown in Figure S21 and Table S5. This step is probably conditions used for the in situ/operando DRIFTS and XAS
followed by O2 adsorption at the oxygen vacancy and its experiments in Figures 2−3. However, since Pt(CO)(O)3 is a
dissociation (facile, see Table S4), resulting in the labile O* minor species under these conditions, varying the CO and O2
species. However, due to the presence of a reaction barrier partial pressures should affect its concentration. Therefore, to
(72−78 kJ/mol) for creating the vacancy, the concentration of further test the fidelity of the reaction pathways in Scheme 1,
labile O* will be dependent on CO concentration and we studied the effect of CO and O2 partial pressures on the
temperature and this in turn will affect the dominant reaction surface coverage of Pt(CO)(O)3 using operando DRIFTS. As
pathway and most abundant Pt1 intermediate complex, which shown in Figure S22a, higher CO partial pressure leads to a
is discussed below. higher surface coverage of Pt(CO)(O)3 (peak at ∼2093 cm−1).
A complete understanding of the kinetics requires more The formation of Pt(O)3 from Pt(O)4 is highly sensitive to
detailed kinetic Monte Carlo simulations, which are beyond PCO and therefore the concentration of Pt(CO)(O)3 increases
the scope of this work. However, the putative reaction as the PCO ratio increases. We note that the higher CO partial
pathways in Figure 5 provide important insights into the CO pressure also affects the CeO2 surface vacancy formation rate
oxidation mechanism on Pt1/CeO2. To simplify the reaction and consequently the concentration of labile O*, but it appears
pathways and connect theory to experiments, we summarize in that the O* concentration around Pt is limited by its diffusion/
Scheme 1 the key reactions and intermediates from the blue mobility and therefore less sensitive to the CO and O2 partial
and green pathways in Figure 5. As shown in Scheme 1, the pressures, as shown in Figure S22b. The results in Figure S22
CO adsorption on Pt(O)4 is equilibrated, but Pt(O)4 is are consistent with the ability to vary the relative rates of the
favored due to the weak CO binding. Consequently, the pathways depicted in Scheme 1 and the dynamic equilibrium
formation of Pt(O)3 is limited by the low concentration of between Pt(CO)(O)3 and Pt(O)4.
Pt(CO)(O)4 and the barrier for CO2 formation/desorption. In summary, through a combination of detailed in situ/
Once Pt(O)3 is formed, its oxidation to Pt(O)4 can go through operando characterizations, reaction kinetics, and DFT
two possible pathways (blue or green) depending on the calculations, we provided a detailed account of how Pt1
temperature and the concentrations of CO and labile O*. catalyzes CO oxidation and the important role of labile O*
Notably, the regeneration of Pt(O)4 from Pt(O)3 through from CeO2 in the reaction cycle. More importantly, the
either of the pathways in Scheme 1 involves low reaction participation of labile O* provided by CeO2 reconciles the
barriers. Consequently, the availability/concentration of presence of Pt1 intermediates with strong and weakly adsorbed
nearby labile O* dictates which pathway the oxidation of CO, namely, Pt(CO)(O)3 and P(O)4.
Pt(O)3 will follow (blue or green in Figure 5 and Scheme 1)
and the relative amounts of Pt(CO)(O)3 and Pt(O)4.
Specifically, if the availability of labile O* is unlimited, the
■ DISCUSSION
Pt single-atom catalysts provide exciting opportunities for
reaction would follow the blue pathway in Figure 5/Scheme 1, understanding the reaction mechanism at the atomic level and
where Pt(O)4 is the dominant intermediate Pt1 complex. how to tune the local structure of Pt atoms to design more
Additionally, the overall reaction rate would be limited by the active catalysts. Recent reports show that the reactivity of Pt1 is
adsorption of CO on Pt(O)4 and the reaction to form Pt(O)3. strongly dependent on the support. For example, Pt single
On the other hand, if the concentration of labile O* is very atoms on small CeO2 clusters supported on TiO2 are more
low, the reaction would follow the green pathway in Figure 5/ active than Pt single atoms on TiO2 or CeO2.62 Additionally,
Scheme 1, the oxidation of Pt(CO)3 will be the rate-limiting CO oxidation reactivity on Pt1/TiO2 was dependent on the
step, and Pt(CO)(O)3 would become the most abundant Pt1 local coordination of Pt after different pretreatments.30
intermediate complex. However, the local coordination of Pt single atoms and their
The above analysis is consistent with our in situ/operando binding with CO/O2, especially under CO oxidation reaction
characterization and the measured reaction orders. First, the conditions, can be complicated, leading to what seems to be
positive CO reaction order measured experimentally is contradictory results in the literature. For Pt1/CeO2 as an
consistent with the rate-limiting step either involving weak example, Resasco et al. showed that the highly uniform and
CO adsorption/reaction on Pt(O)4 (blue pathway) or CO low-loading Pt1/CeO2 single-atom catalyst adsorbs CO very
reaction to generate a vacancy on CeO2 (green pathway). weakly, which is difficult to be observed by CO chemisorption
Additionally, the zeroth order in O2 is consistent with the facile in DRIFTS.40 Our microcalorimetry, EXAFS, and DRIFTS
O2 activation on CeO2 (Table S4). Second, our in situ and results confirm the weak adsorption of CO on Pt1 due to the
operando EXAFS and DRIFTS measurements confirm the saturated coordination of Pt1 present as Pt(O)4 on CeO2(111)
existence of both Pt(O)4 and Pt(CO)(O)3 complexes. step edge sites. However, in other studies, CO was reported to
8709 https://doi.org/10.1021/acscatal.1c01900
ACS Catal. 2021, 11, 8701−8715
ACS Catalysis pubs.acs.org/acscatalysis Research Article

be adsorbed strongly on Pt1/CeO2 during CO oxidation.25,31 modifying the local environment of Pt1 to overcome the weak
For example, multiple studies show a strong CO adsorption CO adsorption and increase the availability of labile O*. The
peak at 2090−2100 cm−1 over Pt1/CeO2,25,31,45 but the results will be used to further guide the future design of more
activities of these reported Pt1/CeO2 catalysts are totally active Pt single-atom catalysts.


different. This cannot be explained by the typical MvK or
Langmuir−Hinshelwood (LH) mechanism if Pt(CO)(O)3 is CONCLUSIONS
the only active species. Our results suggest this is because
Pt(O)4 and Pt(CO)(O)3 coexist as intermediates in the same We studied the mechanism of CO oxidation on a single-atom
reaction cycle. Specifically, we demonstrate that the complex Pt/CeO2 catalyst and its binding with adsorbates using
observed in the Fourier transform infrared (FTIR) spectrum microcalorimetry, operando spectroscopy, kinetic measure-
with a peak located between 2090 and 2100 cm−1 is ments, and DFT calculations. We show that in the precatalyst
Pt(CO)(O)3, which is an intermediate along the reaction Pt single atoms are stabilized as Pt(O)4 on CeO2(111) step
cycle that starts with Pt(O)4. However, the Pt(O)4 complex edge sites. Pt(O)4 is coordinatively saturated, and micro-
cannot be observed in the FTIR spectrum due to weak CO calorimetry and DFT calculations reveal that it binds weakly to
adsorption. Furthermore, despite the strong binding of CO to O2 and CO, −24 and −42 kJ/mol, respectively. During CO
Pt in Pt(CO)(O)3 (−260 kJ/mol), the reaction with labile O* oxidation, operando EXAFS and DRIFTS show that Pt remains
can regenerate Pt(O)4. Additionally, the relative amounts of as single atoms and is present as two intermediate complexes,
Pt(O)4 and Pt(CO)(O)3 depend on the reducibility of Pt(O)4 Pt(O)4 and Pt(CO)(O)3, where the latter binds CO strongly
and the concentration of labile O* supplied by CeO2. These (−260 kJ/mol from DFT calculation). A reaction mechanism
results show that despite detecting the same intermediate was proposed based on our operando characterization and
Pt(CO)(O)3 complex in FTIR on different catalysts, the reactivity measurements to explain the kinetics and role of
intermediate complexes. DFT calculations show that the
activity depends strongly on the reducibility of both Pt(O)4
reaction is limited by the initial reaction between CO and a
and CeO2 and on the concentration of labile O*. Our work
lattice oxygen coordinated to Pt(O)4 to produce Pt(O)3, which
therefore reconciles the seemingly contradictory results in the
can react with labile O* to regenerate Pt(O)4 or react with CO
literature, especially about the connections between the FTIR-
to form Pt(CO)(O)3. More importantly, our DFT calculation
observed structure and the catalytic performance, and
results show that despite the strong CO binding to Pt in
demonstrates that the differences observed are directly related
Pt(CO)(O)3 (−260 kJ/mol), a labile O* from the CeO2
to the presence of more than one resting state (reaction
surface reacts with CO with a low barrier and completes the
intermediate complex) of Pt1 during the reaction and the
reaction cycle. The proposed mechanism is consistent with the
difficulty in quantifying the reducibility of CeO2 and its supply
reaction orders measured experimentally. The results show the
of labile O*.
complex dynamic evolution of intermediate single-atom
We should re-emphasize that both Pt1 complexes detected
complexes along the reaction cycle and provide mechanistic
experimentally in our work are part of the reaction cycle where
insights to help guide the design of Pt-based single-atom
the Pt(O)4 complex adsorbs CO weakly, and therefore need to
catalysts.


be activated into Pt(O)3 by a reaction with CO, which appears
to be the rate-limiting step. Then, to regenerate Pt(O)4,
depending on temperature, [O*], and [CO], labile O* from EXPERIMENTAL AND COMPUTATIONAL DETAILS
the CeO2 surface can either directly fill the vacancy of Pt(O)3 Catalyst Preparation and Pretreatment. High-surface-
or reacts with Pt(CO)(O)3, as shown in Scheme 1. DFT area ceria powder (obtained from Solvay, grade HSA 5) was
calculations complemented by kinetic modeling by Su et al.58 used as a support. The Pt/CeO2 catalysts (0.2 and 1 wt % Pt,
on Pt1/CeO2 showed that more than one intermediate state of nominal) were prepared by incipient wetness impregnation.
the same Pt1 active site could be present under reaction The pore volume of ceria (0.6 mL/g) was determined by first
conditions and the relative concentration of these intermedi- wetting the dry ceria powder with water. Impregnation of the
ates is temperature-dependent. Furthermore, using DFT platinum precursor was done to fill the pores of ceria. The
combined with first-principles molecular dynamics calcula- tetraamine platinum nitrate solution was loaded on ceria and
tions, Daelman et al.63 showed that Pt dynamically switches dried at 110 °C for 4 h until all of the water evaporated in air
coordination and oxidation state during the reaction cycle. after each impregnation. The sample was then calcined in a
Therefore, the experimentally measured oxidation state and Pt tube furnace with 100 mL/min air for 10 h in air at 800 °C.
coordination are a population/time-average. Additionally, the The furnace temperature was ramped up at 1 °C per minute.
work by Daelman et al.63 showed that, starting from Pt(O)4, Aberration-Corrected High-Angle Annular Dark-Field
the highest barrier in the reaction cycle is the initial reduction Scanning Transmission electron microscopy. The
of Pt(O)4 to Pt(O)3 consistent with our results and suggesting catalyst powder was mounted on Cu TEM holey carbon
that Pt(O)4 has a longer lifetime. Our results therefore films by simply dipping the TEM grid into the powder. The
highlighted the importance of using more than one in situ/ electron microscope was a JEOL ARM200CF (NEOARM)
operando characterization technique coupled with first- operated at 200 kV. The microscope is equipped with a
principles calculations for correctly assigning the (average) spherical aberration corrector (will have to look up the model
local structure of the Pt single atoms and determining whether used). The images were obtained using the high-angle annular
more than one intermediate state is present. More importantly, dark-field (HAADF) mode.
by detecting the intermediate Pt single-atom reaction Diffuse-Reflectance Infrared Fourier Transform Spec-
complexes and connecting the reaction kinetics with the troscopy (DRIFTS). DRIFTS was used to characterize the
reaction energies from DFT calculations, we provided a interaction of the supported Pt catalyst with CO, O2, and the
complete picture of the reaction mechanism and the limiting support. The in situ DRIFTS experiments were performed
steps. In particular, our results showed the potential for using a Thermo Scientific IS-50R FTIR equipped with an
8710 https://doi.org/10.1021/acscatal.1c01900
ACS Catal. 2021, 11, 8701−8715
ACS Catalysis pubs.acs.org/acscatalysis Research Article

MCT/A detector. A spectral resolution of 4 cm−1 was used to continuous Cauchy wavelet transform68 (CCWT) was used to
collect spectra, which are reported in the Kubelka−Munk decompose the EXAFS amplitude terms in R- and k-space,
(KM) units. Approximately 50 mg of sample (25−90 μm giving insight into scattering path contributions and subtle
diameter particles) was loaded in a Harrick Praying Mantis variations in the paths that may not be apparent looking at
high-temperature DRIFTS reaction chamber. The chamber only the k- or R-space. CCWT was implemented through the
was sealed and connected to a flow system with temperature Larch package.69 The k2-weighted CCWT used a k-range of
control, and gases were flown through the sample at 0−12.5 Å−1 and an R-range of 0.2−6 Å with Nfft set to 2048.
atmospheric pressure. Each reported spectrum is an average Microcalorimetry. Microcalorimetry was performed on a
of 32 scans. Pt1/CeO2 was pretreated in situ in the DRITS cell Setaram SENSYS Evo DSC calorimeter with a self-built U-
before collecting the spectra. The gas pretreatment procedure shape sample tube. The fork-shape sample tube was connected
was the same as mentioned above. For the in situ pretreated to a Micromeritics 3Flex for the adsorption quantity
catalyst, a spectrum under N2 after the pretreatment was measurement and temperature control as reported previ-
collected as the background. ously.13 For each experiment, approximately 70 mg of catalyst
X-ray Absorption Spectroscopy (XAS). The catalysts was loaded in the reactor. The gas pretreatment procedure was
were characterized by in situ XAS at the Pt L3-edge (11564 eV) the same as mentioned above in the soak mode under 760
using an in-house built cell with a 4 mm ID glassy carbon mmHg O2. After pretreatment, the catalyst was exposed to
tube.64 The XAS measurements were performed at the vacuum with pressure below 1 × 10−5 mmHg. The CO/O2
Stanford Synchrotron Radiation Light Source (SSRL) at adsorption experiment was measured at −75 °C with pressure
beamline 9-3 in fluorescence mode. Beamline 9-3 is a 16- between 0 and 300 mmHg for O2 calorimetry and 0 and 20
pole, 2-Tesla wiggler side station with a vertically collimating mmHg for CO calorimetry.
mirror for harmonic rejection and a cylindrically bend mirror CO Oxidation Kinetic Measurements. CO oxidation
for focusing. The photon energy was selected using a liquid- kinetic measurements were performed under differential
nitrogen-cooled, double-crystal Si (220) ϕ = 90° mono- conditions (<4% conversion) in a conventional laboratory
chromator. The catalyst and standard samples were scanned tubular plug flow reactor (7 mm ID quartz tube). Dilution
simultaneously in transmission and fluorescence detection experiments were performed according to the Koros−Nowak
modes using ion chambers and a 100-element solid-state Ge test to determine the necessary dilution ratio for measurements
monolith detector (Canberra). A Pt standard (Pt foil) was under strict kinetic control without mass- and heat-transfer
scanned simultaneously with each sample for energy effects.70−72 The catalysts (after intraparticle dilution) were
calibration. Step-scanning X-ray absorption spectra were pressed and sieved into a 106−250 μm diameter fraction. The
measured from 11334 to 12554 eV, corresponding to dilution ratio test to eliminate transport limitations was
photoelectron wavenumber k = 14.5 Å−1. The catalyst was performed on a more active 1% Pt/Al2O3 catalyst using SiO2
pretreated similar to before the catalytic measurements and as the diluent. Intraparticle dilutions ratio of 1:40, 1:200, and
then exposed to CO oxidation conditions. XANES and EXAFS 1:1000 showed no difference in activity under different CO
data processing and analysis were performed using Athena and and O2 conditions between 145 and 170 °C, which indicates
Artemis programs of the Demeter data analysis package.65,66 the measured catalytic activity was evaluated under kinetic
For each catalyst, four scans were collected and merged after control without transport artifacts. The 1:10 dilution ratio with
alignment. χ(k) was obtained by subtracting the smooth SiO2 was applied for the 1% Pt/CeO2 single-atom catalyst
atomic background from the normalized absorption coefficient based on its activity. Each catalyst pretreatment was the same
using the AUTOBK code. The theoretical EXAFS signals for as mentioned above, and the total flow rate during pretreat-
Pt−O, Pt−Ce, and Pt−Olong for Pt1 adsorbed on CeO2 were ment was kept at 50 sccm. The catalyst temperature was
generated using the FEFF6 code67 from a manually measured by a K-type thermocouple (OMEGA) attached at
constructed Pt atom adsorbed on CeO2. Additionally, the the center of the catalyst bed on the outside of the tube. After
theoretical EXAFS spectra were also calculated using the pretreatment, CO (5%, balance N2, certified-grade Airgas) was
optimized DFT model reported in Figure 1, and the fit results mixed with pure O2 (99.999% Airgas) and He (99.999%
were almost identical. The theoretical EXAFS signals were Airgas, equipped with moisture/O2 trap, Agilent OT3-2) at 35
fitted to the data in R-space using Artemis by varying the °C to control the partial pressures of CO and O2 (by varying
coordination numbers of the single scattering paths, the their flowrates using Brooks mass flow controllers SLA5800
effective scattering lengths, the bond length disorder, and the series). The CO gas line was equipped with a metal carbonyl
correction to the threshold energy, ΔE0 (common for all paths purifier (Matheson, NanoChem Metal-X) to remove the trace
since they are all from the same FEFF calculation). S02 (the amount of metal carbonyl and a molecular sieve 3 Å (8−12
passive electron reduction factor) was obtained by first mesh) trap to remove trace (ppm) levels of CO2. No CO2 was
analyzing the spectrum for the Pt foil, and the best fit value detected in the mass spectrometer with the CO or O2 flowing
(0.84) was fixed in the fit. The k-range used for fitting was 3− during a blank test. The reactor was heated from room
13 Å−1, while the R-range was 1.2−3.4 Å for the model that temperature to the reaction temperature at 3 °C/min in 0.5
includes only the Pt−O scattering shell and 1.2−3.4 Å for the kPa CO and 10 kPa O2 balanced with He. During the steady-
extended model that also includes the Pt−Ce and Pt−Olong state kinetic measurements, the conversion of CO was always
scattering paths. The best parameters fit using k-weights of 1, below 4% by varying the total flow rate between 50 and 100
2, and 3 (simultaneously) in Artemis are reported; however, sccm. To investigate the effect of CO partial pressure on
the results were very similar to those using only a k-weight of 2. reaction rate, the partial pressure of O2 was held constant at 10
Additionally, small differences in the k-range did not have a kPa and the partial pressure of CO was varied between 0.3 and
significant effect on the best fit values of the model fit but did 1.0 kPa. To investigate the effect of O2 partial pressure, the CO
affect their uncertainties. The wavelet transform provides partial pressure was held constant at 1.0 kPa and the partial
information for both the radical distance and in k-space. The pressure of O2 was varied between 2 and 14 kPa and then the
8711 https://doi.org/10.1021/acscatal.1c01900
ACS Catal. 2021, 11, 8701−8715
ACS Catalysis pubs.acs.org/acscatalysis Research Article

measurements were repeated at the initial condition at 1.0 kPa Authors


CO partial pressure. We note that we rigorously monitored the Yubing Lu − Department of Chemical Engineering, Virginia
catalyst stability by remeasuring the catalyst activity periodi- Polytechnic Institute and State University, Blacksburg,
cally throughout the experiment under the first condition Virginia 24060, United States; orcid.org/0000-0002-
measured. The activity was stable (<10% deactivation) for 5791-9697
each catalyst tested during the entire kinetic measurement. The Shulan Zhou − Department of Chemistry and Chemical
composition of the effluent gases was measured by a gas Biology, University of New Mexico, Albuquerque, New
chromatograph (Inficon Micro GC Fusion with two modules Mexico 87131, United States; School of Chemistry and
each with a separate carrier gas, injector, column, and thermal Chemical Engineering, Shandong University of Technology,
conductivity detector (TCD)). Column A, Rt-Molsieve 5 Å, Zibo 255049, China
0.25 mm ID (10 m) using Ar as the carrier gas. Column B, Rt- Chun-Te Kuo − Department of Chemical Engineering,
Q-Bond 0.25 mm ID (12 m) using He as the carrier gas. The Virginia Polytechnic Institute and State University,
reaction rates were normalized by the total moles of Pt, Blacksburg, Virginia 24060, United States
assuming that all Pt are on the surface. Deepak Kunwar − Center for Microengineered Materials,
DFT Calculations. All spin-polarized DFT calculations Department of Chemical and Biological Engineering,
were performed using the Vienna Ab initio Simulation Package University of New Mexico, Albuquerque, New Mexico 87131,
(VASP)73,74 with the gradient-corrected PW9175 exchange− United States; orcid.org/0000-0002-7378-0218
correlation functional. The electronic wave functions were Coogan Thompson − Department of Chemical Engineering,
expanded in plane waves up to a cutoff energy of 400 eV and Virginia Polytechnic Institute and State University,
the ionic core electrons were approximated by the projector Blacksburg, Virginia 24060, United States
augmented wave (PAW)76 method. To describe properly the Adam S. Hoffman − Stanford Synchrotron Radiation Light
behavior of Ce f electrons, the DFT + U method77,78 with U = Source, SLAC National Accelerator Laboratory, Menlo Park,
4.5 eV was used. A low-energy CeO2 monolayer-high step was California 94025, United States
selected as the slab model with the edge oriented toward the Alexey Boubnov − Stanford Synchrotron Radiation Light
[11̅0] direction, containing nine atomic layers, which is Source, SLAC National Accelerator Laboratory, Menlo Park,
consistent with the models of step II and step-U used in refs California 94025, United States
79 and 80, respectively. The unit cell was about 7.65 × 13.25 Sen Lin − Department of Chemistry and Chemical Biology,
Å2, containing 20 Ce and 40 O atoms. In this model, Pt1 University of New Mexico, Albuquerque, New Mexico 87131,
adsorbed on the step and two excess oxygen atoms were added United States; State Key Laboratory of Photocatalysis on
to obtain a stable square-planar configuration of PtO4. During Energy and Environment, College of Chemistry, Fuzhou
the structural optimization, the bottom three atomic layers University, Fuzhou 350002, China; orcid.org/0000-
were fixed, while the others as well as the adsorbates were fully 0002-2288-5415
relaxed. A vacuum space of 14 Å was employed to avoid the Abhaya K. Datye − Center for Microengineered Materials,
interaction between the neighboring interleaved slabs. To Department of Chemical and Biological Engineering,
sample the Brillouin zone, a 2 × 1 × 1 Monkhorst−Pack81 k- University of New Mexico, Albuquerque, New Mexico 87131,
point grid was adopted, which was tested to be converged. United States; orcid.org/0000-0002-7126-8659
The adsorption energy of a pertinent species was calculated Complete contact information is available at:
as follows: Eads = E(adsorbate + slab) − E(adsorbate) − E(slab). The https://pubs.acs.org/10.1021/acscatal.1c01900
transition states (TSs) between initial states (ISs) and final
states (FSs) were determined using the climb image-nudged Author Contributions
elastic band (CI-NEB) method.82 The total energy difference ∇
Y.L. and S.Z. contributed equally.
was less than 10−4 eV, and the convergence of relaxation was
checked with a 0.05 eV/Å criterion. Notes


The authors declare no competing financial interest.

*
ASSOCIATED CONTENT
sı Supporting Information
The Supporting Information is available free of charge at
■ ACKNOWLEDGMENTS
This research was primarily sponsored by the Army Research
https://pubs.acs.org/doi/10.1021/acscatal.1c01900. Office and was accomplished under Grant numbers W911NF-
Additional results (S1), possible formation mechanisms 16-1-0400 and W911NF-19-1-0308. C.-T.K. and A.M.K. also
for labile oxygens (S2), and additional evidence of the acknowledge support from the Army Research Office under
proposed reaction mechanism (S3) (PDF) Grant number W911NF-20-2-0058. The views and conclu-


sions contained in this document are those of the authors and
should not be interpreted as representing the official policies,
AUTHOR INFORMATION either expressed or implied, of the Army Research Office or the
Corresponding Authors U.S. Government. The U.S. Government is authorized to
Hua Guo − Department of Chemistry and Chemical Biology, reproduce and distribute reprints for Government purposes
University of New Mexico, Albuquerque, New Mexico 87131, notwithstanding any copyright notation herein. H.G. and
United States; orcid.org/0000-0001-9901-053X; A.K.D. acknowledge the support from the Air Force Office of
Email: hguo@unm.edu Scientific Research (Grant number FA9550-18-1-0413). We
Ayman M. Karim − Department of Chemical Engineering, also acknowledge NSF Grant DMR-1828731 for acquisition of
Virginia Polytechnic Institute and State University, the AC-STEM. S.Z. thanks the support from the National
Blacksburg, Virginia 24060, United States; orcid.org/ Natural Science Foundation of China (Grant number
0000-0001-7449-542X; Email: amkarim@vt.edu 21962007), Jiangxi Provincial Natural Science Foundation
8712 https://doi.org/10.1021/acscatal.1c01900
ACS Catal. 2021, 11, 8701−8715
ACS Catalysis pubs.acs.org/acscatalysis Research Article

(Grant number 2020BABL203009), and the Foundation of (15) Yan, H.; Cheng, H.; Yi, H.; Lin, Y.; Yao, T.; Wang, C.; Li, J.;
Jiangxi Educational Committee (Grant number GJJ190697). Wei, S.; Lu, J. Single-Atom Pd/graphene catalyst achieved by atomic
S.L. thanks the National Natural Science Foundation of China layer deposition: remarkable performance in selective hydrogenation
(Grant numbers 21973013 and 21673040). Use of the of 1,3-butadiene. J. Am. Chem. Soc. 2015, 137, 10484−10487.
(16) Lin, L.; Yao, S.; Gao, R.; Liang, X.; Yu, Q.; Deng, Y.; Liu, J.;
Stanford Synchrotron Radiation Light Source (SSRL, beamline Peng, M.; Jiang, Z.; Li, S.; Li, Y. W.; Wen, X. D.; Zhou, W.; Ma, D. A
9-3, user proposal 4645), SLAC National Accelerator highly CO-tolerant atomically dispersed Pt catalyst for chemoselective
Laboratory is supported by the U.S. Department of Energy, hydrogenation. Nat. Nanotechnol. 2019, 14, 354−361.
Office of Basic Energy Sciences under Contract No. DE-AC02- (17) Shan, J.; Li, M.; Allard, L. F.; Lee, S.; Flytzani-Stephanopoulos,
76SF00515. Additional support by the Consortium for M. Mild oxidation of methane to methanol or acetic acid on
Operando and Advanced Catalyst Characterization via supported isolated rhodium catalysts. Nature 2017, 551, 605−608.
Electronic Spectroscopy and Structure (Co-ACCESS) at (18) Matsubu, J. C.; Yang, V. N.; Christopher, P. Isolated metal
SLAC is acknowledged. Co-ACCESS is supported by the active site concentration and stability controls catalytic CO2 reduction
selectivity. J. Am. Chem. Soc. 2015, 137, 3076−3084.
U.S. Department of Energy, Office of Science, Office of Basic
(19) Herzing, A. A.; Kiely, C. J.; Carley, A. F.; Landon, P.;
Energy Sciences, Chemical Sciences, Geosciences and Hutchings, G. J. Identification of active gold nanoclusters on iron
Biosciences under Contract DE-AC02-76SF00515. oxide supports for CO oxidation. Science 2008, 321, 1331−1335.

■ REFERENCES
(1) Liu, L. C.; Corma, A. Metal catalysts for heterogeneous catalysis:
(20) Cui, X.; Li, W.; Ryabchuk, P.; Junge, K.; Beller, M. Bridging
homogeneous and heterogeneous catalysis by heterogeneous single-
metal-site catalysts. Nat. Catal. 2018, 1, 385−397.
(21) Tang, Y.; Li, Y. T.; Fung, V.; Jiang, D. E.; Huang, W. X.; Zhang,
from single atoms to nanoclusters and nanoparticles. Chem. Rev. 2018, S. R.; Iwasawa, Y.; Sakata, T.; Nguyen, L.; Zhang, X. Y.; Frenkel, A. I.;
118, 4981−5079. Tao, F. Single rhodium atoms anchored in micropores for efficient
(2) Sun, S. H.; Zhang, G. X.; Gauquelin, N.; Chen, N.; Zhou, J. G.; transformation of methane under mild conditions. Nat. Commun.
Yang, S. L.; Chen, W. F.; Meng, X. B.; Geng, D. S.; Banis, M. N.; Li, 2018, 9, No. 1231.
R. Y.; Ye, S. Y.; Knights, S.; Botton, G. A.; Sham, T. K.; Sun, X. L. (22) Chen, Y. X.; Huang, Z. W.; Ma, Z.; Chen, J. M.; Tang, X. F.
Single-atom catalysis using Pt/graphene achieved through atomic Fabrication, characterization, and stability of supported single-atom
layer deposition. Sci. Rep. 2013, 3, No. 1775. catalysts. Catal. Sci. Technol. 2017, 7, 4250−4258.
(3) Wang, A. Q.; Li, J.; Zhang, T. Heterogeneous single-atom (23) Chen, Y. J.; Ji, S. F.; Chen, C.; Peng, Q.; Wang, D. S.; Li, Y. D.
catalysis. Nat. Rev. Chem. 2018, 2, 65−81. Single-atom catalysts: synthetic strategies and electrochemical
(4) Ji, S.; Chen, Y.; Wang, X.; Zhang, Z.; Wang, D.; Li, Y. Chemical applications. Joule 2018, 2, 1242−1264.
synthesis of single atomic site catalysts. Chem. Rev. 2020, 120, 11900− (24) Kunwar, D.; Zhou, S.; De La Riva, A.; Peterson, E.; Xiong, H.;
11955. Pereira Hernandez, X. I.; Purdy, S. C.; ter Veen, R.; Brongersma, H.
(5) Qiao, B. T.; Wang, A. Q.; Yang, X. F.; Allard, L. F.; Jiang, Z.; Cui, H.; Miller, J. T.; Hashiguchi, H.; Kovarik, L.; Lin, S.; Guo, H.; Wang,
Y. T.; Liu, J. Y.; Li, J.; Zhang, T. Single-atom catalysis of CO oxidation Y.; Datye, A. Stabilizing high metal loadings of thermally stable
using Pt1/FeOx. Nat. Chem. 2011, 3, 634−641. platinum single atoms on an industrial catalyst support. ACS Catal.
(6) Lin, J.; Wang, A. Q.; Qiao, B. T.; Liu, X. Y.; Yang, X. F.; Wang, 2019, 9, 3978−3990.
X. D.; Liang, J. X.; Li, J. X.; Liu, J. Y.; Zhang, T. Remarkable (25) Jones, J.; Xiong, H. F.; Delariva, A. T.; Peterson, E. J.; Pham,
performance of ir-1/feox single-atom catalyst in water gas shift H.; Challa, S. R.; Qi, G. S.; Oh, S.; Wiebenga, M. H.; Hernandez, X. I.
reaction. J. Am. Chem. Soc. 2013, 135, 15314−15317. P.; Wang, Y.; Datye, A. K. Thermally stable single-atom platinum-on-
(7) Chen, Z. W.; Chen, L. X.; Yang, C. C.; Jiang, Q. Atomic (single, ceria catalysts via atom trapping. Science 2016, 353, 150−154.
double, and triple atoms) catalysis: frontiers, opportunities, and (26) Zhang, Z.; Zhu, Y.; Asakura, H.; Zhang, B.; Zhang, J.; Zhou,
challenges. J. Mater. Chem. A 2019, 7, 3492−3515. M.; Han, Y.; Tanaka, T.; Wang, A.; Zhang, T.; Yan, N. Thermally
(8) Liu, J. Catalysis by Supported Single Metal Atoms. ACS Catal. stable single atom Pt/m-Al2O3 for selective hydrogenation and CO
2017, 7, 34−59. oxidation. Nat. Commun. 2017, 8, No. 16100.
(9) Fu, J.; Lym, J.; Zheng, W.; Alexopoulos, K.; Mironenko, A. V.; Li, (27) Wang, Y.; Yuan, H.; Li, Y.; Chen, Z. Two-dimensional iron-
N.; Boscoboinik, J. A.; Su, D.; Weber, R. T.; Vlachos, D. G. C−O phthalocyanine (Fe-Pc) monolayer as a promising single-atom-
bond activation using ultralow loading of noble metal catalysts on catalyst for oxygen reduction reaction: a computational study.
moderately reducible oxides. Nat. Catal. 2020, 3, 446−453. Nanoscale 2015, 7, 11633−11641.
(10) Lu, Y.; Wang, J.; Yu, L.; Kovarik, L.; Zhang, X.; Hoffman, A. S.; (28) Kistler, J. D.; Chotigkrai, N.; Xu, P.; Enderle, B.; Praserthdam,
Gallo, A.; Bare, S. R.; Sokaras, D.; Kroll, T.; Dagle, V.; Xin, H.; Karim, P.; Chen, C. Y.; Browning, N. D.; Gates, B. C. A single-site platinum
A. M. Identification of the active complex for CO oxidation over CO oxidation catalyst in zeolite KLTL: microscopic and spectroscopic
single-atom Ir-on-MgAl2O4 catalysts. Nat. Catal. 2019, 2, 149−156. determination of the locations of the platinum atoms. Angew. Chem.,
(11) Lu, Y.; Kuo, C.; Kovarik, L.; Hoffman, A. S.; Boubnov, A.; Int. Ed. 2014, 53, 8904−8907.
Driscoll, D. M.; Morris, J. R.; Bare, S. R.; Karim, A. M. A versatile (29) Yang, M.; Li, S.; Wang, Y.; Herron, J. A.; Xu, Y.; Allard, L. F.;
approach for quantification of surface site fractions using reaction Lee, S.; Huang, J.; Mavrikakis, M.; Flytzani-Stephanopoulos, M.
kinetics: The case of CO oxidation on supported Ir single atoms and Catalytically active Au-O(OH)x-species stabilized by alkali ions on
nanoparticles. J. Catal. 2019, 378, 121−130. zeolites and mesoporous oxides. Science 2014, 346, 1498−1501.
(12) Zhang, L. L.; Zhou, M. X.; Wang, A. Q.; Zhang, T. Selective (30) DeRita, L.; Resasco, J.; Dai, S.; Boubnov, A.; Thang, H. V.;
hydrogenation over supported metal catalysts: from nanoparticles to Hoffman, A. S.; Ro, I.; Graham, G. W.; Bare, S. R.; Pacchioni, G.; Pan,
single atoms. Chem. Rev. 2020, 120, 683−733. X.; Christopher, P. Structural evolution of atomically dispersed Pt
(13) Kuo, C.; Lu, Y.; Kovarik, L.; Engelhard, M.; Karim, A. M. catalysts dictates reactivity. Nat. Mater. 2019, 18, 746−751.
Structure sensitivity of acetylene semi-hydrogenation on Pt single (31) Nie, L.; Mei, D. H.; Xiong, H. F.; Peng, B.; Ken, Z. B.;
atoms and subnanometer clusters. ACS Catal. 2019, 9, 11030−11041. Hernandez, X. I. P.; DeLariva, A.; Wang, M.; Engelhard, M. H.;
(14) Yan, X.; Duan, P.; Zhang, F.; Li, H.; Zhang, H.; Zhao, M.; Kovarik, L.; Datye, A. K.; Wang, Y. Activation of surface lattice oxygen
Zhang, X.; Xu, B.; Pennycook, S. J.; Guo, J. Stable single-atom in single-atom Pt/CeO2 for low-temperature CO oxidation. Science
platinum catalyst trapped in carbon onion graphitic shells for 2017, 358, 1419−1423.
improved chemoselective hydrogenation of nitroarenes. Carbon (32) Therrien, A. J.; Hensley, A. J. R.; Marcinkowski, M. D.; Zhang,
2019, 143, 378−384. R. Q.; Lucci, F. R.; Coughlin, B.; Schilling, A. C.; McEwen, J. S.;

8713 https://doi.org/10.1021/acscatal.1c01900
ACS Catal. 2021, 11, 8701−8715
ACS Catalysis pubs.acs.org/acscatalysis Research Article

Sykes, E. C. H. An atomic-scale view of single-site Pt catalysis for low- Nachtegaal, M.; Votsmeier, M.; Grunwaldt, J. D. Tuning the Pt/CeO2
temperature CO oxidation. Nat. Catal. 2018, 1, 192−198. interface by in situ variation of the Pt particle size. ACS Catal. 2018, 8,
(33) Yang, M.; Liu, J.; Lee, S.; Zugic, B.; Huang, J.; Allard, L. F.; 4800−4811.
Flytzani-Stephanopoulos, M. A Common Single-site Pt(II)-O(OH)x‑ (49) Gänzler, A. M.; Casapu, M.; Doronkin, D. E.; Maurer, F.; Lott,
species stabilized by sodium on “active” and “inert” supports catalyzes P.; Glatzel, P.; Votsmeier, M.; Deutschmann, O.; Grunwaldt, J.-D.
the water-gas shift reaction. J. Am. Chem. Soc. 2015, 137, 3470−3473. Unravelling the different reaction pathways for low temperature CO
(34) Cao, L.; Liu, W.; Luo, Q.; Yin, R.; Wang, B.; Weissenrieder, J.; oxidation on Pt/CeO2 and Pt/Al2O3 by spatially resolved structure−
Soldemo, M.; Yan, H.; Lin, Y.; Sun, Z.; Ma, C.; Zhang, W.; Chen, S.; activity correlations. J. Phys. Chem. Lett. 2019, 10, 7698−7705.
Wang, H.; Guan, Q.; Yao, T.; Wei, S.; Yang, J.; Lu, J. Atomically (50) An, K.; Alayoglu, S.; Musselwhite, N.; Plamthottam, S.; Melaet,
dispersed iron hydroxide anchored on Pt for preferential oxidation of G.; Lindeman, A. E.; Somorjai, G. A. Enhanced CO Oxidation rates at
CO in H2. Nature 2019, 565, 631−635. the interface of mesoporous oxides and Pt nanoparticles. J. Am. Chem.
(35) Li, Z.; Ji, S. F.; Liu, Y. W.; Cao, X.; Tian, S. B.; Chen, Y. J.; Niu, Soc. 2013, 135, 16689−16696.
Z. G.; Li, Y. D. Well-Defined Materials for heterogeneous catalysis: (51) Bunluesin, T.; Gorte, R. J.; Graham, G. W. CO oxidation for
from nanoparticles to isolated single-atom sites. Chem. Rev. 2020, 120, the characterization of reducibility in oxygen storage components of
623−682. three-way automotive catalysts. Appl. Catal., B 1997, 14, 105−115.
(36) Maurer, F.; Jelic, J.; Wang, J.; Gänzler, A.; Dolcet, P.; Wöll, C.; (52) Bunluesin, T.; Gorte, R. J.; Graham, G. W. Studies of the water-
Wang, Y.; Studt, F.; Casapu, M.; Grunwaldt, J.-D. Tracking the gas-shift reaction on ceria-supported Pt, Pd, and Rh: implications for
formation, fate and consequence for catalytic activity of Pt single sites oxygen-storage properties. Appl. Catal., B 1998, 15, 107−114.
on CeO2. Nat. Catal. 2020, 3, 824−833. (53) Pereira-Hernández, X. I.; DeLaRiva, A.; Muravev, V.; Kunwar,
(37) Montini, T.; Melchionna, M.; Monai, M.; Fornasiero, P. D.; Xiong, H.; Sudduth, B.; Engelhard, M.; Kovarik, L.; Hensen, E. J.
Fundamentals and catalytic applications of CeO2-based materials. M.; Wang, Y.; Datye, A. K. Tuning Pt-CeO2 interactions by high-
Chem. Rev. 2016, 116, 5987−6041. temperature vapor-phase synthesis for improved reducibility of lattice
(38) Nagai, Y.; Dohmae, K.; Ikeda, Y.; Takagi, N.; Tanabe, T.; Hara, oxygen. Nat. Commun. 2019, 10, No. 1358.
N.; Guilera, G.; Pascarelli, S.; Newton, M. A.; Kuno, O.; Jiang, H. Y.; (54) Xia, Z.; Zhang, H.; Shen, K.; Qu, Y.; Jiang, Z. Wavelet analysis
Shinjoh, H.; Matsumoto, S. In situ redispersion of platinum of extended X-ray absorption fine structure data: theory, application.
autoexhaust catalysts: an on-line approach to increasing catalyst Phys. B 2018, 542, 12−19.
lifetimes? Angew. Chem., Int. Ed. 2008, 47, 9303−9306. (55) Timoshenko, J.; Kuzmin, A. Wavelet data analysis of EXAFS
(39) Kottwitz, M.; Li, Y.; Palomino, R. M.; Liu, Z.; Wang, G.; Wu, spectra. Comput. Phys. Commun. 2009, 180, 920−925.
Q.; Huang, J.; Timoshenko, J.; Senanayake, S. D.; Balasubramanian, (56) Ding, K.; Gulec, A.; Johnson, A. M.; Schweitzer, N. M.; Stucky,
M.; Lu, D.; Nuzzo, R. G.; Frenkel, A. I. Local structure and electronic G. D.; Marks, L. D.; Stair, P. C. Identification of active sites in CO
state of atomically dispersed Pt supported on nanosized CeO2. ACS
oxidation and water-gas shift over supported Pt catalysts. Science
Catal. 2019, 9, 8738−8748.
2015, 350, 189−192.
(40) Resasco, J.; DeRita, L.; Dai, S.; Chada, J. P.; Xu, M. J.; Yan, X.
(57) Alexopoulos, K.; Wang, Y. F.; Vlachos, D. G. First-principles
X.; Finzel, J.; Hanukovich, S.; Hoffman, A. S.; Graham, G. W.; Bare, S.
kinetic and spectroscopic insights into single-atom catalysis. ACS
R.; Pan, X. Q.; Christopher, P. Uniformity is key in defining structure-
Catal. 2019, 9, 5002−5010.
function relationships for atomically dispersed metal catalysts: the
(58) Su, Y.-Q.; Wang, Y.; Liu, J.-X.; Filot, I. A. W.; Alexopoulos, K.;
case of Pt/CeO2. J. Am. Chem. Soc. 2020, 142, 169−184.
Zhang, L.; Muravev, V.; Zijlstra, B.; Vlachos, D. G.; Hensen, E. J. M.
(41) Tang, Y.; Wang, Y.-G.; Li, J. Theoretical investigations of Pt1@
CeO2 single-atom catalyst for CO oxidation. J. Phys. Chem. C 2017, Theoretical approach to predict the stability of supported single-atom
121, 11281−11289. catalysts. ACS Catal. 2019, 9, 3289−3297.
(42) Lu, Y. B.; Thompson, C.; Kunwar, D.; Datye, A. K.; Karim, A. (59) Sravan Kumar, K. B.; Deshpande, P. A. On Identification of
M. Origin of the high CO oxidation activity on CeO2 supported Pt labile oxygen in ceria-based solid solutions: which oxygen leaves the
nanoparticles: weaker binding of CO or facile oxygen transfer from lattice? J. Phys. Chem. C 2015, 119, 8692−8702.
the support? ChemCatChem 2020, 12, 1726−1733. (60) Migani, A.; Vayssilov, G. N.; Bromley, S. T.; Illas, F.; Neyman,
(43) Ferré, G.; Aouine, M.; Bosselet, F.; Burel, L.; Aires, F. J. C. S.; K. M. Greatly facilitated oxygen vacancy formation in ceria
Geantet, C.; Ntais, S.; Maurer, F.; Casapu, M.; Grunwaldt, J. D.; nanocrystallites. Chem. Commun. 2010, 46, 5936−5938.
Epicier, T.; Loridant, S.; Vernoux, P. Exploiting the dynamic (61) Di Benedetto, A.; Landi, G.; Lisi, L. CO reactive adsorption at
properties of Pt on ceria for low-temperature CO oxidation. Catal. low temperature over CuO/CeO2 structured catalytic monolith. Int. J.
Sci. Technol. 2020, 10, 3904−3917. Hydrogen Energy 2017, 42, 12262−12275.
(44) Meunier, F. C.; Cardenas, L.; Kaper, H.; Smid, B.; Vorokhta, (62) Yoo, M.; Yu, Y. S.; Ha, H.; Lee, S.; Choi, J. S.; Oh, S.; Kang, E.;
M.; Grosjean, R.; Aubert, D.; Dembele, K.; Lunkenbein, T. Synergy Choi, H.; An, H.; Lee, K. S.; Park, J. Y.; Celestre, R.; Marcus, M. A.;
between metallic and oxidized Pt sites unravelled during room Nowrouzi, K.; Taube, D.; Shapiro, D. A.; Jung, W.; Kim, C.; Kim, H.
temperature CO oxidation on Pt/ceria. Angew. Chem., Int. Ed. 2021, Y. A tailored oxide interface creates dense Pt single-atom catalysts
60, 3799−3805. with high catalytic activity. Energy Environ. Sci. 2020, 13, 1231−1239.
(45) Xiao, Y. C.; Li, H.; Xie, K. Activating lattice oxygen at the (63) Daelman, N.; Capdevila-Cortada, M.; Lopez, N. Dynamic
twisted surface in a mesoporous CeO2 single crystal for efficient and charge and oxidation state of Pt/CeO2 single-atom catalysts. Nat.
durable catalytic CO oxidation. Angew. Chem., Int. Ed. 2021, 60, Mater. 2019, 18, 1215−1221.
5240−5244. (64) Karim, A. M.; Howard, C.; Roberts, B.; Kovarik, L.; Zhang, L.;
(46) Feng, Y.; Wan, Q.; Xiong, H.; Zhou, S.; Chen, X.; Pereira King, D. L.; Wang, Y. In Situ X-ray Absorption Fine Structure studies
Hernandez, X. I.; Wang, Y.; Lin, S.; Datye, A. K.; Guo, H. Correlating on the effect of pH on Pt electronic density during aqueous phase
DFT calculations with CO oxidation reactivity on Ga-doped Pt/CeO2 reforming of glycerol. ACS Catal. 2012, 2, 2387−2394.
single-atom catalysts. J. Phys. Chem. C 2018, 122, 22460−22468. (65) Ravel, B.; Newville, M. ATHENA, ARTEMIS, HEPHAESTUS:
(47) Gänzler, A. M.; Casapu, M.; Vernoux, P.; Loridant, S.; Aires, F. data analysis for X-ray absorption spectroscopy using IFEFFIT. J.
J. C. S.; Epicier, T.; Betz, B.; Hoyer, R.; Grunwaldt, J. D. Tuning the Synchrotron Radiat. 2005, 12, 537−541.
structure of platinum particles on ceria insitu for enhancing the (66) Newville, M. IFEFFIT: interactive XAFS analysis and FEFF
catalytic performance of exhaust gas catalysts. Angew. Chem., Int. Ed. fitting. J. Synchrotron Radiat. 2001, 8, 322−324.
2017, 56, 13078−13082. (67) Zabinsky, S. I.; Rehr, J. J.; Ankudinov, A.; Albers, R. C.; Eller,
(48) Gänzler, A. M.; Casapu, M.; Maurer, F.; Stormer, H.; Gerthsen, M. J. Multiple-scattering calculations of X-ray-absorption spectra.
D.; Ferre, G.; Vernaux, P.; Bornmann, B.; Frahm, R.; Murzin, V.; Phys. Rev. B 1995, 52, 2995−3009.

8714 https://doi.org/10.1021/acscatal.1c01900
ACS Catal. 2021, 11, 8701−8715
ACS Catalysis pubs.acs.org/acscatalysis Research Article

(68) Muñoz, M.; Argoul, P.; Farges, F. Continuous Cauchy wavelet


transform analyses of EXAFS spectra: a qualitative approach. Am.
Mineral. 2003, 88, 694−700.
(69) Newville, M. Larch: An Analysis Package for XAFS and related
spectroscopies. J. Phys.: Conf. Ser. 2013, 430, No. 012007.
(70) Bonet, F.; Grugeon, S.; Dupont, L.; Urbina, R. H.; Boudart, M.;
Djega-Mariadassou, G. Kinetics of Heterogeneous Catlytic Reactions;
Princeton University Press: Princeton, N.J., 1984.
(71) Koros, R. M.; Nowak, E. J. A diagnostic test of the kinetic
regime in a packed bed reactor. Chem. Eng. Sci. 1967, 22, 470.
(72) Madon, R. J.; Boudart, M. Experimental criterion for the
absence of artifacts in the measurement of rates of heterogeneous
catalytic reactions. Ind. Eng. Chem. Fundam. 1982, 21, 438−447.
(73) Kresse, G.; Furthmuller, J. Efficient iterative schemes for ab
initio total-energy calculations using a plane-wave basis set. Phys. Rev.
B 1996, 54, 11169−11186.
(74) Kresse, G.; Furthmuller, J. Efficiency of ab-initio total energy
calculations for metals and semiconductors using a plane-wave basis
set. Comput. Mater. Sci. 1996, 6, 15−50.
(75) Perdew, J. P.; Chevary, J. A.; Vosko, S. H.; Jackson, K. A.;
Pederson, M. R.; Singh, D. J.; Fiolhais, C. Atoms, molecules, solids,
and surfaces: applications of the generalized gradient approximation
for exchange and correlation. Phys. Rev. B 1992, 46, 6671−6687.
(76) Blöchl, P. E. Projector Augmented-Wave Method. Phys. Rev. B
1994, 50, 17953−17979.
(77) Krcha, M. D.; Janik, M. J. Challenges in the use of density
functional theory to examine catalysis by M-doped ceria surfaces. Int.
J. Quantum. Chem. 2014, 114, 8−13.
(78) McFarland, E. W.; Metiu, H. Catalysis by doped oxides. Chem.
Rev. 2013, 113, 4391−4427.
(79) Dvořák, F.; Camellone, M. F.; Tovt, A.; Tran, N.-D.; Negreiros,
F. R.; Vorokhta, M.; Skála, T.; Matolínová, I.; Mysliveček, J.; Matolín,
V.; Fabris, S. Creating single-atom Pt-ceria catalysts by surface step
decoration. Nat. Commun. 2016, 7, No. 10801.
(80) Liu, J.-C.; Wang, Y.-G.; Li, J. Toward Rational Design of oxide-
supported single-atom catalysts: atomic dispersion of gold on ceria. J.
Am. Chem. Soc. 2017, 139, 6190−6199.
(81) Monkhorst, H. J.; Pack, J. D. Special points for Brillouin-zone
integrations. Phys. Rev. B 1976, 13, 5188−5192.
(82) Henkelman, G.; Uberuaga, B. P.; Jónsson, H. A climbing image
nudged elastic band method for finding saddle points and minimum
energy paths. J. Chem. Phys. 2000, 113, 9901−9904.

8715 https://doi.org/10.1021/acscatal.1c01900
ACS Catal. 2021, 11, 8701−8715

You might also like