Surface Density Dependent Catalytic Activity of Single Palladium Atoms Supported On Ceria

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Research Articles Angewandte

Chemie

How to cite: Angew. Chem. Int. Ed. 2021, 60, 22769– 22775

15213773, 2021, 60, Downloaded from https://onlinelibrary-wiley-com.ezproxy.lib.vt.edu By Virginia Tech- on [28/10/2021]. Re-use and distribution is strictly not permitted, except for Open Access articles
Single-Atom Catalysis Hot Paper International Edition: doi.org/10.1002/anie.202105750
German Edition: doi.org/10.1002/ange.202105750

Surface Density Dependent Catalytic Activity of Single Palladium


Atoms Supported on Ceria**
Yongseon Kim+, Greg Collinge+, Mal-Soon Lee, Konstantin Khivantsev, Sung June Cho,*
Vassiliki-Alexandra Glezakou, Roger Rousseau,* Janos Szanyi,* and Ja Hun Kwak*

Abstract: The analogy between single-atom catalysts (SACs) synergistic nonlocal effects. This phenomenon, however, is
and molecular catalysts predicts that the specific catalytic support-dependent as Pd SACs supported on non-reducible
activity of these systems is constant. We provide evidence that g-Al2O3 exhibit constant specific activity for CO oxidation.
this prediction is not necessarily true. As a case in point, we Numerous SACs have been characterized and reported to
show that the specific activity over ceria-supported single Pd exhibit high activity for the CO oxidation reaction including
atoms linearly increases with metal atom density, originating the seminal Pt/FeOx catalyst.[4] The charge state of the single-
from the cumulative enhancement of CeO2 reducibility. The atom centers is commonly invoked to explain this activity,[5]
long-range electrostatic footprints (  1.5 nm) around each Pd reportedly enhancing CO and/or O2 adsorption. Subsequent-
site overlap with each other as surface Pd density increases, ly, the choice of support and its redox state has been shown to
resulting in an observed deviation from constant specific impact CO oxidation activity greatly,[6] with irreducible
activity. These cooperative effects exhaust previously active O supports, such as g-Al2O3 or MgO, identified as comparatively
atoms above a certain Pd density, leading to their permanent inactive[7] or unstable[8] without modification such as the
removal and a consequent drop in reaction rate. The findings induction of defect sites[9] or the addition of La[10] or other
of our combined experimental and computational study show heteroatom substituents.[11] Conversely, single atoms on
that the specific catalytic activity of reducible oxide-supported reducible supports show considerable CO oxidation activi-
single-atom catalysts can be tuned by varying the surface ty.[12] Some supports have even been reported to allow the
density of single metal atoms. stabilization of metallic single atoms.[13] Spezzati et al. re-
ported high CO-oxidation activity of isolated Pd atoms
Introduction supported on CeO2 (111) and identified a PdOx species as
the active site.[14] However, despite considerable research
Supported single-atom catalysts (SACs) have attracted efforts on SACs, the effect of surface density of single metal
considerable attention owing to their unique activity, efficient atoms on catalytic activity has not been elucidated.
utilization of supported noble metals, and potential to bridge In this study, we report the synergistic effect between
the gap between homogeneous and heterogeneous catalysis. [1] single Pd atoms supported on CeO2. H2 -temperatrue pro-
Increasing the surface metal density of SACs is sought after to grammed reduction (TPR) and in situ Raman studies em-
increase the density of active sites and the mass activity of ployed in this work clearly show that the reducibility of the
industrial catalysts. [2] Meanwhile, it is commonly assumed that support is enhanced as the density of single Pd atoms
the active sites of SACs exhibit constant specific activity and increases. Density functional theory (DFT) calculations
the activity decreases at high surface metal density due to confirm the lowering in oxygen vacancy formation energies
agglomeration of single atoms and nanoparticle formatio- as the surface Pd density increases, while CO and O2
n.[1a, 2g, 3] Here using CO oxidation as a probe reaction, we adsorption energies remain unaffected. This suggests that
show that, contrary to this assumption, the specific activity of the observed specific activity trends are due to the activation
Pd/CeO2 SACs instead increases with Pd density due to of lattice O by Pd. Charge and spin density analysis allow us to

[*] Y. Kim, [+] Prof. J. H. Kwak E-mail: Janos.Szanyi@pnnl.gov


Department of Chemical Engineering Prof. S. J. Cho
Ulsan National Institute of Science and Technology (UNIST) Department of Chemical Engineering
50 UNIST-gil, Ulsan 44919 (Republic of Korea) Chonnam National University
E-mail: jhkwak@unist.ac.kr 77 Yongbong-ro, Buk-gu, Gw angju 61186 (Republic of Korea)
Dr. G. Collinge,[+] Dr. M.-S. Lee, Dr. V.-A. Glezakou, Dr. R. Rousseau E-mail: sjcho@chonnam.ac.kr
Physical and Computational Sciences Directorate and Institute for [+] These authors contributed equally to this work.
Integrated Catalysis, Pacific Northwest National Laboratory
[**] A previous version of this manuscript has been deposited on
Richland, WA 99354 (USA) a preprint server (https ://doi.org/10.26434/chemrxiv.14312780.v1).
E-mail: Roger.Rousseau@pnnl.gov
Supporting information and the ORCID identification number(s) for
Dr. G. Collinge, [+] Dr. M.-S. Lee, Dr. K. Khivantsev, Dr. V.-A. Glezakou,
the author(s) of this article can be found under :
Dr. R. Rousseau, Dr. J. Szanyi https://doi.org/10.1002/anie.202105750.
Institute for Integrated Catalysis
Pacific Northwest National Laboratory
Richland, WA 99354 (USA)

Angew. Chem. Int. Ed. 2021, 60, 22769 – 22775  2021 Wiley-VCH GmbH 22769
Angewandte
Research Articles Chemie

15213773, 2021, 60, Downloaded from https://onlinelibrary-wiley-com.ezproxy.lib.vt.edu By Virginia Tech- on [28/10/2021]. Re-use and distribution is strictly not permitted, except for Open Access articles
identify the active site as an overoxidized Pd>2+ in a [PdO4 ]
square planar complex consistent with X-ray absorption
spectroscopy (XAS). The Pd atom is overoxidized, past + 2,
and its coordinating oxygens are valence unsaturated, making
[PdO 4] an excellent oxidizer with a considerable and impor-
tantly, cumulative, range of effect. Thus, as the surface density
of [PdO4] complexes increases, their individual oxidation
power also increases. The support mediates this process,
allowing for the shuttling of excess charge to nearby [PdO4],
explaining why g-Al2 O 3, a non-reducible support, does not
exhibit the same synergy.

Results and Discussion

CO oxidation on Pd/CeO 2

The effect of surface single-atom density on the catalytic


properties of Pd/CeO2 (for characterization of CeO2 see the
Supporting Information, Figures S1–S3) and Pd/Al2O 3 cata-
lysts was investigated in the model reaction of CO oxidation
as a function of surface Pd density, displayed in Figure 1 a
(Figure S4a shows the relationship as a function of Pd weight
loading).
The specific activity of Pd/CeO2 catalysts increased
linearly with surface Pd density up to  0.8 Pd/nm2. This
behavior was observed for both initial (Figure S4b) and
steady state activities (Figure 1 a). In the contrasting case of
Pd/Al 2O3 catalysts, the specific activity was constant below
 0.034 Pd/nm 2, in agreement with previously reported re- Figure 1. a) Specific activity of CO oxidation at steady state as
sults.[1a, 3a,b] Specific activity as a function of time for a selected a function of surface Pd density of Pd/CeO2 and Pd/Al2 O3 . b) Specific
series of Pd/CeO 2 catalysts is also shown in Figure 1 b (for the activity of Pd/CeO2 as a function of time. c) Arrhenius plots of 0.1, 0.2
entire series of Pd/CeO2 and Pd/Al2O3 catalysts, see Fig- and 0.4 Pd/CeO2 catalysts for CO oxidation.
ure S4c,d). The specific activity was obtained while the CO
conversion is kept below 10% (Figure S4e,f). Pd/CeO2
catalysts with Pd surface densities in the range of 0.1– between Pd (z = 46) and Ce (z = 58), we could not observe
0.8 Pd/nm 2 deactivated slowly while catalysts with surface atomically dispersed Pd supported on CeO2 (Figures S5–
densities between 1.2–4 Pd/nm 2 underwent rapid deactiva- S8).[16] For 0.068 Pd/Al2O 3, atomically dispersed Pd were
tion. Sintering of Pd atoms/small metal clusters and, as we will observed (Figure S9). At higher Pd surface densities (e.g.,
show, the permanent removal of previously catalytically 0.20 Pd/Al2 O 3) distinctive Pd nanoparticles were observed
active lattice O atoms are primarily responsible for this (Figure S10).
activity loss. The apparent activation energies for Pd/CeO2 Since we could not directly confirm the presence of
catalysts with Pd surface densities of 0.1, 0.2 and 0.4 Pd/nm 2, atomically dispersed Pd on CeO2 with AC -HAADF-STEM,
estimated from Arrhenius plots in Figure 1 c, were 45, 35 and additional structural analysis was carried out. Pd K-edge
40 kJ mol 1, respectively. These values are consistent with X-ray absorption near-edge structure (XANES) (Figure S11)
a Mars-van Krevelen (MvK) mechanism. [15] It is important to and extended x-ray absorption fine structure (EXAFS) data
note, however, that based on these data, it is the preexpo- (Figures S12 and S13) were further used to estimate structural
nential factor that is enhanced as surface Pd density increases, parameters (Table 1). The Fourier transformed EXAFS
suggesting the observed effect is entropic in origin and not spectra show primarily a Pd-O scattering at 1.99  with
due to a lowering of the activation
barrier.
Aberration-corrected high an- Table 1: Structural parameters from Pd K-edge EXAFS analysis for 0.8 and 4 Pd/CeO2 .[a]
gle annular dark field scanning Sample Pair CN [b] r s2 DE R-factor
[c]
transmission electron microscopy [] [pm2] [d] [eV][e] [%]
(AC-HAADF-STEM) imaging 0.8 Pd/CeO 2 Pd-O 3.8  0.6 1.99  0.02 52  11 4.9  2.5 3.0
was used to visualize Pd atoms/ 4 Pd/CeO 2 Pd-O 3.7  0.5 1.99  0.01 11  18 4.3  1.2 1.3
particles in the CeO2 and Al2O 3- [a] Many-body reduction factor ; S 2 was set to 0.86. The number of the free fitting parameter for EXAFS
0
supported catalysts (Figures S5– data ; 11 was always less than the number of independent points given by the Nyquist Theorem.
S10). Because of low z-contrast [b] Coordination number. [c] Bond distance. [d] The Debye-Waller factor and [e] energy shift.

22770 www.angewandte.org  2021 Wiley-VCH GmbH Angew. Chem. Int. Ed. 2021, 60, 22769 – 22775
Angewandte
Research Articles Chemie

15213773, 2021, 60, Downloaded from https://onlinelibrary-wiley-com.ezproxy.lib.vt.edu By Virginia Tech- on [28/10/2021]. Re-use and distribution is strictly not permitted, except for Open Access articles
coordination number of 3.8  0.6 and 3.7  0.5 for 0.8 and
4 Pd/CeO2, respectively. In bulk PdO, Pd-Pd scattering
appears around 3.06 and 3.45  with 4 and 8 coordination
number, respectively. [17] However, no Pd-Pd bond was
observed in either 0.8 or 4 Pd/CeO 2 samples. Interestingly,
the corresponding XANES is shifted to higher energies
relative to that of PdO, indicating Pd to be in an oxidation
state greater than + 2. These results suggest that Pd is highly
dispersed with Pd-O coordination number around 4 in an
overoxidized charge state.

CeO 2 Redox Properties Modified by Pd loading

Since CeO 2-based CO oxidation catalysts follow the MvK


mechanism, [15e, 18] the activity must be related to the redox
properties of the catalysts. To gain information about this, we
conducted H2-TPR (Figure 2 a for Pd/CeO2). On bare CeO 2,
surface and bulk reduction peaks at 48088C and above 700 8C
were observed, respectively.[15e] With Pd present, the TPR
profile dramatically changes. The intensity of the reduction
peak at 480 8 C decreased and a new reduction peak appeared
at 162 8C for 0.1 Pd/CeO2 . Interestingly, the reduction peak
gradually shifted toward lower temperature with increasing
Pd surface density (Figure 2 b), ultimately appearing at 93 8C
for 0.4 Pd/ CeO 2 .[19] At even higher surface Pd density, very Figure 2. a) H2 -TPR profiles of Pd/CeO2 and b) reduction temperatures
sharp and intense reduction peak arose under 50 8C. as a function of surface Pd density.
H2-TPR clearly reveals that the hydrogen consumption
(Supporting Information, Table S1) far exceeded the amount
of Pd loaded as CeO2 support is reduced. This result suggests Raman spectroscopy on the 0.2, 0.4 and 4 Pd/CeO2 catalysts at
a causal relationship between reducibility of CeO 2 support room temperature (RT), 6288C and 125 8 C. The results
and surface Pd density. To investigate the origin of initial CO obtained are displayed in Figure 3 (results of 0.1 and
oxidation activity drop shown in Figure S4c, H2 -TPR was also 1.2 Pd/CeO 2 can be found in Figure S19). All samples were
conducted on spent catalysts after CO oxidation for 2 h at pretreated at 35088C in 20 % O 2/He for 0.5 h. A Raman band
75 8C (Figure S14, Table S2). At higher Pd loading ( 0.8 representative of CeO2 at 450–470 cm 1 was observed along
Pd/nm 2), the reduction profile significantly changed, which with weak bands at 270, 600 and 1174 cm 1 (Figures S20 and
can be explained by sintering and partial reduction of Pd S21). [18a, 22] In CO + O2 , a distinct band at 830 cm1 was
during reaction (Table S3, Figures S15–S17). This may be also observed for 0.2 and 0.4 Pd/CeO 2 catalysts. On 4 Pd/CeO2
due to lattice oxygen deactivation, which we will
return to shortly. We also conducted H 2-TPR on
Pd/Al 2O3 (Figure S18). We could hardly observe
H2 consumption on 0.017–0.068 Pd/Al2O3 due to
very low Pd loading. At higher Pd loading
(0.20 Pd/Al2O 3 and 0.68 Pd/Al2O 3), a broad re-
duction peak centered at  87 8 C was observed.
At the same time, additional sharp reduction peak
at 23.5 8 C was observed on 0.68 Pd/Al 2O3, well
beyond the point at which nanoparticles form. A
negative peak at 60 8C of 0.68 Pd/Al 2O3 is attrib-
uted to decomposition of palladium hydride
species which is hardly formed on small parti-
cles.[20] The absence of reduction peak at 23.5 8 C
and feature of hydride decomposition on 0.017–
0.20 Pd/Al2 O3 indicates that Pd is highly dispersed
Figure 3. In situ Raman spectra of a) 0.2 Pd/CeO2, b) 0.4 Pd/CeO 2, c) 4 Pd/CeO2.
on 0.017–0.20 Pd/Al 2O 3.[20b, 21] Notably, Pd/Al2O 3
Conditions : (1) room temperature (RT)—He, (2) RT 1 %CO + 2.5 %O2 /He, (3) RT
catalysts do not show systematic change of
2.5 %O2/He, (4) 62 8 C/%CO + 2.5 %O2/He, (5) 62 8C2.5 %O 2/He, (6) 125 8C/
reducibility with Pd loading. %CO + 2.5 %O2/He, and (7) 125 8C2.5 %O 2 /He). Raman intensity was normalized
To investigate the observed CeO2 reduction with F2g (450–470 cm 1) band. The same scale of y-axis is used across Figure
under reaction conditions, we performed in situ panels.

Angew. Chem. Int. Ed. 2021, 60, 22769 – 22775  2021 Wiley-VCH GmbH www.angewandte.org 22771
Angewandte
Research Articles Chemie

15213773, 2021, 60, Downloaded from https://onlinelibrary-wiley-com.ezproxy.lib.vt.edu By Virginia Tech- on [28/10/2021]. Re-use and distribution is strictly not permitted, except for Open Access articles
catalyst, this Raman band was rather intense after pre- is present as an adsorbed species (sharp IR band at 2351 cm1 )
treatment and He flushing, but this band decreased in the on the atomically dispersed Pd/CeO2 samples, we only
presence of CO + O2 at room temperature. The 830 cm 1 observe the formation of gas phase CO2 (2349 cm 1) over
band has been attributed to adsorbed peroxo species (O 22) the Pd particles. We propose that the origin of CO 2 formed
on isolated two-electron defect sites, indicating oxygen over the single-atom-containing samples is fundamentally
vacancies are formed on Pd/CeO 2 under reaction condi- different from that formed on supported Pd particles. The
tions. [22a] As CO is removed from the gas stream, the results of DFT calculations, which will be discussed below,
characteristic Raman band for the peroxo species disap- clearly demonstrate that the presence of single Pd atoms
peared on the single-atom-containing catalysts (0.2 and significantly reduces the energy of oxygen vacancy formation
0.4 Pd/CeO2 ). At 62 8 C, the Raman band of peroxo species on CeO2 . This, in turn, means that certain oxygen species on
re-appeared in the CO + O 2 stream. The most noticeable the CeO2 surface in the vicinity of single Pd atoms become
feature is the development of new bands at 1104 and very reactive toward CO, and readily produce CO2. Since
1514 cm1 for 0.4 Pd/CeO2 . These bands are assigned to these experiments are carried out at room temperature the
superoxide (O 2) and weakly bound dioxygen (O 2d) species, thus-formed CO 2 stays on the CeO2 surface interacting with
respectively.[23] According to Hess et al., [23] those two dioxy- the oxygen vacancies. In contrast, when Pd particles are
gen species can be attributed to the creation of oxygen present on the CeO2 support, CO 2is formed by the reaction of
vacancies during CO oxidation and the transfer of electron to CO with the surface oxide layer formed on the Pd particles
adsorbed oxygen molecule. As the temperature was further during the 623 K oxidation prior to CO adsorption. This
increased to 125 8C, the peroxo peak (830 cm 1) appeared in reaction has two consequences: 1. Metallic Pd formation upon
a weak intensity under CO + O2 flow. Features for superoxide the removal of the surface oxide layer (the IR bands of
(1104 cm 1) and weakly bound dioxygen (1514 cm 1) ap- adsorbed CO are consistent with metallic adsorption sites)
peared on 0.2 Pd/CeO2 , but their intensities were much lower and 2. Fo rmation of gas phase CO2 (and even some surface
than in the 0.4 Pd/CeO2 catalyst. When CO was taken out of carbonates via the reaction between CO2 and the ceria
the gas stream, the two features became very prominent on support). However, no adsorbed CO2 is seen, as the concen-
both 0.2 and 0.4 Pd/CeO2 catalysts, which might be due to tration of highly labile surface oxygen species in this catalyst
reaction of the charged oxygen species (O2 or O 2 d) with gas is very low due to the low number of single Pd atoms.
phase CO or removal of adsorbed CO on oxygen vacancies.
On the other hand, as the surface Pd density further increased
to 4, the intensity of the dioxygen species became significantly Calculated Properties of the Pd/CeO2 system
lower than over 0.4 Pd/CeO2 . The comparison of Raman
spectra of 0.2 and 0.4 Pd/CeO2 catalysts clearly shows the Density functional theory calculations were employed to
facile formation of oxygen vacancies on CeO2 surface with provide molecular level insights into the effect of surface Pd
higher surface density of single Pd atoms in reaction density on CO oxidation activity. Since experiments use
conditions. This result manifests itself in improved oxygen specifically CeO2 nanocubes (Figure S3) as support, the
vacancy formation with surface density of atomic Pd, dominant oxygen-terminated CeO2 (100) facet was chosen as
consistent with results from H2-TPR. the model system. The 4-fold hollow sites of this surface can
In situ transmission FTIR spectroscopy was used to stabilize Pd adatoms. However, as shown in Figure S23, upon
monitor both the nature of Pd species present on the CeO2 deposition of single Pd atoms, exposing the stoichiometric
support after oxidation at 623 K and the variation of the CeO2(100) surface to any partial pressure of oxygen will result
reducibility of CeO 2 as a function of Pd loading using CO as in the irreversible adsorption of one additional O2 molecule
a probe molecule. The IR spectra collected during sequential per Pd atom to form (PdO2) 1 single-metal-centers, see Fig-
CO adsorption over Pd/CeO2 catalysts with Pd loadings of 0.2, ure S23a. These (PdO2 ) 1 reside in a square planar config-
0.4 and 4 Pd/nm2 are displayed in Figure S22. IR bands uration, with a 4-fold Pd coordination to surface-bound
characteristic of Pd ions are observed exclusively for the oxygen atoms, labeled O1–O4 in Figure 4 a, Pd O bond
0.2 Pd/CeO2 sample with characteristic IR features centered lengths of 1.94–1.98 . This is in excellent agreement with the
between 2098 and 2154 cm1. The 0.4 Pd/CeO 2 catalyst EXAFS results shown in Table 1 (CN of 3.8 and PdO bond
exhibits similar IR features of adsorbed CO. However, the length of 1.99 ). Increasing the surface Pd density therefore
intensity ratio of the high and low frequency bands changes corresponds to decreasing the spatial separation of these
dramatically as the Pd loading increases, indicating changes in [PdO4 ] sites.
the population of Pd ions in different environments on(in) the Through Bader analysis of the electron and spin density
CeO 2 surface. The series of IR spectra collected from the around each atom in and surrounding the [PdO 4] complex
4 Pd/CeO2 sample is fundamentally different from those of (for details, see the Supporting Information), we are able to
the other two catalysts: they are dominated by IR features of determine their electronic states. We find that Pd is in an
CO adsorbed on metallic Pd sites (Pd particles). IR bands at unusually overoxidized + 2.6 state, which is consistent with
2082–2098 and at 1917–1952 cm 1 represent CO molecules our XANES results, and its surrounding O atoms are in
bound linearly and in a bridging configuration to Pd0 centers markedly unsaturated charge states of  1.7. The incom-
of metal particles, respectively. It is very interesting to note plete valence saturation of Pd 4d and O 2p orbitals can be
that upon CO introduction onto all three samples the seen in Figure 4 b, clearly showing the presence of spin density
formation of CO2 is clearly observed. However, while CO 2 (i.e., unpaired charge) in these orbitals, indicative of partially

22772 www.angewandte.org  2021 Wiley-VCH GmbH Angew. Chem. Int. Ed. 2021, 60, 22769 – 22775
Angewandte
Research Articles Chemie

15213773, 2021, 60, Downloaded from https://onlinelibrary-wiley-com.ezproxy.lib.vt.edu By Virginia Tech- on [28/10/2021]. Re-use and distribution is strictly not permitted, except for Open Access articles
surface, nearby O atoms are made dramatically more reactive
(OVFEs of 22–43 kJmol1). The effect is diminished around
a Pd-O distance of  8  (OVFE of 70 kJ mol1) reaching
a plateau (of  88 kJ mol1) at  12 . As a difference in
reducibility between 0.1 and 0.2 Pd/nm2 is seen in our H2-TPR
experiments, and our model only allows for a minimum
surface density of  0.19 Pd/nm2, we suspect a significantly
larger distance is required to completely recover pristine
surface behavior. Using the average of 8  and 12  (i.e.,
10 ) as an approximate “radius of [PdO4] effect”, we overlay
an illustrative “zone of influence” (semi-transparent blue
circles in Figure 4 d) around each [PdO4 ] at different surface
Pd densities wherein O atoms are likely activated. Interest-
ingly, areas begin to overlap as surface Pd density increases,
encompassing neighboring [PdO4 ] by 1.13 Pd/nm2. This
suggests a cumulative effect on oxygen activity as surface
Pd density increases.
To test this presumption, we compute OVFEs for the five
most proximal (out to  8 ) O atoms to Pd (including the O
atoms in [PdO4]) as a function of surface Pd density (Fig-
ure 5 a). The results indicate that oxygen activity is progres-
sively enhanced as surface Pd density increases, with a general
downward trend in OVFEs seen for the O atoms not directly

Figure 4. a) Top-down view of the square planar [PdO4 ] active site on


the CeO 2(100) surface with atoms near Pd labeled to aid discussion
found in the Supporting Information (not to be confused with the
labels used in Figure 5). b) Spin density plot of the [PdO4 ] active site.
c) Oxygen vacancy formation energies (E OVF ) as a function of distance
from Pd. d) Approximate zone(s) of effect (semi-transparent blue
circles, shown with a 10  radius) of each [PdO4], illustrating how
overlapping zones produce a cumulative effect as Pd surface density
increases. Nominally symmetrically distinct O atoms within 8  of the
central Pd atom have been indicated with the colored and labeled
circles of Figure 5 for future reference. Atom key : Pd (brown), Ce
(green), and O (red) atoms on the CeO 2(100) surface ; O adatoms
(magenta).

charged states. This overoxidation occurs because the two


additional O atoms (O1 and O2 in Figure 4 a) require four
total electrons to reach valence saturation, while neither Pd
nor the Ce4+ of the surface can be oxidized sufficiently to
provide all four of these electrons. Instead, Pd provides  2.6
electrons (apparently the limit of its reducing power) and the
remaining electron deficiency is shared equally amongst the 4
Pd-coordinated O atoms (O1–O4 in Figure 4 a) via charge
delocalization.
This overall 2 e deficiency of the PdO4 moiety suggests
that it would be an excellent oxidizer whenever the nearby Figure 5. Quantities computed as a function of Pd loading : a) oxygen
CeO 2 surface is reduced as in CO oxidation via the MvK vacancy formation energies, b) oxygen adsorption energies, and
mechanism. We confirm this by calculating the oxygen c) average CO adsorption energies (inset : side view of the correspond-
vacancy formation energy (OVFE) of O atoms as a function ing ball-and-stick model). d) Ball-and-stick models of the four surface
Pd densities computed here. Symmetrically distinct O atoms are
of their distance from Pd. The results, shown in Figure 4 c, are
labeled 1–5 in each panel of (d). Each series in (a) and (b)
compared against the calculated OVFE of the pristine
corresponds to removal of the labeled O atoms or adsorption of an
CeO 2(100) surface (  102 kJ mol1/O). Intriguingly, while additional O atom to the labeled O atoms to form a superoxo (O 2).
the O atoms in [PdO 4] are more difficult to remove (OVFEs The series in (c) corresponds to the adsorption of CO on all single Pd
of  133 and  143 kJ mol1 /O) than those of the pristine atom centers available in each system in (d).

Angew. Chem. Int. Ed. 2021, 60, 22769 – 22775  2021 Wiley-VCH GmbH www.angewandte.org 22773
Angewandte
Research Articles Chemie

15213773, 2021, 60, Downloaded from https://onlinelibrary-wiley-com.ezproxy.lib.vt.edu By Virginia Tech- on [28/10/2021]. Re-use and distribution is strictly not permitted, except for Open Access articles
coordinated to Pd (labeled O3–O5 in Figure 4). Regardless of Raman spectroscopy, FTIR, and DFT calculations, we
surface Pd density, the Pd-coordinated O atoms (labeled O1 demonstrate that this improvement is due to increased
and O2 in Figure 4) remain more tightly bound than O atoms activity of O atoms proximal to Pd, with a surprisingly large
in pristine CeO 2(100), indicating these O atoms may not be area of influence (at minimum 2.0 nm2 ) and O activity
very active in the CO oxidation reaction. In Figure 5 b,c, accumulation as surface Pd density increases. This unique
respectively, we show that O2 and CO adsorption energies are behavior, characteristic of CeO 2 but absent in Al 2O 3, empha-
minimally affected by surface Pd density, indicating that the sizes the importance of the support.
primary cause of the enhanced CO oxidation activity, as seen XANES, EXAFS, DFT and charge analysis allowed us to
in experiment, is driven by the increased O reactivity. An characterize the Pd/CeO2 active site as an overoxidized
analysis of the charge distribution, seen in Figures S27–S29, square planar [PdO 4 ] complex with Pd in an approximate
reveals that [PdO4] is consistently reduced after O atom + 2.6 oxidation state and unsaturated valence O atoms in
removal, regardless of which O atom is removed. As can be approximately 1.7 oxidation states. The oxidative power of
seen in Figure S30a–c, for the O3 atom most proximal to [PdO4 ] is also shown to have a large lateral range of effects on
[PdO 4], the most likely participant in CO oxidation, the oxygen activity that accumulates as surface density of [PdO4 ]
OVFEs correlate well with the change in work function (DF), increases. This stems from the ability of CeO2 to shuttle
with large DF corresponding to the highest OVFEs. Fig- charge from oxygen vacancies to the [PdO4] complex, which
ure S30d,e shows that geometric relaxations do not correlate reduces it to a more stable oxidation state. Both Raman
with the OVFEs. This confirms that nonlocal electrostatic spectroscopy and DFT calculations reveal that the limits of
effects are the source of the computed OVFE trends. this cumulative effect at high surface Pd densities result in
The OVFE and O2 adsorption energy of O3 at surface Pd overactivation and permanent removal of previously active O
density of 1.13 Pd/nm2 (Figure 5 a,b) deserves highlighting. atoms. Overall, the most interesting feature is direct corre-
We find that the OVFE is thermoneutral to within expected lation of specific activity to CeO2 reducibility rather than Pd.
errors, suggesting that this O atom should be very active. This work invites further investigation into the origin and
However, the subsequent re-adsorption of O2 is lacking an manifestation of nonlocal effects in single-atom catalysts that
enthalpic driving force so once this O vacancy forms, it will have hitherto received little attention.
very seldom reform. This is consistent with experimental
results that showed decreased oxygen species adsorption on
oxygen vacancies on 4 Pd/CeO2 (Figure 3). As can be seen in Acknowledgements
Figure S31, up to 4 such proximal O atoms (all equivalent to
O3 in Figure 5) can be removed, corresponding to two thirds We acknowledge the financial support from the National
of the available [PdO4 ] active sites. This is a significant loss of Research Foundation (NRF) (No. 2016R1A5A1009405,
active O atoms and may explain the experimentally observed 2017R1A2B4007310). Work at the Pacific Northwest Na-
loss in specific activity (Figure 1) in addition to simple Pd tional Laboratory (PNNL) was supported by the U.S. Depart-
agglomeration. This is an important observation because Pd ment of Energy, Office of Science, Basic Energy Sciences,
agglomeration does not necessarily remove the OVFE low- Chemical Sciences, Geosciences, and Biosciences Division.
ering effect, as supported metal nanoparticles have been PNNL is a multiprogram national laboratory operated for
shown to reduce OVFEs at the nanoparticle/support interface DOE by Battelle under Contract DE-AC05-76RL01830.
to a significant degree. [18d, 24] Computational Resources were provided by a user proposal
Lastly, we wish to emphasize that the preceding results at the National Energy Research Scientific Computing
should not be construed to suggest that O atoms more distant Center (NERSC), a U.S. Department of Energy Office of
than the most proximal O3 atoms are directly active in CO Science User Facility located at Lawrence Berkley National
oxidation. This enhancement must be entropic in nature since, Laboratory (LBNL).
as previously stated, it is the preexponential factor and not the
activation energy that is appreciably changed as surface Pd
density increases (Figure 1 c). Since we show that the radius of Conflict of Interest
effect where O reducibility is enhanced is quite large, the
diffusivity of O vacancies is likely to be higher in these The authors declare no conflict of interest.
regions, increasing the probability of O vacancies being filled
and propelling the reaction forward. This is admittedly Keywords: CO oxidation · electrostatic effects · Pd/CeO 2 ·
grounds for a much more ambitious study involving kinetic reducible oxide supports
Monte Carlo to confirm this conjecture.

Conclusion [1] a) X.-F. Yang, A. Wang, B. Qiao, J. Li, J. Liu, T. Zhang, Acc.
Chem. Res. 2013, 46, 1740 – 1748 ; b) S. Mitchell, E. Vorobyeva, J.
Prez-Ramrez, Angew. Chem. Int. Ed. 2018, 57, 15316 – 15329 ;
Our study highlights a synergy between single Pd atoms
Angew. Chem. 2018, 130, 15538 – 15552; c) X. Li, X. Yang, Y.
on CeO2 and enhanced catalytic activity, in contrast to the Huang, T. Zhang, B. Liu, Adv. Mater. 2019, 31, 1902031; d) X.
constant-activity behavior typically assumed for single-atom Cui, W. Li, P. Ryabchuk, K. Junge, M. Beller, Nat. Catal. 2018, 1,
catalysts. By combining activity measurements, H2-TPR, 385 – 397.

22774 www.angewandte.org  2021 Wiley-VCH GmbH Angew. Chem. Int. Ed. 2021, 60, 22769 – 22775
Angewandte
Research Articles Chemie

15213773, 2021, 60, Downloaded from https://onlinelibrary-wiley-com.ezproxy.lib.vt.edu By Virginia Tech- on [28/10/2021]. Re-use and distribution is strictly not permitted, except for Open Access articles
[2] a) J. Wang, Z. Li, Y. Wu, Y. Li, Adv. Mater. 2018, 30, 1801649 ; [14] a) G. Spezzati, Y. Su, J. P. Hofmann, A. D. Benavidez, A. T.
b) J. Wu, L. Xiong, B. Zhao, M. Liu, L. Huang, Small Methods DeLaRiva, J. McCabe, A. K. Datye, E. J. M. Hensen, ACS Catal.
2020, 4, 1900540; c) C. Tang, Y. Jiao, B. Shi, J.-N. Liu, Z. Xie, X. 2017, 7, 6887 – 6891; b) G. Spezzati, A. Benavidez, A. T. DeLaR-
Chen, Q. Zhang, S.-Z. Qiao, Angew. Chem. Int. Ed. 2020, 59, iva, Y. Su, J. P. Hofmann, S. Asahina, E. J. Olivier, J. H.
9171 – 9176 ; Angew. Chem. 2020, 132, 9256 – 9261; d) J.-C. Liu, Neethling, J. T. Miller, A. K. Katye, E. J. M. Hensen, Appl.
H. Xiao, J. Li, J. Am. Chem. Soc. 2020, 142, 3375 – 3383 ; e) J. Li, Catal. B 2019, 243, 36 – 46.
S. Chen, N. Yang, M. Deng, S. Ibraheem, J. Deng, J. Li, L. Li, Z. [15] a) A. I. Boronin, E. M. Slavinskaya, I. G. Danilova, R. V.
Wei, Angew. Chem. Int. Ed. 2019, 58, 7035 – 7039 ; Angew. Chem. Gulyaev, Y. I. Amosov, P. A. Kuznetsov, I. A. Polukhina, S. V.
2019, 131, 7109 – 7113; f) L. Zhao, Y. Zhang, L.-B. Huang, X.-Z. Koscheev, V. I. Zaikovskii, A. S. Noskov, Catal. Today 2009, 144,
Liu, Q.-H. Zhang, C. He, Z.-Y. Wu, L.-J. Zhang, J. Wu, W. Yang, 201 – 211; b) C. Dessal, T. Len, F. Morfin, J.-L. Rousset, M.
L. Gu, J.-S. Hu, L.-J. Wan, Nat. Commun. 2019, 10, 1278; g) X. Aouine, P. Afanasiev, L. Piccolo, ACS Catal. 2019, 9, 5752 – 5759;
Zhang, M. Zhang, Y. Deng, M. Xu, L. Artiglia, W. Wen, R. Gao, c) M. Cargnello, V. V. T. Doan-Nguyen, T. R. Gordon, R. E.
B. Chen, S. Yao, X. Zhang, M. Peng, J. Yan, A. Li, Z. Jiang, X. Diaz, E. A. Stach, R. J. Gorte, P. Fornasiero, C. B. Murray,
Gao, S. Cao, C. Yang, A. J. Kropf, J. Shi, J. Xie, M. Bi, J. A. Science 2013, 341, 771 – 773 ; d) H. Jeong, J. Bae, J. W. Han, H.
van Bokhoven, Y.-W. Li, X. Wen, M. Flytzani-Stephanopoulos, Lee, ACS Catal. 2017, 7, 7097 – 7105 ; e) Z. Hu, X. Liu, D. Meng,
C. Shi, W. Zhou, D. Ma, Nature 2021, 589, 396 – 401. Y. Guo, Y. Guo, G. Lu, ACS Catal. 2016, 6, 2265 – 2279.
[3] a) D. Kunwar, S. Zhou, A. DeLaRiva, E. J. Peterson, H. Xiong, [16] a) H. Jeong, O. Kwon, B. -S. Kim, J. Bae, S. Shin, H.-E. Kim, J.
X. I. Pereira-Hernndez, S. C. Purdy, R. ter Veen, H. H. Bron- Kim, H. Lee, Nat. Catal. 2020, 3, 368 – 375 ; b) X. Liu, P. Ning, L.
gersma, J. T. Miller, H. Hashiguchi, L. Kovarik, S. Lin, H. Guo, Xu, Q. Liu, Z. Song, Q. Zhang, RSC Adv. 2016, 6, 41181 – 41188 ;
Y. Wang, A. K. Datye, ACS Catal. 2019, 9, 3978 – 3990 ; b) B. c) E. M. Slavinskaya, T. Y. Kardash, O. A. Stonkus, R. V.
Qiao, J.-X. Liang, A. Wang, C.-Q. Xu, J. Li, T. Zhang, J. J. Liu, Gulyaev, I. N. Lapin, V. A. Svetlichnyi, A. I. Boronin, Catal.
Nano Res. 2015, 8, 2913 – 2924 ; c) M. Yang, S. Li, Y. Wa ng, J. A. Sci. Technol. 2016, 6, 6650 – 6666.
Herron, Y. Xu, L. F. Allard, S. Lee, J. Huang, M. Mavrikakis, M. [17] A. F. Lee, S. F. J. Hackett, J. S. J. Hargreaves, K. Wilson, Green
Flytzani-Stephanopoulos, Science 2014, 346, 1498 – 1501; d) L. Chem. 2006, 8, 549 – 555.
Lin, W. Zhou, R. Gao, S. Yao, X. Zhang, W. Xu, S. Zheng, Z. [18] a) Z. Wu, M. Li, S. H. Overbury, J. Catal. 2012, 285, 61 – 73; b) J.-
Jiang, Q. Yu, Y.-W. Li, C. Shi, X.-D. Wen, D. Ma, Nature 2017, X. Liu, Y. Su, I. A. W. Filot, E. J. M. Hensen, J. Am. Chem. Soc.
544, 80 – 83. 2018, 140, 4580 – 4587; c) H. Ha, S. Yoon, K. An, H. Y. Kim, ACS
[4] B. Qiao, A. Wang, X. Yang, L. F. Allard, Z. Jiang, Y. Cui, J. Liu, J. Catal. 2018, 8, 11491 – 11501; d) B. Liu, J. Liu, T. Li, Z. Zhao, X.-
Li, T. Zhang, Nat. Chem. 2011, 3, 634 – 641. Q. Gong, Y. Chen, A. Duan, G. Jiang, Y. Wei, J. Phys. Chem. C
[5] T. Yang, R. Fukuda, S. Hosokawa, T. Tanaka, S. Sakaki, M. 2015, 119, 12923 – 12934; e) M. Lohrenscheit, C. Hess, Chem-
Ehara, ChemCatChem 2017, 9, 1222 – 1229. CatChem 2016, 8, 523 – 526.
[6] a) Y.-G. Wang, D. C. Cantu, M.-S. Lee, J. Li, V.-A. Glezakou, R. [19] a) X. Sun, J. Lin, Y. Zhou, L. Li, Y. Su, X. Wang, T. Zhang,
Rousseau, J. Am. Chem. Soc. 2016, 138, 10467 – 10476 ; b) S. F. AIChE J. 2017, 63, 4022 – 4031; b) M.-F. Luo, Z.-Y. Hou, X.-X.
Yuk, G. Collinge, M.-T. Nguyen, M.-S. Lee, V.-A. Glezakou, R. Yuan, X.-M. Zheng, Catal. Lett. 1998, 50, 205 – 209.
Rousseau, J. Chem. Phys. 2020, 152, 154703. [20] a) S. Lin, L. Yang, X. Yang, R. Zhou, Appl. Surf. Sci. 2014, 305,
[7] Y. Lou, J. Liu, Ind. Eng. Chem. Res. 2017, 56, 6916 – 6925. 642 – 649; b) R. Strobel, J.-D. Grunwaldt, A. Camenzind, S. E.
[8] a) D. Mei, J. H. Kwak, J. Hu, S. J. Cho, J. Szanyi, L. F. Allard, Pratsinis, A. Baiker, Catal. Lett. 2005, 104, 9 – 16.
C. H. F. Peden, J. Phys. Chem. Lett. 2010, 1, 2688 – 2691; b) B. B. [21] A. A. Vedyagin, A. M. Volodin, V. O. Stoyanovskii, I. V. Mis-
Sarma, P. N. Plessow, G. Agostini, P. Concepcin, N. Pfnder, L. hakov, D. A. Medvedev, A. S. Noskov, Appl. Catal. B 2011, 103,
Kang, F. R. Wang, F. Studt, G. Prieto, J. Am. Chem. Soc. 2020, 397 – 403.
142, 14890 – 14902; c) S. Abbet, U. Heiz, H. Hkkinen, U. [22] a) Z. Wu, M. Li, J. Howe, H. M. Meyer, S. H. Overbury,
Landman, Phys. Rev. Lett. 2001, 86, 5950. Langmuir 2010, 26, 16595 – 16606 ; b) A. Filtschew, K. Hofmann,
[9] Z. Zhang, Y. Zhu, H. Asakura, B. Zhang, J. Zhang, M. Zhou, Y. C. Hess, J. Phys. Chem. C 2016, 120, 6694 – 6703.
Han, T. Tanaka, A. Wang, T. Zhang, N. Yan, Nat. Commun. 2017, [23] C. Schilling, M. V. Ganduglia-Pirovano, C. Hess, J. Phys. Chem.
8, 16100. Lett. 2018, 9, 6593 – 6598.
[10] E. J. Peterson, A. T. DeLaRiva, S. Lin, R. S. Johnson, H. Guo, [24] a) A. R. Puigdollers, G. Pacchioni, ChemCatChem 2017, 9,
J. T. Miller, J. H. Kwak, C. H. F. Peden, B. Kiefer, L. F. Allard, 1119 – 1127; b) S. W. Hoh, L. Thomas, G. Jones, D. J. Willock,
F. H. Ribeiro, A. K. Datye, Nat. Commun. 2014, 5, 4885. Res. Chem. Intermed. 2015, 41, 9587 – 9601; c) Y. Hinuma, T.
[11] P. Venkataswamy, D. Jampaiah, F. Lin, I. Alxneit, B. M. Reddy, Toyao, N. Hamamoto, M. Takao, K.-i. Shimizu, T. Kamachi, J.
Appl. Surf. Sci. 2015, 349, 299 – 309. Phys. Chem. C 2020, 124, 27621 – 27630 ; d) P. Schlexer, D.
[12] a) B. Qiao, J. Lin, A. Wang, Y. Chen, T. Zhang, J. Liu, Chin. J. Widmann, R. J. Behm, G. Pacchioni, ACS Catal. 2018, 8, 6513 –
Catal. 2015, 36, 1505 – 1511; b) L. Nie, D. Mei, H. Xiong, B. Peng, 6525.
Z. Ren, X. I. P. Hernandez, A. DeLaRiva, M. Wa ng, M. H.
Engelhard, L. Kovarik, A. K. Datye, Y. Wang, Science 2017, 358,
1419 – 1423 ; c) X. Zhang, J. Lei, D. Wu, X. Zhao, Y. Jing, Z.
Zhou, J. Mater. Chem. A 2016, 4, 4871 – 4876. Manuscript received: April 28, 2021
[13] A. J. Therrien, A. J. R. Hensley, M. D. Marcinkowski, R. Zhang, Revised manuscript received: June 11, 2021
F. R. Lucci, B. Coughlin, A. C. Schilling, J.-S. McEwen, E. C. H. Accepted manuscript online: June 28, 2021
Sykes, Nat. Catal. 2018, 1, 192 – 198. Version of record online: July 19, 2021

Angew. Chem. Int. Ed. 2021, 60, 22769 – 22775  2021 Wiley-VCH GmbH www.angewandte.org 22775

You might also like