You are on page 1of 8

Articles

https://doi.org/10.1038/s41929-019-0328-1

Catalyst deactivation via decomposition into


single atoms and the role of metal loading
Emmett D. Goodman1, Aaron C. Johnston-Peck2, Elisabeth M. Dietze   3, Cody J. Wrasman1,
Adam S. Hoffman   4, Frank Abild-Pedersen1,4, Simon R. Bare   4, Philipp N. Plessow   3 and
Matteo Cargnello   1*

In the high-temperature environments needed to perform catalytic processes, supported precious metal catalysts lose their
activity severely over time. Generally, loss of catalytic activity is attributed to nanoparticle sintering or processes by which
larger particles grow at the expense of smaller ones. Here, by independently controlling particle size and particle loading
using colloidal nanocrystals, we reveal the opposite process as an alternative deactivation mechanism: nanoparticles rapidly
lose activity for methane oxidation by high-temperature decomposition into inactive single atoms. This deactivation route is
remarkably fast, leading to severe loss of activity in as little as 10 min. Importantly, this deactivation pathway is strongly depen-
dent on particle density and the concentration of support defect sites. A quantitative statistical model explains how, for certain
reactions, higher particle densities can lead to more stable catalysts.

I
ncreased catalyst stability, especially in automotive emissions Results
control applications, is crucial to decrease the loading of rare Independent control of particle size and loading. In this work,
and precious noble metals1,2. Unfortunately, due to declining we study Pd catalysts because of their relevance in emission control
catalytic activity during operation, emissions control catalysts are catalysis, but we anticipate this observation to be of general valid-
loaded with up to 10 g of precious metals to ensure effective cataly- ity and interest in many supported systems. To vary the nanopar-
sis throughout the lifetime of the material3,4. Two mechanisms are ticle density without affecting particle size, different amounts of
universally proposed for the loss of reactive surface area: particle the same 7.9 ± 0.6 nm pre-formed colloidal Pd nanoparticles were
migration and coalescence and atomic (or Ostwald) ripening. The added to the same mass of stabilized gamma-alumina (γ-Al2O3)
former involves the movement and coalescence of entire nanopar- support to obtain catalysts with Pd loadings of 0.659, 0.067 and
ticles, and the latter involves the motion of atomic species from 0.007 wt%. These loadings correspond to dense, intermediate
smaller to larger particles. In both mechanisms, researchers observe and sparse nanoparticle densities on the support (Fig. 1a–c and
an emergence of larger particle aggregates, a clear indication of par- Supplementary Fig. 1). It should be noted that, before nanopar-
ticle sintering, which is correlated with the corresponding loss of ticle impregnation, the γ-Al2O3 support was calcined at 900 °C
activity. Given that it is difficult to distinguish between sintering for 24 h to ensure that it would be stable throughout our aging
processes, researchers aim to create nanostructures that maximize experiments, which were performed at lower temperatures for
distances between catalytic nanoparticles in hope of minimizing much shorter durations. Organic ligands were removed from
particle growth5–7. Although it is very challenging to isolate a spe- the particle surfaces via rapid heating treatment11, leaving the
cific deactivation mechanism, a clear understanding of specific cat- original nanoparticle size uniformity unchanged (Supplementary
alyst degradation mechanisms is important for the rational design Fig. 2). High-angle annular dark-field scanning transmission elec-
of stable heterogeneous catalysts8–10. tron microscopy (HAADF-STEM) characterization demonstrated
In this work, instead of catalyst deactivation due to particle the random distribution of Pd nanoparticles on the alumina
growth, we identify and explore an alternative deactivation mecha- support and the control over particle loading and density. An
nism characterized by nanoparticle decomposition into inert single obvious change in nanoparticle density is observed as the overall
atoms. Importantly, the extent and severity of this deactivation Pd mass loading on alumina is varied, leading to average nanopar-
mechanism is a direct function of nanoparticle size and spatial ticle densities of 22, 2.2 and 0.23 NPs per square micrometre for
arrangement. However, using traditional catalyst synthesis tech- dense, intermediate and sparse samples, respectively. These values
niques, the spatial arrangement of particles and its influence on cat- correspond to average interparticle distances of 1,100, 340 and
alyst properties are difficult to control and study, as catalyst metallic 110 nm for the three samples, respectively, according to Monte
loading and particle size are two strongly connected parameters. Carlo simulations paired with Voronoi analyses (Supplementary
Here, utilizing colloidal nanocrystals to independently control par- Table 1 and Supplementary Fig. 3). Based on characterization
ticle size and particle loading—two parameters crucial to catalyst and simulation, we conclude that these materials represent a set
stability—we demonstrate an unexpected result: higher particle of powder catalysts with identical nanoparticle size yet different
loadings can result in a more stable catalyst. nanoparticle spatial arrangements.

Department of Chemical Engineering and SUNCAT Center for Interface Science and Catalysis, Stanford University, Stanford, CA, USA. 2Material
1

Measurement Laboratory, National Institute of Standards and Technology, Gaithersburg, MD, USA. 3Institute of Catalysis Research and Technology,
Karlsruhe Institute of Technology, Eggenstein-Leopoldshafen, Germany. 4Stanford Synchrotron Radiation Lightsource, SLAC National Accelerator
Laboratory, Menlo Park, CA, USA. *e-mail: mcargnello@stanford.edu

Nature Catalysis | www.nature.com/natcatal


Articles Nature Catalysis

a b c

50 nm 22 NP per µm2 50 nm 2.2 NP per µm2 50 nm 0.23 NP per µm2

d e
100 800 100

700
Dense
75 600 75

Temperature (°C)
CH4 conversion (%)

CH4 conversion (%)


500
Aging

Fresh

Aged
50 (no CH4) Intermediate 400 50

300

25 200 25

Sparse 100

0 0
0 50 100 150 200 250 Dense Intermediate Sparse
Time (min)

Fig. 1 | Effect of different Pd nanoparticle densities on catalytic stability. a–c, Representative HAADF-STEM images of dense (0.659 wt%) (a),
intermediate (0.067 wt%) (b) and sparse (0.007 wt%) (c) Pd/Al2O3 samples. d, CH4 conversion profiles for Pd/Al2O3 catalysts with different nanoparticle
loadings following the temperature profile (black line and right axis). e, Averaged CH4 conversion values at 460 °C for the Pd/Al2O3 catalysts before
(‘Fresh’) and after (‘Aged’) aging. Error bars represent the minimum and maximum results of at least three repeat experiments.

We set out to compare the thermal stability of these systems Density-dependent decomposition of nanoparticles into single
using methane combustion as a probe reaction, because this reac- atoms. The surprising catalytic stability of the dense Pd/Al2O3
tion requires stable materials that can withstand high temperatures material was confirmed via HAADF-STEM analysis (Fig. 2a and
for emissions control applications4,12. Additionally, methane com- Supplementary Fig. 4). Comparing particle size distributions before
bustion is useful because combustion rates are proportional to the and after aging in oxygen demonstrates a lack of particle agglomera-
exposed Pd surface area, so combustion activity is a good proxy for tion (Fig. 2a,d), although the peak in particle size shifted to 9.3 nm
exposed reactive surface area13,14. The protocol to measure catalyst due to oxidation of Pd to PdO, and a population of smaller par-
deactivation included a measurement of catalytic activity at 460 °C ticles was formed. In the case of the intermediate and sparse density
(a temperature where Pd is unaffected by water poisoning15), in situ catalysts, HAADF-STEM reveals the presence of Pd species in the
aging in dilute oxygen at 775 °C for 1 h and a subsequent measure- form of Pd clusters distinctly smaller than the original 7.9 nm par-
ment of catalytic activity at 460 °C in the same reaction mixture as ticles (Fig. 2b,c,e,f). However, the intermediate material still shows
before (Fig. 1d). Stable materials would show the same conversion a fraction (10%) of larger nanoparticles, particularly in areas of high
before and after the aging treatment, but unstable materials would nanoparticle density (Supplementary Fig. 4), whereas only one sin-
show decreased conversion after the aging treatment. Although this gle nanoparticle larger than 4 nm was observed in the sparse sam-
process is an approximation of realistic conditions, it allows for isola- ple. A similar density-dependent stability was found when aging
tion of the environmental conditions responsible for catalyst deacti- the materials under methane combustion conditions instead of just
vation. The masses of each catalyst and Al2O3 diluent were chosen to oxygen. In this case, direct observation of Pd single atoms was pos-
load the same amount of Pd into the reactor while keeping the total sible (Supplementary Fig. 5).
bed volume constant. Before the aging treatment, each test showed Extended X-ray absorption fine structure (EXAFS) analysis
similar methane conversion (~85%), indicating no significant confirms the HAADF-STEM findings: for the sparse catalyst, the
low-temperature interaction between Pd nanoparticles (Fig. 1d,e). Pd–Pd coordination contribution disappears after aging, and the
Conventional wisdom would predict that the catalyst with lowest average first-shell Pd–Pd coordination number approaches zero,
Pd particle density should be the most stable, due to a lower prob- significantly less than that for bulk Pd or PdO (12 or 4, respectively)
ability of particle migration and coalescence or interparticle atomic (Fig. 2h and Supplementary Table 2). X-ray photoelectron spectros-
exchange. Surprisingly, after high-temperature aging, the dense Pd/ copy (XPS) confirms that although dense and sparse Pd catalysts
Al2O3 catalyst showed completely stable activity, while the sparse both start off in the same metallic Pd state due to the colloidal syn-
Pd/Al2O3 catalyst showed dramatic deactivation, with conversion thesis forming metallic Pd nanoparticles, the sparse catalyst evolves
decreasing from 85 to 20%. Interestingly, the intermediate sample towards a highly disperse and highly oxidized state, characterized by
showed an intermediate loss of activity to ~55% conversion after a binding energy ~1 eV higher than bulk PdO16 (Fig. 2g). Inductively
the aging treatment. Counterintuitively, although the most isolated coupled plasma mass spectrometry (ICP-MS) quantitatively shows
Pd nanoparticles lost nearly all combustion activity, the Pd/Al2O3 that all Pd is conserved after the aging treatment and that none is
sample with the highest particle density remained stable. lost to the gas phase (Supplementary Table 3). The fact that Pd could

Nature Catalysis | www.nature.com/natcatal


Nature Catalysis Articles
Binding energy (eV)
350 346 342 338 334 330
80
a d Before
g
After 60
N = 389 Dense fresh

40

Intensity (a.u.)
PdO Dense aged
20
Sparse fresh
10 nm
0

Percentage of nanoparticles (%)


b e Sparse
Before aged
After Pd(II)
60
N = 367 Pd(0)

40

h
20
Pd foil
10 nm
0
Sparse fresh

Intensity (a.u.)
c f Before
After
60
N = 34 Sparse catalysis

40 Sparse aged

20 Dense aged

10 nm PdO powder
0
0 4 8 12 16 20 0 1 2 3 4 5
Nanoparticle R (Å)
diameter (nm)

Fig. 2 | Particle density-dependent conversion of Pd nanoparticles into Pd single atoms. a–c, Representative HAADF-STEM images of dense (a), intermediate
(b) and sparse (c) Pd/Al2O3 samples after aging in 4 vol% O2/Ar for 1 h at 775 °C. d–f, Corresponding particle size distributions for the samples before and
after aging for dense (d), intermediate (e) and sparse (f) catalysts; N indicates the number of measurements made post-aging. g, XPS of dense and sparse
Pd/Al2O3 before (fresh) and after (aged) the aging treatment. h, EXAFS of sparse and dense Pd/Al2O3 before (fresh), after catalysis at 460 °C (catalysis)
and after the aging treatment (aged). Dark traces are fits; thicker light traces are experimental data; solid traces are real components; dotted traces are
imaginary components.

be dissolved for ICP analysis while the bulk of Al2O3 remains intact (refs. 18,19). Stabilization of atomically dispersed Pd has been identi-
further suggests that Pd is still located on the Al2O3 surface and has fied before, but on CexZr1 − xO2-Al2O3 and La-Al2O3 supports, which
not entered the support lattice. Taken together, the data suggest that, are purported to stabilize atomic species20,21. Here, we report that, at
at low Pd density, nanoparticles decompose into single atomic spe- low nanoparticle densities, complete decomposition of 7.9 nm par-
cies, while at denser particle loadings, catalysts maintain their size ticles (consisting of ~18,000 atoms) into stable single-atom species
and particle dispersion. on unmodified commercial Al2O3 takes place. Although the same
If deactivation was due to either particle migration/coalescence previous reports demonstrate Pd atomic species as good CO oxida-
or classic atomic ripening processes, as traditionally posited in tion catalysts, we here show that these species are very poorly active
high-temperature applications, an increase in particle size would be for methane combustion and are a significant cause of catalyst deac-
expected17. The emergence of smaller particles suggests that atomic tivation21. Thus, our data help us to gain a fundamental understand-
processes are still active, but instead of Pd atoms moving from ing of a previously unrecognized deactivation mechanism and the
smaller to larger particles, Pd atoms are stabilized by the support, activity of single-atom species for methane combustion.
leading to atomically dispersed species. Importantly, the single-
atoms that form are much less active for the methane combustion Emission-limited decomposition kinetics. To further understand
reaction, probably due to the need for a critical ensemble size to trends in the observed density-dependent particle decomposition,
adsorb oxygen and fully dehydrogenate methane. Comparing the we analysed the behaviour of catalysts with smaller (2.5 ± 0.4 nm)
reaction rate of the sparse sample with single-atomic species formed and larger (14.7 ± 1.5 nm) diameter Pd nanoparticles, synthesized
after the aging treatment to that of the dense Pd/Al2O3 catalyst com- using the same colloidal strategies to independently tune particle
posed of Pd particles of 9 nm in size from previous work13, the dif- size and density (Fig. 3a–c and Supplementary Fig. 2). For each size,
ference is very large, with single-atomic species showing a methane reaction conditions were chosen to maintain similar conversion
combustion rate that is more than two orders of magnitude lower (~85%) before aging. The same density-dependent stability phe-
than that of Pd particles. These data shed new light onto the poor nomenon, with the trend of denser catalysts being more stable, was
activity of single-atom catalysts for hydrocarbon combustion, an observed for all sizes after an aging treatment performed in diluted
important result in the fiercely debated area of single-site catalysis. oxygen at 775 °C (Supplementary Fig. 6). Although, at higher den-
Although single atoms are rarely described on pure γ-Al2O3 (that is, sity, an increase in particle size was observed in the 2.5 nm sam-
compared to reducible oxides such as CeO2 or Fe2O3), recent works ples, there was significantly worse deactivation for the sample with
have identified the stability of atomically dispersed Pt on Al2O3 lowest particle density, for which EXAFS analysis revealed highly

Nature Catalysis | www.nature.com/natcatal


Articles Nature Catalysis

100 a b
a d

CH4 conversion at 460 °C (%)


100
50 nm 75
Ar O2

O2 signal (a.u.)
50 80
100

CH4 + O2

O2 + H2O
25

Ar
50 60

H2O

O2
0
Percentage of nanoparticles (%)

100

CH4 conversion at 460 °C (%)


b e 40
50 nm 75 Fresh Aged 500 600 700 800 900
Temperature (°C)
50
100 Fig. 4 | Conditions of catalyst decomposition. a, Effect of aging in different
50
25 flowing environments (pure Ar; 4.2 vol% H2O/Ar; 4 vol% O2/Ar; 0.5 vol%
CH4 + 4.2 vol% O2 in Ar; and 4 vol% O2 + 4.2 vol% H2O in Ar) at 775 °C
0
100 on the stability of intermediate-density 7.9 nm Pd/Al2O3 catalyst.
c f b, Temperature-programmed desorption of oxygen on dense 7.9 nm
50 nm 75 Pd/Al2O3 catalyst in pure Ar environment (green) or 4% O2/Ar (purple).

50
100
Ar, or Ar saturated with 10 vol% H2O, the catalyst actually showed
minor improvements in activity, probably due to reconstruction
25
50 of the Pd facets at high temperature. However, when aged in diluted
O2, the activity drop was substantial, and further deactivation
0 0
0 4 8 12 16 20 500 600 700 800 was observed when aging in methane combustion conditions, or
Nanoparticle diameter (nm) Aging temperature (°C) in a combination of O2 and 4.2 vol% H2O. As γ-Al2O3 alone releases
significant water (adsorbed during exposure to air) via dehydra-
Fig. 3 | Interface-limited atomic emission demonstrated by nanoparticle tion and dehydroxylation processes when heated (Supplementary
size control. a–c, HAADF-STEM images and particle size distributions Fig. 9), directly co-feeding H2O is not necessary for catalyst deac-
(N = 200) of intermediate-density Pd/Al2O3 catalysts with 2.5 nm (a), tivation24. Furthermore, temperature-programmed desorption of
7.9 nm (b) and 14.7 nm (c) Pd nanoparticles. d–f, Temperature-dependent oxygen experiments in either Ar or O2 demonstrate that it is prob-
stability of sparse 2.5 nm (d), 7.9 nm (e) and 14.7 nm (f) Pd/Al2O3 catalysts. ably the PdO phase that is present during the activity loss accompa-
The x axis indicates the aging temperature before returning to catalysis nying nanoparticle decomposition (Fig. 4b). Past work has indeed
conditions at 460 °C. suggested the PdO phase to be highly sensitive to the binding of
H2O25. These experiments demonstrate that O2 and H2O are both
required for particle decomposition and the catalyst deactivation
dispersed Pd species (Supplementary Fig. 7). This result highlights phenomenon, which are conditions commonly found in many cata-
that the formation of atomically dispersed Pd species can be a more lytic environments.
detrimental deactivation mechanism to methane combustion cata- Given that size-dependence studies identify atomic Pd emission,
lytic activity than the loss of surface area via particle aggregation. In and not Pd surface diffusion, as the rate-determining step of the
the case of the largest 14.7 nm Pd particles, the trend of deactivation nanoparticle decomposition process, the chemical nature of the
in samples with lower particle density is still observed, but is much involved Pd atomic adsorbates was further investigated. To under-
less severe than in the case of 2.5 and 7.9 nm Pd particles. stand which Pd species are responsible for this unexpected deactiva-
Interestingly, when different nanoparticle sizes at the same tion, density functional theory (DFT) calculations were performed
Pd loading were sequentially aged with increasing temperature, considering Pd, PdO and Pd(OH)2 as adsorbates on the (100) and
we found that the minimum temperature required to observe (110) γ-Al2O3 surfaces, including experimentally relevant temper-
nanoparticle decomposition varied with size: smaller particles ature-dependent H2O coverages (Supplementary Fig. 10). In these
required lower temperatures (<600 °C) and larger particles required calculations, only molecular species with one Pd atom were con-
much higher temperatures (>700 °C) (Fig. 3d–f). This size depen- sidered due to the direct observation of Pd single atoms and lack
dence strongly suggests that the atomic ripening process is limited of Pd–Pd coordination in EXAFS analysis of post-aging samples
by atomic emission rather than surface diffusion of atomic species (Supplementary Table 2). Free energies of adsorption of various
to a nearby site, the latter process being independent of the initial metal adatom complexes (∆Gads) were calculated and referenced to
nanoparticle size. Previous work has also predicted that noble met- 7.9 nm PdO particles, as temperature-programmed desorption of
als will operate in this regime22. The increased stability of larger oxygen experiments revealed that PdO is probably the phase present
particles, caused by lower rates of adatom emission, can be due during particle decomposition (Fig. 4b). On both γ-Al2O3 surfaces,
to a lower surface energy, which decreases adsorbate formation single Pd atoms have the least favourable adsorption energy (>1.5 eV
(Supplementary Fig. 8)23. This analysis presents experimental proof uphill) and therefore probably do not constitute the most stable
that controlling the rate of atomic emission and adsorption to the adsorbed species—in line with the fact that aging in Ar, which con-
support is key to limiting atomic redispersion or growth processes verts PdO to Pd above 500 °C, does not cause deactivation (Fig. 4a).
in many supported catalysts. Instead, DFT reveals the existence of a lowest-energy Pd(OH)2
adsorbate atop tri-coordinated Al atoms (AlIII sites, Fig. 5a,b), with
Loading-dependent nanoparticle decomposition. Different a free energy of adsorbate formation ∆Gads ≈ 0.5 eV uphill. Note that
aging treatments at 775 °C were performed on the intermediate for temperatures up to 700 °C, Pd(OH)2 formation free energies are
density 7.9 nm Pd/Al2O3 sample to understand the conditions that independent of temperature, as at these temperatures a hydrated
trigger nanoparticle decomposition (Fig. 4a). When aged in pure Al2O3 surface is most stable and the formation of adsorbed Pd(OH)2

Nature Catalysis | www.nature.com/natcatal


Nature Catalysis Articles
3.0
a Sparse nanoparticles b

∆Gads(relative to PdO, eV)


Nanoparticle disintegration 2.5
Pd
2.0

PdO 1.5

1.0
Pd(OH)2
0.5
Minimum energy
0
100
c

Percentage of Pd as Pd(OH)2
e
ars 80
Sp

60
AlIII
40
te
rm edia
Inte 20
Dense nanoparticles
Dense
Stable atomic exchange
0
400 500 600 700 800
Temperature (°C)

Fig. 5 | Statistical mechanics model of density-dependent particle decomposition. a, Schematic illustrating the density dependence of atomic
redispersion. Open blue circles, empty AlIII defect sites; filled blue circles, AlIII defect sites populated by Pd atomic species. The inset shows the
most stable adsorbate structure showing Pd(OH)2 binding to an AlIII site. b, Free energies of adsorbate formation of various atomic Pd species on the
γ-Al2O3(110) surface, calculated by DFT. Free energies are referenced to adsorption on a 7.9 nm PdO nanoparticle, and calculated using partial
pressures of pO2  = 4,000 Pa and pH2 O = 10,400 Pa. c, Percentage of Pd in AlIII defects as a function of temperature and nanoparticle density, according
I using the temperature-dependent
to equation (1), I energies in b. Band widths correspond to the model sensitivity towards various alumina defect site
densities (0.75 to 1.5 nm−2).

from PdO and adsorbed H2O would occur without a change in compared to many cases of metal single sites trapped in reducible
the number of gas-phase molecules. Above 700 °C instead, a dehy- oxide supports, where favourable formation energy calculations
drated Al2O3 surface is most stable, at which point the adsorption suggest strong-binding sites29. Importantly, consistent with this
free energy becomes temperature dependent because H2O needs thermodynamic model, when the aged sparse sample is subjected
to adsorb from the gas phase. These AlIII sites are common to the to long times and moderate temperatures (~10 h, 460 °C), atoms
Al2O3(110) facet, which constitutes ~80% of the support surface, recondense into small nanoclusters, and activity slowly begins to
and are the same sites identified by Sautet and co-workers as being return (Supplementary Fig. 11).
reactive for CH4 and H2 splitting and N2 adsorption26. The insights gained from this work have important implications
Given an understanding of the defect site and atomic emission for realistic powder catalysts synthesized by traditional impregna-
energetics, we apply statistical mechanics to explain the observed tion approaches. To demonstrate that the severity of nanoparticle
density-dependent stability phenomenon. At an aging temperature decomposition is not overstated in the use of uniform catalysts, we
of 775 °C, we expect nearly all AlIII sites (~1.5 sites per nm2) to be synthesized artificially non-uniform materials, prepared by mix-
free of H2O and accessible for Pd atoms, although we still consider ing nanoparticles of different sizes (Supplementary Fig. 12). These
the possibility of site densities between 0.75 and 1.5 sites per nm2. materials exhibit the same strong density-dependent stability, with
Given an AlIII site density and the energy of adsorbate formation the highest loaded material maintaining near complete activity,
∆Gads (eV), the percentage of Pd atoms occupying AlIII defect sites and the sparsest catalyst losing >75% of its methane conversion.
is calculated according to a Boltzmann distribution (equation (1)): Experiments to compare particle redispersion to particle sinter-
ing were also carried out. We thus targeted conditions in which we
NAlIII �ΔGads could observe nanoparticle growth by sintering in the dense cata-
NPd e
kT
XPdðOHÞ2 ¼ N ΔGads ´ 100 ð1Þ lyst; eventually, after 1,000 min (17 h), the dense Pd/Al2O3 catalyst
AlIII
NPd e�kT þ1 indeed forms larger aggregates (Supplementary Fig. 13) and loses
~50% of its catalytic activity. Comparatively, a ~50% loss of activity
This equation is plotted in Fig. 5c. Similar equations were occurs in the sparse catalyst after only 10 min, but due to nanopar-
recently derived using particle mass balances to find equilibrium ticle decomposition into single atoms rather than particle growth.
concentrations of atomic adsorbates in traditional ripening pro- In other words, formation of single atoms effects a comparable loss
cesses27,28. Importantly, this analysis holds, independent of the of activity as nanoparticle growth, but 100 times faster. The fact that
specific nature of the defect sites—the Boltzmann distribution the sample with a dense Pd nanoparticle distribution shows particle
always dictates larger equilibrium populations of single-atom spe- growth after long-term aging is in line with atomic emission pro-
cies formed at higher AlIII defect site (NAl_III) to total Pd atom (NPd) cesses occurring for all samples, but with different consequences set
ratios. Despite the single-atomic Pd adsorbate sites being uphill in by the dynamics involving equilibria between Pd atoms in alumina
energy, samples with a lower particle density populate a larger frac- defects and Pd atoms in nanoparticles (Supplementary Fig. 14).
tion of Pd in the defect sites at increasing temperature because of Across nanoparticle sizes, metals and supports, colloidally syn-
the larger NAl_III/NPd ratio in these materials (Fig. 5c). This system is thesized catalysts with controllable nanoparticle densities represent
a case where the formation of single sites is driven by entropy due to a general tool to reveal sintering and deactivation mechanisms in
a large quantity of defect sites, rather than by enthalpy, which would powder catalysts. Here, we report the dramatic effects of a possi-
be the case if adsorption energies were favourable. This is unique bly overlooked deactivation mechanism that occurs at low particle

Nature Catalysis | www.nature.com/natcatal


Articles Nature Catalysis

loadings, and demonstrate that the rapid formation of atomically dis- X-ray absorption spectroscopy. X-ray absorption spectra were collected at the
persed catalysts may cause severe deactivation for certain reactions. Stanford Synchrotron Radiation Lightsource at wiggler beamline 9–3 using a
Si(220) double-crystal monochromator. The storage ring was operated at 3 GeV
Additionally, increasing nanoparticle loading was demonstrated to with a ring current of 494–500 mA in top-up mode. Data were collected at the
reduce the deactivating effects of nanoparticle decomposition and Pd K-edge (24,350 eV) in fluorescence mode using 32- and 100-element liquid-
trapping as atomic species. The demonstrated atomic emission fur- nitrogen-cooled Ge detectors (Canberra). Catalysts were investigated as self-
ther suggests that eventual growth in particle size is due to atomic supporting pellets with an absorption length that prevented self-absorption at the
Pd edge. A Pd reference foil was scanned simultaneously for energy calibration.
ripening, rather than particle migration processes.
The raw data were energy-calibrated, merged and normalized using the Athena
interface of the Demeter software package. EXAFS results were extracted in
Methods k-space and the k3-weighted EXAFS function was used for Fourier transformation.
Synthesis of colloidal nanoparticles. All syntheses were performed with standard Phase shifts and amplitudes for relevant back-scattering paths were calculated
air-free Schlenk techniques using previously published procedures13,30. Pd(acac)2 using FEFF6. EXAFS modelling was carried out taking into account k1-, k2- and
(35% Pd, Acros Organics) was used as the metal precursor; 1-octadecene (ODE, k3-weighting using the Artemis interface of Dimeter software. S02 was calculated
90%, Acros Organics), 1-tetradecene (TDE, 94%, Alfa Aesar) and 1-dodecene, to be 0.81 ± 0.04 by fitting Pd foil. The EXAFS data from all catalysts of a given Pd
(DDE, 93–95%, Acros Organics) were used as solvents; 1-oleylamine (OLAM starting size were fit simultaneously to minimize errors.
70%, Aldrich), oleic acid (OLAC 90%, Aldrich) and trioctylphosphine (TOP,
97%, Aldrich) were used as surfactants and reducing agents. All chemicals were XPS. XPS spectra were measured using a PHI VersaProbe 3 Scanning XPS
used as received. The general synthetic methodology is as follows. Pd(acac)2 was Microprobe equipped with a hemispherical electron analyser using Al(Kα)
added to a three-neck flask together with solvents, oleylamine and oleic acid at radiation (1,486.3 eV). All samples were deposited onto conductive carbon tabs
room temperature. The mixture was evacuated for 15 min (<270 Pa), TOP was on top of an aluminium holder, and were outgassed at 10–2 Pa and transferred
added, and the mixture was degassed for another 30 min at 50 °C. At this point, to the ion-pumped analysis chamber. The pressure was kept below ~5 × 10−7 Pa
all solutions were a pale translucent light yellow. The reaction was flushed with throughout data acquisition. For most samples, the incident X-ray spot size was
nitrogen and quickly heated (~40 °C min−1) to the desired reaction temperature. 100 um, and an excitation of 100 W at 20 kV was applied. For all samples, a pass
After 15 min of reaction at the appropriate reflux temperature under vigorous energy of 280 eV was used. An Ar+ neutralizer and electron flood gun were used to
magnetic stirring, the solution was allowed to cool to room temperature. compensate for Al2O3 charging, and binding energies were referenced to the C 1s
The exact amounts of reagents used to prepare the different sizes are given in peak (284.8 eV) to account for small charging effects.
the following. For 3.3 nm Pd particles: 157 mg of Pd(acac)2, 11 ml of DDE, 9 ml of
TDE, 1.64 ml of OLAM, 570 µl of TOP; degas at 50 °C for 30 min; reaction at 230 °C Catalytic characterization. Due to the high temperatures required for this
for 15 min under nitrogen. For 8.8 nm Pd particles: 157 mg of Pd(acac)2, 6.8 ml reaction and aging, Al2O3 was calcined at 900 °C for 24 h before Pd nanoparticle
of TDE, 13.2 ml of ODE, 3.4 ml of OLAM, 0.8 ml of OLAC, 2.4 ml of TOP; degas impregnation to minimize support changes throughout the reaction and aging.
at 50 °C for 30 min; reaction at 290 °C for 15 min under nitrogen. For 16.2 nm Pd In general, each catalyst, post-ligand removal, was sieved below 180 μm grain size
particles: 77 mg of Pd(acac)2, 5 ml of ODE, 5 ml of OLAC, 0.56 ml of TOP; degas at and mixed with a certain amount of Al2O3 diluent (also calcined at 900 °C for 24 h);
60 °C for 30 min; reaction at 280 °C for 5 min under nitrogen. this was found to be sufficient to eliminate thermal effects from repeated tests. A
The particles were isolated by precipitation with the addition of isopropanol, 200 mg sample of this diluted mixture was loaded into the reactor to give a bed
ethanol and methanol, then centrifugation (838 rad s−1, 3 min); they were then length of ~1.0 cm. When comparing stability across catalyst weight loadings for
redissolved in hexanes. This procedure was repeated three times, after which the a given nanoparticle size, testing was performed under conditions such that each
particles were dissolved in ~5 ml hexanes, with the addition of ~20 μl of OLAM to reactor bed had the same mass of Pd. For example, 200 mg of 0.007 wt% Pd/Al2O3
ensure colloidal stability. For further size-selective purification of 16.2 nm particles, was compared to a bed of 20 mg of 0.067 wt% Pd/Al2O3 mixed with 180 mg Al2O3
1 ml aliquots of isopropanol were added to the black colloidal solution, followed diluent. This bed rested between two layers of granular quartz, which were used to
by centrifugation at 838 rad s−1, until the majority of particles precipitated, leaving prevent displacement of the catalyst powder and for preheating the reactant gases.
a slightly grey supernatant. The supernatant was discarded, the precipitate was The reactor was heated in a square furnace (Micromeritics) and the temperature of
redispersed in hexanes and the procedure was repeated two more times. Finally, the catalyst was measured with a K-type thermocouple inserted inside the reactor,
the larger Pd nanoparticles were dispersed in ~5 ml hexanes with ~20 μl of OLAM touching the catalytic bed. All experiments were conducted at a total pressure of
to ensure colloidal stability. 1 atm. Aging experiments consisted of an oxidative pretreatment (45 ml min−1 of
5 vol% O2/Ar) at 275 °C to remove residual carbon compounds from synthesis and
Synthesis of supported catalysts. Metal concentrations of synthesized colloidal to activate the catalyst. Next, the catalyst was ramped in 0.5 vol% CH4, 4.0 vol%
nanoparticle solutions were determined by thermogravimetric analysis (TGA). O2, bal Ar to 460 °C at 20 °C min−1 and held until activity was stable. The methane
A nanoparticle solution was added dropwise into an aluminium TGA pan, was removed from the reaction mixture, and the catalyst was ramped to 775 °C at
which was heated on a hot plate at ~80 °C until 150 μl had been added. The 20 °C min−1, held for 1 h, and ramped back to 460 °C at 20 °C min−1, at which point
pan was then heated in the TGA system in flowing air up to 500 °C, and held methane was reintroduced into the reaction mixture. Stability was assessed by
until a steady mass was reached, indicating complete removal of solvents and comparing the conversion before and after the aging procedure.
organics. Dividing this final mass by the volume of solution added gave the metal
concentration. Using serial dilution, an appropriate amount of nanoparticles Analysis of particle density. Voronoi analyses were performed using the Voronoi
was dispersed in hexanes and added to a dispersion of ~2 g support stirred in function in Matlab software. Here, we studied supported uniform 2.5, 7.9 and
hexanes. For a given nanoparticle size, the same nanoparticle batch was used 14.7 nm Pd nanoparticles on γ-Al2O3 with a surface area of 96 m2 g−1. We simulated
for all catalysts to ensure targeted metallic loadings and identical nanoparticle size Al2O3 grains with nanoparticles impregnated onto the grain surface. Nanoparticles
and morphology across different weight loadings. Complete adsorption occurred (one-dimensional (1D) points) were randomly generated on the surface of a
immediately, but the solutions were left stirring for 1 h. The solid was recovered by grain as 2D uniformly distributed random variables. In each simulated Pd/Al2O3
centrifugation (838 rad s−1, 1 min) and dried at 60 °C overnight. Before composite, input parameters included the number of points (nanoparticles) and the
catalytic tests, all samples were sieved below 180 μm grain size, fast-treated surface area of the Al2O3 (grain surface area). For a system with n nanoparticles,
at 700 °C for 30 s in a furnace to remove ligands from synthesis as previously an n × n matrix of all interparticle distances was calculated. For each nanoparticle,
described, then sieved again below 180 μm grain size11. Particle size distribution the minimum distance to a nearest-neighbour nanoparticle was selected (the
analysis after rapid thermal treatment demonstrated that the particle size nearest-neighbour distance, NND). By averaging NNDs across the entire grain, we
distributions were maintained compared to the original colloidal nanoparticles calculated the average expected NND. For each nanoparticle density, we simulated
but that nanoparticle sizes were slightly smaller after deposition, probably due a large enough grain such that the nanoparticle densities and NNDs converged. By
to removal of organic impurities from the nanoparticles after the fast thermal recording all NND and cell area values, we tallied a distribution of these values.
treatment (Supplementary Fig. 2).
Computational tools. DFT calculations were performed with the software
Microscopy characterization techniques. Low-magnification TEM was package VASP 5.432–36 and the projector-augmented wave (PAW) method using
performed on a Tecnai system operating at 200 kV. STEM data were collected using standard PAWs and the functional PBE37 with the zero damping DFT-D3 method
a FEI Titan electron microscope equipped with an aberration corrector for the of Grimme38. An energy cutoff of 400 eV and Gaussian smearing with a width of
probe-forming optics. The microscope was operated at an accelerating voltage of 0.1 eV were used in all calculations. The cell parameters of γ-Al2O3 were optimized
300 kV and the convergence angle was ~14 mrad. HAADF images were collected using an increased cutoff of 800 eV and the optimized lattice constants were
with a Fischione Model 3000 annular detector using a camera length that set to the a = 5.593 Å, b = 8.416 Å and c = 8.075 Å. The structure was based on the model
inner collection angle to ~71 mrad. Samples were prepared by dropcasting dilute proposed by Digne and others39. A Γ-centred (4 × 3 × 1)-k-point mesh was used
nanoparticle solutions or isopropanol dispersions of powder catalysts directly onto for the γ-Al2O3 (100) surface and (2 × 2 × 1) for the γ-Al2O3 (110) surface. The
carbon-coated Cu grids. Particle size distributions were calculated by measuring surfaces were modelled as eight-layer slabs separated by 16 Å vacuum. The five
more than 100 nanoparticle diameters, by hand, using ImageJ software31. lowest layers were kept frozen. The calculations were carried out spin-polarized

Nature Catalysis | www.nature.com/natcatal


Nature Catalysis Articles
only for PdO and Pd(OH)2, but the most stable adsorbate structures showed no 19. Zhang, Z. et al. Thermally stable single atom Pt/m-Al2O3 for selective
spin polarization. As a reference value for PdO bulk, ∆fH° = −1.217 eV (ref. 40) hydrogenation and CO oxidation. Nat. Commun. 8, 16100 (2017).
was used and the temperature-dependent chemical potential was calculated 20. Newton, M. A., Belver-Coldeira, C., Martínez-Arias, A. & Fernández-García,
according to ∆fG(PdO bulk) = ∆fH° − 0.5 𝜇O2(T). The particle size dependence of M. Dynamic in situ observation of rapid size and shape change of supported
PdO was accounted for using the Gibbs–Thomson equation28 with a surface energy Pd nanoparticles during CO/NO cycling. Nat. Mater. 6, 528–532 (2007).
of 3.30 eV nm−2 (ref. 41) and a volume of 0.024 nm3 per PdO unit. Temperature 21. Peterson, E. J. et al. Low-temperature carbon monoxide oxidation catalysed
dependence was taken into account for the chemical potentials of O2 and H2O by regenerable atomically dispersed palladium on alumina. Nat. Commun. 5,
(𝜇O2(T) and 𝜇H2O(T)) which were calculated relative to their value at T = 0 K using 4885 (2014).
the ideal gas approximation with the Atomic Simulation Environment (ASE)42. Due 22. Challa, S. R. et al. Relating rates of catalyst sintering to the disappearance of
to DFT uncertainty, all adsorption energies are corrected using Coupled Cluster individual nanoparticles during Ostwald ripening. J. Am. Chem. Soc. 133,
single point energy calculations with the def2-QZVPP basis set at the CCSD(T) 20672–20675 (2011).
level of theory with Turbomole 7.1 (TURBOMOLE V7.1 2016, University 23. Campbell, C. T., Parker, S. C. & Starr, D. E. The effect of size-dependent
of Karlsruhe and Forschungszentrum Karlsruhe 1989–2007, TURBOMOLE nanoparticle energetics on catalyst sintering. Science 298, 811–814 (2002).
since 2007; available from http://www.turbomole.com) by comparison of the 24. Huang, W., Goodman, E. D., Losch, P. & Cargnello, M. Deconvoluting
adsorption energies of Pd, PdO and Pd(OH)2 on AlOH3 + H2O. AlOH3 + H2O is a transient water effects on the activity of Pd methane combustion catalysts.
reasonable model of the γ-Al2O3 surfaces because of similarities in their molecular Ind. Eng. Chem. Res. 57, 10261–10268 (2018).
structures. The correction weakens the adsorption of Pd and strengthens that of 25. Schwartz, W. R., Ciuparu, D. & Pfefferle, L. D. Combustion of methane
PdO and Pd(OH)2. The formation energies ∆Eform of PdO, PdO2 and Pd(OH)2 over palladium-based catalysts: catalytic deactivation and role of the support.
from atomic Pd, H2O and O2 in the gas phase were also computed using single J. Phys. Chem. C 116, 8587–8593 (2012).
point energy calculations at the CCSD(T)/def2-QZVPP level of theory. The 26. Wischert, R., Laurent, P., Copéret, C., Delbecq, F. & Sautet, P. γ-Alumina: the
following values were obtained: ∆Eform(PdO) = 0.24 eV, ∆Eform(PdO2) = 1.13 eV and essential and unexpected role of water for the structure, stability and
∆Eform(Pd(OH)2) = −1.61 eV. Due to the low adsorption energy and the high energy reactivity of ‘defect’ sites. J. Am. Chem. Soc. 134, 14430–14449 (2012).
of formation of PdO2 compared to PdO und Pd(OH)2 on the γ-Al2O3 (100) surface, 27. Ouyang, R., Liu, J. & Li, W. Atomistic theory of ostwald ripening and
PdO2 was not investigated further. disintegration of supported metal particles under reaction conditions. J. Am.
Chem. Soc. 135, 1760–1771 (2013).
Data availability 28. Parker, S. C. & Campbell, C. T. Kinetic model for sintering of supported
The datasets generated during and/or analysed during the current study are metal particles with improved size-dependent energetics and applications to
available from the corresponding author on reasonable request. Au on TiO2 (110). Phys. Rev. B 75, 035430 (2007).
29. Bruix, A. et al. Maximum noble-metal efficiency in catalytic materials:
atomically dispersed surface platinum. Angew. Chem. Int. Ed. 53,
Received: 16 May 2019; Accepted: 25 June 2019; 10525–10530 (2014).
Published: xx xx xxxx 30. Wu, L. et al. Tuning precursor reactivity toward nanometer-size control in
palladium nanoparticles studied by in situ small angle X-ray scattering.
References Chem. Mater. 30, 1127–1135 (2018).
1. Argyle, M. D. & Bartholomew, C. H. Heterogeneous catalyst deactivation and 31. Schneider, C. A., Rasband, W. S. & Eliceiri, K. W. NIH image to ImageJ: 25
regeneration: a review. Catalysts 5, 145–269 (2015). years of image analysis. Nat. Methods 9, 671–675 (2012).
2. Bartholomew, C. H. Mechanisms of catalyst deactivation. Appl. Catal. A 212, 32. Hu, C. H. et al. Modulation of catalyst particle structure upon support
17–60 (2001). hydroxylation: ab initio insights into Pd13 and Pt13/γ-Al2O3. J. Catal. 274,
3. Tollefson, J. Worth its weight in platinum. Nature 450, 334–335 (2007). 99–110 (2010).
4. Robert, J. Farrauto Low-temperature oxidation of methane. Science 337, 33. Kresse, G. & Hafner, J. Ab initio molecular dynamics for liquid metals.
659–661 (2012). Phys. Rev. B 47, 558–561 (1993).
5. Liu, J., Ji, Q., Imai, T., Ariga, K. & Abe, H. Sintering-resistant nanoparticles in 34. Kresse, G. & Hafner, J. Ab initio molecular-dynamics simulation of the
wide-mouthed compartments for sustained catalytic performance. Sci. Rep. 7, liquid-metal–amorphous-semiconductor transition in germanium.
41773 (2017). Phys. Rev. B 49, 14251–14269 (1994).
6. Prieto, G., Zecevic, J., Friedrich, H., de Jong, K. & de Jongh, P. Towards stable 35. Kresse, G. & Furthmüller, J. Efficiency of ab-initio total energy calculations
catalysts by controlling collective properties of supported metal nanoparticle. for metals and semiconductors using a plane-wave basis set. Comput. Mater.
Nat. Mater. 12, 34–39 (2013). Sci. 6, 15–50 (1996).
7. Prieto, G., Meeldijk, J. D., de Jong, K. P. & de Jongh, P. E. Interplay between 36. Kresse, G. & Furthmüller, J. Efficient iterative schemes for ab initio
pore size and nanoparticle spatial distribution: consequences for the stability total-energy calculations using a plane-wave basis set. Phys. Rev. B 54,
of CuZn/SiO2 methanol synthesis catalysts. J. Catal. 303, 31–40 (2013). 11169–11186 (1996).
8. Goodman, E. D., Schwalbe, J. A. & Cargnello, M. Mechanistic understanding 37. Perdew, J. P., Burke, K. & Ernzerhof, M. Generalized gradient approximation
and the rational design of sinter-resistant heterogeneous catalysts. ACS Catal. made simple. Phys. Rev. Lett. 77, 3865–3868 (1996).
7, 7156–7173 (2017). 38. Grimme, S., Antony, J., Ehrlich, S. & Krieg, H. A consistent and accurate ab
9. Scott, S. L. A matter of life(time) and death. ACS Catal. 8, 8597–8599 (2018). initio parametrization of density functional dispersion correction (DFT-D)
10. Hansen, T. W., Delariva, A. T., Challa, S. R. & Datye, A. K. Sintering for the 94 elements H-Pu. J. Chem. Phys. 132, 154104 (2010).
of catalytic nanoparticles: particle migration or Ostwald ripening? 39. Digne, M., Sautet, P., Raybaud, P., Euzen, P. & Toulhoat, H. Use of DFT to
Acc. Chem. Res. 46, 1720–1730 (2013). achieve a rational understanding of acid–basic properties of γ-alumina
11. Cargnello, M. et al. Efficient removal of organic ligands from supported surfaces. J. Catal. 226, 54–68 (2004).
nanocrystals by fast thermal annealing enables catalytic studies on 40. Nell, J. & O’Neill, H. S. C. Gibbs free energy of formation and heat capacity
well-defined active phases. J. Am. Chem. Soc. 137, 6906–6911 (2015). of PdO: a new calibration of the Pd–PdO buffer to high temperatures and
12. Monai, M., Montini, T., Gorte, R. J. & Fornasiero, P. Catalytic oxidation of pressures. Geochim. Cosmochim. Acta 60, 2487–2493 (1996).
methane: Pd and beyond. Eur. J. Inorg. Chem. 2018, 2884–2893 (2018). 41. Rogal, J., Reuter, K. & Scheffler, M. Thermodynamic stability of PdO surfaces.
13. Willis, J. J. et al. Systematic structure–property relationship studies in Phys. Rev. B 69, 075421 (2004).
palladium-catalyzed methane complete combustion. ACS Catal. 7, 42. Larsen, A. H., Mortensen, J. J., Blomqvist, J. & Jacobsen, K. W. The
7810–7821 (2017). atomic simulation environment—a Python library for working with atoms.
14. Zhu, G., Han, J., Zemlyanov, D. Y. & Ribeiro, F. H. The turnover rate for the J. Phys. Condens. Matter 29, 273002 (2017).
catalytic combustion of methane over palladium is not sensitive to the
structure of the catalyst. J. Am. Chem. Soc. 126, 9896–9897 (2004). Acknowledgements
15. Schwartz, W. R. & Pfefferle, L. D. Combustion of methane over palladium- The authors acknowledge support from the US Department of Energy, Chemical Sciences,
based catalysts: support interactions. J. Phys. Chem. C 116, 8571–8578 (2012). Geosciences and Biosciences Division of the Office of Basic Energy Sciences, via grant
16. Otto, K., Haack, L. P. & DeVries, J. E. Identification of two types of oxidized no. DE-AC02-76SF00515 to the SUNCAT Center for Interface Science and Catalysis.
palladium on γ-alumina by X-ray photoelectron spectroscopy. Appl. Catal. B E.D.G. acknowledges support from the National Science Foundation Graduate Research
Environ. 1, 1–12 (1992). Fellowship under grant no. DGE-1656518. M.C. acknowledges support from the School of
17. Datye, A. K., Xu, Q., Kharas, K. C. & Mccarty, J. M. Particle size distributions Engineering at Stanford University and from a Terman Faculty Fellowship. Part of this work
in heterogeneous catalysts: what do they tell us about the sintering was performed at the Stanford Nano Shared Facilities (SNSF), supported by the National
mechanism? Catal. Today 111, 59–67 (2006). Science Foundation under award no. ECCS-1542152. Use of the Stanford Synchrotron
18. Kwak, J. H., Hu, J., Mei, D., Yi, C. & Kim, D. H. Coordinatively unsaturated Radiation Lightsource, SLAC National Accelerator Laboratory, is supported by the US
Al3+ centers as binding sites for active catalyst phases of platinum on γ-Al2O3. Department of Energy, Office of Science, Office of Basic Energy Sciences under contract
Science 325, 1670–1673 (2009). no. DE-AC02-76F00515. A.S.H. and S.R.B. acknowledge support from the Department of

Nature Catalysis | www.nature.com/natcatal


Articles Nature Catalysis
Energy, Basic Energy Sciences Funded Consortium for Operando and Advanced Catalyst Competing interests
Characterization via Electronic Spectroscopy and Structure (Co-ACCESS) at SLAC. O. The authors declare no competing interests.
Mueller is thanked for beamtime assistance. E.M.D. and P.N.P. acknowledge support from
the state of Baden-Württemberg, Germany through bwHPC (bwunicluster and JUSTUS,
RV bw16G001 and bw17D011) and financial support from the Helmholtz Association. Additional information
Supplementary information is available for this paper at https://doi.org/10.1038/
Author contributions s41929-019-0328-1.
E.D.G. and M.C. conceived and designed the experiments. E.D.G. performed catalyst Reprints and permissions information is available at www.nature.com/reprints.
synthesis and testing. A.C.J.-P. performed HAADF-STEM characterization. E.M.D. Correspondence and requests for materials should be addressed to M.C.
performed DFT calculations with support from P.N.P. C.J.W. performed X-ray
absorption spectroscopy analysis with support from A.S.H. and S.R.B. F.A.-P. contributed Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in
to the discussion of atomic energetics. E.D.G. and M.C. wrote the manuscript with published maps and institutional affiliations.
contributions and discussions from all authors. © The Author(s), under exclusive licence to Springer Nature Limited 2019

Nature Catalysis | www.nature.com/natcatal

You might also like