ArticlePDF Available

Abstract and Figures

A series of Cu-Pd alloy nanoparticles supported on Al2O3 were prepared and tested as catalysts for deNO x reactions. XRD, HAADF-STEM, XAFS, and FT-IR analyses revealed that a single-atom alloy structure was formed when the Cu/Pd ratio was 5, where Pd atoms were well isolated by Cu atoms. Compared with Pd/Al2O3, Cu5Pd/Al2O3 exhibited outstanding catalytic activity and N2 selectivity in the reduction of NO by CO: for the first time, the complete conversion of NO to N2 was achieved even at 175 °C, with long-term stability for at least 30 h. High catalytic performance was also obtained in the presence of O2 and C3H6 (model exhaust gas), where a 90% decrease in Pd use was achieved with minimum evolution of N2O. Kinetic and DFT studies demonstrated that N-O bond breaking of the (NO)2 dimer was the rate-determining step and was kinetically promoted by the isolated Pd.
This content is subject to copyright. Terms and conditions apply.
Chemical
Science
rsc.li/chemical-science
ISSN 2041-6539
Volume 10 Number 36 28 September 2019 Pages 8275–8492
EDGE ARTICLE
Shinya Furukawa et al.
A Cu–Pd single-atom alloy catalyst for highly e cient
NO reduction
ACuPd single-atom alloy catalyst for highly
ecient NO reduction
Feilong Xing,
a
Jaewan Jeon,
a
Takashi Toyao,
ab
Ken-ichi Shimizu
ab
and Shinya Furukawa *
ab
A series of CuPd alloy nanoparticles supported on Al
2
O
3
were prepared and tested as catalysts for deNO
x
reactions. XRD, HAADF-STEM, XAFS, and FT-IR analyses revealed that a single-atom alloy structure was
formed when the Cu/Pd ratio was 5, where Pd atoms were well isolated by Cu atoms. Compared with
Pd/Al
2
O
3
,Cu
5
Pd/Al
2
O
3
exhibited outstanding catalytic activity and N
2
selectivity in the reduction of NO
by CO: for the rst time, the complete conversion of NO to N
2
was achieved even at 175 C, with long-
term stability for at least 30 h. High catalytic performance was also obtained in the presence of O
2
and
C
3
H
6
(model exhaust gas), where a 90% decrease in Pd use was achieved with minimum evolution of
N
2
O. Kinetic and DFT studies demonstrated that NO bond breaking of the (NO)
2
dimer was the rate-
determining step and was kinetically promoted by the isolated Pd.
Introduction
The reactions of nitric oxide (NO) have garnered intense interest
from researchers in the human health,
1
and bioinorganic,
2
industrial,
3
and environmental chemistry elds.
4
Specically,
NO removal has long been studied as an indispensable process
for exhaust-gas purication.
5
Platinum-group metals (PGMs)
such as Pt, Pd, and Rh are known to be ecient catalysts for the
reduction of NO using CO,
6,7
H
2
,
8
NH
3
,
9
and hydrocarbons
10
as
reductants. The recent challenges in this eld involve devel-
oping catalytic systems that function (1) at low temperatures
under cold-start conditions,
11
(2) with minimum use of
PGMs,
12,13
and (3) without emitting N
2
O,
1416
which is a potent
greenhouse gas.
17
These issues have been individually studied
using dierent materials. The development of a single material
that enables (1)(3) is therefore highly desirable. To the best of
our knowledge, no such material has been reported. In partic-
ular, achieving both (1) and (3) is dicult because N
2
O reduc-
tion to N
2
on PGMs requires relatively high temperatures (>300
C).
18
Therefore, an appropriate catalyst design is needed to
obtain not only outstanding catalytic activity toward NO
reduction but also high selectivity to N
2
with minimal incor-
poration of PGMs.
A promising approach that overcomes these challenges is the
single-atom alloying concept,
19
which is relevant to single-atom
chemistry.
20,21
The dilution of PGM atoms with less active metal
atoms not only substantially reduces the use of PGMs but also
enables drastic modication of the electronic and geometric
structures for enhanced catalysis.
22
For example, the isolation of
Pt or Pd with group 11 metals (Au, Ag, and Cu) enables molec-
ular transformations that hardly proceed in the absence of
single-atom alloying, such as selective hydrogenation,
2327
for-
mic acid dehydrogenation,
28
and hydrosilylation.
29
In these
systems, the group 11 metals act as inert elements but modify
the electronic and geometric factors of the PGM and, thus, its
catalytic properties. Conversely, for NO reduction, the group 11
elements are known to be capable of NO activation.
3032
There-
fore, applying the single-atom alloying concept to NO reduction
systems should provide an unprecedented synergistic eect for
ecient NO conversion.
In this study, we focused on Cu as a main component
because of its intrinsic activity toward NO reduction and its
high earth abundance. We found that CuPd/Al
2
O
3
(Cu/Pd ¼5)
acts as a highly ecient catalyst for NO reduction at low
temperatures (>150 C), without generating N
2
O emissions.
Herein, we report both an innovative catalytic system for e-
cient NO reduction and novel catalytic chemistry of single-atom
alloys.
Experimental details
Catalyst preparation
Boehmite (g-AlOOH) was supplied by SASOL Chemicals. g-Al
2
O
3
was prepared by the calcination of boehmite at 900 C for 3 h.
Pd/Al
2
O
3
(Pd: 2 wt%) and CuPd/Al
2
O
3
(Cu: 6 wt%, Cu/Pd ¼1)
were prepared by a conventional impregnation method. The g-
Al
2
O
3
support was added to a vigorously stirred aqueous
a
Institute for Catalysis, Hokkaido University, N-21, W-10, Sapporo 001-0021, Japan.
E-mail: furukawa@cat.hokudai.ac.jp; Fax: +81-11-706-9163
b
Elements Strategy Initiative for Catalysts and Batteries, Kyoto University, Katsura,
Kyoto 615-8520, Japan
Electronic supplementary information (ESI) available: Details of
characterization, kinetic analysis, and DFT calculations. See DOI:
10.1039/c9sc03172c
Cite this: Chem. Sci.,2019,10,8292
All publication charges for this article
have been paid for by the Royal Society
of Chemistry
Received 27th June 2019
Accepted 5th August 2019
DOI: 10.1039/c9sc03172c
rsc.li/chemical-science
8292 |Chem. Sci.,2019,10,82928298 This journal is © The Royal Society of Chemistry 2019
Chemical
Science
EDGE ARTICLE
solution containing Pd(NH
3
)
2
(NO
2
)
2
(Kojima Chemicals,
4.765 wt% in HNO
3
) and/or Cu(NO
3
)
2
$3H
2
O (Sigma-Aldrich,
99%), followed by stirring for 3 h at room temperature (50 ml
H
2
O per gram of Al
2
O
3
). The mixture was dried under a reduced
pressure at 50 C, followed by reduction under owing H
2
(30
ml min
1
) at 400 C (Pd) or 800 C (CuPd) for 1 h. The Cu/Al
2
O
3
(Cu: 6 wt%) and CuPd/Al
2
O
3
(Cu: 6 wt%, Cu/Pd ¼3 and 5)
catalysts were prepared by a depositionprecipitation method
using urea. The g-Al
2
O
3
support was added to a vigorously
stirred aqueous solution of Cu(NO
3
)
2
$3H
2
O (50 ml H
2
O per
gram of Al
2
O
3
). Then, an aqueous solution of urea (Kanto, 99%)
was added dropwise to the stirred mixture at room temperature
(urea/Cu ¼30). The mixture was sealed with a plastic lm and
kept with stirring at 70 C for 10 h. Aer completing the
precipitation of Cu(OH)
2
, the supernatant was decanted and the
resulting crude product was washed with deionized water three
times, followed by drying under a reduced pressure at 50 C and
calcination at 500 C for 1 h. For Cu/Al
2
O
3
, the resulting CuO/
Al
2
O
3
was reduced under owing H
2
at 400 C for 1 h. For Cu
Pd/Al
2
O
3
(Cu/Pd ¼3 and 5), the resulting CuO/Al
2
O
3
was used
for successive impregnation of Pd in a similar fashion to that
mentioned above. The resulting PdCuO/Al
2
O
3
was reduced
under owing H
2
at 400 C for 1 h.
Reaction conditions
The catalyst (0.05 g) diluted with quartz sand (1.95 g, Miyazaki
Chemical, 99.9%) was treated under owing hydrogen (50
ml min
1
) at 400 C for 0.5 h prior to the catalytic reactions. NO
reduction by CO was performed in a xed-bed continuous ow
system by feeding NO (5000 ppm), CO (5000 ppm), and He
(balance) with a total ow rate of 96 ml min
1
(GHSV ¼80 000
h
1
). The gas phase was analyzed using an online thermal
conductivity detection gas chromatograph (Shimadzu GC-8A,
column: SHINWA SHINCARBON ST) located downstream. A
stability test was done using twice the amount of catalyst (0.10
g) under similar conditions (GHSV ¼40 000 h
1
). Aer a time-
on-stream of 24 h, the catalyst was regenerated by owing
hydrogen (50 ml min
1
) at 400 C for 0.5 h, followed by
continuing the catalytic run. A kinetic study was performed by
changing the concentration of NO and CO between 0.3 and
0.6% with that of the counterpart xed at 0.5%. The reaction
temperature was maintained at 150 C so that NO conversion
did not exceed 30%, and the reaction rate (mol s
1
mol
Pd
1
) was
calculated on the basis of NO conversion. NO + CO + O
2
and NO
+CO+O
2
+C
3
H
6
reactions were performed under stoichio-
metric conditions as follows: NO (5000 ppm), CO (10 000 ppm),
O
2
(2500 ppm), He (balance) with a total ow rate of 96
ml min
1
(GHSV ¼80 000 h
1
), and NO (5000 ppm), CO (5000
ppm), O
2
(5625 ppm), C
3
H
6
(1250 ppm), and He (balance) with
a total ow rate of 96 ml min
1
(GHSV ¼80 000 h
1
),
respectively.
Characterization
The crystal structure of the prepared catalyst was examined by
powder X-ray diraction (XRD) using a Rigaku MiniFlex II/AP
diractometer with Cu Karadiation. High-angle annular dark
eld scanning transmission electron microscopy (HAADF-
STEM) was carried out using a JEOL JEM-ARM200 M micro-
scope equipped with an energy dispersive X-ray (EDX) analyzer
(EX24221M1G5T). STEM analysis was performed at an acceler-
ating voltage of 200 kV. To prepare the TEM specimen, all
samples were sonicated in ethanol and then dispersed on a Mo
grid supported by an ultrathin carbon lm.
The Fourier-transformed infrared (FT-IR) spectra of adsor-
bed CO were obtained with a JASCO FTIR-4200 spectrometer
equipped with an MCT detector in transmission mode (reso-
lution 4 cm
1
). The samples were prepared as self-supporting
wafers (2.0 cm diameter, <0.5 mm thickness) and were placed
inside an IR cell with CaF
2
windows. A custom glass manifold
was connected to the cell to control the gas for pretreatment and
the amount of CO introduced. The cell was rst purged with He,
and the sample was reduced under owing hydrogen (50
ml min
1
) at 400 C for 30 min. Aer reduction, the wafer was
cooled to 40 C under owing He. The wafer was exposed to CO
(0.5%) and He (balance) with a total ow rate of 50 ml min
1
for
20 min. Aer the CO exposure, He was owed for 5 min to
remove the gas phase and weakly adsorbed CO, followed by IR
spectral measurements.
X-ray absorption ne structure (XAFS) spectra were recorded
on the BL14B2 station at SPring-8 of the Japan Synchrotron
Radiation Research Institute. A Si(311) double-crystal mono-
chromator was used. Energy calibration was performed using
Pd foil. The spectra were recorded at the edges of Pd K in
a transmission mode at room temperature. The pelletized
sample was pre-reduced with H
2
at 400 C for 0.5 h, and then
sealed in a plastic pack under a N
2
atmosphere together with an
ISO A500-HS oxygen absorber (Fe powder). The obtained XAFS
spectra were analyzed using Athena and Artemis soware ver.
0.9.25 included in the Demeter package. The Fourier trans-
formation of the k
3
-weighted EXAFS from kspace to Rspace was
performed over a krange of 3.015 ˚
A
1
. Some of the Fourier-
transformed EXAFS spectra in the Rrange of 1.23.0 ˚
A were
inversely Fourier transformed, followed by an analysis using
a usual curve tting method in a krange of 315 ˚
A
1
. The back-
scattering amplitude or phase shiparameters were simulated
with FEFF 6L and used to perform the curve tting procedure.
For PdCu scattering, intermetallic Cu
3
Pd with a Pm
3mstruc-
ture was considered for the FEFF simulation. The amplitude
reduction factor (S
02
) of Pd was determined to be 0.775 by tting
the spectra of Pd black and then used for tting of other EXAFS
spectra.
Computational details
Periodic DFT calculations were performed using the CASTEP
code
33
with Vanderbilt-type ultrasopseudopotentials
34
and the
PerdewBurkeErnzerhof exchangecorrelation functional
based on the generalized gradient approximation.
35
The plane-
wave basis set was truncated at a kinetic energy of 350 eV,
and a Fermi smearing of 0.1 eV was utilized. Dispersion corre-
lations were considered using the TkatchenkoScheer
method with a scaling coecient of s
R
¼0.94 and a damping
parameter of d¼20.
36
The reciprocal space was sampled using
This journal is © The Royal Society of Chemistry 2019 Chem. Sci.,2019,10,82928298 | 8293
Edge Article Chemical Science
ak-point mesh with a spacing of typically 0.04 ˚
A
1
, as generated
by the MonkhorstPack scheme.
37
Geometry optimization was
performed on supercell structures using periodic boundary
conditions. The surface was modeled based on Cu(211)-(2 3)
(for NO and N
2
O related reactions), Cu(111)-(2 2) (for CO
oxidation), and Cu(111)-(3 3) (for N
2
O decomposition) slabs
that were six atomic layers thick with 13 ˚
A of vacuum spacing.
The convergence criteria for structural optimization and energy
calculation were set to (a) an SCF tolerance of 1.0 10
6
eV per
atom, (b) an energy tolerance of 1.0 10
5
eV per atom, (c)
a maximum force tolerance of 0.05 eV ˚
A
1
, and (d) a maximum
displacement tolerance of 1.0 10
3
˚
A. The transition state
search was performed using the complete linear synchronous
transit/quadratic synchronous transit (LST/QST) method.
38,39
Linear synchronous transit maximization was performed, fol-
lowed by energy minimization in the directions conjugate to the
reaction pathway. The approximated TS was used to perform
QST maximization with conjugate gradient minimization
renements. This cycle was repeated until a stationary point
was found. Convergence criteria for the TS calculations were set
to root-mean-square forces on an atom tolerance of 0.10 eV ˚
A
1
.
Results and discussion
The monometallic Cu (6 wt%) and Pd (2 wt%) and the bimetallic
CuPd (Cu: 6 wt%, Cu/Pd ¼1, 3, or 5; hereaer, Cu
x
Pd, x¼1, 3,
or 5) catalysts were prepared using g-Al
2
O
3
as a support by
depositionprecipitation and/or impregnation methods. X-ray
diraction (XRD) patterns of the prepared catalysts revealed
that CuPd solid-solution alloy phases with bimetallic compo-
sitions similar to the metal ratio in the feed were formed
(Fig. S1and Table 1).
The crystallite sizes estimated using Scherrer's equation
were 34 nm for all of the catalysts. Fig. 1a and b show a high-
angle annular dark eld scanning transmission electron
microscopy (HAADF-STEM) image of Cu
5
Pd/Al
2
O
3
and the size
distribution of the nanoparticles, respectively. A relatively
narrow size distribution between 2 and 6 nm with an area-
weighted mean diameter of 4.2 nm was obtained, consistent
with the crystallite size estimated by XRD (Table 1).
The energy-dispersive X-ray spectroscopy (EDS) analysis of
a single nanoparticle revealed that the Cu and Pd atoms
comprising the nanoparticle were homogeneously dispersed
(Fig. 1c). The high-resolution HAADF-STEM image shows an fcc
crystal structure viewed along the [100] direction, consistent
with the formation of a solid-solution alloy (Fig. 1d). Moreover,
the isolation of Pd atoms by Cu was indicated by the presence of
atoms with distinct Z contrasts. Note that the corresponding
HAADF-STEM image of Cu/Al
2
O
3
showed a weak and uniform Z
contrast compared with that of Cu
5
Pd/Al
2
O
3
(Fig. S2).
The degree of Pd isolation was further investigated by
Fourier-transform infrared (FT-IR) and extended X-ray absorp-
tion ne structure (EXAFS) analyses (Fig. 2). As shown in Fig. 2a,
the FT-IR spectra of CO adsorbed onto Pd/Al
2
O
3
exhibited
absorption peaks assigned to the stretching vibration of C]O
Table 1 Detailed information on the catalyst prepared in this study
Pd CuPd Cu
3
Pd Cu
5
Pd Cu
Pd loading (wt%) 2.0 10.1 3.4 2.0 0
Cu loading (wt%) 0 6.0 6.0 6.0 6.0
Cu fraction in the catalyst 0 0.50 0.75 0.83 1.0
Cu fraction in particles 0 0.44 0.72 0.89 1.0
Crystallite size/nm
a
3.2 3.3 3.8 3.8 (4.2)
b
4.0
a
Estimated from Scherrer's equation using a Scherrer constant of 0.477
for the area-weighted mean diameter.
b
Area-weighted mean diameter
obtained from TEM images.
Fig. 1 (a) HAADF-STEM image of Cu
5
Pd/Al
2
O
3
and (b) size distribution
of the nanoparticles. (c) Elemental map of the Pd + Cu overlayer, as
acquired by EDS. (d) High-resolution image of a single nanoparticle.
Fig. 2 (a) FT-IR spectra of CO adsorbed on the prepared catalysts. (b)
Fourier transforms of the Pd K-edge EXAFS spectra of the Pd-based
catalysts.
8294 |Chem. Sci.,2019,10,82928298 This journal is © The Royal Society of Chemistry 2019
Chemical Science Edge Article
adsorbed on top (2086 cm
1
), bridge (1975 cm
1
), and hollow
sites (1880 cm
1
).
40
Similar absorption peaks were observed for
CuPd/Al
2
O
3
, suggesting that the PdPd ensembles largely
remain even aer 1 : 1 alloying. For Cu-rich samples, an
absorption band assignable to CO adsorbed on metallic Cu was
also observed at 2100 nm
1
.
41
The peak intensities for the bridge
and hollow CO substantially decreased and disappeared in the
spectra of Cu
3
Pd/Al
2
O
3
and Cu
5
Pd/Al
2
O
3
, respectively, indicating
that Pd atoms at the surface were isolated upon 5 : 1 alloying.
There remained a weak absorption band for linear CO on Pd
in the spectrum for Cu
5
Pd/Al
2
O
3
, suggesting that the isolated Pd
atoms are also present at the surface. Fig. 2b shows the Fourier
transforms of the Pd K-edge EXAFS spectra of the Pd-based
catalysts (the X-ray absorption near edge structure spectra,
raw EXAFS oscillations, curve-tting, and summary of EXAFS
curve tting are shown in Fig. S3S5 and Table S1,respec-
tively). CuPd/Al
2
O
3
showed both PdPd and PdCu bonds, while
Cu
5
Pd/Al
2
O
3
exclusively showed PdCu bonds, suggesting that
the Pd atoms in the bulk were also isolated by Cu upon 5 : 1
alloying. Thus, small CuPd nanoparticles with a single-atom
alloy structure were successfully formed on the Al
2
O
3
support.
Considering the limited sensitivity of EXAFS and FT-IR, we
cannot completely exclude the presence of PdPd interaction.
However, only a small number of PdPd sites, if any, seem not to
contribute to the overall catalytic performance.
We next tested the catalytic activity of Cu
x
Pd/Al
2
O
3
in NO
reduction by CO (GHSV ¼80 000 h
1
), as a model reaction for
exhaust-gas purication. Fig. 3a shows the NO conversion to N
2
(C
N
2
) for the prepared catalysts as a function of reaction
temperature. Here, C
N
2
was obtained by multiplying the NO
conversion and the N
2
selectivity (Fig. S6). Pd/Al
2
O
3
gave the
lowest C
N
2
, because of the poor N
2
selectivity <40% (Fig. S6b).
Cu/Al
2
O
3
exhibited a higher C
N
2
than Pd/Al
2
O
3
because of its
much higher N
2
selectivity (Fig. S6b). CuPd/Al
2
O
3
showed a C
N
2
trend similar to that of Cu/Al
2
O
3
because of the consequence of
increased NO conversion and decreased N
2
selectivity (Fig. S6).
Thus, the 1 : 1 alloy of Cu and Pd gave an insucient catalytic
performance for selective NO reduction. Interestingly, however,
both NO conversion and N
2
selectivity increased when the
alloying ratio was increased to 3 : 1 and 5 : 1 (Fig. S6), which
resulted in great enhancements in C
N
2
(Fig. 3a). NO was
completely converted to N
2
over Cu
5
Pd/Al
2
O
3
without gener-
ating N
2
O emissions even at 200 C, where Pd showed a C
N
2
of
only 5%. Notably, on going from CuPd to Cu
5
Pd, the catalytic
activity increased even though the Pd content was decreased to
1/5 (Table 1). Therefore, a specic synergistic eect between Cu
and Pd likely contributed to the unique properties of the single-
atom alloy catalyst. We emphasize that using an excess amount
of Pd/Al
2
O
3
(0.50 g) with 10 times the equimolar Pd included in
Cu
5
Pd/Al
2
O
3
(labeled as Pd 10) still resulted in poor perfor-
mance (Fig. 3b), highlighting the outstanding performance of
the single-atom alloy catalyst. We also tested the long-term
stability of Cu
5
Pd/Al
2
O
3
in NO reduction by CO under stan-
dard conditions (GHSV ¼40 000 h
1
), where 100% C
N
2
was
maintained at 175 C. Although a number of bimetallic catalysts
for NO reduction have been reported thus far,
1216,4248
to the best
of our knowledge, the present work represents the rst success
in complete NO
x
removal at a temperature less than 200 C. At
150 C, although C
N
2
decreased slightly at the initial stage
because of N
2
O formation, it recovered aer a short H
2
treat-
ment. This result implies that the accumulation of oxygen
species at the catalyst surface triggers the loss of N
2
selectivity
and that the catalytic performance could be recovered under
rich conditions. We next examined the catalytic performance of
Cu
5
Pd/Al
2
O in NO reduction in the presence of O
2
and O
2
+
C
3
H
6
; these conditions more closely resemble those encoun-
tered in practical use. Cu/Al
2
O
3
delivered poor performance
under NO + CO + O
2
conditions. By contrast, Cu
5
Pd/Al
2
O
exhibited much higher performance than Cu/Al
2
O and Pd/Al
2
O.
Notably, Cu
5
Pd/Al
2
O still exhibited a performance better than or
comparable to Pd 10even in the presence of O
2
or O
2
+
C
3
H
6
, respectively (Fig. 3b and S7;a comparison with
temperature dependence and T
50
is presented in Fig. S7).
Furthermore, N
2
O evolution was suciently suppressed over
Cu
5
Pd, where the C
N
2
O
(NO conversion to N
2
O, Fig. 3b and S7)
was much lower than those for Pd. Thus, the single-atom alloy
catalyst enabled not only a decrease in the noble metal use to 1/
10 but also highly selective NO
x
removal. In the reactions con-
ducted in the presence of O
2
and O
2
+C
3
H
6
, reaction temper-
atures greater than 200 C were needed to achieve sucient
catalytic performance (Fig. 3 and S7). A possible explanation is
that the number of active sites for NO reduction decreased
because of the involvement of other reactions such as CO and/or
C
3
H
6
oxidation.
Fig. 3 (a) NO conversion to N
2
during the NO reduction by CO over
Pd, Cu, and CuPd catalysts as a function of reaction temperature
(NO, CO: 0.5%, GHSV ¼80 000 h
1
). (b) Comparison between C
N
2
and
C
N
2
O
in NO reduction in the presence of O
2
and C
3
H
6
. (c) Stability test
for Cu
5
Pd/Al
2
O
3
in the NO + CO reaction at low temperatures (NO,
CO: 0.5%, GHSV ¼40 000 h
1
).
This journal is © The Royal Society of Chemistry 2019 Chem. Sci.,2019,10,82928298 | 8295
Edge Article Chemical Science
Next, to clarify the nature of the synergistic eect, we con-
ducted a mechanistic study based on kinetic analysis and
density functional theory (DFT) calculations. First, the apparent
activation energy (E
A
) for NO reduction by CO was estimated for
the representative catalysts. The corresponding Arrhenius-type
plots and the resulting E
A
values are shown in Fig. 4 and
Table 2, respectively. Cu
5
Pd gave an E
A
value lower than those of
Pd and Cu, which is consistent with the observed catalytic
activity. In addition, we estimated the reaction orders for NO
and CO pressures (P
NO
and P
CO
, respectively) to consider the
rate-determining step (RDS). Both Cu and Cu
5
Pd showed
negative and positive orders for P
NO
and P
CO
, respectively (see
Table 2 and Fig. S8for details). Unlike the case for Pd-based
catalysts,
49
bond dissociation of NO has been speculated to
occur via (NO)
2
dimer formation on Cu surfaces.
5052
Therefore,
in the present study, an extended LangmuirHinshelwood
model that includes (NO)
2
dimer formation, the subsequent
NO scission (N
2
O formation), and N
2
O decomposition (N
2
formation) was considered. We solved the rate equation of each
step regarded as the RDS by using the overall site balance and
equilibrium constants for the other steps (see the ESI,kinetic
analysis). In most cases, the reaction order for P
NO
is positive,
which is inconsistent with the observed experimental results.
Conversely, when the NO scission of the (NO)
2
is considered as
the RDS, the orders for P
NO
and P
CO
range from 2 to 0 and
from 0 to 2, respectively, in agreement with the experimental
values. Thus, our kinetic study suggests that the NO scission
was the RDS in NO reduction by CO. Upon incorporation of Pd
atoms into pure Cu, the order for P
NO
becomes less negative,
while that for P
CO
decreases substantially. This result indicates
that the strong adsorption of NO inhibits CO adsorption onto
Cu, while the latter is enhanced in the presence of Pd. We
performed DFT calculations for the relevant elemental steps on
pure and Pd-doped Cu surfaces. On the basis of the literature,
53
the step site of the (211) surface was considered the active site
for NO scission (Fig. S9). The corresponding energy diagrams
are shown in Fig. 5.
NO adsorption was weakened by the addition of Pd, which
is consistent with the change in the reaction orders (Table 2).
Dimerization occurs at the terrace site adjacent to the step site,
with a negligible energy barrier. The subsequent NO scission
is triggered by capture of an oxygen atom by the edge Cu
atoms, resulting in the formation of an on-top N
2
OwithE
A
values of 59.8 and 47.9 kJ mol
1
for pure and Pd-doped Cu,
respectively. The calculated E
A
values agree with the experi-
mental values (Table 2). The lower E
A
for the Pd-doped Cu
appears to originate from the destabilized adsorption of the
(NO)
2
dimer by Pd.
We also considered the CO oxidation process (CO + O /
CO
2
), which is necessary for the catalytic cycle (Fig. S10). The
CO + O reaction over pure and Pd-doped Cu(111) surfaces gave
E
A
values of 60.9 kJ mol
1
and 34.1 kJ mol
1
, respectively (Table
2 and Fig. S10). These values are very similar to or lower than
those for NO scission, which is consistent with the RDS being
the scission of NO. We also simulated the N
2
O decomposition
process (N
2
O/N
2
+ O) on Cu(211) and (111) surfaces to
understand the intrinsic high N
2
selectivity of Cu (Fig. 6 and
S11; see S12for the pictures of the optimized structures).
The monodentate linear N
2
O was bent to form a bidentate
N
2
O at the edge site of the Cu(211) plane without an energy
barrier. The bidentate N
2
O was subsequently decomposed into
N
2
and O with a low E
A
of 33.4 kJ mol
1
, indicating that the N
2
O
once formed could be smoothly decomposed into N
2
to aord
high N
2
selectivity. Although the Cu(111) surface was also active
Fig. 4 Arrhenius-type plots obtained in the NO + CO reaction over
Cu
5
Pd/Al
2
O
3
, Cu/Al
2
O
3
, and Pd/Al
2
O
3
catalysts.
Table 2 Activation energies estimated from experiments and from
DFT calculations, along with reaction orders
E
A
/kJ mol
1
(/eV)
Pd Cu Cu
5
Pd
Experiment 99.5 60.8 42.7
DFT: NO 100.2
a
59.8 47.9
DFT: CO + O 100.1
b
60.9 34.1
Reaction order
P
NO
0.27 0.02
P
CO
1.93 0.44
a
NO dissociation at the step of Pd(511).
16
b
CO oxidation on Pd(111).
16
Fig. 5 Energy diagrams of NO adsorption, dimerization, and the
dimer's decomposition over pure and Pd-doped Cu(211) surfaces. The
total energy of the slab and free NO was set to zero.
8296 |Chem. Sci.,2019,10,82928298 This journal is © The Royal Society of Chemistry 2019
Chemical Science Edge Article
for N
2
O decomposition in a similar fashion, the energy barrier
was higher than that of Cu(211) (51.6 kJ mol
1
, Fig. S11).
Because large CuCu ensembles are present on the surface of
the Cu and Cu-rich catalysts (Cu
5
Pd and Cu
3
Pd), N
2
O decom-
position could also be enhanced on these catalysts. However,
for CuPd, this eect is limited because of the dilution of CuCu
ensembles and the increase of PdPd ensembles. Thus, our
calculation rationalized the substantial enhancement in cata-
lytic activity on the basis of the formation of the CuPd single-
atom alloy and the origin of the excellent selectivity for N
2
formation. The elucidated mechanism diers completely from
those proposed for other bimetallic alloy systems. For example,
for the PtCo system, Sato et al. reported that alloying with Co
makes Pt electron-rich, which enhances back donation to
adsorbed NO, inducing bond breaking.
13
Therefore, Co likely
acts as a promoter for Pt. By contrast, in our system, the isolated
Pd acts as an ecient promotor for Cu.
Conclusion
We prepared a series of CuPd/Al
2
O
3
catalysts for selective NO
reduction at low temperatures. Alloying of Pd with a large
amount of Cu (Cu/Pd ¼5) isolates Pd and drastically improves
both the catalytic activity and N
2
selectivity, aording
outstanding catalytic performance. In the NO reduction by CO,
NO is completely converted to N
2
even at 175 C, with long-term
stability for at least 30 h. The high catalytic performance is also
achieved in the presence of O
2
and C
3
H
6
, where the amount of
Pd needed for a comparable performance can be reduced to 1/
10, with minimum evolution of N
2
O. For Cu/Al
2
O
3
and Cu
5
Pd/
Al
2
O
3
, the NO bond scission of the (NO)
2
dimer is the RDS in
NO reduction by CO. This step is kinetically facilitated by the
isolated Pd atoms. N
2
O decomposition to N
2
smoothly proceeds
on the Cu surface, which contributes to the excellent N
2
selec-
tivity observed for Cu and Cu-rich catalysts. The key to this
ecient catalysis is the sucient isolation of Pd atoms by Cu,
highlighting the importance of catalyst design based on single-
atom alloy structures. The insights gained in this study provide
not only a highly ecient deNO
x
system with substantially
reduced noble-metal content, but also open a new path for the
chemistry of single-atom alloys.
Conicts of interest
There are no conicts to declare.
Acknowledgements
This work was supported by JSPS KAKENHI (Grant Numbers
17H01341 and 17H04965) and by MEXT within the projects
Integrated Research Consortium on Chemical Sciences
(IRCCS)and Elements Strategy Initiative to Form Core
Research Center, as well as by the JST CREST project
JPMJCR17J3. We deeply appreciate Dr H. Asakura of Kyoto
University for help with XAFS measurement. We thank the
technical staof the Research Institute for Electronic Science,
Hokkaido University for help with HAADF-STEM observation.
Computation time was provided by the supercomputer systems
at the Institute for Chemical Research, Kyoto University. X-ray
absorption measurements were carried out at the BL-14B2
beamline of SPring-8 at the Japan Synchrotron Radiation
Research Institute (JASRI; 2018A1757).
References
1 P. Pacher, J. S. Beckman and L. Liaudet, Physiol. Rev., 2007,
87, 315424.
2 B. C. Berks, S. J. Ferguson, J. W. B. Moir and D. J. Richardson,
Biochim. Biophys. Acta, Bioenerg., 1995, 1232,97173.
3 F. Rezaei, A. A. Rownaghi, S. Monjezi, R. P. Lively and
C. W. Jones, Energy Fuels, 2015, 29, 54675486.
4 C. Monn, Atmos. Environ., 2001, 35,132.
5 J. H. Wang, H. Chen, Z. C. Hu, M. F. Yao and Y. D. Li, Catal.
Rev., 2015, 57,79144.
6 J. H. Holles, R. J. Davis, T. M. Murray and J. M. Howe, J.
Catal., 2000, 195, 193206.
7 A. Srinivasan and C. Depcik, Catal. Rev., 2010, 52, 462493.
8 J. Shibata, M. Hashimoto, K. Shimizu, H. Yoshida, T. Hattori
and A. Satsuma, J. Phys. Chem. B, 2004, 108, 1832718335.
9 W. Z. An, Q. L. Zhang, K. T. Chuang and A. R. Sanger, Ind.
Eng. Chem. Res., 2002, 41,2731.
10 R. Burch, J. P. Breen and F. C. Meunier, Appl. Catal., B, 2002,
39, 283303.
11 V. Tomasic, Catal. Today, 2007, 119, 106113.
12 S. Hosokawa, K. Matsuki, K. Tamaru, Y. Oshino, H. Aritani,
H. Asakura, K. Teramura and T. Tanaka, Mol. Catal., 2017,
442,7482.
13 K. Sato, A. Ito, H. Tomonaga, H. Kanematsu, Y. Wada,
H. Asakura, S. Hosokawa, T. Tanaka, T. Toriyama and
T. Yamamoto, ChemPlusChem, 2019.
14 T. Tanabe, T. Imai, T. Tokunaga, S. Arai, Y. Yamamoto,
S. Ueda, G. V. Ramesh, S. Nagao, H. Hirata, S. Matsumoto,
T. Fujita and H. Abe, Chem. Sci., 2017, 8, 33743378.
15 T. Imai, S. Ueda, S. Nagao, H. Hirata, K. R. Deepthi and
H. Abe, RSC Adv., 2017, 7, 96289631.
Fig. 6 Energy diagrams of N
2
O bending (initial (IS) to intermediate
(MS) states) and its subsequent decomposition to N
2
and O (MS to nal
state (FS)) over the Cu(211) surface. The total energy of slab and free
N
2
O was set to zero.
This journal is © The Royal Society of Chemistry 2019 Chem. Sci.,2019,10,82928298 | 8297
Edge Article Chemical Science
16 J. Jeon, K. Kon, T. Toyao, K. Shimizu and S. Furukawa, Chem.
Sci., 2019, 10, 41484162.
17 M. V. Twigg, Appl. Catal., B, 2007, 70,215.
18 P. Granger, C. Dujardin, J. F. Paul and G. Leclercq, J. Mol.
Catal. A: Chem., 2005, 228, 241253.
19 G. Kyriakou, M. B. Boucher, A. D. Jewell, E. A. Lewis,
T. J. Lawton, A. E. Baber, H. L. Tierney, M. Flytzani-
Stephanopoulos and E. C. H. Sykes, Science, 2012, 335,
12091212.
20 A. Wang, J. Li and T. Zhang, Nat. Rev. Chem., 2018, 2, 65.
21 M. J. H¨
ulsey, J. Zhang and N. Yan, Adv. Mater., 2018, 30,
1802304.
22 M. T. Greiner, T. E. Jones, S. Beeg, L. Zwiener, M. Scherzer,
F. Girgsdies, S. Piccinin, M. Armbruster, A. Knop-Gericke
and R. Schlogl, Nat. Chem., 2018, 10, 10081015.
23 P. Aich, H. J. Wei, B. Basan, A. J. Kropf, N. M. Schweitzer,
C. L. Marshall, J. T. Miller and R. Meyer, J. Phys. Chem. C,
2015, 119, 1814018148.
24 G. X. Pei, X. Y. Liu, A. Q. Wang, A. F. Lee, M. A. Isaacs, L. Li,
X. L. Pan, X. F. Yang, X. D. Wang, Z. J. Tai, K. Wilson and
T. Zhang, ACS Catal., 2015, 5, 37173725.
25 X. X. Cao, A. Mirjalili, J. Wheeler, W. Xie and B. W. L. Jang,
Front. Chem. Sci. Eng., 2015, 9, 442449.
26 F. R. Lucci, J. L. Liu, M. D. Marcinkowski, M. Yang,
L. F. Allard, M. Flytzani-Stephanopoulos and
E. C. H. Sykes, Nat. Commun., 2015, 6, 8550.
27 J. L. Liu, J. J. Shan, F. R. Lucci, S. F. Cao, E. C. H. Sykes and
M. Flytzani-Stephanopoulos, Catal. Sci. Technol., 2017, 7,
42764284.
28 M. D. Marcinkowski, J. L. Liu, C. J. Murphy, M. L. Liriano,
N. A. Wasio, F. R. Lucci, M. Flytzani-Stephanopoulos and
E. C. H. Sykes, ACS Catal., 2017, 7, 413420.
29 H. Miura, K. Endo, R. Ogawa and T. Shishido, ACS Catal.,
2017, 7, 15431553.
30 P. Bera, K. C. Patil and M. S. Hegde, Phys. Chem. Chem. Phys.,
2000, 2, 37153719.
31 P. Miguel, P. Granger, N. Jagtap, S. Umbarkar, M. Dongare
and C. Dujardin, J. Mol. Catal. A: Chem., 2010, 322,9097.
32 H. Iwamoto, S. Kameoka, Y. Xu, C. Nishimura and A. P. Tsai,
J. Phys. Chem. Solids, 2019, 125,6473.
33 M. D. Segall, P. J. D. Lindan, M. J. Probert, C. J. Pickard,
P. J. Hasnip, S. J. Clark and M. C. Payne, J. Phys.: Condens.
Matter, 2002, 14, 27172744.
34 D. Vanderbilt, Phys. Rev. B: Condens. Matter Mater. Phys.,
1990, 41, 78927895.
35 J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett.,
1996, 77, 38653868.
36 A. Tkatchenko and M. Scheer, Phys. Rev. Lett., 2009, 102,
073005.
37 H. J. Monkhorst and J. D. Pack, Phys. Rev. B, 1976, 13, 5188
5192.
38 N. Govind, M. Petersen, G. Fitzgerald, D. King-Smith and
J. Andzelm, Comput. Mater. Sci., 2003, 28, 250258.
39 T. A. Halgren and W. N. Lipscomb, Chem. Phys. Lett., 1977,
49, 225232.
40 W. K. Kuhn, J. Szanyi and D. W. Goodman, Surf. Sci., 1992,
274, L611L618.
41 O. Dulaurent, X. Courtois, V. Perrichon and D. Bianchi, J.
Phys. Chem. B, 2000, 104, 60016011.
42 S. Zhou, B. Varughese, B. Eichhorn, G. Jackson and
K. McIlwrath, Angew. Chem., Int. Ed., 2005, 44, 45394543.
43 M. Fern´
andez-Garcia, A. Martinez-Arias, C. Belver,
J. Anderson, J. Conesa and J. Soria, J. Catal., 2000, 190,
387395.
44 F. Gao, Y. Wang and D. W. Goodman, J. Catal., 2009, 268,
115121.
45 T. Hirano, Y. Ozawa, T. Sekido, T. Ogino, T. Miyao and
S. Naito, Catal. Commun., 2007, 8, 12491254.
46 A. Hungrıa, A. Iglesias-Juez, A. Martınez-Arias,
M. Fern´
andez-Garcıa, J. Anderson, J. Conesa and J. Soria, J.
Catal., 2002, 206, 281294.
47 T. Komatsu, H. Kobayashi, K. Kusada, Y. Kubota, M. Takata,
T. Yamamoto, S. Matsumura, K. Sato, K. Nagaoka and
H. Kitagawa, Chem.Eur. J., 2017, 23,5760.
48 A. Hungr´
ıa, M. Fern´
andez-Garc´
ıa, J. Anderson and
A. Mart´
ınez-Arias, J. Catal., 2005, 235, 262271.
49 D. R. Rainer, S. M. Vesecky, M. Koranne, W. S. Oh and
D. W. Goodman, J. Catal., 1997, 167, 234241.
50 W. Brown, R. Sharma, D. King and S. Haq, J. Phys. Chem.,
1996, 100, 1255912568.
51 A. Bogicevic and K. Hass, Surf. Sci., 2002, 506, L237L242.
52 N. Takagi, K. Ishimura, H. Miura, T. Shishido, R. Fukuda,
M. Ehara and S. Sakaki, ACS Omega, 2019, 4, 25962609.
53 R. Burch, S. Daniells and P. Hu, J. Chem. Phys., 2004, 121,
27372745.
8298 |Chem. Sci.,2019,10,82928298 This journal is © The Royal Society of Chemistry 2019
Chemical Science Edge Article
Article
With heterogeneous catalysts, chemical promotion takes place at their surfaces. Even in the case of single-atom alloys, where small quantities of a reactive metal are dispersed within the main host, it is assumed that both elements are exposed and available to bond with the reactants. Here, we show, on the basis of in situ X-ray absorption spectroscopy data, that in alloy catalysts made from Pt highly diluted in Cu the Pt atoms are located at the inner interface between the metal nanoparticles and the silica support instead. Kinetic experiments indicated that these catalysts still display better selectivity for the hydrogenation of unsaturated aldehydes to unsaturated alcohols than the pure metals. Density functional theory calculations corroborated the stability of Pt at the metal–support interface and explained the catalytic performance as being due to a remote lowering of the activation barrier for the dissociation of H2 at Cu sites by the internal Pt atoms.
Article
Single-atom alloys (SAAs) have attracted considerable attention as promising electrocatalysts in reactions central to energy conversion and chemical transformation. In contrast to monometallic nanocrystals and metal alloys, SAAs possess unique and intriguing physicochemical properties, positioning them as ideal model systems for studying structure–property relationships. However, the field is still in its early stages. In this Perspective, we first review and summarize rational synthesis methods and advanced characterization techniques for SAA nanoparticle catalysts. We then emphasize the extensive applications of SAAs in a range of electrocatalytic reactions, including fuel cell reactions, water splitting, and carbon dioxide and nitrate reductions. Finally, we provide insights into existing challenges and prospects associated with the controlled synthesis, characterization, and design of SAA catalysts.
Article
Considering the current energy scenario, it is of great importance to design and develop innovative, economically feasible electrocatalysts for the various energy applications. The high cost, and low availability of noble metal-based (Pd, Pt, Ru, etc.) electrocatalysts limit their widespread implementation of electrochemical reactions. Earth-abundant copper-based single-atom electrocatalysts (Cu-based SAEs) possess desired electronic, morphological, and physicochemical properties that have been extensively deployed for the energy applications. In the context of the progress of copper-based SAEs, herein we reviewed the notable advancement in fabrication and applications of Cu-based SAEs for the production of fuels, hydrocarbons, and ammonia. We also addressed the stability of developed electrocatalysts and active sites present in the structure of single-atom copper electrocatalysts. The challenges, and potential insights into the mechanism of action are also described including the ways to enhance the overall SAEs activities by tailoring the active site chemistry on the basis of computational studies and designing the advanced synthesis strategy.
Article
Full-text available
Low‐loaded platinum‐group single‐atom catalysts on CeO2 (M1/CeO2) were synthesized via high‐temperature atom trapping (AT) and tested for the NO+CO reaction under dry and wet conditions. The activity of these catalysts for NO+CO reaction follows the order Rh>Pd≈Ru>Pt>Ir. For Rh, Ru, and Pd single‐atom catalysts, the N2O byproduct is formed but not clearly observed in Ir and Pt cases, which may result from the higher reaction temperature (>200 °C) required for Pt and Ir catalysts. The presence of water can promote the activity of these M1/CeO2 catalysts for the NO+CO reaction. Under wet conditions, significant NH3 formation occurred during the reaction, which is due to the co‐existence of water‐gas‐shift reaction on these catalysts. Compared with Pt, Pd and Ir, the Rh and Ru single‐atom catalysts show higher selectivity to NH3 species, resulting from the hydride species on the surface. Among all tested catalysts, Ru1/CeO2 shows the highest production of ammonia and highest CO conversion due to excellent water‐gas‐shift activity, whereas Pd1/CeO2 shows lowest ammonia production. Rh1/CeO2 shows the best low temperature NO reduction activity among all tested catalysts.
Article
Full-text available
Density functional theory is used to compare the catalytic performance of PtPdRhFeCo(100) high entropy alloy (HEA) three‐way catalyst (TWC) to the conventional Pt(100) in the NO reduction step during NH3 production that supplies to passive NH3‐SCR. Stronger adsorption of NO on the HEA(100) surface is beneficial to capture NO. During adsorption, the catalyst surface acts as an electron donor while the adsorbate is the acceptor on both HEA(100) and Pt(100) systems. Herein, the reaction mechanism of NO reduction can be classified into two steps: 1) NO activation and 2) product formation. During NO activation, direct NO dissociation is the preferable pathway on both HEA(100) and Pt(100) surfaces with the same Ea, whereas HNO and NOH pathways on HEA(100) are suppressed. For NH3, N2, and N2O production on HEA(100) is found to be more difficult than on Pt(100). However, the thermodynamic driving force of all reactions on HEA(100) is more spontaneous than on Pt(100). Also, the rate‐determining step on HEA(100) is found to be NH3 formation different from the Pt(100), while difficult H diffusion on HEA(100) is the key factor that reduces NH3 production.
Article
Full-text available
Invited for this month's cover is the group of Dr. Katsutoshi Sato and Prof. Dr. Katsutoshi Nagaoka (Kyoto University) and collaborators at Oita and Kyushu Universities. The cover picture shows the proposed mechanism for automotive exhaust purification over a Pt–Co alloy nanoparticle catalyst with an extremely low Pt/Co molar ratio. In the catalyst, the isolated electron‐rich Pt atoms are present on the surface of the nanoparticles and play an important role in NOx capture and activation, which are important elementary steps in exhaust purification. Read the full text of the article at 10.1002/cplu.201800542.
Article
Full-text available
The development of Pd-based alloy catalysts for highly active and selective reduction of NO by CO was investigated. A survey of Pd-based bimetallic catalysts (PdM/Al2O3: M = Cu, In, Pb, Sn, and Zn) revealed that the PdIn/Al2O3 catalyst displayed excellent N2 selectivity even at low temperatures (100% at 200 °C). The catalytic activity of PdIn was further improved by substituting a part of In with Cu, where a Pd(In1-x Cu x ) pseudo-binary alloy structure was formed. The optimized catalyst, namely, Pd(In0.33Cu0.67)/Al2O3, facilitated the complete conversion of NO to N2 (100% yield) even at 200 °C and higher, which has never been achieved using metallic catalysts. The formation of the pseudo-binary alloy structure was confirmed by the combination of HAADF-STEM-EDS, EXAFS, and CO-FT-IR analyses. A detailed mechanistic study based on kinetic analysis, operando XAFS, and DFT calculations revealed the roles of In and Cu in the significant enhancement of catalytic performance: (1) N2O adsorption and decomposition (N2O → N2 + O) were drastically enhanced by In, thus resulting in high N2 selectivity; (2) CO oxidation was promoted by In, thus leading to enhanced low-temperature activity; and (3) Cu substitution improved NO adsorption and dissociation (NO → N + O), thus resulting in the promotion of high-temperature activity.
Article
Full-text available
Density functional theory calculations here elucidated that Cu38-catalyzed NO reduction by CO occurred not through NO dissociative adsorption but through NO dimerization. NO is adsorbed to two Cu atoms in a bridging manner. NO adsorption energy is much larger than that of CO. N-O bond cleavage of the adsorbed NO molecule needs a very large activation energy (δG°‡). On the other hand, dimerization of two NO molecules occurs on the Cu38 surface with small δG°‡ and very negative Gibbs reaction energy (δG°) to form ONNO species adsorbed to Cu38. Then, a CO molecule is adsorbed at the neighboring position to the ONNO species and reacts with the ONNO to induce N-O bond cleavage with small δG°‡ and very negative δG°, leading to the formation of N2O adsorbed on Cu38 and CO2 molecule in the gas phase. N2O dissociates from Cu38, and then it is readsorbed to Cu38 in the most stable adsorption structure. N-O bond cleavage of N2O easily occurs with small δG°‡ and significantly negative δG° to form the N2 molecule and the O atom adsorbed on Cu38. The O atom reacts with the CO molecule to afford CO2 and regenerate Cu38, which is rate-determining. N2O species was experimentally observed in Cu/I-Al2O3-catalyzed NO reduction by CO, which is consistent with this reaction mechanism. This mechanism differs from that proposed for the Rh catalyst, which occurs via N-O bond cleavage of the NO molecule. Electronic processes in the NO dimerization and the CO oxidation with the O atom adsorbed to Cu38 are discussed in terms of the charge-transfer interaction with Cu38 and Frontier orbital energy of Cu38.
Article
Full-text available
There is interest in minimizing or eliminating the use of Pt in catalysts by replacing it with more widely abundant and cost‐effective elements. The alloying of Pt with non‐noble metals is a potential strategy for reducing Pt use because interactions between Pt and non‐noble metals can modify the catalyst structure and electronic properties. Here, a γ‐Al2O3‐supported bimetallic catalyst [Pt(0.1)Co(1)/Al2O3] was prepared which contained 0.1 wt % Pt and 1 wt % Co and thus featured an extremely low Pt : Co ratio (<1 : 30 mol/mol). The Pt and Co in this catalyst formed alloy nanoparticles in which isolated electron‐rich Pt atoms were present on the nanoparticle surface. The activity of this Pt(0.1)Co(1)/Al2O3 catalyst for the purification of automotive exhaust was comparable to the activities of 0.3 and 0.5 wt % Pt/γ‐Al2O3 catalysts. Electron‐rich Pt and metallic Co promoted activation of NOx and oxidization of CO and hydrocarbons, respectively. This strategy of tuning the surrounding structure and electronic state of a noble metal by alloying it with an excess of a non‐noble metal will enable reduced noble metal use in catalysts for exhaust purification and other environmentally important reactions.
Article
Full-text available
Alloying provides a means by which to tune a metal catalyst's electronic structure and thus tailor its performance; however, mean-field behaviour in metals imposes limits. To access unprecedented catalytic behaviour, materials must exhibit emergent properties that are not simply interpolations of the constituent components' properties. Here we show an emergent electronic structure in single-atom alloys, whereby weak wavefunction mixing between minority and majority elements results in a free-atom-like electronic structure on the minority element. This unusual electronic structure alters the minority element's adsorption properties such that the bonding with adsorbates resembles the bonding in molecular metal complexes. We demonstrate this phenomenon with AgCu alloys, dilute in Cu, where the Cu d states are nearly unperturbed from their free-atom state. In situ electron spectroscopy demonstrates that this unusual electronic structure persists in reaction conditions and exhibits a 0.1 eV smaller activation barrier than bulk Cu in methanol reforming. Theory predicts that several other dilute alloys exhibit this phenomenon, which offers a design approach that may lead to alloys with unprecedented catalytic properties.
Article
The relationship between the oxidation state of bulk Cu and catalytic activity during the NO + CO and N2O + CO reactions were investigated, using commercially available Cu, Cu2O and CuO powders. Metallic Cu powder was found to catalyze both reactions in a stable manner without changing its oxidation state. At elevated temperatures, the catalytic activities of the Cu2O and CuO increased with time, in conjunction with reduction by CO to metallic Cu in both cases. The data confirm that metallic Cu exhibits much high activity and stability than either Cu2O or CuO for both the NO + CO and N2O + CO reactions. In addition, Cu2O species (denoted as Cu2O*) formed via oxidation of metallic Cu species obtained by reduction of CuO with CO at low temperature (<300 °C) showed catalytic activity during both reactions as a result of the high surface area and structural disorder of this material.
Article
Research on single‐atom catalysts (SACs), or atomically dispersed catalysts, has been quickly gaining momentum over the past few years. Although the unique electronic structure of singly dispersed atoms enables uncommon—sometimes exceptional—activities and selectivities for various catalytic applications, developing reliable and general procedures for preparing stable, active SACs in particular for applications under reductive conditions remains a major issue. Herein, the challenges associated with the synthesis of SACs are highlighted semiquantitatively and three stabilization techniques inspired by colloidal science including steric, ligand, and electrostatic stabilization are proposed. Some recent examples are discussed in detail to showcase the power of these strategies in the synthesis of stable SACs without compromising catalytic activity. The substantial further potential of steric, ligand, and electrostatic effects for developing SACs is emphasized. A perspective is given to point out opportunities and remaining obstacles, with special attention given to electrostatic stabilization where little is done so far. The stabilization strategies presented herein have a wide applicability in the synthesis of a series of new SACs with improved performances.
Article
Single-atom catalysis has arguably become the most active new frontier in heterogeneous catalysis. Aided by recent advances in practical synthetic methodologies, characterization techniques and computational modelling, we now have a large number of single-atom catalysts (SACs) that exhibit distinctive performances for a wide variety of chemical reactions. This Perspective summarizes recent experimental and computational efforts aimed at understanding the bonding in SACs and how this relates to catalytic performance. The examples described here illustrate the utility of SACs in a broad scope of industrially important reactions and highlight the advantages these catalysts have over those presently used. SACs have well-defined active centres, such that unique opportunities exist for the rational design of new catalysts with high activities, selectivities and stabilities. Indeed, given a certain practical application, we can often design a suitable SAC; thus, the field has developed very rapidly and afforded promising catalyst leads. Moreover, the control we have over certain SAC structures paves the way for designing base metal catalysts with the activities of noble metal catalysts. It appears that we are entering a new era of heterogeneous catalysis in which we have control over well-dispersed single-atom active sites whose properties we can readily tune.
Article
Silica supported and unsupported PdAu single atom alloys (SAAs) were investigated for the selective hydrogenation of 1-hexyne to hexenes under mild conditions. The catalysts were prepared by adding a trace amount of Pd (0.4 at. %) into the surface of pre-formed Au nanoparticles through a sequential reduction method. TEM and XRD analyses indicate the formation of PdAu nanoparticles and ATR-IR confirms the single atom dispersion of Pd in the Au matrix. In time-resolved batch reactor studies, we found that the Pd single atoms improved the hydrogenation activity of Au by nearly 10-fold but did not decrease the high selectivity to partial hydrogenation products. The enhanced reactivity is attributed to the Pd single atoms (isolated Pd atoms in the Au surface) facilitating molecular hydrogen dissociation leading to the availability of weakly bound atomic hydrogen on the otherwise inert gold surface. Higher than 85% selectivity to hexenes was observed, which is significantly greater than that of monometallic Pd catalysts. Model catalyst studies were conducted to investigate the formation and reactivity of the Pd/Au(111) SAAs. Scanning tunneling microscopy of Pd-Au(111) surfaces confirms the formation of PdAu single atom alloys at low Pd coverage with the Pd preferentially located in the vicinity of the herringbone elbows of the reconstructed Au(111) surfaces. Temperature-programmed desorption experiments confirm that single Pd atom sites dissociate hydrogen and bind both CO and H atoms more weakly as compared to extended Pd surfaces.