Bottom-Up Precise Synthesis of Stable Platinum Dimers On Graphene
Bottom-Up Precise Synthesis of Stable Platinum Dimers On Graphene
Supported metal clusters containing only a few atoms are of great interest. Progress has been
made in synthesis of metal single-atom catalysts. However, precise synthesis of metal dimers
on high-surface area support remains a grand challenge. Here, we show that Pt2 dimers can
be fabricated with a bottom–up approach on graphene using atomic layer deposition, through
proper nucleation sites creation, Pt1 single-atom deposition and attaching a secondary Pt
atom selectively on the preliminary one. Scanning transmission electron microscopy, x-ray
absorption spectroscopy, and theoretical calculations suggest that the Pt2 dimers are likely in
the oxidized form of Pt2Ox. In hydrolytic dehydrogenation of ammonia borane, Pt2
dimers exhibit a high specific rate of 2800 molH2 molPt−1 min−1 at room temperature, ~17- and
45-fold higher than graphene supported Pt single atoms and nanoparticles, respectively.
These findings open an avenue to bottom–up fabrication of supported atomically precise
ultrafine metal clusters for practical applications.
1 Department of Chemical Physics, University of Science and Technology of China, Hefei, Anhui 230026, China. 2 Hefei National Laboratory for Physical
Sciences at the Microscale, University of Science and Technology of China, Hefei, Anhui 230026, China. 3 CAS Key Laboratory of Materials for Energy
Conversion, University of Science and Technology of China, Hefei, Anhui 230026, China. 4 National Synchrotron Radiation Laboratory, University of Science
and Technology of China, Hefei, Anhui 230029, China. 5 Collaborative Innovation Center of Chemistry for Energy Materials (iChEM), University of Science
and Technology of China, Hefei 230026, China. Huan Yan and Yue Lin contributed equally to this work. Correspondence and requests for materials should
be addressed to S.W. (email: [email protected]) or to J.L. (email: [email protected])
S
upported metal catalysts are among the most important anchor sites atom-by-atom in a sequential manner using Pt ALD.
categories of heterogeneous catalysts in many reactions The dominant presence of isolated Pt1 single atoms and Pt2
including chemical upgrading, automobile exhaust treat- dimers in the corresponding samples were confirmed by both
ment, Fischer-Tropsch synthesis, biomass conversions, and many aberration-corrected high-angle annular dark-field scanning
other processes1–7. Decreasing metal particle size is desirable for transmission electron microscopy (HAADF-STEM) and X-ray
improving metal utilization, since catalytic reactions take place on absorption fine structure spectroscopy (XAFS). Their structures
the surface of metal nanoparticles (NPs). When a metal cluster were determined through a combination of density function
contains only a few metal atoms, it could have a discrete energy theory (DFT) calculations and XAFS spectra simulations. In
band structure, tightly correlated with the number of metal hydrolysis of ammonia borane (AB) for hydrogen generation,
atoms. Changing one atom in the ultrafine cluster might largely graphene supported Pt2 dimers (Pt2/graphene) exhibited a strik-
alter the electronic structure and drastically change its catalytic ing activity, which is ~17- and 45-fold higher than that of gra-
properties. Such atom-dependent catalytic behaviors have been phene supported Pt1 single atoms and Pt NPs, respectively.
successfully demonstrated by the model catalysts of mass-selected Compared to Pt1 single atoms and Pt NPs, the decreased
metal clusters, which were fabricated by soft landing of mass- adsorption energies of both AB and H2 molecules on Pt2 dimers
selected ions from their physical vapor under ultrahigh vacuum are likely the major reason for the high activity. More impor-
conditions8–14. However, such complicated approach is only tantly, the Pt2 dimers were stable under the current reaction
limited to model catalyst studies and is not applicable to high- condition and in the inert environment at below 300 °C.
surface area supports for practical applications.
Recently, synthesis of supported metal single-atom catalysts
(SACs) has been extensively explored and a number of successful Results
examples have been demonstrated15–23. Nonetheless, synthesis of Synthesis and morphology of Pt1/graphene and Pt2/graphene.
atomically precise ultrafine metal clusters, such as dimers, on Based on our recent strategy22, the nucleation sites of isolated
high-surface area supports, remains a grand challenge. The phenols or phenol–carbonyl pairs suggested by Shenoyl et al40.
decisive limitation is the lack of precise control over the aggre- for Pt ALD were first created on pristine graphene nanosheets
gation process, which often causes metal NPs formation with a through acid oxidation followed by high-temperature thermal
broad size distribution. Protecting metal clusters with a strong reduction, as illustrated in the schematic model in Fig. 1. The
ligand can certainly inhibit metal aggregation to a large extent, such reduced graphene oxide support was defect-rich multilayered
as in the case of thiolate-protected Au magic clusters24. However, graphene films with a thickness of about a few nanometers
these strong protective ligands typically poison the metal clusters, (Supplementary Fig. 1). It had a surface area of 570 m2/g. Next, Pt
and decrease their catalytic activities considerably25–30. Alter- ALD was performed on the graphene support by alternately
natively, Gates et al. demonstrated that precisely defined iridium exposing trimethyl(methylcyclopentadienyl)-platinum(IV)
and rhodium clusters were achieved by grafting the corresponding (MeCpPtMe3) and molecular O2 at 250 °C. The self-limiting
carbonyl complexes with a specific number of metal atoms onto surface reactions between MeCpPtMe3 and the support ensure
oxide supports31–33. But the success is limited. As a consequence, a nucleation of one MeCpPtMe3 molecule on one phenol-related
general bottom–up approach to synthesize atomically precise metal nucleation site during the saturated MeCpPtMe3 exposure
clusters on high-surface area supports is still missing. Atomic layer (Supplementary Fig. 2). It should be noted that Pt nucleation on
deposition (ALD) relies on two sequential self-limiting surface graphene defect sites, such as edges and line defects, was inhibited
reactions at the molecular level, which are separated by inert gas at 250 °C, although it is possible at 300 °C (Supplementary
purging34–36. This unique character makes ALD possible to Table 1)41. A similar temperature effect on inhibiting metal ALD
bottom–up construct catalytic materials on a high-surface area on oxides was also observed by Elam et al42. Next, the ligands
substrate uniformly and precisely37–39. were removed through a combustion reaction during the O2
Here, we show that Pt2 dimers can be bottom–up fabricated on exposure step43–45 and Pt1 single atoms are formed (denoted as
a graphene support by depositing Pt on phenol-related oxygen Pt1/graphene).
MeCpPtMe3
O2
O3
Self-limiting
Pt ALD
2nd cycle
Pt dimers Pt single atoms
Fig. 1 Schematic illustration of bottom–up synthesis of dimeric Pt2/graphene catalysts. Controlled creation of isolated anchor sites on pristine graphene;
one cycle of Pt ALD on the anchor sites for Pt single atoms formation by alternately exposing MeCpPtMe3 and molecular O2 at 250 °C; second cycle of Pt
ALD on the Pt1/Graphene to selectively deposit the secondary Pt atoms on the preliminary ones for Pt2 dimers formation at 150 °C. The balls in cyan,
white, red, and blue represent carbon, hydrogen, oxygen, and platinum while the ball in gray represents carbon atoms in the graphene support
Next, the formed Pt1 single atoms was further utilized as ALD reactor quickly to perform the second cycle (Pt-MeCpPtMe/
nucleation sites for anchoring the secondary MeCpPtMe3 graphene) at 150 °C. In nine independent trials, the ICP-AES
molecule in the following cycle. Again, the steric hindrance results showed that the ratio of the Pt loadings of Pt-MeCpPtMe/
between MeCpPtMe3 molecules restricts chemisorbing one graphene to those of the corresponding MeCpPtMe/graphene
MeCpPtMe3 molecule only on one isolated Pt1 atom. However, were all very close to one (Fig. 2h). Therefore, there was no
we noticed that a considerable amount of Pt NPs were additional Pt deposited on MeCpPtMe/Graphene, reflecting the
formed after two successive cycles of Pt ALD at 250 °C (denoted saturated self-limiting reaction character of ALD34–36. Second, we
as 2cPt/graphene, Supplementary Fig. 3). Therefore, the deposi- found that exposure of pristine graphene to O2 at 250 °C did not
tion temperature was decreased to 150 °C for the second ALD cause any detectable Pt deposition either (Supplementary Table 1).
cycle to avoid any metal aggregation. Meanwhile, ozone (O3), a This is very important to ensure that the oxygen pulse at 250 °C
stronger oxidizing regent was utilized to remove the ligand in the first ALD cycle did not create any additional nucleation
efficiently to form Pt2 dimers (Pt2/graphene) (Fig. 1)46. sites. Taken together, the Pt1 single atoms formed on graphene,
The formed Pt1 single atoms and Pt2 dimers are expected to be confirmed by HAADF-STEM in Fig. 2a–c, are the only nucleation
in the oxidized forms, since they were exposed to O2 and O3 sites for the following ALD cycle. It is worthy to note that Pt1
during synthesis, respectively. single atoms well isolated from each other could be crucial to
Aberration-corrected HAADF-STEM measurements were car- make all the Pt1 single atoms accessible for chemisorbing the
ried out to investigate the morphologies of the single-atom second MeCpPtMe3 molecule in the second ALD cycle without
Pt1/graphene and dimeric Pt2/graphene catalysts. Compared to steric hindrance.
the naked graphene (Supplementary Fig. 1), HAADF-STEM During the second ALD cycle, one MeCpPtMe3 molecule
images illustrated that one cycle of Pt ALD on the graphene anchors on one Pt1 atom in the Pt1/graphene sample due to the
support at 250 °C resulted in a formation of atomically dispersed steric hindrance effect, which doubles the Pt loading.
Pt1 atoms without presence of any visible clusters or NPs This was confirmed by the ratios of two for the Pt loadings of
(Fig. 2a–c, and Supplementary Fig. 4). These Pt1 single atoms Pt2/Graphene to Pt1/graphene in nine independent trials (Fig. 2i
were well isolated from each other with a distance >2 nm in and Supplementary Table 2), hence providing strong evidence of
average, which is significantly larger than the effective diameter of the formation of Pt2 dimers. The Pt1 single atoms observed in
the MeCpPtMe3 molecule of ~0.96 nm47, confirming the steric Fig. 2e and Supplementary Figs. 6–9 were likely formed by
effect during synthesis. Similar to our recent study of Pd1 single- uncoupling of Pt2 dimers under the high flux electron beam
atom growth on graphene22, we found that complete removal of during STEM measurements49. On the other hand, once Pt NPs
other oxygen-contained functional groups, such as carboxyl were formed during ALD, the ratio of Pt loading of the two-ALD
groups, from graphene by carefully tuning the reduction cycle sample to the one-cycle sample was apparently off from the
temperature and time, is the key to eliminate any Pt clusters or stoichiometric value of two (Supplementary Table 2).
NPs formation (Supplementary Fig. 5). These findings suggest Performing an additional cycle on Pt2/graphene to form Pt3
that metal atoms anchored on carboxyl groups have a general trimers might be possible. However, we noticed that the O3 in the
weak interaction with the graphene support, thus aggregate second cycle can create additional nucleation sites on graphene.
aggressively to larger metal NPs under the ALD conditions, in As a consequence, selective deposition was not achieved for the
line with literature48. third cycle and resulted in a mixture of Pt1, Pt2 and Pt3.
After performing another cycle of Pt ALD on Pt1/graphene at Therefore, we mainly focused on the Pt2 dimers in this work.
150 °C (Pt2/graphene), Pt2 dimers were dominantly formed along
with a certain number of Pt1 single atoms (Fig. 2d–f, and
Supplementary Fig. 6), where neither Pt clusters nor NPs were XAFS characterization and DFT calculations. Figure 3a shows
observed. Very interestingly, we noticed that Pt2 dimers the X-ray absorption near-edge structure (XANES) spectra of
frequently rotated by specified angles of 30, 60, and 90° under MeCpPtMe/graphene, Pt1/graphene, and Pt2/graphene at the Pt
the electron beam during STEM measurements and then split L3-edge, along with Pt foil, PtO2, and MeCpPtMe3 as references.
into two isolated Pt1 atoms (see more details in Supplementary Evidently, the XANES white line peaks of these three samples
Figs. 7–9). Such characteristic rotations might be related with the (11,567 eV) located at between Pt foil and PtO2, indicating that
geometry of the graphene support and the size of carbon defect by the Pt in MeCpPtMe/graphene, Pt1 single atoms, and Pt2 dimers
considering the aforementioned Pt2 dimer structure (Supplemen- were all in a similar oxidation state between Pt0 and Pt4+. The
tary Fig. 10). This observation provides strong evidence of the MeCpPtMe3 reference sample exhibits two well-resolved peaks at
presence of Pt2 dimers rather than the projection coincidence of 1.62 and 1.99 Å in the Fourier transformed (FT) k3χ(k) curve in
two isolated Pt1 atoms at different Z positions. Statistical analysis the real-space (R-space) (Fig. 3b), assigned to the shorter Pt–C
of more than 80 pairs of Pt2 dimers showed a Pt–Pt distance of bonds (1.99−2.14 Å) in the three Pt–Me groups and longer Pt–C
0.30 ± 0.02 nm for Pt2 dimers (Fig. 2g), which is significantly bonds (2.26−2.36 Å) in the Pt–MeCp group, respectively (Sup-
longer than the Pt–Pt bond in Pt bulk. This indicates that the Pt2 plementary Fig. 11a)47,50. Apparently, the two split peaks in the
dimers are in the oxidized form as expected. MeCpPtMe/graphene curve suggests that the MeCp group stayed
To step-wise elucidate the selective deposition of secondary Pt on Pt in this sample. This observation is in line with both the
atom onto the preliminary ones for the formation of Pt2 dimers in experimental result of the formation of MeCpPtMe2 surface
the second ALD cycle, a set of control experiments were further species on oxides43,51 and theoretical calculations where
performed using inductively coupled plasma-atomic emission MeCpPtMe3 on epoxydated and hydroxylated graphene surfaces
spectroscopy (ICP-AES). First, the influence of the Pt precursor liberates either one or two methyl groups depending on the
ligand on the second Pt ALD cycle was examined. In this case, available surface groups52.
half cycle of Pt ALD was executed on the graphene support by The first shell FT peak in the Pt1/graphene spectrum had a
performing the MeCpPtMe3 pulse step only (denoted as higher intensity and slightly shifted to 1.65 Å, while the peak at
MeCpPtMe/graphene) at 250 °C. After that, the ALD reactor 1.99 Å disappeared. Clearly, the MeCp ligand was combusted off
was cooled to near room temperature and half amount of the after the O2 exposure step at 250 °C. The first shell peak is
MeCpPtMe/graphene sample was taken out of reactor for ICP- assigned to Pt−O and/or Pt−C coordinations. Similar to
AES analysis; while the rest of the sample was put back to the MeCpPtMe/graphene, a very weak peak at 2.4 Å was visible in
a b c
d e f
g h i
D = 0.30 ± 0.02 nm 1.01 ± 0.03 1.98 ± 0.07
2.5 2.5
30
Ratio of Pt loadings
Ratio of Pt loadings
2.0 2.0
Frequency (%)
20 1.5 1.5
1.0 1.0
10
0.5 0.5
0 0.0 0.0
0.0 0.2 0.4 0.6 0.8 0 2 4 6 8 10 0 2 4 6 8 10
Pt–Pt distance (nm) Trials Trials
Fig. 2 Morphology of the single-atom Pt1/graphene and dimeric Pt2/graphene catalysts. Aberration-corrected HAADF-STEM images of Pt1/graphene
(a–c) and dimeric Pt2/graphene (d, e). Scale bars, 20 nm (a, d), 2 nm (b, e), and 1 nm (c, f). Pt single atoms in b and c and dimers in e and f are
highlighted by white and yellow circles, respectively. g Statistical Pt–Pt distance in the observed Pt2 dimers. h The ratio of Pt loading of in Pt-MeCpPtMe/
graphene to that in MeCpPtMe/graphene, and i the ratio of Pt loading in Pt2/graphene to that in Pt1/graphene in nine independent trials determined by
ICP-AES
the spectrum of Pt1/graphene. However, this peak is significantly support containing a carbon vacancy along with either isolated
different from the Pt–Pt coordination peak (at ~2.62 Å) in Pt foil, phenol group or phenol–carbonyl pairs40 was employed as the
thus assigned to the second nearest C/O neighbors of Pt. This reduced graphene oxide surface. The structural models optimized
suggests the absence of Pt NPs in Pt1/graphene, consistent by DFT calculations were further examined by EXAFS
with our STEM observation (Fig. 2a–c and Supplementary Fig. 4). simulations.
The dimeric Pt2/graphene sample showed a similar FT curve Regarding the previous work52, the structures of MeCpPtMe2
with Pt1/graphene, implying a similar local C/O coordinations in and MeCpPtMe were both considered for the MeCpPtMe/
these two samples. In the Pt2/graphene spectrum, there was no graphene sample (Supplementary Fig. 11). Compared to the
discernible peak for the Pt–Pt coordination, suggesting the Pt2 MeCpPtMe3 molecule, the five Pt–C bonds in the Pt–MeCp
dimers are in the oxidized form after the ozone exposure step at group in these two structures both changed significantly.
150 °C. According to the EXAFS simulations for these two structures,
Considering the difficulties in discriminating the C/O neigh- MeCpPtMe bonded to the graphene support through two
bors by EXAFS fittings, we resorted to the combination of interfacial O atoms might be the promising structure for the
DFT calculations with EXAFS simulations to determine the MeCpPtMe/graphene sample (Fig. 3c and Supplementary
optimized structures of these three samples. Here, a graphene Figs. 11b and 12). Compared to the spectrum of MeCpPtMe3,
a c
Pt L3-edge MeCpPtMe/graphene
11568 eV Simulation
Pt foil
11566 eV
MeCpPtMe-
/graphene
11567 eV
Pt1/graphene
A 0 1 2 3 4 5 6
Pt2/graphene
R (Å)
d
Pt1/graphene
Simulation
b
Pt-O shell Pt-Pt shell
1.65 Å 2.62 Å
1.99 Å 0 1 2 3 4 5 6
Pt foil*1/3 R (Å)
Pt-C shell
e
|FT(k 3χ(k ))| (a.u.)
1.62 Å PtO2*1/2
MeCpPtMe3 Pt2/graphene
Simulation
|FT(k 3χ(k ))| (a.u.)
1.65 Å
3Å
MeCpPtMe-
1.60 Å
3.0
/graphene
Pt1/graphene
Pt2/graphene
0 1 2 3 4 5 6 0 1 2 3 4 5 6
R (Å) R (Å)
Fig. 3 XAFS structural characterization and spectra simulations. a The XANE spectra and b the K2-weighted Fourier transform spectra of MeCpPtMe/
graphene, Pt1/graphene, and Pt2/graphene at the Pt L3-edge. The reference samples of Pt foil, PtO2, and MeCpPtMe3 are also shown for comparison.
Comparison of the EXAFS simulations based on the corresponding DFT calculated structural models (insets) with the experimental EXAFS spectra of
MeCpPtMe/graphene (c), Pt1/graphene (d) and Pt2/graphene (e). The ball in gray, white, red, and dark blue represent carbon, hydrogen, oxygen, and
platinum, respectively
the remarkably attenuated peak at 1.99 Å in the MeCpPtMe/ exposure step. Taking this information into account, a Pt2O6
graphene spectrum is due to the considerable distortion of the chain structure with O atoms alternating between the terminal
MeCp group. and bridge positions was constructed (inset in Fig. 3e). After
When oxygen combusts off the ligand, additional oxygen optimization, our calculations show that the Pt–Pt bond distance
chemisorbs on the Pt1 atom in the Pt1/graphene sample53. in the Pt2O6 chain is 3.03 Å (the inset of Fig. 3e), consistent with
Indeed, Pt1 atom with one chemisorbed O2 molecule at the the experimental results very well (Fig. 2g). The lengths of the
terminal position (the Pt–O bond distance: 2.00 Å) and two O Pt–O bonds in the Pt2O6 chain are very close to each other,
atoms at the interface (the Pt–O bond distance: 2.02 Å) produces ~1.93–2.03 Å. Moreover, XAFS spectrum simulation for this
an EXAFS spectrum in good agreement with the experimental Pt2O6 chain structure also agrees very well with the experimental
result (Fig. 3d and Supplementary Fig. 12). On the contrary, the result (Fig. 3e and Supplementary Fig. 12). This chain structure is
Pt1 atom with one O and one C atom at the interface generates found to be similar to the suggested structure models for PtxOy (x
split FT peaks in the first shell, in contrast with the experimental = 1–3) clusters by Schneider et al. previously54. Interestingly, we
results (Supplementary Fig. 13). Nonetheless, this structure might also noticed that the tilted angle of the Pt2O6 chain could vary
not be completely ruled out. largely from ~8 to ~50°, depending on both the size of carbon
During the second Pt ALD cycle, a secondary Pt atom anchors vacancies and the configurations of two interfacial O atoms
on the preliminary one and then becomes oxidized during the O3 (Supplementary Fig. 14). The largely varied angles tilted from the
a 25
The activity of 2cPt/graphene, synthesized by two successive
Pt1/graphene
cycles of Pt ALD on graphene at 250 °C, was rather close to
Pt2/graphene
Pt1/graphene, by generating 7.5 mL H2 in 10.5 min. Obviously,
Pt2/graphene and 2cPt/graphene were distinctly different in
Volume of H2 generation (mL)
20
2cPt/graphene structure. As a comparison, the activities of the Pt NP catalysts of
Pt/graphene-WI Pt/graphene-WI, Pt/carbon, and Pt/SiO2, as well as the
15 Pt/carbon commercial PtO2 were tested. The Pt/graphene-WI catalyst with
Pt/SiO2 a Pt particle size of 1.8 ± 0.5 nm (Supplementary Fig. 15 and
PtO2 Table 3), showed a very poor activity of 6.6 mL H2 release in
10 15.2 min. The commercial Pt/carbon catalyst with a Pt particle
size of 2.3 ± 0.7 nm (Supplementary Fig. 16 and Table 3)
was considerably better, generating 23.4 mL H2 in 9.7 min. The
5 Pt/SiO2 ALD catalyst with a Pt particle size of 1.9 ± 0.3 nm
(Supplementary Fig. 17 and Table 3) was also very active,
releasing 23.4 mL H2 in 6.8 min. The commercial PtO2 powder
0 (Supplementary Fig. 18) generated ~ 21 mL H2 in 15 min. In this
case, a reduction of PtO2 into Pt occurred during the reaction, in
0 3 6 9 12 15 18
Time (min) line with literature56.
b 3000
The specific rates of these samples were calculated based on the
2800 Pt contents. The rates were 160 and 110 molH2 MolPt−1 min−1, for
Pt1/graphene and 2cPt/graphene, respectively (Fig. 4b and
Specific rate (molH2 molPt–1 min–1)
2500 Supplementary Fig. 19). For the Pt NP samples, the rates were
62, 380, and 440 molH2 MolPt−1 min−1, for Pt/graphene-WI, Pt/
2000
carbon, and Pt/SiO2, respectively, close to the values for Pt
catalysts reported in the literature (Supplementary Table 4).
The rate of PtO2 was 197 molH2 MolPt−1 min−1. Obviously,
1500 hydrolytic dehydrogenation of AB on Pt catalysts is a structure
sensitive reaction, the size, and electronic properties of Pt NPs
1000 might both play important roles57,58. In sharp contrast with the
above samples, the Pt2 dimers exhibited the highest rate of
2800 molH2 MolPt−1 min−1 ever reported in literature, which
500 380 440 was ~17 and 45 times higher than the corresponding single-
160 110
197 atom Pt1/graphene and Pt/graphene-WI samples, respectively.
62
0 When the mole ratio of Pt to the AB substrate was increased,
ne ne ne WI arbon Pt2/graphene could preserve the high specific rate to a large
phe phe phe ne- SiO 2 PtO 2
/1 gra /gra t / gra phe Pt/
c Pt/ extent (Supplementary Fig. 20). Note that all the Pt samples
t t P ra
produced a similar product of BO2− in the spent reaction
P P 2
2c g
Pt/
a 25
b 25
c
15 15 2000
1670
10
10
1037
5 1000
5 Pt2/graphene
Pt2/graphene-300C
0
Pt2/graphene-400C
1st 2nd 3th 4th 5th 0
0
0 1 2 3 4 5 6 7 0 1 2 3 Fresh 300°C 400°C
Time (min) Time (min)
d e f
g h i
Fig. 6 Stability of the dimeric Pt2/graphene catalyst. a Five recycles in hydrolytic dehydrogenation of AB at room temperature over the dimeric Pt2/
graphene catalyst by adding additional 0.325 mmol of pure AB into the reaction flask after each run. b Plots of time vs volume of hydrogen gas generated
from AB hydrolysis and c the corresponding specific rates at room temperature over the dimeric Pt2/graphene catalysts after different pretreatments: as-
prepared, annealing in helium at 300 and 400 °C for 1 h, respectively. d, g Representative HAADF-STEM images of the used Pt2/graphene catalyst after
the recyclability test, scale bars, 20 nm (d), 1 nm (g). e, h Representative HAADF-STEM images of the Pt2/graphene catalyst after annealing in helium at
300 °C for 1 h, scale bars, 10 nm (e), 2 nm (h). f, i Representative HAADF-STEM images of the Pt2/graphene catalyst after annealing in helium at 400 °C
for 1 h, scale bars, 10 nm (f), 2 nm (i). Pt2 dimers in g–i are highlighted by yellow circles
Ltd., Chinese Academy of Sciences. All materials were used as received without Synthesis of Pt1/graphene. Pt ALD was carried out on a viscous flow reactor
further purification. (GEMSTAR-6 Benchtop ALD, Arradiance) by alternatively exposing to
MeCpPtMe3 precursor and O2 (99.999%) at 250 oC46,63,64. Ultrahigh purity N2
(99.999%) was used as the carrier gas at a flow rate of 200 mL/min. The Pt pre-
Preparation of reduced graphene oxide. Pristine graphene nanosheet was first cursor was heated to 65 °C to get a sufficient vapor pressure. The reactor inlets were
oxidized to graphene oxide according to the procedure described previously62. In held at 110 °C to avoid any precursor condensation. The timing sequence was 90,
brief, 0.6 g graphene nanosheet and 0.3 g sodium nitrate was sequentially added 120, 60, and 120 sec for the MeCpPtMe3 exposure, N2 purge, O2 exposure, and N2
into concentrated sulfuric acid (H2SO4, 15 mL) and stirred at room temperature for purge, respectively (90-120-60-120).
22 h. Then the mixture was cooled down to 0 °C to add 1.8 g potassium per-
manganate (KMnO4). After stirring at room temperature and 35 °C for 2 and 3 h,
respectively, the mixture was heated to 98 °C and kept at this temperature for
Synthesis of Pt2/graphene. The second Pt ALD cycle was performed on the Pt1/
another 30 min. Next, it was cooled down to 40 °C, and 90 mL of water and 7.5 mL
graphene SAC at 150 °C. Here, O3 was used as the oxidant to make sure that the
of hydrogen peroxide (H2O2, 30%) were slowly added into the mixture. After that
precursor ligand can be fully removed65. The timing sequence was 90-120-60-120.
the precipitate was filtered out by washing with HCl (5%) and ultrapure water, it
was dried in a vacuum oven at 45 °C overnight. The dry material was grinded to
obtain graphene oxide powder. Finally, the reduced graphene oxide was obtained
by thermal deoxygenation of graphene oxide powder at 1050 °C for 2 min under Synthesis of 2cPt/graphene. Two consecutive cycles of Pt ALD was also per-
helium at a flow rate of 50 mL/min. formed on the reduced graphene oxide at 250 °C using the same timing sequence.
Synthesis of Pt/SiO2. Pt ALD was performed on the silica gel support for one catalysts was adjusted to keep the same amount of Pt with Pt2/graphene. Typically,
cycle at 250 °C using the same timing sequence. 5 mL aqueous AB solution (6.5 × 10−2 M) were introduced into the glass container
via a syringe. For the commercial PtO2 catalyst, the mole ratio of Pt to AB was kept
as the same with other samples, and the result was normalized to other samples
Synthesis of Pt/graphene-WI. A Pt/graphene NP catalyst was synthesized by the
based the amount of Pt. The AB solution and the catalyst were well-mixed by using
wetness impregnation method (Pt/graphene-WI). In this case, 100 mg graphene
a magnetic stirrer at a speed of 800 r/min to eliminate any mass-transfer issue. The
support was slowly added into a 1.93 × 10−2 M H2PtCl6 aqueous solution (0.9 mL).
generated volume of H2 was measured by a water-filled gas burette, where the
Then, the mixture was stirred for 30 min, and then dried in air at room tem-
volume of water discharged was converted into the volume of hydrogen
perature for 12 h. The dried material was first calcined in air at 120 °C for 12 h, and
generated69.
then reduced in 10% H2 in argon at 300 °C for another 2 h to get the Pt/graphene-
The specific rates (r) of these catalysts were calculated according to the Eq. (2):
WI catalyst.
nH2
r¼ ð2Þ
Catalyst characterization. The Pt loadings in these samples were determined by nPt ´ t
ICP-AES measurements; therein all samples were dissolved in hot fresh aqua regia. Here nH2 is the mole of generated H2, while nPt is the total mole of Pt in the
The BET surface area was measured on a Micromeritics ASAP 2020 system. Raman sample. t is the reaction time in min.
spectra were recorded on a LabRAM HR Raman spectrometer with a 514 nm Ar
laser in backscattering geometry. Aberration-corrected HAADF-STEM measure-
ments were taken on a JEM-ARM200F instrument (University of Science and Data availability. All the relevant data are available from the authors upon
Technology of China) at 200 keV. XAFS measurements at the Pt L3-edge request.
(11,564 eV) were performed in the transmission mode with the Si (111) mono-
chromator at the BL14W1 beamline of the Shanghai Synchrotron Radiation Received: 21 February 2017 Accepted: 4 September 2017
Facility (SSRF), China. The storage ring of SSRF worked at 3.5 GeV with a max-
imum current of 210 mA.
XAFS data analysis and simulation. The acquired EXAFS data were processed
according to the standard procedures using the ATHENA module implemented in
the IFEFFIT software packages61.The EXAFS oscillation functions χ(k) were References
obtained by subtracting the post-edge background from the overall absorption 1. Bartholomew, C. H. & Farrauto, R. J. Fundamentals of Industrial Catalytic
spectra and then normalized with respect to the edge-jump step. The Rbkg value of
Processes. 2nd edn, (Wiley, 2006).
1.0 was used for all samples. Subsequently, k3-weighted χ(k) fucntions in the k
2. Funabiki, M., Yamada, T. & Kayano, K. Auto exhaust catalysts. Catal. Today 10,
range of 2.2–13.5 Å−1 were FT to the R space by using a Hanning window of dk =
33–43 (1991).
3.0 Å−1.
3. Guo, Z. et al. Recent advances in heterogeneous selective oxidation catalysis for
EXAFS simulations were performed with the FEFF8.4 code62 using the
structural models suggested by DFT calculations. The simulated EXAFS χ(k) sustainable chemistry. Chem. Soc. Rev. 43, 3480–3524 (2014).
functions were also k3-weighted and FT into the R-space by using the same k range 4. Shanks, B. H. Conversion of biorenewable feedstocks: new challenges in
of 2.2–13.5 Å−1 as that in the experimental data. During simulations, the heterogeneous catalysis. Ind. Eng. Chem. Res. 49, 10212–10217 (2010).
coordination numbers were set to the values of the model structures generated 5. Khodakov, A. Y., Chu, W. & Fongarland, P. Advances in the development of
by DFT calculations. The amplitude reduction factor S02 was fixed at the value novel cobalt Fischer-Tropsch catalysts for synthesis of long-chain hydrocarbons
of 0.86 which was determined by fitting the reference metal Pt foil. The and clean fuels. Chem. Rev. 107, 1692–1744 (2007).
Debye–Waller factors for the nearest Pt−O/C and Pt−Pt pairs were set at the 6. Alonso, D. M., Wettstein, S. G. & Dumesic, J. A. Bimetallic catalysts for
typical values of 0.0030 and 0.0065 Å2 determined from the fittings of PtO2 upgrading of biomass to fuels and chemicals. Chem. Soc. Rev. 41, 8075–8098
and Pt foil references, respectively, and they were set at 0.008 Å2 for all the (2012).
other distant paths which contributed barely discernible signals as seen from the 7. Calle-Vallejo, F., Koper, M. T. M. & Bandarenka, A. S. Tailoring the catalytic
experimental data in the R-space. To further improve the match between the activity of electrodes with monolayer amounts of foreign metals. Chem. Soc.
simulation and the experimental data, for the MeCpPtMe/graphene sample, the Rev. 42, 5210–5230 (2013).
two nearest Pt–C and Pt–O interatomic distances were optimized to 2.00 and 8. Kaden, W. E., Wu, T. P., Kunkel, W. A. & Anderson, S. L. Electronic structure
2.02 Å, respectively, both of which are within ~ 2% error level as compared to controls reactivity of size-selected Pd clusters adsorbed on TiO2 surfaces.
the optimized structure by DFT calculations (2.05 and 2.06 Å, respectively). Science 326, 826–829 (2009).
For the other two samples, the simulated EXAFS spectra based on the DFT- 9. Palmer, R. E., Pratontep, S. & Boyen, H. G. Nanostructured surfaces from size-
generated structures match well with the experimental data, thus no further selected clusters. Nat. Mater. 2, 443–448 (2003).
structure optimization was performed during EXAFS simulations. 10. Li, Z. Y. et al. Three-dimensional atomic-scale structure of size-selected gold
nanoclusters. Nature 451, 46–48 (2008).
DFT Calculations. All spin-polarized calculations were performed by using the 11. Abbet, S. et al. Acetylene cyclotrimerization on supported size-selected Pd-n
DFT method. The DFT Semi-core Pseudopotential method66 with a double clusters (1<=n<=30): one atom is enough. J. Am. Chem. Soc. 122, 3453–3457
numerical basis set together with polarization functions (DNP) were adopted to (2000).
form the Perdew-Burke-Ernzerhof (PBE) exchange-correlation functional within 12. Yoon, B. et al. Charging effects on bonding and catalyzed oxidation of CO on
the generalized gradient approximation67, implemented in DMol3 package (DMol3 Au-8 clusters on MgO. Science 307, 403–407 (2005).
is a density functional theory quantum mechanical package available from Accelrys 13. Nesselberger, M. et al. The effect of particle proximity on the oxygen reduction
Software Inc.)68. A DFT-D semi-empirical correction with Tkatchenko-Scheffler rate of size-selected platinum clusters. Nat. Mater. 12, 919–924 (2013).
(TS) method is applied with the PBE functional to account for the dispersion 14. Vajda, S. et al. Subnanometre platinum clusters as highly active and selective
interaction. Conductor-like screening model (COSMO) with a dielectric constant catalysts for the oxidative dehydrogenation of propane. Nat. Mater. 8, 213–216
of 78.54 is adopted to consider the water solvent effect regarding the adsorption of (2009).
molecule and fragment in AB hydrolysis. A smearing of 0.001 Ha to the orbital 15. Yang, X. F. et al. Single-atom catalysts: a new frontier in heterogeneous
occupation is applied to achieve electronic convergence. The real-space global catalysis. Acc. Chem. Res. 46, 1740–1748 (2013).
cutoff radius is set to be 4.5 Å. A hexagonal supercell containing (8 × 8) unit cells of 16. Yang, M. et al. Catalytically active Au-O(OH)(x)-species stabilized by
graphene monolayer with about 17 Å vacuum layer was used as a support. The alkali ions on zeolites and mesoporous oxides. Science 346, 1498–1501 (2014).
convergence tolerances of energy, force, and displacement for the geometry 17. Flytzani-Stephanopoulos, M. Gold atoms stabilized on various supports
optimization were 1 × 10−5 Ha, 0.002 Ha/Å, and 0.005 Å, respectively. 1 × 1 × 1 k- catalyze the water-gas shift reaction. Acc. Chem. Res. 47, 783–792 (2014).
points grid is used to describe the Brillouin zone for geometric optimization. 18. Qiao, B. T. et al. Single-atom catalysis of CO oxidation using Pt-1/FeOx. Nat.
The adsorption energy is defined by the formula: Eads(AB) = E(AB/catalyst)–(Ecatalyst + Chem. 3, 634–641 (2011).
EAB) and Eads(H2) = E(H2/catalyst)–(Ecatalyst + EH2,gas), where E(AB/catalyst), E(H2/catalyst), 19. Peterson, E. J. et al. Low-temperature carbon monoxide oxidation catalysed by
and Ecatalyst are the total energies for the optimized equilibrium configurations regenerable atomically dispersed palladium on alumina. Nat. Commun. 5, 4885
of catalyst with and without AB or H2, respectively; and EAB (EH2,gas) is the (2014).
energy of the AB (gas phase H2) molecule in its ground state. For Pt(111), the 20. Vile, G. et al. A stable single-site palladium catalyst for hydrogenations. Angew.
supercell is (4×4). Chem. Int. Ed. 54, 11265–11269 (2015).
21. Liu, P. X. et al. Photochemical route for synthesizing atomically dispersed
Hydrolytic dehydrogenation of AB. As a probe reaction, hydrolytic dehy- palladium catalysts. Science 352, 797–801 (2016).
drogenation of AB was performed in a three-necked flask at 27 °C under atmo- 22. Yan, H. et al. Single-atom pd-1/graphene catalyst achieved by atomic layer
spheric pressure. The flask was immersed in a water bath to control the reaction deposition: remarkable performance in selective hydrogenation of 1,3-
temperature. 10 mg of the Pt2/graphene catalyst was used, while the weight of other butadiene. J. Am. Chem. Soc. 137, 10484–10487 (2015).
23. Ding, K. et al. Identification of active sites in CO oxidation and water-gas shift 51. Setthapun, W. et al. Genesis and evolution of surface species during pt atomic
over supported Pt catalysts. Science 350, 189–192 (2015). layer deposition on oxide supports characterized by in situ xafs analysis and
24. Jin, R. C., Zeng, C. J., Zhou, M. & Chen, Y. X. Atomically precise colloidal metal water-gas shift reaction. J. Phys. Chem. C 114, 9758–9771 (2010).
nanoclusters and nanoparticles: fundamentals and opportunities. Chem. Rev. 52. Karasulu, B., Vervuurt, R. H. J., Kessels, W. M. M. & Bo, A. A. Continuous and
116, 10346–10413 (2016). ultrathin platinum films on graphene using atomic layer deposition: a
25. Yoskamtorn, T. et al. Thiolate-mediated selectivity control in aerobic alcohol combined computational and experimental study. Nanoscale 8, 19829 (2016).
oxidation by porous carbon-supported Au-25 clusters. ACS Catal. 4, 3696–3700 53. Moses-DeBusk, M. et al. CO oxidation on supported single pt atoms:
(2014). experimental and ab initio density functional studies of co interaction with Pt
26. Fang, J. et al. The support effect on the size and catalytic activity of thiolated Atom on theta-Al2O3(010) surface. J. Am. Chem. Soc. 135, 12634–12645
Au-25 nanoclusters as precatalysts. Nanoscale 7, 6325–6333 (2015). (2013).
27. Das, S. et al. Reductive deprotection of monolayer protected nanoclusters: an 54. Xu, Y., Shelton, W. A. & Schneider, W. F. Thermodynamic equilibrium
efficient route to supported ultrasmall au nanocatalysts for selective oxidation. compositions, structures, and reaction energies of PtxOy (x = 1-3) clusters
Small 10, 1473–1478 (2014). predicted from first principles. J. Phys. Chem. B 110, 16591–16599 (2006).
28. Liu, J. et al. Ligand-stabilized and atomically precise gold nanocluster catalysis: 55. Hamilton, C. W., Baker, R. T., Staubitz, A. & Manners, I. B-N compounds for
a case study for correlating fundamental electronic properties with catalysis. chemical hydrogen storage. Chem. Soc. Rev. 38, 279–293 (2009).
Chem. Eur. J. 19, 10201–10208 (2013). 56. Chandra, M. & Xu, Q. A high-performance hydrogen generation system:
29. Nie, X. T., Qian, H. F., Ge, Q. J., Xu, H. Y. & Jin, R. C. CO oxidation catalyzed Transition metal-catalyzed dissociation and hydrolysis of ammonia-borane.
by oxide-supported Au-25(SR)(18) nanoclusters and identification of perimeter J. Power Sources 156, 190–194 (2006).
sites as active centers. ACS Nano 6, 6014–6022 (2012). 57. Chen, W. Y. et al. Mechanistic insight into size-dependent activity and
30. Wu, Z. L. et al. Thiolate ligands as a double-edged sword for CO oxidation on durability in pt/cnt catalyzed hydrolytic dehydrogenation of ammonia borane.
CeO2 Supported Au-25(SCH2CH2Ph)(18) Nanoclusters. J. Am. Chem. Soc. J. Am. Chem. Soc. 136, 16736–16739 (2014).
136, 6111–6122 (2014). 58. Chen, W. Y. et al. Unique reactivity in Pt/CNT catalyzed hydrolytic
31. Xu, Z. et al. Size-dependent catalytic activity of supported metal-clusters. dehydrogenation of ammonia borane. Chem. Commun. 50, 2142–2144 (2014).
Nature 372, 346–348 (1994). 59. Andrews, G. C. & Crawford, T. C. Synthetic utility of amine borane reagents in
32. Argo, A. M., Odzak, J. F., Lai, F. S. & Gates, B. C. Observation of ligand effects the reduction of aldehydes and ketones. Tetrahedron Lett. 21, 693–696 (1980).
60. Mahmoodinia, M. et al. Structural and electronic properties of the Pt-n-PAH
during alkene hydrogenation catalysed by supported metal clusters. Nature 415,
complex (n=1, 2) from density functional calculations. Phys. Chem. Chem.
623–626 (2002).
Phys. 16, 18586–18595 (2014).
33. Gates, B. C. Supported metal-clusters - synthesis, structure, and catalysis. Chem.
61. Mahmoodinia, M., Astrand, P. O. & Chen, D. Influence of carbon support on
Rev. 95, 511–522 (1995).
electronic structure and catalytic activity of pt catalysts: binding to the CO
34. George, S. M. Atomic layer deposition: an overview. Chem. Rev. 110, 111–131
Molecule. J. Phys. Chem. C 120, 12452–12462 (2016).
(2010).
62. Hummers, W. S. & Offeman, R. E. Preparation of graphitic oxide. J. Am. Chem.
35. Puurunen, R. L. Surface chemistry of atomic layer deposition: a case
Soc. 80, 1339–1339 (1958).
study for the trimethylaluminum/water process. J. Appl. Phys. 97, 121301
63. King, J. S. et al. Ultralow loading Pt nanocatalysts prepared by atomic layer
(2005). deposition on carbon aerogels. Nano Lett. 8, 2405–2409 (2008).
36. Knez, M., Niesch, K. & Niinisto, L. Synthesis and surface engineering of 64. Zhou, Y., King, D. M., Liang, X. H., Li, J. H. & Weimer, A. W. Optimal
complex nanostructures by atomic layer deposition. Adv. Mater. 19, 3425–3438 preparation of Pt/TiO2 photocatalysts using atomic layer deposition. Appl.
(2007). Catal. B Environ. 101, 54–60 (2010).
37. Lu, J. L., Elam, J. W. & Stair, P. C. Atomic layer deposition-Sequential self- 65. Hamalainen, J., Puukilainen, E., Sajavaara, T., Ritala, M. & Leskela, M. Low
limiting surface reactions for advanced catalyst “bottom-up” synthesis. Surf. Sci. temperature atomic layer deposition of noble metals using ozone and molecular
Rep. 71, 410–472 (2016). hydrogen as reactants. Thin Solid Films 531, 243–250 (2013).
38. O’Neill, B. J. et al. Catalyst design with atomic layer deposition. ACS Catal. 5, 66. Delley, B. Hardness conserving semilocal pseudopotentials. Phys. Rev. B 66,
1804–1825 (2015). 155125 (2002).
39. Detavernier, C., Dendooven, J., Sree, S. P., Ludwig, K. F. & Martens, J. A. 67. Perdew, J. P., Burke, K. & Ernzerhof, M. Generalized gradient approximation
Tailoring nanoporous materials by atomic layer deposition. Chem. Soc. Rev. 40, made simple. Phys. Rev. Lett. 77, 3865–3868 (1996).
5242–5253 (2011). 68. Delley, B. From molecules to solids with the DMol(3) approach. J. Chem. Phys.
40. Bagri, A. et al. Structural evolution during the reduction of chemically derived 113, 7756–7764 (2000).
graphene oxide. Nat. Chem. 2, 581–587 (2010). 69. Jiang, H. L. & Xu, Q. Catalytic hydrolysis of ammonia borane for chemical
41. Kim, K. et al. Selective metal deposition at graphene line defects by atomic layer hydrogen storage. Catal. Today 170, 56–63 (2011).
deposition. Nat. Commun. 5, 4781 (2014).
42. Lu, J. L. et al. Toward atomically-precise synthesis of supported
bimetallic nanoparticles using atomic layer deposition. Nat. Commun. 5, 3264 Acknowledgements
(2014). This work was supported by the National Natural Science Foundation of China
43. Kessels, W. M. M., Knoops, H. C. M., Dielissen, S. A. F., Mackus, A. J. M. (21473169, 21673215, 21533007, and 21233007), Innovative Research Groups of the
& van de Sanden, M. C. M. Surface reactions during atomic layer deposition of National Natural Science Foundation of China (11621063), the One Thousand Young
Pt derived from gas phase infrared spectroscopy. Appl. Phys. Lett. 95, 013114 Talents Program under the Recruitment Program of Global Experts, the Young Scientists
Fund of the National Natural Science Foundation of China (11404314), Anhui Provincial
(2009).
Natural Science Foundation (1708085MA06), and the Fundamental Research Funds for
44. Aaltonen, T., Rahtu, A., Ritala, M. & Leskela, M. Reaction mechanism studies
the Central Universities (WK2060030017). The calculations were performed on the
on atomic layer deposition of ruthenium and platinum. Electrochem. Solid State
supercomputing system in USTC-SCC and Guangzhou-SCC. We gratefully thank the
Lett. 6, C130–C133 (2003).
BL14W1 beamline at the Shanghai Synchrotron Radiation Facility (SSRF), China.
45. Mackus, A. J. M., Leick, N., Baker, L. & Kessels, W. M. M. Catalytic combustion
and dehydrogenation reactions during atomic layer deposition of platinum.
Chem. Mater. 24, 1752–1761 (2012). Author contributions
46. Hamalainen, J., Ritala, M. & Leskela, M. Atomic layer deposition of noble J.L. conceived the idea and designed the experiments. H.Y. performed catalyst synthesis
metals and their oxides. Chem. Mater. 26, 786–801 (2014). and catalytic performance evaluations. Y.L. performed the STEM measurements. S.W.,
47. Xue, Z. L. et al. Characterization of (Methylcyclopentadienyl)trimethylplatinum Z.S., H.C., W.L., and T.Y. performed the XAFS measurements. W.Z., H.W., and J.Y.
and low-temperature organometallic chemical vapor-deposition of platinum performed the DFT calculations. C.W., J.L., and X.H. assisted catalyst characterization
metal. J. Am. Chem. Soc. 111, 8779–8784 (1989). and catalytic performance tests. J.L. wrote the manuscript. All the authors contributed
48. Gong, T. et al. Activated carbon supported palladium nanoparticle catalysts and commented on the manuscript. H.Y. and Y.L. contributed equally to this work.
synthesized by atomic layer deposition: genesis and evolution of nanoparticles
and tuning the particle size. J. Phys. Chem. C 119, 11544–11556 (2015).
49. Lacroix, L. M., Arenal, R. & Viau, G. Dynamic HAADF-STEM observation of a
Additional information
Supplementary Information accompanies this paper at 10.1038/s41467-017-01259-z.
single-atom chain as the transient state of gold ultrathin nanowire breakdown.
J. Am. Chem. Soc. 136, 13075–13077 (2014). Competing interests: The authors declare no competing financial interests.
50. Shen, J., Muthukumar, K., Jeschke, H. O. & Valenti, R. Physisorption of an
organometallic platinum complex on silica: an ab initio study. New J. Phys. 14 Reprints and permission information is available online at http://npg.nature.com/
(2012). reprintsandpermissions/
Publisher's note: Springer Nature remains neutral with regard to jurisdictional claims in material in this article are included in the article’s Creative Commons license, unless
published maps and institutional affiliations. indicated otherwise in a credit line to the material. If material is not included in the
article’s Creative Commons license and your intended use is not permitted by statutory
regulation or exceeds the permitted use, you will need to obtain permission directly from
Open Access This article is licensed under a Creative Commons the copyright holder. To view a copy of this license, visit http://creativecommons.org/
Attribution 4.0 International License, which permits use, sharing, licenses/by/4.0/.
adaptation, distribution and reproduction in any medium or format, as long as you give
appropriate credit to the original author(s) and the source, provide a link to the Creative
Commons license, and indicate if changes were made. The images or other third party © The Author(s) 2017