You are on page 1of 15

Integrating Interactive Noble Metal Single Atom Catalysts into Transition

Metal Oxide Lattices

Jieqiong Shan†1,†#,
Chao Ye†1,†#
,
Chongzhi Zhu‡2,†#
,
Juncai Dong§3,#†
,
Wenjie Xu∥4
,
Ling Chen†1
,
Yan Jiao†1
,
Yunling Jiang†1
,
Li Song∥4
,
Yaning Zhang§3
,
Mietek Jaroniec⊥5
,
Yihan Zhu*,‡2
,
Yao Zheng*,†1
and
Shi-Zhang Qiao*,†1

†1 School of Chemical Engineering and Advanced Materials, The University of Adelaide, Adelaide, SA 5005, Australia.

‡2

Center for Electron Microscopy, State Key Laboratory Breeding Base of Green Chemistry Synthesis Technology and College of Chemica
Zhejiang University of Technology Hangzhou Zhejiang 310014 China
, , , , .

§3 Beijing Synchrotron Radiation Facility, Institute of High Energy Physics, Chinese Academy of Sciences, Beijing 100049,
China
.

4 National Synchrotron Radiation Laboratory, University of Science and Technology of China, Hefei 230029, China.
5 Department of Chemistry and Biochemistry &and Advanced Materials and Liquid Crystal Institute, Kent State University,
Kent OhioOH 44242 United StatesUSA
, , .

*Email: yihanzhu@zjut.edu.cn.

*Email: yao.zheng01@adelaide.edu.au.

*Email: s.qiao@adelaide.edu.au.

†#
J.S., C.Y., C. Z., and J.D. contributed equally to this paper.

AbstractBSTRACT

Noble metals have broad prospects for catalytic applications yet are restricted to a few packing modes with limited structural

flexibility. Here, we achieved geometric structure diversification of noble metals by integrating spatially correlated noble metal

single atoms (e.g., Pt, Pd, Ru) into the lattice of transition metal oxides (TMOs, e.g., Co3O4, Mn5O8, NiO, Fe2O3). The obtained

noble metal single atoms exhibited distinct topologies (e.g., crs, fcu-hex-pcu, fcu, and bcu-x) from those of conventional metallic

phases. As an example, Pt single atoms with a crs topology (Ptcrs-Co3O4) are endowed with tunable Pt–Pt correlations via the

synergy of metal–metal and metal–support interactions. A quantitative relationship between various Pt topologies determined by

TMOs substrates and their electrocatalytic activities was established. We anticipate that this type of interactive single atom catalysts

can bridge the geometric, topological, and electronic structure gaps between the ‘“close-packed’ packed” nanoparticles and isolated

single atoms as two common categories of heterogeneous catalysts.

KeywordsEYWORDS
single atom catalysts

noble metals

metal-–metal interactions

electrocatalysis

transition metal oxide

Introduction
Noble metals have received an enormous attention as highly efficient heterogeneous catalysts because of their unique electronic
1-6
structure and adsorption behaviorbehaviour for various reaction intermediates. Although great progress has been made in the
7-10
synthesis strategies and structural analysis, the geometric structures and underlying topologies of noble metal catalysts are
generally restricted to a few thermodynamically stable close-packing modes (e.g., fcu, hcp, etc.). 11-13 For example, most of the Pt-
14
based alloys generally exhibit analogous geometric structure to the Pt parent. The geometric structural inflexibility of noble metal

catalysts significantly limits the substantial electronic structure modulation and performance enhancement towards a wider range of

catalytic applications. Currently, it represents a cutting-edge research topic to eliminate such geometric restrictions and artificially

cast new noble metal structures with tunable atomic interactions.

There are generally two types of atomic interactions that are critical for most noble metal catalysts, namely, the metal–metal

interaction between two neighboringneighbouring metal atoms and metal–support interaction between metal atom and the substrate
15,16
(e.g., carbon, metal oxides, etc.). The former is governed by the local co-ordination structure of metal atoms and is tunable in
14,17 18 19 20,21 22,23
terms of composition, size , and geometry. While the latter is dictated by the charge transfer, strain , and bonding
24,25
pattern across the metal–support interface. Each of them is closely related to the apparent catalytic performance of noble metal

catalysts. However, there is a lack of a simple and general chemical strategy to finely tune these two types of interactions in one
26,27
catalyst to remarkably improve the catalytic performance. On one hand, the noble metal-based bulk crystals and nanoparticles

exhibit a very strong metal–metal interaction, which leads to restricted structural flexibility. On the other hand, in the emerging

single atom catalysts (SACs), although the regulation of intramolecular metal–support interactions has been extensively
25,28-30
investigated, the isolated single atom sites are far apart from each other and fail to achieve a sufficient metal–metal interaction
31,32
to trigger the synergistic effect. Very recently, it has been demonstrated that correlated metal single atoms with shortened inter-

site distance can exhibit metal–metal interaction, leading to distinct electronic structures and catalytic performance compared with
33-37
to isolated single atom counterparts. Nevertheless, it remains a grand challenge to achieve artificially designed topological

structure of noble metal single atoms with tunable metal–metal interactions, which may eliminate the current geometric restriction

of noble metals catalysts.

Following this concept, here we propose a general approach to diversify the geometric structure of noble metals by integrating

noble metal single atoms into the sublattice of metal oxide hosts with specific underlying topologies. Firstly, we achieved the

topological structure diversification of Pt single atoms, including crs (cristobalite), fcu-hex-pcu (face centered cubic-hexagonal-

primitive cubic), fcu, and bcu-x (body centered cubic), which were dictated in a series of transition metal oxides (TMOs, Co3O4,

Mn5O8, NiO, and Fe2O3) host structures. Such chemical strategy also permits the formation of a chosen topology and flexible

metal–metal coordination for various noble metal single atoms (i.e., Pt, Pd, and Ru) within cobalt spinel oxide. Among them, the

Ptcrs-Co3O4 catalysts exhibit an excellent electrocatalytic activity and stability for hydrogen evolution and oxidation reactions

(HER/HOR), which represent two critical and fundamental reactions in the hydrogen economy. A volcano-shaped plot can be

established between the intrinsic catalytic activity and hydrogen adsorption ability that is achieved on catalysts with various Pt

topologies. This chemical strategy creates a noble metal library with artificially designed topologies in comparison with to the

conventional noble metal metallic phases and alloys.

Results and Discussion


Structure Pprediction of Pt-TMOs Ccatalysts
To investigate the structural compatibility between Pt single atoms and diverse types of TMOs host structures, we constructed

ternary phase diagram calculated by using the density functional theory (DFT) to predict the thermodynamic stability of the hybrid

materials. We propose the local structural models by replacing a transition metal atom with a Pt atom within various TMOs
Figures 1a-d
matrices (Co3O4, Mn5O8, NiO, Fe2O3, CuO, and ZnO). The presence of orange regions in a–d and Figure S1 indicates

the thermodynamically stable Pt-Co11O16, Pt-Mn9O16, Pt-Ni15O16, and Pt-Fe11O18 structures. In contrast, for the Pt-incorporated CuO

and ZnO systems, Pt-Cu7O8 and Pt-Zn7O8 compounds cannot exist against the competing phases (Figure S2). We propose that this is

related to the coordinatcoordinatedion environment of transition metal sites in the TMOs. It has been suggested that the noble metal
38
sites prefer to substitute octahedral metal sites instead of the tetrahedral ones. Considering that in contrast to the other TMOs that

possess octahedral metal sites, the CuO and ZnO only possess tetrahedral sites, and therefore their incompatibility with Pt species

can be possibly ascribed to the difficulty to accommodate large Pt dopants in the tetrahedral voids in comparison to with the

octahedral ones in other TMOs. As a proof-of-concept, we integrated Pt species via a cation exchange process with different

metal-–organic frameworks (MOFs) as hosts. Followed by pyrolysis, the correlated Pt substitution whas been achieved in the

cationic sublattice of various TMO derivatives to achieve diverse topological structures (Pt contents in Pt-TMOs are summarized in

Table S1). The hybrid samples demonstrate typical X-ray diffraction (XRD) patterns of the corresponding metal oxides without Pt
Figure 1e
nanoparticles ( e).

Figure 1. Structural prediction and identification of various noble metal-substituted TMOtransition metal oxides. (a-–d) The pProjected

ternary phase diagrams of Pt-substituted TMOs for (a) Pt-–Co-–O, (b) Pt-–Mn-–O, (c) Pt-–Ni-–O, and (d) Pt-–Fe-–O in the (ΔμPt, ΔμTM)

planes. The oOrange regions represent the existence of stable Pt-Co11O16, Pt-Mn9O16, Pt-Ni15O16, and Pt-Fe11O18. (e) The XRD patterns of

Pt-substituted TMOs including Pt-Co3O4 (Co3O4: PDF#42-1467), Pt-NiO (NiO: PDF#47-1049), Pt-Mn5O8 (Mn5O8: PDF#39-1218), Pt-

Fe2O3 (Fe2O3: PDF#33-0664), Pt-ZnO (ZnO: PDF#36-1451), and Pt-CuO (CuO: PDF#65-2309).

Structure Iidentification of Ccorrelated Pt Ssites


Using carbon monoxide (CO) as a molecular probe, we conducted diffuse reflectance Fourier-transform infrared (DRIFT)

spectroscopy to identify the structure of Pt sites in various matrices. For comparison, a catalyst with isolated Pt single atoms

integrated in Co3O4 (PtSA-Co3O4, Figure S3) and commercial Pt/C catalyst were also included in the measurements. As shown in
Figure 2a
a, the DRIFT spectrum of adsorbed CO on PtSA-Co3O4 reveals a band centered at ∼~2200 cm−-1 and a weak band at 2115
39
cm−-1, which can be attributed to the CO adsorption oon Co3O4 (Figure S4) and isolated Pt single atoms, respectively. In addition,
39
Pt/C exhibits a band at ∼~2065 cm−-1 corresponding to a linear CO adsorption on planar Pt sites. Accordingly, the hybrids can be

classified into two groups. First, Pt-Co3O4, Pt-Mn5O8, Pt-NiO, and Pt-Fe2O3 samples show all the three sets of bands, demonstrating

that apart from the CO adsorption on TMOs and isolated Pt single atoms, there exists an extra type of CO adsorption mode.

Considering the absence of Pt nanoparticles in these hybrids, we propose that the band at ∼~2065 cm−-1 can be attributed to the

correlated Pt single atom sites within various TMO hosts, which demonstrate Pt bulk-like CO adsorption behaviorbehaviour. In
contrast, the DRIFT spectra of CO-adsorbed Pt-CuO and Pt-ZnO only exhibit the CO adsorption on TMOs and a very weak signal

of isolated Pt single atoms, indicating that the absence of the correlated Pt sites in the CuO and ZnO matrices.

Figure 2. Structural identification of Pt-TMOs. (a) DRIFT spectra on CO adsorbed various catalysts. Three sets of bands illustrated by the

blue (∼~2200 cm−-1), graygrey (∼~2115 cm−-1), and orange (∼~2065 cm−-1) areas refer to CO adsorption on TMOs hosts, isolated Pt

single atoms, and correlated Pt sites, respectively. (b) Wavelet transform edge extended X-ray absorption fine structure (WT-EXAFS)

spectra of various catalysts. The hHybrid samples show three intensity maxima at around k = 4.0, 6.0–-7.0 toand 8.5–-10.0 Å−-1, which

can be respectively ascribed to the Pt-–O, Pt-–TM (TM = Co/Mn/Ni/Fe), and Pt-–Pt scatterings.

The synchrotron-based Pt L3-edge extended X-ray absorption fine structure Fourier transform EXAFS (FT-EXAFS) spectra were

measured recorded to characterize the local coordination structure of the Pt sites (Figure S5). Taking Pt-Co3O4 as an example, there

are two major peaks at 1.6 and 2.7 Å, while the first peak can be assigned to Pt–O coordination, the EXAFS Wavelet transform

(WT-EXAFS) analysis reveals that the second peak originates from due to the co-existence of Pt–Co and Pt–Pt interactions (
Figure 2b
b, Table S2). In contrast, the fitting of FT-EXAFS spectra for PtSA-Co3O4 demonstrates the lack of a second-shell Pt-–Pt

coordination, which confirms the isolation of Pt single atoms and is distinct from that of Ptcrs-Co3O4. These results confirm the

integration of correlated Pt sites instead of fully isolated ones in the TMO matrices.

Topological Ddiversification of Ccorrelated Pt Ssingle Aatoms


As a direct evidence, high-angle annular dark-field high-resolution scanning transmission electron microscopy (HAADF-

HRSTEM) images demonstrate show that the Pt sites are atomically incorporated into the TMO lattices (Figures S6, and S7). A

careful inspection of HRSTEM permits the atomic -resolution structural elucidation of the hybrid materials. The presence of

multiple correlation peaks in the calculated partial projected pair distribution function (pPDF) profiles of Pt sites confirms that the

correlated Pt sites adopt an identical spatial correlation with the host cationic sublattice of octahedral cobalt (Cooct) sites (Figure

S8). Such a Cooct sublattice adopts a crs topology with a 3D network composed of corner-shared tetrahedra in a staggered fashion,
Figure 3a
which interpenetrates with a tetrahedral cobalt (Cotet) sublattice that adopts a dia topology ( a). This permits a new Pt

topology of Ptcrs-Co3O4, which eliminates the close-packing limitation in most metallic Pt phases that usually adopt an fcu or hcp

topology. Similar analysis shows that most Pt atoms with a brighter contrast are allocated in the cationic sublattices of Mn5O8, NiO,

and Fe2O3 host structures with an interpenetrating fcu-hex-pcu net, a standalone fcu net, and a bcu-x net, respectively. These

arrangements result in the formation of Ptfcu-hex-pcu-Mn5O8, Ptfcu-NiO, and Ptbcu-x-Fe2O3 catalysts with novel Pt topologies (
Figure 3b-d
b–d). As expected, only a few Pt single atoms can be integrated in the CuO and ZnO host structures, making the formation of

correlated Pt sites with specific topological structures difficult (Figure S9).

Figure 3.Topological diversification of correlated Pt sites. From top to bottom: the topology models, HRSTEM images with Pt-sites

marked with yellow circles, simulated Z2 maps, and structure models of the corresponding red dashed rectangular areas in the images at top

of (a) Pt-Co3O4, (b) Pt-Mn5O8, (c) Pt-NiO, and (d) Pt-Fe2O3, respectively. Correlated Pt sites form polyhedrons in the topology models are

rendered in light green. Two sets of cationic sublattices of Co cations in Pt-Co3O4 were rendered in two different colors, referring to crs
and, dia underlying topologies. Three sets of cationic sublattices of Mn cations in Pt-Mn5O8 are rendered in different colors, which refer to

fcu, hex, and pcu underlying topologies. The cCationic sublattices of Pt-NiO and Pt-Fe2O3 hybrid structures are assigned to fcu and bcu-x

nets, respectively. The cCorresponding zone axes of these HRSTEM images for Co3O4, Mn5O8, NiO, and Fe2O3 are [233], [47 2], [121],

and [8 10 1]h, respectively.

Extension of NM-Co3O4 Ccatalysts


As demonstrated above, Co3O4 can be adopted as an appropriate host to endow Pt single atoms with desired topological

structures. The investigation of structural compatibility is further extended to various noble metal single atoms (Ru, Pd, and Au) by

taking Co3O4 as a model platform. The presence of orange regions in Figure S10 indicates the thermodynamically stable NM-

Co11O16 (NM = Ru and, Pd) structure. In contrast, for an Au-incorporated Co3O4 system, a compound of Au-Co11O16 cannot exist

against the competing phases. The synthesized Ru- and Pd-Co3O4 catalysts exhibit typical XRD patterns including both

characteristic peaks for cobalt spinel oxide with an Fd3 m symmetry and one or more weak diffuse peaks at 2θ of ∼16°, ∼22°,
Figure 4a
and/or ∼24° arising from nanodomains with a broken symmetry ( a). This indicates the spatial correlation of the

incorporated noble metal single atoms within the cationic sublattice of the spinel Co3O4 matrix. HAADF-HRSTEM images

demonstrate show that almost all alien ions in the synthesized Ru- and Pd-Co3O4 samples are atomically incorporated into the spinel
Figure 4b,c
oxide lattice ( b,c).

Figure 4. Structure of Ru and Pd substituted-Co3O4 hybrids. (a) The XRD patterns. HRSTEM images of (b) Ru-Co3O4 and (c) Pd-Co3O4

catalysts reveal that almost all Ru and Pd atoms (indicated by yellow circles) are incorporated in the Co3O4 lattice. K3-weighted FT-EXAFS

spectra of (d) Ru K-edge data for Ru-Co3O4, (e)€ Pd K-edge data for Pd-Co3O4, and (f) Au L3-edge data for Au-Co3O4 together with the

corresponding spectra for references.

As indicated by the X-ray absorption near-edge structure (XANES) spectra, most of the noble metal single atoms in the NM-

Co3O4 samples possess higher valence states than those in their bulk references, indicating the integration of noble metal atoms into

the cationic sublattice of Co3O4 (Figure S11). This phenomenon is absent in Au-Co3O4 due to the segregation of Au species. The

FT-EXAFS spectra and corresponding curve-fitting demonstrates the co-existence of Ru(Pd)-Cooct and Ru(Pd)-Ru(Pd) interactions

in the second-shell coordination, confirming that the Ru and Pd atoms can be perfectly integrated into the cationic sublattice of the

spinel oxide host structure to construct composite active sites with simultaneously tunable metal–metal and metal–support
Figure 4d,e
interactions ( d,e). In contrast, the spectra of Au-Co3O4 indicate that the Au sites exist mainly as Au nanoparticles with
Figure 4g
the typical Au–Au bond ( g). Our recent findings suggested that the aggregation of Au sites can be attributed to the
40
relatively difficult transition of first-shell coordination from the Au-–ligand to Au-–O during pyrolysis in air.

Catalytic Pproperties of Ptcrs-Co3O4


As demonstrated above, various TMOs can be adopted as appropriate hosts to endow the correlated Pt single atoms with desired

topological structures. It is expected that the topologically structured noble metal active sites with spatial correlation would exhibit
distinct catalytic performance from either a conventional single atom or nanoparticle catalysts. HER/HOR redox couple is one of
41
the best-established reaction mechanisms and Pt/C is the a state-of-the-art electrocatalyst for this process. As the Pt-Co3O4 sample

shows the best electrocatalytic performance among different noble metals-based catalysts (Figure S12), we focus on the Pt-TMOs

catalysts to highlight the effect of different Pt topologies towardstoward their electrocatalytic activities. HER/HOR polarization

curves normalized to the electrochemical active surface area (ECSA) demonstrate that among the Pt-TMOs catalysts with various

Pt topologies, the Ptcrs-Co3O4 delivers the highest specific activity (Figures S13, and 14).

To investigate the effect of the topological structure of Ptcrs-Co3O4 on the electrocatalytic performance, PtSA-Co3O4 and Co3O4

supported Pt nanoparticles (PtNP@Co3O4, Figure S15) catalysts were investigated. The Ptcrs-Co3O4 catalyst exhibits a remarkably

higher specific HER/HOR activity than the PtNP@Co3O4 and PtSA-Co3O4 counterparts, indicating the importance of topologically
Figures 5a
structured Pt active sites in enhancing the overall catalytic performance ( a, and Figure S16, 17). By fitting polarization

curves with the Butler-–Volmer equation with minimizing the reactant’s/product’s mass transport effect, the exchange current
Figures 5b 42
density (j0) values of the catalysts can be extracted ( b, and Figure S18). The Ptcrs-Co3O4 catalyst delivers a specific j0 of

3.78 mA cmPt−2, significantly outperforming the PtSA-Co3O4 (0.75 mA cmPt−2) and PtNP@Co3O4 (0.15 mA cmPt−2) catalysts, and

even the-state-of-the-art Pt/C (1.06 mA cmPt−2) with the same Pt mass loading (Figure S19). It is also shown that at the same Pt

loading of ∼~10 μgPt cm−-2, the Ptcrs-Co3O4 catalyst delivers an 8.3-fold higher turn-over frequency (TOF) of 9.41 s−-1 (at an

overpotential of −-15 mV) than that of Pt/C (1.13 s−-1), outperforming the recent reported HER catalysts evaluated in acidic

environments (Table S3).

Figure 5. HER/HOR performance evaluation of Ptcrs-Co3O4 catalyst. (a) HER/HOR polarization curves on various catalysts in H2-satuated

0.1 M HClO4. (b) Tafel plots of Ptcrs-Co3O4 and Pt/C catalysts (solid curves) and fitting lines to the Butler–Volmer equation (dashed

curves). (c) In situ Co-K edge and Pt-L3 edge FT-EXAFS spectra of Ptcrs-Co3O4 catalyst under open circle potential (OCP) and different

applied potentials (vsvs. Ag/AgCl). (d) Percentages of metal species dissolved in electrolyte over durability test detected by inductively

coupled plasma mass spectrometry (ICP-MS).

In addition, the strong dual interactions offer the Ptcrs-Co3O4 catalyst a robust HER/HOR redox stability (Figure S20). In situ Co-

K and Pt-L3 edge EXAFS spectra demonstrate that the local co-ordination environment of Co and Pt species remains almost
Figure 5c
unchanged during the reaction ( c). The corresponding XANES spectra further indicate that the oxidization state of Pt

remains unchanged and no leaching of the metallic Pt species from the host lattice can be observed (Figure S21). Further, the

inductively coupled plasma mass spectrometry (ICP-MS) results demonstrate that the loss of Pt and Co metal species on the Ptcrs-

Co3O4 catalyst after 600 mins of the stability test was keptremained at a low level of around 1.89 wt %wt% and to 0.84 wt %wt%,
Figure 5d
respectively ( d). By contrast, the PtSA-Co3O4 and PtNP@Co3O4 catalysts suffered from a significant metal loss (∼~4 wt

%wt% and to ∼~ 10 wt %wt%, respectively) after 180 mins of continuous HER electrocatalysis.

Construction of Sstructure-–Aactivity Rrelationship on Pt-TMOs


To elucidate the unique geometric and electronic properties of the topologically structured Pt sites endowed by various TMO

hosts, the chemical interaction between Pt atoms was investigated by using the crystal overlap Hamilton population (COHP)
43 Figure 6a
method. As shown in a, this chemical strategy permits a widely extended range of the integrated-COHP (ICOHP) values,

suggesting a more flexible regulation of the Pt-–Pt atomic interaction in comparison with to those of the conventional Pt metallic
43
phases and alloys. Specifically, three neighboring Pt atoms (Pt1, Pt2, and Pt3) were allocated in various TMOs matrices with the

bonding and antibonding interactions between each two Pt atoms (Pt1-–Pt2, Pt1-–Pt3, and Pt2-–Pt3) investigated by using the

COHP plots (Figure S22). The strength of the corresponding bonds can be evaluated by the ICOHP values recorded at the Fermi

level. Taking Pt-Fe2O3 as an example, the COHP plots corresponding to various Pt-–Pt bonds demonstrate different bonding and

antibonding interactions. As supported by ICOHP values, the strength of the Pt-–Pt bonds can be regulated to as high as 1.28 eV

for the Pt1-–Pt3 bond and as low as 0.20 eV for the Pt2-–Pt3 bonds, respectively. Similar results can be observed for the other

hybrid structures, indicating that the formation of Pt topological structures by integrating them in TMO hosts provides a promising

venue to flexibly regulate the Pt-–Pt interactions, therefore significantly modify the catalytic properties towards various

applications.

Figure 6. Structure-–property relationship of Pt-TMOs catalysts. (a) Calculated values of ICOHP for Pt-TMO structures, which indicate the

strength of Pt-–Pt interactions. Each data point represents an ICOHP value between each two Pt atoms in a structure model with three Pt

substitutions as shown in Figure S22. The Pt-TMO structures demonstrate a widely extended ICOHP ranges in comparison with those of

conventional Pt metal and Pt alloys. (b) Tafel plots of Pt-TMOs catalysts (solid curves) and fitting lines to the Butler–Volmer equation

(dashed curves). (c) The rRelationship between the experimentally determined j0 on Ptcrs-Co3O4, Ptfcu-hex-pcu-Mn5O8, Ptfcu-NiO, Ptbcu-x-Fe2O3,

and Pt/C and the corresponding ΔGH* values on Pt-TMO models with two Pt substitutions and Pt (111). (d) The mMainly the linear

relationship between ΔGH* values and d-band center of Pt-TM domain containing central Pt atom and the nearest neighboring TM/Pt sites

on Pt-TMOs models with two Pt substitutions.

To demonstrate the effect of Pt-–Pt interaction on the catalytic performance, we performed experimental measurements and

theoretical calculations to establish a structure-–activity relationship on the Pt-TMOs catalysts and reveal the origin of HER

catalytic activity. The Pt-TMOs catalysts with various Pt topologies deliver different j0 values of 3.78 mA cm-2Pt, 1.95 mA cm-2Pt,
Figure 6b
0.84 mA cm-2Pt, and 3.06 mA cmPt−2 for Ptcrs-Co3O4, Ptfcu-hex-pcu-Mn5O8, Ptfcu-NiO, and Ptbcu-x-Fe2O3, respectively ( b). We

calculated the ΔGH* values for a series of the Pt-TMOs structure models with 1–-2 Pt atoms by DFT calculations (Figure S23). A

volcano shaped plot can be established between the intrinsic activity j0 of Ptcrs-Co3O4, Ptfcu-hex-pcu-Mn5O8, Ptfcu-NiO, Ptbcu-x-Fe2O3, and

Pt/C catalysts and the ΔGH* values of the corresponding structured Pt-TMOs models with two Pt single atoms and Pt (111) (
Figure 6c
c). It is shown that the topologically structured Pt catalysts with the ΔGH* values closer to the volcano vertex possess a higher j0.

In addition, a mainly linear relationship between ΔGH* values and d-band centers of the Pt-TM active domain can be established on
Figure 6d
the Pt-TMOs catalysts ( d). Therefore, this study shows that the remarkable catalytic activity of Pt-TMOs can be attributed

to the optimized hydrogen adsorption behavior of the Pt sites, which originates from the modified geometric and electronic

structures of unique Pt topologies.


Conclusions
In summary, we report a general approach to construct unique noble metals single atom’s configuration, which show a distinct

electronic structure and enhanced catalytic performance in comparison with to the ‘“close-packed’ packed” nanoparticles and

conventional SACs. By applying this strategy, the Pt single atoms are integrated in the cationic sublattice of various TMOs hosts to

exhibit artificially designed topological structures (e.g., crs, fcu-hex-pcu, fcu, and bcu-x). The various topologically structured Pt

single atoms demonstrate a more flexible regulation of Pt-–Pt atomic interaction in comparison with conventional Pt metallic

phases and alloys. The resultant Ptcrs-Co3O4 catalyst delivers an 8.3-fold higher TOF of 9.41 s−-1 (at an overpotential of −-15 mV)

towards HER than that of Pt/C (1.13 s−-1) at the same Pt loading. Therefore, this work presents a general strategy to create a noble

metal catalyst library with desired topologies for a wide variety of challenging catalytic applications.

Materials and Methods

Material Ssynthesis
Co-based zeolitic imidazolate framework (ZIF-67) nanocubes were synthesized by a surfactant-mediated method reported
44
previously. Afterwards an ion exchange process was conducted on the obtained ZIF-67 nanocubes in the presence of the platinum

precursor in aqueous solution. Typically, 100 mg of ZIF-67 nanocubes was first dispersed in 50 mL of deionized (DI) water, and

then 10 mL of aqueous solution containing 9.7 mg of potassium hexachloroplatinate (IV) (K2PtCl6) was added under stirring

conditions. The suspension was centrifuged, and 3 hhours later the precipitate was collected and repeatedly washed with ethanol

and DI water. After being dried overnight in a vacuum oven at 60 °oC, the precipitate was pyrolyzed at 300 °C in air for 4 hhours.

The Pt content under the stoichiometry of Co3O4 was determined by ICP-MS to be 0.06. The pure Co3O4 was synthesized by

pyrolyzing the ZIF-67 nanocubes under similar procedure without performing the ion exchange process. The PtNP@Co3O4 catalyst

was synthesized by adding 10 mL of aqueous solution containing 7.5 mg of K2PtCl6 into 50 mL of aqueous solution containing

100 mg of Co3O4 nanoparticles. Later 10 mL of aqueous solution containing 11.4 mg of sodium borohydride was added into the

mixture to reduce K2PtCl6 and immobilize Pt species on the surface of Co3O4 nanoparticles.

HRSTEM Iimaging and Ssimulations


HAADF-STEM images were recorded using a Thermo Fisher Spectra 300 electron microscope operated at 300 kV and equipped
45
with a probe corrector. The projected Z2-map and probe profile simulations were carried out by using the QSTEM program with

different supercell configurations. For the 3D probe profile simulation, a focal series from −-30 to 30 nm with an interval of 1 nm

was conducted, using a convergence angle of 30 mrad, third order spherical aberration (C3) of 1 μm, and fifth order spherical

aberration (C5) of 300 μm. After flat-field correction and background subtraction, the position searching, refining, and labelling of
46
the projected Pt sites with brighter contrast in the HRSTEM image were carried out using the CalAtom software with a mMultiple
47
-Eellipse-Ffitting (MEF) method. Instead of directly locating the individual Pt sites within the cationic sublattice of the host
structure, identifying “projected Pt clusters” from as-labeledlabelled Pt doped brighter atomic columns with an intensity penalty

(>70%) offers a sensible evaluation of the 3D clustering probability of the Pt sites over ultrathin regions. Based on the simulated 3D

probe profile using the imaging parameters and aberration coefficients similar to experimental imaging conditions, it is clear that

the identified Pt atoms at the cationic sites are most likely within a few unit cells in height and can be regarded as correlated sites

with a short-range order.

X-ray absorption spectroscopyXAS Aanalysis


The X-ray absorption spectroscopy (XAS) measurements were performed in fluorescence mode at the beamlines of 14 W1 in

Shanghai Synchrotron Radiation Facility and 4B9A and 1W1B in Beijing Synchrotron Radiation Facility. For Pt and Au L3-edge,

the X-ray was monochromatized by a double-crystal Si (111) monochromator, while for Pd and Ru K-edge, the X-ray was

monochromatized by a double-crystal Si (311) monochromator. The incident and fluorescence X-ray intensities were monitored by

using the standard ion chambers and Lytle-type detector, respectively, and the monochromator was detuned to reject higher

harmonics. While the XAS raw data were background subtracted, normalized, and Fourier transformed by standard procedures

within the ATHENA program, the least-squares curve fitting analysis of the EXAFS data was carried out using the ARTEMIS
48
program. The in situ XAS measurements were performed using a home-made three-electrode electrochemical cell and a

computer-controlled electrochemical analyzer. A catalyst modified carbon paper was used as the working electrode, Pt wire as the

counter electrode, and Ag/AgCl electrode as the reference electrode. The electrochemical cell was filled with N2-saturated 0.1 M

HClO4 electrolyte solution.

Electrochemical Mmeasurements
For the Ptcrs-Co3O4 samples, the electrocatalyst ink was prepared by dispersing a freshly synthesized catalyst powder (1 mg) and

carbon black (Vulcan-XC72, 1 mg) in a solution containing distilled water (Milli-Q, 965 μL) and 5 weight % Nafion solution (35

μL) followed by ultrasonication for 2 hhours. We decreased the mass loading of Pt to minimize the mass transport limitation. 40

μL of catalyst ink was then deposited onto a polished glassy carbon electrode (diameter = 5 mm, area = 0.196 cm2, Pine

Research Instrument). All electrochemical experiments were carried out using rotating disk electrode method in a three-electrode

glass cell with an Au-wire as the counter electrode and an Ag/AgCl as the reference electrode (Pine Research Instrument). The

reference electrode was calibrated by cyclic voltammetry method in H2-saturated 0.1 M HClO4 solution. The HER and HOR

measurements were conducted in H2-saturated 0.1 M HClO4 electrolyte with a CHI potentiostat (CHI 760D) at different rotating

speeds. The establishment of the butterfly curve and the extraction of exchange current density would facilitate a precise

assessment of HER/HOR catalytic performance. The polarization curves of the catalysts were obtained with scan rates of 2 mV s−-
49
1
. The ECSAs of the catalysts were determined by CO stripping voltammograms. The electrolyte was bubbled with 20% CO in

argon for 30 min with the working electrode held at 0.1 V vsvs. reversible hydrogen electrode (RHE), followed by argon purging

for 20 min to remove the excess CO. Then the CO stripping voltammograms were collected in the potential range from around 0 V
to 1.2 V vsvs. RHE during argon bubbling. The multi-potential steps process was performed with 40 μL of Ptcrs-Co3O4 catalyst ink

deposited onto carbon paper (area = 0.33 cm2) and polarized alternatively under HER (−-0.1 V vsvs. RHE) and HOR (0.5 V

vsvs. RHE) potentials.

DFT Ccalculations
DFT calculations were carried out using the Vienna ab-initio sSimulation Package (VASP). 50,51 The exchange-correlation
52
interaction was described by generalized gradient approximation (GGA) with the Perdew-–Burke-–Ernzerhof (PBE) functional.
53
The DFT-TS method of Grimme was employed to treat the VDW interaction. The Ddenser 8 × 8 × 2 K-points were used for

the density of states (DOS) calculations. In an acidic solution, a simple method was used to compute the free energy based on the
54
hydrogen adsorption strength, as in previous calculations :

ΔGH* = ΔEH + 0.24 eV

ΔEH = Etotal –− 1/2EH2 –− Es

where Etotal, EH2, and Es are the energies of the whole system, hydrogen, and substrate, respectively.

ASSOCIATED CONTENT.

Supporting Information

This material is available free of charge via the Internet at http://pubs.acs.org.

 Detailed computational and experimental data (PDF).

AUTHOR INFORMATION.

Corresponding Author.

yihanzhu@zjut.edu.cn

yao.zheng01@adelaide.edu.au

s.qiao@adelaide.edu.au

Funding Sources
This work was financially supported by the Australian Research Council (FL170100154, DP220102596, and DP190103472). Y.H.Z.

acknowledges the financial support from the National Key Research and Development Program of China (2022YFE0113800) and the

National Natural Science Foundation of China (22,122,505 and, 22,075,250), Zhejiang Provincial Natural Science Foundation of China
(LR18B030003). J.D. acknowledgements the financial support from Youth Innovation Promotion Association of Chinese Academy of

Sciences (2018017).

Notes.

Authors declare no competing interests.

AcknowledgmentCKNOWLEDGMENT
DFT computations were performed by using the services offered from by the National Computational Infrastructure (NCI) and Phoenix

High Performance Computing, which are supported by the Australian Government and The University of Adelaide.

ReferencesEFERENCES
1. Mavrikakis, M.; Stoltze, P.; Nørskov, J. K. Making gold less noble. Catal. Lett. 2000, 64, 101– 106, 10.1023/A:1019028229377.

2. Shi, Q.; Zhu, C.; Du, D.; Lin, Y. Robust noble metal-based electrocatalysts for oxygen evolution reaction. Chem. Soc. Rev. 2019, 48, 3181–
3192
10.1039/C8CS00671G
, .

3. Shi, Y.; Lyu, Z.; Zhao, M.; Chen, R.; Nguyen, Q. N.; Xia, Y.
Noble-Metal Nanocrystals with Controlled Shapes for Catalytic and Electrocatalytic Applications
Chem. Rev. 2021 121 649 735 10.1021/acs.chemrev.0c00454
. , , – , .

4. Stephens, I. E. L.; Bondarenko, A. S.; Grønbjerg, U.; Rossmeisl, J.; Chorkendorff, I.


Understanding the electrocatalysis of oxygen reduction on platinum and its alloys
Energy Environ. Sci. 2012 5 6744 6762 10.1039/c2ee03590a
. , , – , .

5. Corma, A.; Garcia, H. Supported gold nanoparticles as catalysts for organic reactions. Chem. Soc. Rev. 2008, 37, 2096– 2126,
10.1039/b707314n
.

6. Stamenkovic, V. R.; Mun, B. S.; Arenz, M.; Mayrhofer, K. J.; Lucas, C. A.; Wang, G.; Ross, P. N.; Markovic, N. M.
Trends in electrocatalysis on extended and nanoscale Pt-bimetallic alloy surfaces
Nat. Mater. 2007 6 241 247 10.1038/nmat1840
. , , – , .

7. Zhao, Z.; Chen, C.; Liu, Z.; Huang, J.; Wu, M.; Liu, H.; Li, Y.; Huang, Y. Pt-Based Nanocrystal for Electrocatalytic Oxygen Reduction.
Adv. Mater.
2019 31 e1808115 10.1002/adma.201808115
, , , .

8. Wu, B.; Zheng, N. Surface and interface control of noble metal nanocrystals for catalytic and electrocatalytic applications. Nano Today 2013, 8,
168 197 10.1016/j.nantod.2013.02.006
– , .

9. Li, Y.; Sun, Y.; Qin, Y.; Zhang, W.; Wang, L.; Luo, M.; Yang, H.; Guo, S.
Recent Advances on Water-Splitting Electrocatalysis Mediated by Noble-Metal-Based Nanostru
Adv. Energy Mater.2020 10 1903120 10.1002/aenm.201903120
. , , , .

10. Shan, J.; Zheng, Y.; Shi, B.; Davey, K.; Qiao, S.-Z.
Regulating Electrocatalysts via Surface and Interface Engineering for Acidic Water Electrooxidation
ACS Energy Lett. 2019 4 2719 2730 10.1021/acsenergylett.9b01758
. , , – , .

11. Nesselberger, M.; Roefzaad, M.; Hamou, R. F.; Biedermann, P. U.; Schweinberger, F. F.; Kunz, S.; Schloegl, K.; Wiberg, G. K.; Ashton, S.;
Heiz, U. Mayrhofer, K. J. Arenz, M.
; ;
The effect of particle proximity on the oxygen reduction rate of size-selected platinum clusters
Nat. Mater. 2013 12 919 924 10.1038/nmat3712
. , , – , .

12. Liu, L.; Corma, A. Metal Catalysts for Heterogeneous Catalysis: From Single Atoms to Nanoclusters and Nanoparticles. Chem. Rev. 2018, 118,
4981 5079 10.1021/acs.chemrev.7b00776
– , .
13. Chen, H. S.; Benedetti, T. M.; Goncales, V. R.; Bedford, N. M.; Scott, R. W. J.; Webster, R. F.; Cheong, S.; Gooding, J. J.; Tilley, R. D.

Preserving the Exposed Facets of Pt3Sn Intermetallic Nanocubes During an Order to Disorder Transition Allows the Elucidation of the Effect
J. Am. Chem. Soc. 2020 142 3231 3239
. , , – ,
10.1021/jacs.9b13313
.

14. Greeley, J.; Stephens, I. E.; Bondarenko, A. S.; Johansson, T. P.; Hansen, H. A.; Jaramillo, T. F.; Rossmeisl, J.; Chorkendorff, I.; Norskov, J. K.
Alloys of platinum and early transition metals as oxygen reduction electrocatalysts Nat. Chem. 2009 1 552 556
. , , – ,
10.1038/nchem.367
.

15. Yang, J.; Li, W.; Wang, D.; Li, Y. Electronic Metal-Support Interaction of Single-Atom Catalysts and Applications in Electrocatalysis.
Adv. Mater.
2020 32 2003300 10.1002/adma.202003300
, , , .

16. Li, X.; Yang, X.; Huang, Y.; Zhang, T.; Liu, B. Supported Noble-Metal Single Atoms for Heterogeneous Catalysis. Adv. Mater.2019, 31,
1902031 10.1002/adma.201902031
, .

17. Stamenkovic, V. R.; Fowler, B.; Mun, B. S.; Wang, G.; Ross, P. N.; Lucas, C. A.; Markovic, N. M.
Improved Oxygen Reduction Activity on Pt3Ni(111) via Increased Surface Site Availability
Science 2007 315 493 497 10.1126/science.1135941
. , , – , .

18. Lee, S. W.; Chen, S.; Sheng, W.; Yabuuchi, N.; Kim, Y. T.; Mitani, T.; Vescovo, E.; Shao-Horn, Y.
Roles of Surface Steps on Pt Nanoparticles in Electro-oxidation of Carbon Monoxide a
J. Am. Chem. Soc. 2009 131 15669 15677 10.1021/ja9025648
. , , – , .

19. Tsung, C.; Kuhn, J.; Huang, W.; Aliaga, C.; Hung, L.; Somorjai, G.; Yang, P.
Sub-10 nm Platinum Nanocrystals with Size and Shape Control: Catalytic Study for E
J. Am. Chem. Soc. 2009 131 5816 5822 10.1021/ja809936n
. , , – , .

20. Lykhach, Y.; Kozlov, S. M.; Skala, T.; Tovt, A.; Stetsovych, V.; Tsud, N.; Dvorak, F.; Johanek, V.; Neitzel, A.; Myslivecek, J.; Fabris, S.;
Matolin, V. Neyman, K. M. Libuda, J. Counting electrons on supported nanoparticles Nat. Mater. 2016 15 284 288
; ; . , , – ,
10.1038/nmat4500
.

21. Pacchioni, G. Electronic interactions and charge transfers of metal atoms and clusters on oxide surfaces. Phys. Chem. Chem. Phys. 2013, 15,
1737 1757 10.1039/c2cp43731g
– , .

22. Zhou, Z.; Zaman, W. Q.; Sun, W.; Zhang, H.; Tariq, M.; Cao, L.; Yang, J.
Effective Strain Engineering of IrO2 Toward Improved Oxygen Evolution Catalysis thr
ChemElectroChem 2019 6 4586 4594 10.1002/celc.201901037
. , , – , .

23. Du, M.; Cui, L.; Cao, Y.; Bard, A. J.


Mechanoelectrochemical catalysis of the effect of elastic strain on a platinum nanofilm for the ORR exerted by a shape me
J. Am. Chem. Soc. 2015 137 7397 7403 10.1021/jacs.5b03034
. , , – , .

24. Sanz, J. F.; Marquez, A. Adsorption of Pd Atoms and Dimers on the TiO2 (110) Surface: A First Principles StudyJ.
. Phys. Chem. C 2007, 111,
3949 3955 10.1021/jp0639952
– , .

25. Yu, N.-F.; Tian, N.; Zhou, Z.-Y.; Sheng, T.; Lin, W.-F.; Ye, J.-Y.; Liu, S.; Ma, H.-B.; Sun, S.-G.
Pd Nanocrystals with Continuously Tunable High-Index Facets as a Model Nanocatalyst
ACS Catal. 2019 9 3144 3152 10.1021/acscatal.8b04741
. , , – , .

26. Wang, H.; Wang, L.; Lin, D.; Feng, X.; Niu, Y.; Zhang, B.; Xiao, F.-S.
Strong metal–support interactions on gold nanoparticle catalysts achieved through Le Chatelier’s prin
Nat. Catal. 2021 4 418 424 10.1038/s41929-021-00611-3
. , , – , .

27. Dong, C.; Li, Y.; Cheng, D.; Zhang, M.; Liu, J.; Wang, Y.-G.; Xiao, D.; Ma, D.
Supported Metal Clusters: Fabrication and Application in Heterogeneous Catalysis
ACS Catal. 2020 10 11011 11045 10.1021/acscatal.0c02818
. , , – , .

28. Yang, X.; Wang, A.; Qiao, B.; Li, J.; Liu, J.; Zhang, T. Single-Atom Catalysts: A New Frontier in Heterogeneous Catalysis. Acc. Chem. Res.
2013 46 1740 1748 10.1021/ar300361m
, , – , .
29. Zhang, Y.; Wu, C.; Jiang, H.; Lin, Y.; Liu, H.; He, Q.; Chen, S.; Duan, T.; Song, L.
Atomic Iridium Incorporated in Cobalt Hydroxide for Efficient Oxygen Evolution
Adv. Mater.2018 30 e1707522 10.1002/adma.201806345
. , , , .

30. Li, J.; Guan, Q.; Wu, H.; Liu, W.; Lin, Y.; Sun, Z.; Ye, X.; Zheng, X.; Pan, H.; Zhu, J.; Chen, S.; Zhang, W.; Wei, S.; Lu, J.
Highly Active and Stable Metal Single-Atom C
J. Am. Chem. Soc. 2019 141
. , ,
14515 14519 10.1021/jacs.9b06482
– , .

31. Hahn, C.; Hatsukade, T.; Kim, Y. G.; Vailionis, A.; Baricuatro, J. H.; Higgins, D. C.; Nitopi, S. A.; Soriaga, M. P.; Jaramillo, T. F.
Engineering Cu surfaces for th
.
Proc. Natl. Acad. Sci. U. S. A.
2017 114 5918 5923 10.1073/pnas.1618935114
, , – , .

32. Sandberg, R. B.; Montoya, J. H.; Chan, K.; Nørskov, J. K. CO-CO coupling on Cu facets: Coverage, strain and field effects. Surf. Sci. 2016,
654 56 62 10.1016/j.susc.2016.08.006
, – , .

33. Li, H.; Wang, L.; Dai, Y.; Pu, Z.; Lao, Z.; Chen, Y.; Wang, M.; Zheng, X.; Zhu, J.; Zhang, W.; Si, R.; Ma, C.; Zeng, J.
Synergetic interaction between neighbouring platinum monomers in CO2 hyd
Nat. Nanotechnol. 2018 13 411 417
. , , – ,
10.1038/s41565-018-0089-z
.

34. Jin, Z.; Li, P.; Meng, Y.; Fang, Z.; Xiao, D.; Yu, G.
Understanding the inter-site distance effect in single-atom catalysts for oxygen electroreduction
Nat. Catal. 2021 4 615 622 10.1038/s41929-021-00650-w
. , , – , .

35. Bakandritsos, A.; Kadam, R. G.; Kumar, P.; Zoppellaro, G.; Medved, M.; Tucek, J.; Montini, T.; Tomanec, O.; Andryskova, P.; Drahos, B.;
Varma, R. S. Otyepka, M. Gawande, M. B. Fornasiero, P. Zboril, R.
; ; ; ;
Mixed-Valence Single-Atom Catalyst Derived from Functionalized Graphene
Adv. Mater.2019 31 e1900323 10.1002/adma.201900323
. , , , .

36. Wang, Y.; Park, B.; Paidi, V. K.; Huang, R.; Lee, Y.; Noh, K.-J.; Lee, K.-S.; Han, J. W.
Precisely Constructing Orbital Coupling-Modulated Dual-Atom Fe Pair Sites fo
ACS Energy Lett. 2022 7 640 649
. , , – ,
10.1021/acsenergylett.1c02446
.

37. Bai, L.; Hsu, C. S.; Alexander, D. T. L.; Chen, H. M.; Hu, X. A Cobalt-Iron Double-Atom Catalyst for the Oxygen Evolution Reaction.
J. Am. Chem. Soc.
2019 36 14190 14199 10.1021/jacs.9b05268
, , – , .

38. Shan, J.; Ye, C.; Chen, S.; Sun, T.; Jiao, Y.; Liu, L.; Zhu, C.; Song, L.; Han, Y.; Jaroniec, M.; Zhu, Y.; Zheng, Y.; Qiao, S.-Z.

Short-Range Ordered Iridium Single Atoms Integrated into Cobalt Oxide Spinel Structure for Highly Efficient Electrocatalytic Wate
J. Am. Chem. Soc. 2021 143 5201 5211 10.1021/jacs.1c01525
. , , – , .

39. Ding, K.; Gulec, A.; Johnson, A. M.; Schweitzer, N. M.; Stucky, G. D.; Marks, L. D.; Stair, P. C.
Identification of active sites in CO oxidation and water-gas shift over supported Pt catalysts
Science 2015 350 189 192 10.1126/science.aac6368
. , , – , .

40. Shan, J.; Liao, J.; Ye, C.; Dong, J.; Zheng, Y.; Qiao, S.-Z.
The Dynamic Formation from Metal-Organic Frameworks of High-Density Platinum Single-Atom Ca
Angew. Chem., Int. Ed.Angew. Chem. Int. Ed. 2022 61 e202213412
. , , ,
10.1002/anie.202213412
.

41. Zheng, Y.; Jiao, Y.; Jaroniec, M.; Qiao, S. Z.


Advancing the electrochemistry of the hydrogen-evolution reaction through combining experiment and theory
Angew. Chem., Int. Ed.Angew. Chem. Int. Ed. 2015 54 52 65 10.1002/anie.201407031
. , , – , .

42. Sheng, W.; Bivens, A. P.; Myint, M.; Zhuang, Z.; Forest, R. V.; Fang, Q.; Chen, J. G.; Yan, Y.
Non-precious metal electrocatalysts with high activity for hydrogen oxidation r
Energy Environ. Sci. 2014 7 1719 1724
. , , – ,
10.1039/C3EE43899F
.
43. Ranjan, C.; Hoffmann, R.; DiSalvo, F. J.; Abruna, H. C. D.
Electronic Effects in CO Chemisorption on Pt-Pb Intermetallic Surfaces: A Theoretical Study
J. Phys. Chem. C 2007 111 17357 17369 10.1021/jp0746603
. , , – , .

44. Hu, H.; Guan, B. Y.; Lou, X. W.


Construction of Complex CoS Hollow Structures with Enhanced Electrochemical Properties for Hybrid Supercapacitors
Chem 2016 1 102 113 10.1016/j.chempr.2016.06.001
. , , – , .

45. Koch, C. Determination of core structure periodicity and point defect density along dislocations:. Arizona State University, 2002.

46. Zhang, Q.; Zhang, L. Y.; Jin, C. H.; Wang, Y. M.; Lin, F.
CalAtom: A software for quantitatively analysing atomic columns in a transmission electron microscope image
Ultramicroscopy 2019 202 114 120 10.1016/j.ultramic.2019.04.007
. , , – , .

47. Zhang, Q.; Jin, C. H.; Xu, H. T.; Zhang, L. Y.; Ren, X. B.; Ouyang, Y.; Wang, X. J.; Yue, X. J.; Lin, F.
Multiple-ellipse fitting method to precisely measure the p
Micron 2018 113 99 104
. , , – ,
10.1016/j.micron.2018.06.016
.

48. Ravel, B.; Newville, M. ATHENA, ARTEMIS, HEPHAESTUS: data analysis for X-ray absorption spectroscopy using IFEFFIT.
J. Synchrotron Radiat.
2005 12 537 541 10.1107/S0909049505012719
, , – , .

49. Durst, J.; Simon, C.; Hasché, F.; Gasteiger, H. A.


Hydrogen Oxidation and Evolution Reaction Kinetics on Carbon Supported Pt, Ir, Rh, and Pd Electrocatalysts
J. Electrochem. Soc. 2014 162 F190 F203 10.1149/2.0981501jes
. , , – , .

50. Kresse, G.; Furthmiiller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996,
54 11169 11186 10.1103/PhysRevB.54.11169
, – , .

51. Kresse, G.; Furthmiiller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set.
Comput. Mater. Sci. 1996 6 15 50 10.1016/0927-0256(96)00008-0
, , – , .

52. Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865– 3868,
10.1103/PhysRevLett.77.3865
.

53. Tkatchenko, A.; Scheffler, M. Accurate molecular van der Waals interactions from ground-state electron density and free-atom reference data.
Phys. Rev. Lett. 2009 102 073005 10.1103/PhysRevLett.102.073005
, , , .

54. Nørskov, J. K.; Bligaard, T.; Logadottir, A.; Kitchin, J. R.; Chen, J. G.; Pandelov, S.; Stimming, U.
Trends in the Exchange Current for Hydrogen Evolution
J. Electrochem. Soc. 2005 152 J23 J26 10.1149/1.1856988
. , , – , .

TOC Graphic.

You might also like