Sei sulla pagina 1di 8

Article

Cite This: J. Am. Chem. Soc. XXXX, XXX, XXX−XXX pubs.acs.org/JACS

Ultrasmall Abundant Metal-Based Clusters as Oxygen-Evolving


Catalysts
Xin-Bao Han,†,# Xing-Yan Tang,†,# Yue Lin,‡,# Eduardo Gracia-Espino,∥ San-Gui Liu,†
Hai-Wei Liang,§ Guang-Zhi Hu,∥,⊥ Xin-Jing Zhao,† Hong-Gang Liao,† Yuan-Zhi Tan,*,†
Thomas Wagberg,*,∥ Su-Yuan Xie,† and Lan-Sun Zheng†

State Key Laboratory for Physical Chemistry of Solid Surfaces, and Department of Chemistry, College of Chemistry and Chemical
Engineering, Xiamen University, Xiamen, 361005, China

Hefei National Laboratory for Physical Sciences at the Microscale and §Department of Chemistry, University of Science and
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Technology of China, Hefei 230026, P. R. China



Department of Physics, Umeå University, Umeå 90187, Sweden

Downloaded via UNIV DE TARAPACA on December 25, 2018 at 18:35:24 (UTC).

Key Laboratory of Chemistry of Plant Resources in Arid Regions, State Key Laboratory Basis of Xinjiang Indigenous Medicinal
Plants Resource Utilization, Xinjiang Technical Institute of Physics and Chemistry, Chinese Academy of Sciences, Urumqi 830011,
China
*
S Supporting Information

ABSTRACT: The oxygen evolution reaction is a crucial step in water


electrolysis to develop clean and renewable energy. Although noble metal-
based catalysts have demonstrated high activity for the oxygen evolution
reaction, their application is limited by their high cost and low availability.
Here we report the use of a molecule-to-cluster strategy for preparing
ultrasmall trimetallic clusters by using the polyoxometalate molecule as a
precursor. Ultrafine (0.8 nm) transition-metal clusters with controllable
chemical composition are obtained. The transition-metal clusters enable
highly efficient oxygen evolution through water electrolysis in alkaline media,
manifested by an overpotential of 192 mV at 10 mA cm−2, a low Tafel slope of
36 mV dec−1, and long-term stability for 30 h of electrolysis. We note,
however, that besides the excellent performance as an oxygen evolution catalyst, our molecule-to-cluster strategy provides a
means to achieve well-defined transition-metal clusters in the subnanometer regime, which potentially can have an impact on
several other applications.

■ INTRODUCTION
Electrocatalytic water splitting is widely considered to be a
roles in governing their OER activities. Small metal clusters,
with controlled composition and size, have, however, proven to
promising and sustainable approach to the production of clean be particularly difficult to synthesize and stabilize. Herein we
H2 fuel and O2.1−3 Water oxidation, represented by the half- report a novel and efficient synthesis of transition-metal
reaction 2H2O → 4H+ + 4e− + O2, constitutes a kinetic clusters (TMCs) from well-defined polyoxometalate (POM)
bottleneck in water splitting.4,5 Therefore, the development of molecule precursors via a molecule-to-cluster strategy. The
highly efficient water oxidation catalysts is crucial in the prepared TMCs are characterized by their ultrafine size (0.8 ±
renewable energy field.6−8 Noble-metal catalysts such as 0.2 nm) and tunable chemical composition, all of which were
iridium and ruthenium oxides are the highly efficient oxygen inherited from their parent POMs. Our experimental and
evolution reaction (OER) electrocatalysts in acidic media but theoretical results demonstrate that well-defined Fe-doping of
suffer from high-cost and relative scarcity9 and also require a POMs allow us to precisely modulate the chemical
substantial overpotential to reach the desired current density of compositions of the corresponding TMCs, resulting in a
≥10 mA cm−2 for OER in alkaline media.10 In alkaline media, significant enhancement of their OER activities. The best
on the other hand, a number of research groups have TMC catalyst exhibited an overpotential of 192 mV at a
demonstrated the use of earth-abundant transition-metal current density of 10 mA cm−2, had a low Tafel slope of 36 mV
(TM)-based catalysts, including oxides,11,12 oxyhydroxides,13 dec−1, and was shown to remain stable for 6000 cyclic
borides,14 phosphides,15,16 and molecular complexes,17,18 but voltammetry (CV) cycles and 30 h of electrolysis. The
there is still room for significant improvement in OER systems molecule-to-cluster approach applied here could serve as a
to obtain good energy efficiency and cost effectiveness in the
alkaline electrolysis systems. More importantly, these studies Received: August 23, 2018
suggested that the catalyst size and composition play crucial Published: December 12, 2018

© XXXX American Chemical Society A DOI: 10.1021/jacs.8b09076


J. Am. Chem. Soc. XXXX, XXX, XXX−XXX
Journal of the American Chemical Society Article

general and effective strategy for establishing ultrasmall TM Moreover, HAADF-STEM imaging of the clusters demon-
clusters where practically all atoms reside at the surface and strated a clear speckle pattern (Figure 2e,f). The speckle
therefore are tunable for high-efficiency oxygen evolution and pattern is created by the positive correlation of the spot
other important electrocatalytic reactions. brightness in HAADF-STEM images and the atomic number

■ RESULTS AND DISCUSSION


Synthesis of TMCs. A schematic view of catalyst
of the visualized atoms,21,22 indicating a well-mixed W−Co
system seen as contiguous brighter and dimmer spots,
respectively. These results are also supported by our theoretical
preparation is presented in Figure 1a−d. The self-assembly simulations where W−Co clusters in the range of 0.7−0.9 nm
exhibit a chemical order parameter that is close to zero,
meaning that no significant segregation is observed.23
Furthermore, STEM-coupled energy-dispersive X-ray spectros-
copy (STEM-EDXS) elemental mapping of 1-CoW (Figure
2g) also showed that Co and W were homogeneously
distributed on the carbonic support. It is noteworthy to
indicate that such small TMCs (0.8 nm) have ∼90% of the
atoms located at the surface and are significantly under-
coordinated (Tables S1 and S2), indicating that they are
feasibly surface-passivated by O/OH as suggested by DFT,
discussed later. We therefore suggest that the catalytically
active sites of TMCs are closely related to metal oxy-
hydroxides, in agreement with previous studies,24−26 which is
also confirmed by the X-ray photoelectron spectroscopy (XPS)
analysis of the chemical valence of Co, Fe, W, and O atoms in
TMCs (Figures S4−S7, Supporting Information).
OER Performance of 1-CoW. The OER performance of
Figure 1. Schematic illustration of synthesizing the TMCs. (a) Ball the carbon-supported TMCs was compared to that of IrO2 by
representation of the structure of POMs 1−3. (b) Self-assembly of
POMs and EDA-C60 by electrostatic interaction. (c) Schematic
using a typical three-electrode system in 1 M KOH solution at
structure of the POMs/EDA-C60 composites after freeze-drying. (d) a scan rate of 5 mV s−1. Linear sweep voltammetry (LSV) was
TMC catalysts on a carbonic support. O, gray; Co/Fe, violet; W, red; employed to obtain the polarization curves of the electrode. All
and P, yellow. potentials were calibrated in reference to the reversible
hydrogen electrode (RHE). Before measurements, the as-
prepared electrodes were repeatedly swept in the electrolyte
of [{Co4(OH)3PO4}4(SiW9O34)4]32− (POM 1) (Figure 1a and until a steady voltammogram curve was obtained, and in this
Figure S1a,b in Supporting Information (SI)) and ethylenedi- process, the TMCs would be oxidized. Each catalyst was
amine-grafted C60 (EDA-C60), driven by electrostatic inter- coated onto carbon cloth at a density of 0.64 mg cm−2, and the
action, resulted in the formation of POM 1/EDA-C60 hybrid
generated OER currents were measured. As shown in Figure
composites (Figure 1b). Images obtained by in situ liquid cell
3a, 1-CoW produced a small overpotential of 240 mV at a
transmission electron microscopy (TEM)19,20 showed that the
current density of 10 mA cm−2 (Table 1, all current densities
composites contained discrete POMs (Figure 2a), suggesting
were based on the projected geometric area), which was much
that a satisfactory encapsulation of POMs by EDA-C60 is
smaller than that of IrO2 (305 mV) under the same conditions.
achieved, ultimately preventing self-aggregation. Similarly,
TEM imaging of freeze-dried hybrid composites provided Enhancing OER Performance by Fe Doping. Doping
additional visual evidence for the embedding of discrete POMs additional elements into metallic nanoparticles has proven to
in the EDA-C60 matrix (Figure S2, Supporting Information). be an effective strategy for enhancing the performance of
We speculate that the spatial confinement and single- metallic catalysts.24,27,28 The molecule-to-cluster protocol
molecule dispersion of POMs in the EDA-C60 matrix favor the enabled convenient chemical doping of TMCs by modifying
formation of TMCs during thermal annealing. Indeed, the structure of the POM precursors. By this strategy, we
annealing of the hybrid composite under a 5% H2/Ar synthesized POM 2 [{Fe2Co2(OH)3PO4}4(SiW9O34)4]24− and
atmosphere for 3 h at 900 °C led to the transformation of POM 3 [{FeCo3(OH)3PO4}4(SiW9O34)4]28− (Figure S1,
the embedded POMs into TMCs and the formation of a Supporting Information) by doping POM 1 in situ with
carbonic support from EDA-C60 (Figure 1d). The resulting different amounts of Fe. POM 2 and POM 3 were shown by
material was labeled 1-CoW. single-crystal X-ray crystallography to share structures almost
The morphology of 1-CoW was studied by scanning identical to that of POM 1, albeit with different extents of Co
electron microscopy (SEM) and TEM. The SEM image of displacement with Fe. We then generated two differently Fe-
1-CoW showed curved sheets with few visible nanoparticles on doped TMCs on a carbon support, designated as 2-CoFeW
the surfaces (Figure 2b) which was consistent with the TEM and 3-CoFeW, from POM 2 and POM 3 by the same synthesis
image (Figure 2c). In contrast, abundant clusters were route as that for 1-CoW, respectively. As expected, both 2-
observed to spread over the carbonic supports in images CoFeW and 3-CoFeW were shown by X-ray diffraction (XRD)
generated by TEM (Figure 2d) and high-angle annular dark- patterns, SEM, TEM, and STEM-HAADF characterizations
field scanning transmission electron microscopy (HAADF- (Figures S8−S11, Supporting Information) to share morphol-
STEM) (Figure 2e,f and Figure S3 in Supporting Information). ogies similar to that of 1-CoW and were confirmed by STEM-
The size distribution of the clusters was in the range of 0.6 to EDXS elemental mapping to be uniformly doped with Fe
3.0 nm, with a mean diameter of 0.8 nm (Figure 2f, inset). (Figure S12, Supporting Information).
B DOI: 10.1021/jacs.8b09076
J. Am. Chem. Soc. XXXX, XXX, XXX−XXX
Journal of the American Chemical Society Article

Figure 2. Morphology of TMCs on a carbonic support. (a) In-situ liquid cell TEM image of the solution in which POM 1 self-assembles with EDA-
C60. (b) SEM image of 1-CoW. (c and d) TEM images of 1-CoW. (e and f) HAADF-STEM images of 1-CoW showing nanoparticle speckling of
TMCs. The inset image in f illustrates the corresponding size distribution of the TMCs in 1-CoW. (g) HAADF-STEM image and STEM-EDXS
element mapping of 1-CoW.

Both 2-CoFeW and 3-CoFeW showed that the OER importance of fine tuning the chemical composition of the
overpotentials decreased by 16 and 35 mV, respectively, at a TMC to improve its catalytic performance in OERs.
current density of 10 mA cm−2 when compared with 1-CoW. Substrate Influence on OER Performance. To assess
This demonstrated that well-balanced Fe doping could the effect of the substrate on the catalytic efficiency of the
improve the catalytic performance of the corresponding electrode, we loaded the carbon-supported TMC catalysts at a
TMCs (Figure 3a and Table 1), but it is noteworthy that density of 0.64 mg cm−2 onto gold foam (a nickel foam that
the Fe ratio in 2-CoFeW already is “on the down slope” of was covered with gold to avoid any spurious effects arising
OER activity, as manifested by an overpotential of 224 mV at from the interaction of the catalyst with Ni), which has a
10 mA cm−2 (19 mV higher than that of 3-CoFeW). This greater conductivity than carbon cloth. This switch in the
behavior matches the observations in our DFT studies where substrate significantly boosted the catalytic activities of all
an Fe content that is too high leads to an adsorption energy of three TMCs without affecting the beneficial impact of Fe
doping (Figure 3c). Specifically, 3-CoFeW had a remarkably
O-intermediates that is too high and thereby an increase in the
low overpotential of 192 mV at 10 mA cm−2 on the gold foam
overall OER overpotential. In addition, insight into the kinetic
(Table 1). The enhanced conductivity of the substrate also led
behaviors of all TMC catalysts in OERs was obtained by
to a substantial improvement in the OER kinetics, as evidenced
analyzing their Tafel slope. As such, the superior OER by the decrease in the Tafel slope for all three catalysts (Figure
performance of 3-CoFeW was supported by its smaller Tafel 3d).
slope (38 mV dec−1, Figure 3b) compared to those of 1-CoW In addition, electrochemical impedance spectroscopy (EIS)
(53 mV dec−1) and 2-CoFeW (43 mV dec−1). The turnover of different samples was performed to provide further insight
frequency (TOF) was also calculated at an overpotential of into the electrode kinetics within the OER process. As shown
250 mV in 1 M KOH solution, assuming that all of the Co and in Figure S13 and Table S5 in the Supporting Information, the
Fe atoms account for the catalytic activity of the OER (Table solution resistance (Rs) values of 1-CoW, 2-CoFeW, and 3-
S3, Supporting Information). 3-CoFeW has a larger TOF of CoFeW coated on gold foam are similar, while the charge-
0.377 s−1 compared to 0.114 and 0.259 s−1 for 1-CoW and 2- transfer resistance (Rct) of 3-CoFeW coated on gold foam
CoFeW (Table S4, Supporting Information), respectively. fitted according to the high-frequency semicircle was only 0.93
Taken together, these data lent credence to the critical Ω, which is smaller than for 1-CoW (2.2 Ω) and 2-CoFeW
C DOI: 10.1021/jacs.8b09076
J. Am. Chem. Soc. XXXX, XXX, XXX−XXX
Journal of the American Chemical Society Article

Figure 3. OER polarization curves of catalysts loaded on (a) carbon cloth or (c) gold foam at a scan rate of 5 mV s−1, without iR correction. Tafel
plots for different catalysts loaded on (b) carbon cloth or (d) gold foam in 1 M KOH solution.

Table 1. Comparison of Electrocatalytic Activity at a (ICP-MS) measurements of the three catalysts to detect the
Current Density of 10 mA cm−2 amount of Co, Fe, or W of the reaction solution after a long-
term cycling test. The data of ICP-MS indicated that there is
on carbon cloth on gold foam
less than 0.56% of W; 2.0% of Fe and 1.3% of Co may leach
samples overpotential/mV overpotential/mV out of the catalyst (Table S6, Supporting Information). The
1-CoW 240 (±5) 222 (±3) enhanced operational stability of such small clusters can be
2-CoFeW 224 (±3) 205 (±2) attributed to the large cohesive energy of tungsten as well as
3-CoFeW 205 (±3) 192 (±2) the strength of the W−C bond probably found at the cluster−
IrO2 305 (±4) carbonic substrate interface, and similar results have been
observed for PdW/C systems.29 These results revealed the
good electrochemical stability of the TMC catalysts. Although
(1.4 Ω). Such a low Rct value of 3-CoFeW suggested faster OER catalysts supported on carbonic supports are widely used
charge transfer kinetics compared to those of 1-CoW and 2- for fundamental study,30−34 it is worth noting that carbonic
CoFeW. When 3-CoFeW was coated on carbon cloth, Rs and supports suffer from carbon corrosion for practical application
Rct skyrocketed from 2.4 to 5.2 Ω and 0.93 to 9.9 Ω, under prolonged anodic bias in an alkaline electrolyzer.35,36 We
respectively, which was consistent with the disparity in catalytic
are continuing our efforts to prepare TMC catalysts supported
activity on different substrates.
on more stable supports for practical applications.
Stability. We next evaluated the electrochemical stability of
Theoretical Description of the OER. The catalytic
the carbon-supported TMC catalysts by long-term cycling. We
first ran water oxidation on the catalyst deposited on carbon activity of CoFeW clusters was also investigated by DFT.
cloth (Figure 4a) and gold foam (Figure 4b) under a long-term Three different clusters sizes were prepared by a modified
cycling test in 1 M KOH solution. Gratifyingly, the OER basin-hopping Monte Carlo (BHMC) method with average
polarization curve of 3-CoFeW remained almost unchanged diameters of 0.7, 0.8, and 0.9 nm and chemical compositions of
throughout the experiment and exhibited a negligible potential W13Co16, W23Co16, and W36Co16, respectively (Figures S17
change at 10 mA cm−2 after 6000 CV cycles, despite the and S18, Supporting Information). The resulting clusters
relatively harsh experimental conditions. Chronoamperometric exhibited an average coordination number of ∼7.2, irrespective
responses (i−t) of 3-CoFeW coated on carbon cloth (Figure of the particle size. Interestingly, the addition of Fe does not
4c) and gold foam (Figure 4d) retained their activity for 30 h have a significant effect on the average coordination number.
of long-term electrolysis. The STEM-HAADF and high- However, the individual coordination number of Co is
resolution transmission electron microscope (HRTEM) increased in the presence of Fe (Table S2, Supporting
images of 3-CoFeW after the OER were found to be largely Information), suggesting a Co migration to the cluster core,
identical to those of the freshly synthesized catalyst (Figures which in a 0.8 nm particle corresponds to ∼4 atoms, while Fe
S14 and S15, Supporting Information). In addition, XRD is more often found at the surface. In any case, the severe
(Figure S16) measurements of catalysts showed almost no undercoordination results in large adsorption energies for OER
change before and after the long-term OER test. Finally, we intermediates (*O, *OH, and *OOH) degrading the catalytic
performed the inductively coupled plasma mass spectrometry activity, but it also indicates the feasibility of forming an
D DOI: 10.1021/jacs.8b09076
J. Am. Chem. Soc. XXXX, XXX, XXX−XXX
Journal of the American Chemical Society Article

Figure 4. Polarization curves of 3-CoFeW loaded on (a) carbon cloth and (b) gold foam before and after 6000 potential cycles in 1 M KOH
solution. Chronoamperometric responses (i−t) of 3-CoFeW coated on (c) carbon cloth and (d) gold foam for 30 h at a constant applied potential
of 1.46 V versus RHE for OER.

oxyhydroxide surface layer and is confirmed by the XPS


analysis of Co, Fe, W, and O atoms in TMCs (Figures S4−S7,
Supporting Information).
The OER activity was investigated only on clusters with an
average diameter of 0.8 nm because these were the most
abundant experimentally. The clusters were initially passivated
with hydroxyl groups until the changes in free energy were
zero. During the passivation process, several water molecules
were formed and removed from the system, resulting in an
OH/O surface passivation (Figure S7, Supporting Informa-
tion) with ratios of 4.2, 4.9, and 6.0 and final compositions of
W23Co16O9(OH)38, W23Co12Fe4O8(OH)39, and
W23Fe16O7(OH)42, respectively. The differences in the OH/ Figure 5. Theoretical catalytic activity for metal oxyhydroxide
O ratio already show the stronger interaction that Fe exhibits clusters. OER activity volcano plot for W23CowFexOy(OH)z clusters.
over the hydroxyl groups compared to a weaker OH
interaction when only Co atoms are present. The catalytic W23Co12Fe4O8(OH)39 cluster. However, we did notice differ-
activity was investigated using the four-electron reaction ences when considering the coordination environment, and
pathway37 and the free-energy difference between the *O those Fe/Co atoms with complete octahedral coordination
and *OH intermediate (ΔGO − ΔGOH) as a universal exhibit the lowest ηOER irrespective of the neighbors’ type.
descriptor.37,38 The latter results in a volcano-like plot such Similar results were observed when using a different set of
as the one seen in Figure 5, where the highest OER activity is atomic models constructed using the γ-FeOOH framework
expected for those materials with a free-energy difference of instead of the BHMC method (W 20 Co 16 O 30 (OH) 51 ,
ΔGO − ΔGOH = 1.6 eV, while those at a lower or higher energy W20Co12Fe4O33(OH)44, and W20Fe16O30(OH)51). See Figures
difference will exhibit a lower OER activity. In our case, the S19 and S20 in Supporting Information for further details. Our
clusters are mostly asymmetric with a large variety of active results are in line with previous publications where CoOOH
sites with diverse OER activity. As seen in Figure 5, all three exhibit a weak interaction with oxygenated species, whereas it
clusters exhibit sites with high activity reflecting the perform- is too strong on FeOOH;24 therefore, a combination of these
ance seen experimentally. However, it can be seen that two ultimately results in a material with an optimal adsorption
W23Co16O9(OH)38 favors weak adsorption energies seen as energy.
ΔGO − ΔGOH > 1.6 eV, with the best site having ηOER = 0.45
V, while W23Fe16O7(OH)42 tends toward a higher adsorption
energy with ΔGO − ΔGOH < 1.6 eV and the lowest is ηOER =
■ CONCLUSIONS
A series of carbon-supported TMCs have been successfully
0.39 V. Trimetallic W23Co12Fe4O8(OH)39 mostly lies around prepared by a molecule-to-cluster strategy and employed as
ΔGO − ΔGOH = 1.6 eV with an optimum ηOER of 0.37 V. It is electrocatalysts for OER in 1 M KOH. These catalysts
noteworthy to mention that we did not observe any significant exhibited high electrocatalytic activity and long-term catalytic
difference between Co and Fe in the mixed stability. In particularly, 3-CoFeW generated a an overpotential
E DOI: 10.1021/jacs.8b09076
J. Am. Chem. Soc. XXXX, XXX, XXX−XXX
Journal of the American Chemical Society Article

of 192 mV at a current density of 10 mA cm−2 when loaded on days at 80 °C under an Ar atmosphere. The excess EDA was removed
gold foam as the working electrode, making it one of the most from the crude reaction mixture by rotary evaporation, followed by
active OER catalysts reported to date. Even more importantly, the addition of 10 mL of deionized H2O to dissolve the residuals. The
our molecule-to-cluster strategy presented here provides a target product was precipitated by adding 500 mL of acetone, washed
several times with acetone, vacuum filtered, and then dried at 60 °C
unique approach to fine tuning the catalytic properties of overnight.
ultrafine clusters by selective and controlled surface doping to Synthesis of the TMCs Catalysts. EDA-C60 (6 mL, 5 mg·mL−1)
address the adsorption properties of reaction intermediates at was added to a 150 mL beaker and further diluted with distilled water
the surface of the TM clusters. We note that our approach has to a final concentration of 0.5 mg·mL−1. The solution was vigorously
implications far beyond oxygen evolution reactions. stirred, and its pH was recorded to be 3.57. Next, POM 1 (20 mL, 0.5


mg·mL−1) was added dropwise. The resultant solution was vigorously
stirred for 1 h at room temperature and further ultrasonicated for 30
EXPERIMENTAL SECTION min. The 1-CoW catalyst was obtained by freeze drying the reaction
Materials and Methods. All reagents and solvents were used of solution and then annealing under a 5% H2/Ar atmosphere for 3 h at
commercial grade, without further purification. The SEM images were 900 °C. The preparation of 2-CoFeW and 3-CoFeW followed a
obtained on a Hitachi S-4800 scanning electron microscope at an similar protocol to that of 1-CoW except that different starting
accelerating voltage of 15 kV. TEM and HAADF-STEM images were materials were used. Specifically, 20 mL of 0.5 mg·mL−1 POM 2 and
collected on a JEOL ARM-200F field-emission transmission electron POM 3 was used in place of POM 1 for the syntheses of 2-CoFeW
microscope operating at a 200 kV accelerating voltage. STEM-EDXS and 3-CoFeW, respectively.
images were collected on a Titan Themis transmission electron X-ray Structural Analysis. Single-crystal data sets and unit cells
microscope (FEI). The in situ TEM observation was carried out using were collected on an Agilent Super Nova X-ray single-crystal
an FEI Tecnai F20 transmission electron microscopy with a CMOS- diffractometer using Cu Kα (λ = 1.54184 Å) microfocus X-ray
based TemCam-XF416 camera (TVIPS GmbH, Germany). sources at 100 K. The structures of POM 2 and POM 3 were solved
Synthesis of POM 1. POM 1 was synthesized according to a by the direct method and refined by the full-matrix least-squares
previously described method.39 Briefly, CoCl2·6H2O (0.78 g, 3.26 method on F2 using the SHELX-2014 crystallographic software
mmol) was dissolved in 40 mL of distilled water, followed by the package.43,44 The crystal data and structure refinements for POM 2
addition of Na10[α-SiW9O34]·18H2O40 (1.18 g, 0.40 mmol). The and POM 3 is shown in Table S7 in Supporting Infromation.
reaction mixture was stirred until a clear purple solution was obtained. Theoretical Calculations. Ab initio calculations were performed
Na3PO4·12H2O (0.60 g, 1.58 mmol) was then added, and the pH was by means of density functional theory as implemented in the Siesta
maintained in the range of 8.5−9.0 with 1.0 M KOH(aq). The code.45 The generalized gradient approximation and the revised
resultant turbid solution was stirred for 3 h at room temperature, and model of Perdew, Burke, and Ernzerhof46 were used to describe the
the purple precipitate was removed by filtration. The obtained filtrate exchange and correlation functional. The valence electrons are
was subsequently mixed with 5 mL of 1.0 M KCl solution, stirred for represented by a linear combination of pseudoatomic numerical
another 30 min, and filtered into a 50 mL beaker to allow slow orbitals using a double-ζ polarized basis. An energy cutoff of 250 Ry
evaporation at room temperature for 1 week, resulting in the was used for the charge and potential integration in real space. The
precipitation of 1 as dark-purple crystals. The crystalline product was Co/Fe content of W−Co and W−Fe−Co clusters was selected to
collected by filtration, washed with cold water, and air dried. emulate precursors POM 1 and POM 3. Additional W−Fe clusters
Synthesis of POM 2. CoCl2·6H2O (0.18 g, 0.75 mmol) and were built to study the extreme case where all Co atoms have been
Fe(NO3)3·9H2O (0.34 g, 0.84 mmol) were dissolved in 50 mL of replaced by Fe. The particle size was adjusted by varying the W
distilled water. Na10[α-SiW9O34]·18H2O (1.18 g, 0.40 mmol) was content until average diameters of 7, 8, and 9 Å were obtained,
then added, and the mixture was stirred for 15 min. Next, Na3PO4· resulting in W13Co16, W23Co16, and W36Co16 clusters, respectively. A
12H2O (0.60 g, 1.58 mmol) was added, and the pH was maintained in modified BHMC method was used to find an energetically favorable
the range of 8.0−8.5 with 1.0 M NaOH(aq). The resultant turbid atomic configuration.47 In brief, an initial structure was randomly
solution was stirred for 2 h at 80 °C, and the precipitate was removed created and geometrically optimized using the conjugate gradient
by filtration. The obtained filtrate was subsequently mixed with 5 mL method, and then the positions of two atoms with different chemical
of 1.0 M NaCl solution, stirred for another 30 min, and transferred to elements were randomly swapped, followed by a geometrical
a 100 mL beaker to allow slow evaporation at room temperature over optimization. A total of 80 Monte Carlo steps were carried out, and
1 or 2 weeks. Eventually, product 2 precipitated as light-brown the lowest-energy configuration was used for the OER study. A
crystals, which were washed with cold water and air dried. Overall, thermal energy kBT of 0.2 eV was used as the criterion for accepting
239 mg of 2 was obtained, and the yield was calculated to be 19.35%. or rejecting a configuration. Finally, the optimized clusters were
Anal. Calcd (%): Na, 4.68; Fe, 3.80; Co, 4.00; W, 56.15. Found: Na, passivated with hydroxyl groups until the change in free energy was
4.76; Fe,3.73; Co, 3.92; W, 56.41. zero, and any created water molecule was removed from the system.
Synthesis of POM 3. CoCl2·6H2O (0.36 g, 1.51 mmol) and The final OH/O ratio was 4.2, 4.9, and 6.0 for W23Co16, W23Co12Fe4,
Fe(NO3)3·9H2O (0.23 g, 0.57 mmol) were dissolved in 50 mL of and W23Fe16, respectively.
distilled water. Na10[α-SiW9O34]·18H2O (1.18 g, 0.40 mmol) was The catalytic activity was studied using the computational
then added, and the mixture was stirred for 15 min. Next, Na3PO4· hydrogen-electrode model48 by setting the reference potential equal
12H2O (0.60 g, 1.58 mmol) was added, and the pH was maintained in to the standard hydrogen electrode. In this way, the chemical
the range of 8.0−8.5 with 1.0 M NaOH(aq). The resultant turbid potential of a proton−electron pair (H+ + e−) in solution is equal to
solution was stirred for 2 h at 80 °C, and the precipitate was removed half of the chemical potential of a gas-phase H2 molecule. Under these
by filtration. The obtained filtrate was mixed with 5 mL of a 1.0 M conditions, the explicit treatment of solvated protons is avoided. The
NaCl solution, stirred for another 30 min, and subsequently adsorption free energy is calculated to be ΔG(i) = ΔE(i) + ΔZPE(i) −
transferred to a 100 mL beaker to allow slow evaporation at room TΔS(i), where ΔE(i) is the adsorption energy of intermediate i
temperature over 1 or 2 weeks. Product 3 precipitated as reddish- evaluated using H2 and H2O as reference states.49 ΔZPE is the change
brown crystals, which were washed with cold water and air dried. in zero-point energies, T is the temperature (298.15 K), and ΔS is the
Overall, 201 mg of 3 was obtained, and the yield was calculated to be change in entropy, with values taken from ref 48. The oxygen
16.27%. Anal. Calcd (%): Na, 5.38; Fe, 1.87; Co, 5.91; W, 55.24. adsorption energy was estimated by using the scaling relation for
Found: Na, 5.45; Fe,1.82; Co, 5.78; W, 55.49. transition-metal oxides given by ΔEOH = 0.55ΔEO − 0.48 in eV.38
Synthesis of EDA-C60. EDA-C60 was synthesized by a modified The universal descriptor, (ΔGO − ΔGOH), for transition metals and
method.41,42 Typically, C60 (500 mg) and freshly distilled EDA (500 OER was used to calculate the theoretical overpotential (ηtheory),
mL) were added to a three-necked flask and stirred in an oil bath for 4 where under standard conditions the overpotential is calculated to be

F DOI: 10.1021/jacs.8b09076
J. Am. Chem. Soc. XXXX, XXX, XXX−XXX
Journal of the American Chemical Society Article

ηtheory = {max[(ΔGO − ΔGOH), 3.2 eV − (ΔGO − ΔGOH)]/e} − 1.23 (7) Blasco-Ahicart, M.; Soriano-López, J.; Carbó, J. J.; Poblet, J. M.;
V.37 Galan-Mascaros, J. R. Polyoxometalate electrocatalysts based on earth

■ ASSOCIATED CONTENT
* Supporting Information
S
abundant metals for efficient water oxidation in acidic media. Nat.
Chem. 2018, 10, 24−30.
(8) Zhang, T.; Wang, C.; Liu, S.; Wang, J.-L.; Lin, W. A biomimetic
copper water oxidation catalyst with low overpotential. J. Am. Chem.
These materials are available free of charge via the Internet at Soc. 2014, 136, 273−281.
The Supporting Information is available free of charge on the (9) Lei, C.; Chen, H.; Cao, J.; Yang, J.; Qiu, M.; Xia, Y.; Yuan, C.;
ACS Publications website at DOI: 10.1021/jacs.8b09076. Yang, B.; Li, Z.; Zhang, X.; Lei, L.; Abbott, J.; Zhong, Y.; Xia, X.; Wu,
X-ray crystallography, supplementary figures and tables, G.; He, Q.; Hou, Y. Fe−N4 sites embedded into carbon nanofiber
and IR spectra (PDF) integrated with electrochemically exfoliated graphene for oxygen
evolution in acidic medium. Adv. Energy Mater. 2018, 8, 1801912.
Co8Fe8Na13O233P4Si4W36 (CIF) (10) Pi, Y.; Shao, Q.; Wang, P.; Lv, F.; Guo, S.; Guo, J.; Huang, X.
C8Co12Fe4Na19O232P4Si4W36 (CIF) Trimetallic oxyhydroxide coralloids for efficient oxygen evolution


electrocatalysis. Angew. Chem., Int. Ed. 2017, 56, 4502−4506.
(11) Suntivich, J.; May, K. J.; Gasteiger, H. A.; Goodenough, J. B.;
AUTHOR INFORMATION Shao-Horn, Y. A perovskite oxide optimized for oxygen evolution
Corresponding Authors catalysis from molecular orbital principles. Science 2011, 334, 1383−
*yuanzhi_tan@xmu.edu.cn. 1385.
*thomas.wagberg@physics.umu.se (12) Tüysüz, H.; Hwang, Y. J.; Khan, S. B.; Asiri, A. M.; Yang, P.
Mesoporous Co3O4 as an electrocatalyst for water oxidation. Nano
ORCID
Res. 2013, 6, 47−54.
Yue Lin: 0000-0001-5333-511X (13) Subbaraman, R.; Tripkovic, D.; Chang, K.-C.; Strmcnik, D.;
Eduardo Gracia-Espino: 0000-0001-9239-0541 Paulikas, A. P.; Hirunsit, P.; Chan, M.; Greeley, J.; Stamenkovic, V.;
Thomas Wagberg: 0000-0002-5080-8273 Markovic, N. M. Trends in activity for the water electrolyser reactions
Author Contributions on 3d M(Ni,Co,Fe,Mn) hydroxyoxide catalysts. Nat. Mater. 2012, 11,
# 550−557.
These authors contributed equally to this work. (14) Dincă, M.; Surendranath, Y.; Nocera, D. G. Nickel-borate
Notes oxygen-evolving catalyst that functions under benign conditions. Proc.
The authors declare no competing financial interest. Natl. Acad. Sci. U. S. A. 2010, 107, 10337−10341.

■ ACKNOWLEDGMENTS
This work was supported by the National Natural Science
(15) Mendoza-Garcia, A.; Zhu, H.; Yu, Y.; Li, Q.; Zhou, L.; Su, D.;
Kramer, M. J.; Sun, S. Controlled anisotropic growth of Co-Fe-P from
Co-Fe-O nanoparticles. Angew. Chem., Int. Ed. 2015, 54, 9642−9645.
(16) Li, D.; Baydoun, H.; Verani, C. N.; Brock, S. L. Efficient water
Foundation of China (21771155, 21721001, and 11874334), oxidation using CoMnP nanoparticles. J. Am. Chem. Soc. 2016, 138,
the Ministry of Science and Technology of China 4006−4009.
(2014CB845603 and 2017YFA0204902), and Anhui Provin- (17) Barnett, S. M.; Goldberg, K. I.; Mayer, J. M. A soluble copper−
cial Natural Science Foundation (1708085MA06). E.G.-E. bipyridine water-oxidation electrocatalyst. Nat. Chem. 2012, 4, 498−
acknowledges the Carl Tryggers Foundation (CTS-16-161) for 502.
financial support. The theoretical simulations were performed (18) Jiang, X.; Li, J.; Yang, B.; Wei, X.-Z.; Dong, B.-W.; Kao, Y.;
on resources provided by the Swedish National Infrastructure Huang, M.-Y.; Tung, C.-H.; Wu, L.-Z. A bio-inspired Cu4O4 cubane:
for Computing (SNIC) at the High Performance Computing effective molecular catalysts for electrocatalytic water oxidation in
Center North (HPC2N). Dr. Dong-Fei Lu, Dr. Yu-Hui Luo, aqueous solution. Angew. Chem., Int. Ed. 2018, 57, 7850−7854.
(19) Liao, H. G.; Cui, L.; Whitelam, S.; Zheng, H. Real-time imaging
Prof. Xin-Long Wang, and Prof. Da-Qi Wang are gratefully of Pt3Fe nanorod growth in solution. Science 2012, 336, 1011−1014.
acknowledged for their help during the refinement of the X-ray (20) Liao, H.-G.; Zherebetskyy, D.; Xin, H.; Czarnik, C.; Ercius, P.;
crystal structures.


Elmlund, H.; Pan, M.; Wang, L.-W.; Zheng, H. Facet development
during platinum nanocube growth. Science 2014, 345, 916−919.
REFERENCES (21) Hueso, J. L.; Sebastián, V.; Mayoral, A.; Usón, L.; Arruebo, M.;
(1) McCrory, C. C. L.; Jung, S.; Ferrer, I. M.; Chatman, S. M.; Santamaría, J. Beyond gold: rediscovering tetrakis-(hydroxymethyl)-
Peters, J. C.; Jaramillo, T. F. Benchmarking hydrogen evolving phosphonium chloride (THPC) as an effective agent for the synthesis
reaction and oxygen evolving reaction electrocatalysts for solar water of ultra-small noble metal nanoparticles and Pt-containing nanoalloys.
splitting devices. J. Am. Chem. Soc. 2015, 137, 4347−4357. RSC Adv. 2013, 3, 10427−10433.
(2) Kornienko, N.; Resasco, J.; Becknell, N.; Jiang, C.-M.; Liu, Y.-S.; (22) Wong, A.; Liu, Q.; Griffin, S.; Nicholls, A.; Regalbuto, J. R.
Nie, K.; Sun, X.; Guo, J.; Leone, S. R.; Yang, P. Operando Synthesis of ultrasmall, homogeneously alloyed, bimetallic nano-
spectroscopic analysis of an amorphous cobalt sulfide hydrogen particles on silica supports. Science 2017, 358, 1427−1430.
evolution electrocatalyst. J. Am. Chem. Soc. 2015, 137, 7448−7455. (23) Aguilera-Granja, F.; Vega, A.; Rogan, J.; Andrade, X.; García, G.
(3) Gao, M. R.; Zheng, Y.-R.; Jiang, J.; Yu, S.-H. Pyrite-type Theoretical investigation of free-standing CoPd nanoclusters as a
nanomaterials for advanced electrocatalysis. Acc. Chem. Res. 2017, 50, function of cluster size and stoichiometry in the Pd-rich phase:
2194−2204. Geometry, chemical order, magnetism, and metallic behavior. Phys.
(4) Hunter, B. M.; Gray, H. B.; Müller, A. M. Earth-abundant Rev. B: Condens. Matter Mater. Phys. 2006, 74, 224405.
heterogeneous water oxidation catalysts. Chem. Rev. 2016, 116, (24) Zhang, B.; Zheng, X.; Voznyy, O.; Comin, R.; Bajdich, M.;
14120−14136. García-Melchor, M.; Han, L.; Xu, J.; Liu, M.; Zheng, L.; Arquer, F. P.
(5) Xu, P.; Huang, T.; Huang, J.; Yan, Y.; Mallouk, T. E. Dye- G. D.; Dinh, C. T.; Fan, F.; Yuan, M.; Yassitepe, E.; Chen, N.; Regier,
sensitized photoelectrochemical water oxidation through a buried T.; Liu, P.; Li, Y.; Luna, P. D.; Janmohamed, A.; Xin, H. L.; Yang, H.;
junction. Proc. Natl. Acad. Sci. U. S. A. 2018, 115, 6946−6951. Vojvodic, A.; Sargent, E. H. Homogeneously dispersed multimetal
(6) Symes, M. D.; Cronin, L. Decoupling hydrogen and oxygen oxygen-evolving catalysts. Science 2016, 352, 333−337.
evolution during electrolytic water splitting using an electron-coupled- (25) Goldsmith, Z. K.; Harshan, A. K.; Gerken, J. B.; Vörös, M.;
proton buffer. Nat. Chem. 2013, 5, 403−409. Galli, G.; Stahl, S. S.; Hammes-Schiffer, S. Characterization of NiFe

G DOI: 10.1021/jacs.8b09076
J. Am. Chem. Soc. XXXX, XXX, XXX−XXX
Journal of the American Chemical Society Article

oxyhydroxide electrocatalysts by integrated electronic structure Chemical Reactivity of C60, Electrophile and Dieno-Polarophile Par
calculations and spectroelectrochemistry. Proc. Natl. Acad. Sci. U. S. Excellence. In Fullerenes. Hammond, G. S., Kuck, V. J., Eds.; ACS
A. 2017, 114, 3050−3055. Symposium Series; American Chemical Society: Washington, D.C.,
(26) Burke, M. S.; Kast, M. G.; Trotochaud, L.; Smith, A. M.; 1992; Vol. 24, pp 161−175.
Boettcher, S. W. Cobalt−iron (Oxy)hydroxide oxygen evolution (43) Sheldrick, G. M. SHELXL97: Program for Crystal Structure
electrocatalysts: the role of structure and composition on activity, Refinement; University of Göttingen: Göttingen, Germany,1997.
stability, and mechanism. J. Am. Chem. Soc. 2015, 137, 3638−3648. (44) Sheldrick, G. M. A short history of SHELX. Acta Crystallogr.,
(27) Smith, R. D. L.; Prévot, M. S.; Fagan, R. D.; Zhang, Z.; Sedach, Sect. A: Found. Crystallogr. 2008, 64, 112−122.
P. A.; Siu, M. K. J.; Trudel, S.; Berlinguette, C. P. Photochemical route (45) Soler, J. M.; Artacho, E.; Gale, J. D.; García, A.; Junquera, J.;
for accessing amorphous metal oxide materials for water oxidation Ordejón, P.; Sánchez-Portal, D. The SIESTA method for ab initio
catalysis. Science 2013, 340, 60−63. order-N materials simulation. J. Phys.: Condens. Matter 2002, 14,
(28) Fan, K.; Chen, H.; Ji, Y.; Huang, H.; Claesson, P. M.; Daniel, 2745−2779.
Q.; Philippe, B.; Rensmo, H.; Li, F.; Luo, Y.; Sun, L. Nickel− (46) Hammer, B.; Hansen, L. B.; Nørskov, J. K. Improved
vanadium monolayer double hydroxide for efficient electrochemical adsorption energetics within density-functional theory using revised
water oxidation. Nat. Commun. 2016, 7, 11981. Perdew-Burke-Ernzerhof functionals. Phys. Rev. B: Condens. Matter
(29) Hu, G.; Nitze, F.; Gracia-Espino, E.; Ma, J.; Barzegar, H. R.; Mater. Phys. 1999, 59, 7413−7421.
Sharifi, T.; Jia, X.; Shchukarev, A.; Lu, L.; Ma, C.; Yang, G.; Wågberg, (47) Wales, D. J.; Doye, J. P. K. Global optimization by basin-
T. Small palladium islands embedded in palladium−tungsten hopping and the lowest energy structures of lennard-jones clusters
bimetallic nanoparticles form catalytic hotspots for oxygen reduction. containing up to 110 atoms. J. Phys. Chem. A 1997, 101, 5111−5116.
Nat. Commun. 2014, 5, 5253. (48) Nørskov, J. K.; Rossmeisl, J.; Logadottir, A.; Lindqvist, L.
(30) Ma, W.; Ma, R.; Wang, C.; Liang, J.; Liu, X.; Zhou, K.; Sasaki, Origin of the overpotential for oxygen reduction at a fuel-cell cathode.
T. A superlattice of alternately stacked Ni−Fe hydroxide nanosheets J. Phys. Chem. B 2004, 108, 17886−17892.
and graphene for efficient splitting of water. ACS Nano 2015, 9, (49) Gracia-Espino, E. Behind the synergistic effect observed on
1977−1984. phosphorus−nitrogen codoped graphene during the oxygen reduction
(31) Fei, H.; Dong, J.; Feng, Y.; Allen, C. S.; Wan, C.; Volosskiy, B.; reaction. J. Phys. Chem. C 2016, 120, 27849−27857.
Li, M.; Zhao, Z.; Wang, Y.; Sun, H.; An, P.; Chen, W.; Guo, Z.; Lee,
C.; Chen, D.; Shakir, I.; Liu, M.; Hu, T.; Li, Y.; Kirkland, A. I.; Duan,
X.; Huang, Y. General synthesis and definitive structural identification
of MN4C4 single-atom catalysts with tunable electrocatalytic activities.
Nat. Catal. 2018, 1, 63−72.
(32) Zhang, G.; Wang, G.; Liu, Y.; Liu, H.; Qu, J.; Li, J. Highly active
and stable catalysts of phytic acid-derivative transition metal
phosphides for full water splitting. J. Am. Chem. Soc. 2016, 138,
14686−14693.
(33) Wang, J.; Gan, L.; Zhang, W.; Peng, Y.; Yu, H.; Yan, Q.; Xia, X.;
Wang, X. In situ formation of molecular Ni-Fe active sites on
heteroatom-doped graphene as a heterogeneous electrocatalyst
toward oxygen evolution. Sci. Adv. 2018, 4, No. eaap7970.
(34) Xue, S.; Chen, L.; Liu, Z.; Cheng, H.-M.; Ren, W. NiPS3
nanosheet−graphene composites as highly efficient electrocatalysts for
oxygen evolution reaction. ACS Nano 2018, 12, 5297−5305.
(35) Lu, X.; Zhao, C. Highly efficient and robust oxygen evolution
catalysts achieved by anchoring nanocrystalline cobalt oxides onto
mildly oxidized multiwalled carbon nanotubes. J. Mater. Chem. A
2013, 1, 12053−12059.
(36) Han, L.; Dong, S.; Wang, E. Transition-metal (Co, Ni, and Fe)-
based electrocatalysts for the water oxidation reaction. Adv. Mater.
2016, 28, 9266−9291.
(37) Man, I. C.; Su, H.-Y.; Calle-Vallejo, F.; Hansen, H. A.;
Martínez, J. I.; Inoglu, N. G.; Kitchin, J.; Jaramillo, T. F.; Nørskov, J.
K.; Rossmeisl, J. Universality in oxygen evolution electrocatalysis on
oxide surfaces. ChemCatChem 2011, 3, 1159−1165.
(38) Fernández, E. M.; Moses, P. G.; Toftelund, A.; Hansen, H. A.;
Martínez, J. I.; Abild-Pedersen, F.; Kleis, J.; Hinnemann, B.;
Rossmeisl, J.; Bligaard, T.; Nørskov, J. K. Scaling relationships for
adsorption energies on transition metal oxide, sulfide, and nitride
surfaces. Angew. Chem., Int. Ed. 2008, 47, 4683−4686.
(39) Han, X.-B.; Zhang, Z.-M.; Zhang, T.; Li, Y.-G.; Lin, W.; You,
W.; Su, Z.-M.; Wang, E.-B. Polyoxometalate-based cobalt-phosphate
molecular catalysts for visible light-driven water oxidation. J. Am.
Chem. Soc. 2014, 136, 5359−5366.
(40) Hervé, G.; Tézé, A. Study of α- and β-enneatungstosilicates and
-germanates. Inorg. Chem. 1977, 16, 2115−2117.
(41) Zhen, J.; Liu, Q.; Chen, X.; Li, D.; Qiao, Q.; Lu, Y.; Yang, S.
Ethanolamine-functionalized fullerene as an efficient electron trans-
port layer for high-efficiency inverted polymer solar cells. J. Mater.
Chem. A 2016, 4, 8072−8079.
(42) Wudl, F.; Hirsch, A.; Khemani, K. C.; Suzuki, T.; Allemand, P.-
M.; Koch, A.; Eckert, H.; Srdanov, G.; Webb, H. M. Survey of

H DOI: 10.1021/jacs.8b09076
J. Am. Chem. Soc. XXXX, XXX, XXX−XXX

Potrebbero piacerti anche