You are on page 1of 11

Article

www.acsnano.org

Two-Dimensional High-Entropy Metal


Phosphorus Trichalcogenides for Enhanced
Hydrogen Evolution Reaction
Ran Wang, Jinzhen Huang, Xinghong Zhang, Jiecai Han, Zhihua Zhang, Tangling Gao, Lingling Xu,
Shengwei Liu, Ping Xu, and Bo Song*
Cite This: ACS Nano 2022, 16, 3593−3603 Read Online
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
Downloaded via ZHEJIANG UNIV on July 19, 2023 at 10:28:56 (UTC).

ABSTRACT: Developing earth-abundant and highly effective electrocatalysts for


hydrogen evolution reaction (HER) is a prerequisite for the upcoming hydrogen
energy society. Two-dimensional (2D) high-entropy metal phosphorus trichalcoge-
nides (MPCh3) have the advantages of both near-continuous adsorption energies of
high-entropy alloys (HEAs) and large specific surface area of 2D materials, which are
excellent catalytic platforms. As a typical 2D high-entropy catalyst,
Co0.6(VMnNiZn)0.4PS3 nanosheets with high-concentration active sites are success-
fully demonstrated to show enhanced HER performance: an overpotential of 65.9
mV at a current density of 10 mA cm−2 and a Tafel slope of 65.5 mV dec−1. Decent
spectroscopy characterizations are combined with density function theory analyses to
show the scenario for the enhancement mechanism by a high-entropy strategy. The optimized S sites on the edge and P sites
on the basal plane provide more active sites for hydrogen adsorption, and the introduced Mn sites boost water dissociation
during the Volmer step. Two-dimensional high-entropy MPCh3 provides an avenue for the combination of HEAs and 2D
materials to enhance the HER performance, which also provides an alternative materials platform to explore and design
superior catalysts for various electrochemical systems.
KEYWORDS: 2D materials, high-entropy, hydrogen evolution reaction, metal phosphorus trichalcogenides, electrocatalysis

INTRODUCTION are active,17 while the sites at the basal plane remain less active
Hydrogen is a carbon-neutral fuel that promises to replace toward HER.18 To regulate the activity of adsorption sites on the
traditional fossil fuels to cope with the increasing greenhouse basal plane, unary or binary doping with only a few elements
effect.1 Hydrogen evolution reaction (HER) in the cathode could be an alternative method, but the lack of broad
plays a key role in electrocatalysis water splitting, which can composition is adverse for further improvements in perform-
effectively convert green electricity into renewable hydrogen ance.19,20 Thus, introducing multiple metal elements to achieve
energy. Although noble metals (Pt,2,3 Ru,4 Rh,5etc.) have continuous tunability for optimal adsorption of reaction
impressive HER activities, their large-scale application is limited intermediates is intuitively a better strategy but remains
by high cost and low elemental abundance. Developing high- unexplored.
performance while cost-effective HER catalysts based on non- As a class of materials system, high-entropy alloys (HEAs)
noble elements is still highly desirable.6 Recently, various earth- have attracted wide interest in the past few years due to their
abundant transition metal compounds have been identified as mechanical properties and potential applications.21−25 In
promising HER catalysts, including metal nitride,7,8 metal contrast to the traditional multicomponent alloys with one or
sulfide,9,10 metal phosphide,11 and metal phosphosulfide.12 two principal elements, HEAs are a kind of alloying strategy that
Among them, the two-dimensional (2D) metal phosphorus involves the combination of five or more elements in a relatively
trichalcogenides (MPCh3) are widely investigated as a
promising catalytic platform with high specific surface area, Received: January 30, 2022
good physicochemical characteristics, tunable charge states, and Accepted: February 23, 2022
appropriate band structures.13,14 Although they have similar Published: February 25, 2022
functional groups of [PxSy]4−,15,16 the intrinsic activity of
MPCh3 is still limited by the improper electron distribution.
Theoretical calculations reveal that only the edge sites in MPCh3

© 2022 American Chemical Society https://doi.org/10.1021/acsnano.2c01064


3593 ACS Nano 2022, 16, 3593−3603
ACS Nano www.acsnano.org Article

high concentration (5−35 at. %).26,27 Because of the abundant reported MPCh3 catalysts. The key merits of
elemental composition, HEAs can provide an extensive Co0.6(VMnNiZn)0.4PS3 NSs are the optimized S sites on the
combinatorial space for designing high-performance cata- edge, P sites on the basal plane, and also the introduced Mn sites
lysts.28−30 Moreover, the continuously tunable adsorption on the edge for efficient water dissociation. These findings
energies of HEAs also optimize the kinetic barrier for the provide an avenue for designing HER electrocatalysts with high
adsorption/desorption of reaction intermediates and thus performance based on high-entropy 2D materials.
increase the catalytic performance.31,32 However, the HEAs
prepared by traditional methods are bulk materials rather than RESULTS AND DISCUSSION
nanostructures.33,34 The limited exposure of active sites makes it The free energy for hydrogen chemisorption (ΔGH*) is an
difficult to further improve the catalytic performance.35 effective descriptor to correlate the theoretical predictions with
Although the preparation of uniform nanostructured HEAs experimental catalytic activities of the hydrogen evolution
with specific equipment and high temperature with fast heating/ systems.1 To evaluate the potential of the high-entropy strategy
cooling rate has been proposed,36−40 achieving a large for improving the HER electrocatalytic activity of MPCh3,
electrochemical active area and full utilization of active sites in density function theory (DFT) calculations were performed to
HEAs is still challenging.41,42 investigate the difference of ΔGH* in normal and high-entropy
Considering both the advantage (layered morphology) and MPCh3.43 Given the structural stabilization and compositional
disadvantage (less active basal plane) of MPCh3, we apply the simplification, the approximately equimolar alloys of Co, V, Mn,
high-entropy strategy to the 2D MPCh3 system and demonstrate Ni, and Zn phosphorus trichalcogenides (CoVMnNiZnPS3)
the successful synthesis of Co0.6(VMnNiZn)0.4PS3 nanosheets were chosen as the representative model of high-entropy
(NSs) with monoclinic structure and high surface area (Figure MPCh3 (Figure 1c), and CoPS3 was considered as a control
1a and b). The Co0.6(VMnNiZn)0.4PS3 NSs achieve a low model. The edge is regarded as a candidate for an electrocatalytic
active site in MPCh3, as shown in Figure 1d. The ΔGH* values of
overpotential (η) of 65.9 mV at the current density of 10 mA
0.79, 0.30, and −0.31 eV are obtained for Co, P, and S sites of
cm−2 and a Tafel slope of 65.5 mV dec−1. To our knowledge, it is
CoPS3 (Figure 1e), respectively, which suggest that the anion
promising that a high-entropy strategy is applied to 2D materials sites (P and S) have better adsorption properties than the cation
for the HER process, and the HER performance of sites (Co). Interestingly, when single metal Co sites are replaced
Co0.6(VMnNiZn)0.4PS3 NSs is one of the best among the by multiple metal atoms to form CoVMnNiZnPS3, positive
shifts of ΔGH* values are observed in P (from 0.30 to 0.35 eV)
and S (from −0.31 to −0.17 eV) sites (Figure 1e), respectively,
suggesting that the S sites at the edge in CoVMnNiZnPS3 are
more favorable for the HER process.
To verify the aforementioned prediction, high-entropy and
single-cation MPCh3 NSs were prepared through a conventional
solid-state reaction (see Experimental Section) followed by
ultrasonication-assisted exfoliation (Figure S1, Supporting
Information). Because of the much lower formation enthalpy
than the configurational entropy in high-entropy MPCh3,44,45
the single-phase monoclinic structure (C2/m) with high
crystallinity inherited from the component is expected, which
is proved by the powder X-ray diffraction (PXRD) pattern
(Figures S2 and S3, Supporting Information).46 A more flexible
composition was present in Cox(VMnNiZn)1−xPS3, allowing the
ratio of Co (x) to be adjusted from 0.2 to 1.0 (Figure S4,
Supporting Information). Taking Co0.6(VMnNiZn)0.4PS3 and
CoPS3 as representative examples, scanning electron micros-
copy (SEM) images displayed the typical layered morphology of
bulk samples with a size of ∼5 μm (Figure S5a and b, Supporting
Information), and the resultant NSs after exfoliating for 3 h also
preserved the lamellar morphology (Figure S5c and d,
Supporting Information). Transmission electron microscopy
(TEM) images further revealed the similar layered morphology
in both Co0.6(VMnNiZn)0.4PS3 (Figure 2a) and CoPS3 NSs
(Figure S6, Supporting Information) with a lateral dimension of
∼400−500 nm. Selected area electron diffraction (SAED)
patterns demonstrated their good crystallinity (inset of Figures
2a and S6, Supporting Information), well consistent with the
corresponding PXRD patterns (Figure S4, Supporting Informa-
Figure 1. DFT calculation on the edge mode. (a) Crystal structure of
CoVMnNiZnPS3 with monoclinic structure (C2/m), and (b) the
tion). The as-prepared NSs are about three to four atomic layers,
structural polymorphs of CoPS3 and CoVMnNiZnPS3 along the top as the average thickness of Co0.6(VMnNiZn)0.4PS3 NSs is ∼2.8
views, respectively. (c) Selection strategy of the element in MPCh3. ± 0.7 nm, determined from the atomic force microscopy (AFM)
(d) Edge models of Co, P, and S sites for CoVMnNiZnPS3 and measurement with the Gauss function fitting (Figure 2b and c).
CoPS3, respectively. (e) HER free-energy diagram of corresponding As shown in the high-resolution TEM (HRTEM) image (Figure
edge sites in (d). 2d), the lattice spacings of 0.167 and 0.173 nm were ascribed to
3594 https://doi.org/10.1021/acsnano.2c01064
ACS Nano 2022, 16, 3593−3603
ACS Nano www.acsnano.org Article

Figure 2. Structural characterizations of Co0.6(VMnNiZn)0.4PS3 and CoPS3 NSs. (a) TEM image, SAED pattern (inset of (a)), (b) AFM image,
and (c) statistical analysis from the AFM image of as-synthesized Co0.6(VMnNiZn)0.4PS3 NSs. HRTEM images of (d) Co0.6(VMnNiZn)0.4PS3
and (e) CoPS3 NSs. (f) EDS mapping images of Co0.6(VMnNiZn)0.4PS3 NSs.

Figure 3. Crystal structure analyses on Cox(VMnNiZn)1−xPS3 bulk samples. (a) Plots of the lattice constant a as a function of the ratio of Co (x).
(b) Plots of the fwhm for the (001) diffraction peak as a function of the ratio of Co (x); the inset is the enlarged (001) peak of CoVMnNiZnPS3
and CoPS3. (c) Lattice strains (ε) obtained by W−H analysis that were extracted from the plots of β cos θ as a function of 4 sin θ. (d) Plots of the
P−P bond length as a function of the Co (x).

3595 https://doi.org/10.1021/acsnano.2c01064
ACS Nano 2022, 16, 3593−3603
ACS Nano www.acsnano.org Article

Figure 4. Electrochemical characterizations of Cox(VMnNiZn)1−xPS3 NSs. (a) LSV curves, (b) corresponding Tafel plots, (c) Cdl extracted from
the CV plots, and (d) overpotentials at the current density of 10 mA cm−2 (left) and the exchange current densities (right). (e) LSV curves
recorded at 25−65 °C and (f) corresponding Aapp data point and error bar of both Co0.6(VMnNiZn)0.4PS3 and CoPS3 NSs.

(330) and (060) planes of Co0.6(VMnNiZn)0.4PS3 NSs, lattice expansion, which is well consistent with HRTEM results.
respectively, which were slightly larger than that of CoPS3 NSs Moreover, the strains in pristine CoPS3 and high-entropy
(d330 = 0.161 and d060 = 0.170 nm), suggesting lattice expansion Cox(VMnNiZn)1−xPS3 (x = 0.2, 0.3, 0.4, 0.5, 0.6, 0.7, and 0.8)
in high-entropy NSs.17 Moreover, compared with CoPS3 NSs bulk samples were further investigated by the full width at half-
(inset of Figure 2e), the atomic-scale disorder from multiple maximum (fwhm) and Williamson−Hall (W−H) analysis. The
transition metals was clearly observed in Co0.6(VMnNiZn)0.4PS3 fwhm of the (001) diffraction peak in CoPS3 (0.083°) and
NSs (inset of Figure 2d), as revealed by the random distribution CoVMnNiZnPS3 (0.132°) convincingly demonstrated that the
in the intensity of atomic columns that is approximately high-entropy strategy distorted the lattices in MPCh3 (Figures
proportional to the squared atomic number (∼Z2).47,48 Energy- 3b and S7, Supporting Information).50 The lattice strains (ε)
dispersive X-ray spectroscopy (EDS) mapping showed a were calculated based on W−H analysis by using the fwhm of
homogeneous spatial distribution of Co, V, Mn, Ni, Zn, P, and the representative peaks (Figure 3c). A positive slope was
S elements throughout Co0.6(VMnNiZn)0.4PS3 NSs (Figure 2f), observed for all samples, as calculated from β cos θ and 4 sin θ,
further confirming no phase segregation. revealing the presence of tensile strain in all the high-entropy
To shed light on the impacts of the high-entropy strategy on samples.51 The ε values indicate that high entropy quintuples
the crystal structures, we used the PXRD data of bulk MPCh3 to the lattice strains from 0.068% to 0.337%. Clearly, the mismatch
evaluate the structural features and strain distribution in the of atomic size with random distribution in high-entropy
corresponding NSs. Decreasing the Co content from 0.8 to 0.2 Cox(VMnNiZn)1−xPS3 distorts the lattice. To deepen the
in Cox(VMnNiZn)1−xPS3, a closer view at the (001) peak at the understanding of the crystal features of high-entropy samples,
2θ range of 13.25−14.25° reveals a continuous shift to lower the Rietveld refinement of PXRD data for bulk
angle relative to that of CoPS3 (Figure S4, Supporting Cox(VMnNiZn)1−xPS3 samples was performed by using the
Information).49 The lattice constant a also exhibits a tendency FullProf software package (Figure S8 and Table S1, Supporting
to increase from 5.901 Å to 5.937 Å (Figure 3a), indicating a Information).52 It was found that the P−P bond length and the
3596 https://doi.org/10.1021/acsnano.2c01064
ACS Nano 2022, 16, 3593−3603
ACS Nano www.acsnano.org Article

Figure 5. Surface-chemical state characterizations on Co0.6(VMnNiZn)0.4PS3, CoPS3, VPS3, MnPS3, NiPS3, and ZnPS3 NSs. XPS spectra for (a)
Co, (b) V, (c) Mn, (d) Ni, and (e) Zn. (f) Binding energy of P and S 2p3/2 orbitals, respectively.

content of Co (x) in Cox(VMnNiZn)1−xPS3 displayed a Co0.8(VMnNiZn)0.2PS3 (9.7 mF cm−2), and CoPS3 (7.9 mF
“volcano” relationship (Figure 3d). Obviously, a longer P−P cm−2) NSs (Figure 4c), demonstrating more catalytic active sites
bond distance was observed in high-entropy exposed in Co0.6(VMnNiZn)0.4PS3 NSs.54 As an inherent
Cox(VMnNiZn)1−xPS3 compared with CoPS3, increasing from parameter of HER activity, the exchange current density (j0)
1.851 Å to 2.120 Å. The tunable P−P bond length suggests that was determined from the intercept of log jvs potential.50 The j0
the local electronic structure of P sites in MPCh3 could be values of 0.33, 0.82, 0.94, 0.20, and 0.06 mA cm−2 could be
regulated via the high-entropy strategy, which may lead to a obtained for CoVMnNiZnPS 3 , Co 0.4 (VMnNiZn) 0.6 PS 3 ,
redistribution of the electron density at each P center and make Co0.6(VMnNiZn)0.4PS3, Co0.8(VMnNiZn)0.2PS3, and CoPS3
them the promising electrocatalytic active sites toward HER. NSs, respectively (Figure 4d). The j0 for Co0.6(VMnNiZn)0.4PS3
To verify the advantages of the high-entropy strategy in NSs (0.94 mA cm−2) is at least ∼15 times higher than that of
MPCh3 NSs, the HER performance was measured using a CoPS 3 NSs (0.06 mA cm −2 ), suggesting a significant
typical three-electrode configuration in a 1 M KOH aqueous contribution of high entropy to enhance the intrinsic catalytic
solution with iR correction (Table S2, Supporting Information). activity. The Co0.6(VMnNiZn)0.4PS3 NSs also showed a high
As expected, Co0.6(VMnNiZn)0.4PS3 NSs exhibited enhanced stability over 3000 CV cycles and a 12 h test at 25 and 65 °C
HER performances with an optimum η of 65.9 mV at the current (Figure S12, Supporting Information), only with a small loss of
density of 10 mA cm−2 and a Tafel slope of 65.5 mV dec−1 electrocatalytic activity that may be caused by a slight
(Figure 4a and b) among a series of MPCh3 NSs (Figure S9, mechanical drop of the catalyst during HER.
Supporting Information), which made it one of the best catalysts To better understand the origin of the enhanced catalytic
among the reported MPCh3 (Table S3, Supporting Informa- activity, the temperature-dependent kinetic analysis was
tion). Furthermore, the electrochemical impedance spectrosco- conducted by linear sweep voltammetry (LSV) in 1 M KOH
py (EIS) data were fitted with a Randles circuit (inset in Figures aqueous solution in the temperature range of 25−65 °C (Figure
S9 and S10, Supporting Information) to extract the resistance S13, Supporting Information).55 It was found that with an
(F i gu r e s S 9 an d S 1 0, S upporting I nformation). increase in temperature the catalytic performance gradually
Co0.6(VMnNiZn)0.4PS3 NSs showed lower charge transfer enhanced (Figure 4e). The pre-exponential factor (Aapp) and
resistance (Rct) (4.8 Ω), confirming its fast electron transfer activation energy (Eapp) were calculated by the Arrhenius
characteristic between electrolyte and catalyst.53 To extract the equation around the onset potential of the respective samples
electrochemical active surface area (ECSA) for the HER (Figure S14, Supporting Information). Compared with CoPS3
process, the double-layer capacitance (Cdl) was calculated via NSs, the maxima of both Aapp and Eapp were observed around a
an analysis of cyclic voltammetry (CV) data (Figure S11, lower η in Co0.6(VMnNiZn)0.4PS3 NSs (Figures 4f and S15,
Supporting Information). The Cdl of Co0.6(VMnNiZn)0.4PS3 Supporting Information).56 The higher Aapp suggests that the
(16.5 mF cm−2) NSs is larger than those of CoVMnNiZnPS3 catalytic active sites in Co0.6(VMnNiZn)0.4PS3 NSs are
(10.6 mF cm−2), Co0.4(VMnNiZn)0.6PS3 (13.0 mF cm−2), enhanced by modifying the entropy of activation, which allows
3597 https://doi.org/10.1021/acsnano.2c01064
ACS Nano 2022, 16, 3593−3603
ACS Nano www.acsnano.org Article

Figure 6. (a) Basal-plane models of P sites (P1−P3) and S sites (S1−S9) in Co0.6(VMnNiZn)0.4PS3. (b) HER free-energy diagram of
corresponding sites in (a). (c) Calculated charge-density difference of the P1 site for Co0.6(VMnNiZn)0.4PS3. The red and blue regions refer to
electron accumulation and depletion, respectively. (d) Calculated reaction energy of water dissociation for Co0.6(VMnNiZn)0.4PS3 and CoPS3,
including Co, V, Mn, Ni, and Zn sites.

more active sites to participate in the HER process, as the at 855.76 (Ni 2p3/2) and 873.86 (Ni 2p1/2) eV (Figure 5c).69 In
predominance on the HER kinetics significantly influences the contrast, the Zn XPS spectrum in ZnPS3 presents only one pair
reaction rate. Meanwhile, the increased reaction entropy needs a of spin−orbit doublets at 1022.06 (2p3/2) and 1045.06 (2p1/2)
higher Eapp as reaction enthalpy to compensate for the change in eV without a shakeup satellite (Figure 5e), which is attributed to
binding energy, due to the stable enthalpy−entropy compensa- Zn 2 + . 7 0 Howe ve r, th e b indin g e ner gy of Zn in
tion effect in alkaline pH condition at both normal and high Co0.6(VMnNiZn)0.4PS3 NSs is higher than that in ZnPS3 NSs.
temperatures (Figure S12, Supporting Information).55 Com- Considering the valence-state ratio and binding-energy shift in
prehensively considering the above electrochemical character- Co0.6(VMnNiZn)0.4PS3 NSs compared with the single-metal
izations, it can be concluded that the high-entropy strategy MPCh3, a trend of electron deficiency states in Co and V
enhanced the HER performance by modifying the reaction indicates the electron density redistribution in high-entropy
entropy and increasing the abundance of active sites during HER NSs.71−73 For the nonmetal element, both P and S exhibit the
in Co0.6(VMnNiZn)0.4PS3 NSs. lowest binding energy in Co0.6(VMnNiZn)0.4PS3 NSs compared
To probe the surface chemical status, high-resolution X-ray with those in single-metal MPCh3 NSs (Figure 5f), confirming a
photoelectron spectroscopy (XPS) measurements were per- higher electron-donating character of [P2S6]4− anions, which
formed on Co0.6(VMnNiZn)0.4PS3 NSs and the corresponding should act as the catalytic active sites and show better hydrogen
single-metal MPCh3 (CoPS3, VPS3, MnPS3, NiPS3, and ZnPS3) adsorption during HER.
NSs as displayed in Figures 5 and S16. The valence-state To verify the stability of the high-entropy structure in MPCh3
proportions of each metal are summarized in Table S4. The during HER, a series of postreaction characterizations were
spin−orbit 2p3/2 and 2p1/2 orbitals of Co modes are observed at performed on Co0.6(VMnNiZn)0.4PS3 NSs after a long-term
binding energies of 779.76 and 795.06 eV (Figure 5a).57−59 In chronopotentiometry test at 10 mA cm−2 in 1 M KOH aqueous
both CoPS3 and Co0.6(VMnNiZn)0.4PS3 NSs, the existence of a solution. The TEM image (Figure S17a, Supporting Informa-
doublet splitting of the Co 2p peaks is attributed to the oxidation tion), SAED patterns (inset of Figure S17a, Supporting
state of Co3+ (779.66/794.76 eV) and Co2+ (781.16/796.36 Information), HRTEM image (Figure S17b, Supporting
eV), respectively.60−62 In the V 2p spectrum, the mode at 514.18 Information), and EDS mapping (Figure S17c, Supporting
eV belongs to V 2p3/2 (V3+) and the mode at 521.98 eV Information) showed that the lamellar morphology of
corresponds to V 2p1/2 (V3+) (Figure 5b). The adjacent modes Co0.6(VMnNiZn)0.4PS3 NSs was preserved, which demonstra-
at 516.78 eV (V 2p3/2) and 524.68 eV (V 2p1/2) are attributed to ted effective exposure of active sites on the catalyst surface
the presence of V4+.63 The Mn 2p3/2 and Mn 2p1/2 orbital modes during the HER process. Moreover, the stability of the crystal
are located at 640.95 and 652.86 eV, respectively, which can be structure and surface chemical status of Co0.6(VMnNiZn)0.4PS3
divided into Mn2+ (640.66/652.16 eV) and Mn3+ (641.96/ NSs was also verified according to PXRD (Figure S18a,
653.46 eV) (Figure 5c), respectively.64,65 The Ni XPS spectrum Supporting Information) and XPS characterizations (Figure
of Co0.6(VMnNiZn)0.4PS3 NSs displays two spin−orbit doublets S18b−h, Supporting Information). Since it is known that the
and two shakeup satellites,66,67 including a Ni2+ doublet at surface of some metal chalcogenides partially dissolves under
854.36 (Ni 2p3/2) and 871.86 (Ni 2p1/2) eV68 and a Ni3+ doublet alkaline HER conditions,74 we used inductively coupled plasma
3598 https://doi.org/10.1021/acsnano.2c01064
ACS Nano 2022, 16, 3593−3603
ACS Nano www.acsnano.org Article

mass spectrometry (ICP-MS) to monitor the catalyst dis- optimize the hydrogen adsorption energy of S sites on the edge
solution rate of the high-performance Co0.6(VMnNiZn)0.4PS3 and P sites on the basal plane. Furthermore, diversified metal
NSs during a 12 h chronopotentiometry test at 25 and 65 °C, sites (especially Mn sites) on the edge boost water dissociation
respectively (Table S5, Supporting Information). 7 5 in Co0.6(VMnNiZn)0.4PS3 NSs during the HER process. In all,
Co0.6(VMnNiZn)0.4PS3 NSs displayed low dissolution ratios at the high-entropy strategy increases the abundance of active sites
both normal and high temperatures. Comprehensively consid- with continuous tenability and optimal adsorption for reaction
ering these postreaction characterizations, it can be concluded intermediates and thus enhances the alkaline HER performance.
that Co0.6(VMnNiZn)0.4PS3 NSs have good stability during This work not only demonstrates the successful application of
alkaline HER. the HEA concept in phosphosulfide material systems but also
To build a better understanding of the fundamental provides a strategy to enhance the electrocatalytic activity for 2D
mechanism for enhanced HER catalytic activity, we further catalysts.
focus on the DFT simulations to study the basal plane of the
catalyst, with the geometry mode optimized by the refinement EXPERIMENTAL SECTION
PXRD data. The detailed local structures on possible proton
acceptors and corresponding calculated ΔGH* are exhibited in Materials. The phosphorus powder (99.99%), sulfur powder
(99.99%), cobalt powder (99.999%), vanadium powder (99.999%),
Figure 6a and b, respectively. As a reference, both P (1.24 eV) manganese (99.999%), iron (99.999%), nickel (99.999%), and zinc
and S (0.63 eV) sites on the basal plane of CoPS3 were calculated powder (99.999%) were purchased from Alfa Aesar. Benchmark 20 wt
(Figure S19, Supporting Information), while the more positive % Pt/C catalysts were purchased from Aladdin. All chemicals were of
ΔGH* values imply a weak hydrogen adsorption strength, analytical grade and were used as received without purification, unless
suggesting its surface is less active toward the HER process. otherwise specified.
Interestingly, in the case of high-entropy, hydrogen adsorptions Synthesis of High-Entropy MPCh3 NSs. Bulk high-entropy
on P sites (P1−P3) are generally better than those of S sites MPCh3 samples were prepared by traditional solid-state reaction with
(S1−S9), and even the P1 site is found to have a near- their nominal composition powders (total mass of 1 g). Before the
thermoneutral ΔGH* value of 0.09 eV, suggesting that solid-state reaction, the elemental powder was ground uniformly to
ensure the reaction happens sufficiently. The elemental powders were
introducing high entropy could effectively activate the basal sealed into quartz tubes under a dynamic vacuum of 5.5 × 10−4 Pa. The
plane of Co0.6(VMnNiZn)0.4PS3. To further explain the impact sealed samples were heated in a muffle furnace from room temperature
of charge redistribution on hydrogen adsorption, the charge- to 610 °C with a heating rate of 1 °C min−1. After it was maintained at
density differences of Co0.6(VMnNiZn)0.4PS3 and CoPS3 610 °C for 168 h, the furnace was cooled to room temperature naturally,
models near active P1 sites were also calculated, as illustrated and a series of bulk samples were obtained. High-entropy MPCh3 NSs
in Figures 6c and S20, respectively. The electron accumulation were obtained from an ultrasonication-assisted exfoliation technique.
on P sites and the electron depletion on S sites indicate that the Bulk high-entropy MPCh3 (50 mg) was added to the water−ethanol
electrons transfer from S to P atoms, which is beneficial to solution (100 mL) and ground for 120 min to ensure that large particles
optimize the P−H, thereby lowing ΔGH*. Moreover, the were sufficiently crushed into small ones. The dispersion liquid was
sonicated for 2 h (500 W) at room temperature in an ultrasonic washer.
number of transferred electrons was calculated by the Mulliken The mixture was centrifuged and the supernatant was decanted. Then,
charge analysis method. Compared with CoPS 3 , the supernatant was put in a typical probe sonicator for 3 h, and the
Co0.6(VMnNiZn)0.4PS3 showed more transferred electrons sonication power was adjusted to 900 W with an on/off cycle of 5/3 s to
(0.51 e−) among the P1 sites (Table S6, Supporting get the exfoliated high-entropy MPCh3 NSs. The uniform solution was
Information), which agrees well with free energy evolution centrifuged at 3000 rpm first. Then, the decanted supernatant was
results.76 Due to the catalyst working in an alkaline solution, the centrifuged again at 9000 rpm, and the high-entropy MPCh3 NSs were
water dissociation barriers of various metal sites on the edge obtained. The samples were washed with absolute ethanol and water
were also performed by DFT calculation (Figure 6d). It is shown sequentially. Finally, the high-entropy MPCh3 NSs were dried at 60 °C
that the energy barrier of 0.69 eV (Mn sites, inset of Figure 6d) for 12 h in a vacuum oven.
Synthesis of Pure MPCh3 NSs. The preparation of precursors for
for Co0.6(VMnNiZn)0.4PS3 to dissociate a water molecule is the synthesis of pure MPCh3 (CoPS3, VPS3, MnPS3, NiPS3, and ZnPS3)
lower than that of Co sites in CoPS3 (1.21 eV), suggesting that followed similar protocols except the amounts of metal powder used
metal site diversification promotes the water dissociation and the holding temperature. The CoPS3 sample was prepared at 550
kinetics during the Volmer step.77 Thus, the calculation results °C for 720 h. The VPS3 sample was synthesized at 470 °C for 720 h. The
prove that the enhanced HER activity in high-entropy MPCh3 ZnPS3 sample was additionally heated at 750 °C for 168 h. The other
originates from the synergistic effects of abundant active sites steps were the same as those mentioned above.
endowed from the high-entropy strategy, with optimized S sites Structural Characterization. PXRD measurements were per-
on the edge and P sites on the basal plane providing more active formed on a Rigaku D/max 2500 X-ray diffractometer using Cu Kα
sites for hydrogen adsorption, as well as the introduced Mn sites radiation (λ = 1.5406 Å). XPS spectra were recorded on an ESCALAB
MKII spectrometer using an Al Kα excitation source. A JEOL ARM
on the edge as an efficient water dissociation center. 200F transmission electron microscope and a field-emission gun (FEG)
scanning electron microscope (JEOL 6500 SEM) were used to
CONCLUSIONS characterize the morphology and structure of the NSs. The thickness of
In conclusion, combining high-entropy and 2D material the MPCh3 NSs was analyzed by AFM on a Bruker DI MultiMode-8
concepts, we have systematically investigated the HER proper- system. The ICP-MS analysis was carried out on an Agilent
ties of a series 2D high-entropy MPCh 3 and proved ICPMS7800 spectrometer.
Electrochemical Measurements. Electrochemical measurements
Co0.6(VMnNiZn)0.4PS3 NSs to be one of the best catalysts
were performed with a standard three-electrode setup (CH Instru-
among reported MPCh3, with a low η of 65.9 mV at a current ments) using Hg/HgO as the reference electrode, a graphite rod (Alfa
density of 10 mA cm−2 and a Tafel slope of 65.5 mV dec−1. Aesar, 99.9995%) as the counter electrode, and a clean carbon fiber
Decent spectroscopy characterizations combined with DFT cloth coated with drop-cast MPCh3 catalysts as the working electrode.
calculation results suggest a redistribution of electron density The catalyst was ultrasonically dispersed in a water−ethanol solution
among the [P2S6]4− anions in the high-entropy samples, which (v/v = 3:7) containing 0.1 wt % Nafion to get the catalyst ink of 5 μg

3599 https://doi.org/10.1021/acsnano.2c01064
ACS Nano 2022, 16, 3593−3603
ACS Nano www.acsnano.org Article

μL−1, and 70 μL of the catalyst ink was drop-casted onto the carbon AUTHOR INFORMATION
fiber cloth electrode with a geometric area of 1 cm2. The amount of
deposited catalyst was calculated to be ∼350 μg. All the measurements Corresponding Author
were performed in a H2-saturated KOH (1.0 M) aqueous solution. CVs Bo Song − National Key Laboratory of Science and Technology
taken at various scan rates (20−80 mV s−1) were collected and used to on Advanced Composites in Special Environments, Harbin
extract the double-layer capacitance. The EIS measurements were Institute of Technology, Harbin 150001, China; orcid.org/
carried out at a 100 mV overpotential in the frequency range from 106 to 0000-0003-2000-5071; Email: songbo@hit.edu.cn
0.1 Hz. The electrochemical stability of the catalyst was evaluated by
cycling the electrodes for 3000 times. All the potentials were referenced Authors
to a reversible hydrogen electrode (RHE). To extract the Aapp and Eapp Ran Wang − National Key Laboratory of Science and
for HER, the electrochemical measurements of the catalysts were Technology on Advanced Composites in Special Environments,
conducted in 1.0 M KOH solution at different temperatures. For the
heterogeneous electrocatalytic reaction, the current density can be
Harbin Institute of Technology, Harbin 150001, China;
expressed from Eapp in the Arrhenius equation:55 orcid.org/0000-0002-2034-7189
Jinzhen Huang − National Key Laboratory of Science and
j = A app exp(− Eapp(RT )−1) Technology on Advanced Composites in Special Environments,
(1)
Harbin Institute of Technology, Harbin 150001, China;
where Aapp is the apparent pre-exponential factor, R is the ideal gas orcid.org/0000-0002-0274-0454
constant (8.314 J K−1 mol−1), and T is the temperature in Kelvin (K). Xinghong Zhang − National Key Laboratory of Science and
Therefore, Eapp can be further calculated from fitting the slope of the Technology on Advanced Composites in Special Environments,
Arrhenius plot using the equation78 Harbin Institute of Technology, Harbin 150001, China
Jiecai Han − National Key Laboratory of Science and
|∂(log10 j)∂(1/T )−1|η = − Eapp(2.303R )−1 (2) Technology on Advanced Composites in Special Environments,
Harbin Institute of Technology, Harbin 150001, China
while the intercept of log10jvs 1/T plot is the logarithm of Aapp. Zhihua Zhang − School of Materials Science and Engineering,
Dalian Jiaotong University, Dalian 116028, China
THEORETICAL CALCULATIONS Tangling Gao − Institute of Petrochemistry, Heilongjiang
Academy of Sciences, Harbin 150040, China
DFT Calculations. All the DFT calculations were performed Lingling Xu − Key Laboratory of Photonic and Electronic
by using the CASTEP code.79 All-electron calculations were Bandgap Materials, Ministry of Education, School of Physics
employed using the generalized gradient approximation and the and Electronic Engineering, Harbin Normal University,
Perdew, Burke, and Ernzerhof functional.80 The ultrasoft Harbin 150080, China
pseudopotentials were used, which are fairly efficient compared Shengwei Liu − School of Environmental Science and
with the norm-conserving pseudopotentials. The plane-wave Engineering, Guangdong Provincial Key Laboratory of
cutoff energy was 450 eV, and the Grimme method for DFT-D3 Environmental Pollution Control and Remediation
was used to account for van der Waals interactions. The Technology, Sun Yat-sen University, Guangzhou 510006,
minimum energies for all structures were obtained until the China; orcid.org/0000-0001-5677-1769
energy became less than 2 × 10−5 eV/atom. The vacuum gap Ping Xu − MIIT Key Laboratory of Critical Materials
between periodic images was set to 15 Å to avoid interaction. Technology for New Energy Conversion and Storage, School of
The Brillouin zone was sampled with a 2 × 2 × 1 gamma- Chemistry and Chemical Engineering, Harbin Institute of
centered special k point grid for geometry optimization. Technology, Harbin 150001, China; orcid.org/0000-
0002-1516-4986
ASSOCIATED CONTENT Complete contact information is available at:
*
sı Supporting Information https://pubs.acs.org/10.1021/acsnano.2c01064
The Supporting Information is available free of charge at
https://pubs.acs.org/doi/10.1021/acsnano.2c01064. Notes
The authors declare no competing financial interest.
Figures of a schematic illustration for the synthesis
process; PXRD patterns of high-entropy MPCh3 bulk ACKNOWLEDGMENTS
samples; PXRD patterns of single-cation MPCh3 bulk
samples; PXRD patterns of bulk Cox(VMnNiZn)1−xPS3; This work was supported by the National Natural Science
SEM images of bulk samples; TEM image of CoPS3 NSs; Foundation of China (Grant Nos. 52072085, 52072078, and
enlarged (001) peaks for bulk Cox(VMnNiZn)1−xPS3; 51902091).
Rietveld refinement; electrochemical characterizations;
EIS plots; CV curves; durability test; LSV curves in the REFERENCES
temperature range; logarithm of the catalytic current (1) Seh, W.; Kibsgaard, J.; Dickens Colin, F.; Chorkendorff, I.;
density; extraction of Eapp; P 2p and S 2p XPS spectra; Nørskov Jens, K.; Jaramillo Thomas, F. Combining Theory and
structural characterizations after stability test; crystal Experiment in Electrocatalysis: Insights into Materials Design. Science
structure and surface-chemical state characterizations 2017, 355, No. eaad4998.
after stability test; HER free-energy diagram; calculated (2) Subbaraman, R.; Tripkovic, D.; Strmcnik, D.; Chang, K.-C.;
charge-density difference; tables of reliability factors in Uchimura, M.; Paulikas Arvydas, P.; Stamenkovic, V.; Markovic Nenad,
Rietveld refinement; electrocatalytic performance param- M. Enhancing Hydrogen Evolution Activity in Water Splitting by
Tailoring Li+-Ni(OH)2-Pt Interfaces. Science 2011, 334, 1256−1260.
eters; comparison of HER activity; valence-state (3) Wang, Z.; Xiao, B.; Lin, Z.; Xu, Y.; Lin, Y.; Meng, F.; Zhang, Q.;
proportions; and dissolution ratios (PDF) Gu, L.; Fang, B.; Guo, S.; Zhong, W. PtSe2/Pt Heterointerface with

3600 https://doi.org/10.1021/acsnano.2c01064
ACS Nano 2022, 16, 3593−3603
ACS Nano www.acsnano.org Article

Reduced Coordination for Boosted Hydrogen Evolution Reaction. (20) Wang, J.; Li, X.; Wei, B.; Sun, R.; Yu, W.; Hoh, H. Y.; Xu, H.; Li,
Angew. Chem., Int. Ed. 2021, 60, 23388−23393. J.; Ge, X.; Chen, Z.; Su, C.; Wang, Z. Activating Basal Planes of NiPS3
(4) Mahmood, J.; Li, F.; Jung, S.-M.; Okyay, M. S.; Ahmad, I.; Kim, S.- for Hydrogen Evolution by Nonmetal Heteroatom Doping. Adv. Funct.
J.; Park, N.; Jeong, H. Y.; Baek, J.-B. An Efficient and pH-Universal Mater. 2020, 30, 1908708.
Ruthenium-Based Catalyst for the Hydrogen Evolution Reaction. Nat. (21) Gludovatz, B.; Hohenwarter, A.; Catoor, D.; Chang Edwin, H.;
Nanotechnol. 2017, 12, 441−446. George Easo, P.; Ritchie Robert, O. A Fracture-Resistant High-Entropy
(5) Zheng, J.; Sheng, W.; Zhuang, Z.; Xu, B.; Yan, Y. Universal Alloy for Cryogenic Applications. Science 2014, 345, 1153−1158.
Dependence of Hydrogen Oxidation and Evolution Reaction Activity (22) Li, Z.; Pradeep, K. G.; Deng, Y.; Raabe, D.; Tasan, C. C.
of Platinum-Group Metals on pH and Hydrogen Binding Energy. Sci. Metastable High-Entropy Dual-Phase Alloys Overcome the Strength−
Adv. 2016, 2, No. e1501602. Ductility Trade-Off. Nature 2016, 534, 227−230.
(6) Voiry, D.; Yamaguchi, H.; Li, J.; Silva, R.; Alves, D. C. B.; Fujita, T.; (23) Lun, Z.; Ouyang, B.; Kwon, D.-H.; Ha, Y.; Foley, E. E.; Huang,
Chen, M.; Asefa, T.; Shenoy, V. B.; Eda, G.; Chhowalla, M. Enhanced T.-Y.; Cai, Z.; Kim, H.; Balasubramanian, M.; Sun, Y.; Huang, J.; Tian,
Catalytic Activity in Strained Chemically Exfoliated WS2 Nanosheets Y.; Kim, H.; McCloskey, B. D.; Yang, W.; Clément, R. J.; Ji, H.; Ceder,
for Hydrogen Evolution. Nat. Mater. 2013, 12, 850−855. G. Cation-Disordered Rocksalt-Type High-Entropy Cathodes for Li-
(7) Jin, H.; Liu, X.; Vasileff, A.; Jiao, Y.; Zhao, Y.; Zheng, Y.; Qiao, S.-Z. Ion Batteries. Nat. Mater. 2021, 20, 214−221.
Single-Crystal Nitrogen-Rich Two-Dimensional Mo5N6 Nanosheets (24) Sun, Y.; Dai, S. High-Entropy Materials for Catalysis: A New
for Efficient and Stable Seawater Splitting. ACS Nano 2018, 12, 12761− Frontier. Sci. Adv. 2021, 7, No. eabg1600.
12769. (25) Löffler, T.; Ludwig, A.; Rossmeisl, J.; Schuhmann, W. What
(8) Jin, H.; Gu, Q.; Chen, B.; Tang, C.; Zheng, Y.; Zhang, H.; Jaroniec, Makes High-Entropy Alloys Exceptional Electrocatalysts? Angew.
M.; Qiao, S.-Z. Molten Salt-Directed Catalytic Synthesis of 2D Layered Chem., Int. Ed. 2021, 60, 26894−26903.
Transition-Metal Nitrides for Efficient Hydrogen Evolution. Chem. (26) Löffler, T.; Meyer, H.; Savan, A.; Wilde, P.; Garzón Manjón, A.;
2020, 6, 2382−2394. Chen, Y.-T.; Ventosa, E.; Scheu, C.; Ludwig, A.; Schuhmann, W.
(9) Sivanantham, A.; Ganesan, P.; Shanmugam, S. Hierarchical Discovery of a Multinary Noble Metal−Free Oxygen Reduction
NiCo2S4 Nanowire Arrays Supported on Ni Foam: An Efficient and Catalyst. Adv. Energy Mater. 2018, 8, 1802269.
Durable Bifunctional Electrocatalyst for Oxygen and Hydrogen (27) Du, Z.; Wu, C.; Chen, Y.; Zhu, Q.; Cui, Y.; Wang, H.; Zhang, Y.;
Evolution Reactions. Adv. Funct. Mater. 2016, 26, 4661−4672. Chen, X.; Shang, J.; Li, B.; Chen, W.; Liu, C.; Yang, S. High-Entropy
(10) Yang, J.; Mohmad, A. R.; Wang, Y.; Fullon, R.; Song, X.; Zhao, F.; Carbonitride Max Phases and Their Derivative Mxenes. Adv. Energy
Bozkurt, I.; Augustin, M.; Santos, E. J. G.; Shin, H. S.; Zhang, W.; Voiry, Mater. 2022, 12, 2103228.
D.; Jeong, H. Y.; Chhowalla, M. Ultrahigh-Current-Density Niobium (28) Ding, Q.; Zhang, Y.; Chen, X.; Fu, X.; Chen, D.; Chen, S.; Gu, L.;
Disulfide Catalysts for Hydrogen Evolution. Nat. Mater. 2019, 18, Wei, F.; Bei, H.; Gao, Y.; Wen, M.; Li, J.; Zhang, Z.; Zhu, T.; Ritchie, R.
1309−1314. O.; Yu, Q. Tuning Element Distribution, Structure and Properties by
(11) Popczun, E. J.; McKone, J. R.; Read, C. G.; Biacchi, A. J.; Composition in High-Entropy Alloys. Nature 2019, 574, 223−227.
Wiltrout, A. M.; Lewis, N. S.; Schaak, R. E. Nanostructured Nickel (29) Wang, T.; Chen, H.; Yang, Z.; Liang, J.; Dai, S. High-Entropy
Phosphide as an Electrocatalyst for the Hydrogen Evolution Reaction. J. Perovskite Fluorides: A New Platform for Oxygen Evolution Catalysis.
Am. Chem. Soc. 2013, 135, 9267−9270. J. Am. Chem. Soc. 2020, 142, 4550−4554.
(12) Cabán-Acevedo, M.; Stone, M. L.; Schmidt, J. R.; Thomas, J. G.; (30) Batchelor, T. A. A.; Löffler, T.; Xiao, B.; Krysiak, O. A.;
Ding, Q.; Chang, H.-C.; Tsai, M.-L.; He, J.-H.; Jin, S. Efficient Strotkötter, V.; Pedersen, J. K.; Clausen, C. M.; Savan, A.; Li, Y.;
Hydrogen Evolution Catalysis Using Ternary Pyrite-Type Cobalt Schuhmann, W.; Rossmeisl, J.; Ludwig, A. Complex-Solid-Solution
Phosphosulphide. Nat. Mater. 2015, 14, 1245−1251. Electrocatalyst Discovery by Computational Prediction and High-
(13) Wang, F.; Shifa, T. A.; Yu, P.; He, P.; Liu, Y.; Wang, F.; Wang, Z.; Throughput Experimentation. Angew. Chem., Int. Ed. 2021, 60, 6932−
Zhan, X.; Lou, X.; Xia, F.; He, J. New Frontiers on Van Der Waals 6937.
Layered Metal Phosphorous Trichalcogenides. Adv. Funct. Mater. 2018, (31) Batchelor, T. A. A.; Pedersen, J. K.; Winther, S. H.; Castelli, I. E.;
28, 1802151. Jacobsen, K. W.; Rossmeisl, J. High-Entropy Alloys as a Discovery
(14) Gusmão, R.; Sofer, Z.; Pumera, M. Metal Phosphorous Platform for Electrocatalysis. Joule 2019, 3, 834−845.
Trichalcogenides (MPCh3): From Synthesis to Contemporary Energy (32) Koo, W.-T.; Millstone, J. E.; Weiss, P. S.; Kim, I.-D. The Design
Challenges. Angew. Chem., Int. Ed. 2019, 58, 9326−9337. and Science of Polyelemental Nanoparticles. ACS Nano 2020, 14,
(15) Gusmão, R.; Sofer, Z.; Pumera, M. Exfoliated Layered 6407−6413.
Manganese Trichalcogenide Phosphite (MnPX3, X = S, Se) as (33) Li, H.; Han, Y.; Zhao, H.; Qi, W.; Zhang, D.; Yu, Y.; Cai, W.; Li,
Electrocatalytic Van Der Waals Materials for Hydrogen Evolution. S.; Lai, J.; Huang, B.; Wang, L. Fast Site-to-Site Electron Transfer of
Adv. Funct. Mater. 2019, 29, 1805975. High-Entropy Alloy Nanocatalyst Driving Redox Electrocatalysis. Nat.
(16) Kang, S.; Kim, K.; Kim, B. H.; Kim, J.; Sim, K. I.; Lee, J.-U.; Lee, Commun. 2020, 11, 5437.
S.; Park, K.; Yun, S.; Kim, T.; Nag, A.; Walters, A.; Garcia-Fernandez, (34) Mori, K.; Hashimoto, N.; Kamiuchi, N.; Yoshida, H.; Kobayashi,
M.; Li, J.; Chapon, L.; Zhou, K.-J.; Son, Y.-W.; Kim, J. H.; Cheong, H.; H.; Yamashita, H. Hydrogen Spillover-Driven Synthesis of High-
Park, J.-G. Coherent Many-Body Exciton in Van Der Waals Entropy Alloy Nanoparticles as a Robust Catalyst for CO2 Hydro-
Antiferromagnet NiPS3. Nature 2020, 583, 785−789. genation. Nat. Commun. 2021, 12, 3884.
(17) Song, B.; Li, K.; Yin, Y.; Wu, T.; Dang, L.; Cabán-Acevedo, M.; (35) Glasscott, M. W.; Pendergast, A. D.; Goines, S.; Bishop, A. R.;
Han, J.; Gao, T.; Wang, X.; Zhang, Z.; Schmidt, J. R.; Xu, P.; Jin, S. Hoang, A. T.; Renault, C.; Dick, J. E. Electrosynthesis of High-Entropy
Tuning Mixed Nickel Iron Phosphosulfide Nanosheet Electrocatalysts Metallic Glass Nanoparticles for Designer, Multi-Functional Electro-
for Enhanced Hydrogen and Oxygen Evolution. ACS Catal. 2017, 7, catalysis. Nat. Commun. 2019, 10, 2650.
8549−8557. (36) Yao, Y.; Huang, Z.; Xie, P.; Lacey Steven, D.; Jacob Rohit, J.; Xie,
(18) Mukherjee, D.; Austeria, P. M.; Sampath, S. Two-Dimensional, H.; Chen, F.; Nie, A.; Pu, T.; Rehwoldt, M.; Yu, D.; Zachariah Michael,
Few-Layer Phosphochalcogenide, FePS3: A New Catalyst for Electro- R.; Wang, C.; Shahbazian-Yassar, R.; Li, J.; Hu, L. Carbothermal Shock
chemical Hydrogen Evolution over Wide pH Range. ACS Energy Lett. Synthesis of High-Entropy-Alloy Nanoparticles. Science 2018, 359,
2016, 1, 367−372. 1489−1494.
(19) Liang, Q.; Zhong, L.; Du, C.; Zheng, Y.; Luo, Y.; Xu, J.; Li, S.; (37) Sarkar, A.; Velasco, L.; Wang, D.; Wang, Q.; Talasila, G.; de Biasi,
Yan, Q. Mosaic-Structured Cobalt Nickel Thiophosphate Nanosheets L.; Kübel, C.; Brezesinski, T.; Bhattacharya, S. S.; Hahn, H.; Breitung, B.
Incorporated N-Doped Carbon for Efficient and Stable Electrocatalytic High Entropy Oxides for Reversible Energy Storage. Nat. Commun.
Water Splitting. Adv. Funct. Mater. 2018, 28, 1805075. 2018, 9, 3400.

3601 https://doi.org/10.1021/acsnano.2c01064
ACS Nano 2022, 16, 3593−3603
ACS Nano www.acsnano.org Article

(38) Kenel, C.; Casati, N. P. M.; Dunand, D. C. 3D Ink-Extrusion (56) Yu, X.; Wang, B.; Wang, C.; Zhuang, C.; Yao, Y.; Li, Z.; Wu, C.;
Additive Manufacturing of CoCrFeNi High-Entropy Alloy Micro- Feng, J.; Zou, Z. 2D High-Entropy Hydrotalcites. Small 2021, 17,
Lattices. Nat. Commun. 2019, 10, 904. 2103412.
(39) Yang, J. X.; Dai, B.-H.; Chiang, C.-Y.; Chiu, I. C.; Pao, C.-W.; Lu, (57) Li, T.; Yao, Y.; Ko, B. H.; Huang, Z.; Dong, Q.; Gao, J.; Chen, W.;
S.-Y.; Tsao, I. Y.; Lin, S.-T.; Chiu, C.-T.; Yeh, J.-W.; Chang, P.-C.; Li, J.; Li, S.; Wang, X.; Shahbazian-Yassar, R.; Jiao, F.; Hu, L. Carbon-
Hung, W.-H. Rapid Fabrication of High-Entropy Ceramic Nanoma- Supported High-Entropy Oxide Nanoparticles as Stable Electro-
terials for Catalytic Reactions. ACS Nano 2021, 15, 12324−12333. catalysts for Oxygen Reduction Reactions. Adv. Funct. Mater. 2021,
(40) Qiao, H.; Saray, M. T.; Wang, X.; Xu, S.; Chen, G.; Huang, Z.; 31, 2010561.
Chen, C.; Zhong, G.; Dong, Q.; Hong, M.; Xie, H.; Shahbazian-Yassar, (58) Jin, Z.; Lyu, J.; Zhao, Y.-L.; Li, H.; Chen, Z.; Lin, X.; Xie, G.; Liu,
R.; Hu, L. Scalable Synthesis of High Entropy Alloy Nanoparticles by X.; Kai, J.-J.; Qiu, H.-J. Top−Down Synthesis of Noble Metal Particles
Microwave Heating. ACS Nano 2021, 15, 14928−14937. on High-Entropy Oxide Supports for Electrocatalysis. Chem. Mater.
(41) Xin, Y.; Li, S.; Qian, Y.; Zhu, W.; Yuan, H.; Jiang, P.; Guo, R.; 2021, 33, 1771−1780.
Wang, L. High-Entropy Alloys as a Platform for Catalysis: Progress, (59) Nguyen, T. X.; Liao, Y.-C.; Lin, C.-C.; Su, Y.-H.; Ting, J.-M.
Challenges, and Opportunities. ACS Catal. 2020, 10, 11280−11306. Advanced High Entropy Perovskite Oxide Electrocatalyst for Oxygen
(42) McCormick, C. R.; Schaak, R. E. Simultaneous Multication Evolution Reaction. Adv. Funct. Mater. 2021, 31, 2101632.
Exchange Pathway to High-Entropy Metal Sulfide Nanoparticles. J. Am. (60) Li, S.; Wang, J.; Lin, X.; Xie, G.; Huang, Y.; Liu, X.; Qiu, H.-J.
Chem. Soc. 2021, 143, 1017−1023. Flexible Solid-State Direct Ethanol Fuel Cell Catalyzed by Nanoporous
(43) Nellaiappan, S.; Katiyar, N. K.; Kumar, R.; Parui, A.; Malviya, K. High-Entropy Al-Pd-Ni-Cu-Mo Anode and Spinel (AlMnCo)3O4
D.; Pradeep, K. G.; Singh, A. K.; Sharma, S.; Tiwary, C. S.; Biswas, K. Cathode. Adv. Funct. Mater. 2021, 31, 2007129.
High-Entropy Alloys as Catalysts for the CO2 and CO Reduction (61) Zhao, J.; Bao, J.; Yang, S.; Niu, Q.; Xie, R.; Zhang, Q.; Chen, M.;
Reactions: Experimental Realization. ACS Catal. 2020, 10, 3658−3663. Zhang, P.; Dai, S. Exsolution−Dissolution of Supported Metals on
(44) Ying, T.; Yu, T.; Shiah, Y.-S.; Li, C.; Li, J.; Qi, Y.; Hosono, H. High-Entropy Co3MnNiCuZnOx: Toward Sintering-Resistant Catal-
High-Entropy Van Der Waals Materials Formed from Mixed Metal ysis. ACS Catal. 2021, 11, 12247−12257.
Dichalcogenides, Halides, and Phosphorus Trisulfides. J. Am. Chem. (62) Wang, Z.; Lin, Z.; Deng, J.; Shen, S.; Meng, F.; Zhang, J.; Zhang,
Soc. 2021, 143, 7042−7049. Q.; Zhong, W.; Gu, L. Elevating the d-Band Center of Six-Coordinated
(45) Jin, H.; Song, T.; Paik, U.; Qiao, S.-Z. Metastable Two- Octahedrons in Co9S8 through Fe-Incorporated Topochemical
Dimensional Materials for Electrocatalytic Energy Conversions. Acc. Deintercalation. Adv. Energy Mater. 2021, 11, 2003023.
Mater. Res. 2021, 2, 559−573. (63) Rout, C. S.; Kim, B.-H.; Xu, X.; Yang, J.; Jeong, H. Y.; Odkhuu,
(46) Li, X.; Fang, Y.; Wang, J.; Wei, B.; Qi, K.; Hoh, H. Y.; Hao, Q.; D.; Park, N.; Cho, J.; Shin, H. S. Synthesis and Characterization of
Sun, T.; Wang, Z.; Yin, Z.; Zhang, Y.; Lu, J.; Bao, Q.; Su, C. High-Yield Patronite Form of Vanadium Sulfide on Graphitic Layer. J. Am. Chem.
Soc. 2013, 135, 8720−8725.
Electrochemical Production of Large-Sized and Thinly Layered NiPS3
(64) Gao, S.; Hao, S.; Huang, Z.; Yuan, Y.; Han, S.; Lei, L.; Zhang, X.;
Flakes for Overall Water Splitting. Small 2019, 15, 1902427.
Shahbazian-Yassar, R.; Lu, J. Synthesis of High-Entropy Alloy
(47) Zhao, X.; Xue, Z.; Chen, W.; Bai, X.; Shi, R.; Mu, T. Ambient
Nanoparticles on Supports by the Fast Moving Bed Pyrolysis. Nat.
Fast, Large-Scale Synthesis of Entropy-Stabilized Metal−Organic
Commun. 2020, 11, 2016.
Framework Nanosheets for Electrocatalytic Oxygen Evolution. J.
(65) Han, M.; Wang, C.; Zhong, J.; Han, J.; Wang, N.; Seifitokaldani,
Mater. Chem. A 2019, 7, 26238−26242.
A.; Yu, Y.; Liu, Y.; Sun, X.; Vomiero, A.; Liang, H. Promoted Self-
(48) Cavin, J.; Ahmadiparidari, A.; Majidi, L.; Thind, A. S.; Misal, S.
Construction of β-NiOOH in Amorphous High Entropy Electro-
N.; Prajapati, A.; Hemmat, Z.; Rastegar, S.; Beukelman, A.; Singh, M.
catalysts for the Oxygen Evolution Reaction. Appl. Catal., B-Environ.
R.; Unocic, K. A.; Salehi-Khojin, A.; Mishra, R. 2D High-Entropy 2022, 301, 120764.
Transition Metal Dichalcogenides for Carbon Dioxide Electrocatalysis. (66) Sivanantham, A.; Lee, H.; Hwang, S. W.; Ahn, B.; Cho, I. S.
Adv. Mater. 2021, 33, 2100347. Preparation, Electrical and Electrochemical Characterizations of
(49) Xu, H.; Zhang, Z.; Liu, J.; Do-Thanh, C.-L.; Chen, H.; Xu, S.; Lin, CuCoNiFeMn High-Entropy-Alloy for Overall Water Splitting at
Q.; Jiao, Y.; Wang, J.; Wang, Y.; Chen, Y.; Dai, S. Entropy-Stabilized Neutral-pH. J. Mater. Chem. A 2021, 9, 16841−16851.
Single-Atom Pd Catalysts Via High-Entropy Fluorite Oxide Supports. (67) Cui, M.; Yang, C.; Li, B.; Dong, Q.; Wu, M.; Hwang, S.; Xie, H.;
Nat. Commun. 2020, 11, 3908. Wang, X.; Wang, G.; Hu, L. High-Entropy Metal Sulfide Nanoparticles
(50) Zhan, C.; Xu, Y.; Bu, L.; Zhu, H.; Feng, Y.; Yang, T.; Zhang, Y.; Promise High-Performance Oxygen Evolution Reaction. Adv. Energy
Yang, Z.; Huang, B.; Shao, Q.; Huang, X. Subnanometer High-Entropy Mater. 2021, 11, 2002887.
Alloy Nanowires Enable Remarkable Hydrogen Oxidation Catalysis. (68) Huang, K.; Zhang, B.; Wu, J.; Zhang, T.; Peng, D.; Cao, X.;
Nat. Commun. 2021, 12, 6261. Zhang, Z.; Li, Z.; Huang, Y. Exploring the Impact of Atomic Lattice
(51) Jiang, B.; Yu, Y.; Cui, J.; Liu, X.; Xie, L.; Liao, J.; Zhang, Q.; Deformation on Oxygen Evolution Reactions Based on a Sub-5 nm
Huang, Y.; Ning, S.; Jia, B.; Zhu, B.; Bai, S.; Chen, L.; Pennycook Pure Face-Centred Cubic High-Entropy Alloy Electrocatalyst. J. Mater.
Stephen, J.; He, J. High-Entropy-Stabilized Chalcogenides with High Chem. A 2020, 8, 11938−11947.
Thermoelectric Performance. Science 2021, 371, 830−834. (69) Nguyen, T. X.; Su, Y.-H.; Lin, C.-C.; Ting, J.-M. Self-
(52) Broge, N. L. N.; Bondesgaard, M.; Søndergaard-Pedersen, F.; Reconstruction of Sulfate-Containing High Entropy Sulfide for
Roelsgaard, M.; Iversen, B. B. Autocatalytic Formation of High-Entropy Exceptionally High-Performance Oxygen Evolution Reaction Electro-
Alloy Nanoparticles. Angew. Chem., Int. Ed. 2020, 59, 21920−21924. catalyst. Adv. Funct. Mater. 2021, 31, 2106229.
(53) Lai, D.; Kang, Q.; Gao, F.; Lu, Q. High-Entropy Effect of a Metal (70) Huang, W.; Zhang, J.; Liu, D.; Xu, W.; Wang, Y.; Yao, J.; Tan, H.
Phosphide on Enhanced Overall Water Splitting Performance. J. Mater. T.; Dinh, K. N.; Wu, C.; Kuang, M.; Fang, W.; Dangol, R.; Song, L.;
Chem. A 2021, 9, 17913−17922. Zhou, K.; Liu, C.; Xu, J. W.; Liu, B.; Yan, Q. Tuning the Electronic
(54) Zhang, L.; Cai, W.; Bao, N. Top-Level Design Strategy to Structures of Multimetal Oxide Nanoplates to Realize Favorable
Construct an Advanced High-Entropy Co−Cu−Fe−Mo (Oxy)- Adsorption Energies of Oxygenated Intermediates. ACS Nano 2020,
Hydroxide Electrocatalyst for the Oxygen Evolution Reaction. Adv. 14, 17640−17651.
Mater. 2021, 33, 2100745. (71) Jia, Z.; Nomoto, K.; Wang, Q.; Kong, C.; Sun, L.; Zhang, L.-C.;
(55) Duan, Y.; Dubouis, N.; Huang, J.; Dalla Corte, D. A.; Pimenta, V.; Liang, S.-X.; Lu, J.; Kruzic, J. J. A Self-Supported High-Entropy Metallic
Xu, Z. J.; Grimaud, A. Revealing the Impact of Electrolyte Composition Glass with a Nanosponge Architecture for Efficient Hydrogen
for Co-Based Water Oxidation Catalysts by the Study of Reaction Evolution under Alkaline and Acidic Conditions. Adv. Funct. Mater.
Kinetics Parameters. ACS Catal. 2020, 10, 4160−4170. 2021, 31, 2101586.

3602 https://doi.org/10.1021/acsnano.2c01064
ACS Nano 2022, 16, 3593−3603
ACS Nano www.acsnano.org Article

(72) Chen, Z.-J.; Zhang, T.; Gao, X.-Y.; Huang, Y.-J.; Qin, X.-H.;
Wang, Y.-F.; Zhao, K.; Peng, X.; Zhang, C.; Liu, L.; Zeng, M.-H.; Yu, H.-
B. Engineering Microdomains of Oxides in High-Entropy Alloy
Electrodes toward Efficient Oxygen Evolution. Adv. Mater. 2021, 33,
2101845.
(73) Feng, G.; Ning, F.; Song, J.; Shang, H.; Zhang, K.; Ding, Z.; Gao,
P.; Chu, W.; Xia, D. Sub-2 nm Ultrasmall High-Entropy Alloy
Nanoparticles for Extremely Superior Electrocatalytic Hydrogen
Evolution. J. Am. Chem. Soc. 2021, 143, 17117−17127.
(74) Hu, C.; Ma, Q.; Hung, S.-F.; Chen, Z.-N.; Ou, D.; Ren, B.; Chen,
H. M.; Fu, G.; Zheng, N. In Situ Electrochemical Production of
Ultrathin Nickel Nanosheets for Hydrogen Evolution Electrocatalysis.
Chem. 2017, 3, 122−133.
(75) Wang, Z.; Zheng, Y.-R.; Montoya, J.; Hochfilzer, D.; Cao, A.;
Kibsgaard, J.; Chorkendorff, I.; Nørskov, J. K. Origins of the Instability
of Nonprecious Hydrogen Evolution Reaction Catalysts at Open-
Circuit Potential. ACS Energy Lett. 2021, 6, 2268−2274.
(76) Jia, Z.; Yang, T.; Sun, L.; Zhao, Y.; Li, W.; Luan, J.; Lyu, F.;
Zhang, L.-C.; Kruzic, J. J.; Kai, J.-J.; Huang, J. C.; Lu, J.; Liu, C. T. A
Novel Multinary Intermetallic as an Active Electrocatalyst for
Hydrogen Evolution. Adv. Mater. 2020, 32, 2000385.
(77) Yao, R.-Q.; Zhou, Y.-T.; Shi, H.; Wan, W.-B.; Zhang, Q.-H.; Gu,
L.; Zhu, Y.-F.; Wen, Z.; Lang, X.-Y.; Jiang, Q. Nanoporous Surface
High-Entropy Alloys as Highly Efficient Multisite Electrocatalysts for
Nonacidic Hydrogen Evolution Reaction. Adv. Funct. Mater. 2021, 31,
2009613.
(78) Suermann, M.; Schmidt, T. J.; Büchi, F. N. Comparing the
Kinetic Activation Energy of the Oxygen Evolution and Reduction
Reactions. Electrochim. Acta 2018, 281, 466−471.
(79) Segall, M. D.; Lindan, P. J. D.; Probert, M. J.; Pickard, C. J.;
Hasnip, P. J.; Clark, S. J.; Payne, M. C. First-Principles Simulation:
Ideas, Illustrations and the Castep Code. J. Phys.: Condens. Matter 2002,
14, 2717−2744.
(80) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient
Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865−3868.

Recommended by ACS
A Bimetallic Phosphide@Hydroxide Interface for High-
Performance 5-Hydroxymethylfurfural Electro-Valorization
Ruipeng Luo, Siyu Ye, et al.
MARCH 02, 2023
THE JOURNAL OF PHYSICAL CHEMISTRY C READ

Electrically Conductive Ni-P Nanoporous Membrane


Reactors for Electrochemical Reductive Dechlorination of
Organic Pollutants
Nan Zhang, Charles-François de Lannoy, et al.
MAY 22, 2023
ACS APPLIED NANO MATERIALS READ

Reaction Mechanism and Selectivity Tuning of Propene


Oxidation at the Electrochemical Interface
Xiao-Chen Liu, Shi-Gang Sun, et al.
NOVEMBER 07, 2022
JOURNAL OF THE AMERICAN CHEMICAL SOCIETY READ

Phosphate Group Dependent Metallic Co(OH)2 toward


Hydrogen Evolution in Alkali for the Industrial Current
Density
Huanhuan Liu, Xiaoqing Lu, et al.
MAY 16, 2022
ACS SUSTAINABLE CHEMISTRY & ENGINEERING READ

Get More Suggestions >

3603 https://doi.org/10.1021/acsnano.2c01064
ACS Nano 2022, 16, 3593−3603

You might also like