DOI:10.30919/esee8c320

Received: 20 May 2019
Revised: 15 Jul 2019
Accepted: 20 Jul 2019
Published online: 29 Jul 2019

Update on Recent Designing Strategies of Transition Metal-Based Layered Double Hydroxides Bifunctional electrocatalysts

Zhengyang Cai,a Ping Wang,a* Junhe Yang a and Xianying Wanga*

aSchool of Materials Science and Technology, University of Shanghai for Science and Technology, Jungong Rd.516, 200093 Shanghai, China.

 

*Corresponding Author:

Email: ping.wang@usst.edu.cn; xianyingwang@usst.edu.cn

 

ABSTRACT

Fabrication of efficient, stable and cost-effective bifunctional electrocatalysts to achieve overall water splitting reaction (OWS) has become a task of high priority, in view of the urgent need for the promotion of renewable energy-based hydrogen production systems. Transition metal-based layered double hydroxides (LDHs) as one of promising bifunctional electrocatalysts have been intensely studied, due to the advantages of unique 2D layered structure and outstanding physicochemical properties. Herein, this review aims to summarize the recently advanced strategies on design of LDHs, including nanostructuring, conductive support-based hybrid, defect engineering, anion intercalation, cation doping and related derivatives. Particularly, a thorough literature overview on OWS performance improvement of LDHs is appraised detailedly. The discussions on challenges and future directions are provided, which will shed light on development of better LDHs bifunctional electrocatalysts and exploration of efficient device units for practical applications.

Table of Content

This review is mainly focused on the recent advances on designing strategies of LDHs bifunctional electrocatalysts for overall water splitting.

 

Keywords: Layered Double Hydroxides; Bifunctional Electrocatalysts; Hydro


1. Introduction

     Hydrogen production via water electrolysis is regarded as being a green and efficient technology.1, 2 However, the oxygen evolution reaction (OER) involving complex four-electron redox process is a challenging in water-splitting due to its sluggish kinetics.3,4 It often requires electrocatalysts to reduce the overpotential and promote the reaction rate. Precious metal-based materials with the virtue of high activity are usually used as electrocatalysts for hydrogen evolution reaction (HER) (such as Pt or Pt/C) and oxygen evolution reaction (OER) (such as RuO2 or IrO2).5-8 However, the disadvantages of high cost, scarcity, and poor stability greatly hinder the large scale implementation.9, 10 In the past few years, transition metal-based materials as electrocatalysts have been extensively studied, such as phosphides,8, 11-15 chalcogenides,16-21 carbides,22,23 and nitrides24,25 for HER, and phosphates,26,27 oxides,28-33 hydroxides,34-38 perovskite,39 nitrides,40-43 and chalcogenides44-46 for OER. However, the ideal OWS system for commercial utilizations should be driven by one electrocatalytic material, which should be highly active and durable for both HER and OER in the same electrolyte solution. Thus, there is an urgent need to find highly efficient, stable and cost-effective bifunctional electrocatalysts that could lower the required overpotential for OWS and thereby reduce electric energy consumption.47-49

    Among the developed electrocatalysts as outlined above, transition metal-based layered double hydroxides (LDHs) bifunctional electrocatalysts have gained a great deal of research attentions. As an important member of two-dimensional (2D) layered materials family, LDHs are typically composed of a positively charged hydrotalcite-like main layer and interlayer anions. The chemical formula can be expressed as [M2+1-xM3+x(OH)2][An-]x/n·zH2O, where M2+ represents a divalent metal cation, M3+ represents a trivalent metal cation, and An- are interlayer anions (Fig. 1b).50-52 The unique 2D layered structure gives LDHs many advantages: i) metal cations can be flexibly tuned in the bulk layer; ii) types of interlayer anions can be exchangeable with controlled interlayer spaces; iii) by aid of external driving forces, like simple ultrasonication procedure, the increase of interlayer spaces and even the exfoliation of bulk LDHs into ultrathin nanosheets can be easily achieved.

      The great achievements of LDHs in related electrocatalytic applications for HER or OER have been appraised in several excellent reviews.51,53-56 The strategies on structural design and chemical modification of LDHs bifunctional electrocatalysts have been widely developed. However, to the best of our knowledge, the summary of LDHs as bifunctional electrocatalysts for OWS is rarely reported. This review will be mainly focused on the recent advances on designing strategies of LDHs bifunctional electrocatalysts for OWS. A brief summary on synthetic methods of LDHs bifunctional electrocatalysts and related electrocatalytic mechanism for OWS will be firstly given. Subsequently, the advanced design strategies of LDHs bifunctional electrocatalysts will be highlighted, including nanostructuring, conductive support-based hybrids, defect engineering, anion intercalation, cation doping and related LDHs derivatives. In the final section, the challenges and perspectives are proposed, which if fully addressed, will lead to substantial breakthroughs in optimized energy system configuration.

 

2. Basic electrocatalytic mechanism for alkaline overall water splitting

     The OWS reaction (Eq. 3) contains two half-reactions, namely the cathodic HER (Eq. 4 and 5) and the anodic OER (Eq. 6 and 7), as shown in Fig. 1a. The catalytic mechanisms of HER and OER take place on the active sites (M) in alkaline media are also illustrated.

     Ideally, an electrocatalytic reaction can occur when the applied voltage is equal to the equilibrium potential (E0). However, due to the existence of reaction energy barriers, the electrocatalytic reaction can be carried out only when the applied potential (E) is higher than E0. The overpotential (η) applied can be formulated as follow:

η = E - E0                                                                   (1)

And the relationship of current density (j) and applied overpotential (η) can be expressed by the Tafel equation as follow:

η = b log(j/j0)                                                              (2)

where b is the Tafel slope, j0 is the exchange current density which can be obtained when η = 0.

     The Tafel slope not only indicates the growth rate of current density, but also can be used for inference of the rate-limiting step of electrocatalytic reactions.54,57 Thus, the evaluation of an excellent bifunctional catalyst should be with a small η and a small b, which is desirable for both the HER and OER.

      Moreover, the long-term stability of catalysts is also a very important index. Two kinds of measurement methods are usually applied to assess the stability of catalysts. The one is the chronopotentionmetry curves at constant current densities for a long period of time. The other one is to compare the LSV curves before and after a large amount of cyclic voltammetry cycles. In the harsh industrial applications the catalysts need to be stable for an extremely long time (tens of thousands of hours).58, 59

OWS:                       

2H2O (l) → 2H2 (g) + O2 (g)                                                                         (3)

HER (acidic solution):

2H+ (aq) + 2e- → H2 (g)          (E0 = 0 V vs. RHE)                                       (4)

HER (neutral or alkaline solution):

2H2O (l) + 2e- → H2 (g) + 2OH- (aq)   (E0 = 0 V vs. RHE)                         (5)

OER (acidic solution):

2H2O (l) → 4H+ (aq) + O2 (g) + 4e-     (E0 = 1.23 V vs. RHE)                   (6)

OER (neutral or alkaline solution):

4OH- (aq) → 2H2O (l) + O2 (g) + 2e-   (E0 = 1.23 V vs. RHE)                   (7)

 

 

Fig. 1. a) Schematic illustration for alkaline overall water splitting reactions.b) The idealized structure of carbonate-intercalated LDHs with different M2+/M3+ molar ratios showing the metal hydroxide octahedra stacked along the crystallographic c-axis, as well as water and anions present in the interlayer region. Reproduced with permission from Ref. 50. Copyright 2014, The Royal Society of Chemistry. c) Schematic representation of the electrocatalytic generation of H2 in alkaline media. FeOOH and Ni-H are the formed surface adsorbed intermediates during HER process, while, γ-NiOOH is the observed surface intermediate under OER process. Reproduced with permission from Ref. 60. Copyright 2019, The Royal Society of Chemistry.

 

    To date, a large amount of research works have been put forward on studying the catalytic mechanism of LDHs, with development of cutting-edge characterization and simulation methods, such as operando X-ray absorption spectrum (XAS),61,62 in-situ Raman,60 in-situ transmission electron microscopy (TEM)63 and density functional theory+U (DFT+U). Particularly for incorporation of heteroatoms in the LDH layers, the optimized doping ratio of different metal or non-metal elements in the LDH layer can greatly improve the intrinsic activity of the catalysts, that is, the electronic configuration of the metal sites (the exposed metal sites serve as active sites) can be adjusted, which can promote the synergistic effect among the active sites and thus improve the charge distribution in the LDH layer and the charge transfer capability. Furthermore, the correlation between Gibbs free energy of the reaction rate determining step and the enhancement in catalytic activity can be well established. For instance, the OWS mechanism of NiFe LDH via employment of in-situ Raman spectro-electrochemistry was proposed by Qiu et al.60 The transformation of different active interfacial species during OER and HER unveiled the synergistic interaction between iron and nickel for facilitating water electrolysis processes (Fig. 1c).

 

3. Synthetic Approaches of LDHs Bifunctional Electrocatalysts

    Several synthetic approaches have been explored for preparation of LDHs, mainly including coprecipitation, hydro/solvothermal synthesis, cathodic electrodeposition and electrospinning process (Table 1).64 One of the most used methods is the coprecipitation method. Homogenous LDHs crystals can be quickly prepared by mixing metal salts mixture as precursors and aqueous alkaline solution at a constant temperature under sufficiently vigorous stirring.65-67 Hydro/solvothermal synthetic methods are also usually employed by heating metal salts mixture and a weak alkali precipitant (such as urea,68, 69 ammonia, methanol70 or NaOH71) at a desired temperature in water/ organic solvents (such as ethanol72 and dimethylformamide73). In the cases, the morphology and structure of LDHs can be easily controlled by pH value, reaction temperature and time. Moreover, cathodic electrodeposition method is more convenient and time-saving. Metal nitrate or sulfate can be used as electrolytes for providing a weak alkaline environment.64, 74 However, it commonly suffers the difficulty in morphology control, leading to rough surfaces of the resulting LDHs.75 The electrospinning technique, which is known to be simple, low-cost, and eco-friendly, is also adopted for preparation of LDHs with controllable nanofiber morphologies.76

Table 1 Common synthetic approaches for LDHs

Approaches

Precursors

Temperature (oC)

Morphology/

Structure

Cost

Coprecipitation

Metal salts, alkali solution

< 100

Bulk sheets, Crystal

Medium

Hydro/Solvothermal

Metal salts, precipitant, and solvent

80-120

Bulk sheets, Crystal

High

Electrodeposition

Metal salts and solvent

Room temperature

Few-layer nanosheets, Amorphous

Low

Electrospinning

Metal salts, conductive polymer and solvent

Room temperature

Nanofibers, Amorphous

Low

 

4. Advanced Design Strategies for LDHs Bifunctional Electrocatalysts

    It is well known that bulk thickness of LDHs with a large lateral size commonly suffered poor electrocatalytic performance, owing to the limited numbers of active sites with poor intrinsic activity and low electron conductivity.77,78 In order to address these shortcomings, several design strategies have been developed (Table 2), specifically for nanostructuring, self-assembling on conductive substrates, defect engineering, ion-intercalation, cation doping and some related LDHs derivatives. In the chapter, the great successes gained with the respect to design strategies will be detailedly discussed.

Table 2 Comparison recently reported LDHs-based electrocatalysts for OWS in 1 M KOH aqueous solution.

Strategies

Electrocatalysts

Synthesis Method

Mass Loading
(mg·cm-2)

η10c (mV)
OER / HER

Overall Voltage (V)
@ 10 mA·cm-2

Tafel Slope (mV·dec−1 )
OER/HER

OWS Stability
Time @ J mA·cm-2

Ref.

Year

Nanostructuring
&
Conductive support-based hybrids

Co@N-CS/N-HCP@CC

ECDa + Pseudomorphic Replication + Pyrolysis

3.20

248 / -66

1.545

68 / 65

24 h @ 30

79

2019

Cu@NiFe LDH

Chemical oxidation + Calcination + Electroreduction + ECD

2.20

199 / -116

1.540

27.8 / 58.9

48 h @ 10

80

2017

CoFe LDH-F

Delamination + Exfoliation

0.20

300 / -255

1.630

40 / 95

35 h @ 1.63 V

81

2016

Ni–Co–P HNBs

Sacrificial template + Etching + Phosphorization

2.00

270 / -107

1.620

76 / 46

20 h @ 1.62 V

82

2018

NiFe LDH@NiCoP/NF

HTb + Phosphorization + HT

2.00

220 / -120

1.570

88.2 / 48.6

100 h @ 10

83

2018

 Co3S4 @MoS2

Sulfidation + Coating + Annealing

0.28

280 / -136

1.580

43 / 74

10 h @ 10

84

2018

CoNi/CoFe2O4/NF

HT + Calcination + ECD

2.10

230 / -82

1.570

45 / 96

48 h @ 10

85

2018

Cu@CoFe LDH

Chemical oxidation + Calcination + Electroreduction + ECD

1.80

240 / -171

1.681

44.4 / 36.4

48 h @ 10

86

2017

NiCo2S4@NiFe LDH

HT + HT

——

201 @ η60 / -200

1.600

46.3 / 101.1

12 h @ 10

87

2017

 NiCo2O4 hollow microcuboids

Solvothermal + Annealing

1.00

230 @ η1 / -50 @ η1

1.650

53.0 / 49.7

36 h @ 10

88

2016

VOOH nanospheres

HT

0.80

270 / -164

1.620

68 / 104

50 h @ 50

89

2017

Ni5Fe LDH@NF

HT

——

210 / -133

1.590

59 / 89

20 h @ 50

90

2017

NiFe-LDH/NiCo2O4/NF

HT + Annealing + HT

4.90

290 @ η50 / -192

1.600

53 / 59

12 h @ 20

91

2017

EG/Co0.85Se/NiFe-LDH

Exfoliation + HT

4.00

270 @ η150 / -260

1.670

57 / 160

10 h @ 20

92

2015

NiFe LDH-NS@DG10

Exfoliation + Stirring

2.00

210 / -115 @ η20

1.5 @ 20 mA·cm-2

52 / 110

——

93

2017

NiCo2Px/CNTs

Coprecipitation + Annealing

0.10

284 / -47

1.610

50.3 / 57.0

48 h @ 10

94

2018

Defect engineering

δ-FeOOH NSs/NF

Wet-chemical

0.16

265 / -108

1.620

36 / 68

60 h @ 1.73 V

95

2018

NiAlδP/NF

HT +  Alkali-etching + Phosphorization

2.00

242 / -80

1.550

65 / 52

——

96

2018

NiCoP/NF

HT + Plasma

1.60

280 / -32

1.580

87 / 37

24 h @ 10

97

2016

Ni3S2@NiV-LDH/NF

HT + HT

0.71

190 / -126

1.530

57 / 90

160 h @ 10

98

2019

N-NiCo LDHs/NCF

HT + Plasma

——

190 / -100

1.500

46 / 123

——

99

2019

Intercalation

e-ICLDH@GDY/NF

HT + In-situ Exfoliation

——

216 / -43

1.430

43.6 / 98.9

60 h @ 1.56 V

100

2018

NiFe LDH-US

ECD +  In-situ intercalation

——

203 / —

——

42 / —

——

101

2017

Na0.08Ni0.9Fe0.1O2

Calcination + Chemical extraction

0.13

260 / —

——

40 / —

——

102

2016

Cation doping

CoFe@NiFe-200/NF

HT + ECD

——

190 / -240

1.590

45.71 / 84.69

24 h @ 10

103

2019

NiFeIr LDH

HT

——

200 / -34

1.41 @ 20 mA·cm-2

— / 32

50 h @ 20

104

2018

 Mo-CoP

Wet-chemical + Phosphorization

2.50

305 / -40

1.560

56 / 65

20 h @ 1.5 V

105

2018

PA-NiO

HT +  Phosphorization

——

290 @ η60 / -138

1.560

36 / 81

7 h @ 1.6 V

106

2018

LDHs Derivative

 FeCoNi-HNTAs

HT

1.00

184 / -58

1.429

49.9 / 37.5

100 h @ 1.59 V

107

2018

Se-(NiCo)S/(OH)

HT

——

155 / -103

1.600

33.9 / 87.3

66 h @ 10

108

2018

NiCo2S4

HT + HT

——

243 / -80

1.580

54.9 / 58.5

72 h @ 1.64 V

109

2019

RuO2/NiO/NF

HT + Impreganation + Calcination

1.10

250 / -22

1.500

50.5 / 31.7

25 h @ 10

110

2018

FeCoP

ECD + Phosphorization

——

260 @ η20 / -188 @ η100

1.600

63 / 76

20 h @ 1.606 V

111

2017

MFN-MOFs/NF

Solvothermal

——

235 @ η50 / -79

1.495

55.4 / 30.1

100 h @ 500

112

2018

Ni2Cr1-LDH/NF

HT

2

319 @ η100 / -67

1.550

22.9 / 61.5

30 h @ 1.55 V

113

2018

Note: a Electrochemical deposition, b Hydrothermal, c Overpotential required at the current density of 10 mA·cm-2.

4.1  Nanostructuring

     Fabrication of hierarchical nanostructures or ultrathin nanosheets is of great benefit to maximize the number of active sites via increase of surface areas, and thus facilitate electrolyte diffusion and gases desorption.78, 114 Hu and co-workers meticulously designed a hierarchical Ni-Co-P hollow nanobrick (HNBs) comprised of oriented nanosheets (Fig. 2a and b).82 By employing 3D Ag2WO4 anisotropic cuboids as sacrificial templates, HNBs were fabricated by growth of oriented 2D Ni-Co nanosheets precursor and subsequent etching/phosphorization treatments (Fig. 2c). The unique hierarchical and hollow structures were favorable for OWS performance enhancement of HNBs, as a result of significant increase in the exposed surface area and abundant mass diffusion pathways. Moreover, it was found that small amount of Ag residual also contributed to the increase in catalytic activity, based on the density functional theory (DFT) calculations (Fig. 2d and e). Upon incorporation of Ag, the ΔGH* became relatively close to 0 eV for the catalyst-H* state. It indicated that the faster proton/electron transfer/hydrogen release rate and the much larger water adsorption energy were realized, promoting HER processes.

Fig. 2 a) FESEM and  b) TEM images of Ni-Co-O HNBs, c) Schematic illustration of construction of hierarchical Ni-Co-P hollow nanobricks, d) Calculated free-energy change and e) water adsorption energies for Ni-Co-P and Ag-incorporated Ni-Co-P. Reproduced with permission from Ref. 82. Copyright 2018, The Royal Society of Chemistry.

 

  Another interesting strategy is to fabricate self-standing 3D core-shell LDHs nanostructures. Cu nanowires@NiFe LDH nanosheets with a 3D core/shell structure were successfully synthesized via electrodeposition by Yu and the co-workers. for highly efficient OWS.80 Cu nanowires (NWs) cores supported on Cu foams were firslty prepared by three-step pre-synthesis procedures, namely chemical oxidation, calcination for phase transformation and electroreduction processes. Subsequently, few-layer NiFe LDH nanosheets were grown on Cu NWs cores (Fig. 3a). As shown in Fig. 3b and c, the NiFe LDH uniformly and vertically grew on Cu NWs, forming a core-shell structure. The Cu@NiFe LDH electrocatalyst exhibited excellent OWS performance, as listed in Table 2. It was attributed to the synergistic combination of NiFe LDHs and Cu NWs, which increased the edge sites exposure of LDHs and facilitated desorption of gases product, but also ensured good conductivity and mechanical stability.

    Besides, an alternative approach to boost the OWS performance is to exfoliate bulk LDHs into ultrathin nanosheets. Liu and co-workers exfoliated CO32- intercalated CoFe LDHs anions (CoFe LDH-C) in DMF-ethanol mixed solvents.81 The resulted CoFe LDH-F possessed much smaller thickness of ~4.5 nm, compared to that of CoFe LDH-C (~15 nm) (Fig. 3d and e). During the exfoliation process, plentiful of coordination-unsaturated transition metals associated with oxygen vacancies were created, as demonstrated by X-ray photoelectron spectroscopy (XPS) and electron spin resonance (ESR) spectroscopy (Fig. 3f and g). The generated oxygen vacancies can decrease energy barrier for adsorption of OH- anions, because of the low-coordination sites of MO6 and improved intrinsic electronic conductivity, thus leading to small OER and HER overpotentials.

 Fig. 3 a) Schematic illustration of the fabrication procedures of the self-standing 3D core-shell Cu@NiFe LDH electrocatalysts. (RT is denoted by room temperature.) b-c) SEM and TEM images of Cu@NiFe LDH. Reproduced with permission from Ref. 80. Copyright 2017, The Royal Society of Chemistry. d) TEM, AFM images and height profiles of CoFe LDH-C. e) TEM, AFM images and height profiles of CoFe LDH-F. f-g) O 1s XPS spectra and EPR spectra of the CoFe LDH-C and CoFe LDH-F samples, respectively. Reproduced with permission from Ref. 81. Copyright 2016, American Chemical Society.

 

4.2 Conductive support-based hybrids

     Self-supporting electrocatalysts on conductive substrates are an attractive approach for enhancing the number of catalytic active sites and the charge transfer ability, but also simplifying the electrode preparation process.107, 115-118 Hou et al. reported a ternary EG (exfoliated graphene foil)/Co0.85Se/NiFe-LDH hybrid (Fig. 4a).92 The field emission scanning electron microscopy (FESEM) images show the thickness of NiFe-LDHs grown on EG/Co0.85Se nanoarrays can be reduced to be about 10 nm (Fig. 4b-e). The catalyst with a high surface area of 156 m2·g-1 can achieve a current density of 20 mA·cm-2 at an external voltage of 1.71 V for OWS in a two-electrode cell (Table 2.). It was attributed to the strong coupling effect of the three components, ensuring low charge transfer resistance and fast vectorial electron-transport. Especially, the 3D layered structure could offer large amounts of active sites and accelerate gas bubbles release.

Fig. 4 a) Schematic illustration for the synthesis process of EG/Co0.85Se/NiFe-LDH. b-c) FESEM images of EG/Co0.85Se, Inset: the enlarged FESEM image of EG/Co0.85Se. d-e) FESEM images of EG/Co0.85Se/NiFe-LDH. Reproduced with permission from Ref. 92. Copyright 2016, The Royal Society of Chemistry.

 

   Taking the plentiful advantages of defective graphene (DG), like fast electron transfer kinetics, excellent stability and specific defect types, DG has been widely employed for OER and HER. Jia et al. reported the exfoliated NiFe LDHs nanosheets (NS)/DG by electrostatic stacking of positively charged NiFe LDH NS and negatively charged DG (Fig. 5a).93 The different types of DG defects derived by the removal of heteroatoms from graphene acted as active sites, which can directly capture the transition metal atoms via strong π-π interaction and thus contribute to activities improvement for three basic electrochemical reactions (e.g., ORR, OER and HER). On the other hand, the charge distribution on the hybrid heterostructure was demonstrated by DFT calculation study (Fig. 5b and c). The electrons were re-distributed after assembling of NiFe LDH-NS and DG, and accumulated at the defect sites (blue dash frames in Fig. 5c). Therefore the outstanding OER and HER activities can be ascribed to the exfoliated NiFe LDH and the charge conductivity of DG.

    Carbon nanotubes (CNTs) are also an effective substrate to enhance the electron transfer of LDHs. The electronegative oxygen functional groups on surface of CNTs can promote efficient dispersion of catalysts. Huang and co-workers constructed 3D nicke-cobalt bimetallic phosphide anchored on CNTs (NiCo2Px/CNTs). The catalyst exhibited enhanced catalytic performance toward both HER and OER (Fig. 5d). 94 The high electrocatalytic activity was attributed to the synergistic effect between NiCo2Px and CNTs (Fig. 5e).

Fig. 5 a) Schematic illustration of the preparation of NiFe LDH-NS@DG nanocomposite. b) The top views of optimized Ni-Fe LDH-NS@DG (DG-5, DG-585, or DG-5775) based composite interfaces. c) The side views of 3D charge density difference plot for the interfaces between a defective graphene sheet (DG-5, DG-585 or DG-5775) and a Ni-Fe LDH-NS layer. Yellow and cyan isosurfaces represent the charge accumulation and depletion layers in the 3D space with an isosurface value of 0.002 e Å-3. Green, brown, silver, and red balls represent C, Fe, Ni, and O atoms, respectively. The different defect types and associated enhanced charge density areas are marked with a blue solid line and a blue dash line, respectively. Reproduced with permission from Ref.  93. Copyright 2017, Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim. d) Schematic illustration of the synthetic route for NiCo2Px/CNTs. e) Mechanism and the synergetic effect of NiCo2Px/CNTs for both the HER and OER. Reproduced with permission from Ref. 94. Copyright 2018, The Royal Society of Chemistry.

 

4.3. Defect engineering

     In recent years, much attention has been paid to the defect engineering of LDHs,71, 114, 119-122 i.e., creation of oxygen vacancies, cation vacancies and their mixture. Liu et al. in 2018 firstly proposed synthesis of ultrathin feroxyhyte nanosheets (δ‐FeOOH NSs) with richful Fe vacancies (VFe).95 The origin of the superior HER and OER reactivity of δ‐FeOOH NSs was clearly validated. Introduction of VFe led to obvious upshift on the Fe K absorption edge of δ‐FeOOH NSs (Fig. 6a), showing a reduction of the electron density of Fe atoms in NSs. As seen from the R space plot (Fig. 6b), first Fe-O shell was calculated to be a coordination number (N) of 5.5, which was close to the theoretical value of 6 for an octahedral and the second Fe-Fe shell in NSs was reduced by 20% to 1.6 in comparison with 2 for the bulk. It indicated the existence of rich VFe in δ‐FeOOH NSs. It was also reflected by the upshift of Fe 2p peaks in δ‐FeOOH NSs (Fig. 6c). The formation of VFe can activate the second neighboring surface Fe atom (Fe2 site). The DFT results also demonstrated that the Fe2 site had a small ΔGH*, giving rise to the enhanced HER performance of δ‐FeOOH NSs (Fig. 6d). The Fe2 site acted as the OER active site, which possessed the smallest rate-determining ΔG value (Fig. 6e) and the four primitive OER steps on the electrocatalyst surface were proposed, as shown in Fig. 6f.

Fig. 6 a) Fe  K-edge XANES spectra. b) the corresponding Fourier transformations of Fe K-edge EXAFS spectra for bulk δ-FeOOH and δ-FeOOH NSs. c) The high-resolution XPS spectra of Fe 2p of bulk δ-FeOOH and δ-FeOOH NSs. d) Standard free energy diagram of the HER process on the O and Fe in δ-FeOOH NSs without VFe (marked as δ-FeOOH (001)) and the O neighboring to VFe and the second neighboring Fe to VFe (Fe2) in δ-FeOOH NSs with VFe (marked as VFe). e) Standard free energy diagram of the OER processes on the δ-FeOOH (001), and Fe1 and Fe2 sites on δ-FeOOH (001) NSs with VFe. Fe1 indicates the first neighboring Fe to VFe, and Fe2 indicates the second neighboring Fe to VFe. f) Primitive steps of the OER process on the surface of δ-FeOOH NSs with VFe (top view). The Roman numerals represent (i) adsorption step, (ii and iii) dissociation steps, and (iv) desorption step during the OER process. Brown: Fe; red: O. Reproduced with permission from Ref. 95. Copyright 2018, Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.

 

    Cheng et al. developed a new type of 3D metal-vacancy-solid-solution NiAlP nanowall array (NiAlδP) for all-pH OWS by combination of selective alkali-etching and phosphorization methods.96 Numerous nanoholes were well distributed throughout each nanowall forming a unique porous structure and massive atomic vacancies are observed over the lattice of the NiAlP nanowalls (Fig. 7b and c). The XPS spectra revealed that the existence of surface aluminumvacancy led to a positive shift of 0.6 eV for Ni 2p3/2 and negative shift of 0.3 eV for P 2p (Fig. 7d and e). The theoretical calculations indicated that the NiAlδP possessed a low ΔGH* value -0.09 eV (close to that of Pt, ~0 eV) and the highest H2O adsorption energy, favoring the water adsorption (Fig. 7f). The oxidized Ni and negatively charged P in NiAlδP might act as the active sites for OER and HER, respectively.

Fig. 7 a-c)   SEM, TEN and HRTEM images for the NiAlδP nanowall array. d-e) Ni 2p and P 2p XPS spectra. f) Theoretical H* adsorption Gibbs free energy and H2O* adsorption energy for NiAlP, NiAlδP, and Pt or RuO2. Reproduced with permission from Ref. 96. Copyright 2018, The Royal Society of Chemistry.

 

4.4. Ion intercalation

   In addition to spontaneously inserting different anions or molecules into LDHs layers during preparation processes, the interlayer anions can be further replaced by using ion-exchange methods. It can enlarge the interlayer spacing, but also can exfoliate the bulk LDHs into monolayer or few-layer nanosheets which can increase the number of active sites and thus enhance catalytic activity.123 Li et al. demonstrated an in-situ intercalation method to increase the interlayer spacing of NiFe LDHs (Fig. 8a).101 As shown in Fig. 8b, XRD pattern shows that after formamide intercalation reaction at 80 °C for 3 h, the (003) diffraction peak of NiFe LDH shifted from 11.6° to t 9.3° (NiFe LDH-80 °C), corresponding to the increase of interlayer distance from 7.8 to 9.5 Å. With the assistance of ultrasound, the interlayer distance can increase to 9.3 Å at 30 ℃ for only 10 min (NiFe LDH-US-30 °C). The extended interlayer spacing offered more inner active sites and more space for diffusion of the reactants, the OER overpotential to achieve the current density of 10 mA·cm-2 was reduced to 210 and 203 mV for NiFe LDH-80 °C and NiFe LDH-US-30 °C, compared to that of 256 mV for pristine NiFe LDH. Similarly, Na intercalated NiFeOx (Na1−xNiyFe1−yO2) was also studied (Fig. 8c), consisting of [MO6] (M = Ni, Fe) octahedral layers with Na atoms residual lying between the octahedral layers.102 The existence of Na led to transformation of Ni and Fe into high chemical states (Fig. 8d and e). Moreover, the extraction of Na gave rise to the exposure of more [MO6] active sites, leading to the improved activity. However, the excessive Na intercalation will deteriorate the structural stability.

Fig. 8 a) Schematic diagram of in-situ intercalation process over electrodeposited NiFe LDH nanosheets on substrate. b) XRD patterns of the fresh NiFe LDH coated electrode (black) and intercalated NiFe LDH coated electrodes at 80 °C (red) and 30 °C with ultrasound (blue).Reproduced with permission from Ref. 101. Copyright 2017, Elsevier B.V. c) Crystal structure of O3-type NaNiyFe1-yO2. d-e) Ni 2p and Fe 2p XPS spectra. Reproduced with permission from Ref. 102. Copyright 2017, The Royal Society of Chemistry.

 

4.5. Cation doping

    Cation doping is considered to be an effective pathway for increasing OWS activity, since doping of the third metal ion can effectively adjust the electronic configuration and conductivity of LDHs and generate synergistic effect between metal ions and LDHs layers.124-126 Chen and co-workers in 2018 reported a novel strategy to accelerate HER kinetics of NiFe LDH by Ir4+-doping (NiFeIr LDHs). The water dissociation process (Volmer step) can be accelerated, thus leading to a robust catalytic activity for OWS.104 The Tafel slope of NiFe LDH was estimated to be about 125 mV·dec-1 (Fig. 9a), indicating that the Volmer reaction (H2O + e- → Had + OH-) was the rate-determining step. While NiFeIr LDH possessed a much smaller Tafel slope of 32 m V·dec-1, in accordance with the Volmer-Tafel mechanism (H2O + e- → Had + OH-, 2Had → H2).

Fig. 9 a) Tafel slopes of NiFeIr LDH (violet), NiFe LDH (green) and Pt/C (black). Reproduced with permission from Ref. 104. Copyright 2018, The Royal Society of Chemistry. b) Schematic fabrication process and the resulting microstructures for hollow Mo-CoP nanoarrays. c) HER free energy diagrams for the P- and Co-sites on pristine and Mo-doped CoP (111) surfaces. d) Standard free energy diagrams for the OER path on pristine (green) and Mo-doped (orange) β-CoOOH (001) surfaces. e) Bader charge analysis for surface Co ions and H in HO*adsorption on pristine and Mo-doped β-CoOOH (001) surfaces. Reproduced with permission from Ref. 105. Copyright 2018, Elsevier B.V.

 

    Rational design of hollow Mo-doped CoP (Mo-CoP) nanoarrays was recently proposed by Guan et al. (Fig. 9b).105 DFT calculation results show the ΔGH* value of the P-site on Mo-CoP (111) surface dropped dramatically to 0.07 eV, indicating much superior HER performance (Fig. 9c). For OER (Fig. 9d) the potential limiting step (PLS) for β-CoOOH was determined to be the deprotonation step (HO* → O* + H+ + e-). Accordingly, the ΔG2 value was reduced from 1.95 eV to 1.82 eV by doping of Mo. As also evidenced by the reduced Bader charge of the Co and H (Fig. 9e), the electrons will transfer from Mo dopants to the nearby Co species and the H in HO* adsorption. The Mo-induced electron transfer towards the H can reduce the strength of the H-O bond in adsorbed HO*, thus facilitating the PLS process and enhancing the OER activity.

Fig. 10 a) HRTEM images of Se-(NiCo)S/OH nanosheets. Reproduced with permission from Ref. 108. Copyright 2018, Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim. b) Schematic representation for synthetic procedure of FeCoNi-HNTAs. c) FESEM images of FeCoNi-HNTAs, Scale bars: 2 μm and inset: 200 nm. d-f) Normalized Fe, Co, and Ni L-edge sXAS spectra in the total electron yield (TEY) mode. g) Under-electrolyte Gas bubble adhesive force measurements of FeCoNi-HNTAs. Reproduced with permission from Ref. 107. Copyright 2018, Springer Nature Publishing AG.

 

4.6. Other LDHs derivatives

    Besides the above-mentioned design strategies, fabrication of LDHs derivatives is also of great interest by combination of two or multi design strategies for modification of LDHs (e.g. using multi-element doping and unique morphology design). Based on the theoretical expectations, transition metal chalcogenides (Se-(NiCo)Sx/(OH)x nanosheets) for highly efficient OWS was demonstrated by Hu et al.108 A large amount of Se-O and S-O bonds can be formed in the nanosheets, which accelerated the charge transfer, and introduced abundant disordering and defects (Fig. 10a). Taking consideration of the remarkable electrocatalytic HER activity of metallic monoclinic 1T phase MoS2 and excellent OER activity of trimetallic cobalt/iron/nickel-based (Co, Fe, Ni-based) sulfides, Li and co-workers reported synthesis of hybrid nanotube arrays (FeCoNi-HNTAs) composed of crystalline structures of 1T MoS2 and (Co, Fe, Ni-based)9S8 (Fig. 10b and c). 107 The prominent activity was attributed to systematic optimization of chemical composition and geometric structure. Interestingly, the synergistic effects among Fe, Co and Ni ions were proved by aid of L-edge soft X-ray absorption spectroscopy (sXAS). Friebel et al.127 further demonstrate that the Fe3+ indeed acted as the highly active sites for OER. The observation indicated that the Co ions can change the electronic state of Fe from Fe2+ to Fe3+ (Fig. 10d). The crystal-field coordination of Co ions can be tuned by Fe ions, that Co3+ ions located at octahedral sites (the small shoulder at 780.4 eV in FeCoNi-HNTAs) are the active centers for OER rather than the Co2+ ions at tetrahedral sites (778.2 eV) (Fig. 10e). And the XPS peak shoulder was also observed, indicating the existence of Ni3+ as the OER active sites (Fig. 10f). Furthermore, the FeCoNi-HNTAs electrodes showed prominent superaerophobicity and superhydrophilicity with no gases bubble adhesive force, which was greatly beneficial for gases evolution reaction and mass transfer (Fig. 10g).

 

4.7. Relevant applications

     Although the electrocatalytic water electrolysis is regarded as a promising pathway for hydrogen production, the main obstacle remains the high energy consumption for potential practice. It is suggested that the effective, reliable and sustainable solution is to combine high-efficiency bifunctional OWS electrocatalysts with renewable energy sources, such as solar energy, wind energy, hydropower, etc. The renewable energy can be converted into hydrogen energy for storage, and then burn directly as fuel or used in more efficient fuel cell devices for electricity generation.

     Many efforts have been devoted to exploiting the renewable energy-conversion systems, especially solar energy-based water splitting (Fig. 11).93, 128-130 Luo and co-workers successfully designed an efficient and intrinsically safe bias-free solar-driven water-splitting device composed of perovskite light harvesters, non-precious metal based catalysts, and a bipolar membrane (Fig. 11a).130 And a current density of 10 mA·cm-2 was obtained at the bias voltage of 1.63 V for electrochemical water splitting and the solar to hydrogen conversion efficiency is up to 12.7%. Other interesting case is to design hybrid LDHs via combined with semiconductor materials (like hematite,131 BiVO4,132 g-C3N4,133, 134 etc.) for photoelectrocatalytic water splitting. For instance, Yan’s group successfully fabricated a triadic photoanode consisting of dual-sized CdTe quantum dots (QDs), Co-based layered double hydroxide (LDH) nanosheets and BiVO4 particles (QD@LDH@BiVO4). The catalyst possessed a remarkably enhanced PEC performance, achieving a current density of 2.23 mA·cm-2 at 1.23 V vs. RHE.69

Fig.11 a) Schematic diagram of the solar-driven water-splitting device composed of perovskite light harvesters, earth-abundant catalysts, and a bipolar membrane. Reproduced with permission from Ref. 130. Copyright 2016, Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim. b) Demonstration of a solar power assisted water-splitting device with a voltage of 1.5 V. Reproduced with permission from Ref. 93. Copyright 2017, Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.

 

5. Conclusions and outlook

      Although the LDHs have been widely investigated for solely electrocatalytic HER or OER, poor electrocatalytic performance of bulk LDHs restrict the practical applications, owing to the limited numbers of active sites with poor intrinsic activity and low electron conductivity. Moreover, it remains challenging to the design criteria of LDHs bifunctional electrocatalysts for OWS. Therefore, this review summarizes the recent advanced design strategies of LDHs as bifunctional electrocatalysts for OWS. In the review, The current cutting-edge technologies for rational design of LDHs and related mechanisms are provided: (1) Fabrication of nanostructures can improve the density of active sites and enhance mass transfer capacity; (2) In-situ growth of LDHs on various conductive supports can enhance charge transport capacity; (3) Manufacture of various defects can regulate the surface electronic configuration of LDHs to increase the active site types and intrinsic activity; (4) Small molecules or ion intercalation can expand the LDHs layer spacing to increase the number of active sites; (5) Cation doping can enhance the synergistic effect between different metal ions and thereby increase the intrinsic activity; (6) Development of related LDHs derivatives. These strategies not only improve the OER activity of LDHs, but also enhance their performance for HER, ultimately achieving the excellent OWS performance.

      Despite these considerable efforts, there are still some challenges: (1) DFT methods are powerful in catalytic mechanism analysis of LDHs. However, it is hard to identify the actual active sites and determine the synergistic effect in more complexed catalyst components systems. Theoretical evaluation techniques still need to be continuously improved; (2) It is also expected that in-situ or operando characterization techniques, such as ambient-Pressure XPS, in situ XANES and in-situ Raman, can be fully applied to support the real-time mechanism analysis of catalytic reactions. Therefore, it should be considered to include highly efficient and well-structured electrolyzer systems; (3) Almost all LDHs bifunctional electrocatalysts can only run in alkaline environments. It is necessary to explore new strategies to make LDHs suitable for neutral or even acidic environments.

 

Conflicts of interest

There are no conflicts to declare.

 

Acknowledgements

We greatly appreciate the financial supports from the National Natural Science Foundation of China (51572173, 51602197, 51771121, 51772297 and 51702212), Shanghai Municipal Science and Technology Commission (16060502300, 16JC402200 and 18511110600), Shanghai Academic/Technology Research Leader Program (19XD1422900) and Shanghai Eastern Scholar Program (QD2016014).


 

References

1.  S. Chu and A. Majumdar, Nature, 2012, 488, 294-303.

2.  G. Wu, A. Santandreu, W. Kellogg, S. Gupta, O. Ogoke, H. Zhang, H. L. Wang and L. Dai, Nano Energy, 2016, 29, 83-110.  

3.  C. C. McCrory, S. Jung, I. M. Ferrer, S. M. Chatman, J. C. Peters and T. F. Jaramillo, J. Am. Chem. Soc., 2015, 137, 4347-4357.

4.  J. Wang, W. Cui, Q. Liu, Z. Xing, A. M. Asiri and X. Sun, Adv. Mater., 2015, 28, 215-230.

5.  T. Audichon, T. W. Napporn, C. Canaff, C. Morais, C. Comminges and K. B. Kokoh, J. Phys. Chem. C, 2016, 120, 2562-2573.

6.  M. Carmo, D. L. Fritz, J. Merge and D. Stolten, Int. J. Hydrogen Energ., 2013, 38, 4901-4934.

7.  S. Park, D. Yoon, S. Bang, J. Kim, H. Baik, H. Yang and K. Lee, Nanoscale, 2015, 7, 15065-15069.

8.  J. Tian, Q. Liu, A. M. Asiri and X. Sun, J. Am. Chem. Soc., 2014, 136, 7587-7590.

9.  W. T. Hong, M. Risch, K. A. Stoerzinger, A. Grimaud, J. Suntivich and Y. Shao-Horn, Energy Environ. Sci., 2015, 8, 1404-1427

10. T. Reier, Z. Pawolek, S. Cherevko, M. Bruns, T. Jones, D. Teschner, S. Selve, A. Bergmann, H. N. Nong, R. Schloegl, K. J. J. Mayrhofer and P. Strasser, J. Am. Chem. Soc., 2015, 137, 13031-13040.

11. D. Y. Chung, S. W. Jun, G. Yoon, H. Kim, J. M. Yoo, K. S. Lee, T. Kim, H. Shin, A. K. Sinha, S. G. Kwon, K. Kang, T. Hyeon and Y. E. Sung, J. Am. Chem. Soc., 2017, 139, 6669-6674.

 12. P. Jiang, Q. Liu, Y. Liang, J. Tian, A. M. Asiri and X. Sun, Angew. Chem. Int. Ed., 2014, 53, 12855-12859.

13. J. Tian, Q. Liu, N. Cheng, A. M. Asiri and X. Sun, Angew. Chem. Int. Ed., 2014, 53, 9577-9581.

14. Z. Xing, Q. Liu, A. M. Asiri and X. Sun, Adv. Mater., 2014, 26, 5702-5707.

15. Z. H. Xue, H. Su, Q. Y. Yu, B. Zhang, H. H. Wang, X. H. Li and J. S. Chen, Adv. Energy Mater., 2017, 7, 1602355. 

16. D. Kong, H. Wang, Z. Lu and Y. Cui, J. Am. Chem. Soc., 2014, 136, 4897-4900.

17. H. Li, C. Tsai, A. L. Koh, L. Cai, A. W. Contryman, A. H. Fragapane, J. Zhao, H. S. Han, H. C. Manoharan, F. Abild-Pedersen, J. K. Nørskov and X. Zheng, Nat. Mater., 2016, 15, 53.  

18. C. Tang, N. Cheng, Z. Pu, W. Xing and X. Sun, Angew. Chem. Int. Ed., 2015, 54, 9351-9355.

19. D. Voiry, R. Fullon, J. Yang, C. de Carvalho Castro e Silva, R. Kappera, I. Bozkurt, D. Kaplan, M. J. Lagos, P. E. Batson, G. Gupta, Aditya D. Mohite, L. Dong, D. Er, V. B. Shenoy, T. Asefa and M. Chhowalla, Nat. Mater., 2016, 15, 1003.

20. A. Manikandan, P. R. Ilango, C. W. Chen, Y. C. Wang, Y. C. Shih, L. Lee, Z. M. Wang, H. Ko and Y. L. Chueh, J. Mater. Chem. A, 2018, 6, 15320-15329.

21.  A. Alarawi, V. Ramalingam, H. C. Fu, P. Varadhan, R. Yang and J. H. He, Opt. Express, 2019, 27, A352-A363. 

22. W. Cui, N. Cheng, Q. Liu, C. Ge, A. M. Asiri and X. Sun, ACS Catal., 2014, 4, 2658-2661.

23. A. Manikandan, L. Lee, Y. C. Wang, C. W. Chen, Y. Z. Chen, H. Medina, J. Y. Tseng, Z. M. Wang and Y. L. Chueh, J. Mater. Chem. A, 2017, 5, 13320-13328.  

24. B. Cao, G. M. Veith, J. C. Neuefeind, R. R. Adzic and P. G. Khalifah, J. Am. Chem. Soc., 2013, 135, 19186-19192. 

25. Y. Wang, L. Chen, X. Yu, Y. Wang and G. Zheng, Adv. Energy Mater., 2017, 7, 1601390.

26. J. Xu, X. K. Wei, J. D. Costa, J. L. Lado, B. Owens-Baird, L. P. L. Gonçalves, S. P. S. Fernandes, M. Heggen, D. Y. Petrovykh, R. E. Dunin-Borkowski, K. Kovnir and Y. V. Kolen’ko, ACS Catal., 2017, 7, 5450-5455.

27. G. Zhang, G. Wang, Y. Liu, H. Liu, J. Qu and J. Li, J. Am. Chem. Soc., 2016, 138, 14686-14693.  

28. J. Li, B. T. Yonemoto, X. Zhou, K. Zhu and F. Jiao, J. Am. Chem. Soc., 2015, 137, 4223-4229.          

29. Y. Jin, H. Wang, J. Li, X. Yue, Y. Han, P. K. Shen and Y. Cui, Adv. Mater., 2016, 28, 3785-3790.

30. Q. Kang, L. Vernisse, R. C. Remsing, A. C. Thenuwara, S. L. Shumlas, I. G. McKendry, M. L. Klein, E. Borguet, M. J. Zdilla and D. R. Strongin, J. Am. Chem. Soc., 2017, 139, 1863-1870.  

31. Y. Lee, J. Suntivich, K. J. May, E. E. Perry and Y. Shao-Horn, J. Phys. Chem. Lett., 2012, 3, 399-404.

32. F. Lyu, Y. Bai, Z. Li, W. Xu, Q. Wang, J. Mao, L. Wang, X. Zhang and Y. Yin, Adv. Funct. Mater., 2017, 27, 1702324. 

33. G. Wu, W. Chen, X. Zheng, D. He, Y. Luo, X. Wang, J. Yang, Y. Wu, W. Yan, Z. Zhuang, X. Hong and Y. Li, Nano Energy, 2017, 38, 167-174.  

34. M. Gao, W. Sheng, Z. Zhuang, Q. Fang, S. Gu, J. Jiang and Y. Yan, J. Am. Chem. Soc., 2014, 136, 7077-7084. 

35. W. Liu, H. Liu, L. Dang, H. Zhang, X. Wu, B. Yang, Z. Li, X. Zhang, L. Lei and S. Jin, Adv. Funct. Mater., 2017, 27, 1603904.  

36. X. Long, J. Li, S. Xiao, K. Yan, Z. Wang, H. Chen and S. Yang, Angew. Chem. Int. Ed., 2014, 53, 7584-7588.  

37. M. Tahir, N. Mahmood, L. Pan, Z. F. Huang, Z. Lv, J. Zhang, F. K. Butt, G. Shen, X. Zhang, S. X. Dou and J. J. Zou, J. Mater. Chem. A, 2016, 4, 12940-12946.  

38Y. Li, L. Zhang, X. Xiang, D. Yan and F. Li, J. Mater. Chem. A, 2014, 2, 13250-13258.

39.  J. I. Jung, H. Y. Jeong, J. S. Lee, M. G. Kim and J. Cho, Angew. Chem. Int. Ed., 2014, 53, 4582-4586.

40.  Y. Fan, S. Ida, A. Staykov, T. Akbay, H. Hagiwara, J. Matsuda, K. Kaneko and T. Ishihara, Small, 2017, 13, 1700099.  

41. K. Xu, P. Chen, X. Li, Y. Tong, H. Ding, X. Wu, W. Chu, Z. Peng, C. Wu and Y. Xie, J. Am. Chem. Soc., 2015, 137, 4119-4125. 

42. F. Yu, H. Zhou, Z. Zhu, J. Sun, R. He, J. Bao, S. Chen and Z. Ren, ACS Catal., 2017, 7, 2052-2057.

43. Y. Zhang, B. Ouyang, J. Xu, G. Jia, S. Chen, R. S. Rawat and H. J. Fan, Angew. Chem. Int. Ed., 2016, 55, 8670-8674.  

44. Y. Liu, H. Cheng, M. Lyu, S. Fan, Q. Liu, W. Zhang, Y. Zhi, C. Wang, C. Xiao, S. Wei, B. Ye and Y. Xie, J. Am. Chem. Soc., 2014, 136, 15670-15675.  

45. T. Yoon and K. S. Kim, Adv. Funct. Mater., 2016, 26, 7386-7393. 

46. J. Wan, W. Ye, R. Gao, X. Fang, Z. Guo, Y. Lu and D. Yan, Inorg. Chem. Front., 2019, 6, 1793-1798. 

47. X. Fan, Y. Liu, S. Chen, J. Shi, J. Wang, A. Fan, W. Zan, S. Li, W. A. Goddard and X. M. Zhang, Nat. Commun., 2018, 9, 1809.  

48. P. W. Menezes, C. Panda, S. Loos, F. Bunschei-Bruns, C. Walter, M. Schwarze, X. Deng, H. Dau and M. Driess, Energy Environ. Sci., 2018, 11, 1287-1298.

49. F. Yu, H. Zhou, Y. Huang, J. Sun, F. Qin, J. Bao, W. A. Goddard, S. Chen and Z. Ren, Nat. Commun., 2018, 9, 2551.

50. G. Fan, F. Li, D. G. Evans and X. Duan, Chem. Soc. Rev., 2014, 43, 7040-7066. 

51. Q. Wang and D. O'Hare, Chem. Rev., 2012, 112, 4124-4155.  

52. S. M. Xu, T. Pan, Y. B. Dou, H. Yan, S. T. Zhang, F. Y. Ning, W. Y. Shi and M. Wei, J. Phys. Chem. C, 2015, 119, 18823-18834.  

53. A. Sengeni, K. K and S. Kundu, Mater. Today Energy, 2017, 6, 1-26.  

54. Y. Wang, D. Yan, S. El Hankari, Y. Zou and S. Wang, Adv. Sci., 2018, 5, 1800064. 

55. R. Gao and D. Yan, Adv. Energy Mater., 2019, 0, 1900954. 

56. A. Alarawi, V. Ramalingam and J. H. He, Mater. Today Energy, 2019, 11, 1-23.

57. Y. Li, H. Wang, L. Xie, Y. Liang, G. Hong and H. Dai, J. Am. Chem. Soc., 2011, 133, 7296-7299.           

58. I. Dincer and C. Acar, Int. J. Hydrogen Energy, 2015, 40, 11094-11111.  

59. A. Li, H. Ooka, N. Bonnet, T. Hayashi, Y. Sun, Q. Jiang, C. Li, H. Han and R. Nakamura, Angew. Chem., 2019, 131, 5108-5112.

60. Z. Qiu, C. W. Tai, G. A. Niklasson and T. Edvinsson, Energy Environ. Sci., 2019, 12, 572-581. 

61. Z. Cai, D. Zhou, M. Wang, S. M. Bak, Y. Wu, Z. Wu, Y. Tian, X. Xiong, Y. Li, W. Liu, S. Siahrostami, Y. Kuang, X. Q. Yang, H. Duan, Z. Feng, H. Wang and X. Sun, Angew. Chem., 2018, 130, 9536-9540. 

62. C. Wei, Z. Feng, M. Baisariyev, L. Yu, L. Zeng, T. Wu, H. Zhao, Y. Huang, M. J. Bedzyk, T. Sritharan and Z. J. Xu, Chem. Mater., 2016, 28, 4129-4133.

63.  J. Yang, Z. Zeng, J. Kang, S. Betzler, C. Czarnik, X. Zhang, C. Ophus, C. Yu, K. Bustillo, M. Pan, J. Qiu, L. W. Wang and H. Zheng, Nat. Mater., 2019, DOI: 10.1038/s41563-019-0415-3.  

64.  F. Dionigi and P. Strasser, Adv. Energy Mater., 2016, 6, 1600621.

65.  O. Diaz-Morales, I. Ledezma-Yanez, M. T. M. Koper and F. Calle-Vallejo, ACS Catal., 2015, 5, 5380-5387.   

66. M. A. Oliver-Tolentino, J. Vázquez-Samperio, A. Manzo-Robledo, R. d. G. González-Huerta, J. L. Flores-Moreno, D. Ramírez-Rosales and A. Guzmán-Vargas, J. Phys. Chem. C, 2014, 118, 22432-22438.     

67.  R. Gao and D. Yan, Nano Research, 2018, 11, 1883-1894. 

68.  Z. Lu, W. Xu, W. Zhu, Q. Yang, X. Lei, J. Liu, Y. Li, X. Sun and X. Duan, Chem. Commun., 2014, 50, 6479-6482. 

69.  Y. Tang, R. Wang, Y. Yang, D. Yan and X. Xiang, ACS Appl. Mater. Interfaces, 2016, 8, 19446-19455.

70. H. Chen, L. Hu, M. Chen, Y. Yan and L. Wu, Adv. Funct. Mater., 2014, 24, 934-942. 

71. R. Liu, Y. Wang, D. Liu, Y. Zou and S. Wang, Adv. Mater., 2017, 29, 1701546.  

72. W. Yang, Z. Gao, J. Wang, J. Ma, M. Zhang and L. Liu, ACS Appl. Mater. Interfaces, 2013, 5, 5443-5454.     

73.  M. Gong, Y. Li, H. Wang, Y. Liang, J. Z. Wu, J. Zhou, J. Wang, T. Regier, F. Wei and H. Dai, J. Am. Chem. Soc., 2013, 135, 8452-8455.   

74.  Y. Tang, X. Fang, X. Zhang, G. Fernandes, Y. Yan, D. Yan, X. Xiang and J. He, ACS Appl. Mater. Interfaces, 2017, 9, 36762-36771. 

75.  Z. Cai, X. Bu, P. Wang, J. C. Ho, J. Yang and X. Wang, J. Mater. Chem. A, 2019, 7, 5069-5089.  

76.  Z. Guo, W. Ye, X. Fang, J. Wan, Y. Ye, Y. Dong, D. Cao and D. Yan, Inorg. Chem. Front., 2019, 6, 687-693.   

77.  J. Ji, L. L. Zhang, H. Ji, Y. Li, X. Zhao, X. Bai, X. Fan, F. Zhang and R. S. Ruoff, ACS Nano, 2013, 7, 6237-6243.  

78. F. Song and X. Hu, Nat. Commun., 2014, 5, 4477.   

79. Z. Chen, Y. Ha, H. Jia, X. Yan, M. Chen, M. Liu and R. Wu, Adv. Energy Mater., 2019, 0, 1803918.  

80. L. Yu, H. Zhou, J. Sun, F. Qin, F. Yu, J. Bao, Y. Yu, S. Chen and Z. Ren, Energy Environ. Sci., 2017, 10, 1820-1827.

81. P. F. Liu, S. Yang, B. Zhang and H. G. Yang, ACS Appl. Mater. Interfaces, 2016, 8, 34474-34481.

82. E. Hu, Y. Feng, J. Nai, D. Zhao, Y. Hu and X. W. Lou, Energy Environ. Sci., 2018, 11, 872-880.

83. H. Zhang, X. Li, A. Hähnel, V. Naumann, C. Lin, S. Azimi, S. L. Schweizer, A. W. Maijenburg and R. B. Wehrspohn, Adv. Funct. Mater., 2018, 28, 1706847.

84. Y. Guo, J. Tang, Z. Wang, Y. M. Kang, Y. Bando and Y. Yamauchi, Nano Energy, 2018, 47, 494-502.

85. S. Li, S. Sirisomboonchai, A. Yoshida, X. An, X. Hao, A. Abudula and G. Guan, J. Mater. Chem. A, 2018, 6, 19221-19230.   

86.  L. Yu, H. Zhou, J. Sun, F. Qin, D. Luo, L. Xie, F. Yu, J. Bao, Y. Li, Y. Yu, S. Chen and Z. Ren, Nano Energy, 2017, 41, 327-336.  

87.  J. Liu, J. Wang, B. Zhang, Y. Ruan, L. Lv, X. Ji, K. Xu, L. Miao and J. Jiang, ACS Appl. Mater. Interfaces, 2017, 9, 15364-15372. 

88.  X. Gao, H. Zhang, Q. Li, X. Yu, Z. Hong, X. Zhang, C. Liang and Z. Lin, Angew. Chem. Int. Ed., 2016, 55, 6290-6294.

89.  H. Shi, H. Liang, F. Ming and Z. Wang, Angew. Chem. Int. Ed., 2017, 56, 573-577.

90.  Y. Zhang, Q. Shao, Y. Pi, J. Guo and X. Huang, Small, 2017, 13, 1700355.

91.  Z. Wang, S. Zeng, W. Liu, X. Wang, Q. Li, Z. Zhao and F. Geng, ACS Appl. Mater. Interfaces, 2017, 9, 1488-1495.

92. Y. Hou, M. R. Lohe, J. Zhang, S. Liu, X. Zhuang and X. Feng, Energy Environ. Sci., 2016, 9, 478-483.

93. Y. Jia, L. Zhang, G. Gao, H. Chen, B. Wang, J. Zhou, M. T. Soo, M. Hong, X. Yan, G. Qian, J. Zou, A. Du and X. Yao, Adv. Mater., 2017, 29, 1700017.

94. C. Huang, T. Ouyang, Y. Zou, N. Li and Z. Q. Liu, J. Mater. Chem. A, 2018, 6, 7420-7427.

95. B. Liu, Y. Wang, H. Q. Peng, R. Yang, Z. Jiang, X. Zhou, C. S. Lee, H. Zhao and W. Zhang, Adv. Mater., 2018, 30, 1803144.

96. W. Cheng, H. Zhang, X. Zhao, H. Su, F. Tang, J. Tian and Q. Liu, J. Mater. Chem. A, 2018, 6, 9420-9427.

97. H. Liang, A. N. Gandi, D. H. Anjum, X. Wang, U. Schwingenschlögl and H. N. Alshareef, Nano Lett., 2016, 16, 7718-7725.

98. Q. Liu, J. Huang, Y. Zhao, L. Cao, K. Li, N. Zhang, D. Yang, L. Feng and L. Feng, Nanoscale,2019,18,8855-8853

99. T. Chen, R. Zhang, G. Chen, J. Huang, W. Chen, X. Wang, D. Chen, C. Li and K. Ostrikov, Catal. Today, 2019 

100. L. Hui, Y. Xue, B. Huang, H. Yu, C. Zhang, D. Zhang, D. Jia, Y. Zhao, Y. Li, H. Liu and Y. Li, Nat. Commun., 2018, 9, 5309.

101. X. Li, X. Hao, Z. Wang, A. Abudula and G. Guan, J. Power Sources, 2017, 347, 193-200.

102. B. Weng, F. Xu, C. Wang, W. Meng, C. R. Grice and Y. Yan, Energy Environ. Sci., 2017, 10, 121-128.

103. R. Yang, Y. Zhou, Y. Xing, D. Li, D. Jiang, M. Chen, W. Shi and S. Yuan, Appl. Catal., B, 2019, 253, 131-139.

104. Q. Q. Chen, C. C. Hou, C. J. Wang, X. Yang, R. Shi and Y. Chen, Chem. Commun., 2018, 54, 6400-6403.

105. C. Guan, W. Xiao, H. Wu, X. Liu, W. Zang, H. Zhang, J. Ding, Y. P. Feng, S. J. Pennycook and J. Wang, Nano Energy, 2018, 48, 73-80.

106. Z. Li, W. Niu, L. Zhou and Y. Yang, ACS Energy Letters, 2018, 3, 892-898.

107. H. Li, S. Chen, Y. Zhang, Q. Zhang, X. Jia, Q. Zhang, L. Gu, X. Sun, L. Song and X. Wang, Nat. Commun., 2018, 9, 2452.

108. C. Hu, L. Zhang, Z. J. Zhao, A. Li, X. Chang and J. Gong, Adv. Mater., 2018, 30, 1705538.

109. Z. Kang, H. Guo, J. Wu, X. Sun, Z. Zhang, Q. Liao, S. Zhang, H. Si, P. Wu, L. Wang and Y. Zhang, Adv. Funct. Mater., 2019, 29, 1807031.

110. J. Liu, Y. Zheng, Y. Jiao, Z. Wang, Z. Lu, A. Vasileff and S. Z. Qiao, Small, 2018, 14, 1704073.

111.  L. Zhou, M. Shao, J. Li, S. Jiang, M. Wei and X. Duan, Nano Energy, 2017, 41, 583-590.

112.  D. Senthil Raja, H. W. Lin and S. Y. Lu, Nano Energy, 2019, 57, 1-13.

113.  W. Ye, X. Fang, X. Chen and D. Yan, Nanoscale, 2018, 10, 19484-19491.

114.  Y. Wang, Y. Zhang, Z. Liu, C. Xie, S. Feng, D. Liu, M. Shao and S. Wang, Angew. Chem. Int. Ed., 2017, 56, 5867-5871.

115.  G. F. Chen, T. Y. Ma, Z. Q. Liu, N. Li, Y. Z. Su, K. Davey and S. Z. Qiao, Adv. Funct. Mater., 2016, 26, 3314-3323.

116.  Z. Lu, W. Zhu, X. Yu, H. Zhang, Y. Li, X. Sun, X. Wang, H. Wang, J. Wang, J. Luo, X. Lei and L. Jiang, Adv. Mater., 2014, 26, 2683-2687.

117.  H. Sun, Z. Yan, F. Liu, W. Xu, F. Cheng and J. Chen, Adv. Mater., 2019, 0, 1806326.

118.  Z. Yan, H. Sun, X. Chen, H. Liu, Y. Zhao, H. Li, W. Xie, F. Cheng and J. Chen, Nat. Commun., 2018, 9, 2373.

119.  M. Asnavandi, Y. Yin, Y. Li, C. Sun and C. Zhao, ACS Energy Letters, 2018, 3, 1515-1520.

120.  Y. Wang, M. Qiao, Y. Li and S. Wang, Small, 2018, 14, 1800136.

121.  M. Q. Yang, J. Wang, H. Wu and G. W. Ho, Small, 2018, 14, 1703323.

122. Y. Zhao, X. Zhang, X. Jia, G. I. N. Waterhouse, R. Shi, X. Zhang, F. Zhan, Y. Tao, L. Z. Wu, C. H. Tung, D. O'Hare and T. Zhang, Adv. Energy Mater., 2018, 8, 1703585.

123. J. Ping, Y. Wang, Q. Lu, B. Chen, J. Chen, Y. Huang, Q. Ma, C. Tan, J. Yang, X. Cao, Z. Wang, J. Wu, Y. Ying and H. Zhang, Adv. Mater., 2016, 28, 7640-7645.

124. H. Liu, Y. Wang, X. Lu, Y. Hu, G. Zhu, R. Chen, L. Ma, H. Zhu, Z. Tie, J. Liu and Z. Jin, Nano Energy, 2017, 35, 350-357.

125. X. Long, S. Xiao, Z. Wang, X. Zheng and S. Yang, Chem. Commun., 2015, 51, 1120-1123.

126. A. L. Wang, H. Xu and G. R. Li, ACS Energy Letters, 2016, 1, 445-453.

127. D. Friebel, M. W. Louie, M. Bajdich, K. E. Sanwald, Y. Cai, A. M. Wise, M. J. Cheng, D. Sokaras, T. C. Weng and R. Alonso-Mori, J. Am. Chem. Soc., 2015, 137, 1305-1313.

128. M. Gong, W. Zhou, M. J. Kenney, R. Kapusta, S. Cowley, Y. Wu, B. Lu, M. C. Lin, D. Y. Wang, J. Yang, B. J. Hwang and H. Dai, Angew. Chem. Int. Ed., 2015, 54, 11989-11993.

129. X. Li, L. Zhang, M. Huang, S. Wang, X. Li and H. Zhu, J. Mater. Chem. A, 2016, 4, 14789-14795.

130. J. Luo, D. A. Vermaas, D. Bi, A. Hagfeldt, W. A. Smith and M. Grätzel, Adv. Energy Mater., 2016, 6, 1600100.

131. C. G. Morales-Guio, M. T. Mayer, A. Yella, S. D. Tilley, M. Grätzel and X. Hu, J. Am. Chem. Soc., 2015, 137, 9927-9936.

132. W. He, R. Wang, L. Zhang, J. Zhu, X. Xiang and F. Li, J. Mater. Chem. A, 2015, 3, 17977-17982.

133. M. Arif, G. Yasin, M. Shakeel, X. Fang, R. Gao, S. Ji and D. Yan, Chem. Asian J., 2018, 13, 1045-1052.

134.  M. Arif, G. Yasin, M. Shakeel, M. A. Mushtaq, W. Ye, X. Fang, S. Ji and D. Yan, Mater. Chem. Front., 2019, 3, 520-531.gen Evolution Reaction; Oxygen Evolution Reaction; Overall Water Splitting