Enhancing the efficiency and stability of electrocatalysts for water splitting: NiCo-LDH nanosheet arrays at high current density

Xiaomei Wang a, Tiantian Wang a, Yongren Yu a, Junhua You a, Fang Hu *a and Depeng Zhao *b
aSchool of Materials Science and Engineering, Shenyang University of Technology, Shenyang 110870, P. R. China. E-mail: hufang25@126.com; youjunhua168@163.com
bSchool of Renewable Energy, Shenyang Institute of Engineering, Shenyang, 110136, P. R. China. E-mail: Hellodepeng@163.com

Received 5th June 2024 , Accepted 6th August 2024

First published on 2nd September 2024


Abstract

The development of low-cost, efficient, and stable bifunctional electrocatalysts is very important for the development of renewable energy. Layered double hydroxides (LDHs) are considered to be promising electrocatalysts due to their flexible ion exchange, abundant structural adjustability, excellent thermal stability, and easy functionalization with other materials. In this work, nickel-cobalt layered double hydroxide (NiCo-LDH) nanosheet arrays are prepared by a simple hydrothermal method. Four samples Ni1Co4-LDH, Ni2Co3-LDH, Ni3Co2-LDH and Ni4Co1-LDH are obtained by adjusting the Ni/Co ratio (5 − x[thin space (1/6-em)]:[thin space (1/6-em)]x). Among them, the Ni2Co3-LDH samples show excellent HER and OER properties. At a current density of 50 mA cm−2 in a 1.0 M KOH electrolyte, the overpotentials of the HER and OER are 227 mV and 303 mV, respectively. In addition, as a bifunctional electrocatalyst for water splitting, the sample provides a voltage of 1.83 V and long-term stability at 50 mA cm−2.


1. Introduction

With the rapid development of today's social economy, energy consumption is intensifying, and environmental pollution is becoming more and more serious, which makes the development of clean and sustainable energy to replace traditional fossil energy a top priority.1–3 Among the many renewable energy sources, hydrogen energy is considered to be the most promising clean energy due to its high calorific value, pollution-free products, and abundant raw material storage.4–6 In recent years, hydrogen production by water electrolysis has become the focus of researchers, because of its advantages of environmental friendliness, convenient preparation processes, and high product purity.7,8 However, its high cost and low energy conversion efficiency have been restricting its commercial development.9,10 Water electrolysis for hydrogen production includes the hydrogen evolution reaction (HER) and oxygen evolution reaction (OER).11–13 In general, proton-coupled electron transfer reactions are slower, and strong O–O double bonds on the anode limit the efficiency of the oxygen evolution reaction (OER).14–16 In order to improve the conversion efficiency, precious metal catalysts such as Pt-based catalysts, RuO2 and IrO2 are widely used. However, due to their scarcity, high cost and weak stability, they have set up obstacles for their large-scale application.17 Therefore, the current research hotspot is the development of low-cost, efficient and stable electrocatalysts to accelerate the chemical reaction kinetic process.

According to previous studies, the modification of the catalyst surface, reasonable structural design (such as layered structures and micro-nano arrays) and material composites can achieve the purpose of increasing the number of catalytic activity reaction sites, enhancing the reaction activity and improving the catalytic performance.18–20 Nickel foam (NF) is often used as a substrate due to its high porosity and good electrical conductivity, and combined with the controlled growth technology of two-dimensional layered double hydroxide (LDH), it is possible to prepare high-efficiency and high-stability non-precious metal catalysts.21,22 Non-noble metal electrocatalysts, represented by transition metal LDH, have been widely studied because of their unique layered structure and relatively open ion diffusion rates.23,24 Compared with a single transition metal, bimetallic hydroxides can expose more reactive sites to promote the occurrence of reactions and accelerate the diffusion of ions.25,26 Therefore, bimetallic hydroxides have better catalytic properties. Among many metallic elements, nickel and cobalt are transition metal elements, and Ni–Co-based catalysts have become the focus of research due to their low price, abundant reserves, and excellent hydrogen and oxygen evolution performance in alkaline environments. In recent years, Bao et al. had prepared CoMo-LDH ultra-thin nanosheets by the co-precipitation method, and the electronic structure of LDH had been changed due to the doping of Mo6+, which had promoted the occurrence of catalytic reactions.27 At a current density of 10 mA cm−2, the OER and HER overpotentials were 290 mV and 115 mV, respectively. Song et al. prepared CoMn-LDH ultrathin nanoplates, and the reaction rate was more than 20 times that of Co(OH)2.28 All of them exhibited better catalytic activity than monometallic hydroxide Co(OH)2. Ye et al. used a simple one-step hydrothermal method to prepare a novel two-dimensional NiCr-LDH nanosheet array with a three-dimensional porous structure, which has excellent durability and exhibits efficient and stable electrocatalytic activity in alkaline media.29 To date, many studies have promoted the understanding and progress of TM LDH nanosheets as electrocatalysts. However, there are still many challenges in optimizing the electrocatalytic performance of TM LDHs, which limits their further commercial electrocatalytic applications.30,31

Herein, we grow NiCo-LDH nanosheet arrays in situ on nickel foam (NF) by a simple hydrothermal method, and the samples prepared by adjusting the ratio of Ni to Co (Ni[thin space (1/6-em)]:[thin space (1/6-em)]Co = 5 − x[thin space (1/6-em)]:[thin space (1/6-em)]x) are denoted as Ni1Co4-LDH, Ni2Co3-LDH, Ni3Co2-LDH and Ni4Co1-LDH, respectively. Through structural characterization and electrochemical performance testing, it can be found that Ni2Co3-LDH/NF exhibits excellent electrocatalytic performance. The HER overpotentials at 50 mA cm−2 are 227 mV and 303 mV in a 1.0 M KOH electrolyte and excellent stability after 12 h of cycling is maintained.

2. Experiment

Before the experiment, a piece of nickel foam was washed with absolute ethanol and deionized water. NiCo-LDH was synthesized on nickel foam (NF) by a simple hydrothermal method. Firstly, NiCl2·6H2O and Co(NO3)2·6H2O were dissolved in a mixture of 60 mL of methanol and deionized water (1[thin space (1/6-em)]:[thin space (1/6-em)]1) according to the molar ratio (5 − x[thin space (1/6-em)]:[thin space (1/6-em)]x) and stirred for 30 min. Then, 3 mM urea was added to the above solution during constant stirring. Finally, the above mixture and pretreated nickel foam (4 × 4 cm) were transferred to a 100 mL reactor and kept at 140 °C for 6 hours. After natural cooling to room temperature, the prepared samples were washed with absolute ethanol and deionized water and dried overnight at 60 °C to prepare NiCo-LDH samples with different Ni/Co ratios. All reagents are of analytical-grade and require no further purification.

Structure characterization

X-ray diffraction (XRD, Shimadzu-7000, Cu Kα), X-ray photoelectron spectroscopy (XPS ESCALAB 250 with Al Kα radiation), scanning electron microscopy (SEM, Gemini 300-71-31), and transmission electron microscopy (TEM, JEM-2100 PLUS) were used to characterize the morphology and crystal structure of the samples in detail.

Electrochemical performance testing

An electrochemical workstation was used to measure the electrochemical performance of a sample. The electrolyte was 1.0 M KOH in aqueous solution, using a standard three-electrode system, the counter electrode was 1 cm2 platinum sheet, the reference electrode was Ag/AgCl, and the working electrode was NiCo-LDH/NF; linear scanning voltammetry (LSV), electrochemical impedance (EIS), cyclic voltammetry (CV), and chronopotentiometry (CP) were performed. According to the Nernst equation ERHE = EAg/AgCl + 0.197 V + 0.059 V × pH, all potentials can be converted into reversible hydrogen electrodes (RHEs). The overpotential is η = ERHE − 1.23 V.

3. Results and discussion

First, the prepared samples are characterized by XRD. As shown in Fig. 1a, the angles of the strong diffraction peaks at 2θ are 11.718, 23.242, 34.108, and 38.309 degrees, which belong to the (003), (006), (012), and (015) crystal planes of the NiCo-LDH phase (JCPDS number: 40-0216). The remaining peaks are 44.622 and 60.054 degrees, belonging to the (260) and (444) planes (JCPDS number: 47-0938). This result strongly proves that NiCo-LDH can be successfully prepared on NF. Fourier transform infrared spectroscopy (FTIR) was conducted to further investigate the structural characteristics of the samples (Fig. 1b); 3518 cm−2 and 1600 cm−2 correspond to O–H bonds, 1000–800 cm−2 correspond to M (metal)–O bonds,32 and 2912 cm−2 and 2198 cm−2 correspond to C–H bonds. In addition to this, 1367 cm−2 and 1605 cm−2 correspond to the C[double bond, length as m-dash]C bond.33 This further indicates the successful preparation of NixCoy-LDH.
image file: d4ce00564c-f1.tif
Fig. 1 (a) XRD spectra of NixCoy-LDH; (b) FTIR spectra of NixCoy-LDH; (c) XPS spectra of NixCoy-LDH and (d) Ni 2p; (e) Co 2p; (f) O 1s.

Next, the elemental composition and chemical state of the Ni2Co3-LDH product are investigated by XPS. The full XPS spectrum of NixCoy-LDH/NF is shown in Fig. 1c, indicating the presence of elements such as Ni, Co, C, and O in the prepared sample. As shown in Fig. 1d, the binding energies of Ni 2p3/2 and Ni 2p1/2 for Ni2+ correspond to the peaks at 854.00 eV and 871.31 eV, and those of Ni 2p3/2 and Ni 2p1/2 for Ni3+ correspond to the peaks at 856.09 eV and 873.78 eV, indicating the coexistence of Ni2+ and Ni3+ ions in the prepared product.34 In addition, the binding energies at the satellite peaks are 860.71 eV and 878.26 eV. In the Ni 2p spectrum, the binding energy of Ni3+ is higher than that of Ni2+, indicating that Ni3+ is the dominant valence state. The Co 2p spectrum (Fig. 1e) can be fitted to Co3+ and Co2+; the binding energies of 780.00 eV and 796.49 eV belong to Co3+, and the binding energies of 782.66 eV and 800.34 eV belong to Co2+, and the two satellite peaks of Co are located at 786.54 eV and 803.05 eV. These results confirm the coexistence of Co3+ and Co2+ and the high atomic ratio of Co3+.35 The O 1s spectrum consists of two characteristic peaks, 529.7 eV and 531.09 eV, which are attributed to metal–oxygen bonds (M–O) and oxygen vacancies (VO), respectively, as shown in Fig. 1f.36

Scanning electron microscopy (SEM) is used to characterize the topography of the sample, as shown in Fig. 2(a) and (b). As you can see, the nanosheets are evenly staggered on the surface of the NF, with an average thickness of 90 nm. This array and the resulting porous nanostructures facilitate the diffusion of ions and the transport of charges. The elemental energy spectrum (EDS) of the sample (Fig. 2c) shows a uniform distribution of the Ni, Co, O, and C elements. Fig. 2(d) and (e) show local TEM images of the nanoflower NixCoy-LDH formed by the combination of many ultrathin nanosheet arrays, further demonstrating the lamellar formation of the nanosheet array. Fig. 2f shows the corresponding HRTEM image with a plane spacing of 0.256 nm that matches well with the (012) plane.


image file: d4ce00564c-f2.tif
Fig. 2 Morphology and structure characterization of Ni2Co3-LDH samples. (a) Low-magnification SEM images; (b) high-magnification SEM images; (c) EDS element mapping; (d) and (e) TEM images; (f) HRTEM image.

The OER performance of the samples is investigated in a 1.0 M KOH electrolyte.

Fig. 3a shows the LSV curves for all catalysts at 2 mV s−1. At 50 mA cm−2, the overpotential of the Ni2Co3-LDH sample is 303 mV, less than that of Ni1Co4-LDH (339 mV), Ni3Co2-LDH (335 mV), Ni4Co1-LDH (381 mV) and IrO2 (313 mV). The results show that the activation energy required for the Ni2Co3-LDH reaction is lower and the kinetic rate of the reaction is faster. The Tafel plot (Fig. 3b) is used to reflect the OER kinetics of the prepared samples, and the Tafel slope of Ni2Co3-LDH is 73.62 mV dec−1, which is significantly lower than those of Ni1Co4-LDH (90.24 mV dec−1), Ni3Co2-LDH (100.92 mV dec−1), Ni4Co1-LDH (135.93 mV dec−1) and IrO2 (88.67 mV dec−1). The results show that the sample Ni2Co3-LDH has a higher electron and ion transfer velocity. Electrochemical impedance spectroscopy (EIS) is performed in the frequency range of 100 kHz to 0.1 Hz to further investigate the NixCoy-LDH electrode reaction kinetics (Fig. 3c). In Fig. 3d, the slope value of the Ni2Co3-LDH electrode is 1.4, which is higher than those of Ni1Co4-LDH (1.06), Ni3Co2-LDH (1.08), and Ni4Co1-LDH (1.13), indicating that the Ni2Co3-LDH sample can provide a faster channel for ions than the other three samples. Ni2Co3-LDH exhibits smaller equivalent resistance than Ni1Co4-LDH, Ni3Co2-LDH and Ni4Co1-LDH, indicating that the Ni2Co3-LDH electrode has excellent conductivity. The electrochemically active surface area (ECSA) is an effective parameter for further investigating catalyst catalytic activity, which is calculated in terms of double-layer capacitance (Cdl), as shown in Fig. 3e. The Cdl value of Ni2Co3-LDH is 0.12 mF cm−2, higher than those of Ni1Co4-LDH (0.08 mF cm−2), Ni3Co2-LDH (0.10 mF cm−2) and Ni4Co1-LDH (0.11 mF cm−2). A higher Cdl value represents a larger electrochemically active surface area. Stability is another important criterion that affects the performance of catalysts. Therefore, Fig. 3f shows the long-term OER cycling stability of the four samples prepared. Among them, the Ni2Co3-LDH catalyst can maintain good stability for up to 12 hours.


image file: d4ce00564c-f3.tif
Fig. 3 OER performance. (a) Polarization curve at a scan rate of 2 mV s−1; (b) Tafel diagram; (c) Nyquist plots; (d) impedance fitting of four samples; (e) linear fit of electric double-layer capacitance (Cdl) measurements; (f) constant voltage cycle diagram.

The HER performance of the samples is investigated using LSV curves, as shown in Fig. 4a. At −50 mA cm−2, the overpotential (227.00 mV) of the Ni2Co3-LDH sample is lower than those of the Ni1Co4-LDH sample (278.7 mV), the Ni3Co2-LDH sample (251.59 mV), and the Ni4Co1-LDH sample (265.7 mV), respectively. The overpotential of commercial Pt/C is 85.7 mV. The corresponding Tafel curves (Fig. 4b) show that the Tafel slope of Ni2Co3-LDH is 99.46 mV dec−1, lower than those of Ni1Co4-LDH (207.56 mV dec−1), Ni3Co2-LDH (126.36 mV dec−1), and Ni4Co1-LDH (205.59 mV dec−1). Commercial Pt/C exhibits the smallest Tafel slope (63.44 mV dec−1). The results show that the sample Ni2Co3-LDH has a rapid HER reaction kinetics.


image file: d4ce00564c-f4.tif
Fig. 4 HER performance. (a) LSV curve at 5 mV s−1; (b) Tafel diagram; (c)Nyquist diagram; (d) impedance fitting curve; (e) electric double-layer capacitance (Cdl); (f) 12 hour constant voltage stability.

The charge transfer kinetics is then investigated by electron diffraction spectroscopy (Fig. 4c), and it can be seen that the semicircular diameter of the Ni2Co3-LDH sample is smaller than those of the other three samples, indicating that the Ni2Co3-LDH catalyst has excellent conductivity. The slope value of the Ni2Co3-LDH electrode is 94.49, which is higher than those of Ni1Co4-LDH (49.73), Ni3Co2-LDH (53.18) and Ni4Co1-LDH (10.02), showing its high ion diffusion ability (Fig. 4d). Fig. 4e shows the electrochemically active surface area (ECSA) of the sample. The Cdl value of Ni2Co3-LDH is 0.012 mF cm−2, which is higher than those of Ni1Co4-LDH (0.007 mF cm−2), Ni3Co2-LDH (0.008 mF cm−2) and Ni4Co1-LDH (0.008 mF cm−2). Fig. 4f shows the cycling stability of the prepared sample catalyst. Among them, the Ni2Co3-LDH catalyst has better durability.

We then perform a structural analysis on the Ni2Co3-LDH samples that have undergone cyclic testing. First, XPS is used to analyze the surface valence state of the sample. As shown in Fig. 5a, the peaks at 855.68 eV and 874.68 eV in the Ni 2p spectrum are attributed to Ni3+, with Ni2+ corresponding to the peaks at 855.65 eV and 871.98 eV binding energies.37 The binding energies of the corresponding peaks of Co3+ in the Co 2p spectrum (Fig. 5b) are 779.58 eV and 794.8 eV, and the corresponding peaks of Co2+ are 782.9 eV and 796.8 eV.38 The O 1s spectrum (Fig. 5c) contains two peaks, a metal–oxygen bond (529.09 eV) and an oxygen ion (530.8 eV).39 The above results show that the position and relative intensity of the XPS peak remain basically unchanged, indicating the chemical stability of the Ni, Co and O components. The XPS data confirms that the three metal elements are still in the trimetal LDH system after cyclic testing, and the composition of Ni2Co3-LDH is stable. The morphology and structure of the cycled Ni2Co3-LDH samples are studied. From the SEM images (Fig. 5d and e), the layered morphology of the nanosheet array can still be maintained in the samples after cycling, indicating that the above catalysts have excellent structural stability. As can be seen in Fig. 5f, the lattice spacing of the HRTEM image is 0.256 nm, which is well matched to the (012) plane. It shows that the structure of the sample does not change significantly after cycling.


image file: d4ce00564c-f5.tif
Fig. 5 Morphology and structure characterization of the Ni2Co3-LDH catalyst after OER cycling. (a) Ni 2p XPS; (b) Co 2p XPS; (c) O 1s XPS; (d) low-magnification SEM image; (e) high-magnification SEM image; (f) HRTEM image.

Finally, the overall hydrolysis performance of NiCo-LDH is studied. Fig. 6a shows the LSV curve of NixCoy-LDH at a scan rate of 5 mV s−1. Compared to Ni1Co4-LDH (1.85 V), the application voltage of the Ni2Co3-LDH sample (1.83 V) is lower at 50 mA cm−2. As the electrolytic reaction progresses, the cathode (H2) and anode (O2) produce many bubbles. The EIS plot (Fig. 6b) shows that the Ni2Co3-LDH sample exhibits a smaller semicircular diameter in the high-frequency region and a larger slope in the low-frequency region compared to the Ni1Co4-LDH sample. This phenomenon indicates that the Ni2Co3-LDH sample has high electrical conductivity. In the total solution water system, the Cdl value of Ni2Co3-LDH (0.104 mF cm−2) is higher than that of Ni1Co4-LDH (0.090 mF cm−2), suggesting that more exposed active sites existed in the sample Ni2Co3-LDH and it exhibited superior catalytic activity as shown in Fig. 6c. To test the cycling stability of the samples, a chronopotentiometric method (CP) is employed, as shown in Fig. 6d. The results show that the current density of the Ni2Co3-LDH catalyst is more stable than that of the Ni1Co4-LDH catalyst after cycling up to 12 h, indicating that the Ni2Co3-LDH catalyst has excellent water splitting performance.


image file: d4ce00564c-f6.tif
Fig. 6 (a) Polarization curve of the catalyst with a scan rate of 5 mV s−1; (b) Nyquist curve of overall hydrolysis; (c) double layer capacitance for fully soluble water; (d) cycling data for overall water splitting.

4. Conclusion

In summary, we synthesize a novel two-dimensional NixCoy-LDH nanosheet array with a three-dimensional porous structure on nickel foam (NF) by a simple hydrothermal method. The Ni2Co3-LDH electrocatalyst exhibits excellent HER and OER properties, and is a highly efficient and stable bifunctional catalyst. At a current density of 50 mA cm−2, Ni2Co3-LDH has an overpotential of 227 mV in the HER and 303 mV in the OER, and has excellent cycling stability in alkaline electrolytes. In addition, Ni2Co3-LDH exhibits excellent electrocatalytic performance. It can be efficiently hydrolyzed at a current density of 50 mA cm−2 at 1.83 V and has long-term durability. Its excellent performance may be related to the special spatial structure, sufficient interfacial reactions, and the synergy between the components. This further expands the scope of non-precious metal electrocatalysts for efficient water splitting.

Data availability

Data are available on request from the authors.

Conflicts of interest

The authors declare no conflict of interest.

Acknowledgements

This work was financially supported by the Liaoning Applied Basic Research Program (No. 2023JH2/101300011, No. 2023JH2/101300018), the Basic Scientific Research Project of Liaoning Province Department of Education (No. LJKZZ20220024), and the Shenyang Science and Technology Project (No. 23-407-3-13).

References

  1. Y. Gong, J. Yao and P. Wang, et al., Perspective of hydrogen energy and recent progress in electrocatalytic water splitting[J], Chin. J. Chem. Eng., 2022, 43, 282–296 CrossRef CAS.
  2. J. Wang, P. Chen and S. Ruan, et al., Construction and optimization of multi-interface nanotube structured phosphorus-doped bimetallic oxide arrays as efficient electrocatalysts for water splitting[J], CrystEngComm, 2024, 26(20), 2692–2703 RSC.
  3. L. Tian, Y. Liu and C. He, et al., Hollow heterostructured nanocatalysts for boosting electrocatalytic water splitting[J], Chem. Rec., 2023, 23(2), e202200213 CrossRef CAS PubMed.
  4. J. Gu, Y. Peng and T. Zhou, et al., Porphyrin-based framework materials for energy conversion[J], Nano Res. Energy, 2022, 1, e9120009 CrossRef.
  5. W. R. Yan, Y. Xue and M. C. Liu, et al., Room temperature and rapid synthesis of two-dimensional bimetallic NiCo-CAT MOFs by an electrochemical strategy for enhancing electrocatalytic oxygen evolution reaction[J], CrystEngComm, 2024, 26, 3185 RSC.
  6. J. Gu, Y. Peng and T. Zhou, et al., Porphyrin-based framework materials for energy conversion[J], Nano Res. Energy, 2022, 1, e9120009 CrossRef.
  7. M. Chatenet, B. G. Pollet and D. R. Dekel, et al., Water electrolysis: from textbook knowledge to the latest scientific strategies and industrial developments[J], Chem. Soc. Rev., 2022, 51(11), 4583–4762 RSC.
  8. J. Xie, X. Zhang and H. Zhang, et al., Intralayered Ostwald Ripening to Ultrathin Nanomesh Catalyst with Robust Oxygen-Evolving Performance[J], Adv. Mater., 2017, 29(10), 1604765 CrossRef.
  9. B. Feng, S. Jin and J. Lang, et al., FeS nanosheets assembled with 1T-MoS2 nanoflowers on iron foam for efficient overall water splitting[J], CrystEngComm, 2024, 26(17), 2269–2276 RSC.
  10. C. Ye, L. Zhang and L. Yue, et al., A NiCo LDH nanosheet array on graphite felt: an efficient 3D electrocatalyst for the oxygen evolution reaction in alkaline media[J], Inorg. Chem. Front., 2021, 8(12), 3162–3166 RSC.
  11. G. Givirovskiy, V. Ruuskanen and T. Väkiparta, et al., Electrocatalytic performance and cell voltage characteristics of 1st-row transition metal phosphate (TM-Pi) catalysts at neutral pH[J], Mater. Today Energy, 2020, 17, 100426 CrossRef.
  12. H. Bai, D. Chen and Q. Ma, et al., Atom doping engineering of transition metal phosphides for hydrogen evolution reactions[J], Electrochem. Energy Rev., 2022, 5(2), 24 CrossRef CAS.
  13. J. Yuan, J. Zhou and Z. Peng, et al., Enhanced electrocatalytic hydrogen evolution in alkaline saline electrolyte by NiCo foam supported iridium nanoclusters[J], J. Mater. Chem. A, 2024, 12(4), 2383–2390 RSC.
  14. D. Zhao, M. Dai and H. Liu, et al., PPy film anchored on ZnCo2O4 nanowires facilitating efficient bifunctional electrocatalysis[J], Mater. Today Energy, 2021, 20, 100637 CrossRef CAS.
  15. Y. Guo, T. Park and J. W. Yi, et al., Nanoarchitectonics fortransition-metal-sulfide-based electrocatalysts for water splitting[J], Adv. Mater., 2019, 31(17), 1807134 CrossRef.
  16. Y. Sun, R. Li and X. Chen, et al., A-site management prompts the dynamic reconstructed active phase of perovskite oxide OER catalysts[J], Adv. Energy Mater., 2021, 11(12), 2003755 CrossRef CAS.
  17. Q. Wang, X. Huang and Z. L. Zhao, et al., Ultrahigh-loading of Ir single atoms on NiO matrix to dramatically enhance oxygen evolution reaction[J], J. Am. Chem. Soc., 2020, 142(16), 7425–7433 CrossRef CAS PubMed.
  18. M. Wang, X. Liu and Y. Sun, et al., High-efficiency NiCo layered double hydroxide electrocatalyst[J], New J. Chem., 2022, 46(38), 18535–18542 RSC.
  19. P. Wang, Y. Luo and G. Zhang, et al., Interface engineering of NixSy@MnOxHy nanorods to efficiently enhance overall-water-splitting activity and stability[J], Nano-Micro Lett., 2022, 14(1), 120 CrossRef CAS PubMed.
  20. X. W. Lv, W. W. Tian and Z. Y. Yuan, Recent advances in high-efficiency electrocatalytic water splitting systems[J], Electrochem. Energy Rev., 2023, 6(1), 23 CrossRef CAS.
  21. J. Yan, L. Chen and X. Liang, Co9S8 nanowires@NiCo-LDH nanosheets arrays on nickel foams towards efficient overall water splitting[J], Sci. Bull., 2019, 64(3), 158–165 CrossRef CAS PubMed.
  22. J. Hou, Y. Wu and S. Cao, et al., Active sites intercalated ultrathin carbon sheath on nanowire arrays as integrated core–shell architecture: highly efficient and durable electrocatalysts for overall water splitting[J], Small, 2017, 1702018 CrossRef PubMed.
  23. Z. Wu, Q. Li and G. Xu, et al., Microwave Phosphine-Plasma-Assisted Ultrafast Synthesis of Halogen-Doped Ru/RuP2 with Surface Intermediate Adsorption Modulation for Efficient Alkaline Hydrogen Evolution Reaction[J], Adv. Mater., 2024, 36(13), 2311018 CrossRef CAS PubMed.
  24. H. Yu, W. Wang and Q. Mao, et al., Pt single atom captured by oxygen vacancy-rich NiCo layered double hydroxides for coupling hydrogen evolution with selective oxidation of glycerol to formate[J], Appl. Catal., B, 2023, 330, 122617 CrossRef CAS.
  25. B. Wang, J. Li and D. Li, et al., Single atom iridium decorated nickel alloys supported on segregated MoO2 for alkaline water electrolysis[J], Adv. Mater., 2024, 36(11), 2305437 CrossRef CAS.
  26. X. Tan, Z. Duan and H. Liu, et al., Core-shell structured MoS2/Ni9S8 electrocatalysts for high performance hydrogen and oxygen evolution reactions[J], Mater. Res. Bull., 2022, 146, 111626 CrossRef CAS.
  27. J. Bao, Z. Wang and J. Xie, et al., The CoMo-LDH ultrathin nanosheet as a highly active and bifunctional electrocatalyst for overall water splitting[J], Inorg. Chem. Front., 2018, 5(11), 2964–2970 RSC.
  28. F. Song and X. Hu, Ultrathin cobalt–manganese layered double hydroxide is an efficient oxygen evolution catalyst[J], J. Am. Chem. Soc., 2014, 136(47), 16481–16484 CrossRef CAS PubMed.
  29. W. Ye, X. Fang and X. Chen, et al., A three-dimensional nickel–chromium layered double hydroxide micro/nanosheet array as an efficient and stable bifunctional electrocatalyst for overall water splitting[J], Nanoscale, 2018, 10(41), 19484–19491 RSC.
  30. J. Cai, Y. Song and Y. Zang, et al., N-induced lattice contraction generally boosts the hydrogen evolution catalysis of P-rich metal phosphides[J], Sci. Adv., 2020, 6(1), eaaw8113 CrossRef CAS PubMed.
  31. Y. Zhao, Z. Niu and J. Zhao, et al., Recent advancements in photoelectrochemical water splitting for hydrogen production[J], Electrochem. Energy Rev., 2023, 6(1), 14 CrossRef CAS.
  32. H. W. Chen and C. Li, PEDOT: fundamentals and its nanocomposites for energy storage[J], Chin. J. Polym. Sci., 2020, 38(5), 435–448 CrossRef CAS.
  33. H. Liu, D. Zhao and M. Dai, et al., PEDOT decorated CoNi2S4 nanosheets electrode as bifunctional electrocatalyst for enhanced electrocatalysis[J], Chem. Eng. J., 2022, 428, 131183 CrossRef CAS.
  34. X. Wang, H. Song and S. Ma, et al., Template ion-exchange synthesis of Co-Ni composite hydroxides nanosheets for supercapacitor with unprecedented rate capability[J], Chem. Eng. J., 2022, 432, 134319 CrossRef CAS.
  35. H. Jia, Z. Wang and X. Zheng, et al., Interlaced Ni-Co LDH nanosheets wrapped Co9S8 nanotube with hierarchical structure toward high performance supercapacitors[J], Chem. Eng. J., 2018, 351, 348–355 CrossRef CAS.
  36. W. Liang, M. Wang and C. Ma, et al., NiCo-LDH Hollow Nanocage Oxygen Evolution Reaction Promotes Luminol Electrochemiluminescence[J], Small, 2023, 2306473 Search PubMed.
  37. Y. Yang, L. Dang and M. J. Shearer, et al., Highly active trimetallic NiFeCr layered double hydroxide electrocatalysts for oxygen evolution reaction[J], Adv. Energy Mater., 2018, 8(15), 1703189 CrossRef.
  38. S. Yao, Y. Jiao and C. Lv, et al., Lattice-strain engineering of CoOOH induced by NiMn-MOF for high-efficiency supercapacitor and water oxidation electrocatalysis[J], J. Colloid Interface Sci., 2022, 623, 1111–1121 CrossRef CAS.
  39. W. Guo, C. Dun and C. Yu, et al., Mismatching integration-enabled strains and defects engineering in LDH microstructure for high-rate and long-life charge storage[J], Nat. Commun., 2022, 13(1), 140 CrossRef PubMed.

This journal is © The Royal Society of Chemistry 2024
Click here to see how this site uses Cookies. View our privacy policy here.