Recent advances in various applications of nickel cobalt sulfide-based materials

Gaofei Xue ab, Tian Bai ab, Weiguo Wang ab, Senjing Wang b and Meidan Ye *ab
aShenzhen Research Institute of Xiamen University, Shenzhen 518057, China. E-mail: mdye@xmu.edu.cn
bResearch Institute for Biomimetics and Soft Matter, Fujian Provincial Key Laboratory for Soft Functional Materials Research, Department of Physics, Xiamen University, Xiamen 361005, China

Received 13th January 2022 , Accepted 7th March 2022

First published on 9th March 2022


Abstract

In recent years, nickel cobalt sulfides (NCSs) have received much attention as promising functional materials in various application fields, mainly due to their much lower price, abundant raw materials and considerable reactive activity with relatively higher electrical conductivity, weaker metal–sulfur bonds and better thermal stability compared to their oxide counterparts. In this review, we will briefly summarize the recent development of NCSs on the aspects of structural design, component optimization and composite preparation. Moreover, we will comprehensively review the recent applications of NCS nanomaterials in various fields (e.g., supercapacitors, batteries, electrocatalysis, photocatalysis, sensors and microwave absorption), and some representative examples will be highlighted for different applications of NCSs. Finally, we will outline the common problems and prospects of NCS nanomaterials used in various applications.


image file: d2ta00305h-p1.tif

Gaofei Xue

Gaofei Xue received his B.S. in Material Physics from Fujian Normal University, China in 2020. He is now a graduate student in the College of Physical Science and Technology, Xiamen University, China. He joined Professor Meidan Ye's group in 2020. His research focuses on flexible materials and devices.

image file: d2ta00305h-p2.tif

Tian Bai

Tian Bai received his B.S. from the Shaanxi University of Science & Technology, China in 2018. He is now a graduate student in the College of Physical Science and Technology, Xiamen University, China. He joined Professor Meidan Ye's group in 2019. His research focuses on 2D material-based flexible devices.

image file: d2ta00305h-p3.tif

Weiguo Wang

Weiguo Wang received his B.S. from Fujian Agriculture and Forestry University, China in 2019. He is now a graduate student in the College of Physical Science and Technology, Xiamen University, China. He joined Professor Meidan Ye's group in 2019. His research focuses on flexible materials and devices.

image file: d2ta00305h-p4.tif

Senjing Wang

Senjing Wang received his B.S. in Measurement Control Technology and Instruments from Xiamen University, China in 2021. He is now a graduate student in the College of Physical Science and Technology, Xiamen University, China. He joined Professor Meidan Ye's group in 2021. His research focuses on energy storage materials and devices.

image file: d2ta00305h-p5.tif

Meidan Ye

Meidan Ye received her PhD from the College of Chemistry and Chemical Engineering at Xiamen University in 2014. She then joined the Research Institute for Soft Matter and Biomimetics, Department of Physics, College of Physical Science and Technology at Xiamen University as an associate Professor in 2014. She was promoted to a Professor in 2021. Her research interests are on multi-functional materials for flexible devices, i.e., electrochemical energy storage devices and wearable sensors.


1. Introduction

In recent years, transition metal chalcogenides have caught much research attention owing to their unique physicochemical characteristics, low cost, and broad application potential.1–7 Among them, ternary sulfides have drawn rather extensive interest in the field of energy conversion and storage (e.g., solar cells, water splitting, hydrogen generation, supercapacitors, Li/Na-ion (Li–S, Li–O2, and Zn–air) batteries, etc.) because of their outstanding electrochemical performance.8,9 In particular, nickel cobalt sulfides (NCSs) have become one of the most concerned ternary sulfides, mainly due to their relatively high electrical conductivity, multiple valence states and rich crystallographic structures.10,11 It is generally considered that nickel holds tetrahedral sites, while cobalt takes over both octahedral and tetrahedral sites within spinel NiCo2S4 (Fig. 1). The mixture of heterogeneous spin states from Ni and Co atoms will result in lattice distortion and subtle atomic arrangement due to the mismatch in the Jahn–Teller distortion degree (Co2+, t2g6eg1, strong John–Teller effect; Ni2+, t2g6eg2, no John–Teller effect), which would produce more exposed active octahedral edge sites of NCSs.12 This is beneficial for more electrochemical reactions (e.g., pseudocapacitor or catalysis) and thus the synergistic effect between Ni and Co can effectively boost the electrochemical performance of NCSs with the coexistence of Ni2+, Ni3+, Co2+, and Co3+ ions. Compared to nickel cobalt oxides (Fig. 1), NCSs present relatively lower optical energy band gaps (NiCo2S4: 1.2 eV vs. NiCo2O4: 2.5 eV) and higher conductivity (∼100 times).13,14 This enables the lowering of the charge-transfer resistance of NCSs and facilitates charge transport when they are applied in energy conversion and storage systems. Theoretical calculation results indicate that NiCo2S4 has a smaller band gap than NiCo2O4 (Fig. 1d and e), suggesting its higher electric conductivity. And NiCo2S4 possesses a lower diffusion energy than NiCo2O4 (Fig. 1c), revealing that ions can more easily diffused in NiCo2S4.15 The replacement of oxygen by sulfur for NCSs will create more flexible structures since the electronegativity of sulfur is lower than that of oxygen.16 This can alleviate the structural deformation originating from the intercalation of electrolyte ions and thus promote the reactions between electrolyte ions and NCSs, yielding much enhanced electrochemical performance.
image file: d2ta00305h-f1.tif
Fig. 1 Schematic diagram of the crystal structures of (a) NiCo2S4 and (b) NiCo2O4. (c) The Li ion diffusion battier energies and (d and e) the density of states (DOS) of NiCo2S4 and NiCo2O4 calculated by using density functional theory (DFT) simulation.15 Reproduced from ref. 10 with permission from Elsevier.

In recent years, much effort has been devoted to the development of NCSs, which can be roughly divided into several aspects.

The first is the design of micro-/nano-structures (Fig. 2 and 3). In addition to conventional nanostructures (e.g., zero-dimensional (0D) nanoparticles,17 one-dimensional (1D) nanowires/nanorods/nanotubes,18,19 and two-dimensional (2D) nanosheets20), some special morphologies as well have been designed, such as nanotube-woven hexagonal microsheets,21 nanosheet-constructed hollow nanocages,22 nanoparticle-stacked microspheres,23 hexagonal nanoplate-assembled microstructures,24 and so on. Such hierarchical structures are beneficial to further amplify the active reaction sites, facilitate the charge transport and promote the ion diffusion, ultimately strengthening the performance of NCSs.25–28 Although the hydrothermal/solvothermal method is still the most common way to synthesize the nanostructures of NCS materials,29 multi-step strategies combining different processes (e.g., solution reactions (hydrothermal/solvothermal, sol–gel and chemical bath reactions), electrodeposition and high-temperature calcination) are becoming more and more popular to construct high-performance morphologies of NCSs.30,31 More impressively, in situ growth of NCS nanostructured films directly onto conductive substrates (e.g., carbon cloth and Ni foam; Fig. 2 and 3d–f) is also well accepted mainly because it enables the simplification of the synthesis processes, removal of the use of binders, avoidance of the aggregation of nanomaterials, and reduction of the interfacial resistance between NCS films and substrates, therefore improving the performance of NCSs.19,32


image file: d2ta00305h-f2.tif
Fig. 2 Schematic diagrams of the crystal structures, nanostructures and applications of NCSs.

image file: d2ta00305h-f3.tif
Fig. 3 Some representative NCS structures: transmission electron microscopy (TEM) images of (a) NiCo2S4 @ N-doped carbon hollow capsules,48 (b) NiCo2S4/Co9S8 submicro-spindles,49 and (c) NiCo2S4 double-shelled ball-in-ball hollow spheres;50 SEM images of (d) NiCo2S4 nanosheets @ Ni foam,51 (e) caterpillar-like NiCo2S4 arrays @ carbon cloth,52 and (f) NiCo2S4 nanosheets @ Ni columns.53 Reproduced from ref. 17, 18, 20 and 21 with permission from the Royal Society of Chemistry. Reproduced from ref. 19 and 22 with permission from American Chemical Society.

The second is the study of the stoichiometric effect. For ternary or quaternary semiconductors, the stoichiometric ratio of their elements is usually adjustable, which also plays an important role in their properties and device performance.33,34 A variety of studies have been performed to investigate this stoichiometric effect on NCSs, and various NCSs with different element stoichiometry have been prepared, such as NiCo2S4,19,35 CoNi2S4,36,37 Ni1.4Co1.6S4,38 Ni1.5Co1.5S4,39 (Co0.5Ni0.5)9S8,40 (Ni0.33Co0.67)9S8,41 Ni1.77Co1.23S4,42 Ni4.5Co4.5S8,43 Ni0.67Co0.33S,44etc. The comparison of the relative performance of NCSs with different element stoichiometry has also been discussed in depth (Fig. 4).45 It is proposed that the high conductivity of NCSs is mainly due to the occupation of Ni in the tetrahedral sites of the spinel structure, the low bandgap and the synergistic coupling effect between Co and Ni with p-type (Co) and n-type (Ni) doping.46 Then, the different ratios of Ni/Co contents in NCSs will influence their performance in applications. For example, three NCSs with varied Ni/Co ratios (Ni0.8Co2.2S4, Ni1.5Co1.5S4, and Ni2.2Co0.8S4) were synthesized to study the composition effect on their electrocatalytic activity.47 It was found that Ni2.2Co0.8S4 had the highest current density and the lowest overpotential for the oxygen evolution reaction (OER) due to its largest content of Ni3+ with the conversion of Ni3+ to Ni4+ for producing more active sites to adsorb –OOH, while Ni1.5Co1.5S4 exhibited the highest current density for the oxygen reduction reaction (ORR) owing to the larger content of Ni2+ and Co2+ with superior conductivity and the synergetic coupling between Ni and Co.47


image file: d2ta00305h-f4.tif
Fig. 4 SEM images (a–d) and electrochemical performance (e–g) of metal sulfides composed of different Ni/Co molar ratios: (a) pure Co, (b) pure Ni, (c) 1/2, and (d) 2/1.45 Electrochemical performance (h–j) of NiCo2S4 films with different concentrations of surface sulfur vacancies.20 Reproduced from ref. 60 with permission from American Chemical Society. Reproduced from ref. 61 with permission from the Royal Society of Chemistry.

The third is the fabrication of composites. NCS-based composites also have been widely fabricated to compensate for the shortcomings of NCSs (e.g., relatively low stability and slow mass transfer/reaction kinetics). NCSs can be rationally integrated with lots of other materials, for example, metal oxides (e.g., CeO2,54 CoMoO4,55 NiCo2O4,43 Co3O4,56etc.), sulfides (e.g., FeCo2S4,57 Co9S8,58 MgS,59 MoS2,60 SnS2,61etc.), selenides (e.g., Co0.85Se,62 MoSe2,63etc.), phosphides (e.g., NiCoP,64 NiFeP,65etc.), carbon-based materials (e.g., carbon nanotubes (CNTs),66 graphene,67–69 carbon dots,70 mesoporous carbon spheres,71etc.), polymers (e.g., polypyrrole,72 polyanilnie,73etc.), layered double hydroxides (LDHs; NiMn-LDHs,74 CoNo-LDHs,75 NiFe-LDHs,76etc.), two-dimensional titanium carbide (MXene),77 and others (e.g., g-C3N4,78 Ag,44etc.). Besides, element doping (e.g., Mn,79 and Cu80) also have been widely used to further enhance the performance of NCSs.

The fourth is the development of applications (Fig. 2). NCSs are one type of the most popular electrode material used in supercapacitors due to their excellent electrochemical activity.53,81 Besides, NCSs have been gradually applied to other electrochemical energy storage systems over the years, mainly including Li/Na-ion batteries,18,82 Li–O2 batteries,83 Zn–air batteries,69 and Li–S batteries.35 NCSs are also able to act as effective counter electrodes in dye/quantum dot-sensitized solar cells (DSCs/QDSCs).84,85 NCSs as well have great activity in photocatalytic reactions, such as the degradation of organic pollution and hydrogen evolution.86,87 Moreover, NCSs can be efficient electrocatalysts for overall water splitting with bi-functions of the oxygen evolution reaction (OER) and hydrogen evolution reaction (HER).88,89 Particularly, NCSs can be used as bio-sensing materials due to their attractive electrochemical characteristics90 and microwave absorbers because of their excellent magnetic loss properties.91

For NCSs, some articles have reviewed their applications in supercapacitors and batteries.1,10,11 However, few reviews comprehensively summarize promising applications for NCSs in different fields. In this review, we focus on the recent applications of NSCs in a variety of fields in the last few years. We discuss some representative studies about the applications of NCSs in energy storage devices, electrocatalysis, photocatalysis, sensors and microwave absorption. We highlight the intercommunity of NCSs in different application fields. Finally, the general problems and future prospects of NCSs for various applications are analyzed and emphasized.

2. Applications of NCSs in energy storage devices

2.1. Supercapacitors

NCSs have been intensively applied in supercapacitors due to their relatively higher conductivity and more redox sites in comparison with their corresponding oxides and binary sulfides. It is commonly considered that the pseudocapacitive performance of NCSs is mainly due to the Faraday redox reactions between Ni2+/Ni3+ and Co2+/Co3+. In an alkaline aqueous electrolyte, the reversible Faraday peaks can be associated with the reversible redox reactions from NiCo2S4 to NiSOH, CoSOH and CoSO species (Fig. 5). The theoretical capacitance of NiCo2S4 in supercapacitors is 2534 F g−1.1
image file: d2ta00305h-f5.tif
Fig. 5 Schematic diagram and reactions of NCSs in supercapacitors and Li-ion batteries.

In order to amplify the charge storage ability of NCSs, a series of strategies have been developed (Table 1), including regulating the element ratio (e.g., NiCo2S4,19,35 CoNi2S436,37 and NiCo2S4−x19), tuning the morphological structures in different dimensions (Fig. 6a and b),45,92–94 doping other elements (e.g., Fe,95 Mn,96 P,29 and V;97Fig. 6d–f), and integrating with other materials (e.g., CNTs,98 graphene,63 Nb2O5,99 Ni(OH)2,100 FeCo2S4,57 MXenes,77etc.). For example, the defect engineering strategy enables the production of rich sulfur vacancies on the surface of NCS crystals through various methods, such as low-temperature plasma induction (Fig. 4h–j),20 Ar/H2 gas-assisted annealing,23 and NaBH4 solution reduction.30 The existence of sulfur vacancies will change the electronic structures of NCSs to increase their conductivity and reactive sites for enhanced electrochemical performance.20 Besides, the 3D hierarchical structures of NCSs (Fig. 6) have been widely designed and fabricated for high-performance supercapacitors,93,101–103 mainly because they have attractive properties to offer sufficient surface areas for more faradaic redox reactions, and build fast channels for effective charge transfer and ion diffusion.

Table 1 Performance of different NCS-based materials in supercapacitors
Material Capacitance/capacity of electrodes Capacitance/capacity of devices Energy density/power density of devices Cycling stability of devices Ref.
Ni3S2/CoNi2S4 porous network 2435 F g−1 @ 2 A g−1 175 F g−1 @ 1 A g−1 40.0 W h kg−1/17.3 kW kg−1 92.8% capacitance retention after 6000 cycles @ 10 A g−1 111
ZnS/NiCo2S4/Co9S8 nanotubes 1618.1 F g−1 @ 1 A g−1 66 W h kg−1/0.75 kW kg−1 62% capacitance retention after 5000 cycles @ 5 A g−1 112
Ni–Co–S nanosheets @ graphene foam 2918.1 F g−1 @ 1 A g−1 209.82 F g−1 @ 1 A g−1 79.3 W h kg−1/0.83 kW kg−1 54.02% capacitance retention after 10[thin space (1/6-em)]000 cycles @ 5 A g−1 113
Zn-doped Ni–Co–S nanocomposite 2668 F g−1 @ 1 A g−1 150 F g−1 @ 1 A g−1 60.2 W h kg−1/0.85 kW kg−1 91.2% capacitance retention after 10[thin space (1/6-em)]000 cycles @ 5 A g−1 114
NiCo2S4/MXene 1028C g−1 @ 1 A g−1 171.2 F g−1 @ 1 A g−1 68.7 W h kg−1/0.85 kW kg−1 89.5% capacitance retention after 5000 cycles @ 5 A g−1 77
NiFeP@NiCo2S4 nanosheets 874.4C g−1 @ 1 A g−1 197.6 F g−1 @ 1 A g−1 87.9 W h kg−1/0.43 kW kg−1 85.2% capacitance retention after 10[thin space (1/6-em)]000 cycles @ 8 A g−1 65
MoS2/NiCo2S4@C hollow microspheres 250 mA h g−1 @ 2 A g−1 120 F g−1 @ 1 A g−1 53.01 W h kg−1/4.20 kW kg−1 90.1% capacitance retention after 10[thin space (1/6-em)]000 cycles @ 10 A g−1 115
Carbon nanofiber/NiCo2S4 nanosheets @ polypyrrole 2961 F g−1 @ 1 A g−1 163.3 F g−1 @ 1 A g−1 44.45 W h kg−1/0.70 kW kg−1 85.1% capacitance retention after 5000 cycles @ 5 A g−1 116
P-doped NiCo2S4−x nanotubes 1806.4C g−1 @ 1 A g−1 186.1 F g−1 @ 1 A g−1 68.2 W h kg−1/0.80 kW kg−1 97% capacitance retention after 10[thin space (1/6-em)]000 cycles @ 30 A g−1 19
NiMn-LDH@NiCo2S4 nanotubes 1018C g−1 @ 1 A g−1 193.0 F g−1 @ 0.5 A g−1 60.3 W h kg−1/0.38 kW kg−1 86.4% capacitance retention after 10[thin space (1/6-em)]000 cycles @ 10 A g−1 74
CoNi2S4 microspheres 1836.6 F g−1 @ 1 A g−1 100.7 F g−1 @ 1 A g−1 38.9 W h kg−1/0.85 kW kg−1 101.2% capacitance retention after 50[thin space (1/6-em)]000 cycles @ 5 A g−1 106
Ag@Ni0.67Co0.33S forest-like nanostructures 296.4 mA h g−1 @ 2 mA cm−2 1104.14 mF cm−2 @ 5 mA cm−2 0.36 mW h cm−2/3.81 mW cm−2 83.6% capacitance retention after 8000 cycles @ 20 mA cm−2 44



image file: d2ta00305h-f6.tif
Fig. 6 Schematic diagram (a) and electrochemical performance (b and c) of NiCo2S4 nanosheets @ Ni column arrays.53 SEM image (d) and electrochemical performance (e and f) of P-doped Ni–Co–S@C hierarchical structures.104 Reproduced from ref. 22 with permission from American Chemical Society. Reproduced from ref. 78 with permission from Elsevier.

It is worth noting that NCSs frequently show obvious battery-like charge storage behavior, that is, displaying typical redox peaks in cyclic voltammogram curves (CVs; vs. near-rectangular shapes or broadened peaks for capacitors; Fig. 6c) and clear discharge platforms in galvanostatic charge/discharge curves (GCDs; vs. quasi-triangular shapes for capacitors; Fig. 6e).105 It means that the reaction kinetics of NCSs in the charge storage processes are partially determined by the diffusion-controlled redox behavior (i.e., image file: d2ta00305h-t1.tif, where I is the current and v is the scan rate in CVs), while the reaction kinetics of pseudocapacitance are dominated by the surface-controlled faradaic process (i.e., Iv).58 Thus, the corresponding devices based on battery-like NCS electrodes are named hybrid supercapacitors due to the capacity contribution both from the diffusion-controlled and surface redox processes (Fig. 6f).44,81,106–109 However, we find that the contribution ratio of these two kinds of electrochemical behavior for NCSs is significantly influenced by the micro-/nano-structures of NCSs. The electrochemical performance of four types of NiCo2S4 hierarchical structures assembled from similar nanorod building blocks was investigated,110 and it was indicated that the hat/flower-like structures with a compact packaging way showed unclear CV redox peaks and GCD discharge platforms, suggesting that the diffusion-controlled process was limited and the capacity was mainly from the surface electrochemical process. In contrast, sphere/brush-like structures with a loose and orderly assembly way presented more obvious CV redox peaks and GCD discharge platforms, revealing a higher utilization of active materials from the increased contribution of diffusion-controlled faradaic reactions in the inner part of the electrode, and thus they had larger capacity.110 Moreover, NCSs and their composites are commonly in situ grown on planar substrates (e.g., carbon cloth, Ni foam and metal meshes), and they usually show clear battery-like behavior,58,81 but when the planar substrates are changed into ultrathin metal wires with a diameter on a scale of tens of micrometers and a high aspect ratio (Fig. 7a–d), clear CV redox peaks and GCD discharge platforms disappear (Fig. 7e and f), and most capacity is from the surface pseudocapacitance because of the relatively large surface area of ultrathin wires (Fig. 7g and h).59 Therefore, it is possible to modify NCS electrodes on different aspects (e.g., components, morphologies, substrates and so on) to achieve high-performance electrochemical energy storage devices.


image file: d2ta00305h-f7.tif
Fig. 7 SEM images (a, c and d), schematic diagram (b) and electrochemical performance (e–h) of NiCo2S4/MgS composites.59 Reproduced from ref. 28 with permission from the Royal Society of Chemistry.

2.2. Batteries

In recent years, NCSs have also been employed as effective electrode materials in other electrochemical energy storage devices (Table 2), such as Li/Na/Zn-ion, Li–S, Li–O2 and Zn–air batteries, similarly because of their intrinsic properties of high electrochemical activity stemming from their rich redox reactions with multi-valence states and relatively high electrical conductivity with the synergetic effect from Ni and Co atoms (Fig. 5). The theoretical capacity of NCSs in Li-ion batteries is 703 mA h g−1.1 For example, NiCo2S4 hollow nanowires (Fig. 8a) were uniformly grown on a 3D N-doped carbon nanosheet substrate derived from a bio-polymer of Chitin (poly(b-(1,4)-N-acetyl-D-glucosamine)),18 which exhibited attractive long-term cycling stability with a revisable capacity of 1198 mA h g−1 at 500 mA g−1 after 500 cycles (Fig. 8b). Besides, (Co0.5Ni0.5)9S8 hollow microspheres (Fig. 8c) were converted from the Ni–Co metal–organic framework (Ni–Co-MOF) and then modified with N-doped carbon (N–C).40 When applied as the anode in Na-ion batteries with Na3V2(PO4)3@C as the cathode, the full cells delivered a capacity of 737.6 mA h g−1 at 0.5 A g−1 within 0.01–3.7 V after 50 cycles. It was found that the composite electrodes had a hybrid electrochemical charge storage behavior and the pseudocapacitive contribution was up to 88.8% at 1 mV s−1 (Fig. 8d). In addition, NiCo2S4 porous nanoneedle arrays were in situ grown on N/S co-doped carbon cloth and displayed high sulfur loading capability due to the rich mesoporous structure and a great ability to absorb and convert polysulfide intermediates owing to the high electronic conductivity and catalytic activity.117 Therefore, NCS-based materials are as well promising for Li–S batteries.
Table 2 Performance of some representative NCS-based materials in different kinds of batteries
Materials Specific capacity Cycling stability Remarks Ref.
Li-ion batteries
NiCo2S4/SnS2 hollow spheres (anode) 1260 mA h g−1 @ 0.1 A g−1 627 mA h g−1 @ 0.5 A g−1 after 300 cycles 99 mA h g−1 @ 0.1 A g−1 within 2.0–5.0 V (full cells with NMC111 cathodes and NiCo2S4/SnS2 anodes) 61
NiCo2S4@C hollow microspheres (anode) 1946.7 mA h g−1 @ 0.2 A g−1 1242.8 mA h g−1 @ 2 A g−1 after 1000 cycles Multi-shelled NiCo2S4 hollow microspheres were obtained by a multi-step method with solvothermal reactions, direct pyrolysis under air atmosphere and calcination in Ar/H2S gas 126
NiCo2S4 hollow nanowires (anode) 1437.9 mA h g−1 @ 0.5 A g−1 1198 mA h g−1 @ 0.5 A g−1 after 500 cycles Chitin-derived N-doped carbon was used as a 3D substrate to support NiCo2S4 nanowires 18
[thin space (1/6-em)]
Na-ion batteries
NiCo2S4 @ N-doped carbon nanoparticles (anode) 742.2 mA h g−1 @ 0.2 A g−1 395.6 mA h g−1 @ 6 A g−1 after 5000 cycles Ether-based electrolyte: sodium trifluoromethanesulfonate (NaCF3SO3)/diethylene glycol dimethyl ether (DEGDME) 120
Flower-like NiCo2S4 structures (anode) 748 mA h g−1 @ 0.1 A g−1 376 mA h g−1 @ 2 A g−1 after 500 cycles The capacitance contribution is up to 89% of total capacity @ 2 mV s−1 82
(Co0.5Ni0.5)9S8 @ N-doped carbon hollow spheres (anode) 945.1 mA h g−1 @ 0.1 A g−1 723.7 mA h g−1 @ 1 A g−1 after 100 cycles 847.5 mA h g−1 @ 0.5 A g−1 within 0.01–3.7 V (full cells with Na3V2(PO4)3@C cathodes and (Co0.5Ni0.5)9S8/N–C anodes) 40
[thin space (1/6-em)]
Zn-ion batteries
NiCoP/NiCo2S4 nanosheets (cathode) 251.1 mA h g−1 @ 10 A g−1; 190.3 mA h g−1 @ 50 A g−1 96.9% capacity retention @ 100 mV s−1 after 5000 cycles for full cells 265.1 mA h g−1 @ 5 A g−1 at 1.7 V (full cells with NiCoP/NiCo2S4 cathodes and Zn anodes) 127
NiCo2S4−x nanotubes @ carbon cloth (cathode) 298.3 mA h g−1 @ 0.5 A g−1 144.4 mA h g−1 @ 5.0 A g−1 after 1000 cycles (85% retention) 146.3 mA h g−1 @ 5 A g−1 and 84.7% retention after 1500 cycles (flexible solid-state NiCo2S4−x@CC//Zn@CC with a sodium polyacrylate hydrogel electrolyte) 34
[thin space (1/6-em)]
Li–S batteries
NiCo2S4 nanosheets @ carbon textile (CT) (cathode) 1600 mA h g−1 @ 0.1C 836 mA h g−1 @ 0.5C after 500 cycles (0.018% per cycle) NiCo2S4 @ CT can be as a bifunctional interlayer for the S@MWCNT cathode with higher adsorptive ability for polysulfides 35
Hollow acicular-like NiCo2S4 microspheres (cathode) 1387 mA h g−1 @ 0.2C 543 mA h g−1 @ 2C after 500 cycles (0.06% per cycle) A flexible Li–S battery with a sulfur loading of 8.9 mg cm−2 has a capacity of 720 mA h g−1 (6.52 mA h cm−2) after 65 cycles at 0.1C 128
NiCo2S4−x hollow microspheres (cathode) 1304.5 mA h g−1 @ 0.2C; 628.9 mA h g−1 @ 5C 598.2 mA h g−1 @ 1C after 500 cycles (0.0754% per cycle) NiCo2S4−x has strong chemical adsorption ability toward LiPSs 23
[thin space (1/6-em)]
Zn–air batteries
(Ni,Co)S2 nanosheets (cathode) Specific capacity: 842 mA h gZn−1 @ 5 mA cm−2; charge/discharge voltage gap: 0.45 V @ 2 mA cm−2; peak power density: 153.5 mW cm−2 480 h @ 2 mA cm−2 Cell charge/discharge voltage: 1.71/1.26 V @ 10 mA cm−2 129
NiCo2S4 nanoparticles @ carbon nitrogen nanosheets (cathode) Specific capacity: 801 mA h gZn−1 @ 10 mA cm−2; charge/discharge voltage gap: 0.69 V @ 10 mA cm−2; energy density: 1025 W h kg−2 180 h @ 10 mA cm−2 (1000 cycles) Cell charge/discharge voltage: 1.94/1.25 V @ 10 mA cm−2 78
[thin space (1/6-em)]
Li–O 2 batteries
NiCo2S4 nanorods @ carbon textile (cathode) Specific capacity: 4506 mA h g−1 @ 750 mA g−1; charge/discharge voltage gap: 0.69 V @ 10 mA cm−2; energy density: 1025 W h kg−2 500 cycles with a cut-off capacity of 1000 mA h g−1 @ 500 mA g−1 Charge overpotential: 1.17 V 83
NiCo2S4 nanoflakes/S-doped carbon nanosheets (cathode) Specific capacity: 14[thin space (1/6-em)]173 mA h g−1 @ 150 mA g−1; charge/discharge voltage gap: 0.69 V @ 10 mA cm−2; energy density: 1025 W h kg−2 1704 h with a limited specific capacity of 1000 mA h g−1 @ 150 mA g−1 Initial overall overpotential: 1.27 V with 1.01 and 0.26 V at charge and discharge stages (1.01/0.26 V) with 68% round-trip efficiency 130



image file: d2ta00305h-f8.tif
Fig. 8 SEM image (a) and cycling performance (b) of NiCo2S4/N-doped carbon.18 SEM image (c) and capacitive contribution (d and e) of (Co0.5Ni0.5)9S8/N-doped carbon.40 Reproduced from ref. 49 and 12 with permission from the Royal Society of Chemistry.

Recently, besides the ex situ measurement ways,118in situ characterization methods (e.g., in situ X-ray diffraction (XRD), photoluminescence (PL) and electrochemical impedance spectroscopy (EIS)) have been extensively developed to reveal more details about the charge storage processes for batteries. Thus, some studies have also reported the in situ characterization and analysis of the working mechanism of NCSs in Li/Na-ion batteries.41,66,119–122 For example, in situ XRD of the NiCo2S4@CNT electrodes in Li/Na-ion batteries showed that in the first discharge process with Li/Na ion intercalation, phase change occurred with the formation of Ni/Co particles, Li2S and Na2S, while in the charge process, NiS and CoS would generate due to the deintercalation of Li/Na ions (Fig. 9).66 In addition, the in situ EIS analysis of the interfacial properties and charge transfer resistance of NiCo2S4@carbon electrodes also confirmed the phase change with the conversion of NiCo2S4 into NiSx and CoSx after 200 cycles.120 The participation of carbon materials enabled the improvement of the electrical conductivity and ensured the high reversibility of structural and phase transition of NiCo2S4 during charge/discharge cycles, thus enhancing the stability of NCS-based batteries.66,120 Therefore, it will be very beneficial to deeply elucidate the working mechanism and key processes in electrochemical energy storage devices by combining in situ and ex situ characterization methods, ultimately guiding the design of more high-performance devices.


image file: d2ta00305h-f9.tif
Fig. 9 Schematic diagram of the charge storage mechanism of NiCo2S4@CNT electrodes in Li/Na-ion batteries.66 Reproduced from ref. 53 with permission from Elsevier.

For the applications of Zn–air and Li–O2 batteries, the electrocatalytic activity and durability in the oxygen reduction reactions (ORR) for the discharge process and the oxygen evolution reactions (OER) for the charge process are very essential for advanced cathode electrode materials. Many studies have confirmed that NCSs are feasible for metal–oxygen batteries as a class of bifunctional cathodes,69,78,123–125 still due to their excellent electrochemical activity and reversibility in redox reactions. For example, NiCo2S4 nanoparticles were integrated with carbon nitrogen nanosheets and demonstrated efficient bifunctional electrocatalytic performance (i.e., a positive half-wave potential of 0.83 V for the ORR and a low overpotential of 360 mV at 10 mA cm−2 for the OER).78 The corresponding Zn–air batteries showed a maximum power density of 92 mW cm−2, a capacity of 801 mA h g−1, an energy density of 1025 W h kg−1, and outstanding cycling durability for 180 h with 1000 cycles at 10 mA cm−2, which were comparable or superior to those (106 mW cm−2, 759 mA h g−1, 925 W h kg−1, and 69 h) of the commercial Pt/C–RuO2-based one.

3. Applications of NCSs in water splitting

Nowadays, electrochemical water splitting is widely recognized as one of the most prospective strategies to generate hydrogen (Fig. 10a), which is mostly accepted as an ideal energy choice to substitute for conversional fossil fuels. There are two important processes involved in the electrocatalysis of water, that is, the hydrogen evolution reaction (HER) and oxygen evolution reaction (OER), which require highly active electrocatalysts to overcome the large overpotential (1.6–2.0 V) of these reactions. To date, IrO2 and Pt have been considered as the best catalysts for the OER and HER; nevertheless, the high price of noble metals restricts their large-scale applications. Thus, alternative electrocatalysts with low cost and high activity have attracted much attention. Because of the co-existence of Ni and Co with rich active sites from the redox couples (i.e., Co3+/Co2+ and Ni3+/Ni2+) in crystal structures, NCSs show higher conductivity and catalytic capability than binary sulfides (Fig. 10b–d), and then have been applied in electrocatalytic water splitting as a promising type of bifunctional catalyst (Fig. 10a and Table 3).131–133
Table 3 Performance of different NCS-based catalysts for water splitting
Catalyst Overpotential of the OER Overpotential of the HER Water splitting Electrolyte Stability Ref.
Core–shell NiCo2S4 nanorods 200 mV vs. 330 mV (IrO2) @ 40 mA cm2 190 mV @ 10 mA cm−2 1.57 V vs. 1.54 V (IrO2/C–Pt/C) @ 10 mA cm2 1.0 M KOH 12 h 141
NiCo2S4@Ni3V2O8 290 mV vs. 320 mV (IrO2) @ 35 mA cm2 1.0 M KOH 10 h 132
NiCo2S4 nanorod arrays 220 mV vs. ∼320 mV (IrO2) @ 30 mA cm2 222 mV vs. ∼30 mV (Pt/C) @ 30 mA cm2 1.64 V @ 20 mA cm−2 1.0 M KOH 18 h 142
NiCo2S4 nanoflakes 319 mV vs. ∼404 mV (NiCo2O4) @ 100 mA cm2 169 mV vs. ∼238 mV (NiCo2O4) @ 10 mA cm2 1.61 V @ 10 mA cm−2 1.0 M KOH 70 h 88
Crossed NiCo2S4 nanowires 256 mV vs. 340 mV (IrO2) @ 40 mA cm2 193 mV vs. ∼244 mV (Ni3S2) @ 10 mA cm2 1.59 V @ 10 mA cm−2 1.0 M KOH 12 h 134
NiCo2S4 nanosheets 247 mV vs. 264 mV (NiCo3S4) @ 10 mA cm2 106 mV vs. ∼181 mV (NiCo3S4) @ 10 mA cm2 1.62 V @ 10 mA cm−2 1.0 M KOH 72 h 89
N-doped carbon–NiCo2S4 hollow nanotubes 280 mV vs. ∼290 mV (IrO2) @ 10 mA cm2 183 mV vs. ∼214 mV (NiCo2S4) @ 10 mA cm2 1.60 V @ 10 mA cm−2 1.0 M KOH 15 h 135
Carbon dots/NiCo2S4/Ni3S2 nanorods 116 mV @ 10 mA cm2 127 mV @ 10 mA cm2 1.50 V @ 10 mA cm−2 1.0 M KOH 12 h 131
NiCo2S4@N,S co-doped reduced graphene oxides 253 mV vs. 267.8 mV (RuO2) @ 10 mA cm2 92.7 mV vs. 61.9 mV (Pt/C) @ 10 mA cm2 1.58 V in 1 M KOH and 1.907 V in 1 M phosphate buffer solution @10 mA cm−2 1.0 M KOH 10 h 136


Much effort has been devoted to the rational engineering of the components and nanostructures of NCSs,88,89,131,132,134,135 in order to further enlarge their reactive surface areas, improve their electrical conductivity and enhance their catalytic stability, ultimately synergistically strengthening their electrocatalytic performance. For example, nanorod assembled NiCo2S4 microspheres were in situ planted onto N,S co-doped reduced graphene oxides,136 which exhibited outstanding bifunctional electrocatalytic ability in the HER and OER with a low cell voltage of 1.58 V in 1 M KOH for overall water splitting. More impressively, they also had electrocatalytic capability in a neutral electrolyte (1 M phosphate buffer solution) with a cell voltage of 1.91 V, which is very meaningful to avoid the corrosion of devices from alkaline electrolytes and reduce the operating cost for practical applications. Besides, NiCo2O4, NiCo2S4 and NiCo2Se4 nanosheets were separately grown on conductive carbon papers to compare their behavior in electrocatalytic water splitting.137 For the HER at 10 mA cm−2 , the overpotential value for NiCo2Se4 (56 mV) was relatively lower than those of NiCo2S4 (87 mV) and NiCo2O4 (141 mV), but still slightly higher than that of Pt/C (41 mV). However, for the OER at 25 mA cm−2, NiCo2S4 possessed the lowest overpotential value of 251 mV as compared to other materials (i.e., NiCo2O4: 372 mV; NiCo2Se4: 321 mV; IrO2/C: 358 mV). Accordingly, NiCo2S4 and NiCo2Se4 were employed as the anode and cathode for overall water splitting, respectively, yielding a low cell voltage of 1.51 V for driving the generation of H2 and O2 bubbles at 10 mA cm−2. Theoretical calculation unraveled that the center position of the d band for electrodes was closely related to the electrocatalytic activity. It was found that NiCo2S4 and NiCo2Se4 had the d band center closer to the Fermi level, suggesting increased unoccupied orbitals due to the existence of S and Se, and the upward movement of the antibonding d state with strengthened adsorption interaction with reactive species (e.g., H2O and H*) for enhanced electrocatalytic performance.137


image file: d2ta00305h-f10.tif
Fig. 10 (a) Schematic diagram of the water splitting process based on NCSs. (b) Nyquist plots, and (c) oxygen reduction reaction and oxygen evolution reaction curves of different catalysts. (d) Stability tests of the NCS catalyst.133 Reproduced from ref. 133 with permission from American Chemical Society.

In addition, due to their excellent electrocatalytic activity, NCSs have also been proved to be effective electrocatalysts for the alkaline urea electrooxidation reaction in urea fuel cells, which are accepted as one promising strategy to use urea from wastewater to produce electricity and hydrogen.52,138–140 For example, NiCo2S4 nanowires were in situ grown on a carbon sponge, which showed great electrocatalytic performance with a large electrochemically active surface area of 330.7 cm−2 and a low activation energy of 15.3 kJ mol−1 towards the urea oxidation reaction.140 It is believed that the excellent electrocatalytic properties of NCSs will make them have good application prospects in other electrocatalysis fields.

4. Applications of NCSs in solar cells

NCSs are frequently used as counter electrode materials in dye-sensitized solar cells (DSCs) and quantum dot-sensitized solar cells (QDSCs).84,143–146 In DSCs and QDSCs, counter electrodes (CEs) play an important role in collecting electrons from the external circuit to electro-catalyze the reduction of the electrolyte and thus realize continuous photoelectric conversion. For DSCs, Pt is an ideal choice as the CE due to its excellent electrocatalytic ability for the reduction of iodide/triiodide (I/I3) redox electrolytes.145 However, the high cost of Pt inspires people to develop cheaper alternatives. For QDSCs, Pt is inefficient for the reduction of polysulfide redox electrolytes because of the strong adsorption of sulfur atoms onto the Pt surface to reduce its conductivity and reduction ability.143 Therefore, effective CE materials have been intensively developed, including carbon-based materials, metal sulfides and so on.

Among them, NCS-based materials are expected to be promising CE alternatives because they provide rich redox reactions and relatively high conductivity from the synergetic effect of Ni and Co atoms. For instance, NiCo2S4 quantum dots (QDs) were decorated onto the surface of nitrogen-doped carbon nanotubes (N-CNTs) as CEs for DSCs (Fig. 11a).147 It was found that such CEs showed comparable electrocatalytic performance to Pt (Fig. 11b and c; photoelectric conversion efficiency of DSCs: 7.65% for NiCo2S4@N-CNTs vs. 7.39% for Pt). This was mainly because of the uniform distribution of NiCo2S4 QDs on CNTs for abundant reaction sites, the special 1D channels of CNTs for rapid charge transport and ion diffusion, and the strong metal–nitrogen bonding between NiCo2S4 QDs and CNTs due to the nitrogen doping.


image file: d2ta00305h-f11.tif
Fig. 11 Schematic diagram of the synthesis process and TEM image for the NiCo2S4 quantum dots @ CNTs (a), and CV curves (b) and JV curves of DSCs (c) for different counter electrodes.147 Reproduced from ref. 110 with permission from American Chemical Society.

5. Applications of NCSs in photocatalysis

Recently, metal sulfides have been widely applied in photocatalysis,148–150 and NCSs are also introduced into the photocatalytic system mainly due to their great photoelectric conductivity and p-type semiconductor characteristics (Eg = ∼2.5 eV). For example, NCS@MXene composites were prepared as a photocatalyst and exhibited great photocatalytic activity and stability in the degradation of rhodamine B under visible light conditions.86 Besides, NiCo2S4/CdS (Fig. 12),87 NiCo2S4@Zn0.5Cd0.5S,151 and NiCo2S4/g-C3N4 (ref. 152) co-catalysts have been fabricated and proved to be efficient in visible-light-irradiation photocatalytic hydrogen generation because the participation of NCSs is able to significantly enhance the carrier separation and migration due to the heterojunction formation in the co-catalysts.
image file: d2ta00305h-f12.tif
Fig. 12 Light absorption plots and real pictures (a), and H2 production (c) of different photocatalysis (a), and schematic diagram of the working process (b) and cycling performance (d) for photocatalytic H2 generation by NiCo2S4/CdS (b).87 Reproduced from ref. 55 with permission from the Royal Society of Chemistry.

Moreover, NiCo2S4 micro-particles were synthesized via a solvothermal method and then applied as noble-metal-free catalysts in the CO2 photoreduction system.122 Photocatalytic CO2 reduction is considered as a promising strategy to realize the direct conversion of solar energy into special chemicals (e.g., CO) with the advantages of using renewable solar energy and alleviating the climate change from CO2 emission. Here, the NiCo2S4 photocatalyst showed considerable activity for visible-light CO2 reduction, yielding a production rate of 43.5 μmol mg−1 h−1 for CO. In situ steady-state photoluminescence and transient photocurrent tests indicated that the NiCo2S4 catalyst enabled the reduction of the recombination and the acceleration of the transport of photo-generated charges during the CO2 reduction process.122

6. Applications of NCSs in sensors

The accurate detection of glucose concentration in various sources (e.g., human blood, foods and pharmaceuticals) is essential and necessary for personal healthcare, diagnosis and therapy of diseases (e.g., diabetes mellitus), drug analysis, food monitoring and so on.153 It requires that glucose-monitoring devices should be efficient, reliable, low-cost, and simple. Enzyme-immobilized electrodes are widely used as glucose sensors due to their high sensitivity. However, such glucose sensors demand complex operating conditions for the attachment of enzymes onto the electrode surface, and the biomass enzymes are usually unstable under high temperature, humidity or acidity and alkalinity conditions.90

To address these problems, enzyme-free glucose sensors based on various inorganic materials (e.g., metal nanoparticles, metal oxides, sulfides, etc.) have been developed recently. NCSs and their composites with different structures (e.g., NiCo2S4 nanosheets/nanowires/hollow spheres,90,154,155 and NiS/CoS/NiCo2S4 micro-flowers156) are also considered as an optimal choice for non-enzymatic glucose sensors based on their good electrocatalytic activity toward glucose oxidation. Moreover, they have several advantages, such as facile synthesis, high sensitivity, great stability and low cost. For example, flower-like NiCo2S4 structures constructed from many interlaced nanosheets were simply electrodeposited onto Ni-coated cellulose filter paper (Fig. 13a),153 which showed satisfactory sensing performance for glucose detection (Fig. 13b–e), including a wide linear range (0.5 μM to 6 mM; Fig. 13d) for the glucose content, considerable sensitivity (283 μA mM−1 cm−2), a low detection limit (50 nM), great selectivity (Fig. 13e), excellent repeatability and electrochemical stability.


image file: d2ta00305h-f13.tif
Fig. 13 Schematic diagrams of the preparation process (a) and sensing process (b), and glucose sensing performance: (c) CV curves in 2 mM glucose and 0.1 M NaOH at 20 mV s−1, (d) linear properties and (e) selectivity with interfering species (AA, UA, DA, U, AP, CA, KCl, NaCl, etc.) and glucose (Glu) for NiCo2S4/Ni/cellulose filter paper.153 Reproduced from ref. 113 with permission from American Chemical Society.

Similarly, the excellent electrocatalytic capability of NCSs also endows them with great sensing performance for the detection of other substances (e.g., sulfadimethoxine (SDM)54 and pyrimethanil157). It was reported that a NiCo2S4@N/S-doped CeO2 composite was able to be an effective signal amplification label in a DNA aptasensor for detecting SDM, which is widely used for the veterinary treatment of coccidiosis and other bacterial infections but harmful to organisms even at a low concentration. This was because the composite had great electrocatalytic performance for the oxygen reduction reaction, and thus excellent electrocatalytic amplification for SDM detection with high sensitivity and selection.54 In addition, NCS-based materials (e.g., NiCo2S4/reduced graphene oxide nanosheets) also can act as gas sensors for ethanol monitoring by recording the resistance change of sensing materials. This is mainly due to the adsorption of O2 molecules on the electrode surface and the reaction between the ethanol molecules and the adsorbed O2 molecules will change the electrode resistance.158

7. Applications of NCSs in microwave absorption

With the improvement of science and technology, more and more electronic products come into our daily life, while the accompanying electromagnetic radiation and interference will affect the human health, influence the normal operation of electronic devices, and even threaten the national defense safety. Thus, a variety of microwave absorption materials have been extensively studied in order to effectively absorb and then transform electromagnetic microwaves into other types of energy (e.g., thermal energy; Fig. 14a).159 Ideal microwave absorbers should have a high absorption capacity, a wide absorption frequency range, low cost and light weight. Carbon-based materials (e.g., carbon nanotubes, graphene and porous carbon materials) have been widely used to absorb microwaves primarily due to their tunable dielectric properties, large specific surface areas, light weight and high chemical stability.160
image file: d2ta00305h-f14.tif
Fig. 14 Proposed microwave absorption mechanism of NCS-based materials (a) and comparison of microwave absorption ability among different materials (b).163 Reproduced from ref. 121 with permission from Elsevier.

Recently, NCSs have also been considered to be a potential microwave absorption material owing to their great magnetic loss performance.161,162 For example, NiCo2S4 nanosheets or microspheres were combined with carbon materials (e.g., biomass derived carbon mesoporous91,163 and hollow carbon microspheres164) to achieve excellent microwave absorption capability (Fig. 14b), such as a minimum reflection loss value of −64.74 dB and a wide absorption range from 9.22 to 14.48 GHz.163 It was found that the assistance of NiCo2S4 in the biomass-derived carbon composite was able to improve the impedance matching and the attenuation constant for strong microwave absorption.91 Therefore, it is promising to develop highly efficient and low-cost microwave absorbers based on NCS materials.

8. Conclusion and prospects

In summary, NCS-based materials have been extensively studied in recent years. They possess great performances in different applications due to their intrinsic characteristics of highly reactive activity from the multivalent states of crystals, and relatively high electronic conductivity from the cooperation of Ni and Co atoms. This review summarizes the recent development of NCSs on various aspects of micro/nano-structure design, element modulation and dopants, heterojunction building and composite integration. It emphasizes the applications of NCSs in several fields, including electrochemical energy storage devices (e.g., supercapacitors, Li/Na/Zn-ion batteries, and Li–S/Zn–air/Li–O2 batteries), water splitting, DSCs/QDSCs, photocatalysis, sensors and microwave absorption. It is indicated that these important applications are all mainly based on the attractive properties of NCSs, such as relatively high conductivity, rich active sites, and tunable components.

Although much progress has been achieved on the development of NCS-based materials for different devices, these materials are still unsatisfactory to meet the requirement for practical applications. For example, the specific capacity, energy density, and stability of the corresponding energy storage devices aren't quite high. Although the energy density of NCS-based supercapacitors is higher than that of the carbon-based ones, their power density and cycling stability are reduced concurrently. Compared to other typical batteries with high energy density, NCS-based batteries are still unsatisfactory. Besides, the high self-discharge rate (10–40%/day) of NCS-based supercapacitors should be addressed for their practical applications. Similarly, the catalytic activity and cycling stability of NCSs for the applications of electrocatalysis, photocatalysis, and sensing devices also need further enhancement. In addition, the manufacturing cost, environmental compatibility, and industrial-scale preparation should be especially concerned in future. Moreover, the relationship between the microscopic characteristics of NCS materials and the macroscopic performance should be understood more for the further rational design of NCS materials.

For the performance improvement, the exploitation and construction of highly effective NCS materials concerning various aspects (e.g., modulation of the element ratio, formation of efficient nanostructures, and integration of other components) are definitely required to expose abundant active surface sites, strength electronic conductivity, build fast channels for charge transfer and mass diffusion, and enhance the cycling durability of NCS-based systems. Besides, deep understand of the material behavior and working mechanisms of NCSs in different application systems is very important for their performance improvement. The combination of ex situ and in situ characterization techniques is very useful to disclose the underlying mechanisms (e.g., component conversion, structural evolution, reaction pathways, and interfacial interaction) of NCS-based applications.41,66,119–121,165 Besides, the optimization of mathematical modeling and theoretical simulation with higher rationality and reliability is also very helpful to promote the practical applications of NCSs.15,69,117,121,166 It is noteworthy that machine learning has been rapidly developed and widely applied in many fields, which is possibly a powerful way for the investigation of NCSs with superior performance in various applications by the purposeful design and simple, large-scale, high efficiency and low-cost preparation strategies.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

The authors gratefully acknowledge the financial support from the Shenzhen Science and Technology Program (No. JCYJ20190809161407424), the National Natural Science Foundation of China (No. 22075237), the Natural Science Foundation of Fujian Province of China (No. 2020J01007), and the Fundamental Research Funds for the Central Universities of China (No. 20720210029).

References

  1. X. Chen, Q. Liu, T. Bai, W. Wang, F. He and M. Ye, Chem. Eng. J., 2021, 409, 127237 Search PubMed.
  2. B. Yan, X. Li, W. Xiao, J. Hu, L. Zhang and X. Yang, J. Mater. Chem. A, 2020, 8, 17848–17882 Search PubMed.
  3. X. Zhao, F. Gong, Y. Zhao, B. Huang, D. Qian, H.-E. Wang, W. Zhang and Z. Yang, Chem. Eng. J., 2020, 392, 123675 Search PubMed.
  4. C. Miao, C. Zhou, H.-E. Wang, K. Zhu, K. Ye, Q. Wang, J. Yan, D. Cao, N. Li and G. Wang, J. Power Sources, 2021, 490, 229532 Search PubMed.
  5. J. Ren, M. Shen, Z. Li, C. Yang, Y. Liang, H.-E. Wang, J. Li, N. Li and D. Qian, J. Power Sources, 2021, 501, 230003 Search PubMed.
  6. J. Zhao, G. Wang, K. Cheng, K. Ye, K. Zhu, J. Yan, D. Cao and H.-E. Wang, J. Power Sources, 2020, 451, 227737 Search PubMed.
  7. T. Zhu, Z. He, Y. Ren, W. Zeng, J. Mao and L. Zhu, Sol. RRL, 2021, 5, 2100021 Search PubMed.
  8. G. Fu and J.-M. Lee, J. Mater. Chem. A, 2019, 7, 9386–9405 Search PubMed.
  9. P. Kulkarni, S. K. Nataraj, R. G. Balakrishna, D. H. Nagaraju and M. V. Reddy, J. Mater. Chem. A, 2017, 5, 22040–22094 Search PubMed.
  10. T. Chen, S. Wei and Z. Wang, ChemPlusChem, 2020, 85, 43–56 Search PubMed.
  11. T.-F. Yi, J.-J. Pan, T.-T. Wei, Y. Li and G. Cao, Nano Today, 2020, 33, 100894 Search PubMed.
  12. X. Xu, H. Liang, F. Ming, Z. Qi, Y. Xie and Z. Wang, ACS Catal., 2017, 7, 6394–6399 Search PubMed.
  13. X. Wang, X. Xu, J. Chen and Q. Wang, ACS Sustainable Chem. Eng., 2019, 7, 12331–12339 Search PubMed.
  14. S. Hyun and S. Shanmugam, ACS Omega, 2018, 3, 8621–8630 Search PubMed.
  15. H. Zhang, J. Liu, X. Lin, T. Han, M. Cheng, J. Long and J. Li, J. Alloys Compd., 2020, 817, 153293 Search PubMed.
  16. Q. Hu, W. Ma, G. Liang, H. Nan, X. Zheng and X. Zhang, RSC Adv., 2015, 5, 84974–84979 Search PubMed.
  17. K. Guo, Y. Ding, J. Luo and Z. Yu, ACS Appl. Mater. Interfaces, 2018, 10, 19673–19681 Search PubMed.
  18. X. Wu, S. Li, B. Wang, J. Liu and M. Yu, Chem. Commun., 2021, 57, 1002–1005 Search PubMed.
  19. L. Kang, M. Zhang, J. Zhang, S. Liu, N. Zhang, W. Yao, Y. Ye, C. Luo, Z. Gong, C. Wang, X. Zhou, X. Wu and S. C. Jun, J. Mater. Chem. A, 2020, 8, 24053–24064 Search PubMed.
  20. X. Wang, R. Zhou, C. Zhang, S. Xi, M. W. M. Jones, T. Tesfamichael, A. Du, K. Gui, K. Ostrikov and H. Wang, J. Mater. Chem. A, 2020, 8, 9278–9291 Search PubMed.
  21. J. Wu, X. Shi, W. Song, H. Ren, C. Tan, S. Tang and X. Meng, Nano Energy, 2018, 45, 439–447 Search PubMed.
  22. V. Ganesan, P. Ramasamy and J. Kim, Int. J. Hydrogen Energy, 2017, 42, 5985–5992 Search PubMed.
  23. W. Wang, J. Li, Q. Jin, Y. Liu, Y. Zhang, Y. Zhao, X. Wang, A. Nurpeissova and Z. Bakenov, ACS Appl. Energy Mater., 2021, 4, 1687–1695 Search PubMed.
  24. F. Lu, M. Zhou, K. Su, T. Ye, Y. Yang, T. D. Lam, Y. Bando and X. Wang, ACS Appl. Mater. Interfaces, 2019, 11, 2082–2092 Search PubMed.
  25. L. Hou, H. Hua, R. Bao, Z. Chen, C. Yang, S. Zhu, G. Pang, L. Tong, C. Yuan and X. Zhang, ChemPlusChem, 2016, 81, 557–563 Search PubMed.
  26. L. Hou, R. Bao, Z. Chen, M. Rehan, L. Tong, G. Pang and C. Yuan, Electrochim. Acta, 2016, 214, 76–84 Search PubMed.
  27. T. Zhu, J. Wang and G. W. Ho, Nano Energy, 2015, 18, 273–282 Search PubMed.
  28. Z. Chen, X. Wang, H. Wang, H. Wang, T. Bai, F. Ren, P. Ren, H. Yan, K. Xiao and Z. Shen, J. Alloys Compd., 2022, 891, 161988 Search PubMed.
  29. J. Lin, Y. Wang, X. Zheng, H. Liang, H. Jia, J. Qi, J. Cao, J. Tu, W. Fei and J. Feng, Dalton Trans., 2018, 47, 8771–8778 Search PubMed.
  30. Y. Liu, Y. Wen, Y. Zhang, X. Wu, H. Li, H. Chen, J. Huang, G. Liu and S. Peng, Sci. China Mater., 2020, 63, 1216–1226 Search PubMed.
  31. G. Xiang, Y. Meng, G. Qu, J. Yin, B. Teng, Q. Wei and X. Xu, Sci. Bull., 2020, 65, 443–451 Search PubMed.
  32. Y. Chen, T. Liu, L. Zhang and J. Yu, ACS Sustainable Chem. Eng., 2019, 7, 11157–11165 Search PubMed.
  33. M. Chuai, K. Zhang, X. Chen, Y. Tong, H. Zhang and M. Zhang, Chem. Eng. J., 2020, 381, 122682 Search PubMed.
  34. C. Han, T. Zhang, J. Li, B. Li and Z. Lin, Nano Energy, 2020, 77, 105165 Search PubMed.
  35. B. Liu, S. Huang, D. Kong, J. Hu and H. Y. Yang, J. Mater. Chem. A, 2019, 7, 7604–7613 Search PubMed.
  36. N. Kurra, C. Xia, M. N. Hedhili and H. N. Alshareef, Chem. Commun., 2015, 51, 10494–10497 Search PubMed.
  37. T. Wang, B. Zhao, H. Jiang, H.-P. Yang, K. Zhang, M. M. F. Yuen, X.-Z. Fu, R. Sun and C.-P. Wong, J. Mater. Chem. A, 2015, 3, 23035–23041 Search PubMed.
  38. X. Zhang, Y. Zheng, J. Zhou, W. Zheng and D. Chen, RSC Adv., 2017, 7, 13406–13415 Search PubMed.
  39. H. Chen, J. Jiang, Y. Zhao, L. Zhang, D. Guo and D. Xia, J. Mater. Chem. A, 2015, 3, 428–437 Search PubMed.
  40. D. Cao, W. Kang, S. Wang, Y. Wang, K. Sun, L. Yang, X. Zhou, D. Sun and Y. Cao, J. Mater. Chem. A, 2019, 7, 8268–8276 Search PubMed.
  41. S. Li, C. Li, W. K. Pang, Z. Zhao, J. Zhang, Z. Liu and D. Li, ACS Appl. Mater. Interfaces, 2019, 11, 27805–27812 Search PubMed.
  42. Y. Tang, S. Chen, S. Mu, T. Chen, Y. Qiao, S. Yu and F. Gao, ACS Appl. Mater. Interfaces, 2016, 8, 9721–9732 Search PubMed.
  43. N. Zhao, H. Fan, J. Ma, M. Zhang, C. Wang, H. Li, X. Jiang and X. Cao, J. Power Sources, 2019, 439, 227097 Search PubMed.
  44. G. Nagaraju, S. C. Sekhar, B. Ramulu and J. S. Yu, Small, 2019, 15, 1805418 Search PubMed.
  45. Y. Wen, Y. Liu, T. Wang, Z. Wang, Y. Zhang, X. Wu, X. Chen, S. Peng and D. He, ACS Appl. Energy Mater., 2021, 4, 6531–6541 Search PubMed.
  46. X. Li, Q. Li, Y. Wu, M. Rui and H. Zeng, ACS Appl. Mater. Interfaces, 2015, 7, 19316–19323 Search PubMed.
  47. Y. Xu, A. Sumboja, Y. Zong and J. A. Darr, Catal. Sci. Technol., 2020, 10, 2173–2182 Search PubMed.
  48. B. Guo, T. Yang, W. Du, Q. Ma, L.-z. Zhang, S.-J. Bao, X. Li, Y. Chen and M. Xu, J. Mater. Chem. A, 2019, 7, 12276–12282 Search PubMed.
  49. L. Hou, Y. Shi, S. Zhu, M. Rehan, G. Pang, X. Zhang and C. Yuan, J. Mater. Chem. A, 2017, 5, 133–144 Search PubMed.
  50. Y. Jiang, X. Qian, C. Zhu, H. Liu and L. Hou, ACS Appl. Mater. Interfaces, 2018, 10, 9379–9389 Search PubMed.
  51. P. Liu, Y. Liu, J. Li, M. Wang and H. Cui, Nanoscale, 2020, 12, 22330–22339 Search PubMed.
  52. W. Song, M. Xu, X. Teng, Y. Niu, S. Gong, X. Liu, X. He and Z. Chen, Nanoscale, 2021, 13, 1680–1688 Search PubMed.
  53. Z. Hao, X. He, H. Li, D. Trefilov, Y. Song, Y. Li, X. Fu, Y. Cui, S. Tang, H. Ge and Y. Chen, ACS Nano, 2020, 14, 12719–12731 Search PubMed.
  54. L. Li, L. Yang, S. Zhang, Y. Sun, F. Li, T. Qin, X. Liu, Y. Zhou and S. Alwarappan, J. Mater. Chem. C, 2020, 8, 14723–14731 Search PubMed.
  55. C. Wang, Z. Guan, Y. Shen, S. Yu, X.-Z. Fu, R. Sun and C.-P. Wong, Chem. Eng. J., 2018, 346, 193–202 Search PubMed.
  56. Y. Ouyang, H. Ye, X. Xia, X. Jiao, G. Li, S. Mutahir, L. Wang, D. Mandler, W. Lei and Q. Hao, J. Mater. Chem. A, 2019, 7, 3228–3237 Search PubMed.
  57. J. Zhu, S. Tang, J. Wu, X. Shi, B. Zhu and X. Meng, Adv. Energy Mater., 2017, 7, 1601234 Search PubMed.
  58. Q. Liu, X. Hong, X. You, X. Zhang, X. Zhao, X. Chen, M. Ye and X. Liu, Energy Storage Materials, 2020, 24, 541–549 Search PubMed.
  59. X. Zhang, X. Chen, T. Bai, J. Chai, X. Zhao, M. Ye, Z. Lin and X. Liu, J. Mater. Chem. A, 2020, 8, 11589–11597 Search PubMed.
  60. X.-Z. Song, F.-F. Sun, Y.-L. Meng, Z.-W. Wang, Q.-F. Su and Z. Tan, New J. Chem., 2019, 43, 3601–3608 Search PubMed.
  61. H. Zhang, A. Hao, Z. Sun, X. Ning, J. Guo, Y. Lv and D. Jia, J. Alloys Compd., 2020, 847, 156505 Search PubMed.
  62. C. Zhang, M. Hou, X. Cai, J. Lin, X. Liu, R. Wang, L. Zhou, J. Gao, B. Li and L. Lai, J. Mater. Chem. A, 2018, 6, 15630–15639 Search PubMed.
  63. J. Shen, J. Wu, L. Pei, M.-T. F. Rodrigues, Z. Zhang, F. Zhang, X. Zhang, P. M. Ajayan and M. Ye, Adv. Energy Mater., 2016, 6, 1600341 Search PubMed.
  64. X. Chang, W. Li, Y. Liu, M. He, X. Zheng, J. Bai and Z. Ren, J. Colloid Interface Sci., 2018, 538, 34–44 Search PubMed.
  65. L. Wan, C. He, D. Chen, J. Liu, Y. Zhang, C. Du, M. Xie and J. Chen, Chem. Eng. J., 2020, 399, 125778 Search PubMed.
  66. S. Fan, H. Liu, S. Bi, C. Gao, X. Meng and Y. Wang, Electrochim. Acta, 2021, 388, 138618 Search PubMed.
  67. S.-x. Yan, S.-h. Luo, J. Feng, P.-w. Li, R. Guo, Q. Wang, Y.-h. Zhang, Y.-g. Liu and S. Bao, Chem. Eng. J., 2020, 381, 122659 Search PubMed.
  68. H. S. Lee, J. Pan, G. S. Gund and H. S. Park, Adv. Mater. Interfaces, 2020, 7, 2000138 Search PubMed.
  69. W. Liu, B. Ren, W. Zhang, M. Zhang, G. Li, M. Xiao, J. Zhu, A. Yu, L. Ricardez-Sandoval and Z. Chen, Small, 2019, 15, 1903610 Search PubMed.
  70. Y. Zhu, J. Li, X. Yun, G. Zhao, P. Ge, G. Zou, Y. Liu, H. Hou and X. Ji, Nano-Micro Lett., 2020, 12, 16 Search PubMed.
  71. Y. Liu, G. Jiang, Z. Huang, Q. Lu, B. Yu, U. Evariste and P. Ma, ACS Appl. Energy Mater., 2019, 2, 8079–8089 Search PubMed.
  72. X. Zhao, Q. Ma, K. Tao and L. Han, ACS Appl. Energy Mater., 2021, 4, 4199–4207 Search PubMed.
  73. X. He, Q. Liu, J. Liu, R. Li, H. Zhang, R. Chen and J. Wang, Chem. Eng. J., 2017, 325, 134–143 Search PubMed.
  74. S. Chen, C. Lu, L. Liu, M. Xu, J. Wang, Q. Deng, Z. Zeng and S. Deng, Nanoscale, 2020, 12, 1852–1863 Search PubMed.
  75. Y. Zhu, S. An, X. Sun, D. Lan, J. Cui, Y. Zhang and W. He, Chem. Eng. J., 2020, 383, 123206 Search PubMed.
  76. 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 Search PubMed.
  77. J. Fu, L. Li, J. M. Yun, D. Lee, B. K. Ryu and K. H. Kim, Chem. Eng. J., 2019, 375, 121939 Search PubMed.
  78. J.-Z. He, W.-J. Niu, Y.-P. Wang, Q.-Q. Sun, M.-J. Liu, K. Wang, W.-W. Liu, M.-C. Liu, F.-C. Yu and Y.-L. Chueh, Electrochim. Acta, 2020, 362, 136968 Search PubMed.
  79. Z. Peng, C. Yang, Y. Hu, F. Bai, W. Chen, R. Liu, S. Jiang and H.-C. Chen, Appl. Surf. Sci., 2022, 573, 151561 Search PubMed.
  80. M. Zhang, H. Zheng, H. Zhu, M. Zhang, R. Liu, X. Zhu, X. Li and H. Cui, J. Alloys Compd., 2022, 901, 163633 Search PubMed.
  81. G. Shao, R. Yu, X. Zhang, X. Chen, F. He, X. Zhao, N. Chen, M. Ye and X. Y. Liu, Adv. Funct. Mater., 2020, 30, 2003153 Search PubMed.
  82. Y. Miao, X. Zhao, X. Wang, C. Ma, L. Cheng, G. Chen, H. Yue, L. Wang and D. Zhang, Nano Res., 2020, 13, 3041–3047 Search PubMed.
  83. A. Hu, J. Long, C. Shu, C. Xu, T. Yang, R. Liang and J. Li, ChemElectroChem, 2019, 6, 349–358 Search PubMed.
  84. S. A. Ansari, S. Goumri-Said, H. M. Yadav, M. Belarbi, A. Aljaafari and M. B. Kanoun, Sol. Energy Mater. Sol. Cells, 2021, 225, 111064 Search PubMed.
  85. H.-J. Kim and V. T.-V. Chebrolu, New J. Chem., 2018, 42, 18824–18836 Search PubMed.
  86. S. Vigneshwaran, C. M. Park and S. Meenakshi, Sep. Purif. Technol., 2021, 258, 118003 Search PubMed.
  87. J. Peng, J. Xu, Z. Wang, Z. Ding and S. Wang, Phys. Chem. Chem. Phys., 2017, 19, 25919–25926 Search PubMed.
  88. J. Yu, C. Lv, L. Zhao, L. Zhang, Z. Wang and Q. Liu, Adv. Mater. Interfaces, 2018, 5, 1701396 Search PubMed.
  89. 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 Search PubMed.
  90. P. K. Kannan, C. Hu, H. Morgan and C. S. Rout, Chem.–Asian J., 2016, 11, 1837–1841 Search PubMed.
  91. P. Hu, S. Dong, X. Li, J. Chen and P. Hu, ACS Sustainable Chem. Eng., 2020, 8, 10230–10241 Search PubMed.
  92. Y. Xu, X. Gao, W. Chu, Q. Li, T. Li, C. Liang and Z. Lin, J. Mater. Chem. A, 2016, 4, 10248–10253 Search PubMed.
  93. Y. Xiao, J. Huang, Y. Xu, H. Zhu, K. Yuan and Y. Chen, J. Mater. Chem. A, 2018, 6, 9161–9171 Search PubMed.
  94. L. Hou, R. Bao, M. Rehan, L. Tong, G. Pang, X. Zhang and C. Yuan, Adv. Electron. Mater., 2017, 3, 1600322 Search PubMed.
  95. P. Phonsuksawang, P. Khajondetchairit, T. Butburee, S. Sattayaporn, N. Chanlek, P. Hirunsit, S. Suthirakun and T. Siritanon, Electrochim. Acta, 2020, 340, 135939 Search PubMed.
  96. P. Phonsuksawang, P. Khajondetchairit, K. Ngamchuea, T. Butburee, S. Sattayaporn, N. Chanlek, S. Suthirakun and T. Siritanon, Electrochim. Acta, 2021, 368, 137634 Search PubMed.
  97. S. Abureden, F. M. Hassan, G. Lui, S. Sy, R. Batmaz, W. Ahn, A. Yu and Z. Chen, J. Mater. Chem. A, 2017, 5, 7523–7532 Search PubMed.
  98. T. Peng, H. Yi, P. Sun, Y. Jing, R. Wang, H. Wang and X. Wang, J. Mater. Chem. A, 2016, 4, 8888–8897 Search PubMed.
  99. M. Zhang, H. Liu, Z. Song, T. Ma and J. Xie, Chem. Eng. J., 2020, 392, 123669 Search PubMed.
  100. M. Liang, M. Zhao, H. Wang, J. Shen and X. Song, J. Mater. Chem. A, 2018, 6, 2482–2493 Search PubMed.
  101. Y. Cui, J. Zhang, C. Jin, Y. Liu, W. Luo and W. Zheng, Small, 2018, 15, e1804318 Search PubMed.
  102. X. Chen, D. Chen, X. Guo, R. Wang and H. Zhang, ACS Appl. Mater. Interfaces, 2017, 9, 18774–18781 Search PubMed.
  103. L. Lin, J. Liu, T. Liu, J. Hao, K. Ji, R. Sun, W. Zeng and Z. Wang, J. Mater. Chem. A, 2015, 3, 17652–17658 Search PubMed.
  104. C. Liu, X. Wu and B. Wang, Chem. Eng. J., 2020, 392, 123651 Search PubMed.
  105. M. L. Aparna, T. Thomas and G. R. Rao, J. Electrochem. Soc., 2022, 169, 020515 Search PubMed.
  106. L. Yang, X. Lu, S. Wang, J. Wang, X. Guan, X. Guan and G. Wang, Nanoscale, 2020, 12, 1921–1938 Search PubMed.
  107. Y. Yan, A. Li, C. Lu, T. Zhai, S. Lu, W. Li and W. Zhou, Chem. Eng. J., 2020, 396, 125316 Search PubMed.
  108. X. Song, C. Huang, Y. Qin, H. Li and H. C. Chen, J. Mater. Chem. A, 2018, 6, 16205–16212 Search PubMed.
  109. W. Chen, X. Zhang, L.-E. Mo, Y. Zhang, S. Chen, X. Zhang and L. Hu, Chem. Eng. J., 2020, 388, 124109 Search PubMed.
  110. X. Chen, C. He, W. Wang, T. Bai, G. Xue and M. Ye, Phys. Rev. Appl., 2021, 15, 064042 Search PubMed.
  111. W. He, C. Wang, H. Li, X. Deng, X. Xu and T. Zhai, Adv. Energy Mater., 2017, 7, 1700983 Search PubMed.
  112. Y. Sui, Y. Zhang, H. Hu, Q. Xu, F. Yang and Z. Li, Adv. Mater. Interfaces, 2018, 5, 1800018 Search PubMed.
  113. C. Zhang, X. Cai, Y. Qian, H. Jiang, L. Zhou, B. Li, L. Lai, Z. Shen and W. Huang, Adv. Sci., 2018, 5, 1700375 Search PubMed.
  114. Z. Peng, L. Gong, J. Huang, Y. Wang, L. Tan and Y. Chen, Carbon, 2019, 153, 531–538 Search PubMed.
  115. Q. Li, W. Lu, Z. Li, J. Ning, Y. Zhong and Y. Hu, Chem. Eng. J., 2020, 380, 122544 Search PubMed.
  116. J. Huang, J. Wei, Y. Xu, Y. Xiao and Y. Chen, J. Mater. Chem. A, 2017, 5, 23349–23360 Search PubMed.
  117. T. Sun, C. Huang, H. Shu, L. Luo, Q. Liang, M. Chen, J. Su and X. Wang, ACS Appl. Mater. Interfaces, 2020, 12, 57975–57986 Search PubMed.
  118. Y. Tang, X. Li, H. Lv, D. Xie, W. Wang, C. Zhi and H. Li, Adv. Energy Mater., 2020, 10, 2000892 Search PubMed.
  119. J. Zhang, K. Song, L. Mi, C. Liu, X. Feng, J. Zhang, W. Chen and C. Shen, J. Phys. Chem. Lett., 2020, 11, 1435–1442 Search PubMed.
  120. S. Li, P. Ge, F. Jiang, H. Shuai, W. Xu, Y. Jiang, Y. Zhang, J. Hu, H. Hou and X. Ji, Energy Storage Materials, 2019, 16, 267–280 Search PubMed.
  121. Y. V. Lim, S. Huang, J. Hu, D. Kong, Y. Wang, T. Xu, L. K. Ang and H. Y. Yang, Small Methods, 2019, 3, 1900112 Search PubMed.
  122. Z. Xiong, L. Huang, J. Peng, Y. Hou, Z. Ding and S. Wang, ChemCatChem, 2019, 11, 5513–5518 Search PubMed.
  123. B. He, J.-J. Song, X.-Y. Li, C.-Y. Xu, Y.-B. Li, Y.-W. Tang, Q.-L. Hao, H.-K. Liu and Z. Su, Nanoscale, 2021, 13, 810–818 Search PubMed.
  124. F. Wang, G. Li, X. Meng, S. Xu and W. Ma, J. Power Sources, 2020, 462, 228162 Search PubMed.
  125. W. Liu, J. Zhang, Z. Bai, G. Jiang, M. Li, K. Feng, L. Yang, Y. Ding, T. Yu, Z. Chen and A. Yu, Adv. Funct. Mater., 2018, 28, 1706675 Search PubMed.
  126. X. Zuo, Y. Song and M. Zhen, Appl. Surf. Sci., 2020, 500, 144000 Search PubMed.
  127. F. Yang, Y. Shen, Z. Cen, J. Wan, S. Li, G. He, J. Hu and K. Xu, Sci. China Mater., 2021, 65, 356–363 Search PubMed.
  128. S. Li, P. Xu, M. K. Aslam, C. Chen, A. Rashid, G. Wang, L. Zhang and B. Mao, Energy Storage Materials, 2020, 27, 51–60 Search PubMed.
  129. J. Zhang, X. Bai, T. Wang, W. Xiao, P. Xi, J. Wang, D. Gao and J. Wang, Nano-Micro Lett., 2019, 11, 2 Search PubMed.
  130. S. Hyun, B. Son, H. Kim, J. Sanetuntikul and S. Shanmugam, Appl. Catal., B, 2020, 263, 118283 Search PubMed.
  131. X. Zhao, H. Liu, Y. Rao, X. Li, J. Wang, G. Xia and M. Wu, ACS Sustainable Chem. Eng., 2019, 7, 2610–2618 Search PubMed.
  132. X. Du, X. Zhang, Y. Li and M. Zhao, Int. J. Hydrogen Energy, 2018, 43, 19955–19964 Search PubMed.
  133. X. Feng, Q. Jiao, H. Cui, M. Yin, Q. Li, Y. Zhao, H. Li, W. Zhou and C. Feng, ACS Appl. Mater. Interfaces, 2018, 10, 29521–29531 Search PubMed.
  134. Y. Gong, Y. Lin, Z. Yang, J. Wang, H. Pan, Z. Xu and Y. Liu, ChemistrySelect, 2019, 4, 1180–1187 Search PubMed.
  135. F. Li, R. Xu, Y. Li, F. Liang, D. Zhang, W.-F. Fu and X.-J. Lv, Carbon, 2019, 145, 521–528 Search PubMed.
  136. H. Li, L. Chen, P. Jin, Y. Li, J. Pang, J. Hou, S. Peng, G. Wang and Y. Shi, Nano Res., 2021, 15, 950–958 Search PubMed.
  137. J. Zhou, Y. Dou, T. He, A. Zhou, X.-J. Kong, X.-Q. Wu, T. Liu and J.-R. Li, Nano Res., 2021, 14, 4548–4555 Search PubMed.
  138. L. Sha, K. Ye, G. Wang, J. Shao, K. Zhu, K. Cheng, J. Yan, G. Wang and D. Cao, Chem. Eng. J., 2019, 359, 1652–1658 Search PubMed.
  139. D. Khalafallah, Q. Zou, M. Zhi and Z. Hong, Electrochim. Acta, 2020, 350, 136399 Search PubMed.
  140. B. Li, C. Song, J. Rong, J. Zhao, H.-E. Wang, P. Yang, K. Ye, K. Cheng, K. Zhu, J. Yan, D. Cao and G. Wang, J. Energy Chem., 2020, 50, 195–205 Search PubMed.
  141. X. Du, W. Lian and X. Zhang, Int. J. Hydrogen Energy, 2018, 43, 20627–20635 Search PubMed.
  142. X.-X. Li, X.-T. Wang, K. Xiao, T. Ouyang, N. Li and Z.-Q. Liu, J. Power Sources, 2018, 402, 116–123 Search PubMed.
  143. S. S. Rao, I. K. Durga, N. Kundakarla, D. Punnoose, C. V. V. M. Gopi, A. E. Reddy, M. Jagadeesh and H.-J. Kim, New J. Chem., 2017, 41, 10037–10047 Search PubMed.
  144. L. Li, X. Zhang, S. a. Liu, B. Liang, Y. Zhang and W. Zhang, Sol. Energy, 2020, 202, 358–364 Search PubMed.
  145. K. S. Anuratha, S. Mohan and S. K. Panda, New J. Chem., 2015, 40, 1785–1791 Search PubMed.
  146. J. Qiu, D. He, H. Wang, W. Li, B. Sun, Y. Ma, X. Lu and C. Wang, Electrochim. Acta, 2021, 367, 137451 Search PubMed.
  147. P. Su, Q. Jiao, H. Li, Y. Li, X. Liu, Q. Wu, D. Shi, Y. Zhao, T. Wang and W. Wang, ACS Appl. Energy Mater., 2021, 4, 4344–4354 Search PubMed.
  148. L. Nie and Q. Zhang, Inorg. Chem. Front., 2017, 4, 1953–1962 Search PubMed.
  149. X.-Q. Qiao, Z.-W. Zhang, Q.-H. Li, D. Hou, Q. Zhang, J. Zhang, D.-S. Li, P. Feng and X. Bu, J. Mater. Chem. A, 2018, 6, 22580–22589 Search PubMed.
  150. X.-Q. Qiao, Z.-W. Zhang, F.-Y. Tian, D.-F. Hou, Z.-F. Tian, D.-S. Li and Q. Zhang, Cryst. Growth Des., 2017, 17, 3538–3547 Search PubMed.
  151. S. Zhao, J. Xu, M. Mao, L. Li and X. Li, Appl. Surf. Sci., 2020, 528, 147016 Search PubMed.
  152. K. Jiang, W. Iqbal, B. Yang, M. Rauf, I. Ali, X. Lu and Y. Mao, J. Alloys Compd., 2021, 853, 157284 Search PubMed.
  153. K. J. Babu, T. R. Kumar, D. J. Yoo, S.-M. Phang and G. G. Kumar, ACS Sustainable Chem. Eng., 2018, 6, 16982–16989 Search PubMed.
  154. Q. Guo, T. Wu, L. Liu, Y. He, D. Liu and T. You, J. Alloys Compd., 2020, 819, 153376 Search PubMed.
  155. D. Chen, H. Wang and M. Yang, Anal. Methods, 2017, 9, 4718–4725 Search PubMed.
  156. D. Li, X. Zhang, L. Pei, C. Dong, J. Shi and Y. Xu, Inorg. Chem. Commun., 2019, 110, 107581 Search PubMed.
  157. Y. He, T. Wu, J. Wang, J. Ye, C. Xu, F. Li and Q. Guo, Talanta, 2020, 219, 121277 Search PubMed.
  158. B. Li, J. Xia, J. Liu, Q. Liu, G. Huang, H. Zhang, X. Jing, R. Li and J. Wang, Chem. Phys. Lett., 2018, 703, 80–85 Search PubMed.
  159. M. Zhou, W. Gu, G. Wang, J. Zheng, C. Pei, F. Fan and G. Ji, J. Mater. Chem. A, 2020, 8, 24267–24283 Search PubMed.
  160. M. Green and X. Chen, J. Materiomics, 2019, 5, 503–541 Search PubMed.
  161. R. Peymanfar and S. Ghorbanian-Gezaforodi, Nanotechnology, 2020, 31, 495202 Search PubMed.
  162. J. Liu, Z. Yang, L. Yang, Y. Zhu, T. Xue and G. Xu, J. Alloys Compd., 2021, 853, 157403 Search PubMed.
  163. S. Dong, P. Hu, X. Li, C. Hong, X. Zhang and J. Han, Chem. Eng. J., 2020, 398, 125588 Search PubMed.
  164. P. Hu, S. Dong, F. Yuan, X. Li and C. Hong, Adv. Compos. Hybrid Mater., 2021 DOI:10.1007/s42114-021-00318-w.
  165. Z. Sun, C. Zhao, X. Cao, K. Zeng, Z. Ma, Y. Hu, J.-H. Tian and R. Yang, Electrochim. Acta, 2020, 338, 135900 Search PubMed.
  166. L. Xing, X. Zheng, M. Schroeder, J. Alvarado, A. von Wald Cresce, K. Xu, Q. Li and W. Li, Acc. Chem. Res., 2018, 51, 282–289 Search PubMed.

This journal is © The Royal Society of Chemistry 2022