Transition metal-based electrocatalysts for alkaline overall water splitting: advancements, challenges, and perspectives

Muhammad Nazim Lakhan a, Abdul Hanan b, Altaf Hussain cd, Irfan Ali Soomro e, Yuan Wang *f, Mukhtiar Ahmed g, Umair Aftab *h, Hongyu Sun i and Hamidreza Arandiyan *jk
aApplied Chemistry and Environmental Science, School of Science, STEM College, RMIT University, Melbourne, Australia
bSunway Center for Electrochemical Energy and Sustainable Technology, SCEEST, Sunway University, Bandar Sunway, Malaysia
cState Key Laboratory of Electroanalytical Chemistry, Changchun Institute of Applied Chemistry, Chinese Academy of Sciences, P. R. China
dUniversity of Science and Technology of China, Hefei, P. R. China
eInstitute of Computational Chemistry, College of Chemistry, Beijing University of Chemical Technology, P. R. China
fDepartment of Chemical Engineering, The University of Melbourne, Melbourne, VIC 3010, Australia. E-mail: helena.wang@unimelb.edu.au
gState Key Laboratory of Multiphase Complex Systems, Institute of Process Engineering, Chinese Academy of Sciences, University of Chinese Academy of Sciences, Beijing, P. R. China
hDepartment of Metallurgy and Materials Engineering, Mehran University of Engineering and Technology, Jamshoro, Pakistan. E-mail: umair.aftab@faculty.muet.edu.pk
iSchool of Resources and Materials, Northeastern University at Qinhuangdao, 066004 Qinhuangdao, P. R. China
jCentre for Advanced Materials and Industrial Chemistry (CAMIC), School of Science, RMIT University, Melbourne, VIC 3000, Australia. E-mail: hamid.arandiyan@rmit.edu.au
kLaboratory of Advanced Catalysis for Sustainability, School of Chemistry, University of Sydney, Sydney, NSW 2006, Australia

Received 10th December 2023 , Accepted 26th March 2024

First published on 4th April 2024


Abstract

Water electrolysis is a promising method for efficiently producing hydrogen and oxygen, crucial for renewable energy conversion and fuel cell technologies. The hydrogen evolution reaction (HER) and oxygen evolution reaction (OER) are two key electrocatalytic reactions occurring during water splitting, necessitating the development of active, stable, and low-cost electrocatalysts. Transition metal (TM)-based electrocatalysts, spanning noble metals and TM oxides, phosphides, nitrides, carbides, borides, chalcogenides, and dichalcogenides, have garnered significant attention due to their outstanding characteristics, including high electronic conductivity, tunable valence electron configuration, high stability, and cost-effectiveness. This timely review discusses developments in TM-based electrocatalysts for the HER and OER in alkaline media in the last 10 years, revealing that the exposure of more accessible surface-active sites, specific electronic effects, and string effects are essential for the development of efficient electrocatalysts towards electrochemical water splitting application. This comprehensive review serves as a guide for designing and constructing state-of-the-art, high-performance bifunctional electrocatalysts based on TMs, particularly for applications in water splitting.


image file: d3cc06015b-p1.tif

Muhammad Nazim Lakhan

Muhammad Nazim Lakhan received his MS degree in Chemical Engineering and Technology from Harbin Engineering University, China. Currently, he is a PhD student at RMIT University, Melbourne. His research focuses on nanomaterials, metal oxides, and 2D materials for electrochemical water-splitting reactions, including the HER and OER, in an alkaline medium.

image file: d3cc06015b-p2.tif

Abdul Hanan

Abdul Hanan has recently completed his Master of Engineering in Materials Science and Engineering from Harbin Engineering University under the supervision of Prof. Dianxue Cao. Currently, he is doing a PhD at Sunway University, Malaysia. His current research focuses on energy materials and nanomaterials for energy conversion via electrochemical water-splitting systems.

image file: d3cc06015b-p3.tif

Yuan Wang

Yuan Wang (Helena) is a Senior Lecturer, Group Leader of Renewable Resources & Sustainability (R2S) and an ARC-DECRA Fellow at the Department of Chemical Engineering at the University of Melbourne. She completed her PhD in Chemical Engineering at the University of New South Wales (UNSW) in 2018. She held prestigious fellowships: the Alfred Deakin Research Fellowship-2022 at Deakin University and the International Hydrogen Research Fellowship-2023 at the National University of Singapore. She was also a DAAD Visiting Scholar at the Fritz Haber Institute of Max Planck-Berlin (2018). Her research focuses on green hydrogen production, carbon dioxide conversion and utilization, metal recovery and recycling, and circular economy.

image file: d3cc06015b-p4.tif

Umair Aftab

Umair Aftab has completed his PhD in Metallurgy and Materials Engineering at the Mehran University of Engineering and Technology, Jamshoro. Currently, he is an Assistant Professor in the same department. His research mainly focusses on electrochemical water splitting systems: HER, OER, and ORR.

image file: d3cc06015b-p5.tif

Hongyu Sun

Hongyu Sun received his PhD degree from the State Key Laboratory of Metastable Materials Science and Technology, Yanshan University, in 2010. He then worked in the Department of Materials Science and Engineering (2010–2012), Beijing National Centre for Electron Microscopy (2012–2015), at Tsinghua University (with Prof. Jing Zhu), and Department of Micro- and Nanotechnology at the Technical University of Denmark (2015–2018, with Prof. Kristian Mølhave). Now, he is a senior application scientist at DENS solution B.V. He received the Robert P. Apkarian Award (Physical Sciences, 2018) from the Microscopy Society of America. His research interests include controllable synthesis of functional structures for energy storage and conversion, liquid cell transmission electron microscopy, and microfabrication.

image file: d3cc06015b-p6.tif

Hamidreza Arandiyan

Hamidreza Arandiyan is a leader of Critical Minerals for Clean Energy Group and has an academic tenure in Applied Chemistry and Environmental Science at RMIT University. He completed his PhD from the School of Environment at Tsinghua University in 2014. He received a Vice-Chancellor's Research Fellowship from the University of New South Wales at the School of Chemical Engineering in 2015. He held a University of Sydney Senior VC Fellowship in the School of Chemistry in 2018. He is a Fellow of the Royal Society of Chemistry (FRSC). His research focuses on resource recovery for environmental remediation and energy applications.


1. Introduction

As energy demand continues to grow, advances in efficient conversion systems have become the focus of the global research and development community.1 In recent years, there has been a sharp rise in global energy demand, with projections indicating a 56% increase between 2010 and 2040.2 The development of modern society needs the maximum use of fossil fuels, which produce harmful greenhouse gases such as carbon dioxide (CO2) and carbon monoxide (CO) in the environment.3 Therefore, the search for alternative energy sources has attracted much attention from researchers.4 Renewable fuels are considered an attractive option for resolving the global environmental and energy crises due to their merits of renewability, greenhouse gas-free emission, and high energy density.5–7 Among renewable energy conversion technologies, electrolysis is a cutting-edge technology that has gained popularity recently and is regarded as an effective method for splitting water into hydrogen and oxygen.8 The effectiveness of electrochemical water splitting (EWS) depends on the efficiency, stability, and cost of the electrocatalyst used in the process.9,10

Along with stability and low cost, excellent charge transfer ability and many active sites are some of the pre-requisites for an effective electrocatalyst.11 EWS comprises two half-cell reactions: the hydrogen evolution reaction (HER) and the oxygen evolution reaction (OER).12 As usual, the low electrochemical efficiency results from the slow kinetics of the HER and OER. Both reactions require more overpotential to break the inherent reaction energy barrier.13 An efficient electrocatalyst should have the features of better electronic conductivity, large active surface area, long-term stability, and strong catalytic activity.14 The best electrocatalysts for splitting water are based on noble metals, including Pt- and Ir/Ru-based compounds.15 However, the limited availability of resources and the high cost significantly restrict the scope of the potential applications of these materials.16,17 Therefore, developing high-quality noble metal-free or noble metal doped transition metal (TM) electrocatalysts is crucial for commercialising large-scale water electrolysis.18 Recently, a lot of work has been done on developing several approaches to increase the stability and activity of noble metal-based materials using nanotechnology to catalyse water splitting. These techniques primarily target the structural design, which is accomplished by single-atom construction, interfacial structure engineering (strain effect), and heteroatom introduction (ligand effect). When heterogeneous materials are incorporated into compounds based on noble metals, the ligand effect happens, causing electrons to move between two distinct atom groups and altering the electronic structure.17 The adsorption/desorption characteristics of the intermediate species are associated with the electronic structures of catalysts, which are crucial for electrocatalytic efficiency. The d-band theory suggests that by altering the electronic structures of catalysts to provide the distinct electronic effect, the oxygen/hydrogen adsorption energy might be efficiently tuned. Heteroatom doping is one of the best methods for fine-tuning the electronic structures of catalysts. It is widely known that alloying noble metal NPs with TMs (Fe, Co, or Ni) that are readily available on Earth might improve the electrocatalytic activity of catalysts. The single atomic effect is a quickly developing field of study, which offers an efficient way to maximise atom utilisation efficiency and catalytic activity by shrinking catalysts to the size of a single atom.16,17 We critically reviewed composites derived from TM-based materials for their effectiveness in the electrochemical HER and OER in alkaline environments, as depicted in Scheme 1.


image file: d3cc06015b-s1.tif
Scheme 1 Schematic illustration of this review theme based on TM electrocatalysts for overall water splitting.

Recently, significant progress has been made in the design and synthesis of TM-based catalysts with outstanding performance.19 These catalysts include TM-based metal oxides (TMOs), phosphides (TMPs), nitrides (TMNs), carbides (TMCs), borides (TMBs), chalcogenides, and dichalcogenides (TMDs). These catalysts have been chosen because of their fascinating properties, which include high cyclability, ionic conductivity, and mechanical strength.20,21 The advantages of transition metals (unique d electron configurations, low cost, synergistic effect of multi-metal atoms, and excellent stability) and their porous features (large surface areas, high pore volumes, and modulated pore structures) perfectly complement one another, which is especially advantageous for the development of next generation electrocatalysts.22

TMOs, TMPs, TMNs, TMCs, TMBs, and chalcogenides are potential catalysts because they can donate electrons to or accept electrons from a reagent, depending on the type of reaction. They are incredibly helpful in electrochemical transformations due to their capacity to access and switch between various oxidation states, form complexes with the reagents, and serve as good sources of electrons.23 These materials have several appealing qualities, including abundance, fascinating physicochemical properties, and a high electrocatalytic potential. Furthermore, the d-electron density of TMs is enhanced by the doping of nitrogen, carbon, and boron atoms. This contraction of the d-band also results in the electronic structures of TMNs, TMCs, and TMBs approaching the Fermi level and resembling those of noble metals. In addition to having good corrosion resistance, TMNs, TMCs, and TMBs bind well during EWS under both acidic and alkaline conditions. Additionally, due to the alteration of the electron density distribution in the hybrid materials with great dispersity and more exposed active sites, there are synergistic effects between TMNs, TMCs, TMBs, and non-metal atoms that lead to good electrocatalytic activity and stability.23 An important class of inorganic materials known as TMDs exhibit various catalytic, magnetic, optical, and electrical properties. Recently, the amazing characteristics of this family of materials have been studied by researchers.24–26 Earth-abundant TMDs with the general formula MX2 (M = transition metal; X = chalcogen) are currently being developed at a rapid rate for water splitting. These TMDs include two-dimensional (2D) layered materials such as MoS2, WS2, WSe2, TaS2, MoSe2, and TiS2 and pyrite phase structured TMDs (CoSe2, CoS2, NiS2, and NiS2).27 These 2D layered MX2 nanosheets exhibit remarkable physical, chemical, and electrical properties compared to their bulk counterparts, particularly due to their 2D shape and atomic thickness. Each layer typically consists of three atomic layers covalently joined in an X–M–X configuration. Weak van der Waals forces connect adjacent MX2 layers to create bulk crystals.28 TMDs have been investigated as HER and OER electrocatalysts in the last ten years and have demonstrated higher catalytic activities similar to noble metal-based catalysts. Atomically thin TMDs have a larger precise surface area with more active sites for the HER and OER than other Earth-abundant electrocatalysts.

This review discusses the TM-based composites for efficient electrochemical HER and OER in alkaline media. The Scopus database was used to observe the progress made in the last 10 years starting from 2014 to 2023 for the following keywords “Noble metals for water splitting”, “Transition metal oxides for water splitting”, “Transition metal chalcogenides for water splitting”, “Transition metal phosphides for water splitting”, “Transition metal dichalcogenides for water splitting”, and “Transition metal carbides, nitrides and borides for water splitting”. From data observation, we have seen that the number of publications for reported materials has increased yearly. In comparison, noble metals have been reported in many publications owing to their efficient water-splitting performance. TMOs stand second in terms of publication, followed by TMPs. It is also noted that the scientists have paid more attention to TM nitrides, carbides, borides, and chalcogenides, which has gradually increased in the last five years. The 10-year recent research progress is presented in Fig. 1. This timely review focuses on all classes of TMs for both water-splitting reactions in alkaline media and differs from previously reported reviews.29–31 This is the first review in which all TM types as electrocatalysts are critically discussed with recent developments. The timeline of the development of TM-based electrocatalysts is shown in Table 1.


image file: d3cc06015b-f1.tif
Fig. 1 10 years of recent progress in TM-based electrocatalysts; data were obtained through Scopus.
Table 1 Timeline showing the history of development of TM-based electrocatalysts32–44
Year Development history
1789 • Discovery of water electrolysis
1888 • Industrial alkaline water electrolysis
1890 • Asbestos membrane
1900–1966 • First pressurized electrolyzer
• Nafion-based proton exchange membrane water electrolyzer
1974–1996 • Electronic configuration of tungsten carbide resembles that of platinum near the Fermi level
• The chemisorption of hydrogen on the (111) and (110)-(1 × 2) surfaces of iridium and platinum
• Hydrogen electrochemistry on a platinum low-index single-crystal surface in alkaline solution
2005 • Molecular chemistry of consequence to renewable energy
2007 • The edge structure of MoS2 as an active site toward the HER and OER revealed
2014–2016 • A Bi2Te3@CoNiMo composite as a high-performance bifunctional catalyst for hydrogen and oxygen evolution reactions
• Nanocrystalline Mo2C as a bifunctional water-splitting electrocatalyst
• Cobalt-oxide-based materials as water oxidation catalysts: recent progress and challenges
2017–2023 • Single-atom catalysts for electrocatalytic applications
• Strategies for developing TMPs as heterogeneous electrocatalysts for water splitting
• Transition metal (Fe, Co and Ni)–carbide–nitride (M–C–N) nanocatalysts: structure and electrocatalytic applications
• Recent advances in 2D materials and their nanocomposites in sustainable energy conversion applications
• Nanoarchitectonics for TMS-based electrocatalysts for water splitting
• Designing TMB-based electrocatalysts for applications in EWS
• Recent progress in transition metal selenide electrocatalysts for water splitting
• Recent strategies to improve the catalytic activity of pristine MOFs and their derived catalysts in EWS
• Phosphorus-doped MoS2 with sulfur vacancy defects for enhanced EWS


2. Mechanisms of electrochemical water splitting reactions

Generally, overall water splitting (OWS) reactions involve two half-cell reactions, the HER and OER. In the HER reaction, water gets reduced to generate H2 at the cathode, while it is oxidized at the anode in the OER reaction to generate O2. The sluggish kinetic reactions of the HER and OER are the primary obstacles preventing the practical application of water splitting due to elevated overpotentials, especially in alkaline media. To generate H2 and O2 efficiently, highly efficient catalysts are required to reduce the potential for the HER and OER. The design of electrocatalysts depends upon the operational conditions of the water electrolysis cell. There are three main categories of electrolysis technology at the moment: (i) electrolysis using a proton exchange membrane (PEM), (ii) electrolysis in an alkaline solution, and (iii) solid oxide water electrolysis at high temperatures. Because the temperature is so high, solid oxide water electrolysis uses a lot of energy. The water splitting is carried out in an acidic environment and using a PEM in a PEM-based electrolysis cell. Other membranes, such as those with lower gas permeability and higher proton conductivity, do not have all of the advantages exhibited by PEMs.45 The acidic medium is necessary because of its quick hydrogen production rate and good energy efficiency. However, cutting-edge noble metal and metal oxide-based catalysts can be used as the OER electrocatalysts. Water splitting occurs in an alkaline environment in an alkaline electrolysis cell. Compared to acidic cells, water splitting in alkaline cells broadens the options for electrocatalysts to include non-noble metals or metal oxides. As a result, superior electrocatalysts for electrolytic water splitting are inexpensive, have high catalytic activity, and are stable.46

2.1. Mechanism of the HER

Hydrogen is considered as an efficient energy fuel to replace fossil fuels since it is the only byproduct of water. Electric current is used in the HER process to separate H2 and O2 from water. The HER is an electrochemical process that involves multiple steps, takes place at the cathode electrode, and makes use of acidic and alkaline solutions (Fig. 2a and b).47 The Volmer, Tafel, and Heyrovsky reaction mechanisms can be used to explain the entire process of the HER. The HER takes place according to the following equations (eqn (1)–(3)) in an acidic solution:
 
Volmer reaction: M + H+ + e → M–H*(1)
 
Heyrovsky reaction: M–H* + H+ + e →M + H2(2)
 
Tafel reaction: 2M–H* → H2 + 2M(3)

image file: d3cc06015b-f2.tif
Fig. 2 HER mechanism route in (a) alkaline and (b) acidic media. OER mechanism route in (c) alkaline and (d) acidic media.48 The corresponding images (c) and (d) have been reproduced with permission from Elsevier © Copyright.

In an alkaline solution, the HER reaction takes place according to the following equations (eqn (4)–(6)):

 
Volmer reaction: M + H2O + e → M–H* + OH(4)
 
Heyrovsky reaction: M–H* + H2O + e → M + OH + H2(5)
 
Tafel reaction: 2M–H* → H2 + 2M(6)

The HER that occurs on the electrode surface is an adsorption–desorption process, as shown by eqn (1)–(6), and M stands for surface metal sites. First, the electrode surface adsorbs H atoms with the deduction of one electron, forming M–H*, which is equivalent to the Volmer process (eqn (1) and (4)). The Heyrovsky reaction and the Tafel reaction, which are both described in Fig. 2a and b, correspond to the formation of M–H*, which can subsequently mix with H+ in an acidic solution (H2O in an alkaline solution) or another M–H* to create H2 molecules. It is necessary to adsorb H atoms regardless of the HER reaction pathway, which is intuitively indicated by the free energy of hydrogen adsorption (GH*). In addition to generating strong bonds with the adsorbed H*, a good HER catalyst should also be weak enough to enable appropriate bond breaking, which aims to liberate the H2 molecules.49

2.2. Mechanism of the OER

The OER, or water oxidation, is a key electrochemical process in water splitting and rechargeable metal–air batteries.50,51 The OER involves the generation of oxygen molecules on the catalyst surface through four discrete electron transfer steps.52Fig. 2c and d shows the OER mechanism under alkaline and acidic conditions. The Sabatier principle is the foundational hypothesis underpinning our understanding of the OER activity of electrocatalysts. The surface metal cations are the primary OER active sites (M). Many M–O-bonded reaction intermediates will develop during the process. According to the concept of Sabatier, a high level of catalytic activity can be observed on a surface if the adsorbed species adhere to it neither too strongly nor poorly. Furthermore, it has been found that the OER is very pH sensitive, and the reaction kinetics in alkaline and acidic media are different. As shown in Fig. 2d, hydroxyl groups (OH) in alkaline environments converted into O2 and H2O molecules, but under acidic and neutral conditions, two H2O molecules converted into four protons (H) and O2. Typically, the OER process has four steps: (i) H2O/OH adsorption on the catalyst surface; (ii) the formation of reactive intermediates; (iii) O2 molecule release; the formation of OH adsorbed species as a result of OH anion adsorption and discharge at the anode surface indicates that, mainly on the metal site, the OER activity in the alkaline solution begins. The OH ion and this adsorbed OH species then react, releasing an electron and creating H2O and adsorbed atomic O. Following this, an OH anion reacts with an adsorbate O atom to form adsorbate OOH species; and (iv) combination of this adsorbed OOH species with additional OH anions, releasing an electron. This reaction also results in the production of adsorbed O2 and H2O. The desorption of O2 occurs last. The production of adsorbed OOH species is considered the rate-restraining step in this series of events. The mechanism can be visually illustrated as shown in Fig. 2c. The hypothesized process differs in specific ways from that in the alkaline medium and is analogous to those under acidic conditions. The reactions involved in the OER reaction in acidic media are described in the following equations (eqn (7)–(11)), and in an alkaline solution, the OER reaction takes place according to the following equations (eqn (12)–(16)):
 
H2O + M → M–OH* + H+ + e(7)
 
M–OH* → M–O* + H+ + e(8)
 
2M–O → 2M + O2(9)
 
M–O* + H2O → M–OOH* + H+ + e(10)
 
M–OOH* → M + O2 + H+ + e(11)
 
OH + M → M–OH* + e(12)
 
M–OH* + OH → M–O* + H2O + e(13)
 
2M–O* → 2M + O2(14)
 
M–O* + OH → M–OOH* + e(15)
 
M–OOH* + OH → M + O2 + H2O + e(16)

Moreover, two primary reaction pathways are also involved in the OER: firstly, the adsorbate evolution mechanism (AEM), where a metal serves as the redox centre, and the electronic state around the Fermi level has metallic traits.53 Secondly, the lattice-oxygen mediated mechanism (LOM) involves oxygen serving as the redox centre, and the electronic state close to the Fermi level displays features like oxygen. In the conventional AEM, the OER involves different reaction intermediates such as *OH, *O, *OOH, and *O2, which are adsorbed on active metal centres.54 The minimal theoretical overpotential for the explored catalysts using the AEM is estimated to be nearly 0.37 V, determined by a linear scaling relationship between the adsorption energies of *OH and *OOH intermediates (ΔG*OOH = ΔG*OH + 3.2 eV).55 Nevertheless, this scaling relationship fails to account for some catalysts that have been shown to have lower overpotentials. The existence of a scaling relationship restricts the advancement of catalysts aimed at reaching the optimal state of the OER. Nevertheless, this scaling relationship also provides certain benefits.56 The scaling relationship-based volcano plot may be used as a universal tool to forecast catalyst activity based on the Sabatier principle, leading to a decrease in experimental and computational costs. Electrocatalysts using the AEM, distinguished by its reaction pathway, showed notable stability.57

3. Significant parameters for HER and OER electrocatalysts

The electrocatalytic activity can be determined using the following fundamental parameters, such as electrolyte, electrode, potential (onset), Tafel slope (b), exchange current density (jo), overpotential (η), stability, Faraday efficiency (FE), and turnover frequency (TOF).58 The rate of HER and OER electrocatalysts for electrochemical performance depends on different structures, degree of wettability, conductivity, and catalyst entry into an electrolyte. Electrodes are divided into two types: flat surface electrodes and 3D electrodes. In the literature, glassy carbon (GC) electrodes include flat surface electrodes (FSEs), which are frequently employed to evaluate the efficiency of electrocatalysts. FSEs provide only a single channel for electrolytes to penetrate, limiting catalysis to the surface of catalysts. Although GC electrodes are straightforward, they only provide a modest quantity of catalyst loading. A binder is needed in GC electrodes to stabilize the catalysts (such as Nafion). However, overusing binders makes the catalyst surface more resistant and blocks the active sites. Therefore, scientists are greatly focused on generating catalyst inks without a binder and directly building catalysts onto 3D conducting electrodes such as carbon cloth (CC), Ni foam (NF), and carbon paper (CP). 3D electrodes are very conductive owing to the various channels for electrolyte diffusion from all sides of catalysts.59 The pH of the electrolytes strongly influences the effectiveness of HER and OER electrocatalysts. Regarding HER and OER electrocatalysis, the alkaline medium is preferable to a neutral or acidic one. Alkaline solutions have mainly been used as electrolytes in recent investigations. This is of utmost importance when preparing electrocatalysts that can operate at any pH. When comparing the catalytic behaviour of different electrocatalysts, the onset potential is the parameter that is utilized most frequently. The point at which a sudden increase in current is noticed is referred to as the onset potential. It is challenging to determine the precise magnitude of the onset potential.60 As a result, the overpotential value of 10 mA cm−2 is thought to be more reliable and frequently employed. For the OER, a rotating ring disc electrode (RRDE) can be used to monitor the precise onset potential. The overpotential is one of the most crucial factors to consider when assessing the HER and OER performances of electrocatalysts. In an ideal situation, the applied potential required to perform the specific reaction would be identical to the optimal value. However, in the real world, to break the energy barrier, an applied potential that is significantly higher than the equilibrium potential is required.61 Because of this, the overpotential is typically defined as the possible distinction between the potential that is applied and the potential that exists when conditions are in equilibrium. Typically, it is expressed in volts (V) or millivolts (mV). Tafel analysis is used to determine the HER and OER mechanisms to compare the catalytic efficiency of several catalysts and reaction kinetics. For water splitting, it is possible to achieve anodic or cathodic overpotential. The concerned parameters including overpotential, Tafel slope value, and reversible hydrogen electrode (RHE) conversion can be determined using the following expressions:

The Nernst equation can be applied to calculate the reference potential (for example, Ag/AgCl) for the conversion into the RHE potential:

image file: d3cc06015b-t1.tif
where image file: d3cc06015b-t2.tif is 0.2412 and overpotential (η) is obtained by subtracting an onset thermodynamic potential of 1.23 V (for the OER) and 0 V (for the HER) of the water splitting system.
Overpotential (η) = Onset potential (ERHE) V − 1.23 V (for the OER)

Moreover, the Tafel slope can be calculated by utilizing the Tafel equation:62

η = b[thin space (1/6-em)]log[thin space (1/6-em)]j + a
where η, b, and j indicate the overpotential, Tafel slope, and current density, respectively.63 Another crucial quantity that may be calculated using the Tafel equation is the exchange current density when it equals zero. It reveals the material's inherent activity in the electrolyte at the reversible potential and shows the rate at which electrons migrate from the electrode to the electrolyte. A higher exchange current density indicates improved electrocatalytic capability. An ideal HER and OER electrocatalyst should have a significant exchange current density and a moderate Tafel slope. FE is the proportion of real products to theoretical products and is also frequently utilized in fuel cells and water-splitting reactions. For water electrolysis, FE was equal to the percentage of produced H2 and theoretical H2, which was another significant parameter for analysing the reaction process of water electrolysis. In addition to the previously specified parameters, stability is crucial for practical application.64 In general, a material's stability can decide whether it can be utilized to produce hydrogen and oxygen by water electrolysis in a practical manner or not.65 As a result, assessing a material's electrochemical stability is crucial. Generally, the electrochemical stability of an HER and OER catalyst can be easily determined by long-term chronoamperometry (CA), chronopotentiometry (CP) and cyclic voltammetry (CV). Furthermore, the scientists have also used TOF to measure the catalytic interaction of every active site by calculating the number of reactants that transformed into desirable products by each active site in unit time.66 Therefore, the TOF value is a significant parameter for evaluating the electrocatalytic performance of the catalyst.

4. Transition metals as electrocatalysts for the HER and OER

The stable electrocatalyst can effectively reduce the reaction energy and accelerate the reaction rate. Exploring highly efficient, lower-cost, and stable electrocatalysts is a feasible strategy for energy conversion.67 As discussed before, the most efficient catalysts for water electrolysis to date are Pt, IrO2, and RuO2 for the HER and OER, but their scarcity and high price have severely limited their use in large-scale industrial applications. Some parameters should be considered for selection of materials, which can help find efficient electrocatalysts, such as high cyclability, high ionic conductance, mechanical and thermal stability, and higher surface area.68 In this regard, TM catalysts, such as copper (Cu), nickel (Ni), cobalt (Co), iron (Fe), and molybdenum (Mo), are known to be active electrocatalytic materials for the HER and OER.69,70 These TMs have unfilled d orbital electrons and unpaired electrons in the d orbital, which are conducive to the adsorption and desorption of reactive groups.71 These materials have been reported to be highly active towards the electrocatalytic HER or OER with both high electrocatalytic activity and stability.72,73 Moreover, some of them have also been revealed to possess bifunctional properties, which could simultaneously accelerate the HER and OER, thus enabling them to be promising electrocatalysts for commercial utilization. In the literature, there are many studies on Ni toward OWS reactions, and Ni is an Earth-abundant metal with exceptional catalytic activity because of its electronic structure. It has high electrical and thermal conductivities due to its remarkable electronic structure.74 In contrast to other TMs like Cu, Co, Fe, and Mo, which are known to be effective catalysts for the HER and OER, use of the Ni dopant with other TM materials is a very active approach toward the HER and OER.75,76 It should also be pointed out that the Ni-based catalyst can be further improved through surface structure and morphology tunning.73 In contrast to Ni-based materials, Cu-based materials are extensively utilized in electrochemical catalysis and other fields because of their cheap cost and high performance. The synergistic activity of Cu and Ni, porous structure, enhancement in surface area and the number of active sites led to a boost-up in intrinsic activity.77 These advanced hybrids take advantage of the intrinsic activity of Ni and Cu as well as hybrid TM characteristics such as strong conductivity, large surface area, easy surface functionalization, and high stability. Due to its superior electrical conductivity and stability under alkaline conditions, Co, one of the numerous non-noble metal elements on Earth, has garnered more attention as a prospective HER and OER catalyst. Co-based metal–organic framework (MOF) compounds and their derivatives, Co oxides, Co chalcogenides, Co phosphides, Co-based layered double hydroxides, and so forth are examples of cobalt-based catalysts.78 Their distinct electrochemical characteristics and distinctive structural diversity make them a hotspot for research and development. Fe is an inexpensive metal widely used in industrial practice. At the same time, it exhibits relatively high HER and OER activities. Pure Fe is occasionally used as a reference material that can show diverse results concerning various preparation methods, morphology or surface area.79 Regrettably, very little data on pure Fe as a cathode material are available. One possible reason is that Fe alone exhibits worse stability into the elevated alkaline environment. Fe-based catalysts have also attracted much attention because of their abundance and relatively low cost. They have been extensively used as dopants with other TMs to enhance the HER and OER performance.80 Mn, Mo, and Se-based electrocatalysts have also been extensively studied in artificial water-splitting systems. Doping TMs can be an effective process for enhancing catalytic behaviour because doping can change the electronic properties of the host metals by adjusting the local charge redistribution without altering the desired intrinsic properties of the host elements.81

Inspired by conventional electrodeposition technology, Niu et al.82 reported a rapid self-etching electrodeposition method for constructing a copper-incorporated three-dimensional Ni–Cu coating on copper sheets as a superior bifunctional catalyst (Fig. 3a). According to this study, the optimal overpotentials for the HER and OER are 76 mV and 290 mV at 10 mA cm−2, respectively. The calculation also shows that Ni–Cu has a higher theoretical hydrogen desorption energy and OER overpotential than pure Ni and Cu. The as-prepared Ni–Cu electrocatalyst showed exceptional stability, particularly in the HER reaction process stability test up to 50 hours. He et al.83 developed a new bifunctional Fe–Cu–N co-doped carbon catalyst toward the HER and OER in alkaline media using a simple hydrothermal annealing method (Fig. 3b). Experimental results showed that high temperatures caused nitrogen to move from degraded g-C3N4 to carbon spheres, shattering the inert atmosphere of the surface of the graphite carbon and enhancing the active surface area. However, the capacity of as-prepared samples to split water was improved by controlling the ratio of Fe and Cu. According to electrochemical measurements, the catalyst with co-doped Fe–Cu had improved electron conductivity and a larger ESCA. Fe–Cu@CN3 showed good catalytic water splitting performance, with overpotentials of 91 mV and 362 mV at 10 mA cm−2 for catalysing the HER and OER, respectively. Qiu et al.84 created an effective, long-lasting, bifunctional Ni–Fe/NiMoNx electrocatalyst on nickel foam (Fig. 3c–e). This electrocatalytic system's HER and OER performances are notable, with 49 and 260 mV overpotentials at 20 mA cm−2, respectively. Ni–Fe/NiMoNx electrodes only need 1.54 V for 10 mA cm−2 to split water overall (Fig. 3f and g). This improved electrocatalytic activity for the HER and OER results from the interaction of many active components created following NH3 treatment. Badarnezhad et al.85 used one-step electrodeposition to create Ni–Mo–Fe as a bifunctional electrocatalyst. Minimal overpotentials of 65 and 161 mV for delivering 10 and 100 mA cm−2 for the HER in alkaline media were accomplished using this electrocatalyst with appropriate composition and current density. For the OER, the as-fabricated electrode demonstrated 344 and 408 mV for supplying 10 and 100 mA cm−2. Furthermore, in long-term stability experiments in an alkaline medium, the as-prepared electrocatalyst exhibited excellent stability and low degradation in overpotential for the HER and OER. The superb performance of the produced Ni–Mo–Fe material may be due to the synergistic effects of Ni, Mo, and Fe components and the non-binding structure.85


image file: d3cc06015b-f3.tif
Fig. 3 (a) Schematic illustration of the self-etching electrodeposition process.82 (b) Schematic of the fabrication process of Fe–Cu@CN and CN.83 (c) Illustration of synthesis and SEM characterization of the Ni–Fe/NiMoNx catalyst. (d) TEM and (e) HRTEM images of Ni–Fe/NiMoNx. (f) Overall water splitting polarization curves of Ni–Fe/NiMoNx. (g) Stability test at 100 mA cm−2 for 40 h.84 The corresponding images have been reproduced with permission from Elsevier © Copyright.

For the preparation of bi-metallic Ni–Mo alloy catalysts for the HER and OER, Sun et al.86 used a high throughput technique involving a straightforward co-sputtering approach. The inherent catalytic performance of Ni–Mo alloy electrocatalysts is increased due to the synergistic action between Ni and Mo. This results in their excellent HER and OER performance. The Ni40Mo60 electrocatalyst shows an overpotential of 258 mV at 10 mA cm−2 in the OER. In contrast, the Ni90Mo10 electrocatalyst exhibits the highest HER performance with an incredibly lower overpotential value of 58 mV at 10 mA cm−2. Cheng et al.87 used an easy and affordable electrodeposition technique to create three-dimensional freestanding porous Cu foam in situ shielded CoNi alloy nanosheet arrays with customizable compositions. These freestanding Co81Ni19 alloy nanosheet arrays display exceptional electrocatalytic characteristics with overpotentials of 132 and 240 mV for the HER and OER to produce a catalytic current density of 10 mA cm−2, respectively. Tian et al.88 used a one-step electrodeposition method to synthesize quaternary Ni–Co–S–P nanoparticles (NPs) directly on carbon fibre fabric. Furthermore, the Ni–Co–S–P NPs demonstrate exceptional stability and longevity for both the HER and OER owing to their super hydrophilic and superhydrophobic features.88 Recently, great attention has been paid to TM oxides (TMOs), phosphides (TMPs), chalcogenides, dichalcogenides (TMDs), nitrides, carbides, and borides due to their Earth-abundance, low-cost and superior electrical and optical properties.30,89–91 These catalysts have proven to be highly active, stable, and promising for the EWS scaling-up process. Thus, it is vital to further understand the reaction pathways and basic concepts of the structure-related performance relationship of HER and OER processes. Then, the design and production of an effective catalyst for EWS can be considered.

4.1. Noble metal-based electrocatalysts

Noble metals such as Pt, Pd, Rh, Ru, and Ir are still the most efficient catalysts for the HER and OER. However, their high price and scarcity prevent them from being used widely. Thus, it is crucial to create noble-metal-based catalysts with low noble metal loading, high activity, and superior tolerance for energy conversion systems.92 Noble metal-based catalysts are inherently flawed, mostly due to their low stability and restricted reserves. To solve these problems, scientists have created several strategies, for example: (i) modifying the shape of noble metal catalysts; (ii) adding new elements; (iii) building single-atom anchors for noble metals on sophisticated support materials; (iv) shrinking the size of noble metal-enhanced nanocatalysts to the single-atom scale; and (v) designing the interface between noble metal and transition metal compounds.93 Pt is a highly favoured active HER catalyst for the generation of H2 because it has the highest HER activity. Noble metal oxides are regarded as cutting-edge materials because of their superior stability and catalytic activity in both basic and acidic environments. It is commonly recognised that IrO2 and RuO2 are thought to be the most advanced OER electrocatalysts, with IrO2 being more stable under harsh acidic and alkaline conditions. Other noble metals such as silver and palladium have also been used for the HER and OER.94 For instance, a simple technique for creating Pt single atoms in N-doped mesoporous carbon was presented by Lou et al.95 According to experimental results, when compared to Pt/C, the resulting Pt single atoms on carbon can exhibit a 25-fold increase in mass activity for the HER. Pt single atoms can contribute to the exceptional HER activity by increasing the density of electronic states and creating a favourable charge density distribution. A simple wet-chemical technique was used by Huang et al.96 to create Mn-doped Ru nanosheet branches, which have an ultrathin thickness of 1.09 nm and are made up of many nanoscale branches with plenty of edges. Taking advantage of their electrical and structural properties that are inherent, when it comes to the HER and OER, the as-prepared RuMn nanosheets can demonstrate exceptional catalytic performance. The Ir–Ni3N heterogeneous interface was formed on the surface of the NF substrate and noble metal Ir NPs were directly implanted into the Ni3N nanosheet structure by the one-step ammonia nitridation procedure reported by Liu et al.97 XPS demonstrated that the promotion of the HER and OER was significantly aided by the heterogeneous interfaces between IrOx and NiOOH that developed during surface self-reconstruction. To further enhance the electrocatalytic performance, Manjunatha et al.98 produced decorated Pd nanoclusters and a nitrogen-doped carbon nanotube-supported Co–Fe nanocomposite. The Pd nanocluster-coated Co–FeNCNTs demonstrated higher trifunctional electrocatalytic activity for the ORR, OER, and HER.

4.2. Transition metal oxide-based electrocatalysts

TMO electrocatalysts have garnered considerable attention in the energy conversion and storage field due to their affordability, abundance in the Earth's crust, excellent corrosion resistance, and enhanced electrochemical activity.99,100 In the past, pure TMOs showed better OER potentials but low reactivity towards the HER due to their unattractive hydrogen desorption ability.101 Therefore, significant efforts have been undertaken to modify TMOs for HER activity in alkaline environments, making the modified TMOs excellent candidates for OWS.102 Hybridisation of TMOs with other electroactive substances (such as TMOs, TMPs, TMs, and TM alloys) can enhance the catalytic performance.103 As a result of the mixed-valence states of Co2+/3+/4+, the electrocatalysis of the HER and OER on cobalt oxide electrodes has long been a subject of intense interest in electrochemistry.104 Min et al.105 created Co- and Ce-based bimetallic organic frameworks that served as self-templates for the production of CeO2-decorated Co4N (Co4N/CeO2) nanostructures with highly linked heterointerfaces as active electrocatalysts for the OER in alkaline media. The corresponding illustration can be seen in Fig. 4a. Xu et al.106 effectively created a porous cobalt oxide material with oxygen vacancies using a ligand-assisted polyol reduction process. Such technology gives a chance to successfully create oxygen vacancies on the surface with a regulated concentration while successfully converting Co(OH)2 nanoplates into CoOx with well-maintained morphology. OER performance benefits greatly from the vast surface area created by the 2D porous structure and the abundance of active sites caused by oxygen vacancies, producing overpotential values as low as 306 mV at 10 mA cm−2.106 Cai et al.107 meticulously developed 3D porous NiFe LDH nanosheets electrodeposited on Ni nanochains. SEM showed the core–shell nanostructure, which is advantageous for increasing the catalyst's surface area and exposing more active sites. The excellent conductivity of the Ni@NiFe LDH is also reflected in its low charge transfer resistance (Rct).Zhang et al.108 hydrothermally created a three-dimensional (3D) self-supporting nickel sponge (SN) with nano-Ni synapses and then deposited 2D CoFeLDH nanosheets on the nickel sponge using the electrodeposition approach. CoFeLDH was also deposited on a NF substrate for comparison using the same deposition procedure. In a 1 M KOH alkaline solution, the as-prepared CoFeLDH/SN showed superior electrocatalytic capabilities with 208 mV at 10 mA cm−2 for the OER and 63 mV at 10 mA cm−2 for the HER. Hanan et al.109 created PdO decorated CoSe2 through a combination of wet-chemical and ultraviolet irradiation methods to get the desired doping of PdO (as a promoting agent) as an active electrocatalyst for the OER in 1.0 M KOH. However, the corresponding XRD patterns of CoSe2 and PdO-based composites can be seen in Fig. 4b. It can be observed that the crystallinity of CoSe2 is disturbed due to the addition of PdO, and a bit of shifting of the peak can be noticed for the structure. Elakkiya and colleagues created a nanoporous NiCo2O4 material with a floral morphology, which served as a bi-functional electrocatalyst for water splitting in a 1.0 M KOH electrolyte. This material demonstrated exceptional performance, with high mass activity, low overpotential (360 mV at 10 mA cm−2 for the OER and HER, respectively) and modest Tafel slopes (150 mV dec−1 and 123 mV dec−1 for the OER and HER, respectively). These extraordinary capabilities were ascribed to the material's unique surface and electrical band structure, including an abundance of active sites and porosity, electrochemical active surface area, quick electron transfer, and structural stability.110 Moreover, the Co2FeO4@rGO composite has been prepared by Hanan et al.111 through the hydrothermal method. The prepared material exhibited activity as a trifunctional electrocatalyst toward the HER, OER, and ORR in an alkaline medium having 320 and 240 mV overpotentials. The structure of composite materials was nanosheets of rGO combined with Co2FeO4, as shown in Fig. 4c and d. Srinivas et al.112 demonstrated a carbon nanotube (CNT) network with remarkable water-splitting performance that was produced from a MOF and inserted in a CoFe2O4/CoO heterostructure.113 Because of the collective advantages of MOF-derived porous carbon with abundant oxygen vacancies, the CoFe2O4/CoO heterostructure embedded in the CNT network, the hybrid catalyst (CoFe/C-650), shows great performance for the OER, HER, and overall water splitting with several exposed catalytic sites and synergistic interactions. The catalyst requires minimum overpotentials of 246 mV (for the OER) and 164 mV (for the HER) at 10 mA cm−2 and has low Tafel slopes of 45.27 and 86.38 mV dec−1, respectively. It is interesting to note that the overpotential needed for complete water-splitting catalysis is 1.614 V (η10), which is close to the overpotential needed for the standard Pt/C (cathode) and RuO2 (anode) combination.112 Additionally, Aftab et al.114 reported a mixed CoS2@Co3O4 composite material through a hydrothermal method as a promising electrocatalyst for the HER in alkaline media. The proposed composite is a flower-like structure embedded with Co3O4 NPs. Synergistic effects between metallic oxides and metallic sulfides provided the composite with an overpotential value 320 mV at 10 mA cm−2 and a Tafel slope value of 42 mV dec−1. In their study, Xin et al.115 described a very effective 3D bulk catalyst made of core–shell nanostructures. This catalyst has a one-of-a-kind design in which flower-shaped CuCoMoOx nanosheets are decorated with CoCu alloy NPs. These nanosheets were wrapped uniformly around the copper nanowire (Cu NW) cores, which were held together by a CF substrate. Meanwhile, the nanosheets connected with Cu NWs demonstrated excellent HER and OER activities and amazing overall water splitting properties in the alkaline medium due to their notable conductivity, leading to quick electron transfer. The electrocatalyst performed well in both the HER and the OER, with overpotentials of 75 mV and 315 mV at 100 mA cm−2, respectively.115 Simultaneously, Solangi et al.116 introduced TMO (Co3O4) with doping of palladium (Pd) oxide as a promoting agent through an aqueous chemical and ultraviolet-irradiation method (for PdO doping). The composite had a nanoneedle-like structure embedded with NPs of Pd. Fig. 4e and f shows the XPS spectra of Co3O4 and its composite with Pd, including Co 2p and O 1s. The as-prepared material successfully catalyzed the electrochemical water splitting reaction in alkaline media with an overpotential value of 250 mV at 20 mA cm−2, a Tafel slope of 58 mV dec−1, and a charge transfer resistance of 48.5 Ω, leading to fast transfer kinetics for the OER.
image file: d3cc06015b-f4.tif
Fig. 4 (a) Schematic illustration of the synthesis of Co and Ce-based bimetallic organic frameworks.105 (b) XRD spectra of pristine CoSe2, PdO, and their composites for the OER.109 TEM and HRTEM images, FFT patterns, and line profiles of (c) GO and (d) Co2FeO4@rGO.111 XPS analysis of Co3O4 and Co3O4–Pd and (e) Co 2p and (f) O 1s.116 All the corresponding images have been reproduced with permission from Elsevier and Royal Society of Chemistry © Copyright.

Digraskar et al.117 conducted a study where they created a composite material consisting of oleylamine-functionalized graphene oxide and Cu2ZnSnS4 (OAm-GO/CZTS), with the morphology as shown in Fig. 5a–c. They examined its potential as a highly efficient electrocatalyst for the HER and OER (Fig. 5d). The OAm-GO/CZTS composite exhibited exceptional electrocatalytic performance and long-lasting durability for generating hydrogen and oxygen, even in acidic and basic environments. At a current density of 10 mA cm−2, the composite displayed overpotentials of 47 mV for the HER and 1.36 V for the OER (Fig. 5e and f). Additionally, it exhibited Tafel slopes of 64 and 91 mV dec−1 for the HER and OER, respectively. These results surpass those achieved by TMCs and are on par with the performance of commercially available precious metal catalysts. Li et al.118 successfully used a rapid two-step electrodeposition approach to create a heterostructure comprising CoNiP and NiFe-layered double hydroxides (CoNiP@NiFe LDHs). CoNiP's electron density may be increased, further optimising the Gibbs free energy for hydrogen adsorption via the electron rearrangement facilitated by the discrepancy in the distinct work functions of CoNiP and NiFe LDHs. To achieve 10 mA cm−2, CoNiP@NiFe LDHs require a low overpotential of 68 mV. Furthermore, the OER performance of the CoNiP@NiFe LDHs is significantly improved. At 50 mA cm−2, they show a drastically lower overpotential of 255 mV. This enhancement may be linked to valence band energy level (EVB) changes and the effective collection of holes in the NiFe LDH layer's space charge region. These CoNiP@NiFe LDHs are used as electrodes in the electrolyser, resulting in an astonishingly low cell voltage of 1.59 V at 50 mA cm−2.118 In contrast, Hanan et al.119 recently reported the PdO@Co2FeO4 nanostructures as a bifunctional electrocatalyst toward the HER and OER in an alkaline medium of 1.0 M KOH. The PdO@Co2FeO4 nanostructure material was prepared through a hydrothermal method under ultraviolet irradiation. This composite material has nanoparticle-like morphology with irregular structure (due to PdO addition). The material showed small crystal structure changes due to the addition of PdO within the crystal of Co2FeO4. Interestingly, the electrocatalyst showed promising bifunctional behaviour with overpotential values of 269 and 259 mV at 10 and 20 mA cm−2 current densities, with Tafel slope values of 49 and 59 mV dec−1 for the HER and OER, respectively. Moreover, Laghari et al.120 produced a magnesium (Mg) supported Co3O4 composite material through the wet chemical growth method. XRD analysis of the proposed composite material revealed a nanoneedle-like structure and the cubic crystal structure of Co3O4. The MgO-based Co3O4 composite material demonstrated a robust electrochemical behaviour, with an overpotential value of 273 mV to reach a current density of 10 mA cm−2, an ECSA of 12.8 cm2, a low charge transfer resistance of 45.96 Ω, and a Tafel slope value of 64 mV dec−1. On the other hand, a composite material based on iron oxide (Fe3O4) and reduced-graphene-oxide (rGO) as a potential candidate for the HER in 1.0 M KOH has been introduced by Hanan et al.121 They have produced the rGO@Fe3O4 nanostructure through a hydrothermal method using hydrazine hydrate (N2H4) as the reducing agent. They introduced graphene nanosheets decorated with Fe3O4 NPs. The addition of 2D reduced graphene-oxide sheets improved the conductivity of Fe3O4, facilitating the HER with an overpotential value of 300 mV at 10 mA cm−2, a Tafel slope value of 80 mV dec−1, and an Rct of 95 Ω. Faid et al.122 used in situ Raman to investigate the reaction mechanism of the alkaline HER on Ni/NiO nanosheets (Fig. 5g and h). They revealed that the active phase was β-Ni(OH)2, which remained unchanged during the HER reaction (Fig. 5i), contributing to the good activity and stability. Overall, numerous TMOs have been thoroughly researched and showed improved HER and OER performances in alkaline media, including binary, ternary, and quaternary compositions. Crystal structure, elemental content, surface morphology, and electrical characteristics all impact the performance of these oxides. Doping, surface changes, and nano-structuring have all been used to improve their electrocatalytic activity and efficiency.123 Understanding structure–function correlations of TMOs and developing better synthesis techniques are critical areas of study to further enhance their electrocatalytic efficiency. These materials have a lot of potential for enhancing energy conversion and storage technologies, thus contributing to a more sustainable and clean energy future.124,125


image file: d3cc06015b-f5.tif
Fig. 5 SEM images of (a) GO, (b) pure CZTS NPs, and (c) OAm-GO/CZTS. (d) Diagram of the OAm-GO/CZTS electrocatalyst on a GCE for the total water-splitting reaction (HER and OER). Polarization curves for (e) HER and (f) OER of OAm-GO/CZTS and related samples.117 (g) Scheme of in situ Raman. (h) Digital photo of in situ Raman. (i) in situ Raman spectra of Ni/NiO after 4000 seconds of chronoamperometry experiments at −0.4 V vs. RHE in 1 M KOH.122 All the corresponding images have been reproduced with permission from Elsevier and ACS © Copyright.

4.3. Transition metal phosphide-based electrocatalysts

TMPs are the most prominent candidates for the alkaline HER and OER due to their superior catalytic efficiency and low cost. The phosphorus (P) atoms in TMPs can accept electrons from neighbouring TMs due to their high electronegativity, which is the primary cause of their exceptional catalytic activity.126 The negatively charged P atoms can trap the positively charged proton as a base. As a result, the moderate bonding between the reaction intermediates and the presence of P atoms lead to product formation on the catalyst surface. Therefore, TMPs often have a higher electrocatalytic activity than other TM-based composites.127–129 Different techniques have been developed to improve the intrinsic catalytic activity and structural stability of TMP materials. These methods include: (i) doping with heteroatoms, a practical approach for changing the electronic structure and creating lattice defects, resulting in optimized adsorption and desorption of intermediates at active sites;130 (ii) constructing composites based on TM–P to create heterogeneous catalysts. To increase the intrinsic catalytic activity, more catalytically active sites must be exposed; (iii) depositing the grafted TM–P active phases on well-conductive supports, and (iv) creation of a core–shell structure with a strong contact between the core material and the shell layer.131 This improves the conductivity and endurance of Co–P-based electrocatalysts and increases the number of active sites. The most often reported TMPs are CoxPy, NixPy, FexPy, MoxPy, CuxPy, and WxPy, and they have demonstrated outstanding performances to date.132

Based on the DFT calculations, the highly conductive 2D TMP monolayers and their corresponding oxidized counterparts are anticipated to be effective HER and OER catalysts. Mo2P-2H and Fe2P-2H show the best electrochemical performance among a group of predicted TMPs, further supported by an experiment showing that bifunctional catalysts exhibit good performance. Consequently, the advanced DFT simulations are excellent tools for finding novel bifunctional electrocatalysts. Bimetallic phosphides possess a better activity for the alkaline HER and OER than single TM-based phosphides by utilizing the synergistic effects of bimetallic sites with modified electrical configurations and structural flexibility. This section will address the application aspects and practical methods for increasing the catalytic activity of TMPs in the alkaline HER and OER. For instance, Zhou et al.133 have developed a workable topological conversion technique for preparing O-doped CoP nanosheets through a moderate phosphidation process on well-defined Co(OH)2 nanosheets. O-CoP nanosheets with moderate O incorporation showed excellent electrocatalytic activity toward the HER and OER, with overpotentials of 98 and 310 mV at a current density of 10 mA cm−2, respectively. This has been revealed by experimental and theoretical studies, which also showed that the O doping level in the O-CoP nanosheets played an important role in determining their electrocatalytic efficiency. According to mechanistic research, the right O dopant considerably increased the electrical conductivity and controlled CoP's electronic structure, which optimized the adsorption energy of intermediates and resulted in exceptional electrocatalytic performance. Chen et al.134 have published a study describing the development of a highly efficient catalyst Ru–MnFeP/NF, which originated from a Prussian blue analogue (Fig. 6a). This catalyst, consisting of an ultralow concentration of ruthenium (1.08 wt%) supported on NF (Fig. 6b and c), demonstrates exceptional performance in both the OER and HER when tested in 1 M KOH (Fig. 6d–i). Remarkably, the Ru–MnFeP/NF electrode achieves current densities of 20 mA cm−2 for the OER and 10 mA cm−2 for the HER, with only minimal overpotentials of 191 mV and 35 mV, respectively (Fig. 6j).


image file: d3cc06015b-f6.tif
Fig. 6 (a) Schematic illustration of the synthesis of an Ru–MnFeP/NF catalyst. XPS profiles of (b) Ru 3p and Fe 2p of Ru–MnFeP/NF. Digital photos of the (d) Ru–MnFeP/NF electrode, (e) water splitting cell, and (f) gas collection device. (g) The volume of H2 and O2 collected during electrolysis. Digital photos of gas collection of (h) O2 and (i) H2. (j) Polarization curves of Ru–MnFeP/NF and control samples for the OER and HER in 1 M KOH.134 All the corresponding images have been reproduced with permission from Wiley © Copyright.

Hao et al.135 presented the methodical design and discovery of a tri-functional electrocatalyst Co/CoP-HNC by embedding Co/CoP within a nitrogen-doped carbon polyhedral structure derived from MOFs. The catalyst exhibited a large electrochemical surface area and high electrical conductivity, leading to notable performance in the HER, OER, and ORR.136 A recent study conducted by Ding et al.101 investigated the potential of three different Prussian blue analogues (PBA) with varying iron-based components as precursors for synthesizing metal phosphides through a simple heat treatment process. By adjusting the temperature of phosphidation, they successfully obtained a series of metal phosphides, namely FeCoP, FeNiP, and FeMnP. Among these, FeCoP exhibited the most pronounced porous structure and the widest range of pore sizes when phosphidated at an optimal temperature of 400 °C. This unique porous structure of FeCoP facilitated efficient mass transfer and oxygen release during the OER.

Furthermore, introducing Co into the metal phosphides resulted in a significant decrease in Rct compared to Ni and Mn. This decrease in Rct enhanced the electron transport during the OER process. Because of its exceptional porous geometric structure and distinctive electronic properties, FeCoP-400 demonstrated outstanding stability and catalytic performance for the OER in 1.0 M KOH solution. At a current density of 10 mA cm−2, FeCoP-400 demonstrated the overpotential of 261 mV, surpassing both commercial RuO2 and most other OER electrocatalysts. Yun et al.137 recently presented a novel approach to synthesize a self-supporting catalyst electrode, P-NiFe@CF, for facilitating the urea oxidation reaction (UOR) and HER. A Ni–Fe alloy was electroplated over a carbon fiber substrate to create the electrode, which was subsequently phosphidated. In the HER process, the P-NiFe@CF electrode displayed exceptional performance with a remarkably low overpotential of 0.023 V (versus RHE) at the current density value of 10 mA cm−2. The analysis encompassed the Fe2p, Ni–Fe@CF, Ni2p, P2p, and XPS peak profiles of P-NiFe@CF, providing valuable insights into the composition and properties of the synthesized catalyst electrode. Liang et al.138 produced NiCoP by converting Ni–Co hydroxide by plasma treatment as a bifunctional catalyst for water splitting. As expected, the NiCoP electrocatalyst on a Ni foam support performed better for the HER in an alkaline solution. The catalyst showed an incredibly lower overpotential of 32 mV at a current density of 10 mA cm−2 with an exchange current density value of 1.36 mA cm−2 and a Tafel slope of 37 mV dec−1. In the case of the OER, the active sites were surface oxides originating from NiCoP that were created during anode scanning, resulting in an overpotential of 280 mV and a Tafel slope of 87 mV dec−1 at a current density of 10 mA cm−2. A two-electrode water electrolysis cell with a cell potential of 1.58 V at 10 mA cm−2 was also demonstrated. NiCoP was used as the bifunctional catalyst in this cell.138 Salvo et al.131 successfully synthesized mixed metal phosphides with various metal ratios, specifically Ni/Co and Ni/Fe phosphides. These phosphides were obtained through the phosphidization process of high-surface-area hydroxides grown on CC via a hydrothermal method. The HER performance of the synthesized materials reached its peak when synergistic interactions were established among the different metal centres. Remarkably, the materials containing the highest proportion of ternary compounds, namely NiCoP and NiFeP, exhibited outstanding performance with the overpotential value of only 50 mV needed to achieve a current density of 10 mA cm−2. The as-prepared catalysts displayed a needle-like structure. To create a bifunctional iron cobalt phosphide (FeCoP/3C) electrocatalyst, carbon black Vulcan XC-72, carbon nanotubes, and graphene oxide (referred to as 3C) were used. The catalytic performance toward the HER and the OER in 1.0 M KOH electrolyte was examined. The FeCoP/3C electrocatalyst exhibits 120 and 220 mV overpotentials at a current density of 10 mA cm−2 with Tafel slopes of 129 and 73 mV dec−1.139 In a groundbreaking achievement, crystalline Cu3P phosphide nanosheets were successfully grown on conductive NF, forming a Cu3P@NF composite material. This composite material exhibits exceptional electrocatalytic properties for water splitting, in terms of overall water splitting by visible light irradiation and electrocatalysis. For OER catalysis, a current density of 10 mA cm−2 can be achieved with an impressively low overpotential of approximately 320 mV and a Tafel plot slope of only 54 mV dec−1 in a KOH solution with a concentration of 1.0 M. While, in HER catalysis, the Cu3P@NF electrode displays excellent performance, requiring a mere overpotential of approximately 105 mV to attain a current density of 10 mA cm−2. Moreover, the Cu3P@NF electrode demonstrates its capability for overall water splitting in a water electrolyser. At an applied voltage of approximately 1.67 V, the electrode exhibits a catalytic current density of 10 mA cm−2, enabling the efficient conversion of water into hydrogen and oxygen. Wang et al.140 devised an innovative microwave-induced plasma synthesis approach for the fabrication of an ordered porous structure containing cobalt phosphate and cobalt oxides, as illustrated in Fig. 7a. In this method, an ordered porous Co3O4 framework with a pore size of approximately 200 nm was initially grown in situ on a conductive graphite plate. Subsequently, a plasma treatment was conducted in the presence of red phosphorus, transforming a portion of Co3O4 into Co3(PO4)2 and CoO (Fig. 7b–j). The plasma treatment not only resulted in the formation of cobalt phosphate but also induced a substantial density of oxygen vacancy defects, which served as active sites for both the HER and OER as shown in Fig. 7(k). This unique combination of features allowed for the achievement of overall water splitting at a low cell voltage of 1.65 V (Fig. 7l).


image file: d3cc06015b-f7.tif
Fig. 7 (a) Schematic illustration of the synthetic strategy of cobalt phosphate and cobalt oxides on a carbon plate via a microwave-induced plasma method. Digital photos of (b) PMMA/CP (a polymethyl methacrylate template on a carbon plate), (c) 3D Co3O4/CP, (d) 3D Co:Pi/CoOx/CP (a 3D cobalt phosphate and cobalt oxide mixture on a carbon plate); the microwave-induced plasma treatment of the carbon plate sample inside a vacuumed round bottom flask. SEM images of (e) and (h) PMMA/CP, (f) and (i) 3D Co3O4/CP, and (g) and (j) 3D Co:Pi/CoOx/CP. (k) Digital photo of a water electrolysis cell comprising 3D Co:Pi/CoOx/CP as both the anode and cathode in 1.0 M KOH. (l) LSV curve and inset of chronopotentiometry responses of 3D CoPi/CoOx/CP at a constant current density of 10 mA cm−2 for 100 h.140 All the corresponding images have been reproduced with permission from ACS © Copyright.

Kumaravel et al.141 and his team proposed that TMPs exhibit remarkable performance in the OER and HER under different pH conditions. Additionally, the various phases of phosphides, such as phosphorus-rich or phosphorus-deficient metal composites like Ni/Co–P, present significant potential for applications involving total water splitting. In this study, a low-crystallinity and micro-spherical CoFe-P/NF catalyst that was produced via potentiostat electrodeposition on an NF substrate demonstrated excellent performance in terms of HER, OER, and water splitting in a 1 M solution of KOH. To create a current density of 10 mA cm−2, the CoFe-P/NF material required overpotentials of 45 mV for the HER and 287 mV for the OER. In addition, it was found that the Tafel slopes for the HER and OER were 35.4 and 43.2 mV dec−1, respectively.142 Wu et al.143 created cobalt phosphide NP modified N-doped porous carbon spheres (CoP@NPCSs) using an innovative and effective coordination–polymerization–pyrolysis (CPP) technique, which served as a potent catalyst for the OWS (Fig. 8a–c). To evaluate its HER performance, LSV polarization curves were obtained from recent reports for various metal phosphide coatings, including CN, Co@NPCSs, CoP NPs, CoP@NPCSs, untreated metal foil and a Pt mesh electrode, in a 1.0 M KOH electrolyte.144


image file: d3cc06015b-f8.tif
Fig. 8 (a) Schematic illustration of the synthesis method for CoP@NPCSs. (b) XRD profile (up) and Raman spectrum (down), (c) SEM image of CoP@NPCSs. (d) Overall water splitting of CoP@NPCSs.143 (e) and (f) TEM images and FFT pattern of Fe–Co–P. (g) Schematic illustration of synthesis of Fe–Co–P.145 (h) and (i) TEM images of Ru–NiCoP/NF. (j) Polarization curves of the HER and OER of Ru–NiCoP/NF. (k) Overall water splitting using Ru–NiCoP/NF||Ru–NiCoP/NF. (l) Stability of Ru–NiCoP/NF||Ru–NiCoP/NF.146 All the corresponding images have been reproduced with permission from Elsevier, ACS © Copyright.

Tafel plots in an alkaline electrolyte were generated for CN, Co@NPCSs, CoP NPs, CoP@NPCSs, and Pt/C to analyze their electrochemical behaviour during the HER. The current–time (It) curves were obtained for CoP@NPCSs in both acidic (upward) and alkaline (downward) electrolytes to assess their performance in the HER at fixed overpotentials of 112 mV and 115 mV, respectively (Fig. 8d). Wang et al.147 used a simple technique to strategically manufacture an iron-nickel phosphide (Fe2P/Ni2P) heterostructure on nickel foam using Fe2P nanoparticles placed on Ni2P nanosheets. The Fe2P/Ni2P heterostructure functions as a bifunctional electrocatalyst, delivering very low overpotentials of 64 and 185 mV to achieve a current density of 10 mA cm−2 for the HER and OER. The permeable 3D heterostructure also helps expose abundant active sites, enhance ion diffusion, and enable gas bubble release, improving catalytic performance.147 In a recent study conducted by Liu et al.,144 the experiments were carried out using a rotating disk electrode in a KOH solution with a concentration of 1 M. The HRTEM images presented in Fig. 8e and f exhibit d-spacings of 2.44 and 2.77 Å. These values can be correlated with the fast Fourier transform (FFT) pattern and the interplanar spacings of (111) and (011) of the Fe–Co–P alloy. The d-spacing values fell between those of FeP and CoP, indicating the production of a nanostructured Fe–Co–P alloy. They created hollow and conductive iron-cobalt phosphide (Fe–Co–P) alloy nanostructures utilising an Fe–Co metal organic complex in a simple and scalable manner. The Fe–Co–P alloy catalyst was subjected to chronopotentiometric and chronoamperometric measurements (Fig. 8g–i). These measurements were conducted at a steady current density of 10 mA cm−2 (Vt) and an overpotential of 252 mV (it). Polarization curves were obtained for various catalysts,145,146 such as Fe–Co–P, Ru–NiCoP/NF, NiCoP/NF, Ni2P/NF, and commercial RuO2/NF, in a KOH solution with a concentration of 1 M (Fig. 8j–l). Lin et al.,148 using a straightforward technique, developed a Mo-doped NiCoP nanosheet array through hydrothermal and phosphation processes for the HER and OER. Notably, Mo doping may successfully modify the electronic structure of NiCoP, increasing the number of electroactive sites and improving each site's intrinsic activity. This highly effective dual-functional Mo-doped NiCoP catalyst has shown relatively low overpotentials and long-term stability. In an alkaline environment, the overpotentials for the HER and OER to actuate a current density of 10 mA cm−2 were 76 and 269 mV in 1.0 M KOH, respectively.148 In conclusion, TMPs show extraordinary stability and catalytic activity for both the HER and OER because of their distinct electrical and structural characteristics. They have a bunch of new sites and have great electrical conductivity, which makes electrochemical processes possible. To improve the electrocatalytic efficiency of different TMPs, researchers have investigated binary, ternary, and quaternary compositions. We can open the door to more effective and ecologically sustainable energy storage and conversion technologies by using the properties of TMPs as bifunctional electrocatalysts.

4.4. Transition metal nitride-, carbide- and boride-based electrocatalysts

TMCs, TMNs and TMBs are also regarded as promising electrocatalysts for general water splitting owing to their unique physical, chemical, and electrical characteristics.149,150 The electronic structure of TMNs is comparable to that of Pd and Pt noble metals. It is up to the Fermi level because of the nitrogen atoms in TMNs, which could boost d-electron density and cause the d-band to constrict. In addition to their strong metallic conductivity, TMNs have high corrosion resistance, which makes them desirable for water splitting.151 Nevertheless, the intrinsically low conductivity, stability, and water-splitting activity of bifunctional electrocatalysts constantly put a cap on their performance. Additionally, with the introduction of interstitial N with a modest radius, most TMNs maintain their close-packed metallic range, leading to high electronic conductivity. These intriguing characteristics of TMNs make them potential replacements for noble metal compounds as catalysts.152 This section focuses on new research initiatives to create enhanced TMNs with high activity for water splitting and emphasizes the impact of structure engineering on electrochemical performance. TMNs exhibit the same catalytic capabilities due to their structural similarities, where either C or N exists in the interstitial sites.

Panda et al.153 showed a new easy way to make copper nitride (Cu3N), which acts as a bifunctional electrocatalyst for the OER, HER, and overall water electrolysis in alkaline media, as shown in Fig. 9a. Interestingly, the electrochemically deposited Cu3N on NF had exceptionally low overpotentials for both the OER and HER, and the overall water splitting cell potential was just 1.62 V, with an incredible durability of more than 10 days (Fig. 9b). Most significantly, to drive effective catalysis, the coordinatively unsaturated Cu in Cu3N changed in situ under reducing and oxidative conditions into a copper-rich shell that serves as the active site over an equally significant electrically conductive Cu3N core. Fang et al.154 discovered a suitable technique to produce cobalt nitride nanosheets encapsulated in graphdiyne (CoNx@GDY NS) to improve the catalytic performance. This method resulted in an impressive current density of 10 mA cm−2 at low overpotentials of only 70 and 260 mV for the HER and OER, respectively, surpassing the performance of commercial Pt/C and RuO2 products. Moreover, when utilized as a bifunctional catalyst in a two-electrode electrolytic cell, it exhibited an unusually low cell voltage of 1.48 volts at a current density of 10 mA cm−2. This was significantly lower than the Pt/CeRuO2 pair and the catalyst maintained stability for more than 20 hours. Wu et al.155 integrated active OER (Ni3N) and HER (NiMoN) catalysts as heterostructures for successful overall water splitting, as shown in Fig. 9c. They used a sufficient ratio to encapsulate Ni3N and NiMoN on CC and modulated the crystalline degree and crystalline size by modifying the preparation parameters. The improved Ni3N–NiMoN electrocatalysts demonstrated good stability and low overpotentials of 277 and 31 mV at a current density of 10 mA cm−2 for both the OER and HER in the alkaline electrolyte. Lu et al.156 produced cobalt–molybdenum nitride nanosheet arrays (CoMoNx NSAs/NF) on NF (Fig. 9d) using a simple hydrothermal and nitridation technique.


image file: d3cc06015b-f9.tif
Fig. 9 (a) Atomic structure of Cu3N. (b) Water splitting by Cu3N||Cu3N on NF.153 (c) Schematic of the formation of the Ni3N–NiMoN heterostructure.155 (d) SEM image of CoMoNx-500 NSAs/NF. (e) Polarization curves of overall water splitting by CoMoNx-500 NSAs/NF. (f) The potential comparison of CoMoNx-500 NSAs/NF with electrodes from the literature.156 (g) Schematic diagram of the fabrication of Co6W6C@NC/CC. (h) SEM image of Co6W6C@NC/CC. (i) Polarization curve of Co6W6C@NC/CC.157 All the corresponding images have been reproduced with permission from Elsevier, ACS, and Wiley © Copyright.

The resultant CoMoNx NSAs/NF catalyst has more active sites, a larger surface area, and also good electrical conductivity, accelerating the reaction rate. Also, the synergistic action of Co2N and Mo2N cooperatively enhances charge transfer, boosting the electrocatalytic activity even further. Due to these advantages, the optimum CoMoNx-500 NSAs/NF catalyst demonstrates good HER and OER activities with overpotential values of 91 and 231 mV at a current density of 10 mA cm−2. Additionally, the CoMoNx-500 NSAs/NF combination has a surprising property. It requires a considerably low cell voltage of 1.55 V at a current density of 10 mA cm−2 (Fig. 9e), outperforming other TMN catalysts reported in the literature (Fig. 9f). Chang et al.158 investigated nanomaterials with highly reactive exposed facets, which have sparked considerable attention due to their large catalytic performance improvement. By performing an in situ N/O exchange, the (100) facet of NiMoN nanowires was preferentially exposed, and the shape was adjusted using a NiMoO4 precursor. Under both alkaline and acidic conditions, the facet-tuned NiMoN nanowires demonstrated remarkable electrocatalytic activity for the HER, identical to the noble metal platinum.

TMCs have gathered a lot of attention for the OWS process because of their excellent corrosion resistance, great mechanical toughness, high electrical conductivity, exceptional stability, and low cost.159 TMCs and TMNs are compounds in which carbon and nitrogen atoms are integrated inside the parent metal's interstitial sites. Face-centered cubic (fcc), hexagonally closed packed (hcp), and even simple hexagonal (hex) crystal structures are often found in these materials. TMCs and TMNs are especially intriguing for various applications in electrocatalytic water splitting due to their unique electrical structure and intrinsic characteristics.160 Numerous research studies have been published in recent years, and EWS has advanced quickly. Despite certain catalysts showing promising characteristics, they cannot meet actual application demands because of the difficult preparation process and their unsatisfactory stability. Modifying the catalyst's structure and shape makes it possible to expose additional active sites, which helps improve the electrocatalytic reaction rate. Due to their distinct electronic structure and characteristics, TMCs, particularly TMNs, offer interesting applications in electrocatalytic water splitting.161 Several methods, including carbothermal reaction, chemical vapour deposition, metallothermic reduction, self-propagating high-temperature synthesis, and mechanical alloying, have been used to produce TMCs and TMNs.162 The most popular techniques for developing active TMCs and TMNs for electrocatalysis typically include carbothermal reaction and chemical vapour deposition. Novel methods also received significant attention due to the development of nanotechnology to produce nanostructures. Giordano et al.163 used a soft urea glass process to manufacture metal carbides and nitrides. Compared to other synthesis methods, this method requires a lower reaction temperature, a faster reaction process, a more homogenous dispersion, and a larger surface area. Li et al.164 introduced nitrogen-doped carbon nanofibers (NCNFs) that were effectively manufactured for use as a bifunctional electrocatalyst using a cost-effective and simple electrospinning approach, followed by a carefully regulated carbonization procedure. These NCNFs were subsequently combined with Ni and Mo2C NPs, resulting in Ni/Mo2C-NCNFs. The optimized Ni/Mo2C (1[thin space (1/6-em)]:[thin space (1/6-em)]2)-NCNFs showed the best electrocatalytic performance for total water splitting in a 1.0 M KOH electrolyte among the several compositions examined. The overpotentials recorded for the HER and the OER at a current density of 10 mA cm−2 were 143 mV and 288 mV, respectively. Kou et al.165 reported a remarkable bifunctional catalytic performance for both the HER and the OER by efficiently anchoring Co single atoms onto the surface of a Mo2C nanosheet (referred to as Co SAs/Mo2C). Notably, compared to Co single atoms supported by N-doped carbon, the Co SAs/Mo2C catalyst demonstrated a greater TOF per active site. This discovery implies that the single Co sites connected to the Mo2C nanosheets have higher intrinsic catalytic activity, emphasizing the Co SAs/Mo2C system's strong catalytic capabilities. Based on theoretical calculations, the author hypothesized that the Co-Mo3 coordination is accountable for the noticeably improved intrinsic catalytic activity and that metal carbides have a greater application potential than the extensively researched carbon-based substrates. In another work, Rabi et al.166 demonstrated the electrocatalytic ability of a hybrid heterostructure of molybdenum carbide (Mo2C) and ZIF-67 produced by high-temperature annealing. This bifunctional electrocatalyst demonstrated exceptional water-splitting performance with high activity for both HER and OER processes. The hybrid material MCZ-2, composed of ZIF-67 and Mo2C in a ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]2, exhibited enhanced catalytic activity for both the HER and OER. At a current density of 10 mA cm−2, MCZ-2 demonstrated remarkably low overpotentials of 169 mV and 208 mV for the HER, while for the OER, the overpotentials were 340 mV and 370 mV at current densities of 10 mA cm−2 and 50 mA cm−2, respectively. Furthermore, MCZ-2 displayed excellent stability over 24 hours at a current density of 100 A cm−2. The improved HER and OER performances of the MCZ-2 hybrid can be attributed to the synergistic interaction between ZIF-67 and Mo2C and the modification of active sites due to the presence of N and Co in the hybrid structure. Chen et al.157 successfully synthesized a highly efficient bifunctional electrocatalyst, Co6W6C@NC, which consists of vertically aligned porous cobalt tungsten carbide nanosheets embedded in an N-doped carbon matrix, using a MOF-derived method (as depicted in Fig. 9g and h). The interaction of Co and W atoms in the Co6W6C@NC system changes the electronic state of the carbide synergistically, resulting in optimum hydrogen-binding energy. Consequently, Co6W6C@NC performs better in the HER, with a low overpotential of 59 mV at a current density of 10 mA cm−2. Furthermore, during the OER, Co6W6C@NC is converted in situ into tungsten-activated cobalt oxide/hydroxide, which acts as an active species for the OER with a 286 mV overpotential at a current density of 10 mA cm−2. The Co6W6C@NC electrocatalyst has a low cell voltage of 1.585 V at 10 mA cm−2 and displays high alkaline stability (Fig. 9i). A solid-state co-reduction approach was used to manufacture the Ni/Mo2C@NC electrocatalyst, involving the synthesis of hybrid nanoparticles composed of multiple-interfacial nickel (Ni) and molybdenum carbide (Mo2C). Following this, the nanoparticles were embedded in N-doped carbon nanosheets (Fig. 10a).167 Because of its high conductivity, numerous interface active sites, and the synergistic interaction of Ni and Mo2C nanoparticles (Fig. 10b), the electrocatalyst exhibits extraordinary activity in the HER. In an alkaline solution, it has an extremely low overpotential of 91 mV at a current density of 10 mA cm−2 and a Tafel slope of 74 mV dec−1. Furthermore, the catalyst is very stable, making it an ideal option for long-term use in electrolysis cells. It acts as a bifunctional electrode with durability and dependability, needing an applied voltage of 1.64 V at a current density of 10 mA cm−2 (Fig. 10c–e).


image file: d3cc06015b-f10.tif
Fig. 10 (a) Schematic illustration of the synthesis process of the Ni/Mo2C@NC hybrid. (b) HRTEM image of Ni/Mo2C@NC. (c) Overall water splitting by Ni/Mo2C@NC. (d) Illustration of the reaction mechanism of the water splitting reaction on Ni/Mo2C@NC. (e) Stability test of Ni/Mo2C@NC.167 (f) and (g) SEM images of FexNi1−xB/SSG. Polarization curves of Fe0.2Ni0.8B/SSG for the (h) OER and (i) HER.168 All the corresponding images have been reproduced with permission from Elsevier © Copyright.

TMBs are of great interest due to their diverse compositions, environmentally benign nature, and excellent bifunctional catalytic performances facilitated by the unique electronic effect between the transition metal (M) atom and B atom.169 The main research elements of TMBs as new electrocatalysts in recent years include Fe, Co, Ni and Mo. Chunduri et al.170 developed a new hybrid Co–P–B electrocatalyst by employing a straightforward chemical reduction technique to combine the phosphide and boride for the water-splitting reactions. Under optimized conditions, the catalyst exhibited overpotentials of 145 mV and 290 mV, respectively, to achieve a current density of 10 mA cm−2 in the HER and the OER. Hu et al.171 used 3D bark-like N-doped carbon (BNC) as a nanoreactor for water splitting to create metal borate Co–Bi and Ru-doped Co–Bi (Ru–Co–Bi) nanomeshes with a lot of nanosized pores. A low voltage of 1.53 V produced a current density of 10 mA cm−2 when Co–Bi/BNC and Ru–Co–Bi/BNC were used as the anode and cathode materials, respectively. These materials outperformed the noble RuO2/Pt/C counterpart (1.61 V). Li et al.172 used a straightforward one-pot NaBH4 reduction process to construct a series of Co-based multi-metal boride nanochains by adding other metal elements (Ni and Fe) at room temperature for the overall water splitting. NiCoFeB nanochains with low overpotentials of 345 and 284 mV at a current density of 10 mA cm−2 and Tafel slopes of 98 and 46 mV dec−1, respectively, outperformed other Co-based metal borides under investigation in terms of electrocatalytic performances for the HER and OER. In a single step, Lao et al.168 effectively created nickel boride (NiB) and iron-doped nickel boride (FexNi1−xB) by growing them on 3D self-supporting graphene (SSG) electrodes (Fig. 10f and g). The Fe0.2Ni0.8B/SSG electrode only needed a relatively small overpotential of 263 mV, which is the best OER activity achieved to date for a metal boride. The NiB/SSG electrode showed strong HER activity with average OER performance (Fig. 10h and i). The Fe0.2Ni0.8B/SSG and NiB/SSG water electrolysers produced a current density of 10 mA cm−2 at a voltage of just 1.62 V. Furthermore, both the NiB/SSG and Fe0.2Ni0.8B/SSG catalysts demonstrated outstanding stability over a 14-hour testing period with no deactivation being noted. Doping electrocatalysts based on TMBs with heteroatoms is a sensible technique for regulating the local disorder and improving catalytic activity, just as for electrocatalysts based on other types of materials. In addition, TMB-based electrocatalysts might be grown on carbon materials via a strong interaction, which would increase both the electrical structure and the electrocatalytic performance of the material.

4.5. Transition metal chalcogenide-based electrocatalysts

Transition metal chalcogenides are compounds composed of transition metals linked to chalcogen atoms including sulfur (S), selenium (Se), tellurium (Te), and polonium (Po). Such materials display a variety of features such as semiconducting, magnetic, and catalytic behaviours. These materials have been used in a range of applications including electronics, catalysis, and energy storage.173 Moreover, TM chalcogenides include a wide range of compounds with different stoichiometries that go beyond having just two chalcogen atoms in every transition metal.174 This category may consist of ternary or more complex compounds. They often demonstrate a variety of characteristics and uses that extend beyond those associated with the dichalcogenides. TMDs are compounds consisting of a transition metal atom bound to precisely two chalcogen atoms. These materials consist of layered structures held together by robust covalent bonds within the layers and less strong van der Waals forces between the layers.175 Specifically, TMDs are of great interest because of their distinctive electrical and optical characteristics, particularly when in a 2D state. Examples include MoS2, WS2, and their analogues, as mentioned in the next section.

Furthermore, TMSs have gained significant attention and have been widely explored as a highly promising group of bifunctional catalysts for efficient and durable overall water splitting, primarily due to their superior performance and stability. TMSs have a higher intrinsic electroconductivity than metal oxides.176 Therefore, various synthesis techniques, such as chemical vapour deposition, solid phase synthesis, template approach, solvothermal method, and others, have been employed to design sulphide heterojunctions. The solvothermal approach is the most popular synthetic technique due to its simplicity in adjusting heterojunction particle size, shape, and crystal plane exposure. Sulphide structures change depending on the coordination mode of the metal atom.40 There are currently two basic methods for boosting the performance of transition metal sulphide catalysts: speeding up electron transport to increase conductivity and maximizing the number of active sites to enhance the electrocatalytic activity. For example, in a straightforward solvothermal procedure, Li et al.177 synthesized N-doped carbon layer-coated NiCo2S4 hollow nanotubes (NCT–NiCo2S4) using polyacrylonitrile (PAN) as the template. Three-dimensional hollow nanotubes with uniform N-doping layers (Fig. 11a–j) were designed to optimize catalytically active sites, promote mass transfer, avoid corrosion from the electrolyte, and improve activity and stability (Fig. 11k and l). The findings show that the efficient NCT–NiCo2S4 catalyst demonstrated low overpotentials of 295 and 330 mV to deliver a current density of 10 mA cm−2 for the HER and OER, respectively, and exhibited outstanding stability. The low Tafel slope also reflected the rapid HER and OER electrocatalytic reactions. Ma et al.178 developed a straightforward, multi-step method for preparing a porous CuCo2S4/NiCo2S4 core–shell material as a bifunctional electrocatalyst for OWS (Fig. 11m). The porous NiCo2S4-sheets and CuCo2S4 rods on NF can also maintain their structural integrity and offer more active sites and a larger specific surface area (Fig. 11n and o). The CuCo2S4/NiCo2S4 core–shell electrode showed reduced overpotentials of 271 and 206 mV for the OER and HER at a current density of 10 mA cm−2, respectively, as shown in Fig. 11p and q. Kong et al.179 developed novel 3D hierarchical NiMo3S4 nanoflowers (NiMo3S4/CTs) with many defects and reactive sites produced directly on carbon textiles. By doping Ni2+ in the Mo–S crystal lattice to create defect-rich NiMo3S4 nanoflakes, it was possible to increase their electrocatalytic activity for the HER and OER under alkaline conditions. Small overpotentials of 149.5 and 126.2 mV for the HER and OER at 10 mA cm−2 are exhibited by the self-supported NiMo3S4/CT electrode, respectively. Wang et al.180 used a unique two-step hydrothermal technique involving hydrazine monohydrate and thiourea treatment to create a new form of heterostructure Co3S4/Fe3S4 (referred to as CoFe-HM-T). This heterostructure was created using CoFe LDHs and exhibited novel characteristics. The CoFe-HM-T electrode demonstrated outstanding catalytic activity with overpotentials of 320 and 200 mV for the OER and HER at a current density of 100 mA cm−2, respectively. CoFe-HM-T also had the lowest Tafel slopes, which were calculated to be 38.1 and 144.9 mV dec−1 for the OER and HER, respectively. In another work, Liu et al.181 successfully synthesized the Fe-doped Ni0.85Se (FexNi0.85−xSe, x = 0.02, 0.04, 0.05, 0.06, 0.08, 0.1, 0.15, 0.2, 0.3 and 0.4) catalysts by a one-step hydrothermal method and investigated them as highly active bifunctional catalysts for the OER and HER in electrocatalytic water splitting. The findings suggest that sufficient doping can enhance catalytic performance by creating a synergistic interaction between various metal ions. Adding CNTs to create the Fe0.08Ni0.77Se/CNT composite significantly improves the electrocatalytic activity due to the high electronic conductivity of the carbon nanotube network. In comparison to other composites for the OER and HER, the resulting Fe0.08Ni0.77Se/CNT3 composite demonstrates reduced overpotentials of 204 and 108 mV at a current density of 10 mA cm−2 and low Tafel slopes of 58 and 59 mV dec−1, respectively. Similar to TMSs, TMSes have drawn much interest in the last ten years. Selenium (in the fourth period) and sulphur (in the third period) have the same valence electrons. However, in contrast to S, Se exhibits distinct chemical properties due to various periods: Se differs from S in several ways, including its slightly bigger atomic radius, higher metallicity, improved conductivity, and lower ionization energy.182 These factors proved that Se showed fascinating activities for the HER and OER. Wang et al.183 constructed stacked flower-like NiTe–NiSe nanosheets using a simple hydrothermal technique, bringing the usage of transition metal tellurides in clean and renewable energy closer (Fig. 12a and b). The optimized electrode material has an active surface area 20 times larger than single NiTe/NFF or NiSe/NFF, which results in low overpotentials of 164 and 76 mV for the OER and HER, delivering a current density of 10 mA cm−2. These results are also in good correlation with Gibbs free energy DFT calculations (Fig. 12c and d). Zhou et al.184 used a two-step hydrothermal approach to create a Co, Fe co-doped Ni3Se4 nano-flake array (Ni0.62Co0.35Fe0.03)3Se4 on conductive CF. Co, Fe co-doping can significantly increase the hydrogen evolution activity of Ni3Se4 in acidic and alkaline environments. This study revealed the overpotentials of 87 and 53.9 mV for the HER and OER, respectively, and the computed Tafel slopes of (Ni0.62Co0.35Fe0.03)3Se4/CC were 122.6 and 262 mV dec−1 at a current density of 10 mA cm−2. Wang et al.185 described a simple method for creating a selenide/sulfide hetero-structured NiCo2Se4/NiCoS4 catalyst for the HER and OER (Fig. 12e–h). The NiCo2Se4/NiCoS4 heterostructure demonstrated outstanding OER and HER performances, as evidenced by low overpotentials of 248 mV and 180 mV at 10 mA cm−2 for the OER and HER, respectively, and superior long-term stability to the benchmark IrO2 and Pt/C catalysts for the OER and HER as shown in Fig. 12i and j. Pan et al.186 reported a simple and efficient polymerization–pyrolysis–selenization (PPS) approach for the on-site manufacture of nitrogen-doped carbon nanosnakes (NCNSs) containing iron-doped cobalt selenide NPs as depicted in Fig. 12k. These NPs are made of trimetallic networks of zinc/iron/cobalt polyphthalocyanine coupled polymers (Fig. 12l–r). The FeCoSe@NCNS catalyst showed good electrocatalytic activity for the HER with minimal overpotentials (142 and 99 mV in 0.5 M H2SO4 and 1.0 M KOH) due to the synergistic effect regulated by Fe atoms and CoSe NPs and the confinement effect of the in situ generated porous conductive carbon nanosnakes for the OER (320 mV in 1.0 M KOH) at a current density of 10 mA cm−2.


image file: d3cc06015b-f11.tif
Fig. 11 (a) and (b) TEM images, (c) SEM images, and (d) EDS mapping images of NCT-NiCo2S4. XPS profiles of NCT-NiCo2S4 for (e) survey, (f) Ni 2p, (g) Co 2p, (h) S 2p, (i) C 1s, (j) N 1s. (k) Overall water splitting by NCT-NiCo2S4. (l) Stability test at 10 mA cm−2 for NCT-NiCo2S4.177 (m) Schematic diagram illustrating the synthesis process of CuCo2S4/NiCo2S4 electrodes directly grown on NF. SEM images of (n) CuCo2S4 and (o) CuCo2S4/NiCo2S4. Polarization curves of CuCo2S4/NiCo2S4 for the (p) OER and (q) HER.178 All the corresponding images have been reproduced with permission from Elsevier © Copyright.

image file: d3cc06015b-f12.tif
Fig. 12 (a) and (b) SEM images of NiTe–NiSe nanosheets. DFT calculation of free energy for the (c) HER and (d) OER.183 (e) The illustration of the synthetic method for NiCo2Se4/NiCoS4. SEM images of (f) NiCo2Se4, (g) NiCoS4, and (h) NiCo2Se4/NiCoS4. The polarization curves for the (i) HER and OER and (j) overall water splitting by NiCo2Se4/NiCoS4.185 (k) The schematic illustration of synthesis of NCNS encapsulated Fe-doped CoSe NPs (FeCoSe@NCNSs). XANES of (l) Co K-edge and (n) Fe K-edge. The fitted average oxidation states of (m) Co and (o) Fe from XANES spectra. k3-weighted FT-EXAFS spectra of (p) Co and (q) Fe at R spaces. (r) Wavelet transform contour plots of Co and Fe samples and reference samples.186 All the corresponding images have been reproduced with permission from Elsevier © Copyright.

Gong et al.187 created a CoSe/Co(OH)2 carbonaceous material (CoSe/Co(OH)2-CM (AE)) by acid etching and in situ selenization. CoSe/Co(OH)2-CM (AE) demonstrates outstanding electrocatalytic activity in the 1.0 M KOH electrolyte, with low overpotentials of 299 mV for the OER and 207 mV for the HER at a current density of 10 mA cm−2.

4.5.1. Transition metal dichalcogenide-based electrocatalysts. A significant class of inorganic materials known as TMDCs exhibit various catalytic, magnetic, electrical, and optical properties. Researchers have widely studied the amazing characteristics of this family of materials. As the name TMDCs suggests they consist of TMs (Mo and W) combined with chalcogenides (S and Se). A TMD like MoS2 has the chemical formula MX2, where M and X are the transition metal and the chalcogen, respectively. Due to the weak van der Waals connections between neighbouring layers, bulk TMDs quickly exfoliate into 2D flakes.188 The chalcogen atoms are arranged in two hexagonal planes of the monolayer TMD's sandwich-like X–M–X structure, which is divided by a plane of metal atoms. Between the two chalcogen atom layers is the layer of transition metal atoms. Strong covalent bonds hold together the atoms of one layer, whereas weak van der Waals bonds attract two different layers, allowing these separate sheets to be segregated. The TMDCs are stable and exist as trigonal (1T), hexagonal (2H), and rhombohedral (3R) structures with various stacking configurations.189 A variety of techniques are employed to synthesize TMDCs. These techniques can be broadly categorized into top-down and bottom-up approaches. The type of material, application, and simplicity of the synthesis process are the key determinants of method selection. The layered 2D TMDCs display several desirable characteristics, ranging from high surface area and electrical conductivity to a heterogeneously fast transfer of electrons (depending on composition and layer thickness), which can be found in semimetals, semiconductors, true metals, and superconductors.71 Due to their high binding energies, robust photoluminescence, and variable bandgaps, TMDCs are excellent materials for several applications due to their distinct characteristics, such as optoelectronic devices, photodetectors, LEDs, phototransistors, solar cells, and photocleavable monomers.190 For instance, Zhu et al.191 described the creation of CoS2C@MoS2 core–shell nanofibers by a one-step hydrothermal reaction and a sulfurization procedure using electrospun Co-carbon nanofibers as templates. The resulting CoS2C@MoS2 core–shell nanofibers exhibit outstanding electrocatalytic activity and stability towards the HER and OER, taking advantage of the distinctive structure and the synergistic impact of the components. Sun et al.192 created a special WS2/WSe2 flower-like heterostructure by using heterojunction engineering (Fig. 13a). Cobalt diselenide laminated with molybdenum diselenide (MOF-CoSe2@MoSe2), a novel core–shell structure, was synthesized by Patil et al.193 Its effectiveness as a bifunctional electrocatalyst for the HER and OER in alkaline media was evaluated. The CC/MOF-CoSe2@MoSe2 core–shell structure showed low overpotentials (η10) of 109.87 and 183.81 mV for the HER and OER, respectively. It was made directly on a flexible carbon cloth substrate. For the first time, Lv et al.194 reported the use of hybrid CoSe2/WSe2/WO3 nanowires (NWs) on CC as a very effective HER catalyst, as depicted in Fig. 13b. The developed CoSe2/WSe2/WO3 NWs/CC catalyst showed superior catalytic characteristics against the HER with a low onset potential and Tafel slope. Unlike WSe2/WO3 NWs/CC, CoSe2/WSe2/WO3 NWs/CC only needs a 115 mV overpotential to reach a current density of 10 mA cm−2. The compositional and structural benefits of the WS2/WSe2 heterojunction are combined in this multiscale morphological control design into a hierarchical architecture, which may alter the electronic structure, dramatically facilitating the exposure of more electrochemical active spots. As expected, the hetero-structured WS2/WSe2 catalyst exhibits exceptional HER performance with a small Tafel slope of 74.08 mV dec−1 and a low overpotential of 121 mV at 10 mA cm−2. Zou et al.195 have created evenly distributed triangular WSe2 utilizing electrospun carbon nanofiber mats as the substrate and a CVD reaction that functioned as the selenation process. WSe2 was created under high-temperature conditions with added Ar flow. As a result, the obtained sample exhibited low overpotential for the HER. Recently, Gao et al.196 fabricated MoSe2 with various shapes (nanoplates, nanosheets, and microparticles) at various temperatures (880 °C, 940 °C, and 980 °C, respectively). Because of this, the MoSe2 nanosheets demonstrated exceptional electrochemical characteristics like minimal overpotential and long-term stability for the HER. Huang et al.197 successfully constructed the NiS/MoS2 complex grown in situ on carbon paper (NiS/MoS2/CP) through a single hydrothermal step. It displays relatively low overpotentials of 119 mV (at a current density of 10 mA cm−2) and 314 mV (at a current density of 100 mA cm−2) for the HER and the OER, respectively. The mixed metal dichalcogenide of WS2–MoS2 was electroless plated onto a Cu substrate for HER investigation.198 Due to the surface-active sites linked to the combination of mixed dichalcogenides, the electrocatalytic activity of the WS2-MoS2/NiMoP coating increased, and a synergistic effect developed between the various elements in the coating system. Niyitang et al.199 used a simple hydrothermal process to prepare MoS2, which is then modified utilizing ambient solution plasma for bifunctional hydrogen and oxygen evolution processes (HER and OER). The synthesized MoS2 displays low overpotentials with small Tafel slopes of 33 and 62 mV dec at 10 mA cm−2, respectively. Gao et al.200 fabricated cobalt-doped NiS@MoS2 core–shell nanorods (Co–NiS@MoS2) as a bifunctional catalyst in a straightforward one-step hydrothermal process. The Co–NiS@MoS2 catalyst showed low overpotentials at 50 mA cm−2 for both the HER (139.9 mV) and OER (170.6 mV) in 1.0 M KOH solution. Li et al.201 created a Cu film that supported an MoS2-based electrocatalyst with a partial 1T phase and 3D architecture using inkjet printing (Fig. 13c–e). Results show that (i) nanosized few-layer MoS2 that has been spatially designed by inkjet printing provides adequate active site exposure, (ii) the 1T-MoS2 and RGO conductive network lowers the charge-transfer impedance, and (iii) the Cu support improves the catalyst-electrode charge injection. The corresponding electrocatalyst exhibits low overpotentials (51 mV at 10 mA cm−2 and 126 mV at 100 mA cm−2) and a very low Tafel slope (32 mV dec−1) for the HER, including high stability (over 5000 cycles) and TOF (Fig. 13f and g). Rai et al.202 presented a quick and easy wet-chemical method to create many exposed edges on few-layered MoSe2 and WSe2 nanoflowers. In contrast to WSe2, temperature-controlled reactions show that MoSe2 can be synthesized more quickly. They created a hierarchical heterostructure of MoSe2@WSe2 using a one-step synthesis approach by taking advantage of the quicker kinetics of the synthesis of MoSe2. Among the produced nanostructures, the hierarchical nanostructure has a higher electrocatalytic activity.203 The hierarchical nanostructure needs overpotentials of 231 mV (HER) and 300 mV (OER) to produce a current density of 10 mA cm−2. Xiong et al.204 developed a method of cobalt covalent doping that can induce HER and OER bifunctionality in MoS2 for effective overall water splitting. The findings show that covalent cobalt doping into MoS2 can significantly increase HER activity while also causing outstanding OER activity. The catalyst can easily reach HER and OER onset potentials of 0.02 and 1.45 V (vs. RHE) in 1.0 m KOH when cobalt doping density is optimized. A brief comparative study has been carried out specifically in alkaline media to explore the materials with high electrochemical efficiency, as tabulated in Table 2.
image file: d3cc06015b-f13.tif
Fig. 13 (a) Schematic illustration of the synthesis of the flower-like WS2/WSe2 heterostructure.192 (b) Scheme illustration of the synthesis of CoSe2/WSe2/WO3 NWs/CC by a hydrothermal method and selenation treatment.194 Atomic structures of (c) Cu-MoS2 and (d) Cu-RGO-MoS2. (e) Scheme illustration of the synthesis of MoS2/PVP/RGO. (f) Overpotential and Tafel slope of MoS2/PVP/RGO compared with literature values. (g) Illustration of the HER process of MoS2 nanostructures surrounded by RGO on Cu film.201 The corresponding images have been reproduced with permission from Elsevier © Copyright.
Table 2 TM-based electrocatalysts for the HER and OER in an alkaline medium
Catalyst Reaction type Electrolyte Overpotential (mV) Tafel slope (mV dec−1) Current density (mA cm−2) Ref.
Noble metal-based electrocatalysts
Au33Pt67 HER 1.0 M KOH 171 81 10 205
PtCo HER 1.0 M KOH 76.2 76 20 206
O-Pt3Co-NWs HER 1.0 M KOH 56.1 30 207
Pt24Cu76 NFs HER 0.5 M KOH 18 52 10 208
CoFe–Pt1% HER 1.0 M KOH 18 29 10 209
Pd3Ru HER 1.0 M KOH 42 10 210
IrFe@NC HER 1.0 M KOH 850 30 1000 211
RuNi HER 1.0 M KOH 15 28 10 212
PtRu NCs/BP HER 1.0 M KOH 22 19 10 213
Pt–NiFe LDH/CC HER 1.0 M KOH 28 39 10 214
Pt/Ni-MOF HER 1.0 M KOH 25 42.1 10 215
Ru@MWCNT HER 1.0 M KOH 17 27 10 216
PdO@Co2FeO4 HER 1.0 M KOH 269 49 10 116
PdO@Co2FeO4 OER 1.0 M KOH 259 59 20 116
PdO@CoSe2 OER 1.0 M KOH 260 57 20 109
Pd NPs/Co3O4 OER 1.0 M KOH 250 58 20 116
RuCo@NC OER 1.0 M KOH 25 10 217
Ru-0/CeO2 OER 1.0 M KOH 420 10 218
RuNi OER 1.0 M KOH 327 10 219
Ru–Ni SNs OER 1.0 M KOH 1.45 Onset 220
Ir–O–Co OER 1.0 M KOH 178 10 221
Ir-NR/C OER 1.0 M KOH 296 10 222
Transition metal oxide-based electrocatalysts
RuO2–Fe2O3/HrGO NSs HER 1.0 M KOH 239 97 10 223
Ni3ZnC0.7/NCNT-700 HER 1.0 M KOH 203 91 10 224
Ru/RuO2@N-rGO HER 1.0 M KOH 11 44.2 10 225
Pt2@Ni-rGO HER 1.0 M KOH ∼200 327.9 226
Pd30Ni70/C HER 1.0 M KOH 393 291 227
Pt–Ni (1[thin space (1/6-em)]:[thin space (1/6-em)]1)/rGO HER 1.0 M KOH 147 72 228
LaCoO3@rGO OER 1.0 M KOH 280 104 10 229
MnO2@Co3O4 OER 1.0 M KOH 310 72 20 230
Co3O4–CoS2 OER 1.0 M KOH 280 74 20 231
Cu0.19Mo0.19/Co0.62O OER 1.0 M KOH 250 61 10 232
NiCo2O4/PGR OER 1.0 M KOH 234 110 100 233
Co3O4/PGR OER 1.0 M KOH 242 132 100 233
NiO/PGR OER 1.0 M KOH 272 115 100 233
CuCo2O4/NrGO OER 1.0 M KOH 360 64 10 234
Co@CoO-PNC/CC OER 1.0 M KOH 289 83 10 235
Co3O4/NrGO OER 0.5 M KOH 270 190 236
Transition metal phosphide-based electrocatalysts
Au/CoP@NC-3 HER 1.0 M KOH 140.9 10 237
Co2P HER 1.0 M KOH 280 10 238
CoP/NiCoP NTs HER 1.0 M KOH 39 239
O-CoP HER 1.0 M KOH 98, 310 10 240
CoP–InNC@CNT HER 1.0 M KOH 270 193
Fe–CoP/CoO OER 1.0 M KOH 219 10 241
Mo2C–MoP/NPC OER 1.0 M KOH 320 44.7 10 242
Ce–Co (MoP)/MoP@C OER 1.0 M KOH 287 74.4 10 243
Co–MoP@NCNNS-600 OER 1.0 M KOH 270 60.3 10 244
MoP@Ni3P/NF OER 1.0 M KOH 331 50 10 245
Transition metal nitride, carbide and boride-based electrocatalysts
Mo2C/N–C HER 1.0 M KOH 189 58 246
W2N/W HER 1.0 M KOH 148.5 47.4 10 247
Co–NC@Mo2C HER 1.0 M KOH 99 65 10 248
Ni3Mo3C@NPC NWs/CC HER 1.0 M KOH 215 74.8 100 249
MoNix–MoCx@NC HER 1.0 M KOH 168 71 10 250
Co–TiN@NG-x/CC HER 1.0 M KOH 208 170 10 251
Ni3ZnC0.7/NCNT HER 1/0.1 M KOH 203/380 91/45 224
Ni/MoN@NCNT/CC HER 1.0 M KOH 207 93 10 252
Co4N@NC HER 1.0 M KOH 62 37 10 253
Co–NC@Mo2C OER 1.0 M KOH 347 61 10 248
W2N/WC OER 1.0 M KOH 320 94.5 10 247
2D-NC Fe2Ni2N/rGO NHSs OER 0.1 M KOH 290 49.1 10 254
Co6W6C@NC/CC OER 1.0 M KOH 220 53.96 10 255
Ni/Ni0.2Mo0.8N@N–C OER 1.0 M KOH 260 46 10 256
Co5.47N@N-rGO OER 1.0 M KOH 143 80 257
Ni/MoN@NCNT/CC OER 1.0 M KOH 252 91.8 10 252
Co4N@NC OER 1.0 M KOH 257 58 10 253
Transition metal chalcogenide/dichalcogenide-based electrocatalysts
MoS2/NiSe2/rGO HER 1.0 M KOH 127 73 10 258
MnSxSe1−x@N,F-CQDs HER 1.0 M KOH 129 78.19 10 259
CSCB HER 1.0 M KOH 150 250 35 260
Fe7Se8/CNT/C HER 1.0 M KOH 138 114 10 261
MoSe2/NG-4 HER 1.0 M KOH 120 69 10 262
FeSe2@CoSe2/rGO-2 OER 1.0 M KOH 260 36.3 10 263
FeNi2Se4–NrGO OER 1.0 M KOH 170 62.1 10 264
Co0.85Se/rGO OER 1.0 M KOH 274 62.4 100 265
MoS2/NiSe2/rGO OER 1.0 M KOH 277 107 20 258
Ni0.85Se/rGO OER 1.0 M KOH 280 40.2 100 266
MoS2–NiS2/NGF HER 1.0 M KOH 172 70 10 267
Ni1.5Co0.5@N–C NT/NF HER 1.0 M KOH 114 117 10 268
NPZFNS@C/NF HER 1.0 M KOH 129 66.8 10 269
1TNi0.2Mo0.8S1.8P0.2NS/CC HER 1.0 M KOH 55 61.5 10 270
NiFeOx(OH)y@MoS2/rGO HER 1.0 M KOH 170 80 10 271
CoNi2S4–CNFs HER 1.0 M KOH 228 42.1 20 272
FeNiS–MWCNTs OER 1.0 M KOH 260 50 10 273
Fe–Co9S8@SNC OER 1.0 M KOH 273 55.8 10 274
CoSx@Cu2MoS4 OER 1.0 M KOH 351.4 41.1 10 275
NiS2–MoS2/rGO/N OER 1.0 M KOH 210 58 10 276
MoS2–NiS2/NGF OER 1.0 M KOH 370 13.7 10 267
Ni1.5Co0.5@N–C NT/NF OER 1.0 M KOH 243 103 10 268

Bifunctional electrocatalysts
Catalyst Electrolyte Overpotential HER Overpotential OER Current density (mA cm−2) Ref.
CoFe-LDH/rGO/NF 1.0 M KOH 110 mV 250 mV 10 277
CoNiN@NiFe LDH 1.0 M KOH 150 mV 227 mV 10 278
NiTe@CoFe-LDH 1.0 M KOH 103 mV 218 mV 10 279
Mo2C@g-C3N4@NiMn-LDH 1.0 M KOH 116 mV 290 mV 10 280
Ti3C2–CoS2 276 mV 376 mV 10 281
Mo–CoP 1.0 M KOH 112 mV 330 mV 100 282
Ni-MOF 1.0 M KOH 330 mV 366 mV 10 283
rGO/Ni3Se2/NF 1.0 M KOH 251 mV 292 mV 10 284
P-30-doped Fe/NF 1.0 M KOH 158 mV 284 mV 10 285
Co-CoO@3DHPG 260 mV 1.59 V 10 286
2D MoS2 1.0 M KOH 248 mV 300 mV 10 287
NiCo nanospheres 105 mV 330 mV 10 288


Additionally, MOFs are crystalline materials that can be 1D, 2D, or 3D. They are produced by coordinating metal ions or clusters with rigid organic molecules.289 They possess high crystallinity, mesoporous or microporous structures, and strong metal–ligand interactions. The coordination bond between the metal and the organic linker is stronger than the hydrogen bond, resulting in increased stability of the produced MOFs.290 MOFs often possess high specific surface areas because of their well-organized crystal structures, enabling many active sites necessary for catalytic reactions and enhancing the turnover frequencies (TOFs).291 Li et al.292 selected a neutral N-donor heterocyclic ligand having unsaturated coordination sites to capture MnO4 and CrO42−via the anion exchange technique in pristine MOFs. The Cr@NiCP and Mn@NiCP MOFs, following anion exchange, showed superior HER performance compared to the NiCP MOFs. Unlike classical metal substitution reactions, the adsorption of metal anions through organic ligands may provide more metal active sites, leading to improved synergistic effects and performance in the HER process. Moreover, Ravi Nivethe et al.293 used a customised hydrothermal technique to produce Fe-based MIL-100 MOFs that are extremely porous, having a BET surface area of 2551 m2 g−1. Pristine MOFs feature several scattered metal sites; however, they are not suitable as effective HER active sites because of their saturated coordination.294

5. Challenges faced by transition metals

TM-based electrocatalysts have some key drawbacks, which need to be addressed effectively to make them competitive catalysts compared to noble metal-based electrocatalysts. The intrinsic activity of any material is the most notable and significant factor. For example, Pt and Ir/Ru are excellent catalysts for the HER and OER and highly selective in converting water into hydrogen at the cathodic and oxygen at the anodic applied potential. Durability in practical operation is another crucial aspect that should be considered for TMs to catalyze a reaction for a long time at high electrochemical reaction rates without decreasing their electrocatalytic performance. It should be noted that for electrolysis to be deemed optimal, electrocatalytic oxidation (or corrosion) must be suppressed. A catalyst is more stable when it operates at a continuously applied current density (ideally at current densities >500 mA cm−2) without performance degradation or even under conditions where the current density is reversed during electrolysis.

In water electrolysis devices, noble metal-based catalysts in the form of NPs are frequently placed on conductive supports to avoid stability issues like peeling, deactivation, dissolution, sintering, or poisoning. The ideal catalyst support exhibits exceptional thermal and chemical stability, mechanical strength, and low surface energy.92 These properties can significantly increase the electrochemically active surface area and catalyst efficiency. TMOs face a number of challenges, including intrinsic poor conductivity, slow reaction kinetics, a lack of adsorption sites for H intermediates, and an inappropriate H-binding energy. One of the biggest challenges faced by TMPs is surface oxidation, which happens when the materials are kept in ambient settings. While the OER performance of TMPs may not be impacted by partial oxidation, the oxide formed in situ may reduce conductivity and obstruct HER active areas.295 The slow OER process observed for TMNs, TMCs and TMBs is a significant barrier reducing their use in overall water splitting.160,296 The lack of sufficient surface-active sites is the primary drawback of TM chalcogenides. Numerous approaches have been tried to enrich reactive sites. Initially, considering architecture, a hierarchical porous linked structure could be able to balance mass diffusion pathways, exposed active surfaces, and charge transfer. The proposed HER and OER reaction mechanisms described in the previous literature are still unclear. Thus, to gain more information regarding the dynamic variations in the HER/OER process further systematically, it is important to develop consistent multi-scale fundamental and theoretical measurement models to define analytically the dynamic reaction involvement of the whole catalyst surface, which could lead to the fabrication of high-efficiency electrocatalysts.

6. Conclusions and future perspectives

Development of low-cost, high-efficiency electrocatalysts for energy transformation and renewable energy storage through the electrochemical process is a key research area. To sum up, we have given an overview of the most recent developments in TM-based electrocatalysts for the HER and OER. TM-based electrocatalysts have emerged as promising candidates for catalysing water splitting owing to their unique physiochemical properties and modulated compositions and structures. It has been demonstrated that strategies like surface engineering, combining with other TMs, and even heteroatom dopant incorporation are effective ways to improve the electrocatalytic performance of TM catalysts and provide more electrochemically active surface area to make them industrially effective. More prominent and stable TM-based catalysts will be developed due to a thorough understanding of the reaction mechanisms and the actual reaction active centres, accelerating the development of water electrolysis as an industrial process. Finally, combining theoretical calculations with experimental results can help us find and create more effective and low cost catalysts, given the rapidly developing field of computational chemistry. Theoretical calculations could guide us in choosing and researching appropriate electrocatalytic materials in addition to assisting us in understanding the mechanism. In the realm of electrocatalysis, enhancing the performance and applicability of TM-based electrocatalysts involves a multifaceted approach. Catalyst design and engineering methods, including nano-structuring, alloying, and surface modification, have emerged as pivotal strategies to enhance activity and stability, particularly in alkaline media where avoiding sluggish behaviour is imperative. Exploring novel TM composition configurations and introducing non-metal elements are other strategies for enhancing electrocatalytic performance. Achieving an atomic scale understanding of catalytic processes becomes paramount, guiding the rational design of electrocatalysts with comprehensive improvements. Diversifying materials and architectures, such as metal oxides, nitrides, phosphides, and organic compounds, alongside innovative catalyst topologies like hierarchical structures and core–shell configurations, presents novel opportunities for heightened catalytic performance. Effectively addressing the stability and conductivity of electrocatalysts involves the development of support materials, a crucial aspect in optimizing their functionality. The scalability and cost-effective synthesis of these electrocatalysts are critical for their commercial viability and broader application, especially in alkaline environments. Furthermore, integrating transition metal-based electrocatalysts with energy storage systems, such as water electrolyzers and fuel cells, holds promise for creating sustainable and efficient energy conversion technologies. Amidst these advancements, a critical perspective involves considering the environmental impact and sustainability of electrocatalysts, encompassing raw material extraction, synthesis, and end-of-life disposal. Scheme 2 presents a schematic illustration that depicts future perspectives on TM-based electrocatalysts for the HER and OER.
image file: d3cc06015b-s2.tif
Scheme 2 Schematic illustration reflecting future perspectives on TM-based electrocatalysts toward the HER and OER.

Author contributions

Muhammad Nazim Lakhan and Abdul Hanan: conceptualization, recent progress analysis, writing – reviewing draft, resources; Altaf Hussain: writing – reviewing and editing of text related to factors affecting the hydrogen evolution reaction and the oxygen evolution reaction in alkaline media; Irfan Ali Soomro: writing – reviewing and editing sections dealing with challenges and future perspectives; Yuan Wang: writing – reviewing and editing, supervision, project administration; Mukhtiar Ahmed: writing – reviewing and editing the section dealing with transition metal oxides, formal analysis; Umair Aftab: writing – reviewing and editing, project administration, supervision; Hongyu Sun: writing – reviewing and editing, project administration; and Hamidreza Arandiyan: writing – reviewing and editing, project administration, supervision.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

The authors acknowledge the support from the DCCEEW International Clean Innovation Researcher Networks Grant (ICIRN000011). Y. Wang acknowledges the International Hydrogen Research Fellowship Program financially supported by CSIRO, DCCEEW, the Australian Hydrogen Research Network (AHRN), ARC DECRA (DE230100327), and ARC Linkage Project (LP220200583).

References

  1. Y. A. Bhutto, A. K. Pandey, R. Saidur, I. A. Laghari, D. Buddhi and V. V. Tyagi, IOP Conf. Ser.: Earth Environ. Sci., 2023, 1281, 012013 CrossRef.
  2. G. Makridou, K. Andriosopoulos, M. Doumpos and C. Zopounidis, Energy Policy, 2016, 88, 573–583 CrossRef.
  3. S. Balu, A. Hanan, H. Venkatesvaran, S.-W. Chen, T. C.-K. Yang and M. J. C. Khalid, Catalysts, 2023, 13, 393 CrossRef CAS.
  4. Y. A. Bhutto, A. K. Pandey, R. Saidur, K. Sharma and V. V. Tyagi, Mater. Today Sustainability, 2023, 100443 CrossRef.
  5. J. D. Butson, A. Sharma, J. Tournet, Y. Wang, R. Tatavarti, C. Zhao, C. Jagadish, H. H. Tan and S. Karuturi, Adv. Energy Mater., 2023, 2301793 CrossRef CAS.
  6. G. Elsa, M. Vijayakumar, R. Navaneethan and M. Karthik, Glob. Chall., 2022, 6, 2100139 CrossRef PubMed.
  7. A. Islam, A. K. Pandey, R. Saidur and V. V. Tyagi, Mater. Today Sustainability, 2024, 25, 100678 CrossRef.
  8. Z. Li, H. Feng, M. Song, C. He, W. Zhuang and L. Tian, Mater. Today Energy, 2021, 20, 100698 CrossRef CAS.
  9. Y. Wang, X. Shen, H. Arandiyan, Y. Yin, F. Sun, X. Chen, M. Garbrecht, L. Han, G. G. Andersson and C. Zhao, J. Power Sources, 2020, 478, 228748 CrossRef CAS.
  10. A. Hanan, H. T. A. Awan, F. Bibi, R. R. R. Sulaiman, W. Y. Wong, R. Walvekar, S. Singh and M. Khalid, J. Energy Chem., 2024, 92, 176–206 CrossRef CAS.
  11. H. Arandiyan, S. S. Mofarah, Y. Wang, C. Cazorla, D. Jampaiah, M. Garbrecht, K. Wilson, A. F. Lee, C. Zhao and T. Maschmeyer, Chem. – Eur. J., 2021, 27(58), 14418–14426 CrossRef CAS PubMed.
  12. J. D. Butson, A. Sharma, H. Chen, Y. Wang, Y. Lee, P. Varadhan, M. N. Tsampas, C. Zhao, A. Tricoli, H. H. Tan, C. Jagadish and S. Karuturi, Adv. Energy Mater., 2022, 2102752 CrossRef CAS.
  13. Y. Wang, H. Arandiyan, K. Dastafkan, Y. Li and C. Zhao, Chem. Res. Chin. Univ., 2020, 36, 360–365 CrossRef CAS.
  14. A. Hanan, M. Nazim Lakhan, R. Walvekar, M. Khalid and C. Prakash, Chem. Eng. J., 2024, 483, 149107 CrossRef CAS.
  15. C. Rong, K. Dastafkan, Y. Wang and C. Zhao, Adv. Mater., 2023, e2211884 CrossRef PubMed.
  16. T. Zhao, Y. Wang, S. Karuturi, K. Catchpole, Q. Zhang and C. Zhao, Carbon Energy, 2020, 2, 582–613 CrossRef CAS.
  17. Y. A. Bhutto, A. K. Pandey, R. Saidur, I. A. Laghari, H. Khir, A. Islam and M. A. Zaed, E3S Web of Conferences, 2024, 488, 01007 CrossRef.
  18. A. H. Al-Naggar, N. M. Shinde, J.-S. Kim and R. S. Mane, Coord. Chem. Rev., 2023, 474, 214864 CrossRef CAS.
  19. H. Wang, S. Zhu, J. Deng, W. Zhang, Y. Feng and J. Ma, Chin. Chem. Lett., 2021, 32, 291–298 CrossRef CAS.
  20. D. Zhang, H. Li, A. Riaz, A. Sharma, W. Liang, Y. Wang, H. Chen, K. Vora, D. Yan, Z. Su, A. Tricoli, C. Zhao, F. J. Beck, K. Reuter, K. Catchpole and S. Karuturi, Energy Environ. Sci., 2022, 15, 185–195 RSC.
  21. C. Qiu, F. Cai, Y. Wang, Y. Liu, Q. Wang and C. Zhao, J. Colloid Interface Sci., 2019, 565, 351–359 CrossRef PubMed.
  22. B. Guo, Y. Ding, H. Huo, X. Wen, X. Ren, P. Xu and S. Li, Nano-Micro Lett., 2023, 15, 57 CrossRef CAS PubMed.
  23. M. N. Lakhan, A. Hanan, Y. Wang, S. Liu and H. Arandiyan, Langmuir, 2024, 40, 2465–2486 CrossRef CAS PubMed.
  24. T. Rao, H. Wang, Y.-J. Zeng, Z. Guo, H. Zhang and W. Liao, Adv. Sci., 2021, 8, 2002284 CrossRef CAS PubMed.
  25. Y. A. Bhutto, A. K. Pandey, R. Saidur, B. Aljafari and V. V. Tyagi, J. Energy Storage, 2023, 73, 109116 CrossRef.
  26. A. Islam, A. Sharma, R. Chaturvedi and P. Kumar Singh, Mater. Today: Proc., 2021, 45, 3670–3673 CAS.
  27. F. Wang, Z. Cheng, X. Zhang, C. Xie, F. Liu, C. Chang and G. Liu, 2022, 10.
  28. Z. Tang and H. Yang, in Advanced Nanomaterials for Electrochemical-Based Energy Conversion and Storage, ed. F. Ran and S. Chen, Elsevier, 2020, pp. 355–391 DOI:10.1016/B978-0-12-814558-6.00011-3.
  29. S. Li, E. Li, X. An, X. Hao, Z. Jiang and G. Guan, Nanoscale, 2021, 13, 12788–12817 RSC.
  30. K. C. Majhi and M. Yadav, ACS Eng. Au, 2023, 3, 278–284 CrossRef CAS.
  31. R. R. Raja Sulaiman, W. Y. Wong and K. S. Loh, Int. J. Energy Res., 2022, 46, 2241–2276 CrossRef CAS.
  32. Z.-Y. Yu, Y. Duan, X.-Y. Feng, X. Yu, M.-R. Gao and S.-H. Yu, Adv. Mater., 2021, 33, 2007100 CrossRef CAS PubMed.
  33. Y. Li, L. Zhou and S. Guo, EnergyChem, 2021, 3, 100053 CrossRef CAS.
  34. K. Yin, Z. D. Cui, X. R. Zheng, X. J. Yang, S. L. Zhu, Z. Y. Li and Y. Q. Liang, J. Mater. Chem. A, 2015, 3, 22770–22780 RSC.
  35. X. Deng and H. Tüysüz, ACS Catal., 2014, 4, 3701–3714 CrossRef CAS.
  36. C. Zhu, Q. Shi, S. Feng, D. Du and Y. Lin, ACS Energy Lett., 2018, 3, 1713–1721 CrossRef CAS.
  37. Y. Wang, B. Kong, D. Zhao, H. Wang and C. Selomulya, Nano Today, 2017, 15, 26–55 CrossRef CAS.
  38. J. Jin, J. Yin, H. Liu, M. Lu, J. Li, M. Tian and P. Xi, ChemCatChem, 2019, 11, 2780–2792 CrossRef CAS.
  39. K. Khan, A. K. Tareen, M. Aslam, Y. Zhang, R. Wang, Z. Ouyang, Z. Gou and H. Zhang, Nanoscale, 2019, 11, 21622–21678 RSC.
  40. Y. Guo, T. Park, J. W. Yi, J. Henzie, J. Kim, Z. Wang, B. Jiang, Y. Bando, Y. Sugahara, J. Tang and Y. Yamauchi, Adv. Mater., 2019, 31, 1807134 CrossRef PubMed.
  41. Y. Jiang and Y. Lu, Nanoscale, 2020, 12, 9327–9351 RSC.
  42. X. Xia, L. Wang, N. Sui, V. L. Colvin and W. W. Yu, Nanoscale, 2020, 12, 12249–12262 RSC.
  43. C. Chen, J. Li, Z. Lv, M. Wang and J. Dang, Int. J. Hydrogen Energy, 2023, 48, 30435–30463 CrossRef CAS.
  44. H. Xue, A. Meng, C. Chen, H. Xue, Z. Li and C. Wang, Sci. China Mater., 2022, 65, 712–720 CrossRef CAS.
  45. H. Sun, X. Xu, H. Kim, W. Jung, W. Zhou and Z. Shao, Energy Environ. Mater., 2023, e12441 CrossRef CAS.
  46. C. Hu, L. Zhang and J. Gong, Energy Environ. Sci., 2019, 12, 2620–2645 RSC.
  47. Y. Yan, T. He, B. Zhao, K. Qi, H. Liu and B. Y. Xia, J. Mater. Chem. A, 2018, 6, 15905–15926 RSC.
  48. A. Vazhayil, L. Vazhayal, J. Thomas, S. AshokC and N. Thomas, Appl. Surf. Sci. Adv., 2021, 6, 100184 CrossRef.
  49. Z. Zhang, X. Wu, Z. Kou, N. Song, G. Nie, C. Wang, F. Verpoort and S. Mu, Chem. Eng. J., 2022, 428, 131133 CrossRef CAS.
  50. P. Mao, H. Arandiyan, S. S. Mofarah, P. Koshy, C. Pozo-Gonzalo, R. Zheng, Z. Wang, Y. Wang, S. K. Bhargava, H. Sun, Z. Shao and Y. Liu, Energy Adv., 2023, 2, 465–502 RSC.
  51. Y. Zhou, Z. Zhou, L. Hu, R. Tian, Y. Wang, H. Arandiyan, F. Chen, M. Li, T. Wan, Z. Han, Z. Ma, X. Lu, C. Cazorla, T. Wu and D. Chu, Chem. Eng. J., 2022, 135561,  DOI:10.1016/j.cej.2022.135561.
  52. K. Zhang and R. Zou, Small, 2021, 17, 2100129 CrossRef CAS PubMed.
  53. X. Xu, Y. Pan, Y. Zhong, C. Shi, D. Guan, L. Ge, Z. Hu, Y.-Y. Chin, H.-J. Lin, C.-T. Chen, H. Wang, S. P. Jiang and Z. Shao, Adv. Sci., 2022, 9, 2200530 CrossRef CAS PubMed.
  54. X. Xu, Y. Pan, L. Ge, Y. Chen, X. Mao, D. Guan, M. Li, Y. Zhong, Z. Hu, V. K. Peterson, M. Saunders, C.-T. Chen, H. Zhang, R. Ran, A. Du, H. Wang, S. P. Jiang, W. Zhou and Z. Shao, Small, 2021, 17, 2101573 CrossRef CAS PubMed.
  55. X. Liu, Z. He, M. Ajmal, C. Shi, R. Gao, L. Pan, Z.-F. Huang, X. Zhang and J.-J. Zou, Trans. Tianjin Univ., 2023, 29, 247–253 CrossRef CAS.
  56. Y. Pan, X. Xu, Y. Zhong, L. Ge, Y. Chen, J.-P. M. Veder, D. Guan, R. O’Hayre, M. Li, G. Wang, H. Wang, W. Zhou and Z. Shao, Nat. Commun., 2020, 11, 2002 CrossRef CAS PubMed.
  57. Z.-J. Zhao, S. Liu, S. Zha, D. Cheng, F. Studt, G. Henkelman and J. Gong, Nat. Rev. Mater., 2019, 4, 792–804 CrossRef.
  58. K. Shah, R. Dai, M. Mateen, Z. Hassan, Z. Zhuang, C. Liu, M. Israr, W.-C. Cheong, B. Hu, R. Tu, C. Zhang, X. Chen, Q. Peng, C. Chen and Y. Li, Angew. Chem., 2022, 61, e202114951 CrossRef CAS PubMed.
  59. L. Zhang, J. Xiao, H. Wang and M. Shao, ACS Catal., 2017, 7, 7855–7865 CrossRef CAS.
  60. W. Zhang, L. Cui and J. Liu, J. Alloys Compd., 2020, 821, 153542 CrossRef CAS.
  61. Y. Z. Wang, M. Yang, Y.-M. Ding, N.-W. Li and L. Yu, Adv. Funct. Mater., 2022, 32, 2108681 CrossRef CAS.
  62. M. Y. Solangi, U. Aftab, A. Tahira, A. Hanan, M. Montecchi, L. Pasquali, M. Tonezzer, R. Mazzaro, V. Morandi, A. J. Laghari and A. Nafady, Int. J. Hydrogen Energy, 2023, 36439–36451 CrossRef CAS.
  63. A. Hanan, A. J. Laghari, M. Y. Solangi, U. Aftab, M. I. Abro, D. Cao, M. Ahmed, M. N. Lakhan, A. Ali and A. Asif, Int. J. Eng. Sci. Technol., 2022, 6(1), 1–10 CrossRef.
  64. M. Wang, W. Fu, L. Du, Y. Wei, P. Rao, L. Wei, X. Zhao, Y. Wang and S. Sun, Appl. Surf. Sci., 2020, 515, 146059 CrossRef CAS.
  65. Z. Shi, Z. Rao, J. Zhao, M. Liang, T. Zhu, Y. Wang and H. Arandiyan, Separations, 2021, 8, 247 CrossRef CAS.
  66. S. Anantharaj, P. E. Karthik and S. Noda, Angew. Chem., 2021, 60, 23051–23067 CrossRef CAS PubMed.
  67. Y. Wei, H. Huang, F. Gao and G. Jiang, Int. J. Hydrogen Energy, 2023, 48, 4242–4252 CrossRef CAS.
  68. J. Feng, F. Lv, W. Zhang, P. Li, K. Wang, C. Yang, B. Wang, Y. Yang, J. Zhou, F. Lin, G.-C. Wang and S. Guo, Adv. Mater., 2017, 29, 1703798 CrossRef PubMed.
  69. S. S. Mofarah, L. Schreck, C. Cazorla, X. Zheng, E. Adabifiroozjaei, C. Tsounis, J. Scott, R. Shahmiri, Y. Yao, R. Abbasi, Y. Wang, H. Arandiyan, L. Sheppard, V. Wong, E. Doustkhah, P. Koshy and C. C. Sorrell, Nanoscale, 2021, 13, 6764–6771 RSC.
  70. A. Rategarpanah, F. Meshkani, Y. Wang, H. Arandiyan and M. Rezaei, Energy Convers. Manage., 2018, 166, 268–280 CrossRef CAS.
  71. A. Hanan, M. N. Lakhan, M. Y. Solangi, M. S. AlSalhi, V. Kumar, S. Devanesan, M. A. Shar, M. I. Abro and U. Aftab, Mater. Today Sustainability, 2023, 24, 100585 CrossRef.
  72. A. Muzammil, R. Haider, W. Wei, Y. Wan, M. Ishaq, M. Zahid, W. Yaseen and X. Yuan, Mater. Horiz., 2023, 10, 2764–2799 RSC.
  73. M. Batool, A. Hameed and M. A. Nadeem, Coord. Chem. Rev., 2023, 480, 215029 CrossRef CAS.
  74. N. Hales, T. J. Schmidt and E. Fabbri, Curr. Opin. Electrochem., 2023, 38, 101231 CrossRef CAS.
  75. C. Qiu, S. He, Y. Wang, Q. Wang and C. Zhao, Chem. – Eur. J., 2020, 26, 4120–4127 CrossRef CAS PubMed.
  76. T. Zhao, Y. Wang, X. Chen, Y. Li, Z. Su and C. Zhao, ACS Sustainable Chem. Eng., 2020, 8, 4863–4870 CrossRef CAS.
  77. R. P. Gautam, H. Pan, F. Chalyavi, M. J. Tucker and C. J. Barile, Catal. Sci. Technol., 2020, 10, 4960–4967 RSC.
  78. A. Shahzad, F. Zulfiqar and M. Arif Nadeem, Coord. Chem. Rev., 2023, 477, 214925 CrossRef CAS.
  79. L. Li, X. Cao, J. Huo, J. Qu, W. Chen, C. Liu, Y. Zhao, H. Liu and G. Wang, J. Energy Chem., 2023, 76, 195–213 CrossRef CAS.
  80. X. Zou, Y. Wu, Y. Liu, D. Liu, W. Li, L. Gu, H. Liu, P. Wang, L. Sun and Y. Zhang, Chem, 2018, 4, 1139–1152 CAS.
  81. X. Xu, C. Su, W. Zhou, Y. Zhu, Y. Chen and Z. Shao, Adv. Sci., 2016, 3, 1500187 CrossRef PubMed.
  82. J. Niu, Y. Yue, C. Yang, Y. Wang, J. Qin, X. Zhang and Z.-S. Wu, Appl. Surf. Sci., 2021, 561, 150030 CrossRef CAS.
  83. Q. He, H. Liu, P. Tan, J. Xie, S. Si and J. Pan, J. Solid State Chem., 2021, 299, 122179 CrossRef CAS.
  84. Y. Qiu, M. Sun, J. Cheng, J. Sun, D. Sun and L. Zhang, Catal. Commun., 2022, 164, 106426 CrossRef CAS.
  85. R. Badrnezhad, F. Nasri, H. Pourfarzad and S. K. Jafari, Int. J. Hydrogen Energy, 2021, 46, 3821–3832 CrossRef CAS.
  86. J. Sun, B. Yu, F. Tan, W. Yang, G. Cheng and Z. Zhang, Int. J. Hydrogen Energy, 2022, 47, 15764–15774 CrossRef CAS.
  87. C. Cheng, F. Zheng, C. Zhang, C. Du, Z. Fang, Z. Zhang and W. Chen, J. Power Sources, 2019, 427, 184–193 CrossRef CAS.
  88. Y. Tian, Z. Lin, J. Yu, S. Zhao, Q. Liu, J. Liu, R. Chen, Y. Qi, H. Zhang, R. Li, J. Li and J. Wang, ACS Sustainable Chem. Eng., 2019, 7, 14639–14646 CrossRef CAS.
  89. W. Xiong, H. Yin, T. Wu and H. Li, Chem. – Eur. J., 2023, 29, e202202872 CrossRef CAS PubMed.
  90. D. Liu, G. Xu, H. Yang, H. Wang and B. Y. Xia, Adv. Funct. Mater., 2023, 33, 2208358 CrossRef CAS.
  91. Z. H. Tan, X. Y. Kong, B.-J. Ng, H. S. Soo, A. R. Mohamed and S.-P. Chai, ACS Omega, 2023, 8, 1851–1863 CrossRef CAS PubMed.
  92. L. Tian, Z. Li, X. Xu and C. Zhang, J. Mater. Chem. A, 2021, 9, 13459–13470 RSC.
  93. Y. Li, Y. Sun, Y. Qin, W. Zhang, L. Wang, M. Luo, H. Yang and S. Guo, Adv. Energy Mater., 2020, 10, 1903120 CrossRef CAS.
  94. T. Noor, L. Yaqoob and N. Iqbal, ChemCatChem, 2021, 8, 447–483 CAS.
  95. H. Zhang, P. An, W. Zhou, B. Y. Guan, P. Zhang, J. Dong and X. W. Lou, Sci. Adv., 2018, 4, eaao6657 CrossRef PubMed.
  96. L. Li, L. Bu, B. Huang, P. Wang, C. Shen, S. Bai, T.-S. Chan, Q. Shao, Z. Hu and X. Huang, Adv. Mater., 2021, 33, 2105308 CrossRef CAS PubMed.
  97. G. Liu, F. Hou, Y. Wang, X. Wang and B. Fang, Appl. Surf. Sci., 2023, 637, 157896 CrossRef CAS.
  98. R. Manjunatha, L. Dong, Z. Zhai, J. Wang, Q. Fu, W. Yan and J. Zhang, Green Energy Environ., 2022, 7, 933–947 CrossRef CAS.
  99. Y. Zhang, W. Liao and G. Zhang, J. Power Sources, 2021, 512, 230514 CrossRef CAS.
  100. R. A. Qureshi, A. Hanan, M. I. Abro, M. S. AlSalhi, M. A. Qureshi, M. Y. Solangi, S. Devanesan, M. A. Shar and U. Aftab, Mater. Today Sustainability, 2023, 23, 100446 CrossRef.
  101. A. Ali, F. Long and P. K. Shen, Electrochem. Energy Rev., 2022, 5, 1 CrossRef CAS.
  102. J. Jayabharathi, B. Karthikeyan, B. Vishnu and S. Sriram, Phys. Chem. Chem. Phys., 2023, 25, 8992–9019 RSC.
  103. Y. Hou, X. Zhuang and X. Feng, Small Methods, 2017, 1, 1700090 CrossRef.
  104. H. Arandiyan, S. S. Mofarah, C. C. Sorrell, E. Doustkhah, B. Sajjadi, D. Hao, Y. Wang, H. Sun, B.-J. Ni, M. Rezaei, Z. Shao and T. Maschmeyer, Chem. Soc. Rev., 2021, 50, 10116–10211 RSC.
  105. K. Min, M. Kim, S. Min, H. Kim and S.-H. Baeck, Appl. Surf. Sci., 2023, 624, 157117 CrossRef CAS.
  106. W. Xu, F. Lyu, Y. Bai, A. Gao, J. Feng, Z. Cai and Y. Yin, Nano Energy, 2018, 43, 110–116 CrossRef CAS.
  107. Z. Cai, X. Bu, P. Wang, W. Su, R. Wei, J. C. Ho, J. Yang and X. Wang, J. Mater. Chem. A, 2019, 7, 21722–21729 RSC.
  108. J. Zhang, A. Wei, J. Liu, J. Zhu, Y. He and Z. Liu, J. Alloys Compd., 2022, 927, 166990 CrossRef CAS.
  109. A. Hanan, M. Y. Solangi, A. Jaleel Laghari, A. A. Shah, U. Aftab, Z. A. Ibupoto, M. I. Abro, M. N. Lakhan, I. A. Soomro, E. A. Dawi, A. Al Karim Haj Ismail, E. Mustafa, B. Vigolo, A. Tahira and Z. H. Ibupoto, RSC Adv., 2023, 13, 743–755 RSC.
  110. R. Elakkiya, R. Ramkumar and G. Maduraiveeran, Mater. Res. Bull., 2019, 116, 98–105 CrossRef CAS.
  111. A. Hanan, D. Shu, U. Aftab, D. Cao, A. J. Laghari, M. Y. Solangi, M. I. Abro, A. Nafady, B. Vigolo and A. Tahira, Int. J. Hydrogen Energy, 2022, 47, 33919–33937 CrossRef CAS.
  112. K. Srinivas, Y. Chen, Z. Su, B. Yu, M. Karpuraranjith, F. Ma, X. Wang, W. Zhang and D. Yang, Electrochim. Acta, 2022, 404, 139745 CrossRef CAS.
  113. Y. Wang, B. Liu, X. Shen, H. Arandiyan, T. Zhao, Y. Li, M. Garbrecht, Z. Su, L. Han, A. Tricoli and C. Zhao, Adv. Energy Mater., 2021, 2003759 CrossRef CAS.
  114. U. Aftab, A. Tahira, A. H. Samo, M. I. Abro, M. M. Baloch, M. Kumar Sirajuddin and Z. H. Ibupoto, Int. J. Hydrogen Energy, 2020, 45, 13805–13813 CrossRef CAS.
  115. W. Xin, B. Liu, Y. Zhao, G. Chen, P. Chen, Y. Zhou, W. Li, Y. Xu, Y. Zhong and Y. A. Nikolaevich, Electrochim. Acta, 2022, 404, 139748 CrossRef CAS.
  116. M. Y. Solangi, U. Aftab, A. Tahira, M. I. Abro, R. Mazarro, V. Morandi, A. Nafady, S. S. Medany, A. Infantes-Molina and Z. H. Ibupoto, Int. J. Hydrogen Energy, 2022, 47, 3834–3845 CrossRef CAS.
  117. R. V. Digraskar, V. S. Sapner, A. V. Ghule and B. R. Sathe, ACS Omega, 2019, 4, 18969–18977 CrossRef CAS PubMed.
  118. Y. Li, H. Xu, P. Yang, R. Li, D. Wang, P. Ren, S. Ji, X. Lu, F. Meng, J. Zhang and M. An, Mater. Today Energy, 2022, 25, 100975 CrossRef CAS.
  119. A. Hanan, M. N. Lakhan, D. Shu, A. Hussain, M. Ahmed, I. A. Soomro, V. Kumar and D. Cao, Int. J. Hydrogen Energy, 2023, 48, 19494–19508 CrossRef CAS.
  120. A. J. Laghari, U. Aftab, A. Tahira, A. A. Shah, A. Gradone, M. Y. Solangi, A. H. Samo, M. Kumar, M. I. Abro, M. W. Akhtar, R. Mazzaro, V. Morandi, A. M. Alotaibi, A. Nafady, A. Infantes-Molina and Z. H. Ibupoto, Int. J. Hydrogen Energy, 2023, 48, 12672–12682 CrossRef CAS.
  121. A. Hanan, M. Ahmed, M. N. Lakhan, A. H. Shar, D. Cao, A. Asif, A. Ali and M. Gul, J. Indian Chem. Soc., 2022, 99, 100442 CrossRef CAS.
  122. A. Y. Faid, A. O. Barnett, F. Seland and S. Sunde, Electrochim. Acta, 2020, 361, 137040 CrossRef CAS.
  123. Y. Jiang, C. Y. Toe, S. S. Mofarah, C. Cazorla, S. L. Y. Chang, Y. Yin, Q. Zhang, S. Lim, Y. Yao, R. Tian, Y. Wang, T. Zaman, H. Arandiyan, G. G. Andersson, J. Scott, P. Koshy, D. Wang and C. C. Sorrell, ACS Sustainable Chem. Eng., 2023, 11, 3370–3389 CrossRef CAS.
  124. H. Fan, P. Mao, H. Sun, Y. Wang, S. S. Mofarah, P. Koshy, H. Arandiyan, Z. Wang, Y. Liu and Z. Shao, Mater. Horiz., 2022, 9, 524–546 RSC.
  125. P. Shen, B. Zhang, Y. Wang, X. Liu, C. Yu, T. Xu, S. S. Mofarah, Y. Yu, Y. Liu, H. Sun and H. Arandiyan, J. Nanostruct. Chem., 2021, 11, 33–68 CrossRef.
  126. X. Sun, P. Wei, Z. He, F. Cheng, L. Tang, Q. Li, J. Han and J. He, J. Alloys Compd., 2023, 932, 167253 CrossRef CAS.
  127. P. Wei, X. Sun, Z. He, F. Cheng, J. Xu, Q. Li, Y. Ren, J. He, J. Han and Y. Huang, Fuel, 2023, 339, 127303 CrossRef CAS.
  128. T. Zhao, X. Shen, Y. Wang, R. K. Hocking, Y. Li, C. Rong, K. Dastafkan, Z. Su and C. Zhao, Adv. Funct. Mater., 2021, 2100614 CrossRef CAS.
  129. D. Zhang, H. Li, A. Riaz, A. Sharma, W. Liang, Y. Wang, H. Chen, K. Vora, D. Yan, Z. Su, A. Tricoli, C. Zhao, F. J. Beck, K. Reuter, K. Catchpole and S. Karuturi, Energy Environ. Sci., 2022, 15, 185–195 RSC.
  130. W. Zhang, P. Shen, L. Qian, Y. Wang, H. Arandiyan, H. Mahmoud, X. Liu, J. Wang, X. Wang, Y. Liu, H. Sun and Y. Yu, ACS Appl. Energy Mater., 2021, 4(5), 4551–4560 CrossRef CAS.
  131. D. Salvò, D. Mosconi, A. Neyman, M. Bar-Sadan, L. Calvillo, G. Granozzi, M. Cattelan and S. Agnoli, Nanomaterials, 2023, 13, 683 CrossRef PubMed.
  132. S. S. A. Shah, N. A. Khan, M. Imran, M. Rashid, M. K. Tufail, A. U. Rehman, G. Balkourani, M. Sohail, T. Najam and P. Tsiakaras, Membranes, 2023, 13, 113 CrossRef CAS PubMed.
  133. G. Zhou, M. Li, Y. Li, H. Dong, D. Sun, X. Liu, L. Xu, Z. Tian and Y. Tang, Adv. Funct. Mater., 2020, 30 Search PubMed.
  134. D. Chen, Z. Pu, R. Lu, P. Ji, P. Wang, J. Zhu, C. Lin, H.-W. Li, X. Zhou, Z. Hu, F. Xia, J. Wu and S. Mu, Adv. Energy Mater., 2020, 10, 2000814 CrossRef CAS.
  135. Y. Hao, Y. Xu, W. Liu and X. Sun, Mater. Horiz., 2018, 5, 108–115 RSC.
  136. G. Lan, H. Fan, Y. Wang, H. Arandiyan, S. K. Bhargava, Z. Shao, H. Sun and Y. Liu, New J. Chem., 2023, 47, 21969–21977 RSC.
  137. W. H. Yun, G. Das, B. Kim, B. J. Park, H. H. Yoon and Y. S. Yoon, Sci. Rep., 2021, 11, 22003 CrossRef CAS PubMed.
  138. H. Liang, A. N. Gandi, D. H. Anjum, X. Wang, U. Schwingenschlögl and H. N. Alshareef, Nano Lett., 2016, 16, 7718–7725 CrossRef CAS PubMed.
  139. L. Shen, Q. Li, L. Yi, S. Tang, L. Yu, Q. Zhang, Q. Huang, T. Zhou and L. Zhang, Solid State Sci., 2021, 121, 106731 CrossRef CAS.
  140. Y. Wang, H. Arandiyan, X. Chen, T. Zhao, X. Bo, Z. Su and C. Zhao, J. Phys. Chem. C, 2020, 124, 9971–9978 CrossRef CAS.
  141. S. Kumaravel, K. Karthick, S. Sam Sankar, A. Karmakar, R. Madhu, K. Bera and S. Kundu, ChemElectroChem, 2021, 8, 4638–4685 CrossRef CAS.
  142. D. Duan, D. Guo, J. Gao, S. Liu and Y. Wang, J. Colloid Interface Sci., 2022, 622, 250–260 CrossRef CAS PubMed.
  143. K. Wu, Z. Chen, W.-C. Cheong, S. Liu, W. Zhu, X. Cao, K. Sun, Y. Lin, L. Zheng, W. Yan, Y. Pan, D. Wang, Q. Peng, C. Chen and Y. Li, ACS Appl. Mater. Interfaces, 2018, 10, 44201–44208 CrossRef CAS PubMed.
  144. C. G. Read, J. F. Callejas, C. F. Holder and R. E. Schaak, ACS Appl. Mater. Interfaces, 2016, 8, 12798–12803 CrossRef CAS PubMed.
  145. K. Liu, C. Zhang, Y. Sun, G. Zhang, X. Shen, F. Zou, H. Zhang, Z. Wu, E. C. Wegener, C. J. Taubert, J. T. Miller, Z. Peng and Y. Zhu, ACS Nano, 2018, 12, 158–167 CrossRef CAS PubMed.
  146. D. Chen, R. Lu, Z. Pu, J. Zhu, H.-W. Li, F. Liu, S. Hu, X. Luo, J. Wu, Y. Zhao and S. Mu, Appl. Catal., B, 2020, 279, 119396 CrossRef CAS.
  147. X. Wang, B. Wang, Y. Chen, M. Wang, Q. Wu, K. Srinivas, B. Yu, X. Zhang, F. Ma and W. Zhang, J. Mater. Sci. Technol., 2022, 105, 266–273 CrossRef CAS.
  148. J. Lin, Y. Yan, C. Li, X. Si, H. Wang, J. Qi, J. Cao, Z. Zhong, W. Fei and J. Feng, Nano-Micro Lett., 2019, 11, 55 CrossRef CAS PubMed.
  149. K. N. Dinh, Q. Liang, C.-F. Du, J. Zhao, A. I. Y. Tok, H. Mao and Q. Yan, Nano Today, 2019, 25, 99–121 CrossRef CAS.
  150. L. Lin, S. Piao, Y. Choi, L. Lyu, H. Hong, D. Kim, J. Lee, W. Zhang and Y. Piao, EnergyChem, 2022, 4, 100072 CrossRef CAS.
  151. J. Xie and Y. Xie, Chem. – Eur. J., 2016, 22, 3588–3598 CrossRef CAS PubMed.
  152. R. Jamil, R. Ali, S. Loomba, J. Xian, M. Yousaf, K. Khan, B. Shabbir, C. F. McConville, A. Mahmood and N. Mahmood, Chem. Catal., 2021, 1, 802–854 CrossRef CAS.
  153. C. Panda, P. W. Menezes, M. Zheng, S. Orthmann and M. Driess, ACS Energy Lett., 2019, 4, 747–754 CrossRef CAS.
  154. Y. Fang, Y. R. Xue, L. Hui, H. D. Yu, Y. X. Liu, C. Y. Xing, F. S. Lu, F. He, H. B. Liu and Y. L. Li, Nano Energy, 2019, 59, 591–597 CrossRef CAS.
  155. A. Wu, Y. Xie, H. Ma, C. Tian, Y. Gu, H. Yan, X. Zhang, G. Yang and H. Fu, Nano Energy, 2018, 44, 353–363 CrossRef CAS.
  156. Y. Lu, Z. Li, Y. Xu, L. Tang, S. Xu, D. Li, J. Zhu and D. Jiang, Chem. Eng. J., 2021, 411, 128433 CrossRef CAS.
  157. J. Chen, B. Ren, H. Cui and C. Wang, Small, 2020, 16, 1907556 CrossRef CAS PubMed.
  158. B. Chang, J. Yang, Y. Shao, L. Zhang, W. Fan, B. Huang, Y. Wu and X. Hao, ChemSusChem, 2018, 11, 3198–3207 CrossRef CAS PubMed.
  159. G. Raj, D. Das, B. Sarkar, S. Biswas and K. K. Nanda, Sustainable Mater. Technol., 2022, 33, e00451 CrossRef CAS.
  160. P. Chen, J. Ye, H. Wang, L. Ouyang and M. Zhu, J. Alloys Compd., 2021, 883, 160833 CrossRef CAS.
  161. Y. Yu, J. Zhou and Z. Sun, Adv. Funct. Mater., 2020, 30, 2070311 CrossRef CAS.
  162. S. Roy, D. Bagchi, L. Dheer, S. C. Sarma, V. Rajaji, C. Narayana, U. V. Waghmare and S. C. Peter, Appl. Catal., B, 2021, 298, 120560 CrossRef CAS.
  163. C. Giordano, C. Erpen, W. Yao, B. Milke and M. Antonietti, Chem. Mater., 2009, 21, 5136–5144 CrossRef CAS.
  164. M. Li, Y. Zhu, H. Wang, C. Wang, N. Pinna and X. Lu, Adv. Energy Mater., 2019, 9, 1803185 CrossRef.
  165. Z. Kou, W. Zang, W. Pei, L. Zheng, S. Zhou, S. Zhang, L. Zhang and J. Wang, J. Mater. Chem. A, 2020, 8, 3071–3082 RSC.
  166. O. Rabi, E. Pervaiz, M. Ali and R. Zahra, Ceram. Int., 2021, 47, 24949–24958 CrossRef CAS.
  167. R. Ge, J. Zhao, J. Huo, J. Qu, J. Yang, Y. Li, M. Zhu, J. M. Cairney, R. Zheng, S. Li, J. Zhang, B. Liu and W. Li, Mater. Today Nano, 2022, 20, 100248 CrossRef CAS.
  168. J. Lao, D. Li, C. Jiang, R. Luo, H. Peng, R. Qi, H. Lin, R. Huang, G. I. N. Waterhouse and C. Luo, Int. J. Hydrogen Energy, 2020, 45, 28616–28625 CrossRef CAS.
  169. D. Wang, Y. Song, H. Zhang, X. Yan and J. Guo, J. Electroanal. Chem., 2020, 861, 113953 CrossRef CAS.
  170. A. Chunduri, S. Gupta, O. Bapat, A. Bhide, R. Fernandes, M. K. Patel, V. Bambole, A. Miotello and N. Patel, Appl. Catal., B, 2019, 259, 118051 CrossRef CAS.
  171. Q. Hu, G. Li, Z. Han, Z. Wang, X. Huang, X. Chai, Q. Zhang, J. Liu and C. He, 2019, 9, 1901130.
  172. Y. Li, B. Huang, Y. Sun, M. Luo, Y. Yang, Y. Qin, L. Wang, C. Li, F. Lv, W. Zhang and S. Guo, Small, 2019, 15, 1804212 CrossRef PubMed.
  173. E. C. Okpara, O. C. Olatunde, O. B. Wojuola and D. C. Onwudiwe, Environ. Adv., 2023, 11, 100341 CrossRef CAS.
  174. A. Silva, J. Cao, T. Polcar and D. Kramer, Chem. Mater., 2022, 34, 10279–10290 CrossRef CAS PubMed.
  175. E. Chen, W. Xu, J. Chen and J. H. Warner, Mater. Today Adv., 2020, 7, 100076 CrossRef.
  176. M. Wang, L. Zhang, Y. He and H. Zhu, J. Mater. Chem. A, 2021, 9, 5320–5363 RSC.
  177. F. Li, R. Xu, Y. Li, F. Liang, D. Zhang, W.-F. Fu and X.-J. Lv, Carbon, 2019, 145, 521–528 CrossRef CAS.
  178. L. Ma, J. Liang, T. Chen, Y. Liu, S. Li and G. Fang, Electrochim. Acta, 2019, 326, 135002 CrossRef CAS.
  179. D. Kong, Y. Wang, S. Huang, Y. V. Lim, M. Wang, T. Xu, J. Zang, X. Li and H. Y. Yang, J. Colloid Interface Sci., 2022, 607, 1876–1887 CrossRef CAS PubMed.
  180. S. Wang, Y. Liu, X. Deng, J. Cao, Y. Han, J. Huang and Y. Li, J. Alloys Compd., 2022, 925, 166787 CrossRef CAS.
  181. G. Liu, C. Shuai, Z. Mo, R. Guo, N. Liu, X. Niu, Q. Dong, J. Wang, Q. Gao, Y. Chen and W. Liu, New J. Chem., 2020, 44, 17313–17322 RSC.
  182. X. Peng, Y. Yan, X. Jin, C. Huang, W. Jin, B. Gao and P. K. Chu, Nano Energy, 2020, 78, 105234 CrossRef CAS.
  183. J. Wang, J. Huang, G. Chen, W. Chen, T. Li, A. Meng and K. Ostrikov, Chem. Eng. J., 2022, 446, 137297 CrossRef CAS.
  184. C. Zhou, H. Wu, F. Zhang and Y. Miao, Crystals, 2022, 12, 666 CrossRef CAS.
  185. K. Wang, Z. Lin, Y. Tang, Z. Tang, C.-L. Tao, D.-D. Qin and Y. Tian, Electrochim. Acta, 2021, 368, 137584 CrossRef CAS.
  186. Y. Pan, M. Wang, M. Li, G. Sun, Y. Chen, Y. Liu, W. Zhu and B. Wang, J. Energy Chem., 2022, 68, 699–708 CrossRef CAS.
  187. C. Gong, W. Li, Y. Lei, X. He, H. Chen, X. Du, W. Fang, D. Wang and L. Zhao, Composites, Part B, 2022, 236, 109823 CrossRef CAS.
  188. U. Aftab, M. Y. Solangi, A. Tahira, A. Hanan, M. I. Abro, A. Karsy, E. Dawi, M. A. Bhatti, R. H. Alshammari, A. Nafady, A. Gradone, R. Mazzaro, V. Morandi, A. Infantes-Molina and Z. H. Ibupoto, RSC Adv., 2023, 13, 32413–32423 RSC.
  189. A. B. Maghirang, Z.-Q. Huang, R. A. B. Villaos, C.-H. Hsu, L.-Y. Feng, E. Florido, H. Lin, A. Bansil and F.-C. Chuang, npj 2D Mater. Appl., 2019, 3, 35 CrossRef.
  190. A. Bagheri, J. Yeow, H. Arandiyan, J. Xu, C. Boyer and M. Lim, Macromol. Rapid Commun., 2016, 37, 905–910 CrossRef CAS PubMed.
  191. Y. Zhu, L. Song, N. Song, M. Li, C. Wang and X. Lu, ACS Sustainable Chem. Eng., 2019, 7, 2899–2905 CrossRef CAS.
  192. L. Sun, H. Xu, Z. Cheng, D. Zheng, Q. Zhou, S. Yang and J. Lin, Chem. Eng. J., 2022, 443, 136348 CrossRef CAS.
  193. S. J. Patil, N. R. Chodankar, S.-K. Hwang, P. A. Shinde, G. Seeta Rama Raju, K. Shanmugam Ranjith, Y. S. Huh and Y.-K. Han, Chem. Eng. J., 2022, 429, 132379 CrossRef CAS.
  194. Y. Lv, Y. Liu, Y. Liu, Z. Chen and M. Zhang, J. Alloys Compd., 2018, 768, 889–895 CrossRef CAS.
  195. M. Zou, J. Chen, L. Xiao, H. Zhu, T. Yang, M. Zhang and M. Du, J. Mater. Chem. A, 2015, 3, 18090–18097 RSC.
  196. D. Gao, B. Xia, Y. Wang, W. Xiao, P. Xi, D. Xue and J. Ding, Small, 2018, 14, 1704150 CrossRef PubMed.
  197. L. Huang, Z. Li, S. Sun, G. Sun, Y. Li, S. Han and J. Lian, J. Alloys Compd., 2022, 926, 166870 CrossRef CAS.
  198. S. Rijith, S. Abhilash, S. Sarika, V. S. Sumi and C. O. Sreekala, Int. J. Hydrogen Energy, 2023, 48, 5783–5800 CrossRef CAS.
  199. T. Niyitanga and H. K. Jeong, J. Electroanal. Chem., 2019, 849, 113383 CrossRef CAS.
  200. H. Gao, J. Zang, Y. Wang, S. Zhou, P. Tian, S. Song, X. Tian and W. Li, Electrochim. Acta, 2021, 377, 138051 CrossRef CAS.
  201. P.-Z. Li, N. Chen, A. Al-Hamry, E. Sheremet, R. Lu, Y. Yang, O. Kanoun, R. R. Baumann, R. D. Rodriguez and J.-J. Chen, Chem. Eng. J., 2023, 457, 141289 CrossRef CAS.
  202. R. K. Rai, B. Sarkar, R. Ram, K. K. Nanda and N. Ravishankar, Sustainable Energy Fuels, 2022, 6, 1708–1718 RSC.
  203. H. Arandiyan, Y. Wang, C. M. A. Parlett and A. F. Lee, Heterogeneous Catalysts, 2021, pp. 161–181 DOI:10.1002/9783527813599.ch9.
  204. Q. Xiong, Y. Wang, P.-F. Liu, L.-R. Zheng, G. Wang, H.-G. Yang, P.-K. Wong, H. Zhang and H. Zhao, Adv. Mater., 2018, 30, 1801450 CrossRef PubMed.
  205. W. Wu, Z. Tang, K. Wang, Z. Liu, L. Li and S. Chen, Electrochim. Acta, 2018, 260, 168–176 CrossRef CAS.
  206. Q. Chen, Z. Cao, G. Du, Q. Kuang, J. Huang, Z. Xie and L. Zheng, Nano Energy, 2017, 39, 582–589 CrossRef CAS.
  207. H. Y. Kim, J. M. Kim, Y. Ha, J. Woo, A. Byun, T. J. Shin, K. H. Park, H. Y. Jeong, H. Kim, J. Y. Kim and S. H. Joo, ACS Catal., 2019, 9, 11242–11254 CrossRef CAS.
  208. X.-Y. Huang, L.-X. You, X.-F. Zhang, J.-J. Feng, L. Zhang and A.-J. Wang, Electrochim. Acta, 2019, 299, 89–97 CrossRef CAS.
  209. Q. Zhang, Y. Kuang, Y. Li, M. Jiang, Z. Cai, Y. Pang, Z. Chang and X. Sun, J. Mater. Chem. A, 2019, 7, 9517–9522 RSC.
  210. X. Qin, L. Zhang, G.-L. Xu, S. Zhu, Q. Wang, M. Gu, X. Zhang, C. Sun, P. B. Balbuena, K. Amine and M. Shao, ACS Catal., 2019, 9, 9614–9621 CrossRef CAS.
  211. P. Jiang, H. Huang, J. Diao, S. Gong, S. Chen, J. Lu, C. Wang, Z. Sun, G. Xia, K. Yang, Y. Yang, L. Wei and Q. Chen, Appl. Catal., B, 2019, 258, 117965 CrossRef CAS.
  212. G. Liu, W. Zhou, B. Chen, Q. Zhang, X. Cui, B. Li, Z. Lai, Y. Chen, Z. Zhang, L. Gu and H. Zhang, Nano Energy, 2019, 66, 104173 CrossRef CAS.
  213. Y. Li, W. Pei, J. He, K. Liu, W. Qi, X. Gao, S. Zhou, H. Xie, K. Yin, Y. Gao, J. He, J. Zhao, J. Hu, T.-S. Chan, Z. Li, G. Zhang and M. Liu, ACS Catal., 2019, 9, 10870–10875 CrossRef CAS.
  214. Q. Yan, P. Yan, T. Wei, G. Wang, K. Cheng, K. Ye, K. Zhu, J. Yan, D. Cao and Y. Li, J. Mater. Chem. A, 2019, 7, 2831–2837 RSC.
  215. C. Guo, Y. Jiao, Y. Zheng, J. Luo, K. Davey and S.-Z. Qiao, Chem, 2019, 5, 2429–2441 CAS.
  216. D. H. Kweon, M. S. Okyay, S.-J. Kim, J.-P. Jeon, H.-J. Noh, N. Park, J. Mahmood and J.-B. Baek, Nat. Commun., 2020, 11, 1278 CrossRef CAS PubMed.
  217. C. Gao, H. Wang, S. Li, B. Liu, J. Yang, J. Gao, Z. Peng, Z. Zhang and Z. Liu, Electrochim. Acta, 2019, 327, 134958 CrossRef CAS.
  218. E. Demir, S. Akbayrak, A. M. Onal and S. Ozkar, J. Colloid Interface Sci., 2019, 534, 704–710 CrossRef CAS PubMed.
  219. N. Liu, Z. Zhai, B. Yu, W. Yang, G. Cheng and Z. Zhang, Int. J. Hydrogen Energy, 2022, 47, 31330–31341 CrossRef CAS.
  220. J. Ding, Q. Shao, Y. Feng and X. Huang, Nano Energy, 2018, 47, 1–7 CrossRef CAS.
  221. C. Wang, P. Zhai, M. Xia, Y. Wu, B. Zhang, Z. Li, L. Ran, J. Gao, X. Zhang, Z. Fan, L. Sun and J. Hou, Angew. Chem. Int. Ed., 2021, 60, 27126–27134 CrossRef CAS PubMed.
  222. F. Luo, L. Guo, Y. Xie, J. Xu, K. Qu and Z. Yang, Appl. Catal., B, 2020, 279, 119394 CrossRef CAS.
  223. H. Mosallaei, H. Hadadzadeh, A. A. Ensafi, K. Z. Mousaabadi, M. Weil, A. Foelske and M. Sauer, Int. J. Hydrogen Energy, 2023, 48, 1813–1830 CrossRef CAS.
  224. R. Li, X. Li, D. Yu, L. Li, G. Yang, K. Zhang, S. Ramakrishna, L. Xie and S. Peng, Carbon, 2019, 148, 496–503 CrossRef CAS.
  225. X. Gao, J. Chen, X. Sun, B. Wu, B. Li, Z. Ning, J. Li and N. Wang, ACS Appl. Nano Mater., 2020, 3, 12269–12277 CrossRef CAS.
  226. W. Xu, J. Chang, Y. Cheng, H. Liu, J. Li, Y. Ai, Z. Hu, X. Zhang, Y. Wang, Q. Liang, Y. Yang and H. Sun, Nano Res., 2022, 15, 965–971 CrossRef CAS.
  227. J. J. de la Cruz-Cruz, M. A. Domínguez-Crespo, E. Ramírez-Meneses, A. M. Torres-Huerta, S. B. Brachetti-Sibaja, A. E. Rodríguez-Salazar, E. Pastor and L. E. González-Sánchez, Data Brief, 2022, 42, 108256 CrossRef CAS PubMed.
  228. Z. Du, Y. Wang, J. Li and J. Liu, J. Nanopart. Res., 2019, 21, 13 CrossRef.
  229. J. Ahmed, T. Ahamad, N. Alhokbany, M. A. Majeed Khan, P. Arunachalam, M. S. Amer, R. M. Alotaibi and S. M. Alshehri, J. Ind. Eng. Chem., 2023, 121, 100–106 CrossRef CAS.
  230. M. Y. Solangi, A. H. Samo, A. J. Laghari, U. Aftab, M. I. Abro and M. I. Irfan, Sukkur IBA Journal of Emerging Technologies, 2022, 5, 32–40 CrossRef.
  231. A. Samo, U. Aftab, D. Cao, M. Ahmed, M. Lakhan, V. Kumar, A. Asif and A. Ali, Digest J. Nanomater. Biostruct., 2022, 17, 109–120 CrossRef.
  232. T. Niyitanga and H. Kim, Chem. Eng. J., 2023, 451, 138649 CrossRef CAS.
  233. Z. H. Ibupoto, A. Tahira, A. A. Shah, U. Aftab, M. Y. Solangi, J. A. Leghari, A. H. Samoon, A. L. Bhatti, M. A. Bhatti, R. Mazzaro, V. Morandi, M. I. Abro, A. Nafady, A. M. Al-Enizi, M. Emo and B. Vigolo, Int. J. Hydrogen Energy, 2022, 47, 6650–6665 CrossRef CAS.
  234. S. K. Bikkarolla and P. Papakonstantinou, J. Power Sources, 2015, 281, 243–251 CrossRef CAS.
  235. X. Fan, Y. Fan, X. Zhang, L. Tang and J. Guo, J. Alloys Compd., 2021, 877, 160279 CrossRef CAS.
  236. A. V. Munde, B. B. Mulik, R. P. Dighole and B. R. Sathe, New J. Chem., 2020, 44, 15776–15784 RSC.
  237. X. Wang, Y. Fei, W. Li, L. Yi, B. Feng, Y. Pan, W. Hu and C. M. Li, ACS Appl. Mater. Interfaces, 2020, 12, 16548–16556 CrossRef CAS PubMed.
  238. H. Li, Q. Li, P. Wen, T. B. Williams, S. Adhikari, C. Dun, C. Lu, D. Itanze, L. Jiang, D. L. Carroll, G. L. Donati, P. M. Lundin, Y. Qiu and S. M. Geyer, Adv. Mater., 2018, 30, 1705796 CrossRef PubMed.
  239. Y. Lin, K. Sun, S. Liu, X. Chen, Y. Cheng, W.-C. Cheong, Z. Chen, L. Zheng, J. Zhang, X. Li, Y. Pan and C. Chen, Adv. Energy Mater., 2019, 9, 1901213 CrossRef.
  240. G. Zhou, M. Li, Y. Li, H. Dong, D. Sun, X. Liu, L. Xu, Z. Tian and Y. Tang, Adv. Funct. Mater., 2020, 30, 1905252 CrossRef CAS.
  241. X. Hu, S. Zhang, J. Sun, L. Yu, X. Qian, R. Hu, Y. Wang, H. Zhao and J. Zhu, Nano Energy, 2019, 56, 109–117 CrossRef CAS.
  242. E. Jiang, J. Li, X. Li, A. Ali, G. Wang, S. Ma, P. Kang Shen and J. Zhu, Chem. Eng. J., 2022, 431, 133719 CrossRef CAS.
  243. T. Chen, Y. Fu, W. Liao, Y. Zhang, M. Qian, H. Dai, X. Tong and Q. Yang, Energy Fuels, 2021, 35, 14169–14176 CrossRef CAS.
  244. N. Li, Y. Guan, Y. Li, H. Mi, L. Deng, L. Sun, Q. Zhang, C. He and X. Ren, J. Mater. Chem. A, 2021, 9, 5868 RSC.
  245. F. Wang, J. Chen, X. Qi, H. Yang, H. Jiang, Y. Deng and T. Liang, Appl. Surf. Sci., 2019, 481, 1403–1411 CrossRef CAS.
  246. M. Ji, S. Niu, Y. Du, B. Song and P. Xu, ACS Sustainable Chem. Eng., 2018, 6, 11922–11929 CrossRef CAS.
  247. J. Diao, Y. Qiu, S. Liu, W. Wang, K. Chen, H. Li, W. Yuan, Y. Qu and X. Guo, Adv. Mater., 2020, 32, 1905679 CrossRef CAS PubMed.
  248. Q. Liang, H. Jin, Z. Wang, Y. Xiong, S. Yuan, X. Zeng, D. He and S. Mu, Nano Energy, 2019, 57, 746–752 CrossRef CAS.
  249. L. Guo, J. Wang, X. Teng, Y. Liu, X. He and Z. Chen, ChemSusChem, 2018, 11, 2717–2723 CrossRef CAS PubMed.
  250. J.-Q. Chi, J.-H. Lin, J.-F. Qin, B. Dong, K.-L. Yan, Z.-Z. Liu, X.-Y. Zhang, Y.-M. Chai and C.-G. Liu, Sustainable Energy Fuels, 2018, 2, 1610–1620 RSC.
  251. Q. Zhu, L. Yao, R. Tong, D. Liu, K. W. Ng and H. Pan, New J. Chem., 2019, 43, 14518–14526 RSC.
  252. P. Wang, J. Qi, C. Li, X. Chen, T. Wang and C. Liang, ChemElectroChem, 2020, 7, 745–752 CrossRef CAS.
  253. W. Yuan, S. Wang, Y. Ma, Y. Qiu, Y. An and L. Cheng, ACS Energy Lett., 2020, 5, 692–700 CrossRef CAS.
  254. S. H. Kwag, Y. S. Lee, J. Lee, D. I. Jeong, S. B. Kwon, J. H. Yoo, S. Woo, B. S. Lim, W. K. Park, M.-J. Kim, J. H. Kim, B. Lim, B. K. Kang, W. S. Yang and D. H. Yoon, ACS Appl. Energy Mater., 2019, 2, 8502–8510 CrossRef CAS.
  255. Q. L. Wang, C. Q. Xu, W. Liu, S. F. Hung, H. B. Yang, J. J. Gao, W. Z. Cai, H. M. Chen, J. Li and B. Liu, Nat. Commun., 2020, 11, 4246 CrossRef CAS PubMed.
  256. T. Li, Y. Hu, X. Pan, J. Yin, Y. Li, Y. Wang, Y. Zhang, H. Sun and Y. Tang, Chem. Eng. J., 2020, 392, 124845 CrossRef CAS.
  257. X. Shu, S. Chen, S. Chen, W. Pan and J. Zhang, Carbon, 2020, 157, 234–243 CrossRef CAS.
  258. X. Bai, T. Cao, T. Xia, C. Wu, M. Feng, X. Li, Z. Mei, H. Gao, D. Huo, X. Ren, S. Li, H. Guo and R. Wang, Nanomaterials, 2023, 13(4), 752 CrossRef CAS PubMed.
  259. B. Sun, G. Dong, J. Ye, D.-F. Chai, X. Yang, S. Fu, M. Zhao, W. Zhang and J. Li, Chem. Eng. J., 2023, 459, 141610 CrossRef CAS.
  260. B. Yu, F. Qi, B. Zheng, W. Hou, W. Zhang, Y. Li and Y. Chen, J. Mater. Chem. A, 2018, 6, 1655–1662 RSC.
  261. X. Kong, L. Wang, Z. Xi, Y. Liu, Y. Zhou, Z. Wan, X. Chen, S. Li and L. Rong, Electrochim. Acta, 2023, 439, 141660 CrossRef CAS.
  262. D. Zheng, P. Cheng, Q. Yao, Y. Fang, M. Yang, L. Zhu and L. Zhang, J. Alloys Compd., 2020, 848, 156588 CrossRef CAS.
  263. G. Zhu, X. Xie, X. Li, Y. Liu, X. Shen, K. Xu and S. Chen, ACS Appl. Mater. Interfaces, 2018, 10, 19258–19270 CrossRef CAS PubMed.
  264. S. Umapathi, J. Masud, A. T. Swesi and M. Nath, Adv. Sustainable Syst., 2017, 1, 1700086 CrossRef.
  265. Z. Feng, L. Ren, Y. Liu and B. Gao, Ionics, 2023, 29, 1515–1521 CrossRef CAS.
  266. Z. Feng, L. Ren, Y. Liu and B. Gao, J. Solid State Electrochem., 2023, 1469–1476 CrossRef CAS.
  267. P. Kuang, M. He, H. Zou, J. Yu and K. Fan, Appl. Catal., B, 2019, 254, 15–25 CrossRef CAS.
  268. T. Li, S. Li, Q. Liu, J. Yin, D. Sun, M. Zhang, L. Xu, Y. Tang and Y. Zhang, Adv. Sci., 2020, 7, 1902371 CrossRef CAS PubMed.
  269. W.-Z. Chen, P.-Y. Liu, L. Zhang, Y. Liu, Z. Liu, J. He and Y.-Q. Wang, Chem. Eng. J., 2021, 424, 130434 CrossRef CAS.
  270. U. N. Pan, T. I. Singh, D. R. Paudel, C. C. Gudal, N. H. Kim and J. H. Lee, J. Mater. Chem. A, 2020, 8, 19654–19664 RSC.
  271. F. Zhou, X. Zhang, R. Sa, S. Zhang, Z. Wen and R. Wang, Chem. Eng. J., 2020, 397, 125454 CrossRef CAS.
  272. K. S. Ranjith, C. H. Kwak, S. M. Ghoreishian, J. S. Im, Y. S. Huh and Y.-K. Han, Nanotechnology, 2020, 31, 275402 CrossRef CAS PubMed.
  273. K. Ram Kumar and T. Maiyalagan, Ceram. Int., 2023, 49, 1195–1202 CrossRef CAS.
  274. W. Wang, Y. Yang, Y. Zhao, S. Wang, X. Ai, J. Fang and Y. Liu, Nano Res., 2022, 15, 872–880 CrossRef CAS.
  275. D. C. Nguyen, D. T. Tran, T. L. L. Doan, D. H. Kim, N. H. Kim and J. H. Lee, Adv. Energy Mater., 2020, 10, 1903289 CrossRef CAS.
  276. W. Xu, Y. Wei, S. Zhou, R. Sun, X. Huang, S. Han, S. Wang and J. Jiang, Electrochim. Acta, 2023, 454, 142377 CrossRef CAS.
  277. J. Guo, Z. Wei, K. Wang and H. Zhang, Int. J. Hydrogen Energy, 2021, 46, 27529–27542 CrossRef CAS.
  278. J. Wang, G. Lv and C. Wang, Appl. Surf. Sci., 2021, 570, 151182 CrossRef CAS.
  279. L. Yao, R. Li, H. Zhang, M. Humayun, X. Xu, Y. Fu, A. Nikiforov and C. Wang, Int. J. Hydrogen Energy, 2022, 47, 32394–32404 CrossRef CAS.
  280. B. Zhang, J. Li, Q. Song and H. Liu, Int. J. Energy Res., 2022, 46, 12406–12416 CrossRef CAS.
  281. S. B. Devi and R. Navamathavan, J. Electrochem. Soc., 2023, 170, 096503 CrossRef.
  282. S. M. Li, L. Bai, H. B. Shi, X. F. Hao, L. Chen, X. J. Qin and G. J. Shao, Ionics, 2021, 27, 3109–3118 CrossRef CAS.
  283. K. B. Patel, B. Parmar, K. Ravi, R. Patidar, G. R. Bhadu, J. C. Chaudhari and D. N. Srivastava, Appl. Surf. Sci., 2023, 616, 156499 CrossRef CAS.
  284. S. J. Huang, S. Balu, N. R. Barveen and R. Sankar, Colloids Surf., A, 2022, 654, 130024 CrossRef CAS.
  285. Y. Ju, S. Y. Feng, X. B. Wang, M. Li, L. Wang, R. D. Xu and J. L. Wang, ChemistrySelect, 2021, 6, 4979–4990 CrossRef CAS.
  286. N. Ullah, W. T. Zhao, X. Q. Lu, C. J. Oluigbo, S. A. Shah, M. M. Zhang, J. M. Xie and Y. G. Xu, Electrochim. Acta, 2019, 298, 163–171 CrossRef CAS.
  287. B. Chen, P. Hu, F. Yang, X. J. Hua, F. F. Yang, F. Zhu, R. Y. Sun, K. Hao, K. S. Wang and Z. Y. Yin, Small, 2023, 19, 2207177 CrossRef CAS PubMed.
  288. S. F. Tan, W. M. Ouyang, Y. J. Ji and Q. W. Hong, J. Alloys Compd., 2021, 889, 161528 CrossRef.
  289. C.-P. Wang, X. Lian, Y.-X. Lin, L. Cui, C.-N. Li, N. Li, A.-N. Zhang, J. Yin, J. Kang, J. Zhu and X.-H. Bu, Small, 2023, 19, 2305201 CrossRef CAS PubMed.
  290. C.-P. Wang, Y. Feng, H. Sun, Y. Wang, J. Yin, Z. Yao, X.-H. Bu and J. Zhu, ACS Catal., 2021, 11, 7132–7143 CrossRef CAS.
  291. A. Hanan, M. N. Lakhan, F. Bibi, A. Khan, I. A. Soomro, A. Hussain and U. Aftab, Chem. Eng. J., 2024, 482, 148776 CrossRef CAS.
  292. B. Li, Q. Lei, T. Qin, X. Zhang, D. Zhao, F. Wang, W. Li, Z. Zhang and L. Fan, CrystEngComm, 2021, 23, 8260–8264 RSC.
  293. R. Nivetha, K. Gothandapani, V. Raghavan, G. Jacob, R. Sellappan, P. Bhardwaj, S. Pitchaimuthu, A. N. M. Kannan, S. K. Jeong and A. N. Grace, ACS Omega, 2020, 5, 18941–18949 CrossRef CAS PubMed.
  294. C.-P. Wang, Y.-X. Lin, L. Cui, J. Zhu and X.-H. Bu, Small, 2023, 19, 2207342 CrossRef CAS PubMed.
  295. X.-P. Li, C. Huang, W.-K. Han, T. Ouyang and Z.-Q. Liu, Chin. Chem. Lett., 2021, 32, 2597–2616 CrossRef CAS.
  296. Y. Yao, Z. Zhang and L. Jiao, Energy Environ. Mater., 2022, 5, 470–485 CrossRef CAS.

Footnote

These authors contributed equally to this work.

This journal is © The Royal Society of Chemistry 2024