DOI:
10.1039/D5RA01184A
(Review Article)
RSC Adv., 2025,
15, 13397-13430
Transition metal phosphide/ molybdenum disulfide heterostructures towards advanced electrochemical energy storage: recent progress and challenges
Received
18th February 2025
, Accepted 8th April 2025
First published on 28th April 2025
Abstract
Transition metal phosphide @ molybdenum disulfide (TMP@MoS2) heterostructures, consisting of TMP as the core main catalytic body and MoS2 as the outer shell, can solve the three major problems in the field of renewable energy storage and catalysis, such as lack of resources, cost factors, and low cycling stability. The heterostructures synergistically combine the excellent conductivity and electrochemical performance of transition metal phosphides with the structural robustness and catalytic activity of molybdenum disulfide, which holds great promise for clean energy. This review addresses the advantages of TMP@MoS2 materials and their synthesis methods—e.g., hydrothermal routes and chemical vapor deposition regarding scalability and cost. Their electrochemical energy storage and catalytic functions e.g., hydrogen and oxygen evolution reactions (HER and OER) are also extensively explored. Their potential within battery and supercapacitor technologies is also assessed against leading performance metrics. Challenges toward industry-scale scalability, longevity, and environmental sustainability are also addressed, as are optimization and large-scale deployment strategies.
1. Introduction
Renewable resources have become the new target, considering the need to substitute existing fossil fuel sources as the overall energy demand is increasing.1,2 This consequently generates a high demand for high-performance energy conversion and storage devices.3 While solar and wind power are very attractive renewables, their intermittent nature necessitates effective and scalable solutions for storing energy.4 Batteries, supercapacitors, and hybrid storage devices are the necessary solutions for the challenge but are typically restricted by their high cost, short life cycle, and more considerably their scalability.5–8
Heterostructure materials, formed of at least two constituents combined for their synergistic functionalities, have drawn substantial interest.9–11 TMP@MoS2 heterostructures are particularly important because they combine the high electrical conductivity and catalytic nature of transition metal phosphides (TMPs) with the structural strength and large molybdenum disulfide (MoS2) surface area.12–14 Together, they create synergy, leading to enhanced charge transfer, electrochemical stability, and catalytic activity, thus rendering these composites promising for both electrocatalysis and energy storage devices.15–17
TMP@MoS2 synthesis and characterization improvements have demonstrated their potential for next-generation energy technologies. Various fabrication techniques, such as hydrothermal synthesis,18,19 electrodeposition,20–22 and chemical vapor deposition,23,24 enable the precise engineering of heterostructures with tailored properties.25,26 The structural and electrochemical properties of the materials are extensively investigated using the techniques of X-ray diffraction (XRD),27–29 scanning electron microscopy (SEM),29–31 transmission electron microscopy (TEM),32 and X-ray photoelectron spectroscopy (XPS)33,34 to confirm their composition, morphology, and chemical states. Moreover, TMP@MoS2 heterostructures possess increased efficiencies in such important electrochemical reactions as hydrogen evolution reaction (HER) and oxygen evolution reaction (OER),2,3,13 that play a critical role in water splitting and fuel cell applications. Though bulk-scale synthesis, interface stability of the materials, and electrochemical durability are needed to commercialize the materials,35,36 these also need to be accounted for. Herein, we review the recent efforts we made on the TMP@MoS2 heterostructure from two aspects: the synthesis strategy of TMP@MoS2 and the electrochemical properties & application. We also discuss the major methodologies used for characterization, highlight the bottlenecks that currently exist in the field, and suggest a set of recommendations for the optimization of the materials for practical applications. By interlinking the fundamental studies with technical applications, the review is focused on delivering useful insights into the role of TMP@MoS2 heterostructures toward green energy storage and catalysis.
2 Synthesis methods of transition metal phosphides (TMPs), molybdenum disulfide, and TMP@MoS2 heterostructures
Design and modification of advanced materials like TMP@MoS2 are mainly done through proper control and modulation of their physicochemical and electrochemical properties to render a well-developed energy storage performance. A suitable choice of synthesis methodology would mark a significant difference in structural stability, conductivity, surface area, and accessibility to active sites. Therefore, materials' scalability, cost, and quality are considered essential concerns for developing and extending the utilization of heterostructures. This section describes in detail various synthesis techniques, their fundamental principles, advantages, and limitations to guide future research and development.
2.1 Ball milling
Ball milling is a mechanochemical method that introduces mechanical force to the materials through grinding balls in rotating or vibrating chambers; breakdown and reformation into nanoscale material occur accordingly. The significant contribution generally comes from high-energy ball collisions with material introducing localized heat and pressure, driving the desired transformations. Therefore, the advantages are: the setup is simple and economy-friendly; it effectively leads to the generation of fine and uniform nanoparticles. Scalability for batch or continuous production is advantageous, whereas potential contamination from milling media is a disadvantage. There is a limitation in tight control of particles' size and morphology, and milling conditions need optimization to avoid degradation.37–39
Enabling specific mechanochemical approaches in one-shot or continuous production. The most common batch methods are the shakers pot methods and the planetary mills, and constant processes are often accomplished with twin-screw extruders. The widespread use of TSE in continuous production is partly responsible for the heightened attention toward the use of these mills.40,41
Several reviews and research studies documented the available commercial instruments' distinct features and corresponding advantages and drawbacks. For example, shaker mills are simple and efficient only on a small scale; planetary mills give high-energy impacts that reduce fine and uniform particle size.42,43 On the other side of the spectrum, TSE-based extruders provide continuous production and, thus, are more scalable but are likely to be more complex to set up and run.44,45
The equipment and additives selected heavily influence the mechanisms underlying these reactions. Some recent developments introduced new methodologies involving additives for improving reaction simulation or for better outcomes or specified reactivities and have advanced the field significantly. Many studies employ such additives to promote efficiency, reaction pathway control, and desired product features, realizing new improvements in fields such as pharmaceuticals through materials science.46,47 The constant development of ball milling techniques and careful strategies in using additives underline the dynamics of mechanochemical processes, thus affording many more possibilities for innovations and applications in science and industry.48–50 Some key parameters that affect ball milling outcomes are Milling Time,51,52 Ball-to-Powder Ratio (BPR), Rotation Speed, and Milling Atmosphere. The longer timer is equal to a smaller particle size and higher surface area. Excessive milling may also lead to agglomeration, amorphization, or degradation of sensitive materials.51,53 Increased BPR increases the impact of energy. Thus, particle refinement occurs at higher rates. On the other hand, a high ratio might lead to contamination due to the wearing of milling media.54,55 Higher collision speeds increase the frequency and energies of the collisions and yield smaller particles,54,56,57 which enable higher frictional heat and the possibility of damaging phase transition by pressure and heat. Oxidation, reduction, and phase formation are strongly dependent upon inert gas (e.g., Ar, N2) or chemically reactive environments (e.g., H2, O2).8,11,55 In the process of ball milling, the continuous collision between particles and balls causes local heat and pressure and results in (1) distortion of the lattice and the formation of defects, which enhance the reactivity of the material. (2) Phase transformations, e.g., amorphization or formation of metastable phases.58–60
Solid-state reactions by the combination of reactants at the atomic level make chemical synthesis independent of the use of solvents.28,61 Ball milling is flexible for batch or continuous production.62,63
Batch processing (Shaker and Planetary Mills) advantages: high-energy collisions allow for rapid particle size reduction.10,64 Limitations: Limited throughput with scaling-up difficulties.65
Continuous processing (Twin-Screw Extruders-TSE) advantages: it offers continuous production and is suitable for industrial use.66,67 Limitations: it involves complex setup and process optimization to regulate the reaction (Fig. 1).71,72
 |
| Fig. 1 Schematic representation of (a) ball milling process, (b) Shaker Mill, (c) Planetary Mill, (d) and (d′) Single and Twin-Screw Extruder. Reproduced from ref. 68–70 with permission from [Wiely], copyright [2024]. | |
2.2 Electrochemical deposition (electrodeposition)
Electrochemical deposition is, in principle, based on reducing metal ions from an electrolyte solution onto an electrically conductive substrate under applied potential. Indeed, this technique has been extensively used in preparing thin films and composite materials since it enables excellent thickness and composition control. Advantages: highly controllable process with accuracy in thickness and morphology, inexpensive and easy to handle for laboratory-scale applications, up-scalable for industrial applications, with minimal waste production. Limitations include being only suitable for conductive substrates, uniform deposition over complex, complicated shapes, and deposition rate and material properties dependent upon electrolyte composition and applied conditions.73–75
Electrochemical deposition is the electrodeposition process, typically requiring two or three electrodes for better control.76 During synthesis, conditions like temperature, solution pH, concentration of reactants, and impurity are modified.77 These parameters influence how fast the material dissolves in an electrolyte, thus affecting deposition.78 These may alter the dissolution rate of the materials in the electrolyte and thereby affect deposition. It is a comparatively low-cost method by which the thickness of the deposit can be managed, and the desired properties of the coating are obtained.79 Electrodeposition can deposit large amounts of various materials, including metal-matrix composites, onto substrates.80 The technique produces materials that enhance the surface properties of corrosion resistance and electrical conductivity for small-scale labs and large industries.81 However, uniform deposition over complex shapes requires constant concentrations of electrolytes.82 Optimizing the process parameters avoids cracking or peeling of deposited layers.83 As they are possibly harmful substances to humans and the environment, electrode position materials must be handled and disposed of safely.84
For the TMP@MoS2 heterostructures, direct deposition with control over film thickness is facilitated by electrodeposition and is a good alternative for hydrothermal and solvothermal methods.85 Uniform distribution of TMP into MoS2 is difficult due to varying nucleation rates across the surface.86 Homogenous coverage and prevention against agglomeration are guaranteed through optimization of the electrolyte concentration, applied potential, and deposition time.87 Electrodeposited TMP@MoS2 composites are found as potential supercapacitors and battery candidates with superior electrochemical performances through synergism between the two (Fig. 2).89,90
 |
| Fig. 2 Schematic of electrodeposition process in three-electrode system. Reproduced from ref. 88 with permission from [IOP Science], copyright [2008]. | |
2.3 Hydrothermal/solvothermal synthesis
This technique precipitates the materials from the solution inside the high-temperature and high-pressure autoclave. It is employed extensively for the synthesis of TMP@MoS2 heterostructures since it is capable of regulating the crystal growth and composition. Hydrothermal and solvothermal synthesis has provided highly uniform and crystalline nanostructures through the regulation of factors like temperature, pressure, and reaction time. In contrast to electrodeposition, the processes here possess better phase purity and uniform structure but require longer synthesis time and higher energy requirements. They have the advantages of being efficient and versatile in the generation of tailored morphologies, being suitable for large-scale fabrication, and allowing size, phase, and crystallinity to be controlled precisely. Also, limitations include the requirement for high-cost equipment with unique features, such as an autoclave. Reaction times can be long, and lower throughput and Surfactants that are not well controlled might cause the introduction of impurities.91,92
One of the most powerful approaches for the synthesis of isolated inorganic compounds in the liquid phase is hydrothermal and solvothermal synthesis.93 Both approaches facilitate the uniform deposition of TMP onto MoS2 layers with enhanced electrochemical performance for the function of energy storage. They consist of the reaction between phosphoric materials and metal nitrates or metal chlorides to yield uniform emulsions. The structure is controlled through the incorporation of surfactants into the mixture. The obtained emulsion is subsequently dissolved and recrystallized under high temperatures and pressures within a hydrothermal reactor, e.g., a stainless-steel autoclave lined with Teflon.94
During the synthesis process, the vapor saturation pressure is higher than 100 °C, with the autogenous pressure becoming 100 MPa. The high-pressure autoclave produces well-crystallised structures, and therefore, highly pure inorganic compounds are synthesized with high yields. The reaction must be adjusted properly for TMP@MoS2 to avoid phase segregation and maximize electrical conductivity. Various inorganic compounds like phosphates and complex oxides are prepared using various techniques.95 Surfactants allow for precise control of the particle morphology and size, improving the material's properties.96 The methods are also expected to be scaled up for industry and are suitable for research and commercial use.92 They are ideal for both research and commercial use. The hydrothermal and solvothermal methods, however, have a few disadvantages. The need for extended reaction times (several hours to days) limits rapid material production, which is a drawback compared to faster methods like electrodeposition. As the high temperatures and pressures demanded require unique apparatuses,97 mainly of an expensive nature and complicated to operate,98 it is time-consuming since the period taken for dissolution and recrystallization runs into hours.98 Moreover, other additives and surfactants can be introduced to induce impurities if not regulated.95 It is also challenging to have homogeneity in TMP@MoS2 heterostructures due to nucleation rate and reaction kinetic discrepancies at different temperatures and pressures. Lastly, such techniques cannot be applied to inorganic materials of some natures that have a propensity to become unstable under required conditions (Fig. 3).100
 |
| Fig. 3 Schematic of solvothermal technique's equipment. Reproduced from ref. 99 with permission from [Elsevier], copyright [2023]. | |
2.4 Chemical vapor deposition
CVD is a standard method for depositing material or thin films onto substrates by chemical reaction between precursors in the vapor phase. It is widely applied in synthesizing TMP@MoS2 heterostructures due to the possibility of depositing uniform, high-quality coatings with defined stoichiometry and morphology. This technique has many advantages in terms of thickness and stoichiometry control, pure material production, large-format uniform coating, and use with complex geometries. Limitations include high processing temperatures, often complex equipment, and well-controlled reaction conditions in many cases. Moreover, specific precursors can be toxic, highly expensive, and even environmentally hazardous, thus creating problems in handling and waste management.101–103
It is an excellent technique for forming ultra-pure, smooth, thin films on heated surfaces through chemical reactions.101 For TMP@MoS2, it enables the growth of MoS2 nanosheets with controlled thickness, facilitating improved electrochemical performance in energy storage applications. Compared with PVD techniques, CVD produces highly uniform and conformal coatings, giving it more flexibility because of the strong bonds to the substrate.104–106 Nevertheless, CVD has its shortcomings. It is generally accompanied by the application of elevated temperatures, which, if poorly optimized, can lead to phase segregation in TMP@MoS2 and is restricted by substrate material type.107 Besides, the choice of TMP precursors is also important in determining nucleation, growth rate, and material properties in general, necessitating optimal selection. Substances used in the process will probably be toxic; therefore, handling and disposal must be done carefully.108 Additionally, integrating TMP with MoS2 requires precise reaction conditions to prevent unwanted secondary phases. The complexity increases the equipment costs compared with PVD methods.109 Despite these limitations, the high regard for CVD has been retained due to its capacity to form exceptionally high-quality films with excellent uniformity, making it virtually indispensable in many industries (Fig. 4).107
 |
| Fig. 4 Scheme of the chemical vapor deposition apparatus. Reproduced from ref. 110 with permission from [Beilstein], copyright [2017]. | |
2.5 Phosphorization methods
The general trend of phosphorization usually includes the conversion of precursors to corresponding phosphides by their reactions in the solid state,111 gas phase,112 or solution with active phosphorus-containing agents.113 Advantages include simplicity, versatility for various phosphides, material properties-conductivity, electrochemical activity, and multiple starting materials.114 Limitations: the methods in the gas phase require handling hazardous phosphorous gases.91 Solid-state reactions involve treatment at high temperatures for extended periods.115 Although very popular, some solution-based techniques always risk producing inhomogeneities with chemical waste.116
Phosphorization is one of the most used methods in materials science, mainly for synthesizing metal phosphides applied in electrocatalysis117 and battery118 electrodes. Phosphorization within TMP@MoS2 heterostructures is accountable for the conversion of transition metals to phosphide phases, hence enhancing electrochemical performance by providing improved conductivity and charge transfer kinetics.
There are three major techniques currently in use: solid-state reactions,119 gas-phase phosphorization,120 and solution-based techniques.113 Though simple to carry out and suitable for large-scale production, solid-state reactions require high temperatures and long processing times.121,122 Requires strict temperature control to prevent unwanted side reactions. Gas-phase phosphorization achieves better control; however, the hazardous gases involved create significant environmental and safety concerns.123–125 Necessitates advanced containment measures due to toxic precursors. Solution-based techniques have potential because the temperature ranges used are lower126 and nanostructured materials are possible; however, they often yield inhomogeneous products127 and chemical waste as by-products.102,118 Enabling nanostructured TMP@MoS2 synthesis with tailored morphologies; however, achieving uniformity remains challenging (Fig. 5).
 |
| Fig. 5 Phosphorization process for synthesizing NiCoP by NaH2PO2. Reproduced from ref. 128 with permission from [Elsevier], copyright [2019]. | |
2.6 Ultrasonic method and calcination
Ultrasonic processing and calcination serve critical roles in material synthesis, particularly in enhancing the structural and electrochemical properties of TMP@MoS2 heterostructures. The application of ultrasonic processing has been extensive in the production of fluidic cavitation for chemical reaction intensification.129 Calcination is generally the heat treatment process for crystallization enhancement and/or removal of impurities from materials.130 Merits of the ultrasonication method it has fast processing with low energy input. It offers highly homogeneous nanoparticles.131 Drawbacks: it is non-scalable for large-scale synthesis. It is expensive in terms of equipment for a particular frequency and setup.132 Ultrasonic processing creates a high-frequency sound that is used in producing cavitation in liquid media, resulting in the collapse of tiny bubbles at a very fast rate. This collapse creates localized pressure and heat. Benefits of calcination: the process enhances material crystallinity and purity,133 eliminating residues of the impurities and solvent residues.134 In TMP@MoS2 synthesis, it facilitates phase transformation and stabilizes the heterostructure. Limitation: the limit to this is high energy consumption with the probability of grain growth; this may reduce the integrity of the nanostructure.135
The process is also referred to as sonochemistry, and it is a process in which material chemistry is altered by exposing it to the application of ultrasonic sound waves with a high frequency.136 Destabilization of the liquid material is presumed, with very small bubbles forming and collapsing afterward.137 The collapse is extremely rapid, generating huge pressure and heat, which deposits highly homogeneous nanoparticles with very little energy input.138 Ultrasonic processing and calcination offer special advantages and restrictions. Ultrasonic processing creates highly uniform ultra-fine nanoparticles with high efficiency. However, restricted sound wave penetration does not support large-scale production. In contrast, calcination creates highly pure, crystalline material. However, it consumes high energy and can degrade product quality due to grain growth and prolonged processing times (Fig. 6).141,142
 |
| Fig. 6 Schematic of (a) ultrasound system (water bath type), (b) calcination process. Reproduced from ref. 139 and 140 with permission from [Wiely], copyright [2022]. | |
2.7 Other methods
Other common methods for preparing metal phosphides include electroless deposition, which enables uniform coatings without external electrical power;143 photochemical deposition, which uses light energy to drive reactions, allowing precise nanoscale modifications;102 or hot injection, provides rapid nucleation and controlled particle growth, essential for tuning morphology.144 In solid–solid and solid/gas–solid processes, temperature-programmed reduction (commonly used in solid-state and gas–solid processes to reduce metal precursors at controlled temperatures) is usually used;102 however, joule thermal shock Utilizes rapid heating to synthesize nanomaterials with unique phase compositions,145 arc melting a high-energy technique to produce dense and crystalline materials,146 and microwave treatment enables fast, energy-efficient synthesis by selectively heating precursors147,148 can also be performed. These techniques have often been extensively applied in fabricating supercapacitors, batteries, and electrodes. A comparison between the advantages and disadvantages is presented in Tables 1 to 3. Also, scheme one comparatively analyses all of the synthesis methods, crystal structure (XRD), and morphology (SEM) for MoP/MoS2, CoP/MoS2, and CoP/Ni2P/MoS2 heterostructures. The dense-packed distribution of the nanosheets suggests the high surface area potential, shown in the SEM image in (a). XRD analysis in (a) shows the successfully synthesized MoS2 and MoP phases with clear evidence of the heterostructure. This architecture's combination of molybdenum disulfide and phosphide is expected to yield unique properties, enhancing catalytic performance. Heterostructures promise much in many electrochemical applications with their enhanced conductivity, stability, and catalytic activity.
Table 1 Advantages and limitations of different synthesis methods for metal phosphide electrodes in energy storage and electrocatalysis
Synthesis method |
Applications (energy storage) |
Advantages |
Limitations |
Ref. |
Electrochemical deposition |
Batteries, supercapacitors, and electrocatalysts |
Fast, low-cost, precise control over synthesis |
Requires careful optimization of parameters, limited to conductive substrates |
73 and 149–152 |
Electroless deposition |
Batteries, supercapacitors, and electrocatalysts |
Simple, scalable, high-quality deposits |
Needs surface pretreatment, limited bath life (frequent solution replacement), and environmental concerns |
152–155 |
Photo deposition (photochemical deposition) |
Primarily electrocatalysts |
Enhanced material properties (crystallinity, conductivity) |
Poor reproducibility (results may vary) |
156–158 |
Sonochemical methods |
Electrocatalysts, supercapacitors |
Improved mass transport, surface cleaning |
Complex setup with specialized equipment, limited adoption |
159–161 |
Hydro/Solvothermal synthesis |
Batteries, supercapacitors, and electrocatalysts |
Efficient, inexpensive, versatile (particle properties) |
High pressure & temperature, time-consuming reactions, limited scalability |
115 and 162–165 |
Vapor-phase hydrothermal method |
Batteries, supercapacitors, and electrocatalysts |
High purity, good scalability |
Low yield of material per batch |
166–169 |
Hot-injection method |
Batteries, supercapacitors, and electrocatalysts |
Narrow size distribution, controlled morphology |
Expensive chemicals, challenging to scale up production |
144 and 170–172 |
Solid-state thermal treatment |
Batteries, supercapacitors, and electrocatalysts |
Potentially simple, scalable, energy-efficient, controlled particle size, wide material range |
High temperatures, long reaction times, non-uniform products, limited control over composition, particle agglomeration (increases with complexity) |
173 |
Ball-milling |
Batteries, supercapacitors, electrocatalysts |
Simple, cost-effective, scalable |
Contamination from milling media, limited control over final particle size |
174–176 |
Temperature-programmed reduction |
Batteries, supercapacitors, and electrocatalysts |
Precise control over reduction conditions, versatile |
Requires careful temperature control, sometimes slow |
177–179 |
High-temperature shock |
Rechargeable batteries, electrocatalysts |
Rapid processing, unique material properties |
Potential material degradation, expensive equipment |
180 |
Arc melting |
Primarily electrocatalysts |
Fast startup, minimal contamination |
Compositional inhomogeneity for large samples |
144, 172 and 181 |
Microwave-assisted |
Electrocatalysts, supercapacitors |
Faster synthesis, lower power consumption |
Expensive equipment and limited scalability for large-scale production |
172, 182 and 183 |
Chemical vapor deposition (CVD) |
Batteries, supercapacitors, and electrocatalysts |
High purity |
High temperatures |
172, 179, 184 and 185 |
Plasma-enhanced CVD |
Batteries, supercapacitors, and electrocatalysts |
Lower temperatures than CVD, better film properties |
High equipment cost, complex process |
186–188 |
Table 2 Comparison of synthesis methods for metal phosphides@MoS2 heterostructures: applications, advantages, and limitations
Synthesis method |
Applications (energy storage) |
Advantages |
Limitations |
Ref. |
Hydrothermal synthesis |
Electrocatalysts, supercapacitors, batteries |
Simple, low-temperature, cost-effective, tailorable structures |
Requires autoclaves, limited crystal growth observation |
12, 189 and 190 |
Phosphorization synthesis |
Batteries, electrocatalysts |
Enhances LiPS anchoring, improves activity and kinetics |
Complex process, limited single-catalyst performance, stability concerns |
12 and 191 |
Emulsion polymerization & pyrolysis |
Electrocatalysts |
Uniform, stable structures, scalable, high surface area nanostructures |
High energy consumption, polymerization control challenges, potential environmental impact |
192 |
Plasma-assisted phosphorization |
Electrocatalysts |
Fast reactions, uniform coatings, improved catalytic properties |
Specialized equipment, high costs, potential safety hazards |
193 |
Electrodeposition |
Supercapacitors, electrocatalysts |
Precise control over composition and thickness; simple, cost-effective, scalable |
Requires conductive substrates, potential adhesion issues, non-uniform deposition |
13 and 20 |
Ultrasonic method & calcination |
Electrocatalysts |
Enhanced mixing, high surface area nanoparticles, relatively low cost |
Uneven particle sizes, high energy consumption, limited scalability |
194 |
Chemical vapor deposition (CVD) |
Electrocatalysts |
High purity, high control over thickness and composition, scalable |
Requires high temperature, expensive equipment, potential safety hazards |
195 |
Solvothermal method |
Electrocatalysts |
High crystallinity, uniform particle size, tailorable structures |
Requires high pressure, long reaction times, limited scalability |
196 and 197 |
Melt-diffusion method |
Batteries |
High-purity products, straightforward process, high control over diffusion |
High-temperature requirement, potential phase separation, limited to specific materials |
198 |
High-temperature carbonization |
Electrocatalysts |
High surface area, high thermal stability, and enhanced conductivity |
High energy consumption, potential environmental impact, complex process |
192 |
Table 3 Synthesis methods for MoS2: applications, benefits, and limitations
Synthesis method |
Applications (energy storage) |
Advantages |
Limitations |
Ref |
Hydrothermal synthesis |
Electrocatalysts, supercapacitors, batteries |
High yield, low cost, controllable morphology |
Prolonged reaction time, high pressure required |
199–201 |
Phosphorization synthesis |
Electrocatalysts, supercapacitors, batteries |
Enhanced electrical conductivity, improved activity |
Complex process, use of toxic phosphorous |
202–204 |
Emulsion polymerization & pyrolysis |
Batteries, supercapacitors |
Uniform particle size, high surface area |
Multiple steps require surfactants |
205,206 |
Microwave-assisted |
Electrocatalysts supercapacitors batteries |
Rapid synthesis, energy-efficient, uniform heating |
Limited scalability, expensive equipment |
207–209 |
Electrodeposition |
Electrocatalysts, supercapacitors, batteries |
Simple, cost-effective, reasonable control over thickness |
Requires conductive substrates, lower purity |
210–212 |
Ultrasonic method |
Electrocatalysts supercapacitors batteries |
Simple, scalable, fast reaction |
In inhomogeneous particle size, potential damage to materials |
213–215 |
Chemical vapor deposition (CVD) |
Electrocatalysts supercapacitors batteries |
High purity, reasonable control over thickness, large-area growth |
High cost, complex setup, high temperature |
216–218 |
Solvothermal method |
Batteries supercapacitors electrocatalyst |
High crystallinity, uniform particle size |
Prolonged reaction time requires autoclaves |
219–221 |
Calcination |
Electrocatalysts supercapacitors batteries |
Simple process, high yield, reasonable control over structure |
High energy consumption, potential loss of material properties |
222–224 |
High-temperature sulfurization |
Electrocatalyst |
High crystallinity, reasonable control over phase |
High temperature required, sulfur vapor hazards |
225 |
Based on previous reports summarized in Tables 1–3, the synthesis methods enormously influence the properties and performance of metal phosphide electrodes in energy storage and electrocatalysis. Hydrothermal and solvothermal methods are highly flexible and inexpensive and allow for control over particle properties; however, they often require a very long reaction time and are difficult to scale up. Electrochemical and electroless deposition is inexpensive and fast but substrate-limited and environmentally detrimental. Microwave, sonochemical, and plasma-based methods improve reaction rate and product quality and render them energy efficient but at the expense of highly specific equipment and conditions. Hydrothermal synthesis is still the most popular method to prepare TMP@MoS2 heterostructures due to its ease and versatility. However, the catalytic properties are auspicious for phosphorization or plasma-assisted techniques. Versatile methods have been used for hydrothermal and chemical vapor deposition in MoS2 for this material. Simultaneously, these approaches are accompanied by their restrictions on reaction conditions, type of equipment, and product properties. The optimal synthesis route is a function of such factors as desired material properties, cost, scalability, and environmental footprint for a specific application.
In conclusion, this section delved into advanced synthesis techniques to establish a basis for comprehending their practical roles in electrochemical processes, particularly in hydrogen evolution reactions (HER). In this case, materials like the TMP@MoS2 heterostructure have key structural characteristics and catalyzing efficacy that take center stage in the optimization of HER performance. The essential characteristics, such as improved electron transfer, improved active site exposure, and better structure stability, offer the fundamental platform for optimization in catalysis. Moreover, studies in the acidic and alkaline media point to the importance of such studies in acquiring essential information that can be used in the design of energy-conversion devices.
3. Physical properties of TMP@MoS2 heterostructures
3.1 Morphology & structural features
The morphologies of the TMP@MoS2 heterostructures are various, including layered morphologies, nanowires, and core–shell morphologies. For the enhanced hydrogen evolution reaction, core–shell structured CoP@MoS2 has been employed as an electrocatalyst. In the heterostructure, Maximization of the active site exposure is an available way to offer enhanced catalytic performances. Typically, electron microscopes analyze morphologies; the morphology of the CoCo-PBA possessed a smooth surface with an average particle size ∼of about 400 nm. The aggressive PH3 released by NaH2PO2 during the phosphorylation on the nanocubes of the CoCo-PBA produces a rough surface. In the current work, the nanocubes' surface of the CoP-350 was completely covered with nanosheets of MoS2.226 Besides, copper phosphate nanowires enriched with phosphorus were reported for use in lithium-ion batteries. XRD pattern of CuP2 nanowire powder in this form has structures identical to those of AgP2 with the same space group (P21/c). All the diffraction peaks easily characterized the crystalline monoclinic form of CuP2 with the following lattice parameters: a = 5.802 Å, b = 4.807 Å, c = 7.525 Å, and β = 112.68°; JCPDS 65-1274. Since CuP2 is a 10-member ring polyphosphide with a shared edge, some XPS analytical results offered evidence that the CuP2 nanowires remained stable and oxide-free after being stored in air over a period of more than one month.227 The triple-layer ternary metal phosphide has been among the layered morphologies that have been utilized to synthesize low-cost efficient bifunctional water-splitting electrocatalysts. Triple-layered ternary metal phosphorite is a typical one of the stacked morphologies that have been utilized in the synthesis of efficient and low-cost bifunctional water-splitting electrocatalysts Ni0.51 Co0.49 P film has been thoroughly explained in detail. Particles are hundreds of nanometers in diameter, and the shapes of the particles are asymmetrical. The mass transfers the particles to form a uniform type of sheet where the charges are permeable. The particles' surface, where the vertically ordered nanosheets are closely stacked and aligned.228 In molybdenum disulfide/nickel phosphide, we can observe that XRD patterns of the MoS2/Ni2P hybrids are presented in curves, but the diffraction patterns of MoS2/Ni2P hybrids consist of a very weak diffraction peak of MoS2, which indicates that the MoS2 is amorphous. In addition, the diffraction peaks related to MoS2 and Ni2P are observed, thus suggesting the successful combination between MoS2 and Ni2P. It may be indexed to proof of the fabrication of the MoS2/Ni2P hybrid. Compared with pure Ni2P, the diffraction intensity of Ni2P crystal for three MoS2/Ni2P hybrids becomes weak. However, with the increase of Ni2P addition, intensities of diffraction peaks attributed to Ni2P become strong, and those of MoS2 show a descending trend, which indicates that Ni2P plays a major role in the crystallinity of the hybrids.229,230
3.2 Electrical conductivity & charge transfer TMP@MoS2
The electrical conductivity and charge transfer behavior of TMP@MoS2 heterostructures are essential to their electrochemical activity. While the intrinsic bandgap (∼1.8 eV) of MoS2 results in low conductivity, the presence of metallic TMPs (e.g., FeP, CoP, Ni2P) enhances the charge transport through minimizing interfacial resistance and enhancing electron mobility. Research has shown that the heterostructures of FeP/MoS2 surpass those of individual FeP and MoS2, owing largely to the synergistic interaction between the semi metallic nanoparticles of FeP and the 2D ultrathin nanosheets of MoS2, acting as a conductive substrate and promoting the transfer of charges. Further, the electrochemical impedance spectroscopy (EIS) verifies that the charge transfer resistance (RCT) in the case of FeP/MoS2 is much lower compared to that in pure MoS2, indicating enhanced electron transfer. Further, the turnover frequency (TOF), the density of catalytic sites, and the electrochemical surface area (ECSA) are higher in the case of FeP/MoS2, which results in enhanced hydrogen evolution reaction (HER) activity. In spite of the lower ECSA due to the possible agglomeration of FeP, the better intrinsic conductivity makes up for the disadvantageous aspect. Density functional theory (DFT) calculations also corroborate the findings, with the Gibbs free energy (|ΔG|H) value for the case of FeP/MoS2 closer to the optimum value, thus increasing the efficiency in catalysis. These findings prove that the judicious dispersion of the TMP on the substrate of MoS2 optimizes the conductivity as well as the dynamics of the transfer of charges, making TMP@MoS2 a suitable candidate for applications involving energy conversion.231
3.3 Mechanical properties & structural stability
Mechanical stability in transition metal phosphides (TMPs) is dominated by stoichiometry and bonding in the M–P system. TMPs exhibit both covalency and ionic bonding, making them thermally and chemically stable with high hardness. In metal-rich phosphides (x
:
y ≥ 1), metallic conductivity, as well as even superconducting properties, develop due to the occurrence of M–M interactions. High-phosphorus-content phosphides (x
:
y < 1), on the other hand, are characterized by low electrical conductivity and lower stability due to the lack of M–M bonding. Nickel phosphides (NixPy), for example, demonstrate varying Ni–P coordination with varying P content. An increase in the content of phosphorus leads to the weakening of the Ni–Ni interactions, increasing the distance between the bonds and the formation of P–P dimers. While Ni2P exhibits higher conductivity, Ni12P5 displays greater resistance to corrosion induced by an acidic environment. These phenomena demonstrate the contribution made by stoichiometry and bonding towards the provision of structural stability for catalysis and for purposes related to energy.232 Mechanical properties in MoS2 depend on thickness, structural phase, and pressure. In-plane properties describe monolayers, but stability and function in the case of MoS2 depend on interlayer interactions in the bulk. MoS2 exists in various phases (2H, 3R, 1T, 1T′, 1T′′) with different electronic and catalytic properties. While the stable semiconductor form exists for the case of 2H–MoS2, the higher charge transfer properties for the case of 1T-MoS2 make it suitable for use in electrochemical applications. Pressure also affects the electronic and mechanical properties of MoS2, making it suitable for lubrication and the storage of energy. Since experimental syntheses of good-quality crystals are difficult, simulations have to be performed computationally to predict mechanical behavior of MoS2. While it is clear that pressure-induced changes in 2H MoS2's properties exist, their effect on other phases is poorly understood, suggesting that more research is needed.233
3.4 Thermal & chemical stability of TMP@MoS2
In catalysis and energy storage, the stability of the TMP@MoS2 heterostructures is the determinant of their efficiency and lifespan.9 Their long-term durability relies on various factors, including surface property, stability performance, and corrosion resistance.102
3.4.1 Phosphidation process & high-temperature effects. The phosphidation process, which forms a critical step in the synthesis of transition metal phosphides (TMPs), takes place at elevated temperatures.234 The high-temperature process impacts the crystallinity, phase composition, and the material's morphology.94 If well controlled, the phosphidation process leads to a highly ordered heterostructure with active sites dispersed uniformly without degrading the layered architecture of the MoS2.235,236
3.4.2 Oxidation resistance. MoS2, with its antioxidative properties, forms a protective shell that prevents oxidation of the TMPs further on.237,238 Passivation enhances the long-term stability of the TMP@MoS2 in the area of electrochemistry and catalysis.239
3.5 Surface area & active site exposure
3.5.1 BET surface area analysis & MoS2 growth. The increase in surface area presents more active sites to the environment, consequently increasing catalytic performance.240 The BET surface area analysis can also be carried out to investigate the role of MoS2 nanosheets in enhancing the porosity and surface area.241 Moreover, the 2D building block nature of MoS2 yields a highly exposed surface that is favorable for charge and ion transport in energy storage applications.242
3.5.2 Edge site exposure & catalytic performance. Compared with the inert basal plane, the edge sites on the MoS2 are more catalytically active.243 Vertical growth of the MoS2 in the heterostructures increases the number of edge sites, which considerably enhances the hydrogen evolution reaction (HER) activity.244 Edge-dominant catalysis has a significant function in the synthesis of high-performance electrocatalysts.245
3.6 Hydrophilicity & wettability in electrochemical reactions
The hydrophilicity of the electrode materials influences their interaction with the electrolyte and, thus, the electrochemical reactions.246 Hydrophilicity increases the diffusion of ions and the transfer of charges, and the efficiency of the overall electrocatalysis becomes higher.247 Supercapacitors, batteries, and fuel cells benefit from optimized kinetics through the better wettability found in MoS2-modified electrodes.248 Since both Sn4P3/Co2P and MoS2 possess this characteristic, a heterostructure with both compounds would be expected to result in better wettability.247,249
4. Characterization techniques
4.1 Physical characterization
4.1.1 X-ray diffraction (XRD). X-ray diffraction (XRD) is a critical technique for the identification of the crystalline phases of the TMPs (e.g., FeP, CoP)232 and the MoS2 (1T/2H polytypes) by the indexation of the peaks, with respect to the reference data, for example, JCPDS 37-1492 for the 2H–MoS2.250 XRD also plays an important role in the stability analysis under thermal and electrochemical cycling conditions.251 The example of XRD analysis results is provided in Fig. 8 in this review paper.
4.1.2 Raman spectroscopy. Raman spectra provide a distinct indication of vibrational modes associated with TMP@MoS2 heterostructures.192 Raman spectra of the composites, as well as pure MoS2, offer valuable information about their structural interaction as well as disorder levels. The appearance of characteristic E12g (381 cm−1) and A1g (406 cm−1) Raman peaks indicates the presence of MoS2, and any shift in the peaks indicates the level of interaction with the supporting matrix.252 Raman's characterizations also prove the existence of CoP. The 367 and 394 cm−1 vibration peaks refer to the in-plane and out-of-plane vibrations of the Mo–S bond, respectively. Moreover, the vibration peak around 228 cm−1 is attributed to the Ag. Mode of the Co–P bond, indicating the existence of CoP on the material surface.253 Raman spectroscopy is a powerful analytical technique for the structural interaction and vibrational properties analysis of TMP@MoS2 heterostructures. The occurrence of characteristic MoS2 and CoP peaks confirms their presence and interaction within the composite, demonstrating the power of the Raman analysis in the analysis of the bonding and the material composition.
4.2 Morphological and microstructural analysis
4.2.1 Scanning electron microscopy (SEM). The morphologies of the samples prepared are normally examined using SEM, and FESEM, and transmission electron microscopy (TEM). SEM findings of the hierarchical tubular MoP/MoS2 composite include information about its tubular nature.254 Additionally, it provides information on particle size, such as Molybdenum disulfide-coated nickel–cobalt sulfide with a nickel phosphide core–shell structure.255
4.2.2 Transmission electron microscopy (TEM). Transmission Electron Microscopy (TEM) is commonly used to analyze the structural and morphological properties of TMP@MoS2 heterostructures. Additionally, Energy Dispersive Spectroscopy (EDS),256 in combination with TEM, provides elemental composition information.257For instance, in syntheses of molybdenum phosphide (MoP) nanorods with graphene assistance on silicon, high-resolution TEM (HR-TEM) micrographs indicate distinct fringes of a well-defined lattice with a spacing of 0.278 nm corresponding to hexagonal Mop's (100) plane. Furthermore, the fast Fourier transform (FFT) pattern also confirms the high crystallinity of MoP nanorods to authenticate their structure.258
4.3 Chemical and electronic structure
4.3.1 X-ray photoelectron spectroscopy (XPS). X-ray photoelectron spectroscopy is a powerful method for the study of various phenomena in gases, liquids, and solids. X-ray photoelectron spectroscopy provides information on local physical phenomena in X-ray photoemission, facilitating spectrum modeling and derivation of vital information such as core-level line shapes, binding energies, built-in potentials, and band-edge discontinuities in complex oxide heterostructures. The binding energy and shape of a photoemission peak are influenced not only by the atomic number, valence, and orbital of the ejected electron but also by intricate many–body interactions that occur during the photoemission process.259XPS analysis showed CoP@MoS2-75. The binding energies of Mo 3d and S 2p in CoP@MoS2-75 were displaced from those of pure MoS2. The Mo and S shifts were approximately consistent (∼1 eV) due to the presence of Mo–S bonds, which is attributable to the strong contact between MoS2 and CoP. The binding energies for the Mo 3d and S 2p in the CoP@MoS2-75 were shifted away from those for pure MoS2. The shifts for the Mo and S were about the same value (∼1 eV) as a result of the presence of the Mo–S bonds due to the close contact between the MoS2 and the CoP. The shift in binding energy and the other fitting peaks in the Co–S bands in the spectra for the Mo 3d and the S 2p (Fig. 6(a) and (b)) attest to the strong bonding contact between the CoP and the MoS2, possibly enhancing the HER activity of the MoS2.15
4.4 Electrochemical characterization
This evaluation typically employs a three-electrode system comprising a saturated calomel reference electrode, a platinum plate counter electrode, and the as-prepared working electrode. These techniques are cyclic voltammetry (CV), galvanostatic charge–discharge (GCD), and electrochemical impedance spectroscopy (EIS). They can be used to evaluate the electrochemical performances of supercapacitors,260 electrocatalysts,261 and batteries.262 These methods are beneficial to detect improvement of electrochemical performances in heterostructure; for example, the CV curves of FeCoP/CC, NiCoP/CC, and FeCoP@NiCoP/CC electrodes at a scan rate of 5 mV s− 1 display a growth on redox peaks with apparent potential separation. Additionally, the GCD curves exhibit a longer discharge time in the case of FeCoP@NiCoP/CC. EIS can evaluate the reaction kinetics and the electrical conductivity. In this case, the FeCoP@NiCoP/CC electrode exhibits the lowest Rs value of 0.624 Ω, compared to 0.638 Ω for the FeCoP/CC electrode and 0.655 Ω for the NiCoP/CC electrode. These results demonstrate the significant improvement in electrochemical performance achieved through heterostructures.263 Therefore, all these measurements are crucial for evaluating the performance and efficiency of energy storage equipment.
4.4.1 Cyclic voltammetry (CV). CV is a fundamental technique to examine the redox behavior, capacitance, and charge storage mechanism of TMP@MoS2 heterostructures.264 In this technique, cyclic sweeps of the working electrode occur at a pre-set voltage range, and the current is recorded.265 The curvature of the CV curves provides valuable information on charge storage kinetics, pseudocapacitive contribution, and reversibility of reactions.266–268For TMP@MoS2 materials, their CV curves will show a combination of faradaic redox peaks and non-faradaic double-layer charging dependent on the heterostructure composition.269,270 The scan rate dependency of the CV curves also provides information on charge transfer kinetics with higher scan rates leading to higher capacitive currents.271
4.4.2 Galvanostatic charge–discharge (GCD). GCD measurements consist of subjecting working electrodes to current densities that are constant and measuring voltage response with time.272 The technique provides valuable information on specific capacitance, energy density, power density, and cycling stability.273 The charge–discharge behavior of TMP@MoS2 heterostructures can be linear (capacitive) or nonlinear (pseudocapacitive/faradaic).274 Higher capacitance is seen in a longer discharge time at a given current density.275 Cycling stability of TMP@MoS2 electrodes is determined by subjecting electrodes to repeated charge–discharge cycling and monitoring capacity retention with time.276
4.4.3 Electrochemical impedance spectroscopy (EIS). Electrochemical Impedance Spectroscopy (EIS) is utilized to examine charge transfer resistance, ionic diffusion kinetics, and interfacial properties of TMP@MoS2 heterostructures.277–279 EIS measurements include imposition of AC voltage perturbations over a range of frequencies and measurements of resultant current response.280 The impedance data is represented in a Nyquist plot with high-frequency semicircle due to charge transfer resistance (RCT), and low-frequency tail due to diffusion phenomena of ions.281 EIS fitting optimizes electrode structure to increase charge transfer and decrease resistance to increase electrochemical performance in batteries and supercapacitors.282
5. General mechanisms of heterostructures in energy storage and catalysis
5.1 General mechanisms in supercapacitors
Energy storage has been the backbone of all renewable energy systems in modern times, acting as a link between generation and consumption.283 Supercapacitors have emerged as a critical component because they can store energy rapidly and release it with a very long cycle life.284 This differs from batteries in that supercapacitor energy storage is done through purely electrostatic mechanisms, consequently allowing faster charge and discharge processes.285 This section describes the supercapacitor performance based on electrical double-layer capacitance, pseudocapacitance, and a hybrid mechanism for their potential role in developing energy storage solutions for high-power and sustainable applications. In general terms, there are three ways through which energy is stored in all electrochemical supercapacitor types, which are introduced below.
5.1.1 Electrical double-layer capacitance (EDLC). The electrical double-layer capacitor (EDLC) is an electrolytic energy storage device that works at the interface between an electrode and an electrolyte. When voltage is applied, ions from the electrolyte collect near the electrode surface and produce two oppositely charged layers. Because the space is so small, the capacitance value becomes very high (Fig. 7). This is why EDLCs are very useful in applications requiring quick storage and energy release, such as quick bursts of power supply.287
 |
| Fig. 7 The structure of electrical double-layer capacitance. Reproduced from ref. 286 with permission from [IOP Science], copyright [2022]. | |
5.1.2 Pseudocapacitance. Pseudo-capacitance mainly involves energy storage via fast reversible redox reactions at or close to the electrode's surface. In contrast, EDLCs store energy through electrostatic charge separation, while pseudocapacitance lies in between where faradaic electron transfer across the electrode–electrolyte interface is supposed to occur (Fig. 8). This type of behavior has been observed in nanomaterials, conducting polymers, and transition metal oxides.289
 |
| Fig. 8 The mechanism structure of pseudocapacitance. Reproduced from ref. 288 with permission from [Royal Society of Chemistry], copyright [2021]. | |
5.1.3 Hybrid supercapacitors. Hybrid supercapacitors have gained the attributes of batteries and symmetric supercapacitors, with the ability to store bulk energy and power capability. This device drives faradaic and non-faradic processes that again eliminate the inadequacy of every technology. In this way, developing new electrode materials in the future, improving electrolytes, and trying asymmetric electrode configurations will become essential strategies (Fig. 9) for further performance enhancement to achieve their applicability in all forms of energy storage devices.290
 |
| Fig. 9 The structure of hybrid supercapacitors, Reproduced from ref. 288 with permission from [Royal Society of Chemistry], copyright [2021]. | |
5.2 General mechanisms in batteries
Batteries operate on the principle of the electrochemical mechanisms converting chemical energy to electrical energy. The major electrochemical mechanisms in batteries are classified into different processes: redox reactions (reduction–oxidation reactions), also called faradaic reactions,291 generate current through a redox reaction on the working electrode surface. The second process is called ion transport through the electrolyte,292 whereby, upon discharge, ions flow through the electrolyte from one electrode to the other to balance the charge made by the electron flow in the external circuit.293 The type of ions that flow depends on the chemistry of the battery. For example, in lithium-ion batteries, lithium ions (Li+) migrate from the anode to the cathode during discharge.294 The third mechanism is electrode processes. In sure batteries, like lithium-ion batteries, the active ions (e.g., lithium ions) are inserted into or extracted from the electrode materials during charging and discharging.295 Here, two phenomena are involved: Intercalation, the insertion of ions inside the electrode matrix, usually during the charging process,296 and deintercalation, the removal of ions from the electrode matrix, usually during the discharging process.297,298 And the electrical double layer (non-faradaic processes) does not include the transfer of charge and occurs inside the electrode.299 During the process, the charges of the ions accumulate on the surface of the working electrode, forming and discharging a double-layer capacitance. The attachment of the target biomarker on the electrode surface modifies the dielectric constant of the double-layer capacitance (Fig. 10 and 11).
 |
| Fig. 10 Faradaic reaction process in batteries, Reproduced from ref. 300 with permission from [JEC], copyright [2021]. | |
 |
| Fig. 11 Ion transport in batteries, Reproduced from ref. 301 with permission from [Nature], copyright [2021]. | |
5.3 General mechanisms in electrocatalysts
5.3.1 General mechanism of HER in acidic and alkaline media. The hydrogen evolution reaction is significant in water electrolysis and is a substantial process for sustainable hydrogen production.302 Hydrogen, a clean energy carrier, is vital on the road to renewable energy systems for global energy demands.303,304 In this respect, understanding the HER mechanism is of prime importance in designing efficient catalysts that operate effectively in different media.305,306 HER has to overcome complex problems, such as the dissociation of water in highly alkaline conditions, when its kinetics are faster in an acid environment.305 These sections discuss the mechanisms of acid and alkaline mediums with and upon metal surfaces and illustrate the balance of adsorption and desorption processes toward efficiency and applicability thanks to catalysts.HER serves as the other half of the water electrolysis process. In an acidic environment, hydrogen atoms are absorbed onto the surface of the applied catalyst during the hydrogen evolution reaction (HER), later combining to form molecular hydrogen. The process is relatively efficient and safe.
|
Low Hads coverage: Hads + e− + H+ (aq) → H2 (g)
| (2) |
|
High Hads coverage: Hads + Hads → H2 (g)
| (3) |
On the other hand, she happens in a much more complex way in alkaline conditions. In this case, water first has to dissociate its molecules to provide the protons needed in the reaction, so the process is done step by step, therefore taking time compared to acidic conditions. Thus, the most efficient catalysts for HER need to take control of the adsorption and desorption of the intermediate molecules at the catalyst surface under alkaline conditions.
|
Volmer step: H2O (aq) + e− → Hads + OH−
| (4) |
|
Heyrovsky step: H2O (aq) + e− + Hads → H2 + OH−
| (5) |
|
Tafel step: Hads + Hads → H2 (g)
| (6) |
Although the pathways followed in acidic and alkaline media differ, HER is the central process of water electrolysis. HER is highly efficient in acidic media, and the reverse is the case in alkaline media, where water dissociation proves deleterious to HER. To develop a catalyst that can proficiently handle the adsorption and desorption of molecular species under an alkaline condition, an effective HER is imperative in further developing this technology.307–309
Following the comprehensive examination of HER mechanisms in Section 3, the focus will shift to the corresponding half-reaction in water electrolysis—the oxygen evolution reaction (OER). As the counterpart to HER in hydrogen production, OER generates oxygen, playing a crucial role in shaping the overall efficiency of the electrolysis process. A thorough understanding of HER mechanisms also sheds light on the obstacles and potential strategies for optimizing OER, as both reactions rely on energy-efficient rates driven by the ideal balance of adsorption and desorption characteristics.
5.3.2 General mechanism of OER in acidic and alkaline media. While protons in acidic mediums may be favorable for the reaction, hydroxide ions under alkaline conditions grow into different problems.310,311 The following section describes the mechanism of OER in both acidic and alkaline conditions and how advanced catalysts have significantly improved reaction efficiency and stability toward sustainable energy technologies.The most likely mechanism for reactions in acidic media is as follows:
|
2H2O → O2 + 4e− + 4H+ the overall reaction
| (7) |
|
2H2O + * →OH* + H2O + H+ + e− (ΔG1)
| (8) |
|
OH* + H2O → O* + H2O + H+ + e− (ΔG2)
| (9) |
|
O* + H2O → OOH* + H+ + e− (ΔG3)
| (10) |
|
OOH* → O2 + H+ + e− (ΔG4)
| (11) |
The oxygen evolution reaction (OER) is a complex process involving four electron transfers. It occurs gradually, sequentially forming intermediate compounds OH*, O*, and OOH*. It is relatively slow. Many articles have investigated the detailed mechanisms of the OER in acidic or alkaline environments.
In alkaline media:
|
4OH− → O2 + 2H2O + e− the overall reaction
| (12) |
|
4OH− → OH + 3 OH−+ e− (ΔG5)
| (13) |
|
OH* + 3OH → O* + 2OH− + H2O + e− (ΔG6)
| (14) |
|
O* + 2OH− + H2O → OOH* + OH− + H2O + e− (ΔG7)
| (15) |
|
OOH* + OH− + H2O → O2 + H2O + e− (ΔG8)
| (16) |
where * is the active site, OH*, O*, and OOH* are the intermediates adsorbed on the site, and
G is the change in Gibbs free energy.
312
6. Specific mechanisms of TMP@MoS2 in different applications
The TMP@MoS2 heterostructures' high efficiency in catalysis and energy storage results due to the synergistic structural and electronic properties in the heterostructures. Transition metal phosphides (TMPs) possess good electrical conductivity, while molybdenum disulfide (MoS2) exhibits high surface area, chemical stability, and active site richness. The mechanisms underlying their efficiency are due to the increased charge transfer, interfacial synergy, and optimized adsorption and desorption kinetics of the ions.
In batteries, the presence of the TMPs enhances the diffusion of the sodium/lithium ions by expanding the interlayer spacing and lowering the resistance to the transportation of the ions, whereas the storage of the charges through the intercalation and the conversion reactions is facilitated by the presence of the MoS2. In the case of electrocatalysis, the electronic band structure of the MoS2 gets regulated by the presence of the TMPs, aligning the d-band center at the Fermi level to lower the energy barrier for the HER/OER. The interaction between the TMPs and the MoS2 becomes strong, enhancing the better adsorption of the reactant as well as the more efficient transfer of the electron, enhancing the efficiency of the catalyst tremendously. In the case of the supercapacitors, the pseudocapacitive storage and the fast diffusion of the ions through the TMP@MoS2 heterostructures enhance the cycling stability and the energy density. The synergy in the integrated system optimizes the storage of the ions, the redistribution of the charges, and the catalysis and is found to be highly suitable for the future generation of electrochemical applications.
6.1 Mechanisms in supercapacitors
Unlike most current literature, which attributes MoS2's capacitance primarily to electric double-layer capacitance, studies on current response and scan rates reveal that the capacitance of few-layered MoS2 mainly originates from intercalation pseudocapacitance. Integrating TMPs further enhances electrical conductivity and introduces additional redox-active sites, improving capacitance and energy density.313 The conductivity is enhanced with the addition of TMP, and more redox sites are added, thereby enhancing the capacitance and the energy density.314
Hybrid supercapacitors (HSCs) are drawing attention due to their ability to combine two unlike electrodes with different mechanisms for the storage of charges with high energy density without compromising power output. Transition metal phosphides, and specifically bimetallic nickel cobalt phosphide (NiCoP), are promising as the HSC positive electrode due to their redox sites, high reversibility in the electrochemical process, and stability even over extended periods of time. The electrochemical properties of NiCoP can be further optimized with rationally designed heterostructures, elemental doping, and nanocomposite morphologies to improve the efficiency in the storage of the charge in HSCs.315
As mentioned above, molybdenum disulfide (MoS2) is made by two sheets of S sandwiched between Mo by van der Waals forces. This configuration creates a large surface area and provides the intercalation of ionic species with the electrolyte without causing any crystal structure deformation, resulting in pseudocapacitive charge transfer.316 So, Combining MoS2 with TMPs in hybrid supercapacitors leverages the high-power density of capacitors and the high energy density of batteries. The result is products with improved performance parameters for a range of applications in multiple areas of energy storage. This simple principle increases the natural conductivity, active sites, and synergistic interactions among the individual units of the TMP@MoS2 heterostructure, making the latter more efficient in the storage and conversion process of energy (the excellent performances of the new-generation batteries, electrocatalysts, and supercapacitors combine the best performances of the three factors).
6.2 Mechanisms in batteries
One of the benefits of transition metal phosphides on batteries is ensuring abundant adsorption–diffusion–conversion interfaces for accelerating LiPS transformation and Li2S deposition, which cause to extremely decreases the accumulation of LiPSs in the electrolyte and, therefore, prevents the migration of LiPSs. The ultrafine Ni2P nanoparticles prove this is happening.317 Also, they suggested good electrical conductivity and dual adsorption-conversion capabilities. By increasing electrical conductivity and providing additional active sites for lithium insertion, they are considered a promising cathode host for new-generation LSBs.318 By using molybdenum disulfides, the less stacked layers can lead to a large specific surface area, enhancing electron transfer and exposure of active sites.319 Moreover, it was mentioned that TMPs contribute to improved capacity and cycling stability.320
In Sodium-Ion Batteries (SIBs) similar to LIBs, transition metal phosphides lead to expanded interlayer spacing, accommodating larger sodium ions and enhancing diffusion kinetics. This structural modification leads to improved electrochemical performance.321
6.3 Mechanisms in electrocatalysts
The exceptional catalytic activity of the transition metal phosphides (TMPs) arises due to their optimized electronic structure, rich active sites, as well as efficient electron transfer capability. Phosphorus adjusts the d-band center and the Fermi level, promoting reactant adsorption and desorption of the products. The HER is regulated by the P atoms and is enhanced by oxy/hydroxides or phosphates that are created in situ to enhance OER. Defect sites are introduced by P doping that enhances active site concentration. Metal-rich phosphides (e.g., Co2P) are more conductive and enhance catalytic efficiency. The blending of the TMPs with MoS2 creates a heterostructure that has high conductivity as well as rich active sites, enhancing catalytic activity and stability towards advanced electrocatalysis.146
The blend of the TMP with the MoS2 forms a heterostructure that takes advantage of the high conductivity of the TMP and the richness in active sites of the MoS2. The catalytic activity and stability are enhanced through this combination, and it has the potential to be a new way forward for advanced electrocatalysis. The nature of the active sites has been treated in Coordination Chemistry Reviews (Volume 506, 1 May 2024, 215715) in the review article The Nature of Active Sites of Molybdenum Sulfide-Based Catalysts for Hydrogen Evolution Reaction by Weifeng Hu and coauthors.322
7. Electrochemical performance of TMP@MoS2
The electrocatalytic performance of MoS2/TMP heterostructures is enhanced through interface engineering and charge redistribution, thereby optimizing the hydrogen evolution reaction (HER)15 and oxygen evolution/reduction reactions (OER/ORR).323 The metallic 1T-MoS2 phase in TMP@MoS2 facilitates hydrogen adsorption (ΔG_H* ≈ 0.06 eV) via sulfur vacancies and strained Mo–S bonds.324 Notably, MoS2@CoP/CC exhibits low overpotentials of 64 mV and 282 mV for HER and OER, respectively, in alkaline solution, along with a HER overpotential of 72 mV@10 mA cm−2 in H2SO4. Furthermore, P–MoS2@CoP/CC, as a bifunctional catalyst, delivers relatively low cell voltages of 1.83 V and 1.97 V@500 mA cm−2 in 30% KOH, demonstrating high catalytic efficiency for overall water splitting.325 Tafel kinetics confirm a Volmer–Heyrovsky mechanism,326 and flower-like Co–Ni–P/MoS2 heterostructure hybrid spheres show excellent overall water splitting performance in an alkaline solution with a low Tafel slope of 71 and 41 mV dec−1 for oxygen and hydrogen evolution reactions.323
Specific studies on MoS2/metal phosphide heterostructures for supercapacitors are limited, but phosphorus doping in MoS2 has shown promising results. However, our previous research paper introduced nickel cobalt phosphide/Molybdenum disulfide on nickel foam as an effective electrode material in a supercapacitor. The capacitance achieved was 2352.40 F g−1 at 1 A g−1, which is exceptionally high. Its high energy density (52 W h kg−1 at 321 W kg−1) and power density of 321 W kg−1 made it great for applying supercapacitors.20 Furthermore, researchers recently suggested that phosphorus-doped molybdenum disulfide regulated by sodium chloride be used for advanced supercapacitor electrodes. This study was synthesized using phosphomolybdate acid as a molybdenum source and an in situ dopant and sodium chloride (NaCl) as a structural regulator. A maximum capacity of 564.8 F g−1 at 1 A g−1 and retaining 56.3% of the original capacity at 20 A g−1 was achieved.327 For this reason, this review paper was suggested to give information on the gap research of this heterostructure on supercapacitors and batteries. This can be the novelty of this work and study on the heterostructure of TMP@MoS2.
A flower-like heterostructured MoP–MoS2/PCNFs hierarchical nanoreactor was introduced in Li–S batteries, which enhanced kinetics. This study revealed that this composite shortened the lithium-ion channel, shuttle, and fast LiPS conversion ability due to abundant anchoring sites of MoP–MoS2 heterojunction nanoflowers.12 This heterojunction exhibits 1090.02 mA h g−1 and a high discharge capacity of 884.67 mA h g−1 even after 300 cycles at 1C, which compared to molybdenum disulfide alone with a high reversible discharge capacity of up to 994.6 mA h g−1 for the MoS2-1 electrode and 930.1 mA h g−1 for the MoS2-2 electrode is incredibly high.328
8. Energy storage applications of TMP@MoS2
8.1 Lithium/sodium-ion batteries
Bimetallic phosphide Ni2P/CoP@rGO heterostructure was used in batteries, and this heterostructure delivers an ultrahigh capacity of 196.4 mA h g−1 at 10 A g−1 after 5000 cycles for lithium-ion batteries and 103.7 mA h g−1 at 3 A g−1 after 800 cycles for sodium-ion batteries.329 Yolk–shell tin phosphide composites demonstrated superior electrochemical performance. The capacity of Li-half cells was 521.2 mA h g−1, which was maintained after 3000 cycles at 5.0 A g−1. This amount for Na-half cells is 203.1 mA h g−1 maintained after 300 cycles at 1.0 A g−1.330 Micro–nanostructure designed CoP@MoS2 delivers a high initial discharge capacity (1321 mA h g−1 at 0.1C), high rate capability (837 mA h g−1 at 2C), and stable cycling performance (0.101% capacity decay after 250 cycles at 0.5C), suggesting great application prospects of the micro–nanostructure catalyst in Li–S batteries.16 CoP–C@MoS2/C heterointerface enhanced Lithium/Sodium Storage exhibits outstanding long-cycle performance of 369 mA h g−1 at 10 A g−1 after 2000 cycles. In SIBs, the composite also displays an excellent rate capability of 234 mA h g−1 at 5 A g−1 and an ultra-high-capacity retention rate of 90.16% at 1 A g−1 after 1000 cycles.331 MoS2/MoP Mott–Schottky heterostructures in lithium–sulfur batteries as promising material deliver an initial capacity of 919.5 mA h g−1 with a capacity of 502.3 mA h g−1 remaining after 700 cycles at 0.5C. Even under higher sulfur loading of 4.31 mg cm−2 and a lower electrolyte to sulfur (E/S) ratio of 8.21 μL mg−1, the MoS2/MoP@rGO@S cathode could still achieve good capacity and cycle stability.332 These observations show the huge promise of metal phosphides when incorporated with MoS2 for improving energy related devices based on lithium-ion, sodium-ion, and lithium–sulfide batteries with much greater electrochemical properties. The improved capacity, cycling stability, and high-rate capability validate the effective implementation of metal phosphides with MoS2 in tackling the larger issues for conductivity and structural stability, offering possible solutions for the evolution of novel energy-storage devices.
8.2 Supercapacitors
Several surveys have been conducted on the heterostructure of transition metal phosphides with metal sulfides; however, few have been explicitly related to molybdenum disulfide. An example is NiCoP@CoS tree-like core–shell nanoarrays on nickel foam, which serve as high-performance electrodes for supercapacitors. These NiCoP nanowire-based electrodes with electrodeposited shell layers of CoS nanosheets possess a high value of specific capacitance of 1796 F g−1 at a current density value of 2 A g−1 with a high value of cycling stability as 91.4% retention for 5000 cycles. The asymmetric supercapacitor based on NiCoP@CoS and activated carbon has an energy density value of 35.8 W h kg−1 at a power density value of 748.9 W kg−1.333 Yet another instance: the use of Ni–Co oxide/phosphide/sulfide (NCOPS) composites in nanowire arrays on Ni foam exhibits a specific capacitance value of 2915.6 F g−1 with a retention value of 80.39% of the specific capacitance value for 4000 cycles of constant current as 5 A g−1.334 Further, the hierarchical porous heterostructure Ni2P/NC@CoNi2S4 with a high value of specific capacitance as 2499 F g−1 at the value of current density as 1 A g−1 is another instance, with a capacitance retention value as 91.89% for the value of 10
000 cycles. In this case, it exhibits an impressive value of value of energy density of 73.68 W h kg−1 at the value of power density of 700 W kg−1.335 However, few papers on the heterostructures of the transition metal phosphide/molybdenum disulfide are found. Out of those, we recently reported the electrochemical synthesis of a hybrid nanostructure of nickel cobalt phosphide/molybdenum disulfide on nickel foam, as reported in our previous research work.20 These findings reflect the impressive advancement in the transition metal phosphide/sulfide heterostructures towards the application in the field of supercapacitors. However, the scarcity of papers on the phosphide/molybdenum disulfide heterostructures reflects the novelty in the current work, establishing the unique promise of NiCoP/MoS2 towards high-performance energy storage devices.
8.3 Electrocatalysts
8.3.1 Hydrogen evolution reaction (HER). Fortunately, there is much more research about TMP@MoS2 and its application to the Hydrogen Evolution Reaction (HER). Hierarchical cobalt–nickel phosphide/molybdenum disulfide is a good example related to HER, and the results showed that this heterostructure led to improved electrochemical performance. The optimized CoP/Ni2P/MoS2–CC sample shows low overpotentials of 335, 119, and 211 mV at 100 mA cm−2 in neutral, alkaline, and acidic electrolytes with small Tafel slopes of 63, 62, and 82 mV dec−1, respectively.13 Hollow CoP@MoS2 hetero-nano frames are another example that showed a low overpotential of 119 mV at 10 mA cm−2, a small Tafel slope of 49 mV dec−1, a large electric double-layer capacitance of 10.28 mF cm−2, and prominent long-term stability.15 MoP/MoS2 heterogeneous structure with rich S-vacancy, which revealed excellent catalytic activity and good cyclic stability, showing a lower Tafel slope of 60 mV dec−1 at a current density of 10 mA.cm−2 and no potential attenuation after the cyclic stability test for 24 h.336 S-scheme boron phosphide/MoS2 heterostructure in water splitting photocatalysts application.337 These results demonstrate the vast potential of TMP@MoS2 heterostructures towards the Hydrogen Evolution Reaction (HER), with their high catalytic activity, low overpotentials, and exceptional stability. The range of reported structures demonstrates the versatility of TMP@MoS2 in electrochemical applications, confirming its significance in sustainable hydrogen production.
8.3.2 Oxygen evolution reaction (OER). The research about this heterostructure in Oxygen Evolution Reaction (OER) is not as much as Hydrogen Evolution Reaction (HER). MoS2‖CoP heterostructure loaded on N, P-doped carbon. The excellent electrocatalytic activity was expressed and also mentioned. This work proposes a novel and facile strategy to prepare the heterostructure compound and serves as a good reference for constructing efficient and low-cost electrocatalysts.194 Flower-like HEA/MoS2/MoP exhibited excellent HER and OER electrocatalytic performance. It showed a low overpotential of 230 mV at the current density of 10 mA cm−2 for OER and 148 mV for HER in alkaline electrolytes, respectively.338 Ni2P–MoS2 HNSAs/CC sample as both anode and cathode for overall water splitting requires an impressively low onset potential of only 1.574 V to attain a current density of 10 mA cm−2 and displays excellent long-term stability. The facile synthesis method and insights into the HER and OER active interfaces reported here will advance the development of high-performance bifunctional overall water-splitting electrocatalysts. They can be used for both OER and HER.339 Although research on TMP@MoS2 heterostructures for the Oxygen Evolution Reaction (OER) is less extensive than for the Hydrogen Evolution Reaction (HER), existing studies demonstrate promising catalytic performance and stability. These findings highlight the potential of TMP@MoS2-based materials as efficient and cost-effective bifunctional electrocatalysts for overall water splitting, paving the way for future advancements in sustainable energy conversion.
9. Comparative electrochemical performance of TMP@MoS2 heterostructures
9.1 Comparison with metal phosphides/oxides/sulfides in electrocatalysts
Table 4 compares the HER and OER electrocatalytic activity for the TMP@MoS2 heterostructure with that for the standard metal phosphides, metal oxides, and metal sulfides. The surface properties of the metal phosphides and the MoS2 thus work in synergy to enhance their catalytic properties, as evidenced by lower overpotentials for HER and OER. It was found that MoP@MoS2 demonstrated higher hydrogen evolution than MoP, while CoP@MoS2 has the lowest overpotential among oxygen evolution catalysts compared to most of those where rGO supported CoP. Compared to most metal oxides, the TMP@MoS2 heterostructure has lower cell voltage values for the water-splitting process and remains constant among all others. However, NiP/MoS2 heterostructures exhibited high stability and distinctive capabilities, even under the most rigorous conditions. Therefore, TMP@MoS2 heterostructure could become one of the best options for renewable energy applications. These hetero-crystals are prospective materials to boost HER, OER, and water-splitting processes. They are also critical potential substances from energy conversion and storage, particularly activity stability.
Table 4 Comparing TMP@MoS2 heterostructures with metal phosphides/oxides/sulfides electrode materials in electrocatalysts
Catalyst |
Substrate |
Overpotential @10 mA cm−2 [mV] |
Voltage for overall @10 mA cm−2 [V] |
Tafel slope [mV dec−1] |
Stability |
Electrolyte |
Ref. |
ηHER |
ηOER |
Vm |
HER |
OER |
MoP/MoS2 |
Ni foam |
96 mV |
— |
1.98 |
48 |
— |
— |
1 M KOH and 1 M PBS |
340 |
MoP |
Ni foam |
114 |
265 |
1.62 |
34.4 |
56.6 |
— |
1 M KOH |
341 |
CoP@MoS2 |
CoCo-PBA |
201 |
— |
— |
70.4 |
— |
3000 cycles |
0.5 M H2SO4 |
342 |
CoP |
RGO |
58 |
284 |
1.66 |
58.3 |
116.4 |
— |
1 M KOH |
343 |
Ni2P/MoS2 |
— |
75 |
— |
— |
76 |
— |
8 h |
0.5 M H2SO4 |
344 |
Ni2P |
N-RGO |
80 |
— |
— |
93.1 |
— |
— |
0.5 M H2SO4 |
345 |
NiCoP/MoS2 |
Ni foam |
148 |
— |
— |
109 |
— |
Highly stable |
1 M KOH |
189 |
NiCoP |
Ni foam |
222 |
— |
— |
153 |
— |
— |
1 M KOH |
189 |
MoS2 |
Ni foam |
181 |
— |
— |
— |
— |
— |
1 M KOH |
189 |
FeP/MoS2 |
— |
110 |
— |
— |
67.8 |
— |
Good stability |
0.5 M H2SO4 |
231 |
FeP |
— |
210 |
— |
— |
76.5 |
— |
— |
0.5 M H2SO4 |
231 |
MoPxSy@NiFePxSy@NPS-C |
MoO3 |
— |
274 |
— |
— |
65.6 |
100 h |
1 M KOH |
346 |
MoS2 |
— |
265 |
— |
— |
86.2 |
— |
— |
0.5 M H2SO4 |
231 |
Co3O4/MoS2 |
Nickel foam |
205 |
230 |
— |
98 |
45 |
10 h |
1 M KOH |
347 |
MoS2–MoO3−x/Ni3S2 |
Nickel foam |
76 |
— |
— |
53.2 |
— |
17 h |
1 M KOH |
348 |
CoS2–MoS2 |
Carbon cloth |
60 |
240 |
1.52 |
86 |
68 |
24 h |
1 M KOH |
349 |
MoO2/MoS2/MoP |
— |
135 |
— |
— |
67 |
— |
1000 cycles |
0.5 M H2SO4 |
191 |
MoO2/MoS2/MoP |
— |
145 |
— |
— |
71 |
— |
1000 cycles |
1 M KOH |
191 |
MoS2/NiS2 |
Carbon cloth |
80 |
303 |
1.63 |
61 |
58 |
Excellent stability |
Alkaline solution |
350 |
MoS2/MoO2 |
— |
157 |
— |
— |
119 |
— |
Half-day |
1 M KOH |
351 |
9.2 Comparison with metal phosphides/oxides/sulfides in supercapacitors
Table 5 shows the comparison, indicating that the electrode materials that form the TMP@MoS2 heterostructures were much better when compared to those formed by metal phosphides, oxides, or sulfides. MoS2 increases the electrochemical properties of the materials and enhances the specific capacitance, the cyclic stability, and the energy density.
Table 5 Comparing TMP@MoS2 heterostructures with metal phosphides/oxides/sulfides electrode materials in supercapacitors
Metal phosphides |
Capacitance (F g−1) |
Cycling stability (%) |
Energy density (W h kg−1) |
Power density (W kg−1) |
Coulombic efficiency (%) |
Ref. |
MnP–MoS2 |
432.3 |
86.2 |
16.7 |
403.9 |
93.4 |
190 |
NiCoP@MoS2 |
2352.407 |
— |
52 |
321 |
— |
20 |
NiCoP |
1279.2 |
— |
45.5 |
124.2 |
— |
352 |
CoP |
447.5 |
96.7 |
19 |
350.8 |
— |
353 |
FeP |
149.11 |
41 |
2.02 mW h cm−3 |
9.02 mW cm−3 |
— |
354 |
Cu3P |
300.9 |
81.9 |
44.6 |
0.017 |
— |
355 |
Carbon@MoS2/MoO2 |
569 |
67.1 |
30.8 |
800 |
91.4 |
356 |
SnS2/MoS2 |
466.6 |
88.2 |
115 |
2230 |
— |
357 |
TiO2@MoS2 |
337 |
Long cycle stability |
— |
— |
— |
358 |
MoS2–RuO2 |
719 |
100 |
35.92 |
0.6 |
— |
359 |
CuO/MoS2 |
268 |
90.02 |
26.66 |
1599.6 |
— |
360 |
MoS2/CeO2 |
166.6 |
— |
— |
— |
— |
361 |
MnO2 nanowire/MoS2 |
212 |
84.1 |
29.5 |
1316 |
— |
362 |
nanofibers/TiO2@MoS2 |
510.4 |
95.7 |
— |
— |
— |
363 |
ZnS/MoS2/NiF |
3540 |
97 |
122 |
2500 |
100 |
364 |
MoS2/NiF |
1666 |
79 |
72 |
250 |
85 |
364 |
MoS2–PbS |
205.50 |
— |
6.95 |
24.82 |
— |
365 |
Core–shell TiNb2O7@MoS2/C |
— |
93.7 |
147.2 |
2470.5 |
— |
366 |
MnCo2O4@MoS2 |
512 |
91.87 |
36 |
19 |
99.57 |
367 |
For instance, NiCoP@MoS2 delivered far superior capacitance performance than NiCoP. Researchers said the layer structure with MoS2 was the primary contributing factor. Such a structure would promote active sites and electron transport to achieve the best electrochemical performance. Likewise, capacitance in ZnS/MoS2/NF was enhanced since MoS2 and system stability provided better energy storage and were also boosted.
The layer structure of MoS2 forms a conductive network for efficient electron transfer throughout all the charge/discharge cycles without any structural changes. This thereby positions MoS2 in the linear way of obtaining high-performance supercapacitors, opening up a new opportunity to advance energy storage technologies.
9.3 Comparison with metal phosphides/oxides/sulfides in batteries
This compares electrochemical performance among TMP@MoS2 heterostructures with metal phosphides, oxides, and sulfides as electrode materials that differ significantly in discharge capacity, cycling performance, and rate capability. The first discharge capacity of MoP–MoS2 heterostructure arrays was 1090.02 mA h g−1 at a mass loading of 3.9 mg cm−3, with a corresponding efficiency that remained above 98% for no less than 1000 cycles—a much better result relative to the stand-alone MoP and MoS2. They show much lower discharge capacities of 495 mA h g−1 and 436 mA h g−1, respectively. This hybridization of MoS2 with other metal compounds adds CoP–C@MoS2/C and FeP@SnP@MoS2 for enhanced discharge capacities of 1000 mA h g−1 and 905.3 mA h g−1, respectively, showing the synergy between MoS2 with metal phosphides and improved cycling stability. In comparison, other heterostructures, such as MoS2–MoO and MoS2–SnS, have even higher discharge capacities of 1531 mA h g−1 and 1504.6 mA h g−1, respectively, further explaining the role of composite materials in enhancing battery performance. These materials have a high-rate capability and excellent cycling performance, ladling out the potential of TMP@MoS2 heterostructures for promising candidates of next-generation energy storage devices, balanced against traditional metal phosphides, oxides, and sulfides (Table 6).
Table 6 Comparing metal phosphide/MoS2 heterostructures with metal phosphides/oxides/sulfides electrode materials in batteries
Metal phosphides |
Discharge capacity (mA h g−1) |
Mass loading (mg cm−2) |
Efficiency (%) |
Cycling performance (mA h g−1) |
Rate capability (mA h g−1/A. g−1) |
Ref. |
MoP–MoS2 |
1090.02 |
3.9 |
98 |
1000 |
— |
12 |
MoP |
495 |
1.0 |
— |
— |
— |
368 |
MoS2 |
436 |
— |
— |
187.48 |
— |
369 |
CoP–C@MoS2/C |
1000 |
— |
99.4 |
500 |
— |
370 |
CoP |
1866.9 |
— |
80.3 |
635.3 |
656.81 |
371 |
Fe2P@SnP0.94@MoS2 |
905.3 |
— |
99.4 |
797.5 |
797.5 |
372 |
Fe2P |
413 |
— |
— |
— |
396 |
373 |
Fe7S8–MoS2 |
1250.5 |
6 |
— |
1000.2 |
674.4 |
374 |
MoS2–MoO3 |
1531 |
5.9 |
92 |
640 |
— |
375 |
CoS/NC@MoS2 |
1417.2 |
0.566 |
74.82 |
1256.6 |
100 |
376 |
NiCo2S4@MoS2 |
398 |
0.65 |
>99 |
400 |
398 |
377 |
MnS–MoS2 |
1246.2 |
— |
59.2 |
397.2 |
— |
378 |
VS4/MoS2 |
1061.4 |
4.6 |
90 |
808.3 |
665.0 |
379 |
MoS2–SnS |
1504.6 |
3.50 |
— |
1083.312 |
690.1 |
380 |
Sb2S3/MoS2 |
701 |
— |
95 |
561 |
— |
381 |
MoS2/FeS2/C |
613.1 |
1.1–1.5 |
100 |
77.2 |
574.6 |
382 |
G/NiS2–MoS2 |
509.6 |
1.2–1.6 |
>99 |
383.8 |
— |
383 |
MoS2/ZnS-NC |
1427.2 |
2.4 |
93.7 |
760.15 |
— |
384 |
MoS2/WS2 |
620 |
— |
64 |
333 |
487 |
385 |
(1T-2H MoS2)/CoS2 |
729.6 |
— |
99 |
300 |
400 |
386 |
9.4 Comparative performance of TMP@MoS2 across energy storage and catalysis
By directly comparing the relative performance of TMP@MoS2 heterostructures in batteries, supercapacitors, and electrocatalysts based on Tables 2–4, it will be evident that every application has its peculiar benefits from these materials.387 In batteries, TMP@MoS2 heterostructures such as MoP–MoS2 and Fe7S8–MoS2 have very high discharge capacities while maintaining excellent cycling stability. For instance, MoP–MoS2 at the end of 1000 cycles delivers a discharge capacity of 1090.02 mA h g−1 with an efficiency of 98%, which for Fe7S8–MoS2 is further increased to 1250.5 mA h g−1. These make the TMP@MoS2 heterostructures uniquely suitable for long-term energy storage, wherein performance during thousands of charge–discharge cycles becomes significant. On the other hand, this TMP@MoS2 heterostructure has significantly helped supercapacitors' capacitance and energy density. NiCoP@MoS2 and ZnS/MoS2/NiF are outstanding materials used in applications; between them, the capacitance of NiCoP@MoS2 was somewhat exceptional in terms of its 2352.407 F g−1 at an energy density of 52 W h kg−1. One beyond is the ZnS/MoS2/NiF-type capacitor, which attains a capacitance value of 3540 F g−1 but tells a far greater cycling stability of 97 percent. These features further make TMP@MoS2 heterostructures ideal for applications requiring high power density and rapid energy discharge. They are better candidates for developing supercapacitors for applications with fast charging and discharging cycles. For example, MoP/MoS2 and CoS–MoS2 heterostructures in TMP@MoS2 substrates showed good catalytic performance, especially hydrogen evolution reactions. In a separate study, MoS2/MoS2 heterostructures demonstrated an overpotential of only 96 mV for HER, suggesting a highly catalytically efficient material. CoS–MoS2 further reduced it to 60 mV, thus being highly stable and efficient as an electrocatalyst. With such properties, the TMP@MoS2 heterostructures will work out the most in low-overpotential, strongly stable energy–conversion processes. Supercapacitors have demonstrated the most dramatic rise in performance among all investigated devices. However, which application can exploit the advantages of TMP@MoS2 heterostructures “best” depends on specific requirements. Provided that long-term energy storage is required, batteries equipped with electrodes configured from TMP@MoS2 heterostructures offer both large discharge capacity and stability of operation. In the case when fast energy delivery is required, the performance of supercapacitors is unmatched. On the other hand, electrocatalysts both derive an enhanced catalytic efficiency and durability related to energy conversion. Though all three applications have some compelling reasons, TMP@MoS2 heterostructures have the most significant advantages when applied for supercapacitors.
This comparative analysis is represented by a figure that shows the different electrochemical performances of TMP@MoS2 heterostructures in various fields, including electrocatalysis, supercapacitors, and batteries. Specifically, in electrocatalysis, this is manifested as the FeP/MoS2 composite standing superior to pure MoS2 and FeP in both current density and Tafel slope for HER, as shown in Fig. 12a–a′′. Specifically, the electrochemical impedance spectroscopy result from Fig. 12a′ shows that FeP/MoS2 has the lowest charge transfer resistance compared to other counterparts. This would indicate that electron transfer was more efficient during catalytic processes. The TOF curves in Fig. 12a′′ confirm that the intrinsic catalytic activity of FeP/MoS2 is the highest among all those materials tested. This NiCoP@MoS2 heterostructure performed very well for supercapacitors concerning specific capacitance and cycling stability, as shown in Fig. 12a–a′′ to b–b′′. Obviously, from Fig. 12b, the NiCoP@MoS2 has much higher specific current and electrochemical reversibility than pure MoS2, according to the CV curves. From the EIS plot in Fig. 12b′, these composite exhibits much lower resistance for NiCoP@MoS2; that is, more rapid charge/discharge cycles. Fig. 12b′′ illustrates the excellent cycling stability for NiCoP@MoS2. It was found to retain, even at a large number of cycles, a pretty large capacitance compared with that for MoS2 or the pristine NiCoP material. Regarding batteries, the CoP@MoS2 composite represented in Fig. 12c to c′′ delivers a distinguished large discharge capacity and long-term cycling stability. Concretely, as shown in Fig. 12c, differential capacity versus scan rate curves are presented; here, CoP@MoS2 has higher capacity retention and better rate capability than MoS2 and pure CoP. This is further supported by the Nyquist plot in Fig. 12c′, where CoP@MoS2 has lower resistance, thus contributing to its better electrochemical performance. Fig. 12c′′ outlines the outstanding cycling stability and coulombic efficiency of CoP@MoS2, during which a high discharge capacity can be maintained for more than 1000 cycles, far superior to that of MoS2 and CoP alone. The components of TMP@MoS2 2D heterostructures have distinctly improved electrochemical performance, specifically FeP/MoS2 for electrocatalysis, NiCoP@MoS2 for supercapacitors, and CoP@MoS2 for batteries. Each composite material has characteristics superior to the others in its way. Still, it tailors to the requirements or needs that an application would need, making TMP@MoS2 heterostructures quite versatile and very practical in energy storage and conversion.
 |
| Fig. 12 Comparative analysis of the electrochemical performance of various TMP@MoS2 (a–a′′, b–b′′ and c–c′′) electrocatalyst. Reproduced from ref. 232 with permission from [American Chemical Society]. | |
Indeed, TMP@MoS2 heterostructures have shown unparalleled versatility and performance in various applications, including electrocatalysis, supercapacitors, and batteries.12,20,388 Their high catalytic activity in electrocatalysis reflects reduced overpotentials for HER and OER reactions.389 Moreover, superior specific capacitances and energy density can be achieved in supercapacitors,20 and batteries will have high discharge capacities and long-term stability during cycling.332 This ensures the prospect of promising candidacy in next-generation energy technology. Nevertheless, the following limitations and challenges somewhat offset these strengths and require further research and optimization.
10. Challenges and future perspectives in TMP@MoS2 research
10.1 Limitations in electrocatalysis
CoP@MoS2 heterostructure is susceptible to instability in equilibrium between MoS2 and CoP phases unless it is optimally engineered. If external MoS2 layers are too thin, there is a lack of sufficient CoP to activate surface inertness and reduce charge and mass conductivity.15 If there is excessive CoP, layered architecture is destabilized. While multiple studies have explored CoP/MoS22 heterojunctions as hydrogen evolution catalysts, a scalable and high-performance synthesis approach remains an open research question. Consequently, much work is still needed to develop efficient, cost-effective, and highly durable CoP@MoS2 catalysts for HER.390 During hydrogen evolution, gas bubbles can accumulate on the heterostructure's surface, blocking active sites and lowering efficiency.391 Even though heterojunction catalysts promise superior catalytic properties compared to monocomponent systems, they often suffer from low stability, weak adhesion to electrodes, and aggregation of active chemicals, which limit their exposure to catalytic sites. The incorporation of strong interfacial bonding, defect engineering, and encapsulation in conductive matrices can improve catalyst durability and efficiency.392
10.2 Limitations in supercapacitors
Transition metal phosphides can also show lower inherent electrical conductivity than materials based on carbon, such as graphene and carbon nanotubes, reducing charge mobility and lower power density in supercapacitors.393 Electrolyte selection is also crucial as some TMPs can degrade with some electrolytes and cause a drop in efficiency. Surface modifications and protective coatings can reduce this by optimizing electrode/electrolyte compatibility.394 Supercapacitor energy density is directly proportional to capacitance and operating voltage range and is still lower for many existing supercapacitors compared to batteries even with high-power-density active electrodes. Hence, merely combining TMPs and MoS2 would not increase energy density significantly unless heterostructure engineering is finely optimized to enhance charge storage mechanisms.274 Further, TMP-MoS2 heterostructure synthesis is still complex and requires stringent conditions and higher temperature and is difficult to industrially scale. Future work would be to design simpler and scalable synthesis methods such as rapid chemical vapor deposition or template-assisted hydrothermal methods.314
10.3 Limitations in batteries
The optimal electrochemical performance in MoS2/MoP Mott–Schottky heterostructures in lithium–sulfur batteries is achieved when MoS2 and MoP are in the proper ratio.332 The excess phosphorus in phosphides (x
:
y < 1) is detrimental to electrical conductivity and reduces them to be less efficient in electronics and batteries. An increased amount of phosphorus also reduces structural stability and leads them to decompose on thermal treatment. This trade-off points towards the need for controlling composition with precision in TMP@MoS2 heterostructures. With a reduced Ni
:
P ratio, metal–metal interactions are minimized and with them, conductivity and catalytic efficiency too diminish. Doping techniques and hybrid structures with conductive additives would overcome these limitations.232
TMP@MoS2 heterostructures are of great potential as a material class based on structural stability, electrical conductance, and large specific capacity.20 They will be promising candidates for improving critical energy storage and electrocatalysis.395 These heterostructures can significantly enhance catalytic activity by leveraging synergistic properties between TMPs and MoS2.392
While TMPs have demonstrated intrinsic catalytic performance in HER and OER, their long-term durability in real-world applications remains insufficient.396 Introducing MoS2 into TMP-based catalysts reduces overpotential, improves electron transport, and exposes additional active sites;15 for instance, CoP@MoS2 heterostructures exhibit a reduced overpotential of 60 mV in HER, attributed to electronic modulation induced by MoS2.
TMP@MoS2 heterostructures offer high performance in energy storage applications at both high-power densities and long-term cycling conditions. Incorporating MoS2 into TMP-based materials in batteries enhances cycling stability and discharge capacity compared to their individual components.370,397 MoP@MoS2,12 for example, delivers a discharge capacity of 1090.02 mA h g−1 with 98% efficiency after 1000 cycles, making it one of the most durable electrode materials for advanced energy storage systems. This improvement arises from the rapid ion transport facilitated by the layered MoS2 structure, enabling enhanced charge–discharge cycling. Similarly, NiCoP@MoS2 has demonstrated a specific capacitance of 2352.4 F g−1 and an energy density of ∼52 W h kg−1, outperforming conventional electrode materials and positioning it as a top candidate for high-performance energy storage applications.20
Beyond their outstanding performance, TMP@MoS2 heterostructures offer industrial scalability and cost-efficiency solutions. Hydrothermal synthesis15 and chemical vapor deposition398 provide scalable methods for producing high-purity TMP@MoS2 heterostructures, allowing precise control over particle morphology and size. These methods enable cost-effective large-scale production, making TMP@MoS2 materials feasible for commercial applications. However, challenges remain.399 Synthesis methodologies must be further optimized for large-scale production without sacrificing material quality.400 The stability and efficiency of TMP@MoS2 for HER and OER can be further enhanced by fine-tuning reaction conditions, improving interfacial engineering, and developing innovative composite structures. Further synthesis and integration techniques refinement is necessary to fully realize TMP@MoS2's potential in practical energy applications.
Overcoming these research and industry challenges will allow TMP@MoS2 heterostructures to revolutionize energy storage and electrocatalysis. Their development aligns with global sustainability goals, contributing to efficient and environmentally friendly energy solutions. By integrating green synthesis techniques, optimizing charge transport, and improving catalytic durability, TMP@MoS2 heterostructures could become the key materials for next-generation energy technologies.
10.4 Effect of size, stability, and other compositions
The structural stability, composition, and size of transition metal phosphide/molybdenum disulfide (TMP@MoS2) heterostructures are responsible for regulating their electrochemical energy storage behavior.401–404 Nanoscale reduction promotes electrochemically active sites, ion diffusion, and charge transfer, leading to structural instability and agglomeration; therefore, there is a requirement for optimal balance.405–407
Stability has been enhanced using such methods as heteroatomic doping,408 hybridization with a conductive matrix,409 and coating with carbon.410
Compositions are crucial in controlling catalytic activity, charge storage, and electronic properties.411,412 Different transition metals (Ni, Co, Fe, Mn) offer different electronic structures and redox behaviors.413 Bimetallic/ternary phosphides (such as Ni–Co–P and Fe–Co–P) offer synergistic effects with improved conductivity and redox capability.414 Integrating materials such as graphene, MXenes, or MOFs to further improve electrode properties and multi-functionalities.415
Modulating these structural factors allows the rational design of heterostructures that exhibit higher energy density as well as improved cycling stability and kinetic behavior for charge storage, and TMP@MoS2 represents a promising material for next-generation batteries and supercapacitors.
10.5 Strategies for overcoming challenges in TMP@MoS2 heterostructures for energy storage applications
The progress in the development of the TMP@MoS2 heterostructures for energy storage devices has been hindered by several issues, such as low interface stability,416 low efficiency in the transfer of charges,94 and scalability problems417 in the fabrication process. To encounter these bottlenecks, interface engineering,418 atomic-scale design,419 doping,408,420 surface modification,421 and multifunctional design methodologies422 are being addressed from the perspective of science.
10.5.1 Interface engineering and atomic-scale design. The interfacial contact between TMPs and MoS2 is crucial for the activity of TMP@MoS2 heterostructures.423 Low interfacial contact leads to resistance and poor transportation of charges.424 Atomic-level processing techniques such as epitaxial growth and heterojunction optimization are required to increase interfacial bonding and minimize resistance.425 Strong chemical bonding and good van der Waals interactions in the interface were found to increase electron mobility and electrochemistry by a large margin.426
10.5.2 Doping and surface modification approaches. Doping has been extensively employed to modify MoS2 electronic band structure and increase conductivity.427 Nitrogen, boron, and phosphorus heteroatom doping have been found to enhance electrochemical activity by offering active sites for kinetic reactions and storage.428 Further, the low electrical conductivity and structural instability issues with MoS2 have also been addressed using surface modification techniques,429 which include functionalization with carbonaceous species or conducting polymers (e.g., CNT, graphene).430
10.5.3 Advanced fabrication and scaling techniques. Scalability remains a challenge in the preparative synthesis of TMP@MoS2 heterostructures.431 Classical wet chemical synthesis methods, although efficient, are plagued by inhomogeneous morphology and variability across different batches.432 CVD (chemical vapor deposition), atomic layer deposition (ALD), and laser-based methods are sophisticated methods that have been explored to offer control of thickness with accuracy and homogeneous material dispersion on a large scale.433–435 High surface area and better ion access and stability can be achieved through self-assembly processes to form heterostructures.436
10.5.4 Design of multifunctional TMP@MoS2 heterostructures. To further improve the electrochemical practicality of TMP@MoS2 nanostructures, researchers have tailored multi-functional heterostructures with synergistic attributes that are proven to cater to the performance of supercapacitors and batteries.437 For example, it has been reported that hierarchical pore structures with optimized ion channel size for diffusion significantly affect the electrochemical activity.438 Energy storage and catalytic properties of hybrid structures are also investigated as future-generation energy storage materials.439
10.6 Recent advances in TMP@MoS2-based electrochemical energy storage
Recent advancements in the area of TMP@MoS2-based electrochemical energy storage reveal the potential that these heterostructures have for batteries and supercapacitors. Studies have established that the coupling between the transition metal phosphides (TMPs) and the MoS2 enhances the storage of charges through the increase in the number of active sites as well as electron transfer. Particularly, self-assembly Cu–Mo sulfide and phosphide compounds possess high capacitance and stability with up to 86.9% retention of their initial capacity when cycled to as many as 4000 cycles. Ternary TMPs and sulfides also possess higher conductivity, facilitating multi-electron redox reactions and hence enhancing the energy density when utilized in the case of supercapacitors.
MoS2-based heterostructures also hold promise in hydrogen evolution reactions (HER). A hierarchical CoP/Ni2P/MoS2 catalyst exhibits outstanding catalytic activity in the presence of neutral, alkaline, as well as acidic media, with the advantage of optimized charge transfer and structural design. In addition, Mo-based compounds, such as phosphides, oxides, and carbides, are instrumental in the development of lithium–sulfur (Li–S) batteries through conductivity, stabilization of the polysulfides, and enhanced electrochemical properties. Despite the progress, more work has to be done to make these materials suitable for commercial applications on a large scale.13,440,441 Metal phosphides are shown to be promising for enhancing the performance of lithium–sulfur batteries (LSBs) by addressing significant issues, including the lithium polysulfide (LiPS) shuttle effect, sluggish sulfur conversion, and growth of lithium dendrites. Their high catalytic activity and tunable cationic character make them active in LiPS adsorption and rapid conversion, promoting sulfur utilization and battery stability. Additionally, their addition to composite materials enhances catalytic efficiency, whereas their application in separator modification and the coating of the lithium anode further increases battery efficiency. Despite the continued issues with conductivity and uniform deposition of lithium, continued developments in metal phosphide design are contributing towards the commercialization of high-performance LSBs.442
Artificial Intelligence (AI) is now a potent tool to expedite the discovery,443 optimization, and design of heterostructure materials for energy storage444 by circumventing time and resource-consuming traditional computational and experimental methods.445–447 AI-enabled methods like machine learning (ML), deep learning (DL), and high-throughput computational modeling enable efficient exploration of large chemical and structural spaces.448–450 AI methods screen large databases to provide predictions for optimal compositions,451 crystal structures,452 and interfacial interactions in TMP@MoS2 heterostructures453 and shorten material discovery time by orders of magnitude.454 AI-powered molecular dynamics (MD) simulations and neural network-based interatomic potentials optimize interface stability455,456 and defect engineering455 and enhance catalytic activity and charge transfer behavior.457 AI-enabled robotic synthesis platforms and high-throughput experiments speed up heterostructure development by identifying optimal synthesis conditions for scalable manufacturing.458,459 AI models based on electrochemical test data also predict long-term stability,460 energy density,461 and rate capability in batteries and supercapacitors462 and optimize electrode architectures and ion diffusion pathway simulations to maximize device efficiency.463,464
11. Conclusion
This research highlights the promise of TMP@MoS2 heterostructures to function as next-generation materials in electrocatalysis and energy storage in electrochemistry. The high conductance and catalytic activity of transition metal phosphides and catalytic efficiency and structure stability of molybdenum disulfide make such heterostructures promising next-generation batteries, supercapacitors, and electrocatalytic applications like HER and OER. The synergy between TMP and MoS2 enhances charge transfer, stability, and exposure of active sites and addresses some of the primary issues with present energy storage technology. Despite such advantages, challenges still persist with regard to optimizing the scalability of synthesis, long-term stability, and application viability at a large scale. Future research will be needed to enhance synthesis techniques, interfacial engineering, and compositional tuning with new methodologies to boost performance. Further research in such areas will be critical to unlocking the full capability of TMP@MoS2 in applications involving sustainable energy.
Data availability
No primary research results, software or code have been included and no new data were generated or analysed as part of this review.
Conflicts of interest
There are no conflicts to declare.
Acknowledgements
This work was supported by the Iranian Research & Development Center for Chemical Industries (IRDCI), and Pharmaceutical Analysis Research Center, Tabriz University of Medical Sciences, Tabriz, Iran [76636].
References
- J. L. Holechek, H. M. Geli, M. N. Sawalhah and R. Valdez, Sustainability, 2022, 14, 4792 CrossRef.
- E. Jamshidi, S. Dalvand, F. Manteghi and S. M. Mousavi Khoshdel, iScience, 2025, 111672 CrossRef CAS PubMed.
- S. A. Mousavianfard, A. Molaei, M. Manouchehri, A. Foroozandeh, A. Shahmohammadi and S. Dalvand, J. Energy Storage, 2025, 109, 115232 CrossRef.
- A. D. A. Bin Abu Sofian, H. R. Lim, H. Siti Halimatul Munawaroh, Z. Ma, K. W. Chew and P. L. Show, Sustainable Dev., 2024, 32, 3953 CrossRef.
- R. T. Yadlapalli, R. R. Alla, R. Kandipati and A. Kotapati, J. Energy Storage, 2022, 49, 104194 CrossRef.
- Z. Zhu, T. Jiang, M. Ali, Y. Meng, Y. Jin, Y. Cui and W. Chen, Chem. Rev., 2022, 122, 16610 CrossRef CAS PubMed.
- T. S. Babu, K. R. Vasudevan, V. K. Ramachandaramurthy, S. B. Sani, S. Chemud and R. M. Lajim, IEEE Access, 2020, 8, 148702 Search PubMed.
- S. Dalvand, A. Foroozandeh, A. Heydarian, F. S. Nasab, M. Omidvar, N. Yazdanfar and A. Asghari, Ionics, 2024, 30, 1857 CrossRef CAS.
- Y. Li, J. Zhang, Q. Chen, X. Xia and M. Chen, Adv. Mater., 2021, 33, 2100855 CrossRef CAS PubMed.
- S. M. Qashqay, J. Rahimi, M.-R. Zamani-Meymian and A. Maleki, J. Energy Storage, 2023, 72, 108548 CrossRef.
- R. Eivazzadeh-Keihan, R. Taheri-Ledari, N. Khosropour, S. Dalvand, A. Maleki, S. M. Mousavi-Khoshdel and H. Sohrabi, Colloids Surf., A, 2020, 587, 124335 CrossRef CAS.
- X. Wang, N. Deng, J. Ju, G. Wang, L. Wei, H. Gao, B. Cheng and W. Kang, J. Membr. Sci., 2022, 642, 120003 CrossRef CAS.
- X. Cheng and Y. Tong, ACS Appl. Energy Mater., 2023, 6, 9577 CrossRef CAS.
- M. Afshari Babazad, A. Foroozandeh, M. Abdouss, H. SalarAmoli, R. A. Babazad and M. Hasanzadeh, TrAC, Trends Anal. Chem., 2024, 180, 117964 CrossRef CAS.
- T. Xia, L. Zhou, S. Gu, H. Gao, X. Ren, S. Li, R. Wang and H. Guo, Mater. Des., 2021, 211, 110165 CrossRef CAS.
- X. Wang, G. Zhang, B. Wang, Y. Wu and S. Guo, ACS Sustain. Chem. Eng., 2024, 12, 14018 CrossRef CAS.
- A. Foroozandeh, M. Abdouss, H. SalarAmoli, M. Pourmadadi and F. Yazdian, Process Biochem., 2023, 127, 82 CrossRef CAS.
- J. Zhang, J. Ma, R. Cui, W. Ling, M. Hong and R. Sun, Chem. Eng. J., 2025, 503, 158427 CrossRef CAS.
- A. Foroozandeh, H. SalarAmoli, M. Abdouss and M. Pourmadadi, Sens. Actuators Rep., 2024, 7, 100195 CrossRef.
- A. Shahmohammadi, S. Dalvand and H. Baheri, Mater. Chem. Phys., 2024, 317, 129150 CrossRef CAS.
- F. Khoramjah, M. Omidvar, M. S. Miresmaieli, S. Dalvand, A. Asghari, M. Kambarani and N. Mohammadi, Diamond Relat. Mater., 2023, 132, 109590 CrossRef CAS.
- S. S. Mirzaei, M. Pourmadadi, A. Foroozandeh, A. A. Moghaddam, M. Soltani, N. Basirhaghighi and M. Ahmadi, J. Appl. Electrochem., 2024, 54, 1887–1900 CrossRef CAS.
- X. Peng, Y. Lv and S. Zhao, Coatings, 2022, 12, 68 CrossRef CAS.
- S. Ghorai and A. Govind Rajan, Chem. Mater., 2024, 36, 2698 CrossRef CAS.
- W. Fu, M. John, T. D. Maddumapatabandi, F. Bussolotti, Y. S. Yau, M. Lin and K. E. Johnson Goh, ACS Nano, 2023, 17, 16348 CrossRef CAS PubMed.
- M. Omidvar, S. Dalvand, A. Asghari, N. Yazdanfar, H. Y. Sadat and N. Mohammadi, Fuel, 2023, 347, 128472 CrossRef CAS.
- Y. Zhu, T.-R. Kuo, Y.-H. Li, M.-Y. Qi, G. Chen, J. Wang, Y.-J. Xu and H. M. Chen, Energy Environ. Sci., 2021, 14, 1928 RSC.
- R. Eivazzadeh-Keihan, R. Taheri-Ledari, M. S. Mehrabad, S. Dalvand, H. Sohrabi, A. Maleki, S. M. Mousavi-Khoshdel and A. E. Shalan, Energy Fuels, 2021, 35, 10869 CrossRef CAS.
- Z. Sadat, R. Eivazzadeh-Keihan, V. Daneshvari-Esfahlan, S. Dalvand, A. Kashtiaray and A. Maleki, Sci. Rep., 2024, 14, 3137 CrossRef CAS PubMed.
- J. Dai, Y. Lv, J. Zhang, D. Zhang, H. Xie, C. Guo, A. Zhu, Y. Xu, M. Fan and C. Yuan, J. Colloid Interface Sci., 2021, 590, 591 CrossRef CAS PubMed.
- A. Asghari, S. Dalvand, M. sadat Miresmaeili, F. Khoramjah, M. Omidvar, M. Kambarani and N. Mohammadi, Int. J. Hydrogen Energy, 2023, 48, 9776 CrossRef CAS.
- M. Yousaf, U. Naseer, Y. Li, Z. Ali, N. Mahmood, L. Wang, P. Gao and S. Guo, Energy Environ. Sci., 2021, 14, 2670 RSC.
- P. Zhang, Y. Zhao, Y. Li, N. Li, S. R. P. Silva, G. Shao and P. Zhang, Advanced Science, 2023, 10, 2206786 CrossRef CAS PubMed.
- S. Dalvand, M. Omidvar, A. Asghari, N. Mohammadi and N. Yazdanfar, J. Porous Mater., 2023, 30, 2069 CrossRef CAS.
- M. J. Kim, I. H. Choi, S. C. Jo, B. G. Kim, Y. C. Ha, S. M. Lee, S. Kang, K. J. Baeg and J. W. Park, Small Methods, 2021, 5, 2100793 CrossRef CAS PubMed.
- S. Dalvand, Z. Khoushab, S. M. Mousavi-Khoshdel, H. Ghafuri, H. R. Esmaili Zand and M. Omidvar, Tungstate-Modified Ionic Liquid Functionalized Magnetic Graphene Oxide: Synthesis and Application as a High-Performance Supercapacitor, SSRN, 2021, preprint, DOI:10.2139/ssrn.3968871.
- J. Joy, A. Krishnamoorthy, A. Tanna, V. Kamathe, R. Nagar and S. Srinivasan, Appl. Sci., 2022, 12, 9312 CrossRef CAS.
- C. C. Piras, S. Fernández-Prieto and W. M. De Borggraeve, Nanoscale Adv., 2019, 1, 937 RSC.
- L. K. Wei, S. Z. Abd Rahim, M. M. Al Bakri Abdullah, A. T. M. Yin, M. F. Ghazali, M. F. Omar, O. Nemes, A. V. Sandu, P. Vizureanu and A. E.-h. Abdellah, Materials, 2023, 16, 4635 CrossRef CAS PubMed.
- W. Li, Y. Li, J.-H. Wang, S. Huang, A. Chen, L. Yang, J. Chen, L. He, W. K. Pang and L. Thomsen, Energy Environ. Sci., 2024, 17(15), 5387–5398 RSC.
- S. A. Getaneh, A. G. Temam, A. C. Nwanya, P. M. Ejikeme and F. I. Ezema, Mater. Sci. Technol., 2024, 40, 185 CrossRef CAS.
- S. N. Alam, G. Arka, S. Nityananda, S. Pankaj, S. Kakara and A. M. Shafdar, Consolidation of Mechanically Alloyed Powders, Elsevier., 2024, pp. 119 Search PubMed.
- S. Arya, A. Singh, A. Ahmed, B. Padha, A. Banotra, U. Parihar, A. K. Sundramoorthy, S. Dixit and N. I. Vatin, J. Energy Chem., 2025, 122(23), 17155–17239 Search PubMed.
- H. Patil, S. K. Vemula, S. Narala, P. Lakkala, S. R. Munnangi, N. Narala, M. O. Jara, R. O. Williams III, H. Terefe and M. A. Repka, AAPS PharmSciTech, 2024, 25, 37 CrossRef PubMed.
- A. Singh, S. S. Shah, A. Dubey, A. Ahmed, S. V. Ranganayakulu, A. K. Sundramoorthy and S. Arya, J. Energy Storage, 2025, 109, 115183 CrossRef.
- D. Chen, B. Liu, G. Sun, W. Xu, Y. Zhu, Y. An, L. Zhu, X. Ding, J. Zhang and X. Lu, Adv. Powder Technol., 2024, 35, 104377 CrossRef CAS.
- O. S. ODEBIYI, G. Yuning, D. Hao, L. Biao and W. Shaona, Mater. Chem. Phys., 2024, 129697 CrossRef CAS.
- P. Gao, X. Fan, D. Sun, G. Zeng, Q. Wang and Q. Wang, Water, 2024, 16, 1639 CrossRef CAS.
- A. Ahmed, Y.-L. Chu, S.-J. Young, M. P. Chavhan, A. K. Sundramoorthy and S. Arya, Ionics, 2025, 1 Search PubMed.
- S. Verma, B. Padha, A. Ahmed, R. Singh, D. P. Dubal and S. Arya, Prog. Energy, 2024, 6, 042002 CrossRef CAS.
- Z. Kong, Z. Wang, B. Chen, Y. Li and R. Li, Materials, 2023, 16, 5763 CrossRef CAS PubMed.
- N. A. Mala, A. Singh, S. Bashir, A. Ghosh, A. Padder, S. Arya, N. Thakur, L. Guganathan and R. N. Ali, Inorg. Chem. Commun., 2025, 174, 113969 CrossRef CAS.
- H. Ismail, H. Mohamad and R. Hussin, Emerging Advances in Integrated Technology, Universiti Tun Hussein Onn Malaysia Publisher’s Office, 2022, 3, 9–14 Search PubMed.
- C. Real and F. J. Gotor, Heliyon, 2019, 5, e01227 CrossRef CAS PubMed.
- F. J. Gotor, M. Achimovicova, C. Real and P. Balaz, Powder Technol., 2013, 233, 1 CrossRef CAS.
- Z. Wu, Y. Liang, E. Fu, J. Du, P. Wang, Y. Fan and Y. Zhao, Metals, 2018, 8, 281 CrossRef.
- A. Foroozandeh, M. A. Babazad, S. Jouybar, M. Abdouss, H. Salar Amoli, K. Dashtian and M. Hasanzadeh, TrAC, Trends Anal. Chem., 2025, 183, 118119 CrossRef CAS.
- L. L. Driscoll, E. H. Driscoll, B. Dong, F. N. Sayed, J. N. Wilson, C. A. O'Keefe, D. J. Gardner, C. P. Grey, P. K. Allan, A. A. L. Michalchuk and P. R. Slater, Energy Environ. Sci., 2023, 16, 5196 RSC.
- L. M. Martínez, J. Cruz-Angeles, M. Vázquez-Dávila, E. Martínez, P. Cabada, C. Navarrete-Bernal and F. Cortez, Pharmaceutics, 2022, 14, 2003 CrossRef PubMed.
- S. Azzaza, S. Alleg and J. J. Suñol, Adv. Mater. Phys. Chem., 2013, 03, 90 CrossRef.
- P. Pattanayak, S. Saha, T. Chatterjee and B. C. Ranu, Chem. Commun., 2024, 61, 247 RSC.
- F. Shi and W. Xie, Miner. Eng., 2016, 86, 66 CrossRef CAS.
- S. Dalvand, Z. Khoushab, S. M. Mousavi-Khoshdel, H. Ghafuri, H. R. E. Zand and M. Omidvar, Int. J. Hydrogen Energy, 2023, 48, 10098 CrossRef CAS.
- S. Reichle and M. Felderhoff, Mechanochemistry and Emerging Technologies for Sustainable Chemical Manufacturing, 2023, p. 151 Search PubMed.
- J. F. Reynes, V. Isoni and F. García, Angew. Chem., Int. Ed., 2023, 62, e202300819 CrossRef CAS PubMed.
- H. Chen, Q. Cao, Z. Ye, B. Lai, Y. Zhang, H. Dong, D. E. Crawford, O. M. Istrate and S. L. James, Adv. Mater. Technol., 2024, 9, 2301780 CrossRef CAS.
- M. Dhaval, S. Sharma, K. Dudhat and J. Chavda, J. Pharmaceut. Innovat., 2022, 17, 294 CrossRef.
- J. N. Tiwari, K. Kumar, M. Safarkhani, M. Umer, A. T. E. Vilian, A. Beloqui, G. Bhaskaran, Y. S. Huh and Y.-K. Han, Advanced Science, 2024, 11, 2403197 CrossRef CAS PubMed.
- R. Breitwieser, U. Acevedo Salas, S. Merah and R. Valenzuela, Ferrite Nanostructures Consolidated by Spark Plasma Sintering (SPS), 2017 Search PubMed.
- A. Saberi, M. Kouhjani, D. Yari, A. Jahani, K. Asare-Addo, H. Kamali and A. Nokhodchi, J. Drug Delivery Sci. Technol., 2023, 86, 104746 CrossRef CAS.
- U. Nandi, V. Trivedi, S. A. Ross and D. Douroumis, Pharmaceutics, 2021, 13, 624 CrossRef CAS PubMed.
- P. Evon, V. Vandenbossche, L. Candy, P.-Y. Pontalier and A. Rouilly, Twin-screw Extrusion: A Key Technology for the Biorefinery, ACS Publications, 2018, pp. 25 Search PubMed.
- L. Chai, S. Liu, S. Pei and C. Wang, Chem. Eng. J., 2021, 420, 129686 CrossRef CAS.
- M. B. Lopez and J. Ustarroz, Curr. Opin. Electrochem., 2021, 27, 100688 CrossRef.
- C. C. Weng, J. T. Ren and Z. Y. Yuan, ChemSusChem, 2020, 13, 3357 CrossRef CAS PubMed.
- T. Sivaranjani, T. Revathy and A. Stephen, Controlled Electrochemical Deposition for Materials Synthesis, CRC Press. 2020, pp. 25 Search PubMed.
- S. Islam, M. M. Mia, S. S. Shah, S. Naher, M. N. Shaikh, M. A. Aziz and A. S. Ahammad, Chem. Rec., 2022, 22, e202200013 CrossRef CAS PubMed.
- U. Mohanty, B. Tripathy, P. Singh, A. Keshavarz and S. Iglauer, J. Appl. Electrochem., 2019, 49, 847 CrossRef CAS.
- S. A. Lee, J. W. Yang, S. Choi and H. W. Jang, Exploration, 2021, 1, 20210012 CrossRef PubMed.
- F. C. Walsh, S. Wang and N. Zhou, Curr. Opin. Electrochem., 2020, 20, 8 CrossRef CAS.
- R. H. Miller, S. Hu, S. J. Weamie, S. A. Naame and D. G. Kiazolu, J. Chem. Eng. Mater. Sci., 2021, 9, 68 CAS.
- J. Pu, Z. Shen, C. Zhong, Q. Zhou, J. Liu, J. Zhu and H. Zhang, Adv. Mater., 2020, 32, 1903808 CrossRef CAS PubMed.
- V. S. Saji, J. Ind. Eng. Chem., 2019, 75, 20 CrossRef CAS.
- S. Rajoria, M. Vashishtha and V. K. Sangal, Environ. Sci. Pollut. Res., 2022, 29, 72196 CrossRef CAS PubMed.
- J. Kim, H. Kim, G. H. Han, S. Hong, J. Park, J. Bang, S. Y. Kim and S. H. Ahn, Electrodeposition: an Efficient Method to Fabricate Self-supported Electrodes for Electrochemical Energy Conversion Systems, Wiley Online Library, 2022, p. 20210077 Search PubMed.
- F. Li, Y. Feng, Z. Li, C. Ma, J. Qu, X. Wu, D. Li, X. Zhang, T. Yang and Y. He, Adv. Mater., 2019, 31, 1901351 CrossRef PubMed.
- B. K. Chakrabarti, M. Gençten, G. Bree, A. H. Dao, D. Mandler and C. T. J. Low, Int. J. Energy Res., 2022, 46, 13205 CrossRef CAS.
- I. Brandt, C. Araujo, V. Stenger, R. Delatorre and A. Pasa, ECS Trans., 2008, 14, 413 CrossRef CAS.
- S. A. Kumar, S. Sahoo, G. K. Laxminarayana and C. S. Rout, Small, 2024, 20, 2402087 CrossRef CAS PubMed.
- S. Ji, J. Yang, J. Cao, X. Zhao, M. A. Mohammed, P. He, R. A. Dryfe and I. A. Kinloch, ACS Appl. Mater. Interfaces, 2020, 12, 13386 CrossRef CAS PubMed.
- J. Theerthagiri, A. P. Murthy, S. J. Lee, K. Karuppasamy, S. R. Arumugam, Y. Yu, M. M. Hanafiah, H.-S. Kim, V. Mittal and M. Y. Choi, Ceram. Int., 2021, 47, 4404 CrossRef CAS.
- S. Kang, C. Wang, J. Chen, T. Meng and J. E, J. Energy Storage, 2023, 67, 107515 CrossRef.
- R. I. Walton, Chem.–Eur. J., 2020, 26, 9041 CrossRef CAS PubMed.
- A. Ray, S. Sultana, L. Paramanik and K. Parida, J. Mater. Chem. A, 2020, 8, 19196 RSC.
- Z. Li, J. Yang, T. Guang, B. Fan, K. Zhu and X. Wang, Small Methods, 2021, 5, 2100193 CrossRef CAS PubMed.
- A. H. Mamaghani, F. Haghighat and C.-S. Lee, Chemosphere, 2019, 219, 804 CrossRef CAS PubMed.
- R. Tyagi, O. Ruzimuradov and J. Prakash, Mater. Chem. Phys., 2023, 307, 128108 CrossRef.
- L. Ndlwana, N. Raleie, K. M. Dimpe, H. F. Ogutu, E. O. Oseghe, M. M. Motsa, T. A. M. Msagati and B. B. Mamba, Materials, 2021, 14, 5094 CrossRef CAS PubMed.
- S. R. Khapate, T. A. J. Siddiqui and R. S. Mane, in Chapter 6 - Solvothermal Technique for the Synthesis of Metal Oxide Nanostructures, ed. R. Mane, V. Jadhav and A. Al-Enizi, Elsevier, 2023, p. 95 Search PubMed.
- S. R. Khapate, T. A. Siddiqui and R. S. Mane, Solvothermal Technique for the Synthesis of Metal Oxide Nanostructures, Elsevier. 2023, p. 95 Search PubMed.
- L. Sun, G. Yuan, L. Gao, J. Yang, M. Chhowalla, M. H. Gharahcheshmeh, K. K. Gleason, Y. S. Choi, B. H. Hong and Z. Liu, Nat. Rev. Methods Primers, 2021, 1, 5 CrossRef CAS.
- S. H. Li, M. Y. Qi, Z. R. Tang and Y. J. Xu, Chem. Soc. Rev., 2021, 50, 7539 RSC.
- L. Tang, J. Tan, H. Nong, B. Liu and H.-M. Cheng, Acc. Mater. Res., 2020, 2, 36 CrossRef.
- A. Tombesi, S. Li, S. Sathasivam, K. Page, F. L. Heale, C. Pettinari, C. J. Carmalt and I. P. Parkin, Sci. Rep., 2019, 9, 7549 CrossRef PubMed.
- D. Vernardou, Advances in Chemical Vapor Deposition, MDPI, 2020, p. 4167 Search PubMed.
- M. Saeed, Y. Alshammari, S. A. Majeed and E. Al-Nasrallah, Molecules, 2020, 25, 3856 CrossRef CAS PubMed.
- B. Qin, H. Ma, M. Hossain, M. Zhong, Q. Xia, B. Li and X. Duan, Chem. Mater., 2020, 32, 10321 CrossRef CAS.
- R. Kumar, N. Goel, D. K. Jarwal, Y. Hu, J. Zhang and M. Kumar, J. Mater. Chem. C, 2023, 11, 774 RSC.
- I. Sayago, E. Hontañón and M. Aleixandre, Tin Oxide Materials, 2020, p. 247 Search PubMed.
- F. Tu, M. Drost, S. Imre, J. Kiss, Z. Kónya and H. Marbach, Beilstein J. Nanotechnol., 2017, 8, 2592 CrossRef PubMed.
- T. T. Nguyen, J. Balamurugan, N. H. Kim and J. H. Lee, J. Mater. Chem. A, 2018, 6, 8669 RSC.
- Z. Liu, S. Yang, B. Sun, X. Chang, J. Zheng and X. Li, Angew. Chem., Int. Ed., 2018, 57, 10187 CrossRef CAS PubMed.
- Y. Jiang, Y. Wang, J. Jiang, S. Liu, W. Li, S. Huang, Z. Chen and B. Zhao, Electrochim. Acta, 2019, 312, 263 CrossRef CAS.
- D. Zhu, Q. Zhen, J. Xin, H. Ma, L. Tan, H. Pang and X. Wang, Sens. Actuators, B, 2020, 321, 128541 CrossRef CAS.
- A. Agarwal and B. R. Sankapal, J. Mater. Chem. A, 2021, 9, 20241 RSC.
- J. Nai and X. W. Lou, Adv. Mater., 2019, 31, 1706825 CrossRef PubMed.
- X. Li and J. Wang, Adv. Mater. Interfaces, 2020, 7, 2000676 CrossRef CAS.
- F. Chen, J. Xu, S. Wang, Y. Lv, Y. Li, X. Chen, A. Xia, Y. Li, J. Wu and L. Ma, Advanced Science, 2022, 9, 2200740 CrossRef CAS PubMed.
- X. Yi, Y. Guo, S. Chi, S. Pan, C. Geng, M. Li, Z. Li, W. Lv, S. Wu and Q. H. Yang, Adv. Funct. Mater., 2023, 33, 2303574 CrossRef CAS.
- X. F. Lu, S. L. Zhang, W. L. Sim, S. Gao and X. W. Lou, Angew. Chem., 2021, 133, 23067 CrossRef.
- R. Schlem, C. F. Burmeister, P. Michalowski, S. Ohno, G. F. Dewald, A. Kwade and W. G. Zeier, Adv. Energy Mater., 2021, 11, 2101022 CrossRef CAS.
- M. Bianchini, J. Wang, R. J. Clément, B. Ouyang, P. Xiao, D. Kitchaev, T. Shi, Y. Zhang, Y. Wang and H. Kim, Nat. Mater., 2020, 19, 1088 CrossRef CAS PubMed.
- B. Wang, C. Zhang, W. Zheng, Q. Zhang, Z. Bao, L. Kong and L. Li, Chem. Mater., 2019, 32, 308 CrossRef.
- X. Zhu, G. H. ten Brink, S. de Graaf, B. J. Kooi and G. Palasantzas, Chem. Mater., 2020, 32, 1627 CrossRef CAS.
- S. Gaan, Evaluation of Gas Phase: Mechanisms and Analyses, Elsevier., 2022, p. 117 Search PubMed.
- G. M. Tomboc, Y. Wang, H. Wang, J. Li and K. Lee, Energy Storage Mater., 2021, 39, 21 CrossRef.
- W. Liu, H. Zhi and X. Yu, Energy Storage Mater., 2019, 16, 290 CrossRef.
- L. Su, H. Li, Y. Xiao, G. Han and M. Zhu, J. Alloys Compd., 2019, 771, 117 CrossRef CAS.
- G. Chen, S. Tang, Y. Song, X. Meng, J. Yin, Y. Xia and Z. Liu, Chem. Eng. J., 2019, 361, 387 CrossRef CAS.
- N. Jiang, S. Shi, Y. Cui and B. Jiang, J. Alloys Compd., 2022, 929, 167229 CrossRef CAS.
- J. H. Bang and K. S. Suslick, Adv. Mater., 2010, 22, 1039 CrossRef CAS PubMed.
- S. V. Ley and C. M. Low, Ultrasound in Synthesis, Springer Science & Business Media, 2012 Search PubMed.
- A. Usman, A. Aris, B. Labaran, M. Darwish, A. Jagaba and J. New Mater, Electrochem. Syst., 2022, 25, 251 CAS.
- X. Lv and L. Xiang, Nanomaterials, 2022, 12, 3021 CrossRef CAS PubMed.
- N. Kumari, S. Kumar, P. Chauhan, G. A. Kaur, I. Kainthla and M. Shandilya, J. Inorg. Organomet. Polym. Mater., 2024, 1 Search PubMed.
- C. R. Bandeira, A. R. Dória, J. Y. C. Ribeiro, L. R. Prado, R. A. de Jesus, H. M. C. Andrade, R. S. de Santana Castro, L. F. R. Ferreira, S. M. Egues and R. T. Figueiredo, Mater. Chem. Phys., 2021, 265, 124521 CrossRef CAS.
- N. A. Neto, A. Lima, R. Wilson, T. Nicacio, M. Bomio and F. Motta, Mater. Sci. Semicond. Process., 2022, 139, 106311 CrossRef.
- P. Mohanty, R. Mahapatra, P. Padhi, C. V. Ramana and D. K. Mishra, Nano-Struct. Nano-Objects, 2020, 23, 100475 CrossRef CAS.
- M. Amar, M. Benzerzour, J. Kleib and N.-E. Abriak, Int. J. Sediment Res., 2020, 36(1), 92–109 CrossRef.
- B. Kalsi, S. Singh and M. Alam, J. Food Process Eng., 2023, 46(6), e14163 CrossRef CAS.
- R. Han, Y. Wang, S. Xing, C. Pang, Y. Hao, C. Song and Q. Liu, Chem. Eng. J., 2022, 450, 137952 CrossRef CAS.
- B. Purnama and A. T. Wijayanta, J. King Saud Univ., Sci., 2019, 31, 956 CrossRef.
- G. Zhu, C. Zhang, C. Zhang and Y. Yi, Chem. Eng. J., 2024, 497, 154512 CrossRef CAS.
- A. K. Soni, R. Joshi and R. S. Ningthoujam, in Hot Injection Method for Nanoparticle Synthesis: Basic Concepts, Examples and Applications, ed. A. K. Tyagi and R. S. Ningthoujam, Springer Singapore, Singapore, 2021, p. 383 Search PubMed.
- J. Yuan, Y. Zhang, F. Chen and Z. Gu, J. Mater. Chem. C, 2024, 14729 RSC.
- X. Li, W. Xing, T. Hu, K. Luo, J. Wang and W. Tang, Coord. Chem. Rev., 2022, 473, 214811 CrossRef CAS.
- F. Bu, W. Chen, M. F. Aly Aboud, I. Shakir, J. Gu and Y. Xu, J. Mater. Chem. A, 2019, 7, 14526 RSC.
- L. Yang, X. Yuan, W. Liang, R. Song, Q. Wang, C. Chen and Z. Bai, Catal. Lett., 2024, 4116 CrossRef CAS.
- J. Zheng, M. S. Kim, Z. Tu, S. Choudhury, T. Tang and L. A. Archer, Chem. Soc. Rev., 2020, 49, 2701 RSC.
- V. T. Chebrolu, B. Balakrishnan, S. Aravindha Raja, I. Cho, J.-S. Bak and H.-J. Kim, New J. Chem., 2020, 44, 7690 RSC.
- H. Wu, X. Li, L. Chen and Y. Dan, Batteries Supercaps, 2019, 2, 144 CrossRef CAS.
- A. Agarwal and T. Soga, in Electroless Assisted Nanostructured Morphologies, ed. B. R. Sankapal, A. Ennaoui, R. B. Gupta and C. D. Lokhande, Springer Nature Singapore, Singapore, 2023, p. 211 Search PubMed.
- A. Ramesh, S. Basu and M. Sterlin Leo Hudson, in Polymer-Metal Phosphide Nanocomposites for Flexible Supercapacitors, ed. R. K. Gupta, Springer Nature Singapore, Singapore, 2023, pp. 283 Search PubMed.
- H.-K. Kang and H.-C. Shin, J. Electrochem. Sci. Technol., 2020, 11, 155 CrossRef CAS.
- S. Battiato, L. Bruno, A. L. Pellegrino, A. Terrasi and S. Mirabella, Catal. Today, 2023, 423, 113929 CrossRef CAS.
- G. Chen, R. Li and L. Huang, Nanoscale, 2023, 15, 13909 RSC.
- A. Kumar, P. Choudhary, A. Kumar, P. H. C. Camargo and V. Krishnan, Small, 2022, 18, e2101638 CrossRef PubMed.
- K. Li, X. Chen, J. Zhao, H. She, J. Huang, L. Wang and Q. Wang, ACS Appl. Energy Mater., 2022, 5, 10207 CrossRef CAS.
- M. Z. Iqbal, M. M. Faisal, M. Sulman, S. R. Ali, A. M. Afzal, M. A. Kamran and T. Alharbi, J. Energy Storage, 2020, 29, 101324 CrossRef.
- R. M. A. Hameed, Nanostructured Phosphides as Electrocatalysts for Green Energy Generation, Am. Chem. Soc., 2022, 191 CAS.
- F. Foroughi, J. J. Lamb, O. S. Burheim and B. G. Pollet, Catalysts, 2021, 11, 284 CrossRef CAS.
- Z. Li, J. Yang, T. Guang, B. Fan, K. Zhu and X. Wang, Small Methods, 2021, 5, 2100193 CrossRef CAS PubMed.
- W. Zhang, J. Bao, C. Lu, X. Zhou, X. Xia, J. Zhang, X. He, Y. Gan, H. Huang, C. Wang, W. Wan, R. Fang and Y. Xia, J. Solid State Electrochem., 2023, 51 Search PubMed.
- H. Tan, L. Sun, Y. Zhang, K. Wang and Y. Zhang, Adv. Sustainable Syst., 2022, 6(9), 2200183 CrossRef CAS.
- Z. A. Zeenat, M. Maqbool, M. Asif Hussain, R. Adel Pashameah, A. Shahzadi, N. Nazar, S. Iqbal, A. K. Alanazi, M. Naeem Ashiq and H. M. Abo-Dief, Fuel, 2023, 331, 125913 CrossRef CAS.
- Y. Yang, Y. Zhou, Z. Hu, W. Wang, X. Zhang, L. Qiang and Q. Wang, J. Alloys Compd., 2019, 772, 683 CrossRef CAS.
- Z. Jia, X. Kong, Z. Liu, X. Zhao, X. Zhao, F. He, Y. Zhao, M. Zhang and P. Yang, ChemSusChem, 2024, 17, e202301386 CrossRef CAS PubMed.
- Y. Jeung, H. Jung, D. Kim, H. Roh, C. Lim, J. W. Han and K. Yong, J. Mater. Chem. A, 2021, 9, 12203 RSC.
- M. Malekzadeh and M. T. Swihart, Chem. Soc. Rev., 2021, 50, 7132 RSC.
- F. W. Eagle, R. A. Rivera-Maldonado and B. M. Cossairt, Annu. Rev. Mater. Res., 2021, 51, 541 CrossRef CAS.
- G. E. Ayom, M. D. Khan, J. Choi, R. K. Gupta, W. E. van Zyl and N. Revaprasadu, Dalton Trans., 2021, 50, 11821 RSC.
- A. K. Tyagi and R. S. Ningthoujam, In Handbook on Synthesis Strategies for Advanced Materials, Springer, 2021 Search PubMed.
- X. Tang, S. Gao, D. Zhang, X. Xia, J. Wang, W. She, B. Yang, X. Meng, K. Wang, Z. Han and B. Wang, J. Alloys Compd., 2022, 923, 166289 CrossRef CAS.
- C. Lin, L. Ouyang, R. Hu, J. Liu, L. Yang, H. Shao and M. Zhu, Prog. Nat. Sci.:Mater. Int., 2021, 31, 567 CrossRef CAS.
- S. Liu, C. Luo, L. Chai and J. Ren, J. Solid State Electrochem., 2021, 25, 1975 CrossRef CAS.
- X. Liu, Y. Guo, W. Zhan and T. Jin, Catalysts, 2019, 9, 240 CrossRef.
- C. A. Downes, K. M. Van Allsburg, S. A. Tacey, K. A. Unocic, F. G. Baddour, D. A. Ruddy, N. J. LiBretto, M. M. O'Connor, C. A. Farberow, J. A. Schaidle and S. E. Habas, Chem. Mater., 2022, 34, 6255 CrossRef CAS.
- S.-B. Guo, W.-B. Zhang, Z.-Q. Yang, X. Bao, L. Zhang, Y.-W. Guo, X.-W. Han and J. Long, Crystals, 2022, 12 Search PubMed.
- Z. Li, Y. Zheng, Q. Liu, Y. Wang, D. Wang, Z. Li, P. Zheng and Z. Liu, J. Mater. Chem. A, 2020, 8, 19113 RSC.
- S. Dou, J. Xu, X. Cui, W. Liu, Z. Zhang, Y. Deng, W. Hu and Y. Chen, Adv. Energy Mater., 2020, 10(33), 2001331 CrossRef CAS.
- H.-M. Zhang, J.-J. Wang, Y. Meng and J. Sun, Int. J. Hydrogen Energy, 2022, 47, 36084 CrossRef CAS.
- P. Zhang, W. Wang, Z. Kou, J. Li, T. Wang and J. Guo, Ionics, 2021, 27, 801 CrossRef CAS.
- X. Wu, S. Chen, Y. Feng, Q. Yuan, J. Gao, Y. Chen, Y. Huang, Y. B. He and W. Gan, Mater. Today Phys., 2019, 9, 100132 CrossRef.
- J. Xu, A. Schulte, H. Schönherr, X. Jiang and N. Yang, Small Struct., 2021, 3(2), 2100183 CrossRef.
- S. Rafai, C. Qiao, Z. Wang, C. Cao, T. Mahmood, M. Naveed, W. Younas and S. Khalid, ChemElectroChem, 2019, 6, 5469 CrossRef CAS.
- S. Song, M. Guo, S. Zhang, K. Zhan, Y. Yan, J. Yang, B. Zhao and M. Xu, Electrochim. Acta, 2020, 331, 135431 CrossRef CAS.
- G.-d. Yi, C.-l. Fan, Z. Hu, W.-h. Zhang, S.-c. Han and J.-s. Liu, Electrochim. Acta, 2021, 383, 138370 CrossRef CAS.
- J. Xu, N. Yang, S. Yu, A. Schulte, H. Schonherr and X. Jiang, Nanoscale, 2020, 12, 13618 RSC.
- M. Wang, H. Xu, L. Zhou, T. Sun and Y. Tang, New J. Chem., 2024, 48, 1200 RSC.
- Y. Li, X. Wang, J. Meng, M. Song, M. Jiao, Q. Qin and L. Mi, New J. Chem., 2023, 47, 15143 RSC.
- L. Huang, Y. Yang, C. Zhang, H. Yu, T. Wang, X. Dong, D. Li and Z. Liu, Nanotechnology, 2020, 31, 225403 CrossRef CAS PubMed.
- X. Wang, J. Dai, H. Xie, C. Yang, L. He, T. Wu, X. Liu, Y. Xu, C. Yuan and L. Dai, Chem. Eng. J., 2022, 438, 135544 CrossRef CAS.
- Z. Fu, Z. Jiang, T. Hu and Z.-J. Jiang, Electrochim. Acta, 2022, 419, 140392 CrossRef CAS.
- J. Cai, X. Zhang, Y. Pan, Y. Kong and S. Lin, Int. J. Hydrogen Energy, 2021, 46, 34252 CrossRef CAS.
- Z. L. Choong, B. T. Goh, M. L. Ooi, K. C. Lau, R. C. S. Wong and K. W. Tan, Thin Solid Films, 2024, 788, 140150 CrossRef CAS.
- H.-b. Zheng, Y.-l. Li, Y.-l. Wang, F. Ma, P.-z. Gao, W.-m. Guo, H. Qin, X.-p. Liu and H.-n. Xiao, J. Alloys Compd., 2022, 894, 162411 CrossRef CAS.
- Q. Luo, L. Sun, Y. Zhao, C. Wang, H. Xin, D. Li and F. Ma, J. Mater. Sci. Technol., 2023, 145, 165 CrossRef CAS.
- M. Song, Y. Liu, J. Hong, X. Wang and X. Huang, J. Adv. Ceram., 2023, 12, 1872 CrossRef CAS.
- S. Duraisamy, A. Ganguly, P. K. Sharma, J. Benson, J. Davis and P. Papakonstantinou, ACS Appl. Nano Mater., 2021, 4, 2642 CrossRef CAS.
- S. Zhao, W. Xu, Z. Yang, X. Zhang and Q. Zhang, Electrochim. Acta, 2020, 331, 135265 CrossRef CAS.
- B. Reddy, M. Premasudha, N. Reddy, H.-J. Ahn, J.-H. Ahn and K.-K. Cho, J. Energy Storage, 2021, 39, 102660 CrossRef.
- Y. Li, Y. Zhang, X. Tong, X. Wang, L. Zhang, X. Xia and J. Tu, J. Mater. Chem. A, 2021, 9, 1418 RSC.
- S. Liu, Y. Yin, M. Wu, K. S. Hui, K. N. Hui, C. Y. Ouyang and S. C. Jun, Small, 2019, 15, 1803984 CrossRef PubMed.
- F. Liu, N. Wang, C. Shi, J. Sha, L. Ma, E. Liu and N. Zhao, Chem. Eng. J., 2022, 431, 133923 CrossRef CAS.
- Y. T. Baheri, M. A. Hedayati, M. Maleki and H. Karimian, J. Energy Storage, 2023, 68, 107682 CrossRef.
- H. Cheng, R. Liu, R. Zhang, L. Huang and Q. Yuan, Nanoscale Adv., 2023, 5, 2394 RSC.
- S. Sarwar, M.-C. Lin, M. R. Ahasan, Y. Wang, R. Wang and X. Zhang, Adv. Compos. Hybrid Mater., 2022, 5, 2339 CrossRef CAS.
- N. H. A. Rosli, K. S. Lau, T. Winie, S. X. Chin, S. Zakaria and C. H. Chia, J. Energy Storage, 2022, 52, 104991 CrossRef.
- Y. Wang, T. Ge, X. Zhan, S. Liu, Y. Wei and Y. Qiao, J. Sulfur Chem., 2024, 1 Search PubMed.
- R. Levinas, N. Tsyntsaru and H. Cesiulis, Electrochim. Acta, 2019, 317, 427 CrossRef CAS.
- A. Teli, S. Beknalkar, S. Mane, T. Bhat, B. Kamble, S. Patil, S. Sadale and J. Shin, Ceram. Int., 2022, 48, 29002 CrossRef CAS.
- J. Zhao, H. Ren, C. Gu, W. Guan, X. Song and J. Huang, J. Alloys Compd., 2019, 781, 174 CrossRef CAS.
- D. Saha, V. Patel, P. R. Selvaganapathy and P. Kruse, Nanoscale Adv., 2022, 4, 125 RSC.
- Q. Zhang, W. He, Y. Wang, D. Pei and X. Zheng, Nano, 2019, 14, 1950055 CrossRef CAS.
- Y. Liu and R. Li, Ultrason. Sonochem., 2020, 63, 104923 CrossRef CAS PubMed.
- H. Wang, L. Geng, Z. Zhang, P. Zhong, F. Liu, Y. Xie, Y. Zhao, P. Li and X. Ma, Nanotechnology, 2023, 34, 375601 CrossRef CAS PubMed.
- L. N. Khandare, D. J. Late and N. B. Chaure, J. Energy Storage, 2023, 74, 109336 CrossRef.
- C. A. Beaudette, J. T. Held, K. A. Mkhoyan and U. R. Kortshagen, ACS Omega, 2020, 5, 21853 CrossRef CAS PubMed.
- D. He, Y. Yang, Z. Liu, J. Shao, J. Wu, S. Wang, L. Shen and N. Bao, Nano Res., 2020, 13, 1029 CrossRef CAS.
- H. Ganesha, S. Veeresh, Y. Nagaraju, M. Vandana, M. Basappa, H. Vijeth and H. Devendrappa, Nanoscale Adv., 2022, 4, 521 RSC.
- Q. Lin, X. Dong, Y. Wang, N. Zheng, Y. Zhao, W. Xu and T. Ding, J. Mater. Sci., 2020, 55, 6637 CrossRef CAS.
- V. Klimas, C. Bittencourt, G. Valušis and A. Jagminas, Mater. Charact., 2021, 179, 111351 CrossRef CAS.
- H. Liu, M. Zhang, Z. Song, T. Ma, Z. Huang, A. Wang and S. Shao, J. Alloys Compd., 2021, 881, 160660 CrossRef CAS.
- C. He, W. Yin, X. Li, J. Zheng, B. Tang and Y. Rui, Electrochim. Acta, 2021, 365, 137353 CrossRef CAS.
- X. Wang, W. You, L. Yang, G. Chen, Z. Wu, C. Zhang, Q. Chen and R. Che, Nanoscale Adv., 2022, 4, 3398 RSC.
- D. Chen, J. Xiao, H. Zhou and A. Yuan, ChemistrySelect, 2020, 5, 3130 CrossRef CAS.
- G.-A. Li, C.-Y. Wang, W.-C. Chang and H.-Y. Tuan, ACS Nano, 2016, 10, 8632 CrossRef CAS PubMed.
- J. Yu, Q. Li, Y. Li, C.-Y. Xu, L. Zhen, V. P. Dravid and J. Wu, Adv. Funct. Mater., 2016, 26, 7644 CrossRef CAS.
- F. Yang, N. Kang, J. Yan, X. Wang, J. He, S. Huo and L. Song, Metals, 2018, 8, 359 CrossRef.
- Q. Zhou, J. Feng, X. Peng, L. Zhong and R. Sun, J. Energy Chem., 2020, 45, 45 CrossRef.
- M. H. Suliman, A. Adam, L. Li, Z. Tian, M. N. Siddiqui, Z. H. Yamani and M. Qamar, ACS Sustain. Chem. Eng., 2019, 7, 17671 CrossRef CAS.
- Y. Shi, M. Li, Y. Yu and B. Zhang, Energy Environ. Sci., 2020, 13, 4564 RSC.
- S. Imani Yengejeh, J. Liu, S. A. Kazemi, W. Wen and Y. Wang, ACS Omega, 2020, 5, 5994 CrossRef CAS PubMed.
- Y. Li, R. Li, D. Wang, H. Xu, F. Meng, D. Dong, J. Jiang, J. Zhang, M. An and P. Yang, Int. J. Hydrogen Energy, 2021, 46, 5131 CrossRef CAS.
- N. Attarzadeh, D. Das, S. N. Chintalapalle, S. Tan, V. Shutthanandan and C. Ramana, ACS Appl. Mater. Interfaces, 2023, 15, 22036 CrossRef CAS PubMed.
- Z. Zhu, Y. Tang, W. R. Leow, H. Xia, Z. Lv, J. Wei, X. Ge, S. Cao, Y. Zhang and W. Zhang, Angew. Chem., 2019, 131, 3559 CrossRef.
- T. Chen, H. Zou, X. Wu, C. Liu, B. Situ, L. Zheng and G. Yang, ACS Appl. Mater. Interfaces, 2018, 10, 12453 CrossRef CAS PubMed.
- X. Hu, P. Jiang, J. Wan, Y. Xu and X. Sun, J. Coat. Technol. Res., 2009, 6, 275 CrossRef CAS.
- X. Feng and H. Yang, J. Vinyl Addit. Technol., 2023, 29, 522 CrossRef CAS.
- H. Chen, X. Liang, Y. Liu, X. Ai, T. Asefa and X. Zou, Adv. Mater., 2020, 32, 2002435 CrossRef PubMed.
- A. Altuntepe, S. Erkan, M. A. Olğar, S. Çelik and R. Zan, Int. J. Hydrogen Energy, 2024, 56, 690 CrossRef CAS.
- J. Park, S. Bhoyate, Y.-H. Kim, Y.-M. Kim, Y. H. Lee, P. Conlin, K. Cho and W. Choi, ACS Nano, 2021, 15, 12267 CrossRef CAS PubMed.
- D. Sahoo, J. Shakya, N. Ali, W. J. Yoo and B. Kaviraj, Langmuir, 2022, 38, 1578 CrossRef CAS PubMed.
- G. Swain, S. Sultana and K. Parida, Nanoscale, 2021, 13, 9908 RSC.
- G. Liu, L. Ding, Y. Meng, A. Ali, G. Zuo, X. Meng, K. Chang, O. L. Li and J. Ye, Carbon Energy, 2024, 6, e521 CrossRef CAS.
- L. Zhao, Y. Li, M. Yu, Y. Peng and F. Ran, Advanced Science, 2023, 10, 2300283 CrossRef CAS PubMed.
- X. Qin, B. Yan, D. Kim, Z. Teng, T. Chen, J. Choi, L. Xu and Y. Piao, Appl. Catal., B, 2022, 304, 120923 CrossRef CAS.
- Q. Wang, F. Jia, S. Song and Y. Li, Sep. Purif. Technol., 2020, 236, 116298 CrossRef CAS.
- Y. Yao, J. He, X. Zhu, L. Mu, J. Li, K. Li and M. Qu, Int. J. Hydrogen Energy, 2024, 51, 207 CrossRef CAS.
- S. Das, G. Swain and K. Parida, Mater. Chem. Front., 2021, 5, 2143 RSC.
- R. Mohili, N. Hemanth, H. Jin, K. Lee and N. Chaudhari, J. Mater. Chem. A, 2023, 11, 10463 RSC.
- F. Fioravanti, S. Martínez, S. Delgado, G. García, J. L. Rodriguez, E. P. Tejera and G. I. Lacconi, Electrochim. Acta, 2023, 441, 141781 CrossRef CAS.
- X. Li, X. Wang, Z. Sun, F. Li, Y. Fu, K. Zhao, G. Zhao, C. Zhu and X. Xu, Chem. Eng. J., 2024, 495, 153381 CrossRef CAS.
- P. Wang, X. Wang, R. Diao, Y. Guo, Y. Wang, C. Zhou, K.-F. Xie, S. Sun and Y.-H. Zhang, J. Mater. Sci.: Mater. Electron., 2021, 32, 14047 CrossRef CAS.
- S. Zhao, J. Xu, Z. Li, Z. Liu and Y. Li, J. Colloid Interface Sci., 2019, 555, 689 CrossRef CAS PubMed.
- V.-D. Hodoroaba, Energy-dispersive X-Ray Spectroscopy (EDS), Elsevier, 2020, p. 397 Search PubMed.
- L. E. Franken, K. Grünewald, E. J. Boekema and M. C. Stuart, Small, 2020, 16, 1906198 CrossRef CAS PubMed.
- S. E. Jun, S. Choi, S. Choi, T. H. Lee, C. Kim, J. W. Yang, W.-O. Choe, I.-H. Im, C.-J. Kim and H. W. Jang, Nano-Micro Lett., 2021, 13, 81 CrossRef CAS PubMed.
- S. A. Chambers, Surf. Sci. Rep., 2024, 79, 100638 CrossRef CAS.
- T. Wang, J. Liu, Y. Ma, S. Han, C. Gu and J. Lian, Electrochim. Acta, 2021, 392, 138976 CrossRef CAS.
- C. Lu, W. An, T. Shen, T. Cao, Y. Gao, K. Wang, Y. Wong, C. Cao, C. Wang, G. Huang and S. Xu, Int. J. Hydrogen Energy, 2024, 77, 203 CrossRef CAS.
- C. Dang, P. Feng, S. He, L. Zhao, A. Shan, M. Li, L. Kong and L. Gao, Electrochim. Acta, 2023, 462, 142771 CrossRef CAS.
- L. Wan, Y. Wang, Y. Zhang, C. Du, J. Chen, M. Xie, Z. Tian and W. Zhang, J. Power Sources, 2021, 506, 230096 CrossRef CAS.
- X. Pu, D. Zhao, C. Fu, Z. Chen, S. Cao, C. Wang and Y. Cao, Angew. Chem., Int. Ed., 2021, 60, 21310 CrossRef CAS PubMed.
- T. Kim, W. Choi, H.-C. Shin, J.-Y. Choi, J. M. Kim, M.-S. Park and W.-S. Yoon, J. Electrochem. Sci. Technol., 2020, 11, 14 CrossRef CAS.
- A. Patra, K. Namsheer, J. R. Jose, S. Sahoo, B. Chakraborty and C. S. Rout, J. Mater. Chem. A, 2021, 9, 25852 RSC.
- S. Aderyani, P. Flouda, S. Shah, M. Green, J. Lutkenhaus and H. Ardebili, Electrochim. Acta, 2021, 390, 138822 CrossRef CAS.
- L. Wang, M. Huang, J. Huang, X. Tang, L. Li, M. Peng, K. Zhang, T. Hu, K. Yuan and Y. Chen, J. Mater. Chem. A, 2021, 9, 15404 RSC.
- Q. Zong, C. Liu, H. Yang, Q. Zhang and G. Cao, Nano Today, 2021, 38, 101201 CrossRef CAS.
- J. Zhang, H. Yu, J. Yang, X. Zhu, M. Hu and J. Yang, J. Alloys Compd., 2022, 924, 166613 CrossRef CAS.
- O. Gharbi, M. T. Tran, B. Tribollet, M. Turmine and V. Vivier, Electrochim. Acta, 2020, 343, 136109 CrossRef CAS.
- A. G. Kenesi, M. Ghorbani and M. S. Lashkenari, Int. J. Hydrogen Energy, 2022, 47, 38849 CrossRef CAS.
- B. Ramulu, J. A. Shaik, A. R. Mule and J. S. Yu, Mater. Sci. Eng., R, 2024, 160, 100820 CrossRef.
- H. Rashid Khan and A. Latif Ahmad, J. Ind. Eng. Chem., 2025, 141, 46 CrossRef CAS.
- Y. Jin, H. Ao, K. Qi, X. Zhang, M. Liu, T. Zhou, S. Wang, G. Xia and Y. Zhu, Mater. Today Energy, 2021, 19, 100598 CrossRef CAS.
- W. Zhao, L. Zou, L. Zhang, X. Fan, H. Zhang, F. Pagani, E. Brack, L. Seidl, X. Ou and K. Egorov, Small, 2022, 18, 2107357 CrossRef CAS PubMed.
- N. O. Laschuk, E. B. Easton and O. V. Zenkina, RSC Adv., 2021, 11, 27925 RSC.
- V. Vivier and M. E. Orazem, Chem. Rev., 2022, 122, 11131 CrossRef CAS PubMed.
- H. S. Magar, R. Y. Hassan and A. Mulchandani, Sensors, 2021, 21, 6578 CrossRef CAS PubMed.
- A. C. Lazanas and M. I. Prodromidis, ACS Meas. Sci. Au, 2023, 3, 162 CrossRef CAS PubMed.
- L. A. Santa-Cruz, F. C. Tavares, L. F. Loguercio, C. I. Dos Santos, R. A. Galvão, O. A. Alves, M. Z. Oliveira, R. M. Torresi and G. Machado, Phys. Chem. Chem. Phys., 2024, 26(40), 25748–25761 RSC.
- M. Y. Perdana, B. A. Johan, M. Abdallah, M. E. Hossain, M. A. Aziz, T. N. Baroud and Q. A. Drmosh, Chem. Rec., 2024, 24, e202400007 CrossRef CAS PubMed.
- A. Kalair, N. Abas, M. S. Saleem, A. R. Kalair and N. Khan, Energy Storage, 2021, 3, e135 CrossRef.
- A. G. Olabi, Q. Abbas, A. Al Makky and M. A. Abdelkareem, Energy, 2022, 248, 123617 CrossRef CAS.
- G. Z. Chen, Int. Mater. Rev., 2017, 62, 173 CrossRef CAS.
- J. Li, T. Xiao, X. Yu and M. Wang, J. Phys.: Conf. Ser., 2022, 2393, 012005 CrossRef CAS.
- J. Wu, Chem. Rev., 2022, 122, 10821 CrossRef CAS PubMed.
- N. Swain, B. Saravanakumar, M. Kundu, L. Schmidt-Mende and A. Ramadoss, J. Mater. Chem. A, 2021, 9, 25286 RSC.
- P. Bhojane, J. Energy Storage, 2022, 45, 103654 CrossRef.
- D. P. Chatterjee and A. K. Nandi, J. Mater. Chem. A, 2021, 9, 15880 RSC.
- T. Schoetz, L. Gordon, S. Ivanov, A. Bund, D. Mandler and R. Messinger, Electrochim. Acta, 2022, 412, 140072 CrossRef CAS.
- A. Hu, F. Li, W. Chen, T. Lei, Y. Li, Y. Fan, M. He, F. Wang, M. Zhou and Y. Hu, Adv. Energy Mater., 2022, 12, 2202432 CrossRef CAS.
- O. C. Esan, X. Shi, Z. Pan, X. Huo, L. An and T. Zhao, Adv. Energy Mater., 2020, 10, 2000758 CrossRef CAS.
- C. Li, B. Liu, N. Jiang and Y. Ding, Nano Res. Energy, 2022, 1(3), 9120031 CrossRef.
- M. Weiss, R. Ruess, J. Kasnatscheew, Y. Levartovsky, N. R. Levy, P. Minnmann, L. Stolz, T. Waldmann, M. Wohlfahrt-Mehrens and D. Aurbach, Adv. Energy Mater., 2021, 11, 2101126 CrossRef CAS.
- G. Wang, M. Yu and X. Feng, Chem. Soc. Rev., 2021, 50, 2388 RSC.
- P. Shi, L.-P. Hou, C.-B. Jin, Y. Xiao, Y.-X. Yao, J. Xie, B.-Q. Li, X.-Q. Zhang and Q. Zhang, J. Am. Chem. Soc., 2021, 144, 212 CrossRef PubMed.
- W. H. Li, Q. L. Ning, X. T. Xi, B. H. Hou, J. Z. Guo, Y. Yang, B. Chen and X. L. Wu, Adv. Mater., 2019, 31, 1804766 CrossRef PubMed.
- Y. Lv, M. Zhao, Y. Du, Y. Kang, Y. Xiao and S. Chen, Energy Environ. Sci., 2022, 15, 4748 RSC.
- R. Yuan, Y. Dong, R. Hou, S. Zhang and H. Song, J. Electrochem. Soc., 2022, 169, 030504 CrossRef CAS.
- A. Roy, M. Sotoudeh, S. Dinda, Y. Tang, C. Kübel, A. Groß, Z. Zhao-Karger, M. Fichtner and Z. Li, Nat. Commun., 2024, 15, 492 CrossRef CAS PubMed.
- H. Wu, C. Feng, L. Zhang, J. Zhang and D. P. Wilkinson, Electrochem. Energy Rev., 2021, 4, 473 CrossRef CAS.
- P. Moriarty and D. Honnery, Int. J. Hydrogen Energy, 2019, 44, 16029 CrossRef CAS.
- S. E. Hosseini and M. A. Wahid, Int. J. Energy Res., 2020, 44, 4110 CrossRef.
- N. Mahmood, Y. Yao, J. W. Zhang, L. Pan, X. Zhang and J. J. Zou, Advanced Science, 2018, 5, 1700464 CrossRef PubMed.
- N. Danilovic, R. Subbaraman, D. Strmcnik, V. Stamenkovic and N. Markovic, J. Serb. Chem. Soc., 2013, 78, 2007 CrossRef CAS.
- S. K. T. Aziz, S. Sultana, A. Kumar, S. Riyajuddin, M. Pal and A. Dutta, Cell Rep. Phys. Sci., 2023, 4, 101747 CrossRef CAS.
- Q. Yan, X. Chen, T. Wei, G. Wang, M. Zhu, Y. Zhuo, K. Cheng, K. Ye, K. Zhu and J. Yan, ACS Sustain. Chem. Eng., 2019, 7, 7804 CrossRef CAS.
- J. F. Callejas, J. M. McEnaney, C. G. Read, J. C. Crompton, A. J. Biacchi, E. J. Popczun, T. R. Gordon, N. S. Lewis and R. E. Schaak, ACS Nano, 2014, 8, 11101 CrossRef CAS PubMed.
- T. Reier, H. N. Nong, D. Teschner, R. Schlögl and P. Strasser, Adv. Energy Mater., 2017, 7, 1601275 CrossRef.
- X. Xie, L. Du, L. Yan, S. Park, Y. Qiu, J. Sokolowski, W. Wang and Y. Shao, Adv. Funct. Mater., 2022, 32, 2110036 CrossRef CAS.
- X. Hu, R. Wang, W. Feng, C. Xu and Z. Wei, J. Energy Chem., 2023, 81, 167 CrossRef CAS.
- C. Niu, G. Han, H. Song, S. Yuan and W. Hou, J. Colloid Interface Sci., 2020, 561, 117 CrossRef CAS PubMed.
- G. Li, Y. Feng, Y. Yang, X. Wu, X. Song and L. Tan, Nano Mater. Sci., 2024, 6, 174 CrossRef CAS.
- S. T. Aziz, S. Kumar, S. Riyajuddin, K. Ghosh, G. D. Nessim and D. P. Dubal, J. Phys. Chem. Lett., 2021, 12, 5138 CrossRef CAS PubMed.
- K. Prakash, S. Harish, K. Silambarasan, T. Logu, R. Ramesh, J. Archana and M. Navaneethan, J. Colloid Interface Sci., 2022, 628, 131 CrossRef CAS PubMed.
- Z. Zhao, Y. Duan, F. Chen, Z. Tian, R. Pathak, J. W. Elam, Z. Yi, Y. Wang and X. Wang, Chem. Eng. J., 2022, 450, 138310 CrossRef CAS.
- L. Liu, X. Yin, W. Li, D. Wang, J. Duan, X. Wang, Y. Zhang, D. Peng and Y. Zhang, Small, 2024, 20, 2308564 CrossRef CAS PubMed.
- X. Qian, G. Zhu, K. Wang, F. Zhang, K. Liang, W. Luo and J. Yang, Chem. Eng. J., 2020, 381, 122651 CrossRef CAS.
- X. Zhang, M. Jia, Q. Zhang, N. Zhang, X. Wu, S. Qi and L. Zhang, Chem. Eng. J., 2022, 448, 137743 CrossRef CAS.
- G. Chang, Y. Zhao, L. Dong, D. P. Wilkinson, L. Zhang, Q. Shao, W. Yan, X. A. Sun and J. Zhang, J. Mater. Chem. A, 2020, 8, 4996 RSC.
- W. Hu, L. Xie, C. Gu, W. Zheng, Y. Tu, H. Yu, B. Huang and L. Wang, Coord. Chem. Rev., 2024, 506, 215715 CrossRef CAS.
- J. Bao, Y. Zhou, Y. Zhang, X. Sheng, Y. Wang, S. Liang, C. Guo, W. Yang, T. Zhuang and Y. Hu, J. Mater. Chem. A, 2020, 8, 22181 RSC.
- L. Li, Z. Qin, L. Ries, S. Hong, T. Michel, J. Yang, C. Salameh, M. Bechelany, P. Miele and D. Kaplan, ACS Nano, 2019, 13, 6824 CrossRef CAS PubMed.
- Y. Hu, H. Yu, L. Qi, J. Dong, P. Yan, T. Taylor Isimjan and X. Yang, ChemSusChem, 2021, 14, 1565 CrossRef CAS PubMed.
- H. K. Kim, H. Jang, X. Jin, M. G. Kim and S.-J. Hwang, Appl. Catal., B, 2022, 312, 121391 CrossRef CAS.
- Y. Li, J. Meng, X. Wang, M. Song, M. Jiao, Q. Qin and L. Mi, Dalton Trans., 2023, 52, 14613 RSC.
- S. Wang, G. Li, G. Du, X. Jiang, C. Feng, Z. Guo and S.-J. Kim, Chin. J. Chem. Eng., 2010, 18, 910 CrossRef CAS.
- H. Zhou, Y. Zhao, Y. Jin, Q. Fan, Y. Dong and Q. Kuang, J. Power Sources, 2023, 560, 232715 CrossRef CAS.
- Z. Kong, Z. Liang, M. Huang, H. Tu, K. Zhang, Y. Shao, Y. Wu and X. Hao, J. Alloys Compd., 2023, 930, 167328 CrossRef CAS.
- Y. Xia, T. Yang, Z. Wang, T. Mao, Z. Hong, J. Han, D.-L. Peng and G. Yue, Adv. Funct. Mater., 2023, 33, 2302830 CrossRef CAS.
- M. Song, Y. Liu, J. Hong, X. Wang, X. Huang and J. Adv, Ceram, 2023, 12, 1872 CAS.
- Z. Xu, C. Du, H. Yang, J. Huang, X. Zhang and J. Chen, Chem. Eng. J., 2021, 421, 127871 CrossRef CAS.
- J. Wu, F. Yan, Z. Huang, J. Liu, H. Huang, Y. Liang, J. Li, F. Yuan, X. Liang, W. Zhou and J. Guo, J. Energy Storage, 2024, 97, 112958 CrossRef.
- H. Tan, Y. Zhang, Y. Geng, H. Li, S. Bi, Z. Xia, Q. Yang, Q. Wei and S. Chen, Inorg. Chem., 2024, 63, 13484 CrossRef CAS PubMed.
- J. Du, Q. Han, C. Liu, H. Wu, L. zheng and Z. Yang, Appl. Surf. Sci., 2024, 649, 159098 CrossRef CAS.
- Y. Jiang, S. Lei and M. Wang, ACS Appl. Mater. Interfaces, 2024, 16, 30521 CrossRef CAS PubMed.
- W. Chen, X. Yan, Z. Liu, X. Zhang and C. Du, Int. J. Hydrogen Energy, 2023, 48, 29969 CrossRef CAS.
- B. Zhang, K. Xu, X. Fu, S. Guan, X. Li and Z. Peng, J. Alloys Compd., 2021, 856, 158094 CrossRef CAS.
- A. Wu, Y. Gu, Y. Xie, C. Tian, H. Yan, D. Wang, X. Zhang, Z. Cai and H. Fu, ACS Appl. Mater. Interfaces, 2019, 11, 25986 CrossRef CAS PubMed.
- Y. Jiang, Y. Lu, J. Lin, X. Wang and Z. Shen, Small Methods, 2018, 2(5), 1700369 CrossRef.
- D. Chen, J. Xiao, H. Zhou and A. Yuan, ChemistrySelect, 2020, 5, 3130 CrossRef CAS.
- Z. Duan, H. Liu, X. Tan, A. Umar and X. Wu, Catal. Commun., 2022, 162, 106379 CrossRef CAS.
- Y.-R. Liu, W.-H. Hu, X. Li, B. Dong, X. Shang, G.-Q. Han, Y.-M. Chai, Y.-Q. Liu and C.-G. Liu, Appl. Surf. Sci., 2016, 383, 276 CrossRef CAS.
- D. Ma, K. Meng, J. Ma, Z. Jia, Y. Wang, L. Liu, G. Zhu and T. Qi, Int. J. Hydrogen Energy, 2019, 44, 31960 CrossRef CAS.
- X. Zeng, H. Zhang, R. Yu, G. D. Stucky and J. Qiu, J. Mater. Chem. A, 2023, 11, 14272 RSC.
- A. Muthurasu, V. Maruthapandian and H. Y. Kim, Appl. Catal., B, 2019, 248, 202 CrossRef CAS.
- M. Luo, S. Liu, W. Zhu, G. Ye, J. Wang and Z. He, Chem. Eng. J., 2022, 428, 131055 CrossRef CAS.
- Q. Wang, Z.-Y. Tian, W.-J. Cui, N. Hu, S.-M. Zhang, Y.-Y. Ma and Z.-G. Han, Int. J. Hydrogen Energy, 2022, 47, 12629 CrossRef CAS.
- J. Xu, J. Rong, Y. Zheng, Y. Zhu, K. Mao, Z. Jing, T. Zhang, D. Yang and F. Qiu, Electrochim. Acta, 2021, 385, 138438 CrossRef CAS.
- Y. Huang, X. Xie, Y. Zhang, J. Ding, L. Liu, Y. Fan, H. Lv, Y. Liu and Q. Cai, Appl. Surf. Sci., 2020, 520, 146340 CrossRef CAS.
- Z. Huang, X. Li, X. Xiang, T. Gao, Y. Zhang and D. Xiao, J. Mater. Chem. A, 2018, 6, 23746 RSC.
- Y. Hu, M. Liu, Q. Yang, L. Kong and L. Kang, J. Energy Chem., 2017, 26, 49 CrossRef.
- B. Liang, Z. Zheng, M. Retana, K. Lu, T. Wood, Y. Ai, X. Zu and W. Zhou, Nanotechnology, 2019, 30, 295401 CrossRef CAS PubMed.
- Y.-C. Chen, Z.-B. Chen, Y.-G. Lin and Y.-K. Hsu, ACS Sustain. Chem. Eng., 2017, 5, 3863 CrossRef CAS.
- J. Tian, H. Zhang and Z. Li, ACS Appl. Mater. Interfaces, 2018, 10, 29511 CrossRef CAS PubMed.
- B. Wang, R. Hu, J. Zhang, Z. Huang, H. Qiao, L. Gong and X. Qi, J. Am. Ceram. Soc., 2019, 103, 1088 CrossRef.
- A. Lathe, A. Ansari, R. Badhe, A. M. Palve and S. S. Garje, ACS Omega, 2021, 6, 13008 CrossRef CAS PubMed.
- M. Manuraj, J. Chacko, K. N. Narayanan Unni and R. B. Rakhi, J. Alloys Compd., 2020, 836, 155420 CrossRef CAS.
- G. P. Awasthi, M. B. Poudel, M. Shin, K. P. Sharma, H. J. Kim and C. Yu, J. Energy Storage, 2021, 42, 103140 CrossRef.
- N. Chaudhary, A. Kumar, S. Imtiyaz and M. Khanuja, ECS J. Solid State Sci. Technol., 2021, 10, 053005 CrossRef CAS.
- D. Sahoo, J. Shakya, S. Choudhury, S. S. Roy, L. Devi, B. Singh, S. Ghosh and B. Kaviraj, ACS Omega, 2022, 7, 16895 CrossRef CAS PubMed.
- F. Lu, J. Wang, X. Sun and Z. Chang, Mater. Des., 2020, 189, 108503 CrossRef CAS.
- A. Raza, A. Rasheed, A. Farid, M. Yousaf, N. Ayub and I. A. Khan, J. Energy Storage, 2024, 84, 110811 CrossRef.
- N. Chaudhary and M. Khanuja, Energy Fuels, 2021, 36, 1034 CrossRef.
- M. Yao, A. Liu, C. Xing, B. Li, S. Pan, J. Zhang, P. Su and H. Zhang, Chem. Eng. J., 2020, 394, 124883 CrossRef CAS.
- P. S. Shukla, A. Agrawal, A. Gaur and G. D. Varma, J. Energy Storage, 2023, 59, 106580 CrossRef.
- X. Liu, W. Yang, Z. Liu, H. Fan and W. Zheng, Acta Metall. Sin., 2021, 34, 401 CrossRef CAS.
- Z. Hu, L. Wang, K. Zhang, J. Wang, F. Cheng, Z. Tao and J. Chen, Angew Chem. Int. Ed. Engl., 2014, 53, 12794 CrossRef CAS PubMed.
- Y. Xia, T. Yang, Z. Wang, T. Mao, Z. Hong, J. Han, D. L. Peng and G. Yue, Adv. Funct. Mater., 2024, 34(6), 2310995 CrossRef.
- Y. Yang, J. Xia, X. Guan, Z. Wei, J. Yu, S. Zhang, Y. Xing and P. Yang, Small, 2022, 18, e2204970 CrossRef PubMed.
- J. Liu, T. Zhou, T. Han, L. Zhu, Y. Wang, Y. Hu and Z. Chen, Chem. Commun., 2022, 58, 5108 RSC.
- Y. Wang, L. Zhang, H. Li, Y. Wang, L. Jiao, H. Yuan, L. Chen, H. Tang and X. Yang, J. Power Sources, 2014, 253, 360 CrossRef CAS.
- Y. Li, X. Yan, Z. Zhou, J. Liu, Z. Zhang, X. Guo, H. Peng and G. Li, Appl. Surf. Sci., 2022, 574, 151586 CrossRef CAS.
- D. Lei, W. Shang, X. Zhang, Y. Li, S. Qiao, Y. Zhong, X. Deng, X. Shi, Q. Zhang, C. Hao, X. Song and F. Zhang, ACS Nano, 2021, 15, 20478 CrossRef CAS PubMed.
- Y. Zhao, M. Bi, F. Qian, P. Zeng, M. Chen, R. Wang, Y. Liu, Y. Ding and Z. Fang, ChemElectroChem, 2018, 5, 3953 CrossRef CAS.
- B. Liu, D. Kong, Y. Wang, Y. V. Lim, S. Huang and H. Y. Yang, FlatChem, 2018, 10, 14 CrossRef CAS.
- F. Chen, D. Shi, M. Yang, H. Jiang, Y. Shao, S. Wang, B. Zhang, J. Shen, Y. Wu and X. Hao, Adv. Funct. Mater., 2020, 31, 2009973 Search PubMed.
- Z. Wang, C. Cui, Y. Zhao, Q. Cui, H. Li, Z. Zhao, C. Wu and J. Wei, J. Alloys Compd., 2023, 967, 171820 CrossRef CAS.
- J. Zhang, W. Xi, F. Yu, Y. Zhang, R. Wang, Y. Gong, B. He, H. Wang and J. Jin, Chem. Eng. J., 2023, 475, 146009 CrossRef CAS.
- Z. Zhang, J. Zhao, M. Xu, H. Wang, Y. Gong and J. Xu, Nanotechnology, 2018, 29, 335401 CrossRef PubMed.
- L. Ma, X. Zhou, J. Sun, P. Zhang, B. Hou, S. Zhang, N. Shang, J. Song, H. Ye, H. Shao, Y. Tang and X. Zhao, J. Energy Chem., 2023, 82, 268 CrossRef CAS.
- J. Huang, Y. Yao, M. Huang, Y. Zhang, Y. Xie, M. Li, L. Yang, X. Wei and Z. Li, Small, 2022, 18, e2200782 CrossRef PubMed.
- T. Gu, J. Ren, S. Zhang, H. Guo, H. Wang, R.-P. Ren and Y.-K. Lv, J. Alloys Compd., 2022, 901, 163650 CrossRef CAS.
- B. Zhao, G. Suo, R. Mu, C. Lin, J. Li, X. Hou, X. Ye, Y. Yang and L. Zhang, J. Colloid Interface Sci., 2024, 668, 565 CrossRef CAS PubMed.
- T. Zhang, Y. Feng, J. Zhang, C. He, D. M. Itkis and J. Song, Mater. Today Nano, 2020, 12, 100089 CrossRef.
- L. Zhang, A. Xu, X. Shi, H. Zhang, Z. Wang, S. Shen, J. Zhang and W. Zhong, RSC Adv., 2024, 14, 19294 RSC.
- M. Wang, H. Xu, L. Zhou, T. Sun and Y. Tang, New J. Chem., 2024, 48, 1200 RSC.
- A. Wu, Y. Gu, Y. Xie, C. Tian, H. Yan, D. Wang, X. Zhang, Z. Cai and H. Fu, ACS Appl. Mater. Interfaces, 2019, 11, 25986 CrossRef CAS PubMed.
- L. Zhang, A. Xu, X. Shi, H. Zhang, Z. Wang, S. Shen, J. Zhang and W. Zhong, RSC Adv., 2024, 14, 19294 RSC.
- D. Meng, S. Ran, L. Gao, Y. Zhang, X. San, L. Zhang, R. Li and Q. Jin, Chem. Res. Chin. Univ., 2024, 40, 490 CrossRef CAS.
- Q. Luo, Y. Lv, P. Zhang, Z. Zhao, X. Bao, L. Guo, H. Luo, X. Fan, F. Ma, P-Mos2 Heterojunction as Self-Supported Electrode Enables Boosted Alkaline Hydrogen Evolution Reaction.
- K. N. Dinh, Q. Liang, C.-F. Du, J. Zhao, A. I. Y. Tok, H. Mao and Q. Yan, Nano Today, 2019, 25, 99 CrossRef CAS.
- X. Li, A. M. Elshahawy, C. Guan and J. Wang, Small, 2017, 13, 1701530 CrossRef PubMed.
- S. Bhat, J. Banday and M. Wahid, Energy Fuels, 2023, 37, 6012 CrossRef CAS.
- Y. Cao, Z. Chen, F. Ye, Y. Yang, K. Wang, Z. Wang, L. Yin and C. Xu, J. Alloys Compd., 2022, 896, 163103 Search PubMed.
- X. Wang, G. Zhang, B. Wang, Y. Wu and S. Guo, ACS Sustain. Chem. Eng., 2024, 12, 14018 Search PubMed.
- Z. Zhao, D. E. Schipper, A. P. Leitner, H. Thirumalai, J.-H. Chen, L. Xie, F. Qin, M. K. Alam, L. C. Grabow and S. Chen, Nano Energy, 2017, 39, 444 CrossRef CAS.
- K. Chen, L. Shi, Y. Zhang and Z. Liu, Chem. Soc. Rev., 2018, 47, 3018 RSC.
- S. H. Choi, S. J. Yun, Y. S. Won, C. S. Oh, S. M. Kim, K. K. Kim and Y. H. Lee, Nat. Commun., 2022, 13, 1484 CrossRef CAS PubMed.
- Y. Zhang, K. Nie, L. Yi, B. Li, Y. Yuan, Z. Liu and W. Huang, Advanced Science, 2023, 10, 2302301 Search PubMed.
- C. Ke, R. Shao, Y. Zhang, Z. Sun, S. Qi, H. Zhang, M. Li, Z. Chen, Y. Wang and B. Sa, Adv. Funct. Mater., 2022, 32, 2205635 CrossRef CAS.
- Y.-Y. Hsieh and H.-Y. Tuan, Energy Storage Mater., 2022, 51, 789 CrossRef.
- T. Li, X. Hu, C. Yang, L. Han and K. Tao, Dalton Trans., 2023, 52, 16640 RSC.
- C.-A. Chen, C.-L. Lee, P.-K. Yang, D.-S. Tsai and C.-P. Lee, Catalysts, 2021, 11, 151 CrossRef CAS.
- Y.-h. Luo, Q.-l. Pan, H.-x. Wei, Y.-d. Huang, L.-b. Tang, Z.-y. Wang, C. Yan, J. Mao, K.-h. Dai and Q. Wu, Mater. Today, 2023, 69, 54 CrossRef CAS.
- D. K. Bhat, H. Bantawal and U. S. Shenoy, Diamond Relat. Mater., 2023, 139, 110312 CrossRef CAS.
- H. Bai, D. Chen, Q. Ma, R. Qin, H. Xu, Y. Zhao, J. Chen and S. Mu, Electrochem. Energy Rev., 2022, 5, 24 CrossRef CAS.
- L. Yue, J. Liang, Z. Wu, B. Zhong, Y. Luo, Q. Liu, T. Li, Q. Kong, Y. Liu and A. M. Asiri, J. Mater. Chem. A, 2021, 9, 11879 RSC.
- H. Sun, X. Chu, Y. Zhu, B. Wang, G. Wang and J. Bai, J. Electroanal. Chem., 2023, 932, 117219 CrossRef CAS.
- X. Xu, Y. Pan, L. Ge, Y. Chen, X. Mao, D. Guan, M. Li, Y. Zhong, Z. Hu and V. K. Peterson, Small, 2021, 17, 2101573 CrossRef CAS PubMed.
- S. Sanati, R. Abazari, J. Albero, A. Morsali, H. García, Z. Liang and R. Zou, Angew. Chem., Int. Ed., 2021, 60, 11048 CrossRef CAS PubMed.
- J. Li, Z. Xiong, Y. Wu, H. Li, X. Liu, H. Peng, Y. Zheng, Q. Zhang and Q. Liu, J. Energy Chem., 2022, 73, 513 CrossRef CAS.
- H. Yan, X. Xiao, C. Hu, X. Liu and Y. Song, Mol. Catal., 2023, 547, 113327 CrossRef CAS.
- P. Forouzandeh and S. C. Pillai, Mater. Today: Proc., 2021, 41, 498 CAS.
- X. Wang, H.-M. Kim, Y. Xiao and Y.-K. Sun, J. Mater. Chem. A, 2016, 4, 14915 RSC.
- Z.-Z. Liu, N. Yu, R.-Y. Fan, B. Dong and Z.-F. Yan, Nanoscale, 2024, 16, 1080 RSC.
- J. Luo, Y. Wang, Y. Mao, Y. Zhang, Y. Su, B. Zou, S. Chen, Q. Deng, Z. Zeng and J. Wang, Chem. Eng. J., 2022, 433, 133549 CrossRef CAS.
- J. Yu, X. Wu, Y. Zhong, G. Yang, M. Ni, W. Zhou and Z. Shao, Chem.–Eur. J., 2018, 24, 13800 CrossRef CAS PubMed.
- Y. Xu, R. Wang, Z. Liu, L. Gao, T. Jiao and Z. Liu, Green Energy Environ., 2022, 7, 829 CrossRef CAS.
- Z. Li, X. Dou, Y. Zhao and C. Wu, Inorg. Chem. Front., 2016, 3, 1021 RSC.
- J.-T. Ren, L. Chen, L. Wang, X.-L. Song, Q.-H. Kong and Z.-Y. Yuan, J. Mater. Chem. A, 2023, 11, 2899 RSC.
- R. Mohili, N. Hemanth, K. Lee and N. K. Chaudhari, MXene-Transition Metal Compound Sulfide and Phosphide Hetero-Nanostructures for Photoelectrochemical Water Splitting, Elsevier. 2023, pp. 129 Search PubMed.
- X. Chen, W. He, L.-X. Ding, S. Wang and H. Wang, Energy Environ. Sci., 2019, 12, 938 RSC.
- S. Zhang, Y. Si, B. Li, L. Yang, W. Dai and S. Luo, Small, 2021, 17, 2004980 CrossRef CAS PubMed.
- W. Li, Q. Song, M. Li, Y. Yuan, J. Zhang, N. Wang, Z. Yang, J. Huang, J. Lu and X. Li, Small Methods, 2021, 5, 2100444 CrossRef CAS PubMed.
- M. Iqbal, E. Elahi, A. Amin, G. Hussain and S. Aftab, Superlattices Microstruct., 2020, 137, 106350 CrossRef CAS.
- Y. Li, M. Chen, B. Liu, Y. Zhang, X. Liang and X. Xia, Adv. Energy Mater., 2020, 10, 2000927 CrossRef CAS.
- C. Meng, X. Chen, Y. Gao, Q. Zhao, D. Kong, M. Lin, X. Chen, Y. Li and Y. Zhou, Molecules, 2020, 25, 1136 CrossRef CAS PubMed.
- M. Crisci, F. Boll, S. Domenici, J. Gallego, B. Smarsly, M. Wang, F. Lamberti, A. Rubino and T. Gatti, Adv. Mater. Interfaces, 2025, 12, 2400621 Search PubMed.
- J. Li, X. Yang, Z. Zhang, W. Yang, X. Duan and X. Duan, Nat. Mater., 2024, 23, 1326 CrossRef CAS PubMed.
- P. Ranadive, Scalable Continuous Synthesis of Metal and Metal-Oxide Based Nanomaterials through Jet-Mixing, The Ohio State University, 2021 Search PubMed.
- J. Kwon, S. Ko, H. Kim, H. J. Park, C. W. Lee and J. Yeo, Mater. Chem. Front., 2024, 2322 RSC.
- Y. Chen, F. Fang and N. Zhang, npj 2D Mater. Appl., 2024, 8, 17 Search PubMed.
- S. Lei, X. Zhao, X. Yu, A. Hu, S. Vukelic, M. B. Jun, H.-E. Joe, Y. L. Yao and Y. C. Shin, J. Manuf. Sci. Eng., 2020, 142, 031005 CrossRef.
- E. H. Hill, J. Mater. Chem. C, 2024, 12(30), 11285–11318 RSC.
- K. Nasrin, V. Sudharshan, K. Subramani and M. Sathish, Adv. Funct. Mater., 2022, 32, 2110267 Search PubMed.
- L. Wu, Y. Li, Z. Fu and B.-L. Su, Natl. Sci. Rev., 2020, 7, 1667 CrossRef CAS PubMed.
- N. H. Solangi, A. Abbas, N. M. Mubarak, R. R. Karri, S. H. Aleithan, J. Kazmi, W. Ahmad and K. Khan, Mater. Today Sustain., 2024, 100896 Search PubMed.
- J. Choi, K. Morey, A. Kumar, D. Neupane, S. R. Mishra, F. Perez and R. K. Gupta, Mater. Today Chem., 2022, 24, 100848 CrossRef CAS.
- H. Dai, L. Wang, Y. Zhao, J. Xue, R. Zhou, C. Yu, J. An, J. Zhou, Q. Chen, G. Sun and W. Huang, Research, 2021, 2021, 5130420 CrossRef CAS PubMed.
- Y. Yang, B. Sun, Z. Sun, J. Xue, J. He, Z. Wang, K. Sun, Z. Sun, H. K. Liu and S. X. Dou, Coord. Chem. Rev., 2024, 510, 215836 CrossRef CAS.
- R. Maleki, M. Asadnia and A. Razmjou, Adv. Intell. Syst., 2022, 4, 2200073 CrossRef.
- D. Y. Kirsanova, M. Soldatov, Z. Gadzhimagomedova, D. Pashkov, A. Chernov, M. Butakova and A. Soldatov, J. Surf. Invest.: X-Ray, Synchrotron Neutron Tech., 2021, 15, 485 CrossRef CAS.
- K. Choudhary, K. F. Garrity, S. T. Hartman, G. Pilania and F. Tavazza, Phys. Rev. Mater., 2023, 7, 014009 CrossRef CAS.
- M. A. F. Afzal and J. Hachmann, High-throughput computational studies in catalysis and materials research, and their impact on rational design, World Sci., 2020, 1 Search PubMed.
- T. Zheng, Z. Huang, H. Ge, P. Hu, X. Fan and B. Jia, Energy Storage Mater., 2024, 103614 CrossRef.
- B. Sridharan, M. Goel and U. D. Priyakumar, Chem. Commun., 2022, 58, 5316 RSC.
- Y. Yang, K. Yao, M. P. Repasky, K. Leswing, R. Abel, B. K. Shoichet and S. V. Jerome, J. Chem. Theory Comput., 2021, 17, 7106 Search PubMed.
- K. Shahzad, A. I. Mardare and A. W. Hassel, Sci. Technol. Adv. Mater.:Methods, 2024, 4, 2292486 Search PubMed.
- F. Qayyum, D.-H. Kim, S.-J. Bong, S.-Y. Chi and Y.-H. Choi, Materials, 2022, 15, 1428 Search PubMed.
- C. Xiouras, F. Cameli, G. L. Quilló, M. E. Kavousanakis, D. G. Vlachos and G. D. Stefanidis, Chem. Rev., 2022, 122, 13006 CrossRef CAS PubMed.
- A. Dhakal, C. McKay, J. J. Tanner and J. Cheng, Briefings Bioinf., 2022, 23, bbab476 CrossRef PubMed.
- J. Li, K. Lim, H. Yang, Z. Ren, S. Raghavan, P.-Y. Chen, T. Buonassisi and X. Wang, Matter, 2020, 3, 393 CrossRef.
- X. Liu, K. Fan, X. Huang, J. Ge, Y. Liu and H. Kang, Chem. Eng. J., 2024, 151625 CrossRef CAS.
- H. Xu, Z. Li, Z. Zhang, S. Liu, S. Shen and Y. Guo, Nanomaterials, 2023, 13, 1352 CrossRef CAS PubMed.
- M. Karthikeyan, D. M. Mahapatra, A. S. A. Razak, A. A. Abahussain, B. Ethiraj and L. Singh, Catal. Rev., 2024, 66, 997 CrossRef CAS.
- S. Iravani, A. Khosravi, E. N. Zare, R. S. Varma, A. Zarrabi and P. Makvandi, RSC Adv., 2024, 14, 36835 RSC.
- T. Ha, D. Lee, Y. Kwon, M. S. Park, S. Lee, J. Jang, B. Choi, H. Jeon, J. Kim and H. Choi, Sci. Adv., 2023, 9, eadj0461 CrossRef CAS PubMed.
- X. Zou, J. Pan, Z. Sun, B. Wang, Z. Jin, G. Xu and F. Yan, Energy Environ. Sci., 2021, 14, 3965 Search PubMed.
- H. Adamu, S. I. Abba, P. B. Anyin, Y. Sani and M. Qamar, Energy Adv., 2023, 2, 615 RSC.
- W. Qiu, Y. Wang and J. Liu, Wiley Interdiscip. Rev.: Comput. Mol. Sci., 2022, 12, e1592 CAS.
- G. Hyun, Y. Ham, J. Harding and S. Jeon, Energy Storage Mater., 2024, 69, 103395 CrossRef.
- Y. Elbaz, D. Furman and M. Caspary Toroker, Adv. Funct. Mater., 2020, 30, 1900778 CrossRef CAS.
|
This journal is © The Royal Society of Chemistry 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.