Zinc-based materials for electrocatalytic reduction reactions: progress and prospects

Baghendra Singh * and Apparao Draksharapu *
Southern Laboratories – 208A, Department of Chemistry, Indian Institute of Technology Kanpur, Kanpur-208016, India. E-mail: baghendras@iitk.ac.in; appud@iitk.ac.in

Received 12th May 2025 , Accepted 13th June 2025

First published on 16th June 2025


Abstract

The persistent energy crisis and environmental pollution pose significant challenges for modern society. Developing efficient methods for electrochemical energy conversion presents a promising solution to address these pressing issues. In the past few years, various electrocatalytic reduction reactions such as the hydrogen evolution reaction (HER), oxygen reduction reaction (ORR), nitrogen reduction reaction (NRR), nitrate reduction reaction (NO3RR), and carbon dioxide reduction reaction (CO2RR) have been investigated to create a pollution free green society and environment. Zn-based materials have garnered significant attention as potential candidates in the electrocatalytic reduction reactions owing to their precisely tuned structural and electronic properties, three-dimensional architectures, large surface areas, abundant active sites, high stability, and enhanced mass transport and diffusion capabilities. Numerous studies have been published investigating the potential of Zn-based materials in various electrocatalytic reduction reactions. However, there is a lack of comprehensive reviews systematically exploring the use of Zn-based materials in electrocatalytic reduction reactions. This review explores the structure–property–performance correlations of zinc-based catalysts, emphasizing their role in various electrocatalytic reduction reactions. We discuss the influence of structural modifications, such as doping, alloying, heterostructure formation, and morphological control, on the catalytic activity, stability, and selectivity of these materials. Special focus is given to the electronic structure modulation, active site optimization, and charge transfer mechanisms that underpin their performance. Recent advancements in synthesis techniques and characterization methods are highlighted to illustrate how tailored design strategies enhance catalytic efficiency. By presenting a comprehensive overview of zinc-based catalysts, this review aims to provide insights into their structure–performance relationships and offer guidance for the rational design of next-generation electrocatalysts for sustainable energy and chemical production.


image file: d5qm00354g-p1.tif

Baghendra Singh

Dr. Baghendra Singh earned his PhD in 2023 from the Indian Institute of Technology (BHU) Varanasi, under the guidance of Dr. Arindam Indra. He is currently serving as a National Post-Doctoral Fellow at the Department of Chemistry at IIT Kanpur, working under the mentorship of Dr. Apparao Draksharapu. Over the course of his academic journey, he has been the recipient of several prestigious fellowships awarded by the Government of India, including the INSPIRE Fellowship and the NPDF. In recognition of his scientific contributions, Dr. Singh was selected by the Japan Society for the Promotion of Science (JSPS) to participate in the 14th HOPE Meeting with Nobel Laureates, representing India among a selected group of young researchers. His ongoing research focuses on the development of advanced electrocatalysts for water splitting, hybrid water electrolysis, and energy storage applications.

image file: d5qm00354g-p2.tif

Apparao Draksharapu

Dr. Apparao Draksharapu obtained his master's degree from the University of Hyderabad, India, where he conducted research on emulsion polymerization. He earned his PhD in 2013 from the University of Groningen, Netherlands, under the supervision of Prof. Wesley R. Browne, focusing on the electrochemical and photochemical properties of inorganic metal complexes with relevance to biological systems. Following his doctoral studies, he pursued postdoctoral research at the University of Minnesota, USA, working with Prof. Lawrence Que, Jr., on bioinspired iron complexes. Currently, Dr. Draksharapu is an Assistant Professor at the Department of Chemistry at IIT Kanpur. His research is primarily directed toward the development of high-valent 3d-metal transient species for catalytic applications and energy conversion processes.


1. Introduction

Electrocatalytic reduction reactions play a pivotal role in energy conversion technologies, critical for transitioning towards sustainable and renewable energy systems.1–8 These reactions are essential in processes such as hydrogen production, fuel cells, carbon capture, and the synthesis of valuable chemicals, all of which contribute to the broader goals of energy efficiency, decarbonization, and environmental sustainability. The electrocatalytic reduction reactions mainly involve the hydrogen evolution reaction (HER), oxygen reduction reaction (ORR), nitrogen reduction reaction (NRR), nitrate reduction reaction (NO3RR), and carbon dioxide reduction reaction (CO2RR).1–4,9 The HER is essential for creating hydrogen gas (H2), a clean fuel that may be utilized in combustion engines and fuel cells, among other energy applications.9 Hydrogen is often referred to as a “green” energy carrier when produced through electrocatalytic water splitting using electricity. This process offers a sustainable alternative to fossil fuels. The ORR is the central process for the operation of fuel cells and metal–air batteries.6,7 These devices are key components of energy conversion systems. The ORR, which converts chemical energy into electricity with high efficiency and minimal emissions, takes place at the cathode in solid oxide and proton exchange membrane fuel cells. In zinc–air batteries (ZABs), lithium–air batteries (LiABs), aluminum air batteries (AlABs), and sodium air batteries (NaABs), the ORR is responsible for the discharge process.6,7 These batteries are attractive for storing energy at a practical level because of their large energy density and potential for grid-scale renewable energy storage.

The CO2RR is essential for mitigating climate change through carbon capture and utilization (CCU).2 The CO2RR aims to convert CO2, a greenhouse gas, into value-added products and fuels such as carbon monoxide (CO), methane (CH4), or formic acid (HCOOH). By converting CO2 into useful products, this reaction helps address the growing concern of global warming and contributes to climate mitigation strategies.2 The NRR allows for the production of ammonia (NH3) electrochemically, which is essential for fertilizers and other industrial processes.1,3,4 Traditional ammonia production, via the Haber–Bosch process, is energy-intensive and contributes significantly to the world's CO2 emissions. The electrocatalytic NRR offers a sustainable path for ammonia production. The NRR can be driven by renewable electricity, drastically reducing the energy input and greenhouse gas emissions associated with ammonia synthesis.3,4 The NO3RR is vital for environmental remediation, particularly in reducing nitrate contamination in water sources, a significant environmental and public health issue.5 Nitrate pollution, commonly caused by agricultural runoff and industrial processes, contaminates water. The NO3RR can convert harmful nitrates into benign products like nitrogen gas (N2) or ammonia (NH3).5 Electrocatalytic reduction reactions are key in transitioning away from fossil fuels by enabling renewable and sustainable alternatives for energy production, chemical synthesis, and storage.1,2,4,6

However, an active electrocatalyst is needed to drive these electrocatalytic reduction reactions efficiently. Researchers have explored a wide range of transition and noble metal-based materials for electrocatalytic reduction reactions in this context.10–20 Over the past few years, zinc-based materials have garnered much interest in electrocatalytic reduction processes because of their readily available, inexpensive, environmentally friendly, and customizable structural and electrical characteristics. These materials, whether in the form of zinc-based oxides, sulfides, phosphides, metal–organic frameworks (MOFs), or alloys, show promise across several key electrocatalytic reduction reactions (HER, ORR, NRR, NO3RR, and CO2RR) that are critical for energy conversion.10–20 In all these reduction reactions, Zn-based materials have shown excellent performance.21–30 Zinc-based materials have gained significant attention in electrocatalysis due to their unique properties, cost-effectiveness, and versatility. These materials play a crucial role in various reduction reactions, including the CO2 reduction reaction, NRR, and ORR, contributing to sustainable energy and chemical production.

Zinc is widely available, reducing reliance on expensive and scarce noble metals such as platinum or palladium.21–28,30–37 Zinc-based catalysts provide an economically viable alternative for industrial applications, especially in large-scale energy and chemical production. Zinc-based materials include oxides, phosphides, sulfides, MOFs, and alloys, each tailored for specific reactions.21–28,30–37 Zinc materials readily accommodate structural defects, which improve adsorption sites and reaction kinetics. Zinc-based catalysts can be functionalized with other elements or compounds to enhance activity, stability, and selectivity. Combining Zn with conductive materials (e.g., carbon nanotubes or graphene) improves charge transfer and catalytic efficiency. Incorporation of transition metals like Fe, Co, or Ni into Zn-based structures fine-tunes electronic properties and catalytic behavior. Zinc-based materials are indispensable for advancing electrocatalytic reduction reactions attributed to their unique structural and electronic properties.21–28,30–37 Their abundance, tunability, and compatibility with green technologies make them a cornerstone for sustainable energy and chemical production.

However, a fundamental understanding of the structure–property–performance correlation in zinc-based materials is essential for their rational design and optimization. The intrinsic properties of zinc, such as its electronic structure, coordination environment, and redox behavior, significantly influence its catalytic activity and stability. Additionally, structural modifications, including doping, alloying, heterostructure formation, and nanoscale engineering, have been employed to fine-tune these materials for enhanced performance.21–28,30–37 These modifications can alter active site availability, improve charge transfer, and modulate binding energies of reaction intermediates, all of which are critical for catalytic efficiency. While Zn-based materials have been extensively studied for various electrocatalytic reduction reactions, much of the research has focused on their structural, morphological, and electronic tuning.21–28,30–37 The challenges remain in achieving a comprehensive understanding of the mechanisms by which zinc-based materials catalyze reduction reactions and in translating their laboratory-scale performance to practical applications. Recent progress in synthesis techniques, advanced characterization methods, and computational modeling has provided valuable insights into the interplay between material structure and catalytic behavior.

However, a thorough review that systematically addresses their design, properties, applications, and challenges in electrocatalytic reduction reactions is still missing. This review aims to fill that gap by focusing on Zn-based materials and critically analyzing their design, properties, and applications in water-splitting processes (Fig. 1). It highlights various Zn-based materials, such as oxides, phosphides, chalcogenides, alloys, and MOFs, with a special focus on their structure, morphology, and electronic properties. This overview aims to thoroughly grasp the properties and uses of various materials through important examples. This review also discusses recent advancements, persistent challenges, and the intricate relationship between structure, activity, and stability while outlining future research directions.


image file: d5qm00354g-f1.tif
Fig. 1 Pictorial depiction showing the various objectives of this review.

This review aims to provide a comprehensive overview of zinc-based materials for electrocatalytic reduction reactions, focusing on their structure–property–performance relationships. We discuss the impact of various structural modifications on catalytic activity, selectivity, and durability, highlighting key advancements in this field. By synthesizing knowledge from experimental and theoretical studies, this review seeks to guide the development of next-generation zinc-based electrocatalysts for sustainable energy and chemical production.

2. Why should zinc be chosen for material design?

Materials based on zinc have become more significant in electrocatalytic reduction reactions due to their unique properties, including abundant availability, low cost, environmental friendliness, and versatile electrochemical behavior.21–28,30–37 They serve as promising candidates in various reduction processes. Zn is an abundant and relatively inexpensive metal compared to other metals.38,39 Its widespread availability makes it an economically viable choice for large-scale applications in designing Zn-based materials for electrocatalytic reduction reactions, particularly for sustainable energy technologies.21–28,30–37 The leaching of zinc by evaporation at high temperatures during the synthesis of catalytic materials can significantly influence the final material's structure, morphology, and catalytic properties.40–42

During synthesis, zinc leaching is often a controlled process and can result in beneficial modifications to the material's structure.40–42 Zinc leaching can create porous structures within the catalyst. When removed during synthesis, zinc leaves voids or pores that can expand the material's surface area, giving catalytic processes additional active sites. The development of metal oxide/hydroxide phases due to leaching might enhance catalytic activity. For example, in zinc–cobalt materials, the leaching of zinc during synthesis can expose more active cobalt sites.40,41 Leaching of zinc can alter the electronic properties of the material, optimizing its catalytic behavior.

The charge distribution, band gap, and electron density in the catalyst can be tuned, improving its ability to adsorb and activate reactants. Zinc can be combined with other metals to create bimetallic or multimetallic catalysts that exhibit synergistic effects.21–28,30–37 When Zn is introduced with other metals, it provides the available binding sites for the process intermediates by changing the electronic states of the different metals.21–28,30–37 Overall, zinc is a versatile element used in the design of various catalytic materials due to its abundance, cost-effectiveness, tunable electronic properties, and ability to form stable compounds.21–28,30–37

3. Properties of Zn-based materials

Owing to its unique properties, Zn has been employed to design a series of various materials such as oxides, phosphides, and chalcogenides for electrocatalytic reduction reactions.21–28,30–37 Zn-based materials offer altered electronic structure, large surface area, tunable properties, and desirable morphology (Fig. 2). In most materials, Zn can be evaporated during pyrolysis causing a porous structure. The porosity provides exposed active sites improving the adsorption and release of reactive intermediates and products, respectively.21–28,30–37 Intrinsic and extrinsic defects in Zn-based materials, such as oxygen vacancies in ZnO or Zn-doping in other materials, enhance catalytic activity by modulating electronic properties and creating more reactive sites. Zinc-based materials can be tailored into various structural forms, such as nanorods, nanowires, nanosheets, and nanoparticles. These designs enhance catalytic performance by increasing surface area and exposing more active sites.21–28,30–37 Porous zinc-based materials, in particular, improve mass transport and provide abundant reactive sites, benefiting catalytic processes. Zinc-based phosphides, on the other hand, exhibit high electrical conductivity and excellent water reduction activity. Moreover, Zn-based metal–organic frameworks (MOFs) are valued for their high porosity, customizable structures, and potential as precursors for derivative materials.21–28,30–37 Zinc-based materials, especially oxides and hydroxides, often have poor electrical conductivity, which hampers charge transfer and reduces overall catalytic efficiency. The need for conductive substrates or additional conductive additives complicates material design and increases production costs. Zinc-based materials can suffer from structural instability under harsh reaction conditions, including acidic or alkaline media. Zinc tends to dissolve or undergo phase changes, particularly during prolonged operation, leading to catalyst degradation and reduced durability.21–28,30–37 Many zinc-based catalysts have limited active sites for catalytic reactions. This results in low turnover frequencies (TOF) and reduced overall catalytic activity. Achieving high selectivity for desired products (e.g., specific reduction products in CO2 or N2 reduction) can be challenging due to competing side reactions, such as hydrogen evolution.21–28,30–37 Tailoring zinc-based materials for specific reaction pathways remains a key challenge.
image file: d5qm00354g-f2.tif
Fig. 2 Schematic illustration depicting the properties of Zn-based materials for electrocatalytic reduction reactions.

Incorporation of conductive supports (e.g., carbon-based materials) in Zn-based materials can improve conductivity. Doping with other metals or non-metals enhances stability and active site density.21–28,30–37 Designing hybrid structures or composites mitigates structural instability and poisoning effects. In-depth mechanistic studies and computational modeling are very helpful in rational catalyst design. Compared to other emerging electrocatalysts such as transition metal chalcogenides, single-atom catalysts (SACs), and metal–nitrogen–carbon (M–N–C) systems, Zn-based materials exhibit several unique advantages and trade-offs. Zn's moderate binding energy with key intermediates makes it particularly selective for reactions like the CO2RR to CO and NRR, while its low hydrogen evolution activity enhances faradaic efficiency. However, Zn-based catalysts generally suffer from lower intrinsic conductivity and limited multi-electron product selectivity. In contrast, transition metal chalcogenides offer higher electrical conductivity and tunable layered structures that are beneficial for the HER and the ORR but often require stabilization strategies under long-term operation. Single-atom catalysts, including Zn–N4 moieties, demonstrate high atom utilization and excellent selectivity, though their synthesis and scalability remain challenging. Meanwhile, M–N–C systems, particularly those based on Fe, Co, or Ni, have demonstrated superior activity and durability in the ORR and CO2RR, attributed to their rich active sites and conductive carbon matrix. Overall, while Zn-based materials show promising selectivity and earth abundance, further advancements in structure engineering and electronic modulation are required to match the performance benchmarks set by more mature catalyst families like SACs and M–N–C materials.

4. Design of Zn-based materials

The design of Zn-based materials for electrochemical reduction reactions involves careful selection of composition, morphology, and structure.

4.1. Approaches of design

Zn-based materials are designed based on several structural, electronic, and morphological approaches so that one can get superior electrocatalytic performance (Fig. 3).
image file: d5qm00354g-f3.tif
Fig. 3 Schematic illustration depicting the designing of Zn-based materials using various approaches.
4.1.1. Structural design approach. It is well established that Zn-based MOFs offer high porosity, tunable structure, and abundant active sites. Therefore, bimetallic MOFs are designed by introducing transition metals (e.g., Zn–Co, Zn–Fe, and Zn–Ni) to improve catalytic activity and stability.21–28 Furthermore, Zn-based MOF-derived materials are synthesized from Zn-MOFs. Doping with heteroatoms (N, S, and P) modifies their electronic structure, introducing localized charge density variations and improving electron transfer kinetics. Nitrogen, sulfur, and phosphorus can donate or withdraw electrons, creating electroactive sites and enhancing electrical conductivity to improve the reactivity. Zn-based oxides are prepared using air calcination to create porous and hollow structures improving the surface adsorption properties.21–28 Porous and hollow structures provide a high surface area-to-volume ratio, exposing more active sites for reactant interaction. These architectures also allow efficient mass transport of ions and gases through the catalyst, minimizing diffusion limitations and improving overall catalytic efficiency. Templating and controlled synthesis of Zn-based hydroxides leads to the formation of layered hydroxides with enhanced electrochemical properties.21–28 Tailored interlayer spacing improves ion accessibility and diffusion. The layered structures also offer more exposed basal planes and edge sites, which are known to be catalytically active.

The formation of heterostructures by integrating Zn-based materials with other materials or conductive substances is also adopted to get excellent catalytic activity and stability. Combining Zn-based materials with conductive or catalytically active phases creates synergistic interfaces that facilitate charge separation, improved conductivity, and optimized adsorption energies of intermediates. These interfacial effects often lower energy barriers and enhance reaction kinetics. Surface engineering such as coating with carbon layers and defect engineering such as vacancy creation are the main approaches utilized to design Zn-based materials. Surface engineering provides unsaturated coordination sites, which often act as active centers by increasing the binding affinity for reactants or intermediates. These defects also modulate the local electronic environment, facilitating charge redistribution and lowering reaction energy barriers.

4.1.2. Morphological design approach. Zn-based materials are designed with varying morphology including 0D, 1D, 2D, and 3D architectures. The morphological tuning offers increased active surface area by designing hierarchical structures.21–28 The core–shell nanostructures are also designed with shell modifications to create a high surface area. As already mentioned, porous morphology provides excellent surface adsorption properties and exposed active sites; designing Zn-based materials with a highly porous nature is adopted to ensure efficient mass diffusion in catalytic processes.30–37
4.1.3. Electronic structure design approach. During the synthesis of Zn-based materials, one should take care of the modulation of electronic structure.30–37 The Zn-based materials with altered electronic features provide excellent catalytic performance. Doping or substitution of the other metals with Zn atoms alters the local electronic environment and redistributes electron density. Doping can modulate the d-band center relative to the Fermi level. A closer d-band center (toward the Fermi level) typically enhances adsorption strength for intermediates, which can be beneficial or detrimental depending on the reaction. The introduction of Zn2+ modifies the local charge density, affecting the electrostatic potential at active sites and thus reactivity.30–37 Doping can introduce new energy states in the bandgap or decrease the bandgap altogether, leading to enhanced conductivity, which is essential for electron transport in electrocatalysis. Introducing vacancies disrupts the periodicity of the lattice, altering local coordination environments and electronic states. Oxygen vacancies create shallow donor levels within the bandgap, facilitating charge carrier generation and improving electronic conductivity.30–37 Vacancies cause uneven electron distribution, resulting in localized charge polarization that can serve as active centers for adsorption and activation of small molecules. The presence of defects changes the symmetry and occupancy of d-orbitals, potentially shifting the d-band center and modifying the binding strength with intermediates. Introducing non-metal heteroatoms into Zn-based catalysts alters the electronegativity landscape, introducing asymmetry in electronic distribution. N, S, or P atoms have different electronegativities and introduce localized dipoles, which affect charge delocalization and the electronic affinity of neighboring Zn centers. The Zn–N or Zn–S coordination creates localized electronic states that tune adsorption energies of reaction intermediates.30–37 Heteroatoms can conjugate with carbon supports, forming sp2-hybridized networks with enhanced π-electron delocalization, leading to better electron mobility. Alloying Zn with transition metals (e.g., Cu, Co, and Ni) combines two or more metals at the atomic level, creating new electronic states through orbital hybridization. Alloying introduces the lattice strain effect that alters the crystal field and electronic orbital overlap, which can either enhance or suppress catalytic activity. Adjacent metal atoms may participate in dual-site catalysis, enabling multi-step electron transfer reactions or stabilizing key intermediates through ensemble effects.30–37

4.2. Methods of design

The synthesis of zinc-based materials can be achieved through a variety of methods, each designed to regulate the morphology, structure, and properties of the desired materials. These techniques are tailored to optimize the performance of zinc-based materials in electrocatalytic reduction reactions.
4.2.1. Hydrothermal/solvothermal methods. The hydrothermal synthesis of Zn-based materials involves heating of metal precursors in an autoclave with different reagents like sulfur, phosphorus, selenium, and others. Various solvents, such as water and ethanol, have been employed to facilitate the synthesis of Zn-based materials. Hydrothermal processes usually take place in autoclaves at high temperatures and pressures (Fig. 4).43,44 This method promotes the growth of zinc-based materials with well-defined morphologies, such as nanorods, nanowires, and nanosheets.43,44 Solvothermal synthesis is similar to hydrothermal synthesis but uses non-aqueous solvents. It is beneficial for preparing zinc chalcogenides and other complex zinc-based materials. In this regard, ZnxCo1−xSe2 was developed using the ZnCo-MOF as the precursor and the hydrothermal method.45 The ZnxCo1−xSe2 catalyst was created by heating the mixture to 200 °C in an autoclave and keeping it there for 12 h. Similarly, ZnInS was also created using the hydrothermal method (120 °C for 4 h).46
image file: d5qm00354g-f4.tif
Fig. 4 Pictorial depiction showing the hydrothermal synthesis of Zn-based materials.

In a different work, Nam and coworkers prepared ZnS using a simple hydrothermal method following a specific procedure.47 Initially, a solution of 5 mM Zn(NO3)2·6H2O was prepared in water following the addition of 5 mmol of S-source such as C2H5NS or CH4N2S. After being cleaned, Zn-foil was heated in an autoclave in a convection oven for 6 h at 120 °C. Following the hydrothermal process, the autoclave was left to cool naturally to ambient temperature, resulting in ZnS. In further works, Ti-ZnCoS was also fabricated using a hydrothermal technique.48

4.2.2. Chemical vapor deposition (CVD) method. Zn-based compounds may be effectively synthesized via chemical vapor deposition (CVD).49–51 Zn-based precursors are treated at high temperatures in an inert environment in a furnace, frequently with reagents such as phosphorus, nitrogen, sulfur, or selenium to form Zn-based chalcogenides, phosphides, or nitrides (Fig. 5). It should be noted that the high-temperature heating of the Zn-precursor leads to the evaporation of zinc, creating a highly porous structure.49–51 Therefore, the synthesis temperature can be carefully regulated based on the desired materials and their specific properties. To fine-tune attributes for catalytic applications, temperature control is essential in establishing the final material's crystallinity, porosity, and chemical makeup.49–51 In this context, Dai et al. explored the ZnS@C material for electrocatalytic ORR reactivity.51 First, they prepared the Zn-precursor in water and ethanol. The Zn-precursor was then heated to 700 °C for two hours at a rate of 5 °C per minute while being exposed to an Ar flow. After cooling, the resulting sample was the ZnS@C catalyst. In another work, the ZnSe/NiSe heterostructured material was also developed employing the CVD technique.52
image file: d5qm00354g-f5.tif
Fig. 5 Pictorial depiction showing the CVD process for the synthesis of Zn-based materials.

Heating N,C-containing Zn-MOFs under inert conditions results in the formation of N-doped carbon frameworks embedded in the final material. For example, Lim and coworkers first prepared a ZnCo-ZIF and then utilized the CVD method for the sulfidation to produce ZnCoS/NCS.49 The ZnCo-ZIF was treated with thiourea in a tubular furnace under an inert atmosphere. The ZnCo-ZIF was pyrolyzed at 700 °C for 2 h with the rate of 10 °C min−1 under Ar-gas. After the reaction, ZnCoS integrated with N-, and S-doped carbon was synthesized. The N-doped carbon was in situ integrated with the sulfide by the pyrolysis of a MOF precursor having a nitrogenous ligand. Similarly, Hu and coworkers developed Cu–ZnSe@NC using a ZnCu-ZIF under CVD.50 In an N2 environment, the ZIF and Se powder were employed and reacted at 500 °C for four hours. The in situ synthesis under CVD resulted in Cu–ZnSe@NC.

4.2.3. Air calcination method. In a related process known as “air calcination,” Zn-precursors are exposed to an oxygen-rich environment, where they are heated to convert into their corresponding oxide forms.43,44 During this process, the precursors thermally decompose, forming highly porous zinc oxide materials. For example, Liang et al. demonstrated the C–ZnO material for hydrogen evolution.53 To produce C–ZnO, the ZIF was calcined at different temperatures, speeds, and durations in an air environment muffle furnace. The calcination procedure involved setting temperatures at 350, 400, 450, and 500 °C and calcining at a rate of 2 °C per minute for three hours. The obtained samples were given the corresponding designations ZnO-350, ZnO-400, ZnO-450, and ZnO-500. As an alternative, a two-step calcination method was used to create C–ZnO. The first phase was heating at 350 °C for two hours at a rate of 2 °C per minute and then heating at 400 °C and 450 °C for one hour at a rate of 1 °C per minute. The resulting samples were designated as ZnO-350–400 and ZnO-350–450, respectively. Similarly, Buckley and coworkers developed ZnO for electrocatalytic hydrogen evolution using air calcination.54 The Zn-precursor was calcined in the air within a muffle furnace for 90 minutes at 600 °C with a 5 °C min−1 rate.

Although the above-mentioned three methods are used for the synthesis of Zn-based materials, there are pros and cons of each method (Table 1).43,44,49–51 Hydrothermal synthesis operates at relatively moderate temperatures and pressures and enables the synthesis of a wide range of structures (nanorods, nanowires, etc.) by adjusting reaction parameters. The hydrothermal technique produces materials with high phase purity.43,44,49–51 The hydrothermal method is suitable for large-scale synthesis with appropriate reactor designs and also allows uniform doping and incorporation of heteroatoms. Besides these merits, the hydrothermal method also has demerits. The hydrothermal method often requires hours to days to complete the reaction and requires specialized high-pressure vessels (autoclaves). This method is not ideal for synthesizing materials that require extreme temperatures or vapor-phase reactions.43,44,49–51

Table 1 Summary of the advantages and disadvantages of various synthesis methods of Zn-based materials
Sr. no. Synthesis method Advantages Disadvantages
1 Hydrothermal or solvothermal method Need for moderate temperature, adjustable reaction parameters, controlled structure design, easier synthesis of various materials Requires high-pressure autoclaves, the danger of exploding
2 Chemical vapor deposition (CVD) method Produces highly crystalline materials, precise control over temperature and gas flow, forms highly porous materials with rough surfaces Requires high temperature and expensive equipment, needs input of gases like N2 and Ar, releases toxic gases like phosphine
3 Air calcination method Need for minimum equipment, no external gas flow, open-air heating, produces highly porous and crystalline materials Requires high temperature, poor control over structure and morphology of the designed materials


The CVD method is suitable for a wide range of materials, including semiconductors and composites, and produces uniform and high-purity thin films or coatings.43,44,49–51 The CVD technique enables precise control over material thickness, composition, and nanostructure. In contrast to these merits, CVD requires expensive equipment and precursors. CVD often operates at elevated temperatures and requires precise control of gas flow rates and temperature. CVD uses toxic, flammable, or hazardous gases in some cases.43,44,49–51 Air-calcination is a straightforward process requiring minimal equipment. This method is suitable for achieving phase transitions or crystallinity in materials and is used for preparing oxides, phosphates, and other stable materials.43,44,49–51 The air calcination method has poor control over nanostructures and surface area and requires high temperatures, leading to significant energy consumption. Each technique has its unique advantages and limitations, making it suitable for specific applications based on the desired material properties and end-use requirements.43,44,49–51

5. Hydrogen evolution reaction (HER)

The electrochemical hydrogen evolution reaction involving the reduction of water (in alkaline) or protons (in acidic) into molecular hydrogen by a two-electron transfer process.55,56 The reaction mechanism of the HER includes the Volmer Tafel and Volmer Heyrovsky pathways (Fig. 6).43,44,57 In most noble metal-based materials, the catalysts reduce water by the Volmer Tafel mechanism. On the other hand, materials based on transition metals reduce water using the Volmer Heyrovsky route (Fig. 6).43,44,57 In recent years, most investigations have focused on designing proficient catalysts in which zinc-based materials have been thoroughly explored for the electrocatalytic HER.
image file: d5qm00354g-f6.tif
Fig. 6 Reaction mechanism of the electrocatalytic water reduction reaction in alkaline and acidic electrolytes. Reproduced from ref. 56 with permission from Wiley-VCH.

Mirabella and coworkers synthesized a ZnOHF-ZnO catalyst using chemical bath deposition.58 After casting onto graphene paper, this catalyst was assessed as a potential material for a full water-splitting cell. At a current density of 10 mA cm−2, it displayed an overpotential (η10 = overpotential required at 10 mA cm−2 current density) of 181 mV. Upon Pt decoration, the catalyst achieved a cell voltage (E10 = cell voltage required at 10 mA cm−2 current density) of 1.7 V at 10 mA cm−2 for overall water splitting. Kumar et al. explored a heterostructured CuO/ZnO catalyst, finding that adding ZnO significantly boosted CuO's reactivity.59 For the OER, CuO/ZnO exhibited an onset potential of 1.40 V, lower than CuO's 1.74 V. For the HER, CuO/ZnO showed an onset voltage of 0.15 V, while CuO alone reached 0.24 V. Furthermore, Kim and coworkers developed a ternary oxide consisting of Gd2O3, In2O3, and ZnO, which displayed a HER η10 of 271 mV, indicating promising catalytic potential.60

In another study, Zn-doped CoP was synthesized to enhance water-splitting performance.61 This catalyst was prepared on cobalt foam through a two-step hydrothermal process followed by phosphating (Fig. 7a). The unchanged PXRD peaks confirmed the successful incorporation of Zn into the lattice. Compared to CoP, the Zn–CoP's Co 2p spectrum showed a negative shift of approximately 1.72 eV, indicating that Zn doping influenced cobalt's electronic state, reducing the Co3+ amount (Fig. 7b). The P 2p spectrum displayed a negative shift of ∼0.15 eV compared to CoP, suggesting a strong electronic interaction between Zn and Co (Fig. 7c). The presence of Co3+ and Zn2+ in Zn–CoP further indicated electronic interactions between these metal ions.


image file: d5qm00354g-f7.tif
Fig. 7 (a) Pictorial depiction for the synthesis of Zn–CoP using the CVD method; (b) Co 2p XPS of CoP and Zn–CoP showing the negative shift in the peak of Co3+ in Zn–CoP; (c) P 2p XPS of CoP and Zn–CoP; (d) electronic interaction between Co 3d and Zn 3d orbitals; (e) electrocatalytic HER reactivity of Zn–CoP compared with CoP, Pt/C and bare CF manifesting the best HER activity of Zn–CoP; (f) CA stability plot of Zn–CoP for the HER; (g) free energy change profile for Zn–CoP and CoP; and (h) PDOS plot of the CoP and Zn–CoP showing the d-band center. Reproduced from ref. 61 with permission from the Royal Society of Chemistry; (i) LSV curves showing the improved HER activity of CoP after Zn-doping. Reproduced from ref. 62 with permission from Elsevier; (j) LSV profiles indicating the enhanced HER activity of P–Zn–CoNiP-2 among the studied catalysts. Reproduced from ref. 63 with permission from Wiley-VCH; and (k) catalytic HER activity of Zn–CoNiP compared with CoNiP showing the best activity of Zn–CoNiP. Reproduced from ref. 64 with permission from Elsevier.

The electronic configurations of Zn and Co were examined to clarify these interactions. Co3+, with a valence electron configuration of 3d6 (t42ge2g), had two unpaired electrons in the t2g orbitals, which interact with O2−via π-donation (Fig. 7d). In contrast, Zn2+, with a configuration of 3d10 (t2g6eg4), had fully occupied t2g orbitals, resulting in electron–electron repulsion with O2−. When Co and Zn are combined, the π-interaction between Co and O was strengthened by the repulsion between O2− and Zn2+, causing partial charge transfer. In Zn–CoP, this redistribution of valence electrons adjusted the binding energies of hydrogen and oxygen intermediates, aligning with the Sabatier principle. The Zn doping altered the electronic structure of CoP and improved the charge transport, significantly boosting the catalytic activity.61

The Zn–CoP catalyst exhibited an η100 (η100 = overpotential required at 100 mA cm−2 current density) of just 166 mV for hydrogen evolution, a notably low value compared to other materials studied (Fig. 7e). A water electrolyzer was assembled using Zn–CoP as both the cathode and anode, achieving an E100 (E100 = cell voltage required at 100 mA cm−2 current density) of 1.71 V and an E1000 (E1000 = cell voltage required at 1000 mA cm−2 current density) of 2.01 V. Tafel plot analysis indicated that Zn–CoP had the fastest reaction kinetics for the HER among the catalysts tested. Furthermore, the catalyst demonstrated excellent stability of 60 h for the HER during CA testing (Fig. 7f). The Gibbs free energy change profile determined through DFT suggested that the free energy change decreased after Zn-doping, which was closer to zero, suggesting the improved HER reactivity in the case of Zn–CoP (Fig. 7g). Furthermore, density of state (DOS) analysis revealed a higher density of states near the Fermi level, indicating the metallic characteristics of CoP (Fig. 5h).

In a separate study, a highly effective HER electrocatalyst was synthesized through a straightforward approach. The Zn–CoP catalyst featured a 3D hierarchical nanostructure, providing a large surface area and abundant catalytic sites to facilitate efficient mass transport of reactants and electrolytes.62 Its unique one-dimensional porous architecture further enhanced electrocatalytic activity by offering numerous active sites for the HER. Zn–CoP demonstrated impressive HER performance in alkaline media with an η10 of 138 mV (Fig. 7i). Cai et al. investigated Zn-doped CoNiP for overall water splitting, noting that Zn doping modified the electronic properties of Ni and Co sites.63 This catalyst achieved a HER η100 of 177 mV, with an E50 of 1.71 V for concurrent O2 and H2 production, respectively (Fig. 7j). Similarly, Wang and coworkers developed a Zn–CoNiP catalyst for overall water splitting, achieving remarkable HER activity with an η10 of 73 mV under acidic and 34 mV under alkaline conditions (Fig. 7k).64 The catalyst required only an E10 of 1.51 V for total water splitting. These above-mentioned studies showed that Zn doping altered the electronic features of Co and Ni sites improving the catalytic HER performance.

To enhance the catalytic performance of Zn-based phosphides, conductive materials have been incorporated to form N- and C-containing phosphides. In several studies, researchers directly integrate N-doped carbons, nanotubes, or graphene into the phosphide catalyst. In most cases, metal–organic frameworks (MOFs) serve as precursors for phosphide synthesis. MOFs with N- and C-containing linkers are phosphorized under high-temperature conditions, creating conductive N-doped carbon materials within the phosphide structure in situ. For Zn-based MOFs, the high-temperature phosphidation process causes zinc to evaporate, resulting in a highly porous structure.

Li et al. demonstrated the Co2P/CoP–NC and NiP/Co2P/CoP–NC catalysts using a Zn-based MOF precursor, where Zn evaporated during synthesis, forming the phosphide heterostructure.65 Additionally, NC was integrated into the heterostructured phosphide catalyst during this process. The NC support facilitated rapid charge transfer in the Co2P/CoP–NC catalyst, and its hierarchical structure provided a large surface area with numerous channels for mass transport, enhancing diffusion in electrochemical reactions.

Improved catalytic kinetics and overall activity were observed due to electronic interactions between the NC nanosheets and dual-phased phosphides, which enhanced charge transfer and lowered reaction barriers for intermediates. The Co2P/CoP hetero-interfaces further optimized electronic topologies, creating multiple pathways for charge movement. The Co2P/CoP–NC-0.1 catalyst exhibited excellent HER activity, with η10 values of 84 mV in acidic and 91 mV in alkaline electrolytes, while CoP–NC showed higher η10 values of 144 mV and 188 mV under the same conditions.

As shown theoretically and experimentally in the study, heterostructured catalysts provide superior catalytic performance compared to individual materials; forming heterostructures involves combining two or more distinct phases or materials to enhance reaction efficiency.66–70 These heterostructures create interfaces between materials that enable the effective movement of charge carriers, facilitating improved charge transfer between phases and enhancing overall catalytic activity.66–70 The materials in a heterostructure often possess complementary electronic properties, such as band structure or work function, which can reduce energy barriers, improve electronic conductivity, and tailor the electronic environment for better reactant interaction.66–70 Additionally, the combination of materials within a heterostructure increases the density of active sites, resulting in more efficient catalytic processes and higher reaction rates for the HER and the OER.

Zhang and coworkers conducted an in-depth study to enhance electrochemical HER activity by incorporating conductive polyaniline (PANI) into ZnS nanoparticles.71 PANI was integrated into ZnS in varied quantities (10, 50, 90, and 110 mg), yielding samples denoted as ZnS-10 mg, ZnS-50 mg, ZnS-90 mg, and ZnS-110 mg. Among these, ZnS-90 mg showed an optimal synergistic effect, achieving an η10 of 168 mV to enable the HER, significantly lower than the 244 mV overpotential required by unmodified ZnS. This comparative study established that PANI notably enhanced charge transfer properties by elevating conductivity, thereby improving HER activity. The mechanism behind PANI's effectiveness lies in its conductive nature, which facilitates electron mobility within the ZnS matrix, thus lowering charge transfer resistance and providing a more favorable electronic environment. This interaction underscores the potential of conductive polymers like PANI in boosting the catalytic properties of semiconductors, highlighting a promising pathway for enhancing efficiency in HER and other electrocatalytic applications.

An intriguing study conducted by Pan et al. focused on the design of CnZnS on copper foam for total water splitting.72 The catalyst was synthesized using a one-step hydrothermal method. SEM images of CoZnS showed the presence of radial nanoflocs and angular nanoparticles. TEM pictures revealed lattice spacings that corresponded to the planes of CoS. The CoZnS achieved an η10 of 130 mV for the HER, showcasing superior reactivity compared to CoS and ZnS. Additionally, CoZnS exhibited an E10 of 1.53 V, indicating its effectiveness in producing both O2 and H2 in a two-electrode system.

In a related study, Zn doping was applied to CoNiS to enhance its HER activity.73 The incorporation of Zn optimized the electronic properties of the catalyst, resulting in an E10 of 1.50 V in a two-electrode configuration. Building on this, researchers have also developed heterostructured Zn-based chalcogenides for water-splitting applications. In this context, Wei and coworkers created a heterostructured polyoxometalate (POM)/ZnCoS catalyst to improve overall water-splitting efficiency.74 The POM@ZnCoS nanowires were synthesized on a 3D spongy Ni substrate using a simple and scalable two-step hydrothermal method.

ZnCoS nanowires were initially synthesized through a reaction between ZnCo-hydroxide and Na2S during hydrothermal processes. Following this, PW12 nanoparticles were anchored onto the ZnCoS nanowires, forming POM@ZnCoS nanowires. SEM pictures of POM@ZnCoS revealed a morphology similar to that of ZnCoS but with noticeable differences. The nanowires appeared thicker, rougher, and shorter, exhibiting increased roughness compared to the original ZnCoS. This significant alteration in shape is likely due to the attachment of PW12 on the surface of the nanowires. TEM pictures of ZnCoS showed surface roughness with consistent size, thickness, and diameter of the nanowires. In contrast, the TEM images of POM@ZnCoS revealed a distinct morphology, where PW12 nanoparticles were anchored onto the ZnCoS nanowires, resulting in a rough structure. HRTEM images displayed clear lattice fringes with observable lattice spacings attributed to the planes of Zn0.76Co0.24S and PW12. Selected area electron diffraction (SAED) further confirmed the crystalline nature and identified various planes within the material.

The ZnCoS catalyst required an η20 of 240 mV to achieve the OER. In contrast, the POM@ZnCoS demonstrated a lower η20 of 200 mV. Furthermore, POM@ZnCoS exhibited total water splitting with an E10 of 1.56 V, which was lower than that of ZnCoS. CA stability tests indicated that POM@ZnCoS maintained stability for 20 hours in a two-electrode setup. The remarkable OER activity of POM@ZnCoS can be attributed to the synergistic effects resulting from the heterostructure formation between POM and ZnCoS nanowires, which facilitates significant electronic and structural modulation of the catalyst. Additionally, various studies have reported a series of heterostructured catalysts, such as ZnFeS/MoS2,75 Zn0.76Co0.24S/CoS2,76 and ZnCoS/MoS2,77 all demonstrating efficient water-splitting activity.

In recent developments, doping a second metal along with a conductive substance has been employed in zinc selenide to enhance HER reactivity. The Cu–ZnSe@NC catalyst was synthesized using a MOF as a precursor for water splitting, employing a CVD method.50 SEM images of Cu–ZnSe@NC displayed cubic characteristics, while HRSEM images revealed a rough surface, indicating the successful formation of a porous nanostructure. TEM images further confirmed that the Cu–ZnSe nanoparticles were well-dispersed within an amorphous carbon matrix. HRTEM images indicated a d-value of 0.327 nm, corresponding to the (111) facet of Cu–ZnSe, thereby providing provisional evidence for the formation of the composite catalyst.

Raman data revealed three prominent peaks corresponding to the G, D, and 2D bands at 1575.9 cm−1, 1333.2 cm−1, and 2807.9 cm−1, respectively. Notably, the peaks for Cu–ZnSe@NC shifted to lower wavenumbers compared to those of pristine ZnSe@NC, indicating that the incorporation of copper may have altered the electrical properties of Cu–ZnSe@NC. The Se 3d spectra showed a shift to lower binding energies by approximately 0.2 eV in Cu–ZnSe@NC compared to ZnSe@NC, suggesting an increase in electron density due to Cu doping. Similarly, the Cu–C 1s spectra exhibited shifts to higher binding energies, implying that electron depletion occurred following Cu incorporation. Deconvolution of the N 1s spectra identified three peaks. As expected, Cu–ZnSe@NC exhibited remarkable hydrogen evolution reactivity with an η10 of 84 mV, significantly outperforming pristine ZnSe@NC, which required 222 mV. Additionally, various studies have reported the introduction of conductive substances into zinc selenide to enhance water-splitting performance.

These studies concluded that Zn-based materials exhibit exceptional reactivity for the electrocatalytic HER due to their altered electronic properties (Table 2).22,31,32,78,79 Despite promising attributes such as earth abundance, tunable electronic properties, and structural versatility, zinc-based materials face several challenges in the HER. Key limitations include relatively low intrinsic conductivity, moderate catalytic activity compared to noble metals, and structural instability under prolonged electrochemical operation, particularly in acidic media where Zn is prone to corrosion and dissolution. Moreover, the limited density of active sites and sluggish reaction kinetics hinder their practical performance. Looking forward, overcoming these barriers require innovative strategies such as electronic structure modulation through doping or alloying, development of Zn-based heterostructures with conductive supports, and stabilization via protective coatings or hybrid architectures. Integration with computational screening and in situ/operando techniques is essential for understanding reaction pathways and guiding rational design.

Table 2 Comparison of electrocatalytic HER performance of various Zn-based materials reported in the literature
Sr. no. Catalysts Overpotential (mV) Current density (mA cm−2) Ref.
1 ZnOHF–ZnO 181 10 58
2 Gd2O3–In2O3–ZnO 271 10 60
3 Zn–CoP 166 100 61
4 Zn–CoP 138 10 62
5 Zn–CoNiP 177 100 63
6 Zn–CoNiP 34 10 64
7 Co2P/CoP–NC 91 10 65
8 ZnS/PANI 168 10 71
9 CoZnS 130 10 72
10 Zn–CoNiS 138 10 73
11 POM/ZnCoS 200 20 74
12 ZnFeS/MoS2 145 10 75
13 Zn0.76Co0.24S/CoS2 238 20 76
14 MoS2@Zn0.76Co0.24S 169 20 77
15 Cu–ZnSe@NC 84 10 50


6. Oxygen reduction reaction (ORR)

One important electrochemical process that takes place in a variety of energy conversion devices is the ORR.2,7,8,16,19 In these systems, the ORR is crucial for generating electricity by reducing O2 molecules at the cathode. Systems that seek to efficiently transform chemical energy into electrical energy depend on this process. The quantity of electrons transported during the reaction is the primary factor distinguishing the several paths by which the ORR might progress.2,7,8,16,19

Two primary mechanisms are widely studied for the oxygen reduction reaction (ORR): (i) the four-electron pathway and (ii) the two-electron pathway (Fig. 8).142,143 The O2 is immediately reduced in the 4e route to OH in an alkaline medium or H2O in an acidic electrolyte. The 4e is the most efficient pathway, as it produces a large amount of energy by immediately reducing oxygen to water. In the 2e process, oxygen is reduced to either O22− in an alkaline medium or H2O2 in an acidic medium, which can be further reduced to water.2,7,8,16,19 This pathway is less efficient and can degrade the performance of the electrochemical device. The ORR is inherently slow and requires highly efficient catalysts to proceed at a reasonable rate, especially in the four-electron pathway.


image file: d5qm00354g-f8.tif
Fig. 8 (a) Associative mechanism of oxygen reduction reaction; and (b) dissociative mechanism of oxygen reduction reaction. Reproduced from ref. 61 with permission from the Royal Society of Chemistry.

Precious metal catalysts like platinum are often used, but alternative catalysts based on transition metals and carbon-based materials are being explored due to high cost and scarcity of noble metals. The mode of oxygen adsorption on the metal surface plays a crucial role in determining the reaction pathway.142 In the four-electron mechanism, oxygen undergoes reduction via a bidentate adsorption mode, leading to the formation of water (H2O). In contrast, the two-electron mechanism involves an end-on adsorption mode, resulting in the production of hydrogen peroxide (H2O2).143,144 The four-electron pathway is preferred in energy conversion applications, such as fuel cells and metal–air batteries, due to its higher efficiency. Meanwhile, the two-electron pathway is primarily utilized in industrial processes for H2O2 production.144

The two-electron ORR process differs under alkaline and acidic conditions. In alkaline media, the active catalytic site initially binds to molecular O2, which is subsequently reduced through the transfer of an electron and a proton, forming the hydroperoxide intermediate (*OOH).2,7,8,16,19,61 This intermediate undergoes further reduction by accepting an additional electron, producing HO2, which then desorbs from the catalyst surface, completing the reaction cycle. In acidic media, the mechanism diverges due to the high concentration of free H+ ions. Here, the *OOH intermediate readily reacts with protons, directly yielding hydrogen peroxide (H2O2) without forming the HO2 species.2,7,8,16,19,61 The four-electron transfer mechanism also differs in acidic and alkaline media. In an acidic medium, molecular O2 adsorbs onto the active catalytic site of the electrode. Furthermore, O2 undergoes the first proton-coupled electron transfer, forming a hydroperoxide intermediate (*OOH). A second proton-coupled electron transfer reduces *OOH, breaking the O–O bond and forming adsorbed oxygen (*O) and water (H2O).2,7,8,16,19,61 The *O is further reduced by sequential proton-coupled electron transfer, forming adsorbed hydroxyl (*OH) and then water (H2O). The final product, H2O, desorbs from the catalyst surface, completing the cycle. In an alkaline medium, O2 adsorbs onto the catalytic active site. After that, the first electron transfer occurs, followed by protonation from water, forming *OOH.

The *OOH undergoes further reduction and splits into adsorbed oxygen (*O) and hydroxide (OH). The *O is reduced to *OH through another proton-coupled electron transfer, followed by a reduction to OH. Finally, OH desorbs from the active site, regenerating the catalyst surface.2,7,8,16,19,61 The distinction between the two-electron and four-electron oxygen reduction pathways arises from variations in oxygen adsorption modes, which influence the reaction mechanism. These differences are primarily governed by the binding strength between O2 and the active catalytic site. When the active site exhibits a strong affinity for O2, the molecule is adsorbed in a parallel orientation, interacting with two adjacent active centers to form a bridging oxygen configuration (A–O–O–A).2,7,8,16,19,61 This adsorbed O2 undergoes proton coupling to form an A–OH intermediate, which is subsequently reduced to water (H2O). In contrast, when O2 is adsorbed in an end-on configuration, it undergoes a proton-coupled electron transfer to form a MOOH intermediate. During this pathway, successive electron transfers lead to the formation of A–O and A–OH intermediates, eventually resulting in H2O. The end-on adsorption mode is associated with higher reaction efficiency and faster kinetics compared to the bridging oxygen configuration. These insights are crucial for the rational design of catalysts with enhanced activity and performance in oxygen reduction processes.2,7,8,16,19,61

In this regard, zinc-based materials are well-studied for the electrochemical ORR. Mostly, zinc-based oxides and alloys have been reported for the electrochemical ORR.2,7,8,16,19,61 However, introducing N-doped carbon (NC) with Zn-based materials has been observed to offer excellent advantages in the ORR. When nitrogen atoms are added to the carbon lattice, catalytically active sites are formed, which makes it easier for oxygen to be adsorbed and reduced. Nitrogen's higher electronegativity than carbon attracts electron density, leading to a more efficient interaction with oxygen molecules.2,7,8,16,19,61 Nitrogen atoms can donate electron density to adjacent carbon atoms, creating localized charge density that enhances the adsorption of oxygen intermediates. Nitrogen doping can change the spin density of carbon atoms, making the catalyst more effective in facilitating the 4-electron pathway for more efficient ORR.

In Zn-based materials with NC, four different types of nitrogen have been observed to be coordinated with carbon (Fig. 9).80 Pyridinic nitrogen provides lone pairs of electrons that can participate directly in the reduction process, contributing significantly to the catalytic activity. It is known to facilitate the reduction of O2 by interacting with intermediates. Pyrrolic nitrogen contributes to the structural stability of the carbon framework. Graphitic nitrogen alters the conductivity of the carbon matrix, improving electron transfer rates, which is essential for raising the ORR's general effectiveness.80


image file: d5qm00354g-f9.tif
Fig. 9 Schematic illustrations showing four types of nitrogen present in Zn-based materials having N-doped carbon.

Oxidized nitrogen (quaternary) can increase the hydrophilicity of the catalyst, promoting better interaction with the electrolyte. To fully capitalize on the advantages of both oxides and N-doped carbon materials and improve electrochemical performance, a layered crystalline spinel nanostructure was hybridized with nitrogen-doped graphene (NG) nanosheets. This hybridization aimed to enhance the catalyst's activity and stability, making it more effective for the ORR. This led to the effective synthesis of a hybrid Zn-based catalyst called Zn2Mn3O8 fixed on NG sheets.23 A hydrothermal technique and post-annealing treatment were used to create the catalyst. During synthesis, the amount of precursor salts may be changed to precisely regulate the hybrid's size, shape, and morphology. A potential contender for ORR applications, the hybrid showed remarkable electrocatalytic activity, excellent cycle stability, and great methanol endurance in alkaline circumstances.

Zn2Mn3O8/NG was first established by the PXRD analysis, which manifested the spinel structure of Zn2Mn3O8 with the peak of NG having significant intensity. The D and G bands indicative of graphitic carbon substrates were represented by two distinctive peaks in the Raman data of the studied materials, which were located at approximately 1350 cm−1 and 1590 cm−1.23 These bands were key indicators of the material's structural properties. While the G band correlated with the in-plane vibration of sp2-connected carbon atoms, the D band was linked to defects in the carbon lattice. Defect concentration in graphitic nanostructures was measured by the D to G band intensity ratio. For NG, the ID/IG ratio was 0.98, notably higher than that of GO (0.90), indicating an increase in defect density. The ID/IG values were further elevated in the hybrid materials: ZnO–NG showed a ratio of 1.01, Mn3O4–NG had 1.0, and Zn2Mn3O8–NG reached 1.02. Due to the stabilizing impact of dense metal nanoparticles on the graphene surface, this rise indicated a higher degree of distortion inside the graphitic composition.23

The electrocatalytic performance of the Zn2Mn3O8–NG hybrid for the ORR was initially assessed using CV in alkaline electrolyte, both under N2 and O2-saturated conditions. The CV curve for the catalyst exhibited a well-defined reduction peak at −0.14 V, indicating superior ORR activity in an alkaline environment. The Zn2Mn3O8–NG hybrid was the most notable catalyst when comparing the CV plots of other catalysts since it showed the greatest peak current and the strongest positive peak potential. In contrast, the Epeak values were −0.213 V for rGO, −0.187 V for NG, −0.184 V for ZnO–NG, and −0.142 V for Mn3O4–NG. These outcomes demonstrated the outstanding catalytic reactivity of the Zn2Mn3O8–NG hybrid. Further analysis using LSV underscored the good reactivity of the Zn2Mn3O8–NG hybrid, showing the highest onset potential (Vos = −0.013 V) and half-wave potential (Vhw = −0.12 V). These metrics outperformed the other catalysts. The uniform distribution of active nanomaterial over graphene's enormous surface area and excellent conductivity significantly enhanced the catalyst's performance. This uniform distribution improved interfacial charge transfer, increased the number of adsorption sites, and created more active centers for the ORR.

In similar studies, ZnFe2O4/rGO25 and Co–ZnO@NC/CNT33 have been employed for the ORR. According to these investigations, ZnO was essential in improving ORR performance by catalyzing the formation of critical active species such as metallic Co0, Co2+ ions, and nitrogen species like pyridinic and graphitic N, all of which were essential for ORR activity. These species provide active sites facilitating oxygen reduction, boosting the overall catalytic efficiency. Like oxides, zinc–cobalt sulfide fabricated on N,S-doped carbon was reported for the electrochemical ORR.49 The exceptional ORR efficacy of the synthesized Zn–Co–S@NSC in an alkaline medium was facilitated by the binary metal system's distinct structure and synergistic effects. This catalyst exhibited impressive electrocatalytic activity, having a Vos of 0.955 V vs. RHE and a Vhw of 0.831 V vs. RHE, comparable to the traditional Pt/C material. DFT analysis supported these observations, indicating that the binary metal system effectively enhanced the ORR activity through the optimization of the free energy, leading to more efficient oxygen reduction. Furthermore, Huang et al. developed ZnS/Co-NSCNTs for the electrocatalytic ORR.81 They also observed a similar effect of N-doped carbon material, as mentioned in the previous literature.

As further progress of Zn-based materials, Zn-based alloys have been well studied in the literature for the electrocatalytic ORR. For example, a work presented by Fu et al. demonstrated the creation of Zn and Fe on the NC material (Zn/Fe–NC) for efficient ORR reactivity.82 This method produced dual-atom Zn and Fe reactive sites that were stabilized inside a carbon framework with a porous structure resembling a honeycomb. As a result of this thermal treatment, a significant portion of the zinc evaporated, and Fe-infused ZIF-8 produced a microporous carbon structure that included both Zn and Fe after breaking down. In the interim, a carbon-based structure was produced as a result of the PVP matrix carbonizing.

There were two noticeable peaks of diffraction for the graphitic carbon planes in the PXRD data of Zn–NC and Zn/Fe–NC. The efficient dispersion of Zn and Fe atoms throughout the NC matrix was confirmed by the notable absence of PXRD peaks associated with any other phases having Zn and Fe. Two separate peaks were visible in the Raman data of materials at 1350 cm−1 for the D band and 1580 cm−1 for the G band. For Zn–NC, Zn/Fe–NC, and Zn/Fe–NC-50, the ID/IG values were determined to be 0.97, 0.94, and 0.96, respectively. According to these ratios, the Zn/Fe–NC sample had a somewhat more significant graphitization level than the rest.82

XAS was examined to learn more about the conditions for coordination in Zn/Fe–NC catalysts. The Zn K-edge value of Zn/Fe–NC was situated in the middle of the metallic Zn foil and ZnO spectra, suggesting that the Zn's chemical valence state lay somewhere in between (Fig. 10a). The Fe K-edge value in Zn/Fe–NC was found to be closer to that of FeO than Fe2O3, indicative of the +2 state of Fe (Fig. 10b). At 1.47 Å, the Zn–N scattering channel was identified as the source of the significant peak in the Zn R space plot of Zn/Fe–NC (Fig. 10c). Crucially, no signal corresponding to the Zn–Zn link was seen, which further supported the idea that atomically distributed Zn species are present and rule out metallic Zn. With no discernible peak for the Fe–Fe bonding, the Fe R space plot showed a prominent peak at 1.50 Å, which corresponded to the Fe–N scattering channel (Fig. 10d). The fact that there was no Fe–Fe signal suggests that Fe was atomically distributed across the Zn/Fe–NC material. These results strongly support the effective creation of atomically scattered Zn and Fe sites in NC.


image file: d5qm00354g-f10.tif
Fig. 10 (a) Zn K edge XANES plot for Zn/Fe–NC, ZnO and ZnPC; (b) Fe K edge XANES plot of Zn/Fe–NC, FeO and FePC; (c) Zn R space profile of Zn/Fe–NC, ZnO and ZnPC; (d) Fe R space profile of Zn/Fe–NC, ZnO and ZnPC; (e) CV profile of Zn/Fe–NC in N2 and O2 containing 0.1 M KOH; (f) LSV curves of Pt/C and Zn/Fe–NC in O2-saturated 0.1 M KOH electrolyte; (g) free energy change profile at U = 0 V; (h) free energy change profile at U = 1.23 V. Reproduced from ref. 82 with permission from Elsevier; and (i) ORR activity profile of A-SAC(Fe,Ni,Zn)/NC, SAC(Fe,Ni,Zn)/NC, and commercial Pt/C in O2-saturated 0.1 M KOH. Reproduced from ref. 83 with permission from Elsevier.

Significantly higher ORR activity was shown by the Zn/Fe–NC as the peak of O2 reduction emerged at 0.810 V (Fig. 10e). This potential was noticeably more positive than the studied materials, including the commercial Pt/C catalyst. The Zn/Fe–NC manifested a Vos of 1.081 V and a Vhw of 0.875 V. These measures performed noticeably better than those of pure Zn–NC, Fe–NC, and commercial Pt/C (Fig. 10f). This improvement demonstrated how the Zn and Fe atoms' synergistic collaboration in the Zn/Fe–NC improved the ORR reactivity. As demonstrated by the Zn/Fe–NC-50, which had a Vos and Vhw of 0.994 V and 0.855 V, respectively, a higher Fe content was observed to impair ORR reactivity.

The reactive sites and their combined impact on the ORR were investigated using DFT. All reaction paths showed that the electron transfer process in all the materials happened exothermically at the input of U = 0 V (Fig. 10g). At 1.08 eV energy, the reduction of OOH* was the potential-determining step (PDS) for the Zn–NC when U climbed to 1.23 V (Fig. 10h). The overpotentials of Zn/Fe–NC and Fe–NC, on the other hand, were considerably 0.282 eV and 0.438 eV, respectively, as the PDS changed to the reduction of OH*. The addition of zinc increased the reactivity of the Fe atom in the Zn/Fe–NC, as evidenced by the difference in the PDS and the significantly reduced overpotential for the Zn/Fe–NC. The electronic features of the Fe and Zn in Zn/Fe–NC and the Zn–NC and Fe–NC were examined to investigate the source of the synergistic effect in more detail.82

The Zn/Fe–NC catalyst exhibited modest electron deficiency around Zn, similar to Zn–NC, while differential charge density analysis revealed increased electron concentration near Fe compared to Fe–NC. Electron transfer from Zn to Fe led to a downshift in the Fe d-band center, reducing the adsorption strength of *OH. The d-band centers of Zn–NC, Zn/Fe–NC, and Fe–NC were −7.00 eV, −2.28 eV, and −1.60 eV, respectively, with a 0.68 eV shift in Zn/Fe–NC enhancing ORR activity. Partial density of states (PDOS) analysis indicated electronic synergy between Zn and Fe, highlighting the cooperative effect of the dual-atom configuration in improving electrocatalytic performance.

Following the above work, Tsai et al. introduced a third metal, Ni, in a ZnFe/NC based catalyst and developed a single atom catalyst consisting of Fe, Zn, and Ni in NC named A-SAC (Fe,Ni,Zn)/NC for the ORR.83 They observed that introducing a trimetallic catalyst with N-doped carbon significantly boosted the performance. Additionally, the study investigated the effectiveness of Fe–Nx, Ni–Nx, and Zn–Nx as ORR reactive sites. Their total catalytic efficacy was greatly improved by the three neighboring Fe, Ni, and Zn-SACs working in concert inside the NC framework. As a result, they obtained a Vhw of 0.88 V and a voltage differential of 0.75 V (Fig. 10i).

In an intriguing study, Lin et al. used a straightforward pyrolysis process of a meticulously crafted MOF precursor for the ORR to establish an efficient ZnCoFe–N–C SAC.84 The XAS analysis specifically elaborated on the formation of SAC. The Zn K-edge XANES data manifested that the absorption edges for Zn–NC and ZnPc were closely aligned, suggesting that Zn was in the +2-valence state. The Zn–N coordination was indicated by a noticeable peak at 1.56 Å that emerged in the Zn R space plot of ZnCoFe–N–C. Importantly, no peak associated with the Zn–Zn bonding was seen, suggesting the absence of Zn clusters within ZnCoFe–NC. The Co K-edge absorption edge for ZnCoFe–N–C was positioned between the edges of CoO and Co2O3, indicating that Co had an electronic state between +2 and +3. The R space plot data for cobalt showed only a peak at about 1.58 Å, confirming Co–N coordination and supporting the conclusion that cobalt atoms were atomically dispersed within the material.84 The absorption edge of ZnCoFe–NC was located in between FeO and Fe2O3 as seen in XANES data, indicating that the electronic state of Fe is between +2 and +3.

The EXAFS for iron displayed a dominant peak around 1.48 Å, which likely corresponded to Fe–N coordination. Notably, no signals indicating Fe–Fe bonding at 2.16 Å were observed, further confirming the atomic level distribution of Fe. The LSV test indicated that the ZnCoFe–N–C afforded the largest Vhw of 0.878 V. This performance surpassed Zn–NC, Co–NC, Fe–NC, ZnCo–NC, and ZnFe–NC, having the Vhw of 0.736 V, 0.838 V, 0.841 V, 0.841 V, and 0.823 V, respectively.

Interestingly, the Vhw of ZnCoFe–NC was 28 mV higher than the Vhw of 0.850 V for the Pt/C. The existence of M–Nx reactive sites and the synergistic impact amongst the three metal single atoms were credited with this remarkable performance. In further reports, a series of Zn-based trimetallic catalysts with N-doped carbon such as ZnCoFe/NC,85 FeCoZn@PANI,86 Ni–Co–Zn–N–PC,87 and so on have been studied for the electrocatalytic ORR.

The literature studies have established that doping heteroatoms such as boron, sulfur, and phosphorus can significantly enhance the electrocatalytic performance of N-doped carbon materials. The electronic features of materials having carbon may be altered by heteroatom doping, which increases the electrical conductivity of certain materials. The chemical characteristics of N-doped carbon may be precisely adjusted by changing the kind and quantity of heteroatoms present.86 This allows for the optimization of materials for specific applications, such as enhancing the selectivity and activity in chemical reactions. Heteroatom doping can improve the structural stability of N-doped carbon materials under operating conditions. In this regard, Liu et al. presented an innovative design featuring atomically distributed Zn and Co dual metals on N,S co-doped 3D dendritic carbon structure optimized for the high-performance ORR.88 The addition of S, which had a bigger atomic size and a smaller electronegativity than N, successfully changed the reactive sites and electronic features. The XAS analysis verified that the electronic characteristics were adjusted by introducing S and N in C. The resulting (Zn,Co)/NSC catalyst exhibited exceptional ORR reactivity. The (Zn,Co)/NSC demonstrated Vhw, which was 67 mV greater than commercial Pt/C. The A-Zn@NSG has been thoroughly investigated for the ORR in subsequent developments.89

Although the approaches mentioned above in developing Zn-based catalysts improve the catalytic ORR performance, few researchers have demonstrated the introduction of Au nanoparticles in Zn-based catalysts to attain impressive performance. Consequently, they obtained efficient ORR activity with Au-doped Zn-based materials. Introducing Au is primarily done to serve as a location for nucleation to lay down other metal nanoparticles, creating catalysts with high surface area and active site density. Furthermore, Au can improve the conductivity of Zn-based materials. Its high electrical conductivity can facilitate better charge transfer during the ORR, contributing to enhanced catalytic efficiency. A multilayer core–shell-shaped Au@ZnFe/C material was created by Lu et al. and utilized for ORR reactivity.90 The ZnFe-MOFs and C material were used in an experiment to wrap a single Au nanoparticle in a porous carbon shell embedded with Zn–Fe compounds. Spectroscopy and microscopy were used to study the core–shell structure of Au@ZnFe/C. The Au@ZnFe/C material manifested a Vos of +0.94 V vs. RHE for ORR reactivity.

In contrast to Pt/C, a follow-up investigation showed an O–PdZn catalyst with multiple Pd atomic layers to triple the mass reactivity for the ORR. Furthermore, adding Au to the O–PdZn (Au–O–PdZn) significantly increased its longevity; after 30[thin space (1/6-em)]000 potential cycles, the mass reactivity decreased by less than 10%.91 The stabilizing effect of the Au atoms and the meticulously planned orderly structure were credited with this improvement. Using spectroscopic techniques, it was shown that Au not only entered the PdZn lattice and produced a homogeneous distribution across the particles but also substituted Pd and Zn at their surfaces through galvanic interchange.

It is evident from the aforementioned research that Zn-based materials may be a viable option for the ORR (Table 3).92–107 However, Zn-based materials face several challenges in terms of the ORR, which include insufficient intrinsic catalytic activity, poor electrical conductivity, and low tolerance to intermediate poisoning (e.g., by peroxide species), particularly in acidic environments. Furthermore, Zn-based catalysts often exhibit limited stability, especially when used in practical devices such as zinc–air batteries. To address these challenges, future strategies may include designing atomically dispersed Zn active sites embedded in conductive carbon matrices (e.g., Zn–N–C structures), engineering Zn-based hybrid systems with transition metals to modulate the d-band center, and introducing structural defects or vacancies to boost active site density.

Table 3 Comparison of the electrocatalytic ORR activity of Zn-based materials reported in the literature
Sr. no. Catalysts Onset potential (V) Half-wave potential (V) Ref.
1 Zn2Mn3O8/NG −0.013 −0.12 23
2 ZnFe2O4/rGO −0.08 25
3 Co–ZnO@NC/CNT 0.90 0.86 33
5 Zn–Co–S@NSC 0.955 0.831 49
6 ZnS/Co–NSCNTs 0.871 81
7 Zn/Fe–NC 1.0801 0.875 82
8 FeNiZn/NC 0.92 0.88 83
9 ZnCoFe–N–C 0.878 84
10 ZnCoFe/NC 0.95 0.878 85
11 Ni–Co–Zn–N–PC 0.864 87
12 Zn–Co/N,S–C 1.07 0.893 88
13 A-Zn@NSG 1.040 0.905 89
14 Au@ZnFe/C 0.94 90


7. Nitrogen reduction reaction (NRR)

The nitrogen reduction reaction (NRR) is the process of reducing molecular nitrogen to ammonia, which is a six-electron transfer process.1,4,13–15 It is a crucial reaction in natural and industrial processes, especially for producing ammonia, an essential chemical for fertilizers and various industrial applications. The NRR mimics the biological nitrogen fixation process carried out by certain bacteria. However, electrochemical or catalytic setup aims to reduce nitrogen more sustainably and efficiently than current industrial methods such as the Haber–Bosch process.1,4,13–15 The electrocatalytic NRR follows two pathways, namely associative and dissociative pathways (Fig. 11). In the first one, both nitrogen atoms are reduced simultaneously. In the later one, the nitrogen–nitrogen bond breaks, and each nitrogen atom is reduced independently.1,4,13–15 Although various advancements have been achieved in the electrocatalytic NRR process, several challenges remain.13–15 The NRR is often hindered by the competing HER, which dominates under standard electrocatalytic conditions. Zn-based materials, due to their unique properties and tunability, help mitigate HER and enhance NRR selectivity through several strategies.13–15 Zn-based materials often exhibit low hydrogen adsorption energy (H*), making the HER less favorable compared to N2 adsorption and activation. Zinc-based active sites promote π-backdonation from d-orbitals to the antibonding orbitals of N2, facilitating selective N2 reduction over H2 evolution.
image file: d5qm00354g-f11.tif
Fig. 11 Reaction pathways illustrating key electrochemical nitrogen reduction reaction (NRR) mechanisms: (a) dissociative mechanism, (b) alternating associative mechanism, (c) distal associative mechanism, and (d) Mars–van Krevelen (MvK) mechanism.

Defect-rich Zn-based materials preferentially bind and activate N2 molecules rather than protons, suppressing the HER.13–15 Zn-based materials doped with heteroatoms (e.g., N,S) create unique coordination environments that lower the activation barrier for N2 reduction while reducing HER activity. Zn-based materials with polar surfaces can stabilize adsorbed N2 through dipole interactions, favoring NRR pathways. In non-alkaline media, Zn-based materials often demonstrate weaker HER kinetics, allowing the NRR to dominate.

The triple bond in molecular nitrogen (N[triple bond, length as m-dash]N) is one of the strongest chemical bonds, requiring significant energy to break. Overcoming this challenge mildly and efficiently is the biggest task for the NRR.13–15 Competing side reactions, especially the HER, often dominate during the NRR. Since the electrochemical potential for the HER is close to that for the NRR, catalysts can preferentially reduce protons to H2, lowering ammonia yield. Efficient catalysts for the NRR are still under development. Transition metal-based catalysts and non-metal catalysts like boron nitride and carbon-based materials are being explored for the NRR due to their ability to bind and activate nitrogen. However, achieving high selectivity, efficiency, and stability remains a challenge. In most systems, the faradaic efficiency is low due to the strong competition from side reactions, particularly the HER.

In this context, Wen and coworkers explored oxygen vacancy containing Zn-doped Co3O4 for the electrocatalytic NRR.108 The Zn–Co3O4 was created using a ZIF under air calcination. The Zn–Co3O4 displayed diffraction peaks similar to those of pure Co3O4, which corresponded to the cubic Co3O4. This showed that the catalysts were effectively created by a low-temperature calcination procedure. A low carbon content inside the catalyst was also shown by the PXRD data of Zn–Co3O4-x (here x denotes 0, 5, 10, 15, 20), which matched the typical spectrum of Co3O4 with a few peaks from C components (Fig. 12a). Furthermore, as the Zn doping level increased, the Zn–Co3O4-x PXRD peaks shifted slightly toward higher angles, indicating a reduction in the d-spacing of Co3O4 due to zinc incorporation (Fig. 12a).


image file: d5qm00354g-f12.tif
Fig. 12 (a) PXRD analysis of Zn–Co3O4 and Co3O4; (b) Co 2p XPS of Zn–Co3O4 and Co3O4; (c) O 1s XPS of Zn–Co3O4 and Co3O4; (d) Zn 2p XPS of Zn–Co3O4; (e) ESR spectra of Zn–Co3O4 and Co3O4; (f) NH3 production rates and FE of Zn–Co3O4 across a range of applied potentials. Reproduced from ref. 108 with permission from the American Chemical Society; (g) Zn R space profiles of Zn Pc and Zn/NC NSs showing the absence of the Zn–Zn peak for Zn/NC NSs; (h) NH3 production rates and FE of Zn/NC NSs across a range of applied potentials; and (i) NH3 production rates and FE of Zn/NC NSs for various cycles. Reproduced from ref. 26 with permission from Wiley-VCH.

XPS analysis confirmed Zn incorporation in Zn–Co3O4, indicated by an additional peak absent in pristine Co3O4. The XPS analysis revealed Co3+ and Co2+ oxidation states (Fig. 12b). A shift to lower binding energies in Zn–Co3O4 suggests electron transfer from Zn to Co, increasing Co's electron density. This shift, along with a rise in Co2+ concentration, created oxygen vacancies (Vo) as some O atoms escape to balance the charge, creating donor–acceptor pairings of Zn and Co that may improve NRR reactivity by encouraging ammonia desorption, nitrogen binding, and dissociation.

In the O 1s XPS spectrum of Zn–Co3O4, three peaks were noted corresponding to O present in the lattice, Vo, and adsorbed oxygen (Fig. 12c).108 The Vo peak was significantly enhanced with Zn doping, indicating increased oxygen vacancies. The Zn 2p spectra verified Zn2+ (Fig. 12d). ESR measurements further explored Vo density in Co3O4 as well as Zn–Co3O4 (Fig. 12e). At g = 2.0, the Zn–Co3O4-10 showed the brightest ESR peak, indicating the highest Vo density. This resulted in a large number of Lewis acid sites that made nitrogen adsorption easier. The bimetallic ZIF-x exhibited weak ESR signals, indicating minimal Vo presence. Raman spectroscopy revealed five characteristic peaks of Co3O4, with increased Vo concentration causing peak broadening and redshift, suggesting structural modifications. Zn–Co3O4 showed the most pronounced broadening and shift, enhancing nitrogen binding, which benefits subsequent hydrogenation for ammonia production.

Under nitrogen-saturated conditions, a marked increase in current density was observed at voltages below −0.1 V, suggesting the significant NRR reactivity of Zn–Co3O4. Furthermore, in contrast to Co3O4, Zn–Co3O4 displayed a more pronounced difference in the polarization curve's current between nitrogen and argon conditions, indicating superior NRR performance for the Zn-doped material. During NRR testing at various applied potentials over a 2-hour duration, the time-dependent current density curves for Zn–Co3O4 showed minimal decay, indicating good catalyst durability in a 0.1 M HCl solution. Additionally, UV-vis absorption spectra confirmed the NRR reactivity of Zn–Co3O4 when voltage varied from −0.1 to −0.5 V. The best results were obtained at −0.3 V, offering a FE of 11.9% and an outstanding rate of 22.71 μg h−1 mg−1 to give ammonia, indicating the most efficient potential of Zn–Co3O4 for the NRR (Fig. 12f).

In an interesting study, atomically dispersed zinc supported on N-doped carbon nanosheets (Zn/NC NSs) was analyzed for electrocatalytic NRR applications.26 XAS results manifested the absence of a peak for Zn–Zn, suggesting the atomic dispersion of Zn and the formation of a single-atom catalyst (Fig. 12g). The Zn/NC NSs delivered a substantial ammonia production rate of 46.62 μg h−1 mg−1 at −0.85 V vs. RHE and an impressive FE of 95.8% at −0.70 V vs. RHE (Fig. 12h). Furthermore, the Zn/NC demonstrated excellent stability and selectivity, showing consistent NH3 production rates and FEs across multiple cycles (Fig. 12i). Structural characterization confirmed that Zn–N4 sites within the catalyst were the key active centers driving the nitrogen reduction reaction.

They conducted DFT calculations to investigate the NRR pathway and the enhanced performance of atomically dispersed Zn/NC. The DFT analysis showed the alternate and distal NRR pathways. The Gibbs free energy profiles for both pathways highlighted a notably lower energy barrier for the alternate pathway, indicating that the NRR in Zn/NC likely followed this route. In the alternate pathway, the hydrogenation of *N2 to form *NNH was identified as the PDS. The energy barrier for *N2 reduction in Zn/NC was significantly lower than in NC. This indicated that the Zn–N4 sites in Zn/NC NSs effectively lowered the energy barrier for forming *NNH intermediates, which enhanced *N2 activation and promoted subsequent hydrogenation steps, ultimately accelerating the NH3 production rate. Based on spectroscopy and theory, this study has confirmed that introducing conductive N-doped carbon substances provided several advantages to improve the catalytic NRR reactivity. Furthermore, it has been shown that ZnFe/NC,109 PdZn–NPs/NHCP,110 and Zn–N2S2-MOF111 exhibit competent reactivity for the NRR.

Historically, Zn-based compounds have predominantly been used in electrochemical energy storage due to Zn's d10 electronic configuration. However, recent defect and interface engineering advances have sparked interest in ZnS as an electrocatalyst, given its low cost and environmental benefits. For example, Feng and coworkers used thiourea in situ sulfurizing Zn foil to provide N-doped-ZnS containing sulfur vacancies (Sv) for the NRR.112 With a FE of 7.92%, the N–ZnS produced NH3 having the rate of 2.42 × 10−10 mol s−1 cm−2, which was ascribed to Sv and Zn–N reactive sites. Further studies conducted by Zhao and coworkers explored defect-rich ZnS supported on reduced graphite oxide (DR ZnS–rGO), created utilizing a Zn-salt and a S-source in a solvothermal process.113 In contrast to pristine ZnS, the ZnS with defects had a much greater NH3 output and FE, owing to enhanced stability and recyclability in nitrogen and argon-saturated environments due to rGO. According to DFT data, the first hydrogenation was the step that determined the potential, and DR ZnS–rGO with two Sv had a lower barrier of energy. The NRR used associative distal and alternative routes with similar energy barriers to go forward. Chen and coworkers developed a ZnS nanoarray via self-organized growth under solvothermal conditions.114 This ZnS@Ni structure benefits from a large surface area and strong adhesion to the 3D Ni foam substrate, resulting in an excellent rate of 5.27 × 10−10 mol s−1 cm−2 to give ammonia having 5.62% FE. Furthermore, Zhang et al. created FeS2/ZnS–NC for proficient NRR reactivity by utilizing 2-methylimidazole to create FeZn-LDH in situ and then converted it to FeS2/ZnS–NC.115 SEM revealed evenly spaced FeS2 nanosheets, generating a large number of active sites. The FeS2/ZnS–NC showed potential for good NRR reactivity, producing ammonia at 58.52 μg h−1 mg−1 having 46.84% FE.

It has been discovered that using Zn-based alloys effectively increases the NRR reactivity. By altering the electrical characteristics, researchers have significantly enhanced the NRR reactivity of Zn-based materials by integrating conductive N-doped carbon. The literature-reported studies have demonstrated that Zn-based materials offer excellent NRR activity due to their altered electronic features (Table 4). However, inherently low electrical conductivity, poor activation of the strong N[triple bond, length as m-dash]N triple bond due to limited d-orbital participation, and low faradaic efficiency arising from competing HER, especially under aqueous conditions are the main concerns with Zn-based materials. Additionally, the identification of true NRR activity is complicated by contamination and the extremely low yield rates of ammonia, making reproducibility and verification a challenge. Looking ahead, strategies such as atomically dispersed Zn coordination, dual-site engineering, and integration with defect-rich carbon supports could enhance N2 adsorption and activation. Furthermore, advanced electrolyte systems and in situ/operando techniques will be critical for mechanistic elucidation and real-time tracking of intermediates.

Table 4 Comparison of the electrocatalytic NRR activity of Zn-based materials reported in the literature (potentials have been presented against RHE)
Sr. no. Catalysts Potential FE NH3 yield rate Ref.
1 Zn–Co3O4 −0.3 V 11.9% 22.71 μg h−1 mg−1 108
2 Zn/NC −0.85 V 95.8% 46.62 μg h−1 mg−1 26
3 PdZn–NPs −0.2 V 16.9% 5.28 μg mg−1 h−1 110
4 Zn–N2S2-MOF −0.3 V 44.5% 25.07 μg h−1 cm−2 111
5 N–ZnS −0.6 V 7.92% 2.42 × 10−10 mol s−1 cm−2 112
6 ZnS–rGO −0.15 V 28.2% 51.2 μg h−1 mg−1 113
7 ZnS@Ni −0.5 V 10.8% 5.27 × 10−10 mol s−1 cm−2 114
8 FeS2/ZnS–NC −0.5 V 46.8% 58.52 μg h−1 mg−1 115


8. Nitrate reduction reaction (NO3RR)

The nitrate reduction reaction (NO3RR) is the process in which nitrate (NO3) is reduced to various nitrogen-containing compounds, such as nitrite (NO2), ammonia (NH3), or nitrogen gas (N2), depending on the conditions and the catalyst used.5,10–12,34 This reaction is vital in both environmental processes, such as denitrification in wastewater treatment and energy-related electrochemical systems, where nitrate reduction is studied for its potential in nitrogen recovery and conversion to valuable chemicals.5,10–12,34 Nitrate reduction can proceed through multiple pathways, each leading to different products (Fig. 13).
image file: d5qm00354g-f13.tif
Fig. 13 Schematic illustration for the reaction mechanism of the electrocatalytic nitrate reduction reaction. Reproduced from ref. 34 with permission from OAE Publishing Inc.

The overall reaction pathways vary depending on the intended result and the quantity of electrons involved. The reduction to nitrite (NO2) is the first step in most nitrate reduction processes, involving a two-electron transfer. In certain catalytic systems, nitrate can be reduced all the way to ammonia, involving an eight-electron transfer.5,10–12,34 Nitrate ions can also be reduced to nitrogen gas (N2), the most stable form of nitrogen. This pathway involves a multi-step process with intermediate species like nitrite (NO2), nitric oxide (NO), and nitrous oxide (N2O), requiring ten electrons. Selectively and efficiently reducing nitrate is an active area of research. Transition metal catalysts and bimetallic systems are commonly employed to achieve this goal. These catalysts must balance activity, selectivity, and stability to perform effectively under the harsh conditions often present in nitrate reduction.

Owing to their excellent properties, Zn-based materials have been well-studied for the electrocatalytic NO3RR. For instance, Huang and coworkers discovered the efficient nitrate reduction activity of a flower-shaped zinc-based ZnCo2O4 catalyst.116 They adopted a simple calcination procedure to develop ZnCo2O4 and further identified the structure of ZnCo2O4 using PXRD (Fig. 14a). The PXRD analysis established the creation of a zinc cobaltite phase that was well aligned with reference data (Fig. 14b). The SEM and TEM analyses showed flower shapes with various planes of cobaltite in HRTEM (Fig. 14c–f). Elemental mapping images clarified the even distribution of the constituting elements (Fig. 14g). As Zn was added to the cobalt oxide material, it was observed to introduce electronic modification in Co sites. These electronic modifications were tracked by XPS and UPS spectroscopy. XPS analysis confirmed Zn2+ incorporation in ZnCo2O4, with Zn 2p spectra. The Co 2p peaks indicated Co3+ and Co2+ oxidation states. A positive shift in Co3O4 peaks after Zn2+ doping suggested electron migration from Co to Zn, reducing Co electron density. Oxygen species in ZnCo2O4 were also observed. ZnCo2O4 showed a higher Vo content (37.6%) than Co3O4 (33.1%). UPS analysis revealed a lower work function (6.37 eV vs. 6.62 eV), enhancing electron transport and NO3RR catalytic activity.


image file: d5qm00354g-f14.tif
Fig. 14 (a) Synthesis of ZnCo2O4; (b) PXRD data of ZnCo2O4; (c) SEM image of ZnCo2O4; (d) and (e) TEM images of ZnCo2O4; (f) HRTEM image of ZnCo2O4; (g) elemental mapping pictures of ZnCo2O4; (h) Co 2p XPS of ZnCo2O4 and Co3O4; (i) O 1s XPS of Co3O4 and ZnCo2O4; (j) UPS valence band plot of ZnCo2O4 and Co3O4; (k) LSV profile for electrocatalytic nitrate reduction reaction with ZnCo2O4 in 0.1 M KOH solution with or without nitrate; (l) CA test with ZnCo2O4 for electrocatalytic nitrate reduction at various potentials; and (m) NH3 production rates of ZnCo2O4 and Co3O4 across a range of applied potentials. Reproduced from ref. 116 with permission from Wiley-VCH.

The fact that ZnCoO4 successfully reduced NO3 was proven by the increased current density. The onset of NO3 reduction was started at the voltage of −0.2 V vs. RHE, while the NO3RR reactivity was also assessed at varying potentials between −0.2 and −0.6 V vs. RHE (Fig. 14h). As the test voltage increased, both ZnCo2O4 and Co3O4 demonstrated higher current densities for the NO3RR, although ZnCo2O4 exhibited a significantly greater current density than Co3O4, indicating that Zn ion doping enhanced reduction current density (Fig. 14h). The NH3 yield rate was increased with increasing voltage, while the FE for NH3 followed a volcano-shaped trend, peaking at −0.4 V. The CA test at varying potentials was also conducted to determine the maximum yield. The ZnCo2O4 outperformed Co3O4, manifesting 95.4% FE to give an excellent rate of 2101.2 μg mg−1 h−1 for ammonia formation, in contrast to 83.2% and 1200 μg mg−1 h−1 for Co3O4, respectively. This enhanced performance highlighted the impact of Zn doping on the catalytic efficacy of ZnCo2O4 for the NO3RR.

In a similar type of analysis, Dong and coworkers reported ZnCr2O4 (denoted as 2Z-ZCO) with oxygen vacancies for nitrate reduction.117 However, there was a difference that the Dong group performed the nitrate reduction in a buffer solution with a neutral pH. They also adjusted the composition of Zn and Cr to get an optimized composition. Furthermore, they performed DFT, explaining that oxygen vacancies significantly altered the d-band center and improved catalytic reactivity. The existence of Zn0 in 2Z-ZCO was verified by XAS analysis, where the magnitude of the signal in 2Z-ZCO was lower compared to ZnO but greater in contrast to Zn foil. The Zn0 arrangement was further confirmed by the Zn–Zn bonding signal and the Zn–O coordination peak.

The LSV data showed a noticeable rise in current density in electrolytes with NO3, suggesting active nitrate electroreduction (Fig. 14i). The 2Z-ZCO had the shortest radius among the analyzed materials, according to EIS Nyquist plots, indicating proficient charge transport. Furthermore, 2Z-ZCO exhibited a minimal slope value compared to the others, suggesting a faster reaction rate. Nitrate reduction on 2Z-ZCO was evaluated over a varying voltage, where NH3 yield progressively increased with increasing voltage, affording a rate of 30.32 mg h−1 mg−1 (Fig. 14j). The FE followed a volcano-shaped trend, recorded highest at −1.2 V. Beyond this potential, the FE decreased slightly.

Thus, −1.2 V was adopted as the ideal potential for assessing selectivity and stability. Ion chromatography (IC) detected an NH3 concentration of 50.8 mg L−1, which closely matched the 51.7 mg L−1 obtained through UV-vis. Compared to other forms, 2Z-ZCO, with the highest concentration of oxygen vacancies, demonstrates superior performance for nitrate reduction.

With a ΔG(NO3) of −2.28 eV, the ZCO (220) surface demonstrated significant contact with nitrate ions. Electron transport from Cr and N to O was shown by the charge density difference profile, strengthening the Cr–O binding while weakening the N–O binding. The free energy of the following stage (*NO3 → *NH3) decreased steadily, necessitating no extra energy for upcoming steps. With an energy boost of 0.62 eV, the NH3 desorption process [*NH → NH3(g)] presented a potential reaction barrier.

As demonstrated by the charge density difference, the ΔG(NO3) dropped to −1.25 eV when the Cr–O pair in ZnCr2O4 was substituted with Zn. This was due to minimal orbital interaction between the O and Zn. Interestingly, the free energy of ammonia desorption was reduced to 0.36 eV, indicating that NH3 was more readily formed and desorbed, supporting a continuous reaction flow. Posing an energy demand of 0.72 eV, the PDS for the ZnO (002) surface happened during the *NO → *N process.

With the introduction of oxygen vacancies (Vo), the PDS changed from *NO2 to *NO. The weaker Zn–O involvement, compared to Cr–O bonds, further facilitated NH3 desorption, enhancing the overall reaction efficiency. The formation energy of oxygen vacancies for ZCO showed a considerable decrease compared to ZnO. This indicated that oxygen vacancies were more readily formed in ZnCr2O4, making them more favorable in this structure. The above two studies have clearly explained based on experimental and theoretical analyses that the catalytic reactivity was significantly increased by O-vacancies. In a recent report, Ye et al. developed a reduced cobalt-based spinel oxide to enhance the electrocatalytic NO3RR.118 Comprehensive structural characterization and electrochemical assessments showed that incorporating zinc and applying a reduction treatment created electron-deficient Co2+ sites in tetrahedral coordination. The modified R-ZnCo2O4 demonstrated impressive ammonia production capabilities, achieving high selectivity and FE for ammonia, having the proficient rate of 26.75 mgNH3 h−1 mgcat−1. DFT calculations and in situ data provided insights into the NO3RR mechanism on R-ZnCo2O4. Additionally, the excellent NO3RR performance of R-ZnCo2O4 was examined within a Zn–NO3 battery, demonstrating an OCV of 1.46 V having 1.40 mW cm−2 power density. These results underscored the material's potential for the NO3RR.

In an interesting study conducted by Du et al., Zn–CuO has been created as an efficient catalyst for excellent NO3RR reactivity.119 This was because Zn doping modified the electronic features of CuO, which greatly improved charge transport. By subjecting commercial brass to an alkaline medium, where it underwent successive corrosion, oxidation, and structural reformation, these Zn–CuO NAs were created. Firstly, the brass was dealloyed, generating nanoscale Cu, rapidly oxidizing to form Cu2O with a lower valence. At the same time as Zn2O was incorporated, this CuO was further oxidized to CuO and reassembled into nanosheets. For the NO3RR, Zn–CuO demonstrated exceptional performance, achieving an efficient rate of 945.1 μg h−1 cm−2 for ammonia creation. This efficiency placed Zn–CuO among the top-performing catalysts, showcasing a straightforward and scalable synthesis method that enabled efficient NH3 production from NO3 reduction. In further work, Yang and coworkers developed Ti/CuZn5O for electrocatalytic nitrate reduction with efficient ammonia yield.120

As already studied, in the final step of NO3 reduction to N2, N–N coupling is necessary and can only occur when two N* species occupy adjacent active sites. Supported single-atom catalysts, which feature isolated active metal centers without neighboring active sites, may hinder this N–N coupling step, thus reducing the selectivity toward N2 production. In this context, Zhao and coworkers developed a Zn-based ZnSA-MNC as an efficient and selective catalyst for the NO3RR.121 This catalyst achieved a NO3 conversion rate of 97.2%, with an NH3 selectivity of 94.9%, producing ammonia at the rate of 2.31 mmol h−1 mg cat−1. DFT calculations suggested that the positively charged Zn atoms, coordinated with N atoms, promoted the NO3RR, aiding in reaction pathway clarification. Two primary reaction pathways guided NO3 conversion, involving Zn–O and Zn–N bond formation that led to *ONO and *NOO intermediates. The rate-limiting step for NO3 to NH3 conversion on ZnSA-MNC was the initial hydrogenation (*NO3 → *NO3H), validating the energy demand of 0.58 eV. In further developments, Zn-doped Cu nanosheets and Zn1/MnO2 were demonstrated for efficient nitrate reduction reactions. Besides these studies, numerous Pt/calcined CuZnAl hydrotalcite,122 Cu/Zn/TiO2,123 Fe/ZnO,124 Cu@ZnO,125 and Zn–Cu catalyst126 have been designed for the electrocatalytic NO3RR. The successful application of zinc-based catalysts for nitrate reduction addresses environmental concerns associated with nitrate pollution and presents opportunities for resource recovery and the generation of value-added products (Table 5).

Table 5 Comparison of the electrocatalytic NO3RR activity of Zn-based materials reported in the literature (potentials have been presented against RHE)
Sr. no. Catalysts Potential FE NH3 yield rate Ref.
1 ZnCo2O4 −0.4 V 95.4% 2101.2 μg mg−1 h−1 116
2 ZnCr2O4 −1.2 V 90.21% 30.32 mg h−1 mg−1 117
3 ZnCo2O4 −0.3 V 92.2% 26.75 mg h−1 mg−1 118
4 Zn–CuO −0.7 V 95.6% 945.1 μg h−1 cm−2 119
6 ZnSA-MNC −1.0 V 97.2% 2.31 mmol h−1 mg−1 121
9 Fe/ZnO −0.7 V 87.0% 31 nmol s−1 cm−2 124
10 Cu@ZnO −0.6 V 89.14% 6.03 mg cm−2 h−1 125
11 ZnCu-alloy −0.85 V 98.4% 5.8 mol h−1 g−1 126


9. Carbon dioxide reduction reaction (CO2RR)

The carbon dioxide reduction reaction (CO2RR) is a chemical process where carbon dioxide (CO2) is electrochemically reduced to produce valuable products such as fuels and chemicals.6,18,29 This reaction is of significant interest in the context of addressing climate change and promoting sustainable energy because it can help mitigate CO2 emissions by converting them into valuable products, utilizing renewable energy in the process. The reaction requires the presence of a catalyst to lower the energy barriers and drive the reduction of CO2.6,18,29 Electrocatalysts are often metals like copper, silver, zinc, or gold. However, non-metallic materials like carbon-based catalysts or metal–organic frameworks (MOFs) are also being explored. The products of CO2RR depend on the catalyst used, the applied potential, and the reaction conditions. They range from carbon monoxide (CO), formic acid (HCOOH), and methane (CH4) to more complex molecules like ethylene (C2H4), ethanol (C2H5OH), or other hydrocarbons and alcohols (Fig. 15).6,18,29 The CO2RR process involves multiple electron transfer steps and intermediate stages, making it complex.
image file: d5qm00354g-f15.tif
Fig. 15 Electrocatalytic CO2RR shows the various products formed after CO2 reduction and their reaction mechanisms. Reproduced from ref. 18 with permission from Wiley-VCH.

Specifically, Zn sites, especially those in oxidized states, exhibit moderate binding to CO2, facilitating its activation without excessive binding that would hinder desorption.6,18,29 Zn's electronic configuration supports partial electron transfer to the CO2 molecule, bending the linear CO2 geometry and forming reactive intermediates like CO2 or CO. Zn has an optimal binding energy for CO, stabilizing *CO intermediates without poisoning the catalyst. This is critical for the subsequent coupling reactions that form C2+ products. In Zn oxides or hydroxides, the presence of surface hydroxyl groups can stabilize intermediates like *COOH and *CH2, facilitating multi-step reductions. Zn surfaces or Zn atoms anchored in a nitrogen-doped carbon matrix (Zn–N4 sites) provide active sites where *CO intermediates are brought into close proximity, increasing the likelihood of C–C coupling.6,18,29 The Zn's coordination environment, particularly in Zn-doped carbon structures, can polarize the adsorbed *CO, facilitating the coupling of two *CO molecules to form C2 intermediates like *OCCO. Hybrid Zn systems (e.g., Zn–Cu or Zn with co-catalysts) enable synergistic effects, where Zn stabilizes intermediates, and neighboring active sites drive C–C coupling. After C–C coupling to form *OCCO, hydrogenation steps catalyzed by Zn sites can lead to *CH2CHO, which further reduces to ethanol. Zn sites can also promote the dehydration of intermediates like *CH2CH2OH, leading to ethylene formation. Hydrophobic surfaces on Zn-based materials can enhance ethylene selectivity by destabilizing polar intermediates, favoring desorption.

The CO2 can be reduced through different pathways, producing various products. CO is one of the simplest and most common products, requiring 2 electrons. Formic acid requires 2 electrons and protons.6,18,29 Methane is a more reduced product that requires 8 electrons and protons. Ethylene formation requires 12 electrons and protons. The selectivity of the catalyst determines which pathway dominates. Controlling the production of specific products is challenging due to the multiple reaction pathways. Achieving high faradaic is critical. Catalysts can degrade over time, and maintaining long-term performance is a major research focus. Enhancing CO2RR's energy efficiency is crucial to making the process commercially feasible because it uses much energy.6,18,29

Exploring the Zn-based oxides, Shviro and coworkers demonstrated ZnO with varying morphologies for efficient CO2RR reactivity.127 The ZnO nanoparticles were prepared in diverse morphologies, such as nanorods, nanosheets, nanoparticles, and random shapes. They observed that the change in the morphology also altered the electronic states of Zn. As a result, ZnO nanorods offered excellent reactivity for the CO2RR. The PXRD studies showed the creation of ZnO hexagonal wurtzite structure for all the morphologies. The Zn 2p XPS manifested that the XPS peaks of nanorods were moved in the positive direction by 0.7 eV in contrast to nanoparticles, suggesting the electronic property modification. This resulted from significant variations in the material's shape, crystallinity, and Ov inside its lattice. In terms of reactivity, all ZnO samples demonstrated excellent CO selectivity with a CO FE of approximately 80%. The Zn nanorods and nanosheets achieved the most outstanding FE values, reaching 90% at higher negative potentials, equivalent to Ag nanoparticle's reactivity. Notably, ZnO nanorods had the greatest CO current density out of all the studied materials. When compared to other ZnO-based gas diffusion electrodes, ZnO nanorods demonstrated the most promising result, affording −150 mA cm−2 current density with the input of −1.2 V voltage while retaining a strong FE of 83%.

Furthermore, Chen and coworkers established the CO2RR mechanism in zinc-based electrocatalysts, a set of thermally oxidized zinc foils prepared to establish a connection between the catalyst's chemical state and product selectivity.128 Findings from in situ Raman spectroscopy, XAS, and PXRD indicated that surface Zn(II) species primarily produced carbon monoxide, while Zn(0) species were linked to formate production. Notably, thermal oxidation, which disrupted a dense oxide layer on the zinc foil surface, resulted in a fourfold increase in the FE for formate in the CO2RR. In situ analyses highlighted that the chemical state of zinc electrocatalysts played a critical role in determining product distribution, offering a viable approach for tuning selectivity in zinc-based CO2RR electrocatalysts. Therefore, it is clear that not only morphology but also the valence state of Zn has a significant impact on CO2RR reactivity.

To reduce the impact of the Zn substrate in XRD measurements, Qin and coworkers investigated a range of Zn catalysts with different ratios of Zn(002) to Zn(101) crystal facets.129 A detailed investigation of catalytic reactivity revealed that the CO/H2 ratio was influenced by the particular electrochemical reduction circumstances as well as the Zn crystal facet ratio. To confirm surface changes on the electrode during the CO2RR, ATR-IRAS analyses were conducted. The CO2 was represented by the signal at 2358 cm−1, whose intensity progressively increased in the early stages, indicating progressive CO2 adsorption on the electrode's surface. After 20 minutes, this intensity stabilized due to CO2 adsorption saturation. The C[double bond, length as m-dash]O of formate was associated with the 1723 cm−1 signal, whereas COO was associated with the 1436 and 1596 cm−1 peaks. Furthermore, the signals for the –OH group in COOH and free –OH were also witnessed in Raman data.

As electrolysis occurred, these grouping intensities surged, signifying their accumulation on the electrode surface and alteration of the electrolytic environment. The CO adsorbed on Zn was identified as the cause of two small signals in Raman. Adsorbed over the electrode surface, CO2 and HCOOO are first interconverted. Then, (i) a reversible transformation from HCOOO to CO32− radicals took place, serving as a proton source; (ii) CO2 accepted electrons to become CO2*; (iii) the CO2* radical underwent further electrochemical reduction by accessing electrons from (i), forming COOH* intermediate; (iv) a small fraction of COOH* was reduced to COOH, desorbing over the catalyst generating HCOOH; and (v) the majority of COOH* was reduced, producing CO*, which then desorbed from Zn to produce CO, the primary product.

In the CO2RR to CO, the energy necessity of COOH* and CO* was verified using DFT analysis. The process of converting CO2 to COOH* was the rate-regulating phase. The CO2RR was inhibited if the relative energy between CO2 and COOH* was too high. As the potential decreased, the CO2RR was initiated over the (101) facet of Zn with −(η + 0.2) V, where the CO2 to COOH* energy was −0.015 eV. On the (002) facet of Zn, however, this energy barrier stayed high at 0.343 eV until the CO2RR happened at −(η + 0.6) V, with the energy of CO2 to COOH* at −0.015 eV.129

DFT data evidenced that the (101) facet of Zn favored the CO2RR at a lower reduction potential due to a 0.358 eV lower CO2 to COOH* relative energy compared to Zn(002). This aligned with experimental observations in the CO2RR. At the height of the FEs of CO, Zn-1 and Zn-2 needed potentials of −1.2 and −1.1 V, respectively, but Zn-3, which had a higher percentage of Zn(101) facets, needed a potential of −0.9 V. Experimental findings and DFT simulations both verified that CO2RR to CO performance could be further enhanced by increasing Zn(101) facets.

In a work, for the efficient CO2RR reactivity, nanoporous ZnO was created by hydrothermal synthesis followed by thermal breakdown.130 The ZnO was converted to Zn throughout the electrochemical conditions used for the CO2RR, according to in situ XAS. Having a very good CO FE of 92.0%, the ZnO showed noticeably better reactivity than Zn foil. This improvement suggested that the nano-porosity of the catalyst boosted the CO2RR reactivity, possibly due to the enormous surface area and a higher number of coordination-unsaturated surface atoms. In another report, Yeo's group reported ZnO nanoparticles for electrocatalytic CO2RR reactivity.

The development of heterostructures and heterointerfaces based on zinc-based materials was crucial to improving CO2RR performance. To comprehend the link between methanol selectivity and structure in the Cu–ZnO system for the CO2RR, a thorough structural and catalytic investigation of ZnO/Cu/Al2O3 catalysts was carried out.131 Two different synthetic procedures were applied to produce ZnO/Cu with varying morphologies, resulting in distinct outcomes in terms of methanol selectivity and production. Smaller copper particles were created by the incipient-wetness impregnation technique, which produced a greater distribution of ZnO on both Cu and Al2O3. On the other hand, zinc oxide was directly deposited over the pre-existing Cu via CVD, forming more organized ZnO and Cu structures.131 Analysis through spectroscopy and microscopy showed that chemical vapor deposition provides a more advantageous alignment of zinc oxide with copper, enhancing ZnO–Cu interface formation, which was essential for selective CO2 conversion to methanol. Meanwhile, the highly dispersed Cu–ZnO phases in the impregnated sample resulted in less effective contact between phases, decreasing both reactivity to generate CH3OH, preferring the reverse water–gas-shift process, and producing more CO. These results demonstrated that, in the Cu–ZnO system, obtaining high reactivity toward methanol required an ideal ZnO–Cu phase interaction, which outweighed the advantage of greater Cu dispersion.

In addition to these studies, researchers have developed bimetallic Zn-based materials for electrocatalytic CO2RR reactivity. In this regard, oxide-based bimetallic CuZnx catalysts have been suggested as viable substitutes for attaining efficient selectivity for a range of essential products.132 The CuO was synthesized using a co-precipitation technique with varying concentrations of ZnO dopants. An aqueous carbonate solution was used to evaluate the electrochemical reactivity. Ethanol production and FE were assessed for various CuO–ZnOx (where x denotes 5, 10, 15, and 20 weight percent). Among the catalysts studied, CuO–ZnO10 produced C2H5OH at a rate of around 121 μmol h−1 L−1, with a maximum FE of 22.27%. The CuO–ZnO10 showed long-term stability for a minimum of 12 h. The catalyst's reactive sites were better understood by further analysis, which showed that CO2 reduction took place on reduced reactive sites instead of on the metal oxides.

After these efforts, a ZnTe-MOF-based ZnTe/ZnO@C material supported by NC was demonstrated with efficient CO2RR reactivity.27 The catalyst was fabricated using the solvothermal method followed by calcination. The XRD patterns manifested diffraction peaks corresponding to ZnTe and ZnO in ZnTe/ZnO@C and ZnO@C. Raman analysis was performed to evaluate the graphitization level in the created materials. Both displayed a distinctive D band at 1336 cm−1 and a G band at 1585 cm−1. The ID/IG content in ZnTe/ZnO@C suggested a high level of graphitization in the C-layer, which was beneficial for electron transport. The ZnTe/ZnO@C demonstrated significantly improved CO2RR reactivity to give formate, attaining an 86% selectivity and stable current density. Compared to previous Zn-based materials, ZnTe/ZnO@C showed greater reactivity and selectivity. According to DFT research, the stabilization of the essential HCOO* on ZnTe promoted the selective production of formate.

In recent developments, Zn-single atom catalysts were studied for the electrocatalytic CO2RR. Han et al. created an electrocatalyst named SA-Zn/MNC for the electroreduction of CO2 to methane to enable good electron transport.133 TEM and HRTEM pictures manifested the creation of SA-Zn/MNC (Fig. 16a and b). Raman spectra and specific surface area analysis indicated that incorporating zinc into the MNC led to an increase in structural defects and consequently enhanced the creation of micropores. According to the Zn K-edge XANES, the extent of the first signal for SA-Zn/MNC after the edge was comparable to that of ZnCl2 but higher in contrast to Zn foil. The pre-edge characteristic also moved to a larger energy value than the reference, more closely matching ZnCl2. These findings indicated the electronic state of Zn to be +2 (Fig. 16c). The Zn–N interaction was responsible for the noticeable signal at 1.55 Å in the Zn R space plot of SA-Zn/MNC (Fig. 16d and e). Moreover, SA-Zn/MNC showed no scattering peaks suggestive of Zn–Zn coordination. These findings validated Zn's atomic dispersion. The N 1s showed the peak for Zn–N bonding in SA-Zn/MNC (Fig. 16f).


image file: d5qm00354g-f16.tif
Fig. 16 (a) and (b) TEM pictures of SA-Zn/MNC; (c) Zn K edge XANES spectra of SA-Zn/MNC compared with ZnO and ZnCl2; (d) Zn R space profile of SA-Zn/MNC compared with ZnO and ZnCl2; (e) fitted Zn R space profile of SA-Zn/MNC; (f) N 1s XPS of SA-Zn/MNC compared with MNS; (g) LSV profile for the CO2RR activity of SA-Zn/MNC compared with MNC and Zn-powder; (h) potential dependent FE of SA-Zn/MNC compared with MNC and Zn-powder; and (i) methane yield rate with SA-Zn/MNC compared with MNC and Zn-powder. Reproduced from ref. 133 with permission from the American Chemical Society.

The LSV was examined to reveal the CO2RR reactivity of SA-Zn/MNC, MNC, and Zn powder (Fig. 16g). The results showed that SA-Zn/MNC afforded current density for the CO2RR that exceeded those of both MNC and Zn powder, indicating it had the highest activity for this reaction. The total FE for methane, carbon monoxide, and hydrogen was nearly 100% throughout the tested potential range, with no other byproducts observed. The FE of 85% was recorded with SA-Zn/MNC for CH4 generation, which was remarkably high among the highest values documented in aqueous environments (Fig. 16h).133 Moreover, 158 ± 4 μmol h−1 cm−2 was determined to be the creation rate of CH4 (Fig. 16i). Higher quantities of KHCO3 were found to increase the production rate of CH4, according to an analysis of the effect of electrolyte concentration on CO2RR reactivity.

This connection between CH4 generation and the amount of Zn in SA-Zn/MNC demonstrated how crucial atomically distributed Zn–N sites were to the CO2RR process. These studies highlighted the role of the N-doped carbon support in improving catalytic reactivity. In further efforts, a series of electrocatalysts having N-doped carbon have been designed for efficient CO2RR reactivity. In this context, ZnO,134 CuZn/rGO,135 Zn/NG,136 N-ZnNiO,137 and Zn-NHPC30 were demonstrated by different research groups exploring the efficient reactivity in terms of the CO2RR. In further advancements, ZnCu-alloy,138 Ag/Zn-alloy,139 CuZn nanoparticles,140 Pd/ZnO,141 Cu5Zn8 catalyst,142 In–Zn-alloy,143 CuZn-catalyst,144 CuZn-alloy,145 CuZn-alloy,146 SnZnOx,147 and CuO/ZnO148 have been demonstrated for efficient reactivity for the CO2RR.

With a fully occupied shell orbital and an electronic structure of d10, Zn2+ is categorized as a moderately hard acid.37 It may easily coordinate with atoms such as S, N, or O. As a result, Zn-based MOFs might have coordination numbers between 4 and 6. Given their potential for a multitude of uses, these MOFs have been the subject of much research. They have recently drawn interest as proficient CO2RR catalysts. Han and coworkers investigated the Zn-BTC MOF's capacity to reduce CO2 in ionic liquid (IL).149 Their results showed that current density and selectivity for CH4 were strongly influenced by the nature of electrodes and the shape of the MOFs. They thoroughly examined how ILs influenced CO2 reduction effectiveness, observing that the viscosity of the liquid and the kind of counterion both impacted ionic liquids' efficiency.

Their research showed that the electrolytes' composition and characteristics affected the reaction route and the end product's activity and selectivity. During the CO2RR, the Zn-MOF was essential for promoting electron transport. Their investigation demonstrated that imidazolium cations first adsorbed onto the Zn-MOF surface during electrolysis. The CO2˙ intermediates were created when one electron was moved to the CO2 the ILs had captured via Zn-MOF. Because CO adsorbed more strongly than CH4 over the Zn-MOF, these intermediates obtained another electron to create CO, which was transformed into CH4 through a six electron transfer process. Therefore, the specific CO2RR to produce CH4 depended on the synergistic action of ILs and the Zn-MOF.

Beyond organic electrolytes, using inorganic counterions has also been shown to impact the reactivity of Zn-MOF. Wang and associates used different zinc sources to create different ZIF-8 and investigated how these sources and electrolytes affected CO2RR reactivity.150 Having a FE of 65%, they discovered that ZIF-8, which was derived from ZnSO4, had the best catalytic activity for reducing CO2 to CO. Weaker contacts between the SO42− and the Zn accounted for the increased reactivity of ZIF-8 including SO42−. This facilitated anion exchange for charge balancing. Furthermore, a drop in the electrolyte's hydrated anionic radius (such as NaCl) was associated with a rise in FE for converting CO2 to CO and a corresponding fall in current density. This implied that smaller hydrated anions restricted the formation of hydrogen while enhancing anion exchange for charge balance. According to CV investigations, ZIF-8 showed a redox peak that matched the Zn center's redox reaction. While the metal nodes of ZIF-8 were considered electrocatalytic active sites, it is important to confirm this since Zn(II) is a d10 system and typically does not undergo redox reactions under the given conditions. Nonetheless, the research suggests that electrocatalytic CO2 reduction using Zn-MOFs is feasible with the appropriate solvent.

Therefore, the ability of zinc-based catalysts to effectively convert CO2 into useful products addresses pressing environmental concerns related to greenhouse gas emissions and contributes to the development of sustainable energy solutions.151–154 Although Zn-based materials have emerged as promising candidates for the electrochemical CO2RR, particularly for the selective production of CO and formate due to their moderate binding strength with *COOH intermediates and high hydrogen evolution suppression in neutral and alkaline media, several key limitations persist. These include limited multi-carbon (C2+) product selectivity, relatively low intrinsic activity compared to noble metals, and instability under long-term electrolysis due to catalyst degradation or restructuring. Moreover, the underlying reaction pathways and intermediate stabilization mechanisms at Zn active sites remain insufficiently understood. To overcome these challenges, future directions may focus on precise tuning of the electronic structure via heteroatom doping or alloying, constructing Zn-based heterostructures with synergistic interfaces and anchoring single Zn atoms on conductive nitrogen-doped carbon frameworks to enhance selectivity and atom utilization. Additionally, the integration of machine learning-guided design and in situ spectroscopic techniques is crucial for improving selectivity, stability, and scalability.

10. Conclusion and future perspectives

Zinc-based materials have emerged as promising candidates for electrocatalytic reduction reactions, demonstrating remarkable potential in the reduction of H2O, CO2, O2, N2, and NO3 into valuable chemicals. The unique properties of zinc, including its favorable coordination chemistry, low cost, and abundance, make it an attractive element for developing efficient catalysts. Recent advancements in the synthesis and characterization of zinc-based nanostructured materials have shown significant improvements in catalytic performance, selectivity, and stability. The exploration of innovative synthetic strategies has allowed researchers to tailor zinc-based catalysts' morphology, electronic structure, and surface characteristics, enhancing their efficiency in electrocatalytic processes. Notably, the integration of zinc with other metal components and the development of heterostructures have further boosted catalytic activity, leading to high faradaic efficiencies for target products.

Numerous studies have investigated the potential of zinc-based materials for various electrocatalytic reduction reactions, including the HER, ORR, NRR, NO3RR, and CO2RR. However, there is noticeable absence of comprehensive reviews that systematically examine their applications in these electrocatalytic processes. This review provides an in-depth analysis of recent advancements in zinc-based materials for key reduction reactions, emphasizing the reaction pathways involved. It discusses the structural, morphological, and electronic characteristics of these materials and how optimizing these features can improve catalytic performance. Additionally, this review investigates the relationship between structure, properties, and performance, focusing on enhancing activity, stability, and efficiency in reduction reactions. It also addresses the challenges faced in the field and outlines future directions, underscoring the significant potential of zinc-based materials.

Besides electrocatalytic reduction reactions, Zn-based materials hold potential for several applications in electrocatalysis beyond the commonly discussed areas such as water splitting and CO2 reduction. Their versatile electronic properties, tunable surface characteristics, and stability in various environments make them suitable for other emerging applications like hydrogen storage, organic pollutant degradation, and more. ZnO-based materials can act as catalysts for generating reactive oxygen species (ROS), such as hydroxyl radicals (*OH), in electrochemical systems, which are effective for breaking down organic pollutants. Zn-based catalysts can degrade complex organics like dyes, pharmaceuticals, and pesticides in wastewater through ROS-driven oxidation. Zn-based semiconductors (e.g., ZnO and ZnS) have demonstrated excellent photocatalytic activity under UV or visible light, making them suitable for removing organic pollutants. Zn-based materials offer immense potential across various electrocatalytic applications due to their versatility, tunable properties, and cost-effectiveness. While some of these applications, like pollutant degradation and nitrogen fixation, are still emerging, ongoing research into material design and performance optimization could unlock new opportunities, making Zn-based materials increasingly relevant for addressing global energy and environmental challenges.

The Zn-based materials also play a key role in commercial applications. For example, zinc oxides are used as co-catalysts with other metals in methanol synthesis and the water–gas shift reaction. These catalysts are critical for large-scale hydrogen production and methanol plants. Zinc oxide (ZnO) is widely applied for photocatalytic degradation of organic pollutants in wastewater treatment. Zn-based sorbents and catalysts are used to remove sulfur-containing compounds from fuels, meeting environmental standards for cleaner energy. Zinc chalcogenides and Zn-doped materials are explored for the HER and OER, offering potential for scalable hydrogen production. Catalysts based on Zn/Fe–NC enhance the ORR efficiency in ZABs, making them suitable for next-generation energy storage devices. Zinc-based electrocatalysts are utilized to convert CO2 into value-added chemicals like CO and formate, aiding in carbon-neutral technology. Zinc ions serve as active centers in enzymes like carbonic anhydrase, which catalyzes the hydration of CO2. Zinc-based catalysts facilitate the two-electron ORR pathway for hydrogen peroxide production, a critical chemical in pulp bleaching, disinfectants, and industrial oxidation processes. These catalysts are cost-effective and scalable for industrial H2O2 synthesis.

Although Zn-based materials have been well studied for a series of valuable applications, there are several challenges with Zn-based materials. As mentioned in the previous sections, stability is a critical challenge for catalytic materials in real-world applications, as harsh operational environments, such as extreme pH, high temperatures, or prolonged cycling, can lead to material degradation. For zinc-based materials, the main stability issues include dissolution in acidic or alkaline media, surface passivation, and structural degradation under electrochemical conditions. Addressing these issues requires careful design and modification of the catalytic materials to achieve a balance between stability and activity. Various strategies like surface engineering, doping and alloying, formation of heterostructures, and morphological tuning have been employed to enhance the stability of Zn-based materials. While enhancing stability, maintaining or even improving catalytic activity is a crucial challenge. The key is to ensure that structural modifications do not block the active sites or impede electron/proton transfer. By adopting these strategies, zinc-based catalytic materials can achieve the necessary balance of activity and durability for real-world applications, paving the way for their widespread use in energy conversion and storage systems. Furthermore, the efficiency of stability tests conducted for Zn-based electrocatalysts to evaluate industrial applicability depends on several factors, including the duration of the tests, the operating conditions, and the evaluation metrics used. The degradation mechanism can be identified during the stability test. Generally, long-term durability tests are conducted under industrially relevant conditions. During this test, in situ techniques (e.g., XPS, TEM, and operando spectroscopy) can be used to monitor structural and chemical changes during operation. Post-test analysis can be performed by ex situ analyses (e.g., SEM, XRD, and ICP-MS) to identify leaching, structural changes, and surface modifications. Moreover, computational methods can be used to predict degradation pathways and correlate them with experimental data.

If the stability tests conducted for Zn-based electrocatalysts did not thoroughly replicate industrial conditions or identify degradation mechanisms, their efficiency for evaluating industrial applicability is questionable. Incorporating comprehensive testing protocols and explicitly addressing degradation mechanisms such as leaching, structural collapse, and passivation can significantly enhance the reliability and applicability of Zn-based electrocatalysts in industrial settings.

The intrinsic catalytic activity of Zn-based materials is often lower compared to noble metal catalysts. The moderate binding energy of intermediates (e.g., *O, *OH, or *OOH) leads to suboptimal catalytic performance in reactions like the ORR and NRR. Achieving high selectivity for desired products (e.g., formate in CO2 reduction or NH3 in NRR) remains a significant challenge due to competing side reactions. Zinc-based materials generally exhibit poor electronic conductivity, which restricts charge transfer and overall catalytic efficiency. Scalable and reproducible synthesis of Zn-based materials with controlled morphology, size, and defect density remains difficult. However, these challenges can be tackled with the help of various techniques. The different techniques such as formation of heterostructures and composites, doping and alloying, and defect engineering can be utilized for the designing of advanced electrocatalysts. A special focus can be provided on understanding reaction pathways and intermediate binding energies through in situ and operando techniques. Zinc-based catalysts are integral to diverse commercial applications, spanning chemical synthesis, environmental protection, energy conversion, and beyond. Ongoing advancements in catalyst design and scalability will unlock their full potential for sustainable industrial processes.

After the significant progress made in the field of Zn-based materials, several outstanding questions are available preventing further progress. The questions are: How can the stability of Zn-based materials be enhanced in acidic media? Zinc leaching remains a critical bottleneck for their use in acidic environments. What are the most effective strategies for improving conductivity? Can advanced hybridization techniques or conductive frameworks make Zn-based materials competitive with noble metals? What governs the selectivity of Zn-based catalysts in multi-electron reduction reactions? What role do defects, doping, and morphology play in optimizing Zn-based catalysts? Can Zn-based materials achieve industrial-level performance?

These questions can be resolved by adopting advanced characterization tools, computational techniques, appropriate electrochemical testing, and creating collaborative platforms. Techniques like X-ray absorption spectroscopy (XAS), Fourier-transform infrared (FTIR) spectroscopy, and Raman spectroscopy can provide insights into reaction intermediates and catalyst behavior under working conditions. Tools like scanning electron microscopy (SEM), transmission electron microscopy (TEM), and atomic force microscopy (AFM) allow visualization of surface morphology and defects. Density functional theory (DFT) is essential for understanding reaction mechanisms, adsorption energies, and electronic structure. Machine learning (ML) provides predictive models that can identify promising Zn-based catalysts and optimize synthesis parameters. Extensive electrochemical investigations like Tafel and EIS analysis should be used to reveal the reaction kinetics and charge transfer properties. Open databases and collaborative research networks can accelerate the discovery of new materials. Zn-based materials hold immense potential for electrocatalytic reduction reactions due to their tunable properties and abundance. However, challenges such as stability, conductivity, and selectivity must be addressed. By leveraging advanced tools, fostering interdisciplinary collaborations, and focusing on sustainable design, future research can unlock the full potential of Zn-based materials for large-scale, efficient, and sustainable energy and chemical production. Overall, the progress in zinc-based electrocatalysts holds significant promise for advancing the field of electrocatalysis and addressing global energy and environmental challenges.

Author contributions

Dr. Baghendra Singh conducted the literature survey and contributed to the manuscript preparation. Dr. Baghendra Singh and Dr. Apparao conceptualized and edited the manuscript.

Conflicts of interest

The authors declare no conflict of interest.

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.

Acknowledgements

This work is gratefully supported by SERB (CRG/2023/001112) and CSIR (01(3050)/21/EMR-II). Baghendra Singh acknowledges the financial support from the Science and Engineering Research Board National Post-Doctoral Fellowship, Govt. of India (PDF/2023/001766) for providing financial assistance.

References

  1. D. Liu, M. Chen, X. Du, H. Ai, K. H. Lo, S. Wang, S. Chen, G. Xing, X. Wang and H. Pan, Development of electrocatalysts for efficient nitrogen reduction reaction under ambient condition, Adv. Funct. Mater., 2021, 31, 2008983 CrossRef .
  2. C. Kim, F. Dionigi, V. Beermann, X. Wang, T. Möller and P. Strasser, Alloy Nanocatalysts for the Electrochemical Oxygen Reduction (ORR) and the Direct Electrochemical Carbon Dioxide Reduction Reaction (CO2RR), Adv. Mater., 2019, 31, 1805617 CrossRef .
  3. I. E. Khalil, C. Xue, W. Liu, X. Li, Y. Shen, S. Li, W. Zhang and F. Huo, The role of defects in metal–organic frameworks for nitrogen reduction reaction: when defects switch to features, Adv. Funct. Mater., 2021, 31, 2010052 CrossRef .
  4. X. Zhao, G. Hu, G.-F. Chen, H. Zhang, S. Zhang and H. Wang, Comprehensive understanding of the thriving ambient electrochemical nitrogen reduction reaction, Adv. Mater., 2021, 33, 2007650 CrossRef .
  5. X. Liang, H. Zhu, X. Yang, S. Xue, Z. Liang, X. Ren, A. Liu and G. Wu, Recent advances in designing efficient electrocatalysts for electrochemical nitrate reduction to ammonia, Small Struct., 2023, 4, 2200202 CrossRef .
  6. W. Yang, K. Dastafkan, C. Jia and C. Zhao, Design of electrocatalysts and electrochemical cells for carbon dioxide reduction reactions, Adv. Mater. Technol., 2018, 3, 1700377 CrossRef .
  7. A. Kulkarni, S. Siahrostami, A. Patel and J. K. Nørskov, Understanding catalytic activity trends in the oxygen reduction reaction, Chem. Rev., 2018, 118, 2302–2312 CrossRef PubMed .
  8. M. Shao, Q. Chang, J.-P. Dodelet and R. Chenitz, Recent advances in electrocatalysts for oxygen reduction reaction, Chem. Rev., 2016, 116, 3594–3657 CrossRef .
  9. C. Li and J.-B. Baek, Recent Advances in Noble Metal (Pt, Ru, and Ir)-Based Electrocatalysts for Efficient Hydrogen Evolution Reaction, ACS Omega, 2020, 5, 31–40 CrossRef .
  10. Z. Wang, D. Richards and N. Singh, Recent discoveries in the reaction mechanism of heterogeneous electrocatalytic nitrate reduction, Catal. Sci. Technol., 2021, 11, 705–725 RSC .
  11. W. Jung and Y. J. Hwang, Material strategies in the electrochemical nitrate reduction reaction to ammonia production, Mater. Chem. Front., 2021, 5, 6803–6823 RSC .
  12. X. Lu, H. Song, J. Cai and S. Lu, Recent development of electrochemical nitrate reduction to ammonia: A mini review, Electrochem. Commun., 2021, 129, 107094 CrossRef CAS .
  13. C. Tang and S.-Z. Qiao, How to explore ambient electrocatalytic nitrogen reduction reliably and insightfully, Chem. Soc. Rev., 2019, 48, 3166–3180 RSC .
  14. T. Wang, Z. Guo, X. Zhang, Q. Li, A. Yu, C. Wu and C. Sun, Recent progress of iron-based electrocatalysts for nitrogen reduction reaction, J. Mater. Sci. Technol., 2023, 140, 121–134 CrossRef CAS .
  15. S. C. Jesudass, S. Surendran, J. Y. Kim, T.-Y. An, G. Janani, T.-H. Kim, J. K. Kim and U. Sim, Pathways of the Electrochemical Nitrogen Reduction Reaction: From Ammonia Synthesis to Metal-N2 Batteries, Electrochem. Energy Rev., 2023, 6, 27 CrossRef CAS .
  16. Y. Li, Q. Li, H. Wang, L. Zhang, D. P. Wilkinson and J. Zhang, Recent progresses in oxygen reduction reaction electrocatalysts for electrochemical energy applications, Electrochem. Energy Rev., 2019, 2, 518–538 CrossRef CAS .
  17. M.-R. Gao, J. Jiang and S.-H. Yu, Solution-based synthesis and design of late transition metal chalcogenide materials for oxygen reduction reaction (ORR), Small, 2012, 8, 13–27 CrossRef CAS PubMed .
  18. W. Zhang, Y. Hu, L. Ma, G. Zhu, Y. Wang, X. Xue, R. Chen, S. Yang and Z. Jin, Progress and Perspective of Electrocatalytic CO2 Reduction for Renewable Carbonaceous Fuels and Chemicals, Adv. Sci., 2018, 5, 1700275 CrossRef PubMed .
  19. J. Stacy, Y. N. Regmi, B. Leonard and M. Fan, The recent progress and future of oxygen reduction reaction catalysis: A review, Renewable Sustainable Energy Rev., 2017, 69, 401–414 CrossRef CAS .
  20. T. Tang, Z. Wang and J. Guan, Optimizing the electrocatalytic selectivity of carbon dioxide reduction reaction by regulating the electronic structure of single-atom M-N-C materials, Adv. Funct. Mater., 2022, 32, 2111504 CrossRef CAS .
  21. S. Yang, X. Xue, C. Dai, X. Liu, Q. Yin, J. Lian, Y. Zhao, Y. Bu and G. Li, Zinc-iron bimetallic-nitrogen doped porous carbon microspheres as efficient oxygen reduction electrocatalyst for zinc-air batteries, Appl. Surf. Sci., 2021, 546, 148934 CrossRef CAS .
  22. A. Nadeema, P. S. Walko, R. N. Devi and S. Kurungot, Alkaline water electrolysis by NiZn-double hydroxide-derived porous nickel selenide-nitrogen-doped graphene composite, ACS Appl. Energy Mater., 2018, 1, 5500–5510 CAS .
  23. J. Gautam, D. T. Tran, N. H. Kim and J. H. Lee, Mesoporous layered spinel zinc manganese oxide nanocrystals stabilized nitrogen-doped graphene as an effective catalyst for oxygen reduction reaction, J. Colloid Interface Sci., 2019, 545, 43–53 CrossRef CAS PubMed .
  24. L. Wen, X. Li, R. Zhang, H. Liang, Q. Zhang, C. Su and Y. J. Zeng, Oxygen Vacancy Engineering of MOF-Derived Zn-Doped Co3O4 Nanopolyhedrons for Enhanced Electrochemical Nitrogen Fixation, ACS Appl. Mater. Interfaces, 2021, 13, 14181–14188 CrossRef CAS .
  25. W. Hong, L. Li, R. Xue, X. Xu, H. Wang, J. Zhou, H. Zhao, Y. Song, Y. Liu and J. Gao, One-pot hydrothermal synthesis of Zinc ferrite/reduced graphene oxide as an efficient electrocatalyst for oxygen reduction reaction, J. Colloid Interface Sci., 2017, 485, 175–182 CrossRef CAS .
  26. Y. Wei, X. Wang, M. Sun, M. Ma, J. Tian and M. Shao, Atomically Dispersed Zinc Active Sites Efficiently Promote the Electrochemical Conversion of N2 to NH3, Energy Environ. Mater., 2024, 7, e12630 CrossRef CAS .
  27. X. Teng, J. Lu, Y. Niu, S. Gong, M. Xu, T. J. Meyer and Z. Chen, Selective CO2 Reduction to Formate on a Zn-Based Electrocatalyst Promoted by Tellurium, Chem. Mater., 2022, 34, 6036–6047 CrossRef CAS .
  28. C. Zhang, C. Cao, Y. Zhang, X. Liu, J. Xu, M. Zhu, W. Tu and Y. F. Han, Unraveling the Role of Zinc on Bimetallic Fe5C2–ZnO Catalysts for Highly Selective Carbon Dioxide Hydrogenation to High Carbon α-Olefins, ACS Catal., 2021, 11, 2121–2133 CrossRef CAS .
  29. R. Zhang, Y. Xue, M. Ma, Y. Han and J. Tian, Cu–Bi Bimetallic Sulfides Loaded on Two-Dimensional Ti3C2Tx MXene for Efficient Electrocatalytic Nitrogen Reduction under Ambient Conditions, Nano Lett., 2024, 24, 10297–10304 CrossRef CAS .
  30. N. Wang, Z. Liu, J. Ma, J. Liu, P. Zhou, Y. Chao, C. Ma, X. Bo, J. Liu, Y. Hei, Y. Bi, M. Sun, M. Cao, H. Zhang, F. Chang, H. L. Wang, P. Xu, Z. Hu, J. Bai, H. Sun, G. Hu and M. Zhou, Sustainability perspective-oriented synthetic strategy for zinc single-atom catalysts boosting electrocatalytic reduction of carbon dioxide and oxygen, ACS Sustainable Chem. Eng., 2020, 8, 13813–13822 CrossRef CAS .
  31. A. Kareem, H. Mohanty, K. Thenmozhi, S. Pitchaimuthu and S. Senthilkumar, Trimetallic Zn–Co–Ni selenide nanoparticles as electrocatalysts for the hydrogen evolution reaction, ACS Appl. Nano Mater., 2024, 7, 4886–4894 CrossRef CAS .
  32. G. Rajeshkhanna, S. Kandula, K. R. Shrestha, N. H. Kim and J. H. Lee, A New Class of Zn1−xFex–Oxyselenide and Zn1−xFex–LDH Nanostructured Material with Remarkable Bifunctional Oxygen and Hydrogen Evolution Electrocatalytic Activities for Overall Water Splitting, Small, 2018, 14, 1803638 CrossRef .
  33. X. Zhao, X. He, B. Chen, F. Yin and G. Li, MOFs derived metallic cobalt-zinc oxide@ nitrogen-doped carbon/carbon nanotubes as a highly-efficient electrocatalyst for oxygen reduction reaction, Appl. Surf. Sci., 2019, 487, 1049–1057 CrossRef CAS .
  34. C. Xing, J. Ren, L. Fan, J. Zhang, M. Ma, S. Wu, Z. Liu and J. Tian, π-d Conjugated Copper Chloranilate with Distorted Cu-O4 Site for Efficient Electrocatalytic Ammonia Production, Adv. Funct. Mater., 2024, 34, 2409064 CrossRef CAS .
  35. J. T. Ren, L. Chen, H. Y. Wang and Z. Y. Yuan, Aqueous Rechargeable Zn–N2 Battery Assembled by Bifunctional Cobalt Phosphate Nanocrystals-Loaded Carbon Nanosheets for Simultaneous NH3 Production and Power Generation, ACS Appl. Mater. Interfaces, 2021, 13, 12106–12117 CrossRef CAS PubMed .
  36. J. Wang, Z. Zhu, X. Wei, Z. Li, J. S. Chen, R. Wu and Z. Wei, Hydrogen-Mediated Synthesis of 3D Hierarchical Porous Zinc Catalyst for CO2 Electroreduction with High Current Density, J. Phys. Chem. C, 2021, 125, 23784–23790 CrossRef CAS .
  37. M. Zabilskiy, V. L. Sushkevich, M. A. Newton and J. A. Van Bokhoven, Copper–zinc alloy-free synthesis of methanol from carbon dioxide over Cu/ZnO/faujasite, ACS Catal., 2020, 10, 14240–14244 CrossRef CAS .
  38. X. Zheng, T. Ahmad and W. Chen, Challenges and strategies on Zn electrodeposition for stable Zn-ion batteries, Energy Storage Mater., 2021, 39, 365–394 CrossRef .
  39. C. Li, X. Xie, S. Liang and J. Zhou, Issues and future perspective on zinc metal anode for rechargeable aqueous zinc-ion batteries, Energy Environ. Mater., 2020, 3, 146–159 CrossRef CAS .
  40. S. Wahl, S. M. El-Refaei, A. G. Buzanich, P. Amsalem, K.-S. Lee, N. Koch, M.-L. Doublet and N. Pinna, Zn0.35Co0.65O – A Stable and Highly Active Oxygen Evolution Catalyst Formed by Zinc Leaching and Tetrahedral Coordinated Cobalt in Wurtzite Structure, Adv. Energy Mater., 2019, 9, 1900328 CrossRef .
  41. M. H. Naveen, T. L. Bui, L. Lee, R. Khan, W. Chung, R. Thota, S.-W. Joo and J. H. Bang, Nanostructuring Matters: Stabilization of Electrocatalytic Oxygen Evolution Reaction Activity of ZnCo2O4 by Zinc Leaching, ACS Appl. Mater. Interfaces, 2022, 14, 15165–15175 CrossRef CAS PubMed .
  42. S. Yuan, Z. Pu, H. Zhou, J. Yu, I. S. Amiinu, J. Zhu, Q. Liang, J. Yang, D. He, Z. Hu, G. Van Tendeloo and S. Mu, A universal synthesis strategy for single atom dispersed cobalt/metal clusters heterostructure boosting hydrogen evolution catalysis at all pH values, Nano Energy, 2019, 59, 472–480 CrossRef CAS .
  43. B. Singh and A. Indra, Designing Self-Supported Metal-Organic Framework Derived Catalysts for Electrochemical Water Splitting, Chem. – Asian J., 2020, 15, 607–623 CrossRef CAS .
  44. B. Singh, R. Kumar and A. Draksharapu, Correlating structure-activity-stability relationship of high-valent 3d-metal-based MOFs and MOF-derived materials for electrochemical energy conversion and storage, Coord. Chem. Rev., 2025, 523, 216239 CrossRef CAS .
  45. A. Zhao, L. Zhang, G. Xu, X. Zhang, S. Zhang and Y. Xiao, Hollow ZnxCo1-xSe2 microcubes derived from Metal–Organic framework as efficient bifunctional electrocatalysts for hydrogen evolution and oxygen evolution reactions, Int. J. Hydrogen Energy, 2020, 45, 2607–2616 CrossRef CAS .
  46. K. Y. Kumar, L. Parashuram, M. K. Prashanth, H. Shanavaz, C. B. P. Kumar, V. S. A. Devi, F. Alharethy, B.-H. Jeon and M. S. Raghu, Tailoring the bandgap of zinc indium sulfide/boroncarbonitride heterostructure for efficient photocatalytic CO2 reduction, J. Environ. Chem. Eng., 2023, 11, 110867 CrossRef CAS .
  47. H. Nam, K. Ryeol Park, Y.-W. Choi, H. Sim, K. Yong Sohn and D.-H. Lim, Electrocatalytic CO2 reduction using self-supported zinc sulfide arrays for selective CO production, Appl. Surf. Sci., 2023, 612, 155646 CrossRef CAS .
  48. C. Zhang, R. Lu, C. Liu, J. Lu, Y. Zou, L. Yuan, J. Wang, G. Wang, Y. Zhao and C. Yu, Trimetallic sulfide hollow superstructures with engineered D-band center for oxygen reduction to hydrogen peroxide in alkaline solution, Adv. Sci., 2022, 9, 2104768 CrossRef CAS .
  49. D. Lim, K. Min, M. Hwang, H. C. Ham, G.-J. Kim and S.-H. Baeck, Hollow hierarchical zinc cobalt sulfides derived from bimetallic-organic-framework as a non-precious electrocatalyst for oxygen reduction reaction, Mol. Catal., 2021, 509, 111614 CrossRef CAS .
  50. L. Hu, P. Zhong, X. Zhang, Y. Xiang, M.-S. Balogun, Y. Tong and H. Yang, Electronic modulation of zinc selenide toward efficient alkaline hydrogen evolution, Appl. Surf. Sci., 2023, 623, 157040 CrossRef CAS .
  51. C. Dai, R. Li, H. Guo, S. Liang, H. Shen, T. Thomas and M. Yang, Nitrogen, sulfur co-doped carbon coated zinc sulfide for efficient hydrogen peroxide electrosynthesis, Dalton Trans., 2021, 50, 5416–5419 RSC .
  52. Y. Chen, S. Gong, Y. Zhang, L. Li, Y. Wang, X. Tan, L. Zhang, X. Guo, X. Lin and L. Hu, Bimetallic MOF-derived ZnSe/NiSe heterostructures toward enhanced hydrogen evolution reactions, Inorg. Chem. Commun., 2022, 142, 109587 CrossRef CAS .
  53. S. Liang, G. He, G. Sui, J. Li, D. Guo, Z. Luo, R. Xu, H. Yao, C. Wang and Z. Xing, ZIF-L-derived C-doped ZnO via a two-step calcination for enhanced photocatalytic hydrogen evolution, J. Mol. Struct., 2023, 1276, 134787 CrossRef CAS .
  54. V. M. Sofianos, J. Lee, D. S. Silvester, P. K. Samanta, M. Paskevicius, N. J. English and C. E. Buckley, Diverse morphologies of zinc oxide nanoparticles and their electrocatalytic performance in hydrogen production, J. Energy Chem., 2021, 56, 162–170 CrossRef CAS .
  55. Z.-Y. Yu, Y. Duan, X.-Y. Feng, X. Yu, M.-R. Gao and S.-H. Yu, Clean and affordable hydrogen fuel from alkaline water splitting: past, recent progress, and future prospects, Adv. Mater., 2021, 33, 2007100 CrossRef CAS .
  56. C.-C. Weng, J.-T. Ren and Z.-Y. Yuan, Transition metal phosphide-based materials for efficient electrochemical hydrogen evolution: a critical review, ChemSusChem, 2020, 13, 3357–3375 CrossRef CAS PubMed .
  57. B. Singh and A. Indra, Prussian blue-and Prussian blue analogue-derived materials: progress and prospects for electrochemical energy conversion, Mater. Today Energy, 2020, 16, 100404 CrossRef .
  58. G. M. Di Mari, M. C. Spadaro, F. Salutari, J. Arbiol, L. Bruno, G. Mineo, E. Bruno, V. Strano and S. Mirabella, Low-cost, high-yield zinc oxide-based nanostars for alkaline overall water splitting, ACS Omega, 2023, 8, 37023–37031 CrossRef CAS PubMed .
  59. M. P. Kumar, N. Kumaresan, R. V. Mangalaraja, I. Zaporotskova, A. Arulraj, G. Murugadoss and A. Pugazhendhi, Zinc oxide nanoflakes supported copper oxide nanosheets as a bifunctional electrocatalyst for OER and HER in an alkaline medium, Environ. Res., 2024, 252, 119030 CrossRef CAS .
  60. P. Ilanchezhiyan, G. Mohan Kumar, S. Tamilselvan, T. W. Kang and D. Y. Kim, Highly efficient overall water splitting performance of gadolinium-indium-zinc ternary oxide nanostructured electrocatalyst, Int. J. Energy Res., 2020, 44, 6819–6827 CrossRef CAS .
  61. X. Li, J. Huang, Z. Liu, Q. Chen, G. Chen, Y. Zhang, K. Kajiyoshi, Y. Zhao, Y. Liu, L. Cao and L. Feng, Electronic modulation of CoP nanosheets array by Zn doping as an efficient electrocatalyst for overall water splitting, Catal. Sci. Technol., 2023, 13, 6550–6560 RSC .
  62. C. D. Nguyen, T. L. M. Pham, T. Y. Vu, V. B. Mai and K. L. Vu-Huynh, Hierarchical Zn-Co-P nanoneedle arrays supported on three-dimensional framework as efficient electrocatalysts for hydrogen evolution reaction in alkaline condition, J. Electroanal. Chem., 2020, 858, 113803 CrossRef CAS .
  63. Z. C. Cai, A. P. Wu, H. J. Yan, C. G. Tian, D. Z. Guo and H. G. Fu, Zn-Doped porous CoNiP nanosheet arrays as efficient and stable bifunctional electrocatalysts for overall water splitting, Energy Technol., 2020, 8, 1901079 CrossRef CAS .
  64. X. Wang, Y. Xie, W. Zhou, X. Wang, Z. Cai, Z. Xing, M. Li and K. Pan, The self-supported Zn-doped CoNiP microsphere/thorn hierarchical structures as efficient bifunctional catalysts for water splitting, Electrochim. Acta, 2020, 339, 135933 CrossRef CAS .
  65. J. Li, Y. Xu, L. Liang, R. Ge, J. Yang, B. Liu, J. Feng, Y. Li, J. Zhang, M. Zhu, S. Li and W. Li, Metal-organic frameworks-derived nitrogen-doped carbon with anchored dual-phased phosphides as efficient electrocatalyst for overall water splitting, Sustainable Mater. Technol., 2022, 33, e00421 CrossRef .
  66. K. T. N. Le, V. H. Hoa, H. T. Le, D. T. Tran, N. H. Kim and J. H. Lee, Multi-interfacial engineering of IrOx clusters coupled porous zinc phosphide-zinc phosphate heterostructure for efficient water splitting, Appl. Surf. Sci., 2022, 600, 154206 CrossRef CAS .
  67. Z. Liu, H. Li, Y. Shao, P. Ge, H. Tan and K. Chen, Structural design of electrocatalytic electrodes of Zn-Ni-S/Ni-Co-P microspheres to improve stability and activity for supercapacitors and water splitting, Energy Fuels, 2023, 37, 614–623 CrossRef CAS .
  68. B. Singh and A. Indra, Surface and interface engineering in transition metal–based catalysts for electrochemical water oxidation, Mater. Today Chem., 2020, 16, 100239 CrossRef CAS .
  69. B. Singh, A. Singh, A. Yadav and A. Indra, Modulating electronic structure of metal-organic framework derived catalysts for electrochemical water oxidation, Coord. Chem. Rev., 2021, 447, 214144 CrossRef CAS .
  70. B. Singh, Y. Arya, G. K. Lahiri and A. Indra, Electronic structure modulation in Prussian blue and its analogs: Progress and challenges in perspective of energy-related catalysis, Coord. Chem. Rev., 2025, 523, 216288 CrossRef CAS .
  71. M. F. Iqbal, Y. Yang, M. U. Hassan, X. Zhang, G. Li, K. N. Hui, M. Esmat and M. Zhang, Polyaniline grafted mesoporous zinc sulfide nanoparticles for hydrogen evolution reaction, Int. J. Hydrogen Energy, 2022, 47, 6067–6077 CrossRef CAS .
  72. H. Pan, Y. Wang, Z. Lu, X. Huang and X. Chen, Free-standing Co/Zn sulfide supported on Cu-foam for efficient overall water splitting, New J. Chem., 2022, 46, 11149–11157 RSC .
  73. J. Du, Z. Zou and C. Xu, Enhanced oxygen and hydrogen evolution reaction by zinc doping in cobalt–nickel sulfide heteronanorods, Electrochem. Sci. Adv., 2021, 1, e2000038 CrossRef CAS .
  74. J. Gautam, Y. Liu, J. Gu, Z. Ma, J. Zha, B. Dahal, L. N. Zhang, A. N. Chishti, L. Ni, G. Diao and Y. Wei, Fabrication of polyoxometalate anchored zinc cobalt sulfide nanowires as a remarkable bifunctional electrocatalyst for overall water splitting, Adv. Funct. Mater., 2021, 31, 2106147 CrossRef CAS .
  75. R. Balu, G. Devendrapandi, P. C. Karthika, O. H. Abd-Elkader, J. R. R, W. K. Kim, V. R. Minnam Reddy, S. Singh and M. Lavanya, Astonishing performance of zinc iron sulfide with MoS2 composite in allium-shaped structure for comprehensive alkaline water splitting, Int. J. Hydrogen Energy, 2024, 78, 492–501 CrossRef CAS .
  76. Y. Liang, Q. Liu, Y. Luo, X. Sun, Y. He and A. M. Asiri, Zn0.76Co0.24S/CoS2 nanowires array for efficient electrochemical splitting of water, Electrochim. Acta, 2016, 190, 360–364 CrossRef CAS .
  77. M. Mekete Meshesha, J. Gautam, D. Chanda, S. Gwon Jang and B. Lyong Yang, Enhancing the electrochemical activity of zinc cobalt sulfide via heterojunction with MoS2 metal phase for overall water splitting, J. Colloid Interface Sci., 2023, 652, 272–284 CrossRef CAS PubMed .
  78. G. Zhang, B. Hu, Y. Luo, Y. Xie, Y. Chen, Y. Zhang, Y. Ling, J. Zou and Y. Shao, ZnCo-ZIF derived CoSe2 on carbon nanotubes: A nanotubular catalyst for enhanced water splitting, Int. J. Hydrogen Energy, 2024, 53, 1226–1232 CrossRef CAS .
  79. M. Wu, Y. Huang, X. Cheng, X. Geng, Q. Tang, Y. You, Y. Yu, R. Zhou and J. Xu, Arrays of ZnSe/MoSe2 Nanotubes with Electronic Modulation as Efficient Electrocatalysts for Hydrogen Evolution Reaction, Adv. Mater. Interfaces, 2017, 4, 1700948 CrossRef .
  80. S. R. Stoyanov, A. V. Titov and P. Král, Transition metal and nitrogen doped carbon nanostructures, Coord. Chem. Rev., 2009, 253, 2852–2871 CrossRef CAS .
  81. K. Huang, W. Zhang, R. Devasenathipathy, Z. Yang, X. Zhang, X. Wang, D. H. Chen, Y. Fan and W. Chen, Co nanoparticles and ZnS decorated N, S co-doped carbon nanotubes as an efficient oxygen reduction catalyst in zinc-air batteries, Int. J. Hydrogen Energy, 2021, 46, 30090–30100 CrossRef CAS .
  82. C. Fu, X. Qi, L. Zhao, T. Yang, Q. Xue, Z. Zhu, P. Xiong, J. Jiang, X. An, H. Chen, J. S. Chen, A. Cabot and R. Wu, Synergistic cooperation between atomically dispersed Zn and Fe on porous nitrogen-doped carbon for boosting oxygen reduction reaction, Appl. Catal., B, 2023, 335, 122875 CrossRef CAS .
  83. J. E. Tsai, W. X. Hong, H. Pourzolfaghar, W. H. Wang and Y. Y. Li, A Fe-Ni-Zn triple single-atom catalyst for efficient oxygen reduction and oxygen evolution reaction in rechargeable Zn-air batteries, Chem. Eng. J., 2023, 460, 141868 CrossRef CAS .
  84. X. Lin, Q. Li, Y. Hu, Z. Jin, K. M. Reddy, K. Li, X. Lin, L. Ci and H. J. Qiu, Revealing Atomic Configuration and Synergistic Interaction of Single-Atom-Based Zn-Co-Fe Trimetallic Sites for Enhancing Oxygen Reduction and Evolution Reactions, Small, 2023, 19, 2300612 CrossRef CAS PubMed .
  85. G. Li, W. Deng, L. He, J. Wu, J. Liu, T. Wu, Y. Wang and X. Wang, Zn, Co, and Fe tridoped N–C core–shell nanocages as the high-efficiency oxygen reduction reaction electrocatalyst in zinc–air batteries, ACS Appl. Mater. Interfaces, 2021, 13, 28324–28333 CrossRef CAS PubMed .
  86. L. M. A. Monzon, K. Rode, M. Venkatesan and J. M. D. Coey, Electrosynthesis of iron, cobalt, and zinc microcrystals and magnetic enhancement of the oxygen reduction reaction, Chem. Mater., 2012, 24, 3878–3885 CrossRef CAS .
  87. K. Huang, C. Rong, W. Zhang, X. Yang, Y. Fan, L. Liu, Z. Yang, W. Chen and J. Yang, MOF-assisted synthesis of Ni, Co, Zn, and N multidoped porous carbon as highly efficient oxygen reduction electrocatalysts in Zn–air batteries, Mater. Today Energy, 2021, 19, 100579 CrossRef CAS .
  88. D. Liu, B. Wang, H. Li, S. Huang, M. Liu, J. Wang, Q. Wang, J. Zhang and Y. Zhao, Distinguished Zn, Co-Nx-C-Sy active sites confined in dentric carbon for highly efficient oxygen reduction reaction and flexible Zn-air Batteries, Nano Energy, 2019, 58, 277–283 CrossRef CAS .
  89. R. Jiang, X. Chen, W. Liu, T. Wang, D. Qi, Q. Zhi, W. Liu, W. Li, K. Wang and J. Jiang, Atomic Zn sites on N and S codoped biomass-derived graphene for a high-efficiency oxygen reduction reaction in both acidic and alkaline electrolytes, ACS Appl. Energy Mater., 2021, 4, 2481–2488 CrossRef CAS .
  90. J. Lu, W. Zhou, L. Wang, J. Jia, Y. Ke, L. Yang, K. Zhou, X. Liu, Z. Tang, L. Li and S. Chen, Core–Shell Nanocomposites Based on Gold Nanoparticle@Zinc–Iron-Embedded Porous Carbons Derived from Metal–Organic Frameworks as Efficient Dual Catalysts for Oxygen Reduction and Hydrogen Evolution Reactions, ACS Catal., 2016, 6, 1045–1053 CrossRef CAS .
  91. Y. Yang, W. Xiao, X. Feng, Y. Xiong, M. Gong, T. Shen, Y. Lu, H. D. Abruña and D. Wang, Golden palladium zinc ordered intermetallics as oxygen reduction electrocatalysts, ACS Nano, 2019, 13, 5968–5974 CrossRef CAS PubMed .
  92. X. Liu, L. Wang, G. Zhang, F. Sun, G. Xing, C. Tian and H. Fu, Zinc assisted epitaxial growth of N-doped CNTs-based zeolitic imidazole frameworks derivative for high efficient oxygen reduction reaction in Zn-air battery, Chem. Eng. J., 2021, 414, 127569 CrossRef CAS .
  93. T. Najam, S. Shoaib Ahmad Shah, M. Sufyan Javed, P. T. Chen, C. Chuang, A. Saad, Z. Song, W. Liu and X. Cai, Modulating the electronic structure of zinc single atom catalyst by P/N coordination and Co2P supports for efficient oxygen reduction in Zn-Air battery, Chem. Eng. J., 2022, 440, 135928 CrossRef CAS .
  94. A. Mahsud, M. Arif, A. Mehmood, M. Azam, X. Wang, Z. Yang and Z. Lu, Self-templated formation of ZnNC microtube with high oxygen reduction reaction overpotential for Zn-based catalyst, J. Environ. Chem. Eng., 2024, 12, 113616 CrossRef CAS .
  95. Z. Lu, B. Wang, Y. Hu, W. Liu, Y. Zhao, R. Yang, Z. Li, J. Luo, B. Chi, Z. Jiang, M. Li, S. Mu, S. Liao, J. Zhang and X. Sun, An isolated zinc–cobalt atomic pair for highly active and durable oxygen reduction, Angew. Chem., 2019, 131, 2648–2652 CrossRef .
  96. P. Song, M. Luo, X. Liu, W. Xing, W. Xu, Z. Jiang and L. Gu, Zn single atom catalyst for highly efficient oxygen reduction reaction, Adv. Funct. Mater., 2017, 27, 1700802 CrossRef .
  97. L. Ye, G. Chai and Z. Wen, Zn-MOF-74 derived N-doped mesoporous carbon as pH-universal electrocatalyst for oxygen reduction reaction, Adv. Funct. Mater., 2017, 27, 1606190 CrossRef .
  98. D. Deng, J. Qian, X. Liu, H. Li, D. Su, H. Li, H. Li and L. Xu, Non-covalent interaction of atomically dispersed Cu and Zn pair sites for efficient oxygen reduction reaction, Adv. Funct. Mater., 2022, 32, 2203471 CrossRef CAS .
  99. G. Chen, P. Liu, Z. Liao, F. Sun, Y. He, H. Zhong, T. Zhang, E. Zschech, M. Chen, G. Wu, J. Zhang and X. Feng, Zinc-Mediated Template Synthesis of Fe-N-C Electrocatalysts with Densely Accessible Fe-Nx Active Sites for Efficient Oxygen Reduction, Adv. Mater., 2020, 32, 1907399 CrossRef CAS .
  100. M. Liu, B. A. Lu, G. Yang, P. Yuan, H. Xia, Y. Wang, K. Guo, S. Zhao, J. Liu, Y. Yu, W. Yan, C. L. Dong, J. N. Zhang and S. Mu, Concave Pt–Zn nanocubes with high-index faceted Pt Skin as highly efficient oxygen reduction catalyst, Adv. Sci., 2022, 9, 2200147 CrossRef CAS PubMed .
  101. Q. Jin, C. Wang, Y. Guo, Y. Xiao, X. Tan, J. Chen, W. He, Y. Li, H. Cui and C. Wang, Axial Oxygen Ligands Regulating Electronic and Geometric Structure of Zn-N-C Sites to Boost Oxygen Reduction Reaction, Adv. Sci., 2023, 10, 2302152 CrossRef CAS .
  102. Y. Jia, Z. Xue, J. Yang, Q. Liu, J. Xian, Y. Zhong, Y. Sun, X. Zhang, Q. Liu, D. Yao and G. Li, Tailoring the electronic structure of an atomically dispersed zinc electrocatalyst: coordination environment regulation for high selectivity oxygen reduction, Angew. Chem., Int. Ed., 2022, 61, e202110838 CrossRef CAS PubMed .
  103. J. Zang, F. Wang, Q. Cheng, G. Wang, L. Ma, C. Chen, L. Yang, Z. Zou, D. Xie and H. Yang, Cobalt/zinc dual-sites coordinated with nitrogen in nanofibers enabling efficient and durable oxygen reduction reaction in acidic fuel cells, J. Mater. Chem. A, 2020, 8, 3686–3691 RSC .
  104. J. Xue, Y. Li and J. Hu, Nanoporous bimetallic Zn/Fe–N–C for efficient oxygen reduction in acidic and alkaline media, J. Mater. Chem. A, 2020, 8, 7145–7157 RSC .
  105. D. K. Yadav, R. Gupta, V. Ganesan, P. K. Sonkar and M. Yadav, Gold nanoparticles incorporated in a zinc-based metal-organic framework as multifunctional catalyst for the oxygen reduction and hydrogen evolution reactions, ChemElectroChem, 2018, 5, 2612–2619 CrossRef CAS .
  106. H. Li, S. Di, P. Niu, S. Wang, J. Wang and L. Li, A durable half-metallic diatomic catalyst for efficient oxygen reduction, Energy Environ. Sci., 2022, 15, 1601–1610 RSC .
  107. J. Xu, S. Lai, D. Qi, M. Hu, X. Peng, Y. Liu, W. Liu, G. Hu, H. Xu, F. Li, C. Li, J. He, L. Zhuo, J. Sun, Y. Qiu, S. Zhang, J. Luo and X. Liu, Atomic Fe-Zn dual-metal sites for high-efficiency pH-universal oxygen reduction catalysis, Nano Res., 2021, 14, 1374–1381 CrossRef CAS .
  108. L. Wen, X. Li, R. Zhang, H. Liang, Q. Zhang, C. Su and Y. J. Zeng, Oxygen Vacancy Engineering of MOF-Derived Zn-Doped Co3O4 Nanopolyhedrons for Enhanced Electrochemical Nitrogen Fixation, ACS Appl. Mater. Interfaces, 2021, 13, 14181–14188 CrossRef CAS .
  109. Z. Liang, C. Liu, M. Chen, M. Luo, X. Qi, S. G. Peera and T. Liang, Theoretical screening of di-metal atom (M = Fe, Co, Ni, Cu, Zn) electrocatalysts for ammonia synthesis, Int. J. Hydrogen Energy, 2020, 45, 31881–31891 CrossRef CAS .
  110. M. Ma, X. Han, H. Li, X. Zhang, Z. Zheng, L. Zhou, J. Zheng, Z. Xie, Q. Kuang and L. Zheng, Tuning electronic structure of PdZn nanocatalyst via acid-etching strategy for highly selective and stable electrolytic nitrogen fixation under ambient conditions, Appl. Catal., B, 2020, 265, 118568 CrossRef CAS .
  111. Y. Sun, W. Fan, Y. Li, N.-L.-D. Sui, Z. Zhu, Y. Zhou and J.-M. Li, Tuning Coordination Structures of Zn Sites Through Symmetry-Breaking Accelerates Electrocatalysis, Adv. Mater., 2024, 36, 2306687 CrossRef CAS PubMed .
  112. D.-M. Feng, Y. Sun, Z.-Y. Yuan, Y. Fu, B. Jia, H. Li and T. Ma, Ampoule method fabricated sulfur vacancy-rich N-doped ZnS electrodes for ammonia production in alkaline media, Mater. Renewable Sustainable Energy, 2021, 10, 8 CrossRef .
  113. J. Zhao, X. Liu, X. Ren, X. Sun, D. Tian, Q. Wei and D. Wu, Defect-rich ZnS nanoparticles supported on reduced graphene oxide for high-efficiency ambient N2-to-NH3 conversion, Appl. Catal., B, 2021, 284, 119746 CrossRef CAS .
  114. X. Chen, Y.-T. Liu, C. Ma, J. Yu and B. Ding, Self-organized growth of flower-like SnS2 and forest-like ZnS nanoarrays on nickel foam for synergistic superiority in electrochemical ammonia synthesis, J. Mater. Chem. A, 2019, 7, 22235–22241 RSC .
  115. T. Zhang, W. Zong, Y. Ouyang, Y. Wu, Y.-E. Miao and T. Liu, Carbon fiber supported binary metal sulfide catalysts with multi-dimensional structures for electrocatalytic nitrogen reduction reactions over a wide pH range, Adv. Fiber Mater., 2021, 3, 229–238 CrossRef CAS .
  116. P. Huang, T. Fan, X. Ma, J. Zhang, Y. Zhang, Z. Chen and X. Yi, 3D Flower-Like Zinc Cobaltite for Electrocatalytic Reduction of Nitrate to Ammonia under Ambient Conditions, ChemSusChem, 2022, 15, e202102049 CrossRef CAS PubMed .
  117. S. Dong, A. Niu, K. Wang, P. Hu, H. Guo, S. Sun, Y. Luo, Q. Liu, X. Sun and T. Li, Modulation of oxygen vacancy and zero-valent zinc in ZnCr2O4 nanofibers by enriching zinc for efficient nitrate reduction, Appl. Catal., B, 2023, 333, 122772 CrossRef CAS .
  118. J. Ye, J. Du, A. Wang, Y. Yang, J. Zhu, W. Li, C. Wan, F. Yin, G. He and H. Chen, Reduced spinel oxide ZnCo2O4 with tetrahedral Co2+ sites for electrochemical nitrate reduction to ammonia and energy conversion, Chem. Eng. J., 2024, 498, 155354 CrossRef CAS .
  119. Z. Du, K. Yang, H. Du, B. Li, K. Wang, S. He, T. Wang and W. Ai, Facile and scalable synthesis of self-supported Zn-doped CuO nanosheet arrays for efficient nitrate reduction to ammonium, ACS Appl. Mater. Interfaces, 2023, 15, 5172–5179 CrossRef CAS .
  120. S. Yang, L. Wang, X. Jiao and P. Li, Electrochemical reduction of nitrate on different Cu-Zn oxide composite cathodes, Int. J. Electrochem. Sci., 2017, 12, 4370–4383 CrossRef CAS .
  121. J. Zhao, X. Ren, X. Liu, X. Kuang, H. Wang, C. Zhang, Q. Wei and D. Wu, Zn single atom on N-doped carbon: Highly active and selective catalyst for electrochemical reduction of nitrate to ammonia, Chem. Eng. J., 2023, 452, 139533 CrossRef CAS .
  122. A. Aristizábal, M. Kolafa, S. Contreras, M. Domínguez, J. Llorca, N. Barrabés, D. Tichit and F. Medina, Catalytic activity and characterization of Pt/calcined CuZnAl hydrotalcites in nitrate reduction reaction in water, Catal. Today, 2011, 175, 370–379 CrossRef .
  123. L. Wang, X. Lv, P. Geng, Z. Wang and J. Sun, Cu and Zn metal particles modified nanocathode based electrocatalytic nitrate reduction: Effective, selective and mechanism, J. Electroanal. Chem., 2024, 954, 118053 CrossRef CAS .
  124. T. S. Bui, Z. Ma, J. A. Yuwono, P. V. Kumar, G. E. P. O’Connell, L. Peng, Y. Yang, M. Lim, R. Daiyan, E. C. Lovell and R. Amal, Enhanced Nitrate-to-Ammonia Activity on Fe/ZnO Nanoparticles via Tuning Intermediate Adsorption in Alkaline Electrolyte, Adv. Funct. Mater., 2024, 34, 2408704 CrossRef CAS .
  125. A. Feng, Y. Hu, X. Yang, H. Lin, Q. Wang, J. Xu, A. Liu, G. Wu and Q. Li, ZnO Nanowire Arrays Decorated with Cu Nanoparticles for High-Efficiency Nitrate to Ammonia Conversion, ACS Catal., 2024, 14, 5911–5923 CrossRef CAS .
  126. L. Wu, J. Feng, L. Zhang, S. Jia, X. Song, Q. Zhu, X. Kang, X. Xing, X. Sun and B. Han, Boosting electrocatalytic nitrate-to-ammonia via tuning of N-intermediate adsorption on a Zn–Cu catalyst, Angew. Chem., Int. Ed., 2023, 62, e202307952 CrossRef CAS PubMed .
  127. I. Stamatelos, C. T. Dinh, W. Lehnert and M. Shviro, Zn-Based Catalysts for Selective and Stable Electrochemical CO2 Reduction at High Current Densities, ACS Appl. Energy Mater., 2022, 5, 13928–13938 CrossRef CAS .
  128. T. L. Chen, H. C. Chen, Y. P. Huang, S. C. Lin, C. H. Hou, H. Y. Tan, C. W. Tung, T. S. Chan, J. J. Shyue and H. M. Chen, In situ unraveling of the effect of the dynamic chemical state on selective CO2 reduction upon zinc electrocatalysts, Nanoscale, 2020, 12, 18013–18021 RSC .
  129. B. Qin, Y. Li, H. Fu, H. Wang, S. Chen, Z. Liu and F. Peng, Electrochemical Reduction of CO2 into Tunable Syngas Production by Regulating the Crystal Facets of Earth-Abundant Zn Catalyst, ACS Appl. Mater. Interfaces, 2018, 10, 20530–20539 CrossRef CAS .
  130. X. Jiang, F. Cai, D. Gao, J. Dong, S. Miao, G. Wang and X. Bao, Electrocatalytic reduction of carbon dioxide over reduced nanoporous zinc oxide, Electrochem. Commun., 2016, 68, 67–70 CrossRef CAS .
  131. S. Saedy, M. A. Newton, M. Zabilskiy, J. H. Lee, F. Krumeich, M. Ranocchiari and J. A. van Bokhoven, Copper–zinc oxide interface as a methanol-selective structure in Cu–ZnO catalyst during catalytic hydrogenation of carbon dioxide to methanol, Catal. Sci. Technol., 2022, 12, 2703–2716 RSC .
  132. D. Ren, B. S. H. Ang and B. S. Yeo, Tuning the Selectivity of Carbon Dioxide Electroreduction toward Ethanol on Oxide-Derived CuxZn Catalysts, ACS Catal., 2016, 6, 8239–8247 CrossRef CAS .
  133. L. Han, S. Song, M. Liu, S. Yao, Z. Liang, H. Cheng, Z. Ren, W. Liu, R. Lin, G. Qi, X. Liu, Q. Wu, J. Luo, H. L. Xin, X. Liu, Q. Wu, J. Luo and H. L. Xin, Stable and Efficient Single-Atom Zn Catalyst for CO2 Reduction to CH4, J. Am. Chem. Soc., 2020, 142, 12563–12567 CrossRef CAS PubMed .
  134. Y. H. Li, P. F. Liu, C. Li and H. G. Yang, Sharp-tipped zinc nanowires as an efficient electrocatalyst for carbon dioxide reduction, Chem. – Eur. J., 2018, 24, 15486–15490 CrossRef CAS .
  135. V. Deerattrakul, P. Dittanet, M. Sawangphruk and P. Kongkachuichay, CO2 hydrogenation to methanol using Cu-Zn catalyst supported on reduced graphene oxide nanosheets, J. CO2 Util., 2016, 16, 104–113 CrossRef CAS .
  136. Z. Chen, K. Mou, S. Yao and L. Liu, Zinc-Coordinated Nitrogen-Codoped Graphene as an Efficient Catalyst for Selective Electrochemical Reduction of CO2 to CO, ChemSusChem, 2018, 11, 2944–2952 CrossRef CAS .
  137. S. Zhang, X.-T. Gao, P.-F. Hou, T.-R. Zhang and P. Kang, Three-dimensional ordered porous N-doped carbon-supported accessible Ni-Nx active sites for efficient CO2 electroreduction, Rare Met., 2023, 42, 1800–1807 CrossRef .
  138. H. Hu, Y. Tang, Q. Hu, P. Wan, L. Dai and X. J. Yang, In-situ grown nanoporous Zn-Cu catalysts on brass foils for enhanced electrochemical reduction of carbon dioxide, Appl. Surf. Sci., 2018, 445, 281–286 CrossRef CAS .
  139. Q. H. Low, N. W. X. Loo, F. Calle-Vallejo and B. S. Yeo, Enhanced Electroreduction of Carbon Dioxide to Methanol Using Zinc Dendrites Pulse-Deposited on Silver Foam, Angew. Chem., 2019, 131, 2278–2282 CrossRef .
  140. S. Zhen, G. Zhang, D. Cheng, H. Gao, L. Li, X. Lin, Z. Ding, Z. J. Zhao and J. Gong, Nature of the Active Sites of Copper Zinc Catalysts for Carbon Dioxide Electroreduction, Angew. Chem., Int. Ed., 2022, 61, e202201913 CrossRef CAS .
  141. M. Zabilskiy, V. L. Sushkevich, M. A. Newton, F. Krumeich, M. Nachtegaal and J. A. van Bokhoven, Mechanistic Study of Carbon Dioxide Hydrogenation over Pd/ZnO-Based Catalysts: The Role of Palladium–Zinc Alloy in Selective Methanol Synthesis, Angew. Chem., Int. Ed., 2021, 60, 17053–17059 CrossRef CAS .
  142. G. Yin, H. Abe, R. Kodiyath, S. Ueda, N. Srinivasan, A. Yamaguchi and M. Miyauchi, Selective electro- or photo-reduction of carbon dioxide to formic acid using a Cu–Zn alloy catalyst, J. Mater. Chem. A, 2017, 5, 12113–12119 RSC .
  143. I. S. Kwon, T. T. Debela, I. H. Kwak, H. W. Seo, K. Park, D. Kim, S. J. Yoo, J. G. Kim, J. Park and H. S. Kang, Selective electrochemical reduction of carbon dioxide to formic acid using indium–zinc bimetallic nanocrystals, J. Mater. Chem. A, 2019, 7, 22879–22883 RSC .
  144. J. Zeng, T. Rino, K. Bejtka, M. Castellino, A. Sacco, M. A. Farkhondehfal, A. Chiodoni, F. Drago and C. F. Pirri, Coupled Copper–Zinc Catalysts for Electrochemical Reduction of Carbon Dioxide, ChemSusChem, 2020, 13, 4128–4139 CrossRef CAS PubMed .
  145. G. Keerthiga and R. Chetty, Electrochemical Reduction of Carbon Dioxide on Zinc-Modified Copper Electrodes, J. Electrochem. Soc., 2017, 164, H164–H169 CrossRef CAS .
  146. Y. Feng, Z. Li, H. Liu, C. Dong, J. Wang, S. A. Kulinich and X. Du, Laser-Prepared CuZn Alloy Catalyst for Selective Electrochemical Reduction of CO2 to Ethylene, Langmuir, 2018, 34, 13544–13549 CrossRef CAS PubMed .
  147. J. Kim, J. Y. Park, W. S. Cho, W. J. Dong, J. W. Lim, K. Kim, P. Fernández and J. L. Lee, Selective Electrochemical Conversion of Carbon Dioxide to Formic Acid on Oxide-Derived SnxZn Bimetallic Catalysts, ACS Appl. Energy Mater., 2022, 5, 11042–11051 CrossRef CAS .
  148. A. C. García, J. Moral-Vico, A. A. Markeb and A. Sánchez, Conversion of Carbon Dioxide into Methanol Using Cu–Zn Nanostructured Materials as Catalysts, Nanomaterials, 2022, 12, 999 CrossRef PubMed .
  149. X. Kang, Q. Zhu, X. Sun, J. Hu, J. Zhang, Z. Liu and B. Han, Highly efficient electrochemical reduction of CO2 to CH4 in an ionic liquid using a metal–organic framework cathode, Chem. Sci., 2016, 7, 266–273 RSC .
  150. Y. Wang, P. Hou, Z. Wang and P. Kang, Zinc Imidazolate Metal–Organic Frameworks (ZIF-8) for Electrochemical Reduction of CO2 to CO, ChemPhysChem, 2017, 18, 3142–3147 CrossRef CAS PubMed .
  151. P. Shao, L. Yi, S. Chen, T. Zhou and J. Zhang, Metal-organic frameworks for electrochemical reduction of carbon dioxide: The role of metal centers, J. Energy Chem., 2020, 40, 156–170 CrossRef .
  152. J. Wu, T. Sharifi, Y. Gao, T. Zhang and P. M. Ajayan, Emerging Carbon-Based Heterogeneous Catalysts for Electrochemical Reduction of Carbon Dioxide into Value-Added Chemicals, Adv. Mater., 2019, 31, 1804257 CrossRef PubMed .
  153. W. Dai, M. Zou, J. Long, B. Li, S. Zhang, L. Yang, D. Wang, P. Mao, S. Luo and X. Luo, Nanoporous N-doped Carbon/ZnO hybrid derived from zinc aspartate: An acid-base bifunctional catalyst for efficient fixation of carbon dioxide into cyclic carbonates, Appl. Surf. Sci., 2021, 540, 148311 CrossRef CAS .
  154. Q. Yang, C. Yang, C. Lin and H. Jiang, Metal–Organic-Framework-Derived Hollow N-Doped Porous Carbon with Ultrahigh Concentrations of Single Zn Atoms for Efficient Carbon Dioxide Conversion, Angew. Chem., 2019, 131, 3549–3553 CrossRef .

This journal is © the Partner Organisations 2025
Click here to see how this site uses Cookies. View our privacy policy here.