Next Article in Journal
A Scalable High-Throughput Deposition and Screening Setup Relevant to Industrial Electrocatalysis
Next Article in Special Issue
Tafel Slope Analyses for Homogeneous Catalytic Reactions
Previous Article in Journal
Rendering Visible-Light Photocatalytic Activity to Undoped ZnO via Intrinsic Defects Engineering
Previous Article in Special Issue
Ir-Sn-Sb-O Electrocatalyst for Oxygen Evolution Reaction: Physicochemical Characterization and Performance in Water Electrolysis Single Cell with Solid Polymer Electrolyte
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Transition Metal Carbide Electrocatalysts for Oxygen Evolution Reaction

College of Environmental Science and Engineering, Nankai University, 94 Weijin Road, Tianjin 300071, China
*
Authors to whom correspondence should be addressed.
The four first authors share same contribution to this article.
Catalysts 2020, 10(10), 1164; https://doi.org/10.3390/catal10101164
Submission received: 15 September 2020 / Revised: 3 October 2020 / Accepted: 5 October 2020 / Published: 11 October 2020
(This article belongs to the Special Issue Electrocatalytic Water Oxidation)

Abstract

:
The electrolysis of water is considered to be a primary method for the mass production of hydrogen on a large scale, as a substitute for unsustainable fossil fuels in the future. However, it is highly restricted by the sluggish kinetics of the four-electron process of the oxygen evolution reaction (OER). Therefore, there is quite an urgent need to develop efficient, abundant, and economical electrocatalysts. Transition metal carbides (TMCs) have recently been recognized as promising electrocatalysts for OER due to their excellent activity, conductivity, and stability. In this review, widely-accepted evaluation parameters and measurement criteria for different electrocatalysts are discussed. Moreover, five sorts of TMC electrocatalysts—including NiC, tungsten carbide (WC), Fe3C, MoC, and MXene—as well as their hybrids, are researched in terms of their morphology and compounds. Additionally, the synthetic methods are summarized. Based on the existing materials, strategies for improving the catalytic ability and new designs of electrocatalysts are put forward. Finally, the future development of TMC materials is discussed both experimentally and theoretically, and feasible modification approaches and prospects of a reliable mechanism are referred to, which would be instructive for designing other effective noble-free electrocatalysts for OER.

1. Introduction

With the development of science and technology, the contradiction between the increasing demand for resources and the limitations of traditional energy has become increasingly prominent. Hydrogen energy has the characteristics of a high energy density and safe by-products, so it is an ideal substitute for carbon-based fuel [1,2,3,4,5]. In recent years, researchers have been trying to find technologies for large-scale hydrogen production, but almost all of these methods, including the alkaline water splitting reaction, metal–air battery, and fuel cell, inevitably rely on a dual electrode system. Generally speaking, this kind of system consists of two parts: The oxygen reduction reaction (ORR: O2 + 4e + 4H+ → 2H2O or O2 + 4e + 2H2O → 4OH) or hydrogen evolution reaction (HER: 4H+ + 4e → 2H2 or 4H2O + 4e → 2H2 + 4OH) that occurs in the cathode part, and the oxygen evolution reaction (OER: 2H2O → 4H+ + 4e + O2 or 4OH → 2H2O + 4e + O2) that takes place in the anode part.
Restricted by the slow four-electron transfer process [6,7,8], OER is more likely to block the total reaction than HER, which is merely a two-electron reaction [9]. The sluggish kinetic process of OER requires a higher potential to cross the energy barrier, thus greatly increasing the cost of the reaction. As a matter of fact, electrocatalysis is considered to be the most effective way to accelerate the kinetic process and reduce the cost [10,11]. However, the original electrocatalysts were noble metals, such as platinum and ruthenium, especially their oxides, which had a satisfactory catalytic effect, but could not be widely used due to their high cost [12,13,14].
Currently, transition metal carbides (TMCs), which have served as typical representatives of burgeoning non-metal electrocatalysts, have attracted much attention due to their high electrical conductivity, satisfactory mechanical strength and hardness, and excellent stability. After much effort over a long period of time [15], researchers have gradually realized that TMCs and their rich hybrid materials exhibit excellent OER performances. For example, Tang et al. [16] successfully synthesized an Mo/Co carbide electrocatalyst, which was applied to OER and displayed a low overpotential and good stability, showing advantages of hybrid electrocatalysts over traditional ones. Very recently, the emerging nanotechnology improves the practical application of TMCs in OER because it helps to overcome the large energy loss caused by the high-temperature calcination synthesis method and is extremely beneficial for the constructing of surface morphology [17]. Therefore, TMC electrocatalysts with a composite morphology perform a good application prospect in the OER field. Benchakar et al. reported a TMC electrocatalyst fabricated by coupling cobalt layered double hydroxide (Co-LDH) on Ti3C2Tx MXene, and it exhibited promising catalytic activity for OER [18]. Munir et al. [19] synthesized Ni/Ni3C nanoparticles (NPs) encapsulated in N-enriched carbon nanotubes (CNTs) and they demonstrated that the exceptional activity for OER were attributed to the synergistic effect between Ni/Ni3C and CNTs. All the above examples support that hybridization and composite morphology play quite an essential role for OER in recent studies [20,21]. Therefore, the rapid development of TMC electrocatalytic materials urgently needs to be comprehensively reviewed in light of recent progress. Wang et al. [22] briefly summarized the synthesis and application of TMCs. At the same time, they paid full attention to the hybridization of TMC materials, which was proved to be one of the decisive factors affecting the catalytic performance. Unfortunately, the research on TMC electrocatalysts is still very limited.
In this paper, TMCs are classified into five sections—NiC, WC, Fe3C, MoC, and MXene—according to their intrinsic properties and main structures. The mechanism analysis, designed synthesis methods, and OER performance of TMCs are discussed in detail. Several new developments of OER are also analyzed from both theoretical and experimental aspects. In the second section, the basic parameters and measurement standards for evaluating electrocatalysts are introduced. In the third part, application examples of the above five types of electrocatalysts and their hybrids in the OER field are described at length. It is worth mentioning that MXene material has been relatively less studied, but it shows a good application prospect of OER, so this paper mainly introduces it. Finally, we make a brief summary of the whole research field, and put forward new views on the material synthesis, morphology and structure, hybrid method, theoretical development, and mechanism in different conditions of OER electrocatalysts, in order to benefit future research.

2. Parameters and Mechanism for OER

Generally speaking, OER electrocatalysts are catalytic materials, which can be coated on the electrode surface or as the electrode itself. With high conductivity and special structure, the catalysts accelerate the sluggish four-charge transfer process by absorbing reactants on the electrode surface, which helps to overcome kinetic obstacles and reduce energy loss. Therefore, in order to objectively analyze the performance of electrocatalysts, we need to introduce a series of specific evaluation systems and parameters. To better explain the whole OER process and TMC catalysts, Section 2.1 briefly introduces the key parameters and Section 2.2 clarifies the measurement standards. Importantly, the following parameters constitute a comprehensive evaluation system of an electrocatalyst based on the standard evaluation conditions established by McCrory et al. [23].

2.1. Parameters and Measurement Criteria for OER

2.1.1. Overpotential (η)

Overpotential (η) is a basic parameter to evaluate the performance of electrocatalysts. Theoretically, according to Gibbs free energy formula (Equation (1)), the potential of OER can be acquired and it is 1.23 V vs. reversible hydrogen electrode (RHE). In this equation, ΔG denotes Gibbs free energy change, n implies the number of transferred electrons, and F and Eeq are the Faraday constant and standard electromotive force, respectively. It is essential to note that this thermodynamic potential (1.23 V) for OER is defined in the case of standard condition (25 °C and 1 atm) and equilibrium state. However, since the OER, which is a four electron-proton coupled reaction, requires more energy to overcome the kinetic barrier caused by electron transfer process, the reaction needs a working potential that is higher than the equilibrium value. Overpotential (η) is defined as the difference between the two potentials in Equation (2), in which E is the applied potential. If η = 0, E = Eeq, the reaction is at equilibrium condition and total net current is zero [24].
∆G = −nFEeq
η = E − Eeq
Therefore, the overpotential (η) can be understood as the excess energy required to achieve a specific current density and the electrocatalysts with a lower overpotential are more suitable for OER. Additionally, the two most commonly used potential values as electrocatalyst evaluation criteria are the onset overpotential of OER (η0) and the overpotential when the current density is equal to 10 mA/cm210). Among them, the onset overpotential is the point that indicates a sudden increase of current density, while η10 implies that the commercially equivalent photoelectrochemical water splitting efficiency equals 12.3%. Therefore, it is appropriate to choose these two values as a widely accepted standard [25]. The main approach to calculating the overpotential of an electrocatalyst is to normalize the polarization curves of cyclic voltammetry (CV) and linear sweep voltammetry (LSV) to obtain the potential values under different exchange current densities, and the overpotential can be acquired by then subtracting the equilibrium potential.

2.1.2. Tafel Slope

In the assessment methods of the properties of electrocatalysts, the Tafel slope (b) plays an essential role. To better interpret the Tafel slope, the exchange current density (i0) will be discussed first. For a certain electrochemical reaction, the total current (j) is the sum of currents in both the anode (ja) and cathode (jc) (Equation (3)); more importantly, according to the famous Bulter–Volmer (B-V) equation (Equation (4)), αa and αc refer to the transfer coefficient in the anode and cathode, respectively. n denotes the number of electrons transferred in the reaction and F indicates the Faraday constant, whilst R and T represent the universal gas constant and absolute temperature, respectively. Moreover, j0 is the exchange current, which can be divided by the electrode area (A), so the exchange current density (i0) is obtained in Equation (5).
j = ja + jc
j = j0[exp(αanFη/RT) − exp(αcnFη/RT)]
j0/A = i0
The B-V Equation (Equation (4)) can be simplified to the form of Equation (6), which is originally put forward by Tafel as an empirical formula, in which a refers to the overpotential at a current density of 1 mA/cm2 and b is the Tafel slope.
η = a + blog(j)
Considering that the Tafel slope (b) represents “the potential rise corresponding to the increase of unit exchange current density”, it is obvious that a lower Tafel slope (b) indicates a higher current increasing efficiency, which refers to a more promising electrocatalyst. Tafel slope (b) is also an indicator clarifying the rate determining step (RDS) of the whole process of water splitting reaction, including Volmer mechanism, Volmer–Tafel mechanism, and Volmer–Heyrovsky mechanism [26]. The profound role of Tafel slope (b) in OER mechanism needs to be further studied.

2.1.3. Electrochemically Active Surface Area (ECSA)

It has to be pointed out that the electrochemically active surface area (ECSA), which can be measured by means of testing double-layer capacitance (Cdl), is an indispensable measurement parameter for a given electrocatalyst, especially for porous substrate materials. Through analysis of CV polarization curve, there is enough evidence to prove that in the non-Faradaic overpotential region (0.1 V window area near the open-circuit voltage region), all the current changes are because of the charge-discharge processes of double-layer capacitance (Cdl). The charging current (ic) varies at different scan rates (v), so Equation (7) is frequently utilized to calculate the value of Cdl; as the slope of ic-v curve equals to Cdl, the method is precise and practical. When Cdl is known, the magnitude of ECSA can be easily obtained with Equation (8). Cs is the specific capacitance of each sample, and for transition metals, Cs often ranges from 20 to 60 μF·cm−2 [27].
ic = vCdl
ECSA = Cdl/Cs
A relatively high ECSA, which certifies that there are sufficient coordination unsaturated atoms and exposing active sites in the catalyst, implicates a better catalytic performance.

2.1.4. Faradaic Efficiency (FE)

Faradaic efficiency (FE), which can be calculated by the ratio between the effective oxygen yield and the theoretical oxygen evolution, is a special parameter employed to evaluate the performance of electrocatalysts and implies the efficiency of electrons involved in an electrochemical reaction. However, FE is often not accounted for by researchers, who just assume it as 100%; as a result, the effect of some catalysts is overestimated, leading to the consequence that a proportion of the research is invalid. In fact, the FE plays an essential role in the assessment procedure as it indicates the ultimate effect of OER. In different cases, the catalytic performance of electrocatalysts may differ somewhat, especially when phase change and substance decomposition exist.

2.1.5. Electrochemical Impedance Spectroscopy (EIS)

Electrochemical impedance spectroscopy (EIS) is a safety disturbance technique that is often used to detect the internal process of an electrochemical system and measure the impedance of a battery in a certain frequency range. These data can determine the state of health (SOH) and state of charge (SOC) of the battery. In the case of its use for OER, it is also one important module of the electrochemical workstation and essential for the conductivity test of electrocatalysts. EIS test is significant for OER kinetics research. By measuring the position and size of semicircle region in EIS, the voltage of solution resistance loss can be obtained, and normalization of LSV curve can then be completed. Apart from that, EIS test also helps to analyze the impedance changes with specific morphological transformation, representing an assisted approach in material evaluation.

2.1.6. Stability

Apart from the above parameters, long-term stability, which has a close connection to the practical use of electrocatalysts, undoubtedly plays an important role in OER evaluation process. Most often, we test the stability of materials by accelerated CV detection or a long-time (>10 h) constant voltage (current) test, and there should be no obvious current falling occurring for acceptable electrocatalysts. For all kinds of electrocatalysts, stability is always a problem needs to be solved [28].

2.2. Mechanism for OER

As introduced in Section 2.1, the applied potential of OER is much higher than the theoretical one at equilibrium state, which is 1.23 V vs. RHE. The total overpotential consists of overpotential in anode (ηa) and cathode (ηc) and the Ohm polarization overpotential (ηo). Ohm polarization overpotential is caused by contact resistance of solution, and can be measured by EIS test mentioned in Section 2.1.5 and reduced by optimizing the design of electrolytic cells (Equation (9))
η = ηa + ηc + ηo
Therefore, ηa and ηc are the main factors which lead to a high applied potential, with the parameters above, we can assess the performance of electrocatalysts through the overpotential test. However, another problem then arises. For OER, the half reaction of water splitting reaction shows different reaction paths in acidic and alkaline conditions. As a result, in different measurement backgrounds, the experimental data, even for the same electrocatalysts, are totally diverse. In order to maintain the consistency of the experimental result of electrocatalysts fabricated in a wide variety of methods, McCrory et al. set a widely accepted standard for the testing environment considering various measurement techniques and conditions. They suggested that, for alkaline electrolytes, the medium is 1.0 M NaOH, while for acid electrolytes, 1.0 M H2SO4 is recommended. The main function of the media is to provide the reaction with the environment that electrons can transfer intermediates through [29].
The detailed mechanism of OER can be obtained through density functional theory (DFT). Yu et al. [30] studied the OER mechanism in acidic solution and proposed the four-step elementary reactions as follows, in which * denotes the active site on electrocatalysts.
H 2 O + * OH * + e + H +
OH * O * + e + H +
H 2 O + O * OOH * + e + H +
OOH * e + O 2 + H +
It is significant to note that in this mechanism model, metal dissolution and corrosion are not considered due to the stability of metal carbides. The author also indicates that MC2 (M refers to metal), with increased carbon content, behave better corrosion resistance towards acid solution. With the protection of carbon atoms, metals are prevented from being exposed to the surface, thus increasing the stability of electrocatalysts. Besides, the mechanism in neutral and alkaline media have been widely investigated as the equations below [31].
OH + * OH * + e
OH * + OH O * + H 2 O + e
O * + OH OOH * + e
OOH * + OH * + O 2 + H 2 O + e
Note that in the above reactions, chemical groups such as O, OH, and OOH, are absorbed on the active sites (*) of catalysts, including physical or chemical absorption with no radicals involved as many literatures reported [31,32,33,34,35,36]. Generally, the formation of intermediates O* (Equation (15)) and OOH* (Equation (16)) has been considered as RDS for OER, and previous literature has reported that TMCs are of great value to decrease overpotential for OER as it usefully accelerates RDS in kinetic process [32]. For instance, Ouyang et al. [33] found out that oxygen was favorable to Co sites adjacent to β-Mo2C, because the doping of β-Mo2C created a heterointerface and thus effectively lowered the energy barrier. In addition, the research of Wang et al. [34] suggests the similar conclusion. Their study revealed that Ni-Ni3C exhibited smaller Gibbs free energy change (1.815 eV) for the RDS than Ni (3.877 eV) and Ni3C (3.701 eV), indicating the important role of Ni3C, which efficiently accelerating kinetic process, thus deceasing the overpotential. Note that in different conditions, carbon atoms play diverse roles but similarly contributing to enhancing OER performance [35]. In the mechanism research, DFT calculation is essential, but the drawback of which is also clear. Since particles like Co and Fe3C may be oxidized at OER potential, the mechanism under harsh oxidative conditions is still not confirmed [36], including oxidation in the bulk or only on the surface of TMC electrocatalysts.
With the understanding of OER evaluation parameters and mechanism, we have lucid and unified criteria to assess the performance of electrocatalysts, which makes it possible to find promising catalysts for OER.

3. Transition Metal Carbide Electrocatalysts for OER

3.1. Synthesis Methods of TMCs

Firstly, the typical synthetic methods of TMCs need to be introduced. Syntheses of TMCs can be divided into template-free method and template-assisted method, both of which are based on a high-temperature carbothermal reaction. For example, in the template-free method, metals and carbon are usually heated directly above 800 °C. The difference is that in the template-assisted method, templates such as metal organic framework (MOF), are first made, after that, the synthesized templates are burned. The distinction between the two methods lies in the control of structure and composition, the template-free method has a poor control effect, while the template-assisted method has a better operability.
Through different synthesis methods, we can obtain various structures of electrocatalysts, such as nanodots, nanowires, two-dimensional (2D) layered structure, core–shell structure, and mosaic structure. A wide variety of examples will be discussed in detail based on experimental results as Table 1 displays.

3.2. Nickel Carbide

Among common transition metal carbides, nickel carbide has the advantages of a high cost efficiency and great electric conductivity. Therefore, it is suitable for the manufacturing of high-performance electrocatalyst for OER. In 1988, Zhao et al. publicized the fabrication of highly disordered hexagonal NiCx filaments by chemical vapor deposition (CVD) [54]. After that, hexagonal NiCx with other forms of nickel carbide constituted the nickel carbide family, and moreover, provided us with valuable alternative materials as promising HER and OER electrocatalysts.
One of the necessary factors for a good catalyst is a large surface area. Unless the material made of nickel carbide can reach a small particle size, its electrocatalytic performance will be reduced. In a way, smaller particle size means larger surface area, leading to more active sites during the process of reaction. Fadil et al. [55] reported atomically ordered Ni3C, synthesized in the form of NPs smaller than 10 nm. Yang et al. [48] reported a facile route for the controllable orientation-dependent crystal growth of high-index faceted dendritic hexagonal NiCx nanosheets on Ni-coated copper foil, with an optimized carbon content of 16.7 at %, by a mild electrodeposition approach (Figure 1). The material obtained was fully exposed {120} high-index facets, contributing to improving the mass/electron transport capability and fully exposed active sites during the OER process. Impressively, the as-prepared material (denoted d-NiC0.2NS/Ni/CF) exhibited remarkable catalytic activity (with overpotential of 121 and 228 mV at 10 mA·cm−2 for the HER and OER, respectively), simultaneously giving a nearly 100% Faradaic yield and affording a superior catalytic stability (beyond 100 h) as a bifunctional electrocatalyst for both the HER and OER in basic media.
Moreover, d-NiC0.2NS-coated copper foil (d-NiC0.2NS/CF) samples, Pt/C-loaded CF and RuO2/C-loaded CF (Pt/C/CF and RuO2/C/CF) were prepared and then tested separately for a comparison. From the results, owing to its dendritic nanosheet morphology, optimized 16.7 at % carbon content, and fully exposed high-index facets, d-NiC0.2NS/Ni/CF displayed an improved mass/electron transport capability and fully exposed active sites, resulting in its high performance in OER and HER.
Although nickel carbide has a good catalytic effect on OER, it is actually more used to dope other elements or form a composite structure, such as Fe-doped [56], nitrogen-doped [57,58], and metal phosphide-mixed structure [47]. These electrochemical catalysts, most of which are doped with iron, nitrogen, or other metals, exhibit better results. For example, Haosen Fan et al. developed Fe-doped Ni3C nanodots in N-doped carbon nanosheets [45]. In their experiment, by employing the method of carburization, 2D nickel-cyanide polymer precursors reacted and formed a structure in which Ni3C nanodots were dispersed in ultra-thin N-doped carbon nanosheets. After the carburization process in Ar/H2 atmosphere, 2D hybrid nanosheets with uniform Ni3C nanodots embedded in N-doped carbon nanosheets were successfully obtained. The hybrid nanosheets had the lateral size of about 200–300 nm and thickness of about 10 nm, with uniform dispersion of 3–5 nm Ni3C nanodots. When further doping heteroatoms of Fe in different amount into Ni3C nanodots, the test results indicated that the 2 at % Fe-doped Ni3C nanodot-based nanosheets (Fe-Ni3C-2%) showed outstanding HER and OER properties, with a low overpotential (292 mV for HER and 275 mV for OER in 1.0 M KOH) and small Tafel slopes (34.6 mV dec−1 for HER and 62 mV dec−1 for OER in 1.0 M KOH). In addition, among these three catalysts, pure Ni3C displayed less of an OER response, indicating that Fe doping can significantly improve the catalytic activity. The FeNi3C-2% catalyst also showed a small overpotential of 275 mV at an anode current density of 10 mA·cm−2, which is lower than those of pure Ni3C and Fe-Ni3C-5% samples.
In another example, Jaison Joy et al. reported the synthesis of nickel-incorporated nitrogen-doped graphene nanoribbon (Ni/NGNRs) through a facile solvothermal process [49]. After incorporating nitrogen functionalities, graphene nanoribbon with tunable nickel content functions efficiently facilitated the water oxidation reaction and accelerated the catalytic reactivity, exhibiting an overpotential of 380 mV with a Tafel slope of 60 mV dec−1 to sustain 10 mA·cm−2 in 1.0 M KOH.
In addition, nickel carbide is often used to form a composite structure with other substances, such as transition metal [59], as well as their oxides and carbides [60], and some of them have intensive electrocatalytic properties for OER. For instance, Shuo Geng et al. synthesized a three-dimensional (3D) heterostructure NiC/MoC/NiMoO4 electrocatalyst in a simple and feasible way [46]. As shown in Figure 2, NiMoO4/NF nanorod arrays were synthesized by hydrothermal method with the precursors of NiMoO4 nanorods and nickel foam (NF), which was the Ni source. Then, at a high annealing temperature, the reaction between NiMoO4 nanorods and dopamine accounted for the conversion of NiMoO4 to NiC/MoC, which led to the formation of NiC/MoC/NiMoO4 heterostructure nanorod arrays.
Through electrode performance test, the optimized NiC/MoC/NiMoO4 behaved a low overpotential of 280 mV at a current density of 10 mA·cm−2 in 1.0 M KOH for OER. Moreover, it showed quite low cell voltage of 1.52 V at 10 mA·cm−2 and remarkable stability for more than 20 h, demonstrating the superiority of the NiC/MoC/NiMoO4 heterostructure for OER. Besides, Qing Qin et al. [59] in situ synthesized Ni/Ni3C core/shell hierarchical nanospheres through an ionic liquid-assisted hydrothermal method at a relatively low temperature. Resulting from the high electrical conductivity, more exposed active sites, and core–shell interface effect, the material obtained exhibited superior OER performance with a low overpotential, small Tafel slope, and fantastic stability (with overpotential of 350 mV at 10 mA·cm−1 and Tafel slopes of 57.6 mV dec−1).
MOF is a crystalline material with a one or multi-dimensional network structure formed by coordination between metal ions and an organic complex. Due to the high surface area of MOFs, they serve as both homogeneous and heterogeneous catalytic sites. Furthermore, MOFs possess controlled pore structure. These peculiarities lead to the result that MOFs are widely applied in many fields, varying from industrial gas separation and storage [61] to heterogeneous catalysis [62] and sensing devices [63], including synthesis of high-performance electrocatalysts for OER. Jia et al. [64] prepared Co-doped Ni3C/Ni embedded in a carbon matrix (Co-Ni3C/Ni@C) by pyrolysis of bimetal-MOF (Co/Ni-MOF) under a flow of Ar/H2 at 350 °C. The synthesis of the precursor bimetal-MOFs is displayed in Figure 3. Uniform cubic Co/Ni-BTC was constructed through a simple method. After carbonization at 350 °C under a flow of Ar/H2 gas, Co-Ni3C/Ni@C was obtained.
For comparison, Ni3C/Ni@C was also prepared and then tested. As shown in Figure 4a, due to surface oxidation of the catalysts during OER, the oxidation peaks in the LSV curves appeared at ~1.35 V vs. RHE, meaning that Co-Ni3C/Ni@C showed a smaller onset overpotential (as introduced in Section 2.1.1, the onset overpotential can be obtained by means of using the potential where there is a sudden increase of exchange current density in the LSV curve to subtract 1.23 V) of 220 mV in comparison to Ni3C/Ni@C (290 mV) in the OER. Co-Ni3C/Ni@C also showed higher activity in the OER, reaching an overpotential of 325 mV at the current density of 10 mA/cm210 can be obtained by means of using the potential at the current density of 10 mA·cm−2 to subtract 1.23 V), which is close to IrO2 (330 mV) and lower than Ni3C/Ni@C (350 mV) (Figure 4b), indicating that Co-doping can significantly improve the catalytic activity. Besides, the Tafel plot (According to Equation (6), Tafel plot is the slope of logi-V curve, and a lower Tafel slope indicates better performance for OER) for Co-Ni3C/Ni@C (67.76 mV/dec) was much lower than Ni3C/Ni@C (112.45 mV/dec) and IrO2 (79.92 mV/dec), verifying the efficient reaction kinetics in the OER (Figure 4c). As can be seen in Figure 4d, the LSV curves showed small negative variation after 1000 cycles, indicating its high stability in OER.

3.3. Tungsten Carbide

Tungsten carbide (WC) was discovered to possess similar catalytic properties to those of platinum group metals for certain reactions by Levy and Boudart [65]. Therefore, it was considered as an attractive non-noble metal catalyst for proton exchange membrane (PEM) fuel cell applications. In general, there are two ways to synthesize tungsten carbide [66]. One of them is to first reduce a tungsten precursor to a reaction intermediate (usually tungsten powder). Then, the powder is carbonized at a high temperature in reductant atmosphere. The other way is to first obtain a highly active tungsten precursor (such as WO3). The precursor powder is usually carbonized in H2 and CH4 mixed atmosphere. The electrocatalytic performance of product WC is affected by many factors, such as source materials, activation process, and preparation method. Therefore, exploring new carbonizing approaches is a method worth considering for improving the performance.
Although the performance of WC is inferior in direct comparison to platinum, WC has received much attention due to its low price and insensitivity to catalyst poisons, such as H2S and CO [67,68]. Therefore, tungsten carbide became one of the most promising alternatives to noble metal electrocatalysts for the HER. Nevertheless, nothing has been reported on its application in the OER until the work of Han et al. [69] in 2018. In the experiment, they synthesized a superaerophobic nitrogen-doped tungsten carbide nanoarray electrode, exhibiting high stability and activity toward OER and effectively accelerating oxygen evolution in acid. In testing process, the performance of N-WC nanoarray toward OER was measured in 0.5 M H2SO4, and for a comparison, commercial 20 wt % Ir/C and IrO2 catalysts were measured. As displayed in Figure 5a, OER starts at about 1.35 V on the nanoarray electrode, while at about 1.5 V on commercial 20 wt % Ir/C and IrO2. The synchronized variation between voltage and O2 concentration in Figure 5b demonstrated that OER really occurs at about 1.35 V vs. RHE. In addition, the current density increases drastically with potential, reaching up to 60 mA·cm−2 at about only 1.7 V vs. RHE.
Except for its instability, N-WC nanoarray is a promising electrocatalyst for OER. The XRD patterns of the nanoarray before and after the OER (Figure 5d) indicate that a small amount of tungsten oxides structured on the nanoarray, resulting in an increase of potential after 1 h of water splitting at 10 mA·cm−2 (Figure 5c). Although its catalytic stability is not satisfactory, N-WC nanoarray provides a research direction for noble-metal-free electrocatalysts in OER.
Regarded as a promising candidate for applications as a non-noble metal electrocatalyst for HER, WC itself is not a suitable catalyst for OER due to its high overpotential, which consequently leads to low electrocatalytic activity. In recent years, most researchers have preferred to study WC as an electrocatalyst substrate combined or doped with other substances, such as cobalt [70] and nitrogen [71]. Most of them are widely used in HER and ORR, while their application in OER is seldom reported. From the experiment carried out by Han et al. [69], tungsten oxides formed on the electrode during the process, which hindered the OER. This may be a possible reason why WC exhibited low electrocatalytic activity toward OER.
With the use of MOFs (Co2+ and W6+ as metal ions, and DMF as a ligand), Tao Zhao et al. [72] produced a material composed of porous nitrogen-doped carbon materials encapsulating cobalt and tungsten carbide (Co/W-C@NCNSs), working as a bifunctional catalyst for overall water splitting. They synthesized nanoscale MOFs through a microwave method and the MOFs obtained were then annealed under N2 atmosphere at the temperature of 600 °C and 800 °C. Finally, the samples were obtained. The results show that, among the two samples obtained, Co/W-C@NCNSs (800 °C) has better electrocatalytic activities, with an overpotential of 323 mV at a current density of 10 mA·cm−2 in the OER. Although the overpotential is not low enough compared with some noble-metal-based electrocatalysts, it has great potential to become an efficient electrocatalyst toward OER.

3.4. Iron Carbide

Iron carbide is a kind of intermetallic compound formed by iron and carbon, owning high conductivity and good stability [73,74,75,76,77]. The chemical formula of iron carbide is Fe3C and it is usually applied for metal–air batteries [78,79,80,81]. The crystal structure of this material is a complex orthorhombic system, with high hardness, zero plasticity, and great brittleness. At present, the most common methods to prepare Fe3C are reduction carburization, thermal decomposition, laser and CVD. The as-obtained morphology of Fe3C is mainly powder, filled with carbon nanotubes or coated with carbon nanospheres. Fe3C has many excellent properties and has been widely studied. How to further improve its performance and expand its application range have gradually become the main points of research in recent years.
Due to the combination of nitrogen atoms in graphite structure, nitrogen-doped carbon materials can perform impressive ORR catalytic activity. In addition, Defilippi et al. [82] took advantage of the sol–gel technique and successfully synthesized HfN NPs (less than 15 nm in diameter), which exhibited a quite low overpotential of 358 mV at 10 mA·cm−2 and long term stability as an OER electrocatalyst. This recent study indicates that metal nitrides possess high conductivity, stability, and melting point, and these properties make them qualified to be active both for ORR and OER. It has been shown that the electrocatalytic ability can also be developed by combining transition metals and metal oxides with carbon nitride materials [83]. At present, graphite carbonitride (g-C3N4) with 2D layered structure has been widely developed for effective OER and ORR electrocatalysts [84]. However, the relatively low conductivity and surface area of g-C3N4 seriously limit its application in electrocatalysis [85]. Therefore, the simple method shown in Figure 6 can be used to prepare N3-doped graphene-encapsulated Fe3C/Fe2O3 heterostructures (named Fe3C/Fe2O3@NGNs) to make it a better catalyst [43].
The polarization curves can be utilized to study the electrocatalytic activity of synthetic samples for ORR and OER. As shown in Figure 7a, the reduction peak of the metal-free N-doped graphene catalyst was positively shifted after the Fe element had been incorporated, thus the as fabricated Fe3C/Fe2O3@NGNs displayed a halfwave potential of 0.86 V, which was the same as commercial Pt/C. Simultaneously, Fe3C/Fe2O3@NGNs, with an Tafel slope of 73 mV/dec, also outperformed NGNs and Pt/C, indicating the gratifying kinetic process for ORR (Figure 7b). As for OER, the electrocatalyst showed an overpotential of 1.69 V at 10 mA·cm−2 and a Tafel slope of 163 mV dec−1 (Figure 7c,d). As a matter of fact, the enhanced ORR and OER performance could be attributed to the combination of the Fe3C/Fe2O3 heterostructure, which provided more inherent active sites, thereby making this structure a promising dual-functional electrocatalyst.
Among many developed materials, Fe-N-C is considered a promising candidate. Especially in recent years, a large number of researchers have reported studies in which NPs, such as iron and/or its carbides, oxides, and nitride, are encapsulated in various N-doped carbon nano structures, such as CNTs, carbon layers, and carbon nanofibers. As a result, these modified materials exhibited satisfactory properties in some fields. Given the fact that some materials with a high surface area and large pore volume have poor electrical conductivity, conductive carbon materials such as graphene and CNTs must be introduced into them, in order to enhance catalytic ability. For instance, Zhao et al. [44] utilized porous, conductive NCNT/NPC hybrids, which possessed high density of Fe–N active sites, to fabricate a new catalyst named Fe3C@NCNT/NPC and it exhibited excellently in OER.
Wang et al. [86] reported a new type of electrocatalyst, the construction of which is based on a CNT framework, with NPs of Fe and carbon embedded in it (called Fe@C-NG/NCNTs). Therefore, it provides us with an idea that it is possible to construct an electrocatalyst as a ‘framework active site’ structure with a large number of Fe–Nx active sites. DE (the potential gap between OER and ORR) is usually measured to reflect bifunctional activity of ORR/OER and a lower DE indicates better performance [87,88,89]. As shown in Figure 8, according to the curve, the DE value of Fe@C–NG/NCNTs is estimated to be 0.84 V, which is better than Pt/C catalyst (1.05 V). Therefore, Fe@C–NG/NCNTs seems to be of great value for research on bifunctional catalysts for ORR and OER.
Carbon materials functionalized with transition metal-based inorganic NPs can effectively improve electrocatalysis. For example, FeCo@NC core-shell nanospheres, with the carrier graphene, can be used as efficient dual-functional OER electrocatalysts [87]. Besides, Cui et al. [90] fabricated a great bimetal carbide-based bifunctional catalyst consisting of iron-molybdenum carbide (Fe3Mo3C) and IrMn nanoalloys, which was not only chemically stable in alkaline medium, but also in acidic environment. Under alkaline condition, the active OH* can be provided by free OH in the solution, while in acidic condition, the OH* just derived from the ionization of water. As a result, the alkaline medium is a more suitable environment for OER, thus the electrocatalysts can be more effective in alkaline condition.

3.5. Molybdenum Carbide

Molybdenum carbide (Mo2C) has been extensively described as a low-cost catalyst for the HER [91,92,93], but, as for OER, it has not been reported until 2015 [50]. In a comparative experiment of transition metal carbides, Yagya N. Regmi et al. indicated that Mo2C is the most active HER catalyst [94]. In addition, Fe2N-type molybdenum carbide (β-Mo2C) showed the best performance in HER activity and the lowest overpotential among materials synthesized via an amine-metal oxide composite intermediate. Morphology and surface structure of electrocatalysts often play an important role in their catalytic activities, and these physical characteristics vary with the synthesis methods. From this perspective, Yagya N. Regmi et al. [50] dived deeper into their research and indicated that Fe2N-type Mo2C displayed bifunctional catalytic activities. Therefore, it is believed to act as an OER electrocatalyst.
Although Mo2C has shown great electrocatalytic performance in oxygen revolution activity, the stability still needs improving. An increasing number of researchers are studying how to convert it into nanomaterials, with an expectation of obtaining more efficient catalysts for OER. Due to excellent characteristics, graphite-like carbon nanomaterials (CNMs) have received great attention. Therefore, in the field of nanotechnology, the synthesis of CNMs is developing rapidly. Benefiting from high surface area, CNMs have been widely researched as a catalyst in electrochemical activities [95,96,97]. As a result, some researchers have been trying to apply CNMs in OER electrocatalysts, thus synthesizing molybdenum carbide coupled on carbon nanomaterials. For example, H. Wang et al. [51] designed highly crystalline Mo2C NPs supported on carbon sheets (Mo2@CS) as a bifunctional catalyst toward HER and OER. The materials were obtained through a one-pot process where Mo was sourced from ammonium molybdate and carbon was from glucose (Figure 9).
The result was that Mo2C@CS displayed excellent electrocatalytic performance, as well as a great stability over 100 h in the OER, with an overpotential of 320 mV at 10 mA·cm−1 in 1.0 M KOH.
According to theoretical calculations, nitrogen-doped carbon nanotubes (N-CNTs) show metallic behavior [98], meaning more active catalytic activity. Thus, a number of researchers were inspired to investigate methods on how to synthesize N-CNTs. Mo2C combined with N-CNTs were designed as effective electrocatalysts for the OER. Das D. et al. [52] developed a Ni/MoxC NPs-supported N-doped graphene/CNT hybrid catalyst, showing great bifunctionally catalytic activity toward both OER and HER. The overpotential was only 328 mV, while achieving a current density of 10 mA·cm−1 in 1.0 M KOH. Apart from N-CNTs, nitrogen-doped carbon materials (N-CNMs), including carbon nanosheets, were applied for the OER. At the same time, materials based on molybdenum carbide combined with N-CNMs are usually doped with other transition metals, such as cobalt [99,100] and cobalt phosphide [101]. For instance, Zhu X. et al. [102] reported cobalt-doped β-molybdenum carbide encapsulated within nitrogen-doped carbon (Co0.1-βMo2C@NC), which showed superior catalytic activity toward OER. In the synthesizing process, bimetallic zeolitic imidazolate framework (ZIFs) of different Co/Zn mole ratios were saturated with a dimethylformamide (DMF) solution of MoO2(acac)2, and the mixture was calcined under nitrogen atmosphere at 900 °C (Figure 10). In the test for its electrocatalytic performance, the result of low overpotential of 262.2 mV at 10 mA·cm−1 (in 1.0 M KOH) was outstanding. However, the large resistance which had a bad effect on water splitting limited its application.
As mentioned earlier, cobalt-doped molybdenum carbide combined with N-CNMs has shown great catalytic activity in OER. Furthermore, it has been reported that materials based on Mo2C-doped with two or more elements possess excellent catalytic activity. Anjum M.A.R. et al. [53] successfully synthesized boron and nitrogen co-doped molybdenum carbide NPs embedded in B,N-doped carbon networks (B,N:Mo2C@BCN). With B and N doped, as well as formation of NPs embedded in BCN networks, the surface area was increased and charge transfer characteristics were improved, resulting in both great catalytic activity and a robust stability. For comparison, they also synthesized IrO2, Mo2C@C, and N:Mo2C@C in the experiment. As displayed in Figure 11, the material obtained displayed excellent electrocatalytic performance and remarkable stability. It had a rather low overpotential, both before and after chronoamperometry (Figure 11a,b,e,f,h), and it still exhibited a relatively low Tafel slope of 61 mV dec−1 (Figure 11c) and satisfactory stability after the durability test (Figure 11d,g).

3.6. MXene

MXene materials and their nanocomposites are considered emerging electrocatalysts for OER due to their high surface area, tunable electric structure, and thermal stability. They are a burgeoning class of 2D transition metal carbide or carbonitride nanosheets which are 3, 5, or 7 atomic layers thick and have some special properties compared to traditional 3D counterparts. Graphene is the most well-known 2D material with only a single atomic carbon layer, which leads to its limited use. Recently, MXene materials have gained great attention for the large variety of compositions, conductive property, and special morphology. A conventional synthesis method of MXenes is via selective etching of the ‘A’ elements in MAX phase precursors (general formula Mn+1AXn, in which ‘M’ represents early transition metals; ‘A’ is a group A element mainly in IIIA and IVA; ‘X’ is carbon or nitrogen; and ‘n’ usually equals 1, 2, or 3, Figure 12) by hydrofluoric acid or fluoride (HF, NH4HF2, and so on) [40]. The other synthesis approaches include a high temperature method and CVD [103]. Meanwhile, method of etching from a non-MAX phase to form MXene has also been reported [104].
MXene materials are mainly used in supercapacitors and lithium-ion batteries. However, in recent years, researchers have gradually realized the merits of MXene as catalyst materials for OER, and they have achieved good results. For a long time, there had been a view that transition metal oxides displayed better catalytic activity than most TMCs. However, some MXene materials, which are stable in aqueous media and immune to oxidation, show good practical application effect compared with transition metal oxides, which proves the potential advantages of TMCs [73]. The mainstream view is that the best strategy for maximizing the utilization of MXene is to use it as the substrate, which makes it easy to hybridize with other nano materials, so as to prepare electrocatalysts with higher catalytic activity.
One defect that limits the practical use of MXene materials is their tendency to intersheet aggregation driven by van der Waals force [105]. To solve this problem, freeze-drying, NPs forming, and polymer and hollow shell structures may be beneficial, but are restricted in large-scale production [106]. Xiu et al. [37] reported a hierarchical 3D architecture MXene material CoP@3D Ti3C2-MXene (Figure 13b) with high resistance towards aggregation. The capillary-forced strategy was utilized, which proved to be successful in forestalling aggregation between intersheets, and highly enhanced the electrochemical activity, thus accelerating the kinetic process. The synthesis method of CoP@3D Ti3C2-MXene is as follows: NPs of Co3O4 are dispersed on 3D architecture Ti3C2 MXene (produced by selectively etching Al from the Ti3AlC2 MAX phase with LiF/HCl), and the obtained product is then modulated via phosphorization at a high temperature to fabricate the CoP@3D Ti3C2-MXene material (Figure 13a).
Via electrochemical tests, the CoP@3D Ti3C2-MXene material displayed an overpotential (ηj=10) value of 220 to 340 mV vs. RHE, with a Tafel slope of 51 to 60 mV decade−1 (changed with the proportion of added CoP), which outperformed commercial RuO2 in identical conditions (Figure 13c,d). Additionally, this electrocatalyst exhibited a constant operating potential for 10 h (Figure 13e), which demonstrated its superior durability. Surprisingly, CoP@3D Ti3C2-MXene/CP, which was fabricated by loading carbon paper (CP) on a CoP@3D Ti3C2-MXene catalyst, was a bifunctional electrocatalyst efficient for the overall water splitting reaction, proving that modulating the architecture of MXene, and/or coating other material on it to improve its electrocatalytic ability is possible. Similarly, another bifunctional MXene material reported by Le et al. [39] was synthesized by combining the known catalytic active phase IrCo with plane-porous Ti3C2 MXene (named IrCo@ac-Ti3C2). The as-fabricated electrocatalyst had an overpotential value of ηj=10 = 220 mV, as shown in Figure 14a, with a long-term stability in 1.0 M KOH, thus it outweighed the activity of the other prepared catalysts in HER (displayed in Figure 14b).
The above two electrocatalysts demonstrate that some MXene compounds display superior performance for OER. Besides, the importance of material modification also lies in porosity engineering, which efficiently shortens the charge transmission passways and speeds up the kinetic process. More significantly, Ti3C2 MXene can be molded into various structures to enhance its conductivity, which is expected meaningful for the future development of OER. For example, Michael et al. [107] reported a clay-shaped Ti3C2 MXene which could be dried into solid or rolled into films with extremely high volumetric capacitance of up to 900 farads per cube centimeter, and the property of this architecture is possibly suitable for OER. We have to indicate that plenty of TMCs with a high conductivity and porous quality have been proven to be efficient electrocatalysts for OER. Therefore, synthesis procedures with apt controlling might be beneficial for yielding ideal electrocatalysts, which gives eloquent proof of the importance of research on synthesis.
Graphene material has also been one of the most popular 2D catalysts in recent study. g-C3N4, as we mentioned in Section 3.4, originating from interaction between metal centers and an N-rich network [108], has similar morphology to MXene. In the study of Tian et al. [109], g-C3N4 was coupled with Ti3C2 MXene nanosheets to form a rich-surface area electrocatalyst (called TCCN) through Ti-NX interaction. The TCCN film, which achieved an onset potential of 1.44 V and an Ej=10 magnitude of 1.65 V with a Tafel slope of 74.6 mV decade−1 in 0.1 M KOH, exhibited a better OER catalytic activity than IrO2/C. Notably, the freestanding TCCN film afforded a high double layer capacitance (Cdl) of 29.7 mF cm, which outperformed the powdery TCCN (ball-milled film TCCN, Cdl = 10.3 mF cm−2), and revealed that the freestanding film structure was more effective in magnifying the active surface area, thus also demonstrating the superiority of MXene materials for OER from another perspective. As they noted, coupling different 2D materials to form hybrid catalysts through interactions between them is a worthwhile way of enhancing their electrocatalytic activity.
Coincidentally, another extensively researched material, named 2D MOF, with porous structure and fast electron transfer property, was also utilized to hybridize Ti3C2 MXene nanosheets, in order to achieve further improved OER performance [110,111]. Zhao et al. [41] focused on in situ hybridizing 2D cobalt 1,4-benzenedicarboxylate (CoBDC) with Ti3C2 MXene nanosheets through an interdiffusion reaction-assisted process (Figure 15a), the aim was to combine the porous structure (CoBDC MOF) and conductivity and hydrophilia character (Ti3C2 MXene nanosheets) to form a composite electrocatalyst, in order to facilitate the four-electron transfer process.
The Ti3C2Tx–CoBDC hybrid material displayed a potential of 1.64 V vs. RHE and a Tafel slope of 48.2 mV decade−1 at a current density of 10 mA/cm2 in 0.1 M KOH (Figure 15b,c). As can be seen in Figure 15d,e, it also exhibited a higher double layer capacitance (9.6 mF/cm2) and smaller semicircle region than its pure components, which was in good agreement with its high surface area of 199.1 m2 g−1 (Figure 15g) caused by the porous morphology. Such a large surface area and high conductivity provided more active sites for the electron transfer process, therefore leading to an extraordinary OER efficiency. Moreover, the authors also pointed out that Ti3C2 MXene nanosheets might prevent the aggregation of CoBDC MOF, and thus put forward a new idea for tackling the problem of assembling intersheets of 2D materials. The work of above two groups suggest that combining diverse 2D catalysts may help to gather their desirable functional properties together to form an outstanding electrocatalyst for OER.
Except for Ti-based MXenes, we intend to introduce other 2D materials which have similar properties to MXenes as comparison to preferably analyse the intrinsic characters of these sorts of catalysts. Liang et al. [42] reported a single crystalline CoNiPS3 incorporated with N-doped carbon nanosheets electrocatalyst (CoNiPS3/C). This electrocatalyst formed a mosaic structure and exhibited an overpotential of 262 mV at 30 mA·cm−2 (Figure 16a,c), and a Tafel slope of 56 mV decade−1 (Figure 16b) in a traditional three-electrode system in 1.0 M KOH. Furthermore, after being separated into CoNiPS3/C nanodots, the electrocatalyst showed a much better catalytic ability; however, the stability was severely decreased (Figure 16d).
Since the first Ti3C2TX MXene has been successfully fabricated, it has developed rapidly based on its electrical and mechanical properties. However, the applications of exfoliated MXene material without appropriate surface morphology modification (denoted “bare MXene”) have rather limited further development [112]. Heterostructure and hybridization approaches for future prospects of MXene are of great importance. We also have to indicate that the hybridization and heterostructure modification methods of MXenes for OER are restricted to a specific form (MXene substrate) and Ti3C2 MXene phase are widely focalized, while the other kinds of MXenes have been less studied. For example, Pt/Nb2CTx MXene material and its compounds have been reported beneficial for water-gas shift light alkane dehydrogenation [113]. Besides, Mo2CTX, which delivered an overpotential of 283 mV at 10 mA·cm−2 in 0.5 M H2SO4 for HER, also deserves more research [114].
Another important MXene material, 2D Mo2C, which has recently attracted the attention of many researchers, displays promising prospects for HER. Geng et al. [115] successfully synthesized a 2D Mo2C-on-graphene film by means of molten copper-catalyzed CVD. In this material, nanometer-thin Mo2C particles grow on the base of graphene and form a vertical heterostructure, which is an extremely excellent electrocatalyst for HER. The as-fabricated Mo2C MXene displayed an overpotential of 236 mV vs. RHE at 10 mA/cm2 and a Tafel slope of 73 mV dec−1, and after 1000 CV cycles, no obvious decrease could be seen. However, despite the satisfactory performance of Mo2C MXene for HER, the CVD synthesis method requires a high temperature to melt the metals, especially the most common catalyst of Cu, which severely wastes energy. Therefore, Chaitoglou et al. proposed a new low-temperature CVD method, by using a liquid Sn-Cu alloy as the catalyst. In this way, the temperature demand could be reduced to 880 °C, compared with the usual 1084 °C, and the cost of energy decreased without sacrificing the quality of the Mo2C/graphene film. The electrocatalyst also exhibited a better result than the pure Mo2C [116]. As the synthetic method has been developed, Mo2C, which is proved to be an excellent catalyst for HER, might also become low-cost and efficient OER electrocatalyst.
The applications of MXenes are few in number on account of the lack of theoretical predictions and experiments [38], the utilization of unexplored MXene materials for OER is worth anticipating.

4. Conclusions and Future Perspectives

Searching for a newly-developing renewable energy system to replace traditional fossil fuels has recently been an essential task for fulfilling the progress of sustainable requirements. Simultaneously, converting new resources into storable energies is also one of the main objects of current research. Therefore, electrocatalysis for OER is of great significance as it provides us with the possibility to address both goals. To enhance the electrocatalysis efficiency, promising electrocatalysts undoubtedly play a critical role in OER system. In this review article, we have outlined fundamental understandings and experimental examples in the field of OER, the concepts of parameters, and measurement criteria for OER. Intrinsic properties, hybridization, and synthesis methods of different types of TMCs are emphasized. Additionally, the morphology, performance, and comparison with precious metal catalysts are proposed to evaluate the effects of recently fabricated TMCs, as well as their hybrids.
There has been great interest in TMC materials for OER because of their amazing catalytic abilities, including high conductivity, malleable property, abundant active sites, and stability. Moreover, their low cost and easy accessibility make them potential candidates for OER and other energy systems in the future. The major TMCs researched in OER field consist of MXene, Ni3C, Fe3C, Mo2C, and WC, with various kinds of morphologies and useful characters, but similarly exhibit excellent activities as catalytic electrodes. TMCs are promising substrates and frameworks, through hybridization, coating, and combination with perovskite spinel and graphene families, their catalytic ability can be highly enhanced due to heterostructure and high electrochemical active sites for the charge transfer of kinetics. Some also exhibit amazing conductivity and can act as bifunctional catalysts, being beneficial for overall water splitting. Transition metal carbide compounds doped with foreign elements exhibit more efficient performance than the pure TMCs, indicating the vital importance of combinations for electrocatalysts. In a manner of speaking, these works have laid the foundation for the development of OER.
Although progress has been achieved, the drawbacks of TMC electrocatalysts are also evident. We have to point out that, compared with precious metals and traditional metal oxides, as catalysts, problems for TMCs are more difficult to address. Here, we put forward some issues that need to be tackled and potential research trends, respectively, as follows.
One of the most basic problems for TMC electrocatalysts for OER is that the mechanism is still debated and needs to be researched. Considering this reaction is under harsh oxidative condition, including the oxidation to the outer surface or inner core and the oxidation of both, the mechanism of which has not been investigated by DFT to date because the complexity caused by the oxides of catalytic materials as confirmed by Yang et al. [36]. Therefore, focusing on mechanism and kinetics in different conditions is of paramount necessity. It mostly depends on the development of advanced technology since they can indicate the accurate catalyst structure and thus helping to analyse the concrete mechanism. Regarding this situation, in situ characterization techniques may help. For example, the in situ X-ray absorption spectroscopy (XAS) once used by Yang et al. [117] has recently become quite popular in OER mechanism research. Besides, Massel et al. [118] took advantage of in situ XAS and XRD to evaluate the actives sites and mechanism of co-phosphate catalysts in solid fuels. Note that under existing experimental conditions, in situ characterization cannot be easily carried out, thus the research of kinetic process and reaction pathways deserve further study.
Secondly, the lack of large-scale fabrication techniques limits the pragmatic applications of TMCs as electrocatalysts for OER. As we have mentioned, the conventional synthesis methods—both the template-free and template-assisted one—require a high-temperature environment, which consumes excessive energy, so they are both unaffordable for industrial production. Herein, effective and energy-saving approaches for TMC synthesis are required, the development of which will certainly promote the applications of TMCs in many fields, particularly in OER. The long and sophisticated process of yielding pure and high-quality TMCs also sets a barrier for the improvement of TMCs. The properties of catalysts, which affect the performance for OER profoundly, depend on the crystalline structure. However, the orientation growth in the surface or inter-surface area of TMC materials is unpredictable and difficult to control. As a result, it takes an extremely long period of time to synthesize satisfying electrocatalysts, even if we already know the specific ideal structure. The situation asks for more accurate manipulation of synthesis process, which requires long-term research. Nevertheless, some hybridization and mosaic methods may help. For example, doping extraneous elements and/or changing the morphology in order to modify pure TMCs is one of the main ways of enhancing the material’s catalytic ability. In addition, combining perovskite or graphene with the original electrocatalyst may have a large effect on the catalytic properties, which may result in high conductivity and more active sites in the catalyst and make it a promising catalyst. Other than this, another approach is to optimize the synthesis approach, or invent new ones, modulating dimensions and crystalline forms, such as 2D structure or nanostructure, in order to adapt to OER.
In short, whilst TMCs have recently been focused on, there is still only a small amount of research on them. We have to point out that TMC electrocatalysts deserve to be deeply explored due to their excellent performance for OER in the overall water splitting reaction and their low-cost property, which are very promising for the future compared with other materials. The development of highly active structures—such as MOF, graphene, and nano technology—will allow TMCs to exhibit varied possibilities. We believe that with the support of principles, theoretical mechanism, and experimental standpoints, applications for practical and large-scale use of TMCs as electrocatalysts for OER can be highly anticipated. Moreover, the advantages of TMCs will facilitate the development of other electrocatalytic systems as they share similar conditions and issues.

Author Contributions

Y.W. wrote the introduction, as well as Section 2 and Section 3.6, in addition to carrying out revision work. Q.W. was responsible for the writing of Section 3.1, Section 3.2, and Section 3.3. L.T. finished Section 3.4 and Section 3.5 and checked all the formats of references. B.Z. wrote the abstract and the conclusion and also revised the manuscript. Corresponding authors K.L. and X.Z. designed the whole work and revised the manuscript. Y.W., Q.W., B.Z. and L.T. share same contribution to this article. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Nankai University Undergraduate Innovation Projects 100 (no. 201910055811) and the State Key Laboratory of Simulation and Regulation of Water Cycle in River Basin (no. IWHR-SKLKF201806).

Acknowledgments

This work was supported by Nankai University Undergraduate Innovation Projects 100 (no. 201910055811) and the State Key Laboratory of Simulation and Regulation of Water Cycle in River Basin (no. IWHR-SKLKF201806).

Conflicts of Interest

There are no conflicts to declare.

References

  1. Yan, X.D.; Tian, L.H.; He, M.; Chen, X.B. Three-dimensional crystalline/amorphous Co/Co3O4 core/shell nanosheets as efficient electrocatalysts for the hydrogen evolution reaction. Nano Lett. 2015, 15, 6015–6021. [Google Scholar] [CrossRef] [PubMed]
  2. Yang, Y.; Fei, H.L.; Ruan, G.D.; Tour, J.M. Porous cobalt-based thin film as a bifunctional catalyst for hydrogen generation and oxygen generation. Adv. Mater. 2015, 27, 3175–3180. [Google Scholar] [CrossRef] [PubMed]
  3. Jeong, D.; Jin, K.; Jerng, S.E.; Seo, H.; Kim, D.; Nahm, S.H.; Kim, S.H.; Nam, K.T. Mn5O8 nanoparticles as efficient water oxidation catalysts at neutral ipH. ACS Catal. 2015, 5, 4624–4628. [Google Scholar] [CrossRef]
  4. Tang, C.; Wang, H.S.; Wang, H.F.; Zhang, Q.; Tian, G.L.; Nie, J.Q.; Wei, F. Spatially confined hybridization of nanometer-sized NiFe hydroxides into nitrogen-doped graphene frameworks leading to superior oxygen evolution reactivity. Adv. Mater. 2015, 27, 4516–4522. [Google Scholar] [CrossRef] [PubMed]
  5. Ramesh, R.; Sawant, S.Y.; Nandi, D.K. Hydrogen Evolution Reaction by Atomic Layer-Deposited MoNx on Porous Carbon Substrates: The Effects of Porosity and Annealing on Catalyst Activity and Stability. ChemSusChem 2020, 13, 4159–4168. [Google Scholar] [CrossRef]
  6. Wu, P.W.; Wu, J.; Si, H.N. 3D Holey-Graphene Architecture Expedites Ion Transport Kinetics to Push the OER Performance. Adv. Energy Mater. 2020, 10, 2001005. [Google Scholar] [CrossRef]
  7. Chen, B.T.; Morlanes, N.; Adogla, E.; Takanabe, K.; Rodionov, V.O. An efficient and stable hydrophobic molecular cobalt catalyst for water electro-oxidation at neutral pH. ACS Catal. 2016, 6, 4647–4652. [Google Scholar] [CrossRef]
  8. Wang, J.; Zhong, H.X.; Wang, Z.L.; Meng, F.L.; Zhang, X.B. Integrated three-dimensional carbon paper/carbon tubes/cobalt-sulfide sheets as an efficient electrode for overall water splitting. ACS Nano 2016, 10, 2342–2348. [Google Scholar] [CrossRef]
  9. Wang, H.; Hsu, Y.Y.; Chen, R.; Chan, T.S.; Chen, H.M.; Liu, B. Ni3+-induced formation of active NiOOH on the spinel Ni–Co oxide surface for efficient oxygen evolution reaction. Adv. Energy Mater. 2015, 5, 5. [Google Scholar] [CrossRef]
  10. Tian, J.Q.; Liu, Q.; Asiri, A.M.; Sun, X.P. Self-supported nanoporous cobalt phosphide nanowire arrays: An efficient 3D hydrogen-evolving cathode over the wide range of pH 0–14. J. Am. Chem. Soc. 2014, 136, 7587–7590. [Google Scholar] [CrossRef]
  11. Wu, Y.Z.; Chen, M.X.; Han, Y.Z.; Luo, H.X.; Su, X.J.; Zhang, M.T.; Lin, X.H.; Sun, J.L.; Wang, L.; Deng, L.; et al. Fast and simple preparation of iron-based thin films as highly efficient water-oxidation catalysts in neutral aqueous solution. Angew. Chem. 2015, 127, 4952–4957. [Google Scholar] [CrossRef]
  12. Chen, P.Z.; Xu, K.; Fang, Z.W.; Tong, Y.; Wu, J.C.; Lu, X.L.; Peng, X.; Ding, H.; Wu, C.Z.; Xie, Y. Metallic Co4N porous nanowire arrays activated by surface oxidation as electrocatalysts for the oxygen evolution reaction. Angew. Chem. Int. Ed. 2015, 54, 14710–14714. [Google Scholar] [CrossRef] [PubMed]
  13. Tian, J.Q.; Liu, Q.; Cheng, N.Y.; Asiri, A.M.; Sun, X.P. Self-supported Cu3P nanowire arrays as an integrated high-performance three-dimensional cathode for generating hydrogen from Water. Angew. Chem. Int. Ed. 2014, 53, 9577–9581. [Google Scholar] [CrossRef] [PubMed]
  14. Bockris, J.O.M. Hydrogen economy in the future. Int. J. Hydrogen Energy 1999, 24, 1–15. [Google Scholar] [CrossRef]
  15. Chen, Q.; Fu, Y.L.; Jin, J.L.; Zang, W.J.; Liu, X.; Zhang, X.Y.; Huang, W.Z.; Kou, Z.K.; Wang, J.; Zhou, L.; et al. In-situ surface self-reconstruction in ternary transition metal dichalcogenide nanorod arrays enables efficient electrocatalytic oxygen evolution. J. Energy Chem. 2021, 55, 10–16. [Google Scholar] [CrossRef]
  16. Tang, Y.J.; Liu, C.H.; Huang, W. Bimetallic Carbides-Based Nanocomposite as Superior Electrocatalyst for Oxygen Evolution Reaction. ACS Appl. Mater. Interfaces 2017, 9, 16977–16985. [Google Scholar] [CrossRef] [PubMed]
  17. Hu, Y.; Jensen, J.O.; Zhang, W. Hollow Spheres of Iron Carbide Nanoparticles Encased in Graphitic Layers as Oxygen Reduction Catalysts. Angew. Chem. Int. Ed. 2014, 53, 3675–3679. [Google Scholar] [CrossRef]
  18. Benchakar, M.; Bilyk, T.; Garnero, C. MXene Supported Cobalt Layered Double Hydroxide Nanocrystals: Facile Synthesis Route for a Synergistic Oxygen Evolution Reaction Electrocatalyst. Adv. Mater. Interfaces 2019, 6, 1901328. [Google Scholar] [CrossRef]
  19. Munir, A.; Haq, T.U.; Saleem, M.; Qurashi, A.; Hussain, S.Z.; Sher, F.; Hamid, A.U.; Jilani, A.; Hussain, I. Controlled engineering of nickel carbide induced N-enriched carbon nanotubes for hydrogen and oxygen evolution reactions in wide pH range. Electrochim. Acta 2020, 341, 136032. [Google Scholar] [CrossRef]
  20. Li, G.L.; Zhang, X.B.; Zhang, H. Ultrathin 2D nanosheet based 3D hierarchical hollow polyhedral CoM/C (M = Ni, Cu, Mn) phosphide nanocages as superior electrocatalysts toward oxygen evolution reaction. Chem. Eng. J. 2020, 398, 125467. [Google Scholar] [CrossRef]
  21. Ren, S.S.; Duan, X.D.; Ge, F.Y. Trimetal-based N-doped carbon nanotubes arrays on Ni foams as self-supported electrodes for hydrogen/oxygen evolution reactions and water splitting. J. Power Sources 2020, 480, 228866. [Google Scholar] [CrossRef]
  22. Wang, H.P.; Zhu, S.; Deng, J.W.; Zhang, W.C.; Feng, Y.Z.; Ma, J.M. Transition metal carbides in electrocatalytic oxygen evolution reaction. Chin. Chem. Lett. 2020. [Google Scholar] [CrossRef]
  23. McCrory, C.C.L.; Jung, S.; Ferrer, I.M.; Chatman, S.M.; Peters, J.C.; Jaramillo, T.F. Benchmarking Hydrogen Evolving Reaction and Oxygen Evolving Reaction Electrocatalysts for Solar Water Splitting Devices. J. Am. Chem. Soc. 2015, 137, 4347–4357. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Suen, N.T.; Hung, S.F.; Quan, Q. Electrocatalysis for the oxygen evolution reaction: Recent development and future perspectives. Chem. Soc. Rev. 2017, 46, 337–365. [Google Scholar] [CrossRef]
  25. Yu, P.; Wang, F.M.; Shifa, T.A.; Zhan, X.Y.; Lou, X.D.; Xia, F.; He, J. Earth abundant materials beyond transition metal dichalcogenides: A focus on electrocatalyzing hydrogen evolution reaction. Nano Energy 2019, 58, 244–276. [Google Scholar] [CrossRef]
  26. You, B.; Jiang, N.; Sheng, M.L.; Gul, S.; Yano, J.; Sun, Y.J. High-Performance Overall Water Splitting Electrocatalysts Derived from Cobalt-Based Metal–Organic Frameworks. Chem. Mater. 2015, 27, 7636–7642. [Google Scholar] [CrossRef]
  27. Fei, H.L.; Dong, J.; Arellano-Jiménez, M.J. Atomic cobalt on nitrogen-doped graphene for hydrogen generation. Nat. Commun. 2015, 6, 8668. [Google Scholar] [CrossRef]
  28. Zhang, Y.; Lu, T.; Ye, Y.K.; Dai, W.J.; Zhu, Y.A.; Pan, Y. Stabilizing Oxygen Vacancy in Entropy-Engineered CoFe2O4-Type Catalysts for Co-prosperity of Efficiency and Stability in an Oxygen Evolution Reaction. ACS Appl. Mater. Interfaces 2020, 12, 32548–32555. [Google Scholar] [CrossRef]
  29. Frydendal, R.; Paoli, E.A.; Knudsen, B.P.; Wickman, B.; Malacrida, P.; Stephens, I.E.L.; Chorkendorff, I. Benchmarking the Stability of Oxygen Evolution Reaction Catalysts: The Importance of Monitoring Mass Losses. ChemElectroChem 2014, 1, 2075–2081. [Google Scholar] [CrossRef] [Green Version]
  30. Yu, Y.; Zhou, J.; Sun, Z. Novel 2D Transition-Metal Carbides: Ultrahigh Performance Electrocatalysts for Overall Water Splitting and Oxygen Reduction. Adv. Funct. Mater. 2020, 2000570. [Google Scholar] [CrossRef]
  31. Anurupa, M. Cobalt-based heterogeneous catalysts in an electrolyzer system for sustainable energy storage. Dalton Trans. 2020, 49, 11430–11450. [Google Scholar]
  32. Cheng, W.; Lu, X.F.; Luan, D.; Lou, X.W. NiMn-Based Bimetal-Organic Framework Nanosheets Supported on Multi-Channel Carbon Fibers for Efficient Oxygen Electrocatalysis. Angew. Chem. Int. Ed. 2020, 59, 18234–18239. [Google Scholar] [CrossRef] [PubMed]
  33. Ouyang, T.; Ye, Y.Q.; Wu, C.Y.; Xiao, K.; Liu, Z.Q. Heterostructures Composed of N-Doped Carbon Nanotubes Encapsulating Cobalt and b-Mo2C Nanoparticles as Bifunctional Electrodes for Water Splitting. Angew. Chem. Int. Ed. 2019, 58, 4923–4928. [Google Scholar] [CrossRef] [PubMed]
  34. Wang, P.; Qin, R.; Ji, P.; Pu, Z.; Zhu, J.; Lin, C.; Zhao, Y.; Tang, H.; Li, W.; Mu, S. Synergistic Coupling of Ni Nanoparticles with Ni3C Nanosheets for Highly Efficient Overall Water Splitting. Small 2020, 16, 2001642. [Google Scholar] [CrossRef] [PubMed]
  35. Liu, Y.; Zhang, J.; Li, Y. Manipulating dehydrogenation kinetics through dual-doping Co3N electrode enables highly efficient hydrazine oxidation assisting self-powered H2 production. Nat. Commun. 2020, 11, 5148–5180. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Yang, C.C.; Zai, S.F.; Zhou, Y.T.; Du, L.; Jiang, Q. Fe3C-Co Nanoparticles Encapsulated in a Hierarchical Structure of N-Doped Carbon as a Multifunctional Electrocatalyst for ORR, OER, and HER. Adv. Funct. Mater. 2019, 29, 1901949. [Google Scholar] [CrossRef]
  37. Xiu, L.Y.; Wang, Z.Y.; Yu, M.Z.; Wu, X.H.; Qiu, J.S. Aggregation-Resistant 3D MXene-Based Architecture as Efficient Bifunctional Electrocatalyst for Overall Water Splitting. ACS Nano 2018, 12, 8017–8028. [Google Scholar] [CrossRef]
  38. Du, C.F.; Dinh, K.N.; Liang, Q.; Zheng, Y.; Luo, Y.; Zhang, J.; Yan, Q. Self-Assemble and In Situ Formation of Ni1−xFexPS3 Nanomosaic-Decorated MXene Hybrids for Overall Water Splitting. Adv. Energy Mater. 2018, 8, 1801127. [Google Scholar] [CrossRef]
  39. Anh, L.T.; Quang, T.N.; Yeseul, H.; Meeree, K.; Hyoyoung, L. Porosity-Engineering of MXene as a Support Material for a Highly Efficient Electrocatalyst toward Overall Water Splitting. ChemSusChem 2020, 13, 945–955. [Google Scholar]
  40. Li, Z.; Wu, Y. 2D Early Transition Metal Carbides (MXenes) for Catalysis. Small 2019, 15, 15. [Google Scholar] [CrossRef] [Green Version]
  41. Zhao, L.; Dong, B.; Li, S. Interdiffusion Reaction-Assisted Hybridization of Two-Dimensional Metal–Organic Frameworks and Ti3C2Tx Nanosheets for Electrocatalytic Oxygen Evolution. ACS Nano 2017, 11, 5800–5807. [Google Scholar] [CrossRef] [PubMed]
  42. Liang, Q.H.; Zhong, L.X.; Du, C.F.; Zheng, Y.; Luo, Y.B.; Xu, J.W.; Li, S.Z.; Yan, Q.Y. Mosaic-Structured Cobalt Nickel Thiophosphate Nanosheets Incorporated N-doped Carbon for Efficient and Stable Electrocatalytic Water Splitting. Adv. Funct. Mater. 2018, 28, 1805075. [Google Scholar] [CrossRef]
  43. Tian, Y.H.; Xu, L.; Qian, J.C.; Bao, J.Y. Fe3C/Fe2O3 heterostructure embedded in N-doped graphene as a bifunctional catalyst for quasi-solid-state zinceair batteries. Carbon 2019, 146, 763–771. [Google Scholar] [CrossRef]
  44. Zhao, P.P.; Hua, X.; Xu, W.; Luo, W.; Chen, S.L.; Cheng, G.Z. Metal–organic framework-derived hybrid of Fe3C nanorod-encapsulated, N-doped CNTs on porous carbon sheets for highly efficient oxygen reduction and water oxidation. Catal. Sci. Technol. 2016, 6, 6365–6371. [Google Scholar] [CrossRef]
  45. Fan, H.S.; Yu, H.; Zhang, Y.F.; Zheng, Y.; Dai, Z.F.; Li, B.; Zong, Y. Fe-Doped Ni3C Nanodots in N-Doped Carbon Nanosheets for Efficient Hydrogen-Evolution and Oxygen-Evolution Electrocatalysis. Angew. Chem. Int. Ed. 2017, 56, 12566–12570. [Google Scholar] [CrossRef]
  46. Geng, S.; Yang, W.W.; Yu, Y.S. Fabrication of NiC/MoC/NiMoO4 Heterostructured Nanorod Arrays as Stable Bifunctional Electrocatalysts for Efficient Overall Water Splitting. Chem. Asian J. 2019, 14, 1013–1020. [Google Scholar] [CrossRef]
  47. He, P.; Yu, X.Y.; Lou, X.W.D. Carbon-Incorporated Nickel-Cobalt Mixed Metal Phosphide Nanoboxes with Enhanced Electrocatalytic Activity for Oxygen Evolution. Angew. Chem. Int. Ed. 2017, 56, 3897–3900. [Google Scholar] [CrossRef]
  48. Yang, H.; Luo, S.; Li, X. Controllable orientation-dependent crystal growth of high-index faceted dendritic NiC0.2 nanosheets as high-performance bifunctional electrocatalysts for overall water splitting. J. Mater. Chem. A 2016, 4, 18499–18508. [Google Scholar] [CrossRef]
  49. Joy, J.; Sekar, A.; Vijayaraghavan, S.; Kumar, P.; Pillai, V.K.; Alwarappan, S. Nickel-Incorporated, Nitrogen-Doped Graphene Nanoribbons as Efficient Electrocatalysts for Oxygen Evolution Reaction. J. Electrochem. Soc. 2018, 165, H141–H146. [Google Scholar] [CrossRef]
  50. Regmi, Y.N.; Wan, C.; Duffee, K.D.; Leonard, B.M. Nanocrystalline Mo2C as a Bifunctional Water Splitting Electrocatalyst. ChemCatChem 2015, 7, 3911–3915. [Google Scholar] [CrossRef] [Green Version]
  51. Wang, H.; Cao, Y.J.; Sun, C.; Zou, G.F.; Huang, J.W.; Kuai, X.X.; Zhao, J.Q.; Gao, L.J. Strongly Coupled Molybdenum Carbide on Carbon Sheets as a Bifunctional Electrocatalyst for Overall Water Splitting. ChemSusChem 2017, 10, 3540–3546. [Google Scholar] [CrossRef] [PubMed]
  52. Das, D.; Santra, S.; Nanda, K.K. In Situ Fabrication of a Nickel/Molybdenum Carbide-Anchored N-Doped Graphene/CNT Hybrid: An Efficient (Pre)catalyst for OER and HER. ACS Appl. Mater. Interfaces 2018, 10, 35025–35038. [Google Scholar] [CrossRef] [PubMed]
  53. Raza, A.M.A.; Hee, L.M.; Sung, L.J. Boron- and Nitrogen-Codoped Molybdenum Carbide Nanoparticles Imbedded in a BCN Network as a Bifunctional Electrocatalyst for Hydrogen and Oxygen Evolution Reactions. ACS Catal. 2018, 8, 8296–8305. [Google Scholar]
  54. Zhao, Y.X.; Bowers, C.W.; Spain, I.L. Graphitic nature of chemical vapor-deposited carbon filaments grown on silicon surfaces from acetylene. Carbon 1988, 26, 291–293. [Google Scholar] [CrossRef]
  55. Fadil, N.A.; Saravanna, G.; Ramesh, G.V. Synthesis and electrocatalytic performance of atomically ordered nickel carbide (Ni3C) nanoparticles. Chem. Commun. 2014, 50, 6451–6453. [Google Scholar] [CrossRef]
  56. Wan, C.; Leonard, B.M. Iron-Doped Molybdenum Carbide Catalyst with High Activity and Stability for the Hydrogen Evolution Reaction. Chem. Mater. 2015, 27, 4281–4288. [Google Scholar] [CrossRef]
  57. Hao, J.; Zhang, G.F.; Zheng, Y.T.; Luo, W.H.; Jin, C.; Wang, R.; Wang, Z.; Zheng, W.J. Controlled synthesis of Ni3C/nitrogen-doped carbon nanoflakes for efficient oxygen evolution. Electrochim. Acta 2019, 320. [Google Scholar] [CrossRef]
  58. Kang, J.; Duan, X.G.; Wang, C.; Sun, H.Q.; Tan, X.Y.; Tade, M.O.; Wang, S.B. Nitrogen-doped bamboo-like carbon nanotubes with Ni encapsulation for persulfate activation to remove emerging contaminants with excellent catalytic stability. Chem. Eng. J. 2018, 332, 398–408. [Google Scholar] [CrossRef]
  59. Qin, Q.; Hao, J.; Zheng, W.J. Ni/Ni3C Core/Shell Hierarchical Nanospheres with Enhanced Electrocatalytic Activity for Water Oxidation. ACS Appl. Mater. Interfaces 2018, 10, 17827–17834. [Google Scholar] [CrossRef]
  60. Zi, Y.Y.; Yu, D.; Min, R.G. A one-dimensional porous carbon-supported Ni/Mo2C dual catalyst for efficient water splitting. Chem. Sci. 2017, 8, 968–973. [Google Scholar]
  61. Rostami, S.; Pour, A.N.; Salimi, A.; Abolghasempour, A. Hydrogen adsorption in metal-organic frameworks (MOFs): Effects of adsorbent architecture. Int. J. Hydrogen Energy 2018, 43, 7072–7080. [Google Scholar] [CrossRef]
  62. Lee, J.Y.; Farha, O.K.; Roberts, J. Metal–organic framework materials as catalysts. Chem. Soc. Rev. 2019, 38, 1450–1459. [Google Scholar] [CrossRef] [PubMed]
  63. Zhao, S.S.; Liu, Y.Y. Fluorescent Aromatic Tag-Functionalized MOFs for Highly Selective Sensing of Metal Ions and Small Organic Molecules. Inorg. Chem. 2016, 55, 2261–2273. [Google Scholar] [CrossRef] [PubMed]
  64. Jia, X.X.; Wang, M.H.; Liu, G. Mixed-metal MOF-derived Co-doped Ni3C/Ni NPs embedded in carbon matrix as an efficient electrocatalyst for oxygen evolution reaction. Int. J. Hydrogen Energy 2019, 44, 24572–24579. [Google Scholar] [CrossRef]
  65. Levy, R.B.; Boudart, M. Platinum-Like Behavior of Tungsten Carbide in Surface Catalysis. Science 1973, 181, 547–549. [Google Scholar] [CrossRef]
  66. Glibin, V.; Svirko, R.; Bashtan-Kandybovich, R. Synthesis, surface chemistry and electrocatalytic properties of oxygen-modified tungsten carbide in relation to the electrochemical hydrogen oxidation. Surf. Sci. 2010, 604, 500–507. [Google Scholar] [CrossRef]
  67. Palanker, V.S.; Gajyev, R. On adsorption and electro-oxidation of some compounds on tungsten carbide; their effect on hydrogen electro-oxidation. Electrochim. Acta 1977, 22, 133–136. [Google Scholar] [CrossRef]
  68. Zhou, X.S.; Qiu, Y.J. Tungsten carbide nanofibers prepared by electrospinning with high electrocatalytic activity for oxygen reduction. Int. J. Hydrogen Energy 2011, 36, 7398–7404. [Google Scholar] [CrossRef]
  69. Han, N.; Yang, K.R.; Lu, Z. Nitrogen-doped tungsten carbide nanoarray as an efficient bifunctional electrocatalyst for water splitting in acid. Nat. Commun. 2018, 9, 924. [Google Scholar] [CrossRef] [Green Version]
  70. Bukola, S.; Merzougui, B.; Akinpelu, A.; Zeama, M. Cobalt and Nitrogen Co-Doped Tungsten Carbide Catalyst for Oxygen Reduction and Hydrogen Evolution Reactions. Electrochim. Acta 2016, 190, 1113–1123. [Google Scholar] [CrossRef]
  71. Kim, D.W.; Li, O.L.; Pootawang, P. Solution plasma synthesis process of tungsten carbide on N-doped carbon nanocomposite with enhanced catalytic ORR activity and durability. RSC Adv. 2014, 4, 16813–16819. [Google Scholar] [CrossRef]
  72. Zhao, T.; Gao, J.K.; Wu, J.; He, P.P.; Li, Y.W.; Yao, J.M. Highly Active Cobalt/Tungsten Carbide@N-Doped Porous Carbon Nanomaterials Derived from Metal-Organic Frameworks as Bifunctional Catalysts for Overall Water Splitting. Energy Technol. 2019, 7, 1800969. [Google Scholar] [CrossRef]
  73. Zhong, Y.; Xia, X.H.; Shi, F.; Zhan, J.Y.; Tu, J.P.; Fan, H.J. Transition Metal Carbides and Nitrides in Energy Storage and Conversion. Adv. Sci. 2016, 3, 1500286. [Google Scholar] [CrossRef] [PubMed]
  74. Chen, J.G. Carbide and Nitride Overlayers on Early Transition Metal Surfaces: Preparation, Characterization, and Reactivities. Chem. Rev. 1996, 96, 1477–1498. [Google Scholar] [CrossRef]
  75. Amiinu, I.S.; Pu, Z.; Liu, X. Multifunctional Mo–N/C@MoS2 Electrocatalysts for HER, OER, ORR, and Zn–Air Batteries. Adv. Funct. Mater. 2017, 27, 1702300. [Google Scholar] [CrossRef]
  76. Qian, Y.H.; Hu, Z.G.; Ge, X.M.; Yang, S.L.; Peng, Y.W.; Kang, Z.X.; Liu, Z.L.; Lee, J.Y.; Zhao, D. A metal-free ORR/OER bifunctional electrocatalyst derived from metal-organic frameworks for rechargeable Zn-Air batteries. Carbon 2017, 111, 641–650. [Google Scholar] [CrossRef]
  77. Jiang, H.; Gu, J.X.; Zheng, X.S.; Liu, M.; Qiu, X.Q. Defect-rich and ultrathin N doped carbon nanosheets as advanced trifunctional metal-free electrocatalysts for the ORR, OER and HER. Energy Environ. Sci. 2019, 12, 322–333. [Google Scholar] [CrossRef]
  78. Arico, A.S.; Bruce, P.; Scrosati, B.; Tarascon, J.M.; Schalkwijk, W.V. Nanostructured materials for advanced energy conversion and storage devices. Nat. Mater. 2005, 4, 366–377. [Google Scholar] [CrossRef]
  79. Debe, M.K. Electrocatalyst Approaches and Challenges for Automotive Fuel cells. Nature 2012, 486, 43–51. [Google Scholar] [CrossRef]
  80. Fu, J.; Cano, Z.P.; Park, M.G.; Yu, A.P.; Fowler, M.; Chen, Z.W. Electrically Rechargeable Zinc–Air Batteries: Progress, Challenges, and Perspectives. Adv. Mater. 2017, 29, 1604685. [Google Scholar] [CrossRef]
  81. Wang, Z.L.; Xu, D.; Xu, J.J.; Zhang, X.B. Oxygen electrocatalysts in metal–air batteries: From aqueous to nonaqueous electrolytes. Chem. Soc. Rev. 2014, 43, 7746–7786. [Google Scholar] [CrossRef] [PubMed]
  82. Defilippi, C.; Shinde, D.V.; Dang, Z.; Manna, L.; Hardacre, C.; Greer, A.J.; D’Agostino, C.; Giordano, C. HfN Nanoparticles: An Unexplored Catalyst for the Electrocatalytic Oxygen Evolution Reaction. Angew. Chem. 2019, 131, 15610–15616. [Google Scholar] [CrossRef]
  83. Song, J.Y.; Ren, Y.R.; Li, J.S.; Huang, X.B.; Cheng, F.Y.; Tang, Y.G.; Wang, H.Y. Coreshell Co/CoNx@C nanoparticles enfolded by Co-N doped carbon nanosheets as a highly efficient electrocatalyst for oxygen reduction reaction. Carbon 2018, 138, 300–308. [Google Scholar] [CrossRef]
  84. Wu, Y.X.; Wang, T.H.; Zhang, Y.D.; Xin, S.; He, X.J.; Zhang, D.W.; Shui, J.L. Electrocatalytic performances of g-C3N4-LaNiO3 composite as bi-functional catalysts for lithium-oxygen batteries. Sci. Rep. 2016, 6, 24314. [Google Scholar] [CrossRef] [PubMed]
  85. Xu, C.Y.; Han, Q.; Zhao, Y.; Wang, L.X.; Li, Y.; Qu, L.T. Sulfur-doped graphitic carbon nitride decorated with graphene quantum dots for an efficient metal-free electrocatalyst. J. Mater. Chem. 2015, 3, 1841–1846. [Google Scholar] [CrossRef]
  86. Wang, Q.C.; Lei, Y.P.; Chen, Z.Y. Fe/Fe3C@C nanoparticles encapsulated in N-doped graphene–CNTs framework as an efficient bifunctional oxygen electrocatalyst for robust rechargeable Zn–air batteries. J. Mater. Chem. A 2017, 6, 516–526. [Google Scholar] [CrossRef]
  87. Wu, N.; Lei, Y.P.; Wang, Q.C.; Wang, B.; Han, C.; Wang, Y.D. Facile synthesis of FeCo@NC core–shell nanospheres supported on graphene as an efficient bifunctional oxygen electrocatalyst. Nano Res. 2017, 10, 2332–2343. [Google Scholar] [CrossRef]
  88. Su, C.Y.; Cheng, H.; Li, W.; Liu, Z.Q.; Li, N.; Hou, Z.; Bai, F.Q.; Zhang, H.X.; Ma, T.Y. Atomic Modulation of FeCo-Nitrogen-Carbon Bifunctional Oxygen Electrodes for Rechargeable and Flexible All-Solid-State Zinc-Air Battery. Adv. Energy Mater. 2017, 7, 1602420. [Google Scholar] [CrossRef]
  89. Liu, Q.; Wang, Y.B.; Dai, L.M.; Yao, J.N. Scalable Fabrication of Nanoporous Carbon Fiber Films as Bifunctional Catalytic Electrodes for Flexible Zn-Air Batteries. Adv. Mater. 2016, 28, 3000–3006. [Google Scholar] [CrossRef]
  90. Cui, Z.M.; Li, Y.T.; Fu, G.T. Robust Fe3Mo3C Supported IrMn Clusters as Highly Efficient Bifunctional Air Electrode for Metal–Air Battery. Adv. Mater. 2017, 29, 1702385. [Google Scholar] [CrossRef] [Green Version]
  91. Chen, W.F.; Wang, C.H.; Sasaki, K. Highly active and durable nanostructured molybdenum carbide electrocatalysts for hydrogen production. Energy Environ. Sci. 2013, 6, 943–951. [Google Scholar] [CrossRef]
  92. Liao, L.; Wang, S.N.; Xiao, J.J. A nanoporous molybdenum carbide nanowire as an electrocatalyst for hydrogen evolution reaction. Energy Environ. Sci. 2014, 7, 387–392. [Google Scholar] [CrossRef] [Green Version]
  93. Wu, H.B.; Xia, B.Y.; Yu, L. Porous molybdenum carbide nano-octahedrons synthesized via confined carburization in metal-organic frameworks for efficient hydrogen production. Nat. Commun. 2015, 6, 6512. [Google Scholar] [CrossRef] [PubMed]
  94. Regmi, Y.N.; Waetzig, G.R.; Duffee, K.D.; Schmuecker, S.M.; Thode, J.M.; Leonard, B.M. Carbides of group IVA, VA and VIA transition metals as alternative HER and ORR catalysts and support materials. J. Mater. Chem. A 2015, 3, 10085–10091. [Google Scholar] [CrossRef]
  95. Zhu, J.; Holmen, A.; Chen, D. Carbon Nanomaterials in Catalysis: Proton Affinity, Chemical and Electronic Properties, and their Catalytic Consequences. ChemCatChem 2013, 5, 378–401. [Google Scholar] [CrossRef]
  96. Chen, C.C.; Lin, C.L.; Chen, L.C. Functionalized Carbon Nanomaterial Supported Palladium Nano-Catalysts for Electrocatalytic Glucose Oxidation Reaction. Electrochim. Acta 2015, 152, 408–416. [Google Scholar] [CrossRef]
  97. Zhao, Z.; Xia, Z. Design Principles for Dual-Element-Doped Carbon Nanomaterials as Efficient Bifunctional Catalysts for Oxygen Reduction and Evolution Reactions. ACS Catal. 2016, 6, 1553–1558. [Google Scholar] [CrossRef]
  98. Podyacheva, O.Y.; Ismagilov, Z.R. Nitrogen-doped carbon nanomaterials: To the mechanism of growth, electrical conductivity and application in catalysis. Catal. Today 2015, 249, 12–22. [Google Scholar] [CrossRef]
  99. Kim, M.; Kim, S.; Song, D.; Oh, S.; Chang, K.J.; Cho, E. Promotion of electrochemical oxygen evolution reaction by chemical coupling of cobalt to molybdenum carbide. Appl. Catal. B Environ. 2018, 227, 340–348. [Google Scholar] [CrossRef]
  100. Jiang, J.; Liu, Q.; Zeng, C. Cobalt/molybdenum carbide@N-doped carbon as a bifunctional electrocatalyst for hydrogen and oxygen evolution reactions. J. Mater. Chem. A 2017, 5, 16929–16935. [Google Scholar] [CrossRef]
  101. Dutta, S.; Indra, A.; Han, H.S. An Intriguing Pea-Like Nanostructure of Cobalt Phosphide on Molybdenum Carbide Incorporated Nitrogen-Doped Carbon Nanosheets for Efficient Electrochemical Water Splitting. ChemSusChem 2018, 11, 3956–3964. [Google Scholar] [CrossRef] [PubMed]
  102. Zhu, X.; Zhang, X.; Huang, L. Cobalt doped beta-molybdenum carbide nanoparticles encapsulated within nitrogen-doped carbon for oxygen evolution. Chem. Commun. 2019, 55, 9995–9998. [Google Scholar] [CrossRef] [PubMed]
  103. Xu, C.; Wang, L.B.; Liu, Z.B.; Chen, L.; Guo, J.K.; Kang, N.; Xiu, L.M.; Cheng, H.M.; Ren, W.C. Large-area high-quality 2D ultrathin Mo2C superconducting crystals. Nat. Mater. 2015, 14, 1135–1141. [Google Scholar] [CrossRef] [PubMed]
  104. Zhou, J.; Zha, X.H.; Chen, F.Y. A two-dimensional zirconium carbide by selective etching of Al3C3 from nanolaminated Zr3Al3C5. Angew. Chem. Int. Ed. 2016, 128, 5092–5097. [Google Scholar] [CrossRef]
  105. Zhao, M.Q.; Xie, X.; Ren, C.E.; Makaryan, T.; Anasori, B.; Wang, G.; Gogotsi, Y. Hollow MXene Spheres and 3D Macroporous MXene Frameworks for Na-Ion Storage. Adv. Mater. 2017, 29, 1702410. [Google Scholar] [CrossRef] [PubMed]
  106. Wu, X.H.; Wang, Z.Y.; Yu, M.Z.; Xiu, L.Y.; Qiu, J.S. Stabilizing the MXenes by Carbon Nanoplating for Developing Hierarchical Nanohybrids with Efficient Lithium Storage and Hydrogen Evolution Capability. Adv. Mater. 2017, 29, 1607017. [Google Scholar] [CrossRef] [PubMed]
  107. Ghidiu, M.; Lukatskaya, M.R.; Zhao, M.Q. Conductive two-dimensional titanium carbide ‘clay’ with high volumetric capacitance. Nature 2014, 516, 78–81. [Google Scholar] [CrossRef] [PubMed]
  108. Ma, T.Y.; Ran, J.; Dai, S.; Jaroniec, M.; Qiao, S.Z. Phosphorus-Doped Graphitic Carbon Nitrides Grown In Situ on Carbon-Fiber Paper: Flexible and Reversible Oxygen Electrodes. Angew. Chem. Int. Ed. 2015, 54, 4646–4650. [Google Scholar] [CrossRef] [Green Version]
  109. Ma, T.Y.; Cao, J.L.; Jaroniec, M.; Qiao, S.Z. Interacting Carbon Nitride and Titanium Carbide Nanosheets for High-Performance Oxygen Evolution. Angew. Chem. Int. Ed. 2016, 55, 1138–1142. [Google Scholar] [CrossRef]
  110. Zhu, Q.L.; Xu, Q. Metal-Organic Framework Composites. Chem. Soc. Rev. 2014, 43, 5468–5512. [Google Scholar] [CrossRef]
  111. Zhao, S.L.; Wang, Y.; Dong, J.C.; He, C.T.; Yin, H.J.; An, P.F.; Zhao, K.; Zhang, X.F.; Gao, C.; Zhang, L.J.; et al. Ultrathin Metal–Organic Framework Nanosheets for Electrocatalytic Oxygen Evolution. Nat. Energy 2016, 1, 16184. [Google Scholar] [CrossRef]
  112. Yu, H.; Wang, Y.H.; Jing, Y. MXene-Based Nanocomposites: Surface Modified MXene-Based Nanocomposites for Electrochemical Energy Conversion and Storage. Small 2019, 15, 1970133. [Google Scholar] [CrossRef]
  113. Li, Z.; Yu, L.; Milligan, C.; Ma, T.; Zhou, L.; Cui, Y.; Qi, Z.; Libretto, N.; Xu, B.; Luo, J.; et al. Two-dimensional transition metal carbides as supports for tuning the chemistry of catalytic nanoparticles. Nat. Commun. 2018, 9, 5258. [Google Scholar] [CrossRef] [PubMed]
  114. Seh, Z.W.; Fredrickson, K.D.; Anasori, B.; Kibsgaard, J.; Strickler, A.L.; Lukatskaya, M.R.; Gogotsi, Y.; Jaramillo, T.F.; Vojvodic, A. Two-dimensional molybdenum carbide (MXene) as an efficient electrocatalyst for hydrogen evolution. ACS Energy Lett. 2016, 1, 589–594. [Google Scholar] [CrossRef]
  115. Geng, D.C.; Zhao, X.X.; Chen, Z.X.; Sun, W.W.; Fu, W.; Chen, J.Y.; Liu, W.; Zhou, W.; Loh, K.P. Direct Synthesis of Large-Area 2D Mo2C on In Situ Grown Graphene. Adv. Mater. 2017, 29, 1700072. [Google Scholar] [CrossRef]
  116. Chaitoglou, S.; Giannakopoulou, T.; Speliotis, T.; Vavouliotis, A.; Trapalis, C.; Dimoulas, A. Mo2C/graphene heterostructures: Low temperature chemical vapor deposition on liquid bimetallic Sn-Cu and hydrogen evolution reaction electrocatalytic properties. Nanotechnology 2018, 30, 125401. [Google Scholar] [CrossRef]
  117. Yang, Y.; Wang, Y.; Xiong, Y. In Situ X-ray Absorption Spectroscopy of a Synergistic Co-Mn Oxide Catalyst for the Oxygen Reduction Reaction. J. Am. Chem. Soc. 2019, 146, 1463–1466. [Google Scholar] [CrossRef]
  118. Massel, F.; Ahmadi, S.; Hahlin, M.; Liu, Y.S.; Guo, J.H.; Edvinsson, T.; Rensmo, H.; Duda, L.C. Transition metal doping effects in Co-phosphate catalysts for water splitting studied with XAS. J. Electron Spectrosc. Relat. Phenom. 2018, 224, 3–7. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of the synthetic route for d-NiC0.2NS/Ni/CF. By using nickel sulfate and polystyrylpyridine resins (PSP) as the nickel and carbon precursor, the samples were obtained via a controllable and mild electrodeposition process with ultrasonic treatment at 60 °C Reproduced with permission from [48]. Copyright 2016 Journal of Materials Chemistry A.
Figure 1. Schematic illustration of the synthetic route for d-NiC0.2NS/Ni/CF. By using nickel sulfate and polystyrylpyridine resins (PSP) as the nickel and carbon precursor, the samples were obtained via a controllable and mild electrodeposition process with ultrasonic treatment at 60 °C Reproduced with permission from [48]. Copyright 2016 Journal of Materials Chemistry A.
Catalysts 10 01164 g001
Figure 2. Schematic illustration of the synthesis process for NiC/MoC/NiMoO4/NF. Reproduced with permission from [46]. Copyright 2019 Chemistry-An Asian Journal.
Figure 2. Schematic illustration of the synthesis process for NiC/MoC/NiMoO4/NF. Reproduced with permission from [46]. Copyright 2019 Chemistry-An Asian Journal.
Catalysts 10 01164 g002
Figure 3. Schematic illustration of the synthetic route for Co/Ni-350. Reproduced with permission from [64]. Copyright 2019 International Journal of Hydrogen Energy.
Figure 3. Schematic illustration of the synthetic route for Co/Ni-350. Reproduced with permission from [64]. Copyright 2019 International Journal of Hydrogen Energy.
Catalysts 10 01164 g003
Figure 4. (a) OER polarization curves tested in 1.0 M KOH of Co-Ni3C/Ni@C. (b) OER polarization curves tested in 1.0 M KOH of Co-Ni3C/Ni@C, Ni3C/Ni@C and IrO2. (c) Tafel plots of different samples. (d) Polarization curves of Co-Ni3C/Ni@C before and after 1000 cycles, and the chronoamperometric response at a constant current of 10 mA/cm2 of Co-Ni3C/Ni@C. Reproduced with permission from [64]. Copyright 2019 International Journal of Hydrogen Energy.
Figure 4. (a) OER polarization curves tested in 1.0 M KOH of Co-Ni3C/Ni@C. (b) OER polarization curves tested in 1.0 M KOH of Co-Ni3C/Ni@C, Ni3C/Ni@C and IrO2. (c) Tafel plots of different samples. (d) Polarization curves of Co-Ni3C/Ni@C before and after 1000 cycles, and the chronoamperometric response at a constant current of 10 mA/cm2 of Co-Ni3C/Ni@C. Reproduced with permission from [64]. Copyright 2019 International Journal of Hydrogen Energy.
Catalysts 10 01164 g004
Figure 5. (a) OER polarization curves tested in 0.5 M H2SO4 of an N-WC nanoarray, IrO2, and 20 wt % Ir/C. (b) O2 concentration in 0.5 M H2SO4 when different potentials are intermittently applied to an N-WC nanoarray electrode. (c) Stability test of an N-WC nanoarray for OER at 10 mAcm−2. (d) X-ray diffraction (XRD) patterns of an N-WC nanoarray before and after an OER test. Reproduced with permission from [69]. Copyright 2018 Nature Communications.
Figure 5. (a) OER polarization curves tested in 0.5 M H2SO4 of an N-WC nanoarray, IrO2, and 20 wt % Ir/C. (b) O2 concentration in 0.5 M H2SO4 when different potentials are intermittently applied to an N-WC nanoarray electrode. (c) Stability test of an N-WC nanoarray for OER at 10 mAcm−2. (d) X-ray diffraction (XRD) patterns of an N-WC nanoarray before and after an OER test. Reproduced with permission from [69]. Copyright 2018 Nature Communications.
Catalysts 10 01164 g005
Figure 6. Synthesis method for Fe3C/Fe2O3@NGNs. Reproduced with permission from [43]. Copyright 2019 Carbon.
Figure 6. Synthesis method for Fe3C/Fe2O3@NGNs. Reproduced with permission from [43]. Copyright 2019 Carbon.
Catalysts 10 01164 g006
Figure 7. (a) ORR LSV curves of N-doped graphene (NGNs), Fe3C/Fe2O3@NGNs, and commercial Pt/C electrocatalysts in O2-saturated 0.1 M KOH condition. (b) The ORR Tafel plots of NGNs, Fe3C/Fe2O3@NGNs, and Pt/C electrocatalysts obtained by fitting LSV curves using linear equation at 1600 rpm. (c) OER LSV curves of Fe3C/Fe2O3@NGNs, NGNs and IrO2 in N2-saturated 0.1 M KOH. (d) The OER Tafel plots of Fe3C/Fe2O3@NGNs, NGNs and IrO2. Reproduced with permission from [43]. Copyright 2019 Carbon.
Figure 7. (a) ORR LSV curves of N-doped graphene (NGNs), Fe3C/Fe2O3@NGNs, and commercial Pt/C electrocatalysts in O2-saturated 0.1 M KOH condition. (b) The ORR Tafel plots of NGNs, Fe3C/Fe2O3@NGNs, and Pt/C electrocatalysts obtained by fitting LSV curves using linear equation at 1600 rpm. (c) OER LSV curves of Fe3C/Fe2O3@NGNs, NGNs and IrO2 in N2-saturated 0.1 M KOH. (d) The OER Tafel plots of Fe3C/Fe2O3@NGNs, NGNs and IrO2. Reproduced with permission from [43]. Copyright 2019 Carbon.
Catalysts 10 01164 g007
Figure 8. OER/ORR performance of Fe@C–NG/NCNTs and Pt/C is shown in the figure. Reproduced with permission from [86]. Copyright 2017 Journal of Materials Chemistry A.
Figure 8. OER/ORR performance of Fe@C–NG/NCNTs and Pt/C is shown in the figure. Reproduced with permission from [86]. Copyright 2017 Journal of Materials Chemistry A.
Catalysts 10 01164 g008
Figure 9. Schematic illustration of the synthetic process of an Mo2C@CS electrocatalyst for the oxygen evolution reaction. Reproduced with permission from [51]. Copyright 2017 ChemSusChem.
Figure 9. Schematic illustration of the synthetic process of an Mo2C@CS electrocatalyst for the oxygen evolution reaction. Reproduced with permission from [51]. Copyright 2017 ChemSusChem.
Catalysts 10 01164 g009
Figure 10. Schematic diagram of Con-β-Mo2C@NC synthesis. Reproduced with permission from [102]. Copyright 2019 Chemical Communications (Cambridge England).
Figure 10. Schematic diagram of Con-β-Mo2C@NC synthesis. Reproduced with permission from [102]. Copyright 2019 Chemical Communications (Cambridge England).
Catalysts 10 01164 g010
Figure 11. (a) OER IR-corrected polarization curves of Mo2C@C, N:Mo2C@NC, and B,N:Mo2C@BCN catalysts in 1.0 M KOH loaded on an NF electrode. (b) Comparison of η10 values with other electrocatalysts based on Mo. (c) Tafel plots. (d) Electrode stability. (e) LSV curves and comparison of η100 values (inset) before and after a 20-h durability test. (f) Polarization curves of a two-electrode electrolyzer using cathode-anode combinations of B,N:Mo2C@BCN-B, N:Mo2C@BCN, and Pt/C-RuO2@NF catalysts at a scanning rate of 2 mV/s and comparison of current densities at a cell voltage of 1.8 V before and after choronoamperometry (inset). (g) Electrolyzer stability test by choronoamperometry. (h) Pulse chronopotentiometric curve for H2/O2 evolution in a B,N:Mo2C@BCN electrolyzer with the current density of −10/10 mA·cm−2. The reversion time interval between the anode and cathode was 600 s in 1.0 M KOH solution. Reproduced with permission from [53]. Copyright 2018 American Chemical Society Catalysis.
Figure 11. (a) OER IR-corrected polarization curves of Mo2C@C, N:Mo2C@NC, and B,N:Mo2C@BCN catalysts in 1.0 M KOH loaded on an NF electrode. (b) Comparison of η10 values with other electrocatalysts based on Mo. (c) Tafel plots. (d) Electrode stability. (e) LSV curves and comparison of η100 values (inset) before and after a 20-h durability test. (f) Polarization curves of a two-electrode electrolyzer using cathode-anode combinations of B,N:Mo2C@BCN-B, N:Mo2C@BCN, and Pt/C-RuO2@NF catalysts at a scanning rate of 2 mV/s and comparison of current densities at a cell voltage of 1.8 V before and after choronoamperometry (inset). (g) Electrolyzer stability test by choronoamperometry. (h) Pulse chronopotentiometric curve for H2/O2 evolution in a B,N:Mo2C@BCN electrolyzer with the current density of −10/10 mA·cm−2. The reversion time interval between the anode and cathode was 600 s in 1.0 M KOH solution. Reproduced with permission from [53]. Copyright 2018 American Chemical Society Catalysis.
Catalysts 10 01164 g011
Figure 12. MXene materials have three different formulas (M2XTX, M3X2TX, and M4X3TX) and three different structures, including mono-M elements, solid solution double-M elements, and ordered double-M elements. In the ordered double-M elements MXene, one transition metal forms the peripheral layers, while the other occupies the central layer, whereas the solid solution double-M MXene has no sequence like this. Reproduced with permission from [40]. Copyright 2019 Small.
Figure 12. MXene materials have three different formulas (M2XTX, M3X2TX, and M4X3TX) and three different structures, including mono-M elements, solid solution double-M elements, and ordered double-M elements. In the ordered double-M elements MXene, one transition metal forms the peripheral layers, while the other occupies the central layer, whereas the solid solution double-M MXene has no sequence like this. Reproduced with permission from [40]. Copyright 2019 Small.
Catalysts 10 01164 g012
Figure 13. (a) Fabrication, hybridization, and advantages of the 3D MXene architecture. (b) Images, dispersion, and elements mapping of CoP@3D Ti3C2-MXene material. Figure (b1,b2) are the scanning electron microscopy (SEM) images of CoP@3D Ti3C2-MXene catalyst. Besides, Figure (b3b5) exhibit the transmission electron microscopy (TEM) images of CoP@3D Ti3C2-MXene material, which demonstrate the dispersion of CoP NPs on the surface of 3D MXene. Figure (b6) is the high resolution transmission electron microscopy (HRTEM) image of CoP NPs, which reveals the clear lattice fringes from the plane of CoP crystal and proves the single crystalline nature of CoP NPs. Figure (b7) is the element mapping that displays the homogeneous scattering of C, Co, Ti, and P elements in the CoP@3D Ti3C2-MXene architecture. (c) The IR-corrected LSV, (d) Tafel plots, and (e) chronopotentiometric response at a current density of 10.0 mA·cm−2 of CoP@3D Ti3C2-MXene, CoP@Ti3C2-MXene (different architecture), CoP@3D reduced graphene oxide (rGO), CoP, 3D Ti3C2 MXene (different hybridization), and RuO2 (commercial catalyst) catalysts for OER in alkaline solution (1.0 M KOH). Reproduced with permission from [37]. Copyright 2018 American Chemical Society Nano.
Figure 13. (a) Fabrication, hybridization, and advantages of the 3D MXene architecture. (b) Images, dispersion, and elements mapping of CoP@3D Ti3C2-MXene material. Figure (b1,b2) are the scanning electron microscopy (SEM) images of CoP@3D Ti3C2-MXene catalyst. Besides, Figure (b3b5) exhibit the transmission electron microscopy (TEM) images of CoP@3D Ti3C2-MXene material, which demonstrate the dispersion of CoP NPs on the surface of 3D MXene. Figure (b6) is the high resolution transmission electron microscopy (HRTEM) image of CoP NPs, which reveals the clear lattice fringes from the plane of CoP crystal and proves the single crystalline nature of CoP NPs. Figure (b7) is the element mapping that displays the homogeneous scattering of C, Co, Ti, and P elements in the CoP@3D Ti3C2-MXene architecture. (c) The IR-corrected LSV, (d) Tafel plots, and (e) chronopotentiometric response at a current density of 10.0 mA·cm−2 of CoP@3D Ti3C2-MXene, CoP@Ti3C2-MXene (different architecture), CoP@3D reduced graphene oxide (rGO), CoP, 3D Ti3C2 MXene (different hybridization), and RuO2 (commercial catalyst) catalysts for OER in alkaline solution (1.0 M KOH). Reproduced with permission from [37]. Copyright 2018 American Chemical Society Nano.
Catalysts 10 01164 g013
Figure 14. (a) OER and (b) HER IR-corrected polarization curves of IrCo@ac-Ti3C2 with a scan rate of 10 mV/s in 1.0 M KOH that demonstrate its bifunctional ability. Reproduced with permission from [39]. Copyright 2020 ChemSusChem.
Figure 14. (a) OER and (b) HER IR-corrected polarization curves of IrCo@ac-Ti3C2 with a scan rate of 10 mV/s in 1.0 M KOH that demonstrate its bifunctional ability. Reproduced with permission from [39]. Copyright 2020 ChemSusChem.
Catalysts 10 01164 g014
Figure 15. (a) Illustration of the hybridization procedures of Ti3C2 MXene nanosheets and a cobalt 1,4-benzenedicarboxylate (CoBDC) MOF. (b) LSV curves, (c) Tafel plots, and (d) Cdl of CoBDC, Ti3C2Tx, and Ti3C2Tx–CoBDC hybrid. (e) Nyquist plots (EIS) of CoBDC, Ti3C2Tx, Ti3C2Tx–CoBDC hybrid, and IrO2. (f) CV curves of Ti3C2Tx–CoBDC hybrid at a scan rate of 20 to 100 mV/s at a potential range of 1.11 to 1.15 V vs. RHE without obvious Faradaic process. (g) The specific surface area of Ti3C2Tx–CoBDC hybrid. Reproduced with permission from [41]. Copyright 2017 American Chemical Society Nano.
Figure 15. (a) Illustration of the hybridization procedures of Ti3C2 MXene nanosheets and a cobalt 1,4-benzenedicarboxylate (CoBDC) MOF. (b) LSV curves, (c) Tafel plots, and (d) Cdl of CoBDC, Ti3C2Tx, and Ti3C2Tx–CoBDC hybrid. (e) Nyquist plots (EIS) of CoBDC, Ti3C2Tx, Ti3C2Tx–CoBDC hybrid, and IrO2. (f) CV curves of Ti3C2Tx–CoBDC hybrid at a scan rate of 20 to 100 mV/s at a potential range of 1.11 to 1.15 V vs. RHE without obvious Faradaic process. (g) The specific surface area of Ti3C2Tx–CoBDC hybrid. Reproduced with permission from [41]. Copyright 2017 American Chemical Society Nano.
Catalysts 10 01164 g015
Figure 16. (a) OER curves, (b) Tafel plots, and (c) polarization curves of NiPS3/C, CoNiPS3/C, and the mosaic CoNiPS3/C nanosheets and nanodots. The inset image in (c) shows the mosaic CoNiPS3/C nanosheets used in water electrolysis. (d) Chronoamperometric curves of nanosheets and nanodots of CoNiPS3/C. Reproduced with permission from [42]. Copyright 2018 Advanced Functional Materials.
Figure 16. (a) OER curves, (b) Tafel plots, and (c) polarization curves of NiPS3/C, CoNiPS3/C, and the mosaic CoNiPS3/C nanosheets and nanodots. The inset image in (c) shows the mosaic CoNiPS3/C nanosheets used in water electrolysis. (d) Chronoamperometric curves of nanosheets and nanodots of CoNiPS3/C. Reproduced with permission from [42]. Copyright 2018 Advanced Functional Materials.
Catalysts 10 01164 g016
Table 1. Important transition metal carbide catalysts employed for water splitting reaction.
Table 1. Important transition metal carbide catalysts employed for water splitting reaction.
CatalystsOverpotential (mV) for OER at a Specific Current DensitySynthesis StrategiesReference
MXene
Co-P/3D Ti3C2 MXene298 @ 10 mA·cm−2Constructing aggregation-resistant 3D Mxene architecture via a capillary-forced assembling strategy (template free).[37]
Ni0.7Fe0.3PS3 @ Ti3C2Tx282 @ 10 mA·cm−2Decorating nickel-based bimetal phosphorus trisulfide nanomosaic on the surface of MXene nanosheets (template free).[38]
IrCo @ ac-Ti3C2220 @ 10 mA·cm−2Incorporating IrCo into basal-plane-porous titanium carbide MXene (template free).[39]
TCCN420 @ 10 mA·cm−2Free-standing flexible films with hierarchically porous structure was constructed from 2D graphitic carbon nitride and titanium carbide nanosheets (template free).[40]
Ti3C2TX−CoBDC410 @ 10 mA·cm−2Hybridizing 2D cobalt 1,4-benzenedicarboxylate (CoBDC) with Ti3C2T x nanosheets via an interdiffusion reaction-assisted process (template free).[41]
CoNiPS3/C262 @ 30 mA·cm−2Simultaneous phosphorization and sulfurization processes and Co-Ni Prussian-blue analogues (PBA) as templating precursor (template assisted).[42]
Fe3C(Cementite)
Fe3C/Fe2O3 @ NGNs-Assembling graphene oxide with graphitic carbon nitride and FeOOH nanorods, nanostructure of carbon layers coated iron species was constructed (template free).[43]
Fe @ C-NG/NCNTs450 @ 10 mA·cm−2Annealing a Fe-based MOF (MIL-88B) loaded with melamine at 800 °C in N2 to get a hybrid structure (template assisted).[44]
NiC
Fe-Ni3C-2%275 @ 10 mA·cm−2Carburizing treatment is carried out in argon/H2 atmosphere and then dope heteroatom of Fe into Ni3C nanodots via a co-precipitation method (template assisted).[45]
NiC/MoC/NiMoO4280 @ 10 mA·cm−2After preparing NiMoO4/NF nanorod arrays, NiMoO4 nanorods react with dopamine at a high annealing temperature to form 3D NiC/MoC/NiMoO4 (template assisted).[46]
Co-Ni3C/Ni @ C325 @ 10 mA·cm−2Co-Ni3C/Ni @ C is prepared via pyrolysis of the bimetallic MOF (Co/Ni-BTC) (template assisted).[47]
d-NiC0.2NS/Ni/CF228 @ 10 mA·cm−2Nanosheets are prepared on nickel-coated copper foil by mild electrodeposition with high index polyhedral dendritic hexagonal NiCx (template free).[48]
Ni/NGNRs380 @ 10 mA·cm−2Mix MWCNTs together with MERCK via a solvothermal process and then collect the resultant mixture for the post treatment of further heating under argon atmosphere (template free).[49]
Mo2C
β-Mo2C267 @ 5 mA·cm−2Reduction of MoO3 with decolorizing carbon (Mo2C-DC) and multiwalled carbon nanotubes (Mo2C-MW) at 950 °C resulted in phase pure Mo2C (template free).[50]
Mo2C @ CS320 @ 10 mA·cm−2Mo2C @ CS is synthesized through a one-pot process with ammonium molybdate and glucose as the Mo and C source (template free).[51]
Ni-Mo2C @ NC328 @ 10 mA·cm−2NiMoO4·xH2O nanobelts serve as a single source for the synthesis of both Ni and MoxC nanoparticles, while melamine simultaneously serves as the precursor and dopant for the formation of Ni-Mo2C @ NC (template assisted).[52]
B,N:Mo2C @ BCN360 @ 100 mA·cm−2Obtaining hybrid structure via an eco-friendly organometallic complex of Mo imidazole and boric acid (template assisted).[53]

Share and Cite

MDPI and ACS Style

Wang, Y.; Wu, Q.; Zhang, B.; Tian, L.; Li, K.; Zhang, X. Recent Advances in Transition Metal Carbide Electrocatalysts for Oxygen Evolution Reaction. Catalysts 2020, 10, 1164. https://doi.org/10.3390/catal10101164

AMA Style

Wang Y, Wu Q, Zhang B, Tian L, Li K, Zhang X. Recent Advances in Transition Metal Carbide Electrocatalysts for Oxygen Evolution Reaction. Catalysts. 2020; 10(10):1164. https://doi.org/10.3390/catal10101164

Chicago/Turabian Style

Wang, Yuanfei, Qimeng Wu, Bicheng Zhang, Lei Tian, Kexun Li, and Xueli Zhang. 2020. "Recent Advances in Transition Metal Carbide Electrocatalysts for Oxygen Evolution Reaction" Catalysts 10, no. 10: 1164. https://doi.org/10.3390/catal10101164

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop