Next Article in Journal
Synthesis of HZSM-5 Rich in Paired Al and Its Catalytic Performance for Propane Aromatization
Next Article in Special Issue
Recent Advances in the Electroreduction of CO2 over Heteroatom-Doped Carbon Materials
Previous Article in Journal
Stereoselective ROP of rac- and meso-Lactides Using Achiral TBD as Catalyst
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Studies on Multifunctional Electrocatalysts for Fuel Cell by Various Nanomaterials

Department of Chemistry, Pusan National University, Busandaehak-ro 63beon-gil, Geumjeong-gu, Busan 46241, Korea
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Catalysts 2020, 10(6), 621; https://doi.org/10.3390/catal10060621
Submission received: 14 May 2020 / Revised: 28 May 2020 / Accepted: 2 June 2020 / Published: 3 June 2020
(This article belongs to the Special Issue Innovative Electrocatalysts for Fuel Cell and Battery Applications)

Abstract

:
Based on nanotechnology, nanocomposites are synthesized using nanoparticles (NP), which have some advantages in terms of multifunctional, economic, and environmental factors. In this review, we discuss the inorganic applications as well as catalytic applications of NPs. Recently, structural defects, heteroatomic doping, and heterostructures of such efficient ideal catalysts and their application as multifunctional catalysts for hydrogen evolution reaction (HER) and oxygen evolution reaction (OER) in water splitting. It has been verified that the catalysts used in oxygen reduction reaction and OER can be used effectively in metal/air batteries. Moreover, it has been reported that high-efficiency catalysts are required to implement urea oxidation reaction (UOR), which involves a six-electron reaction, as an electrochemical reaction. We expect that this review can be applied to sustainable and diverse electrochemistry fields.

1. Introduction

A material comprising two or more constituents is generally referred to as a hybrid material (Scheme 1). Such a hybrid material not only has the unique characteristics of each individual component, but also has additional synergistic characteristics [1,2,3]. Such synergistic characteristics allow their application as catalysts and sensors, as well as for adsorption, desorption, etc., [4]. Because of properties like high selectivity for a particular material, high catalytic efficiency, and high physical/chemical stability compared to catalysts based on one metal, multimetallic fusion nanoparticles (NPs) have been synthesized for use as catalysts in recent years [5,6,7,8,9,10,11,12,13,14]. The synthesis of such multicomponent nanostructures with improved functionalities has been explored by many researchers [15,16,17,18,19,20,21,22,23,24,25].
The realization of a hydrogen-based, renewable, and sustainable energy cycle offers a clear vision to secure resources for our upcoming future. High-performance electrode catalysts for power conversion systems like fuel cells and water splitting reactions are central to this thrust [26,27,28,29].
Hydrogen-producing electrochemical water splitting is a promising technique for renewable energy production and storage [30,31]. In fact, owing to its potential for substituting fossil fuels, the electrochemical conversion of H2O into hydrogen fuel (2H2O → 2H2 + O2) has been successfully achieved (Scheme 2). The water splitting method usually involves two-half processes. One process is the anodic oxygen evolution reaction (OER), and the other is the cathodic hydrogen evolution reaction (HER) [32]. There are literature reviews that show that several electrocatalytic systems for water splitting have been reported with many different elementary formulations and microstructures [33,34,35,36]. With the increased use of wide-throughput combinatorial methods for the development of new OER and HER catalysts in various ternary and quaternary structure spaces, this number is likely to increase rapidly in the future [37,38,39,40,41,42,43].
HER and OER performance can be enhanced based on three aspects (Scheme 3):
  • Defect sites
  • Heteroatoms
  • Heterostructures
Hitherto, surface defect development approaches appear promising for the achievement of high-performance HER electrocatalysis owing to their tunable electrical properties [44,45,46]. Nonetheless, while the quantity of edge sites in the nanostructure increases the HER efficiency, the manufacturing process limits the efficiency [47]. Edge sites enhance the activity by increasing the surface of the catalyst. A common technique involves the design and development of porous, hollow micro, and nanostructured electrocatalysts with a large number of catalytically active sites, reduced weight, and load transportation diffusion [48,49]. This may also lead to the development of the reactivity of the active sites by adding oxygen vacancies, resulting in major improvement in electrocatalytic operations [45].
Over the past decades, a number of earth-abundant metal-based composite materials, including oxides, chalcogenides, nitrides, borides, and phosphides, have been used to catalyze both the HER and OER [50,51,52,53,54,55,56,57,58,59]. Because of their hydrogenase-like catalytic mechanism in the HER and partial creation of active metal oxyhydroxides in the OER, transition metal phosphide (TMP)-based materials demonstrate noteworthy HER and OER activities among these well-explored electrocatalysts [60,61,62,63].
Heteroatom-doping is an effective method for further electrocatalyst activity enhancing by regulating the electronic structure, increasing the number of active sites, and maximizing intermediate formation capacity. In addition, co-doping of binary or ternary heteroatom catalysts can optimize catalytic activity with respect to their single-doped counterparts via synergistic coupling [64].
The heterostructured catalyst comprises two or more kinds of materials, typically bound together physically or chemically. Recently, an attractive variety of heterostructured catalysts were identified, demonstrating exceptional catalytic efficiency against electrochemical water splitting equivalents [65,66,67,68,69]. Most heterostructured catalysts exhibit higher HER behavior than their single-structured counterparts because of the many advantages of the former. Several heterostructured catalysts demonstrate high-quality HER activity over a broad pH spectrum and/or multifunctional HER and OER activity [70].
The oxygen reduction reaction (ORR) and OER are emerging as highly relevant electrochemical reactions owing to the growing emphasis on renewable power generation technologies (Scheme 4) [27,71]. Lithium-oxygen batteries (LOBs) having the extremely high theoretical energy capacity have drawn significant attention in recent years [72,73].
The international research community has recently reported that advanced manufacturing studies, such as transformation from a coal-based system to a H2-based one for effective power storage and sewage disposal will offset many emerging advances in energy and environmental pollutant technologies (Scheme 5) [74,75]. Among the various recently established energy storage electrolysis technologies, urea electrolysis is remarkably intriguing, allowing the simultaneous disposal of urea-rich wastewater through the oxidation of the urea and a healthy, environmentally sustainable mass processing of H2 fuel [76].
In this review, we focus on multifunctional nanomaterials and nanocomposites and their modifications in different fields:
  • Interesting means to increase water splitting in hydrogen and oxygen evolution reactions, such as defect sites, doping of heteroatoms, and heterostructures.
  • Fuel cell technology using ORR and OER.
  • Specific electronic processing with urea decomposition to address environmental problems.

2. Methods of Increasing Water Splitting in HER and OER

2.1. Defect Sites

In case of vacancy defects, one part of the lattice becomes deficient and unstable, thereby enhancing activity with the reactive species [78]. Various defects have been established in NPs such as vacancies and lattice defects. Thus, catalysts with defect sites in their lattice provide more active sites, and consequently, greater reactivity than defect-less catalysts with the same surface area (Scheme 6). In particular, catalysts that intentionally create coalesced defect sites can further enhance the rate of chemical reactions, because the deficiency of the lattice effectively lowers the hydrogen adsorption energy and hydrogen diffusion energy barriers in HER, which are significant for the overall rate of water decomposition [79,80]. In addition, the electrochemical activity of the catalyst should be enhanced by introducing chalcogen into the defective sites, as suggested by the increase in reactivity of the active sites for HER and OER based on the defect sites of oxygen atom [81,82,83,84]. The existing methods of composing defect sites are ball milling, plasma technology, etching, etc. Ball milling is the process of grinding matter in ball mill and turning it into fine powder, and this is the relative adjustment of anti-site defect and vacancies in crystal. The plasma technology is mainly carried out in vacuum chamber, then is applied to gas or electric field to generate plasma after air enters the chamber. From this plasma, the energy of ions often determines defect density. The plasma technology can obtain vacancy defects on a constant basis. Li’s group explain making defect sites through etching method.

2.1.1. Tafel Reaction of MoSe2 through Coalesced Vacancies

A recent study by Li et al. suggested that the sulfur vacancies and tensile strength in molybdenum disulfides can trigger the basal plane with optimization the free energy difference associated with HER [85]. The formation of defect sites has been proven to enhance the HER operation by electrochemical methods, hydrogen annealing, and etching by plasma [85,86,87]. Based on these findings, in this research, the in-situ defect site in MoSe2 has been established via hydrogen reactivity control during the chemical vapor deposition (CVD) process. The vacancy-controlled MoSe2 exhibited outstanding HER behavior, reflected by an unprecedented Tafel reaction. In addition, when MoSe2 has been synthesized in hydrogen and Ar atmospheres as carrier gases, the density of Se vacancies is significantly affected by the concentration of these carrier gases. They can serve either as a reactivity booster by promoting the substitution of MoO3 and Se or as an etching substance for Se atoms in MoSe2, leading to vacancy-controlled MoSe2 [88,89].
Scheme 7 illustrates a synthesis mechanism for MoSe2-X, in which X refers to the relative concentration of H2 in the carrier gas (X = 0, 10, 20, 30, 40, and 50). During the growth process, the vacancy of Se was controlled by regulating the concentration of H2 gas for CVD, as follows:
  • Reducing stable MoO3 to reactive MoO3−x to improve the activity between Se and metal oxides under hydrogen assistance.
  • Selenization of MoO3x to MoSe2 via the substitution reaction of Se and O.
  • Use of the substrate for nucleation and growth of MoSe2.
  • Etching of Se atoms in MoSe2 by generating H2Se gas under hydrogen assistance.
In the growth procedure for MoSe2, hydrogen has significant impact on the etching of Se atoms in MoSe2 (increase in the density of Se vacancies) and on the reduction of MoO3 (reduction in the density of Se vacancies) [88,89]. MoO3 is not completely reduced at low H2 concentration, and MoO3x impurities are produced by the low activity of Se and MoO3 (MoSe2-10). On the other hand, relatively high concentrations of H2 gas (MoSe2-30 and MoSe2-50) facilitates the reaction between Se and MoO3, which inhibits the accumulation of MoO3x impurities.
In the presence of H2 gas, the creation of Se vacancies in MoSe2 from pristine MoSe2 are illustrated in Figure 1a. Annular dark-field-scanning transmission electron microscopy (ADF-STEM) images (Figure 1b) show that MoSe2-30 was synthesized as hexagonal MoSe2 with minimum Se vacancies. However, in MoSe2-50, which was also synthesized as a hexagonal structure with high crystallinity, vacant Se sites were observed more frequently than in MoSe2-30. Further, these vacancy sites were generally formed close to each other, resulting in the formation of coalesced vacancies. The single-Se vacancy sites of Se atoms were described using the mapping of the atomic location depicted in Figure 1c, caused predominantly by the etching of hydrogen.
After dropping vacancy-controlled MoSe2-X materials onto a glassy carbon electrode without using polymeric binding agents, the monolayer’s HER activities were investigated in 0.5 M H2SO4 electrolyte by using a three-electrode framework based on linear sweep voltammetry (LSV) curves (Figure 2a) and Tafel slopes obtained from polarization curves (Figure 2b). The Pt/C polarization curve indicates standard hydrogen production with an onset potential near approximately 0 V vs. RHE and a significantly rising current density. As reported in previous literature, the Tafel slope of Pt/C was 30 mV dec−1, which means that Pt/C follows a Volmer–Tafel mechanism in its HER pathway [90]. For the monolayer electrocatalysts MoSe2-X, the sum of Se vacancies in the MoSe2 lattice is nearly in agreement with the HER activities. Owing to less number of vacancies, MoSe2-30 displayed the lowest HER efficiency with an onset potential of −0.3 V and a gradually rising current density. Although MoSe2-10 contains a large number of vacancies in the MoSe2 lattice, it demonstrates only modest electrochemical performance, possibly owing to the large number of MoO3x impurities present on the surface of the catalyst, interfering with the transition of the carrier needed for the development of hydrogen. Conversely, a dramatic enhancement in HER behavior was observed for MoSe2-50, with a more positive onset potential close to −0.16 V, and a rapidly increasing current density such as in Pt/C. Interestingly, MoSe2-50 with the clean surface exhibited a quite low Tafel slope (33 mV dec−1), close to the onset potential owing to the coalesced vacancies. Further, unlike the typical trend of Tafel plots in transition metal dichalcogenide (TMDs), which increases significantly with increasing voltage, the Tafel plot of MoSe2-50 was retained almost at the same level as 33 mV dec−1 in the high-overpotential area as well. The slope of this Tafel plot is among the smallest recorded to date, including TMDs of the 1T and 2H level, quite similar to that for Pt/C [85,86,87,91,92,93,94]. The HER performance of intermediate samples (MoSe2-20 and MoSe2-40) are also depicted in Figure 2c. By decline of MoO3−x group in MoSe2-20, improved HER performance was triggered compared to MoSe2-10, leading to a lower Tafel slope. The HER activity in MoSe2-40 suggests intermediates between MoSe2-30 and MoSe2-50, primarily owing to less number of active sites than MoSe2-50.
The typical HER in TMDs mostly follows the Heyrovsky mechanism, which requires an extra electron transfer to adsorbed hydrogen on a catalytic surface (H*) to generate H2 (H* + H+ + e → H2). In comparison, the extremely low and steady Tafel slope with a large overpotential for MoSe2-50 is most likely caused by the Tafel reaction, in which only two H* must to be combined to generate H2 without any extra electron transfer (H* + H* → H2). Thus, high HER performance in vacancy-induced MoSe2 relates to the extra Tafel reaction because of the creation of coalesced Se vacancies, as depicted in Figure 2c.

2.1.2. Hollow and Porous Structured Using NiS2(1−x)Se2x in Neutral Condition

In the recent past, neutral-pH overall water electrolysis has been considered the new efficient energy conversion alternative, as it can lower the cost of electrolysis devices without requiring costly membranes, and several electrolysis cells usually operate in neutral pH [95,96,97]. However, owing to low ionic concentrations and significant ohmic loss, electrocatalysts exhibiting excellent HER/OER performances in neutral pH are rare [98,99,100,101]. Only several Co- or Ni-based catalysts, such as CoP, Co3S4, and Ni3Se4, have been known to exhibit moderate performance under neutral media [102,103,104,105]. In addition, most of them exhibit less efficient catalytic performance than noble metals such as Pt- and Ru/Ir-based catalysts and generally suffer major long-term stability losses [104,105,106]. Under such situations, optimizing the relatively low-cost bifunctional electrocatalysts for water splitting with high performance and stability in neutral media is a crucial task, even though it entails great challenges. In recent years, TMDs have drawn considerable interest in terms of their ideal atomic structure, good electrochemical stability, and high electron transport, and many related studies on water splitting have been published [28,107,108,109]. Among them, MS2(1−x)Se2x compounds, referred to as metal (di)sulfoselenide, provide a fascinating promise, enabling accurate customizable electrocatalytic and other physiochemical characteristics. Zhuo et al. reported that ternary MS2(1−x)Se2x can accomplish a relatively low HER overpotential (69 mV at 10 mA cm−2) through customizable HER electrical properties under acidic conditions [110]. Similarly, improved catalytic efficiency was also reported for ternary Mo(Se0.85S0.15)2, WS2(1−x)Se2x, and Co(SxSe1-x) [111,112,113,114]. Although these indicated electrocatalysts can exhibit high HER/OER performances under acid or alkaline conditions, few of them show neutral-pH reactivity.
In this study, Ni-based catalysts, NiS2(1-x)Se2x, were fabricated and investigated to confirm HER/OER activities owing to their advantages such as eco-friendly features and low cost of Ni. Further, Ni dichalcogenides have exhibited relatively high activities for either HER or OER under neutral-pH conditions [97,115]. Furthermore, to obtain a large number of catalytic active sites and decrease diffusion of charge and mass transport, one common strategy is to model and develop porous/hollow micro/nanostructured catalysts [48,49]. With these discussions, ternary nickel sulfoselenides (NiS2(1−x)Se2x) porous/hollow spheres were synthesized via a facile hydrothermal using the L-cysteine and selenization method.
Scheme 8 depicts the fabrication process for the ternary NiS2(1−x)Se2x porous/hollow spheres. Ni(NO3)2·6H2O, urea, and L-cysteine were used to synthesize the NiS2 hollow spheres precursor via heating at 140 °C for 12 h, and the as-synthesized precursors were dried in a vacuum oven at 60 °C for 6 h. Thereafter, for selenization, three different mass ratios of NiS2 and Se powder were used (1:2, 1:5, and 1:10), which were heated to 450 °C in a furnace for 2–5 h.
Figure 3a exhibits the morphological aspects of the as-prepared NiS2 precursor by transmission electron microscopy (TEM), minimizing the interfacial energy of hollow spheres based on the Ostwald ripening process. Subsequently, NiS2(1−x)Se2x porous/hollow spheres were synthesized by thermal selenization using the NiS2 precursor and Se powder in Ar gas. In this process, the Se/S atomic ratio was adjusted by varying the amount of Se powder and the reaction time (Figure 3b). Figure 3c, exhibiting the high-resolution TEM (HRTEM) image and the corresponding line scan outlined in a white rectangle, indicates the well-matched lattice fringes of the (200) plane with a lattice distance of 0.291 nm, slightly larger than that of standard NiS2 (0.284 nm). The HAADF image and the corresponding EDX mapping of Ni(S0.5Se0.5)2 demonstrate the presence and homogeneous distribution of Ni, S, and Se element around the whole hollow structure (Figure 3d).
The powder X-ray diffraction (PXRD) patterns exhibit the NiS2 hollow spheres before and after the thermal selenization process of a Pa3, corresponding to JCPDS no. 88-1709 (Figure 4a). However, the diffraction peaks of Ni(S0.5Se0.5)2 shift to a small angle compared to those of pure NiS2 because of the larger atomic sizes of Se than S (inset of Figure 4a). The ternary NiS2(1−x)Se2x compound is shown to have combined effects. Atoms of the transition metal Ni are offered as superior water splitting sites and have a strong electrostatic affinity with the hydroxide group in neutral media owing to plenty of unfilled d orbitals. In addition, the morphology modified by selenization offers a lough surface area, and the highly porous structure not only provides additional exposed catalytically active sites but also facilitates more active species to access the active sites. In addition, intrinsic catalytic activity can be improved by generating electrolytic diffusion/penetration and mass/electron diffusion systems more effectively. The results have been indicated using the N2 sorption isotherm (Figure 4b). The unusual hierarchical hollow spheres were found to have a wide Brunauer-Emmett-Teller (BET) surface area of 91 m2 g−1 with a pore diameter distribution primarily within the range 4–10 nm. Furthermore, anion doping efficiently modifies the electronic Ni structure, which is helpful in maximizing the adsorption free energy of process intermediates and improving the catalytic ability intrinsic to them. Therefore, this ternary NiS2(1−x)Se2x material was considered to be a highly efficient and stable electrocatalyst for overall water splitting in neutral media.
Figure 5a indicates that, among the samples, Ni(S0.5Se0.5)2 has the highest HER activity. It also displays the lowest Tafel slope (Figure 5b). The lowest charge transfer resistance was demonstrated by electrochemical impedance spectroscopy, performed to investigate more detailed electron-transfer kinetics (Figure 5c). Under neutral conditions, the HER pathway may be identical to that in alkaline conditions via the Volmer–Heyrovsky cycle. As indicated by the abovementioned results, the Ni atom has a strong electrostatic affinity with the hydroxide group for water dissociation ability to produce H atoms for the primary Volmer step. However, in Heyrovsky’s process, the free energy of H adsorption, when H atoms on the layer of catalysts are formed into H2, is such that the disulfided Ni site is too strong, whereas the diselenided Ni site is too weak. Therefore, by changing the Se/S ratio, the free energy of H adsorption, which can increase the HER operation, is optimized. In the chronoamperometry (CA) performed to demonstrate long-term durability, a significant property in electrocatalysts, it remained without degradation for 20 h, and the HER performance was mostly unchanged after 2000 continuous cyclic voltammetry (CV) cycles (Figure 5d).
The electrocatalytic oxygen generation was performed under the same conditions (Figure 6a). Similarly, the polarization curve had an overpotential of 501 mV to reach the current density of 10 mA cm−2 of Ni(S0.5Se0.5)2 among samples, compared to the standard IrO2 catalyst, and the lowest Tafel slope of 94 mV dec−1 (Figure 6b). This resulted in a desirable electronic configuration and specific active sites for adsorption through the anion-doping activity of oxygen-rich groups. Similar to previously reported mechanisms for other metal chalcogenide electrocatalysts, the Ni cation exists as an actual active species with the oxide species (NiOx) on the surface of the OER catalyst and forms the shell/core NiOx/Ni(S0.5Se0.5)2. Similar to the tendency of HER, the Ni(S0.5Se0.5)2 electrode with the smallest Tafel slope indicates the lowest resistance for charge transfer (Figure 6c). Performing CA to demonstrate long-term durability, Ni(S0.5Se0.5)2 electrode demonstrated good durability at a constant applied voltage value of 1.73 V without degradation for 20 h (Figure 6d). The OER performance was almost unchanged after 2000 continuous CV cycles (inset of Figure 6d).
Based on the above mentioned results, the ternary NiS2(1−x)Se2x electrocatalysts were superior to other catalysts in terms of electrocatalytic efficiency under neutral conditions. Ni(S0.5Se0.5)2 hollow/porous spheres were used for both the cathode and the anode to perform the overall water splitting. As depicted in Figure 7a, the Ni(S0.5Se0.5)2‖Ni(S0.5Se0.5)2 electrode requires 1.87 V at a current density of 10 mA cm−2, whereas a typical Pt/C‖IrO2 couple electrode requires 1.94 V at the same current density. In comparison, the Ni-assembled electrolyzer could sustain current density at a steady voltage of 1.87 V for 12 h under neutral conditions, in contrast to the rapid decline in current density for the Pt/C‖IrO2 couple-catalyzed cell under CA (Figure 7b). These results demonstrated that the latest method involving bifunctional electrocatalyst under neutral conditions is a successful practical application for the complete splitting of water.

2.2. Doping of Heteroatoms

Doping of heteroatoms, such as N, O, S, and P, is effective for increasing electrocatalyst activity by optimizing the electronic structure, increasing the number of active sites, and optimizing the intermediate formation regime (Scheme 9). Also, heteroatom doping can give graphene a variety of new or improved electromagnetic, physicochemical, optical, and structural properties. This greatly increases the arsenal of graphene materials and their potential for a wide range of applications. For example, Lee et al. reported that MoP dual-doped with S and N exhibited a current density of 10 mA cm−2 for HER at an overpotential of 104 mV in 0.5 M H2SO4 [117]. Zang et al. recorded excellent catalytic efficiency for the HER with an overpotential of just 89 mV to give 10 mA cm−2 in 0.5 M H2SO4 for the Cu3P NPs coated by an N, P-codoped carbon layer [118]. Furthermore, they reported that MOF-derived N, S, O-C@Co/Co9S8 NPs acted as a multifunctional catalyst producing a current density of 10 mA cm−2 at 216 and 373 mV for HER and OER, respectively, in 1 M KOH [119]. Du’s group reported Ir-doped NiV(OH)2 for overall water splitting [120].

2.2.1. Ir-O-V Group Effect

The overall splitting of water in Ni-based layered double hydroxides (LDHs) occurs at quite high applied voltages, usually greater than 1.55 V [121,122,123]. NiV LDH was the initiating material, and doped Ir to create Ir-O-V and Ir atoms helped dissociate water and adjust the V atom and bridge O adsorption energies. The resulting NiVIr LDH exhibited superior HER and OER performances at overpotentials of 41 and 203 mV, respectively, at 10 mA cm−2. The new NiVIr LDH achieved a total water electrolysis current density of 10 mA cm−2 at an approximate voltage of 1.49 V (as both cathode and anode), far lower than that of the Pt/C-IR/C catalytic complexes (1.60 V at 10 mA cm−2) and seemed to set a new value for the overall water electrolysis cycle.
Figure 8a,b indicates that OER and HER are supposed to occur on V and bridge O, respectively, in the NiV LDH material. As seen in Figure 8b, after one Ni atom had been substituted by Ir, a simple decrease in the charge intensity was determined on the bridge oxygen atom located between the atoms V and Ir. The excessive adsorption of the hydrogen intermediates on the NiV LDH bridge oxygen could be reduced. Doping with Ir, however, seems to enhance charge density on V atoms, thereafter, decreasing upon excess adsorption of OER intermediates. In short, the Ir dopant is supposed to improve both OER and HER performance in the NiVIr LDH simultaneously. Morphologically, all three materials were identical, consisting of vertically grown nanosheet arrays on nickel foam (NF), as illustrated by scanning electron microscopy (SEM) images (Figure 8c). XRD patterns (Figure 8d) indicate that the NiV LDH and NiVIr LDH materials had phase compositions similar to that of α-Ni(OH)2, and no additional diffraction peaks are found, indicating that the structure of the three materials were in one phase [124,125,126]. The (003) and (006) peaks of NiVIr LDH (Figure 8d) were shifted to lower diffractions angles by ~0.3° relative to NiV LDH, which suggests that the insertion of large Ir ions into NiV LDH induced lattice expansion.
The HER and OER performances of the catalysts were analyzed with 1 M KOH in N2- and O2-saturated solutions, respectively. Two reactions were practiced in the three-electrode system—the counter-electrode of the HER is the graphite rod and that of the OER is the platinum electrode [127,128,129]. The LSVs of the OER presented in Figure 9a indicate that the Ir-doped LDH demonstrates a lower overpotential of 203 mV to fulfill a current density of 10 mA cm−2, which is better than those of the Ni(OH)2, NiV LDH, and commercial Ir/C. The Ir-doped LDH material indicated an excellent HER activity, which has been proven by a low overpotential of 41 mV at 10 mA cm−2 (Figure 9b). To compare, the overpotential of the Ir-free LDH depicted at 10 mA cm−2 was 148 mV. Table 1 indicates the effects of the OER and HER performance calculated by Du’s group [120].
Du’s group developed a new bifunctional catalyst comprising surface Ir-O-V catalytic groups, which were very effective for overall water electrolysis. The NiVIr LDH displays good catalytic attributes with an OER overpotential of 203 mV and an HER overpotential 41 mV at 10 mA cm−2. The NiVIr LDH required 1.493 V at 10 mA cm−2, performing better than the noble metallic Pt/C-Ir/C complex. More precisely, Ir, O, and V atoms of NiVIr LDH change the adsorption energy and electronic structure of the reaction course to acceptable values, which remarkably decrease the OER and HER overpotentials. Du’s experiment demonstrates that doping with noble metals is an efficient way for improving the catalytic role of LDHs, which improves their ability and decreases expensive metal consumption.

2.2.2. Effect of Multiplane Ni3S2 Superstructures

Liu’s group suggested a new simple method for the synthesis of 3D Ni3S2 superstructures grown in situ on NFs using a one-step method of chemical etching (Scheme 10). Only the controlling solvent of the reaction allows the superstructures to be transformed into both nanoforest patterns and rod-like arrays. The technique is facile, economical, low-temperature, and has good commercial application prospects. Scheme 10 illustrates Ni3S2 superstructures formed on NF in disparate solvents (DI water, EDA). Arranged from disparate solvents, the 3D superstructures of the two species can be derived from the diverse kinetics of crystallization in disparate solvents. Thus, the product superstructures could be handled by regulating the chemical etching elements.
The Ni3S2 nanoforest and nanorod superstructures were further investigated using TEM and an analysis of the SAED patterns (Figure 10a–d). Figure 10a depicts a diameter of 200 nm and a length of approximately 1.5 μm, which corresponds with the SEM images. The rod tip amplification image (Figure 10b) clearly indicates a lattice at a distance of 0.28 nm, which can be related with the (110) crystal plane of the Ni3S2. Similarly, each nanoforest image (Figure 10c,d) is checked, and the branched structure is surrounded by an ultrathin nanosheets layer, synthesizing 3D hierarchical nanostructures. Since the electroactive sites are abundant, these ultrathin nanosheets offer outstanding electrochemical catalytic efficiencies.
Ni3S2 was commonly used as water electrolysis catalysts because of its remarkable conductivity and activity [131,132,133]. To assess the electrocatalytic reactivity toward HER, tests were conducted on the synthesized 3D Ni3S2/NF superstructures in a 1.0 M KOH solution using a standard three-electrode system, in which the working electrode used the 2.0 cm2 NF. The electrocatalytic reactivities of Pt/C on NF and bare NF were analyzed for comparison. Figure 11a displays their LSV curves. The synthesized superstructures of Ni3S2/NF display exceptional HER performance, in which the current densities are intensified at the used voltage. Rod-like array (RA)-Ni3S2/NF demands overpotentials of 0.146 and 0.255 V to provide current densities of 10 and 100 mA cm−2, whereas NF-Ni3S2/NF demands those of 0.135 and 0.225 V. The NF-Ni3S2/NF has ultrathin leaflets, which decrease the overpotentials because leaflets provide more active sites for HER. Durability is also a significant criteria with respect to the electrocatalyst of the HER reaction. The long-period durability of the electrocatalysts was tested continuously in 1.0 M KOH over 5000 cycles (Figure 11b). Notably, both NF-Ni3S2/NF and RA-Ni3S2/NF exhibited a small loss of the current density of the cathode after 5000 CV cycles confirming the excellent durability. RA-Ni3S2/NF and NF-Ni3S2/NF were maintained at 15 and 20 mA cm−2, respectively, when CA was measured for 50 h at a constant 150 mV vs. RHE (inset of Figure 11b).
Further research was carried out on the 3D Ni3S2 superstructures to assess the OER output in 1.0 M KOH. Figure 11c presents the LSV curves. All the 3D Ni3S2 superstructures demonstrated still higher efficiencies than bare NF. The 100-mA cm−2 current densities of NF-Ni3S2/NF and RA-Ni3S2/NF exhibited 0.32 and 0.34 V overpotentials, respectively. Bare NF, in contrast, provided 100 mA cm−2 at an overpotential of 0.51 V. Similar to the HER results, in OER, NF-Ni3S2/NF displayed a lower overpotential than RA-Ni3S2/NF. The Liu group found that NF-Ni3S2/NF could display better OER performance relative to RA-Ni3S2/NF because of the abundance of particular surface areas of ultrathin nanosheets in NF-Ni3S2/NF, resulting in further active sites toward OER. Both NF-Ni3S2/NF and RA-Ni3S2/NF exhibited overpotentials of 0.23 and 0.27 V at 20 mA cm−2, respectively, which displayed lower overpotential than other existing non-noble metal electrocatalysts [130,134,135,136,137,138,139,140,141,142]. To obtain the long-period durability of NF-Ni3S2/NF and RA-Ni3S2/NF toward OER, a long-period water oxidation was performed at 300 mV value of overpotential, and the OER reactions remained almost unchanged for 50 h (Figure 11d).
The Liu group further constructed an electrolyzer whose cathode and anode were both composed of the synthesized catalysts in a solution of 1.0 M KOH. NF-Ni3S2/NF and RA-Ni3S2/NF impressively produced current densities of 10 mA cm−2 at 1.59 and 1.62 V and combined overpotentials of approximately 360 and 390 mV, respectively (Figure 12a). The 3D Ni3S2/NF performed better than the IrO2/C-Pt/C commercial couple (1.68 V at 10 mA cm−2) and the newly reported overall water splitting electrocatalysts [125,141,143,144,145,146,147,148]. In addition, when the experiments were executed at 1.63 V, the electrolyzer supplied steady current densities (RA-Ni3S2/NF is approximately 20 mA cm−2 and NF-Ni3S2/NF is approximately 30 mA cm−2) and maintained then for more than 50 h (Figure 12b).
The 3D Ni3S2 superstructures of nanorod arrays and nanoforest patterns were well grown in NF through a simple one-step method of chemical etching. The synthesized 3D Ni3S2 superstructures demonstrated considerable efficiency in HER, and the stability remained for more than 50 h. In addition, the 3D Ni3S2 superstructures consisted of a cathode and an anode for bifunctional water electrolysis with outstanding activity and notable long-time duration.

2.3. Heterostructure

The heterostructure idea originates from the physics of semiconductors (Scheme 11). In general, heterostructures can be characterized as hybrid structures consisting of interfaces formed using various solid-state materials, such as conductors, insulators, and semiconductors [70]. Heterostructures provide an ability to produce effective solar cells from highly absorbing thin-film materials without major losses by electron-hole recombination on the front side. Normally, heterostructures of 2D materials offer a way to study different phenomena and open up unprecedented possibilities to combine them for technological use. Such layers are very different from previous 3D semiconductor heterostructures, as each layer acts simultaneously as a bulk material and an interface, reducing the amount of load displacement within each layer. Nevertheless, the movement of charges between the layers may be very large, producing huge electrical fields and providing intriguing possibilities in the field of band-structure engineering. However, the key problem with hetero junctions is the lattice mismatch caused defect, which usually degrades the output of the system. It is normal for two different materials to have specific lattice parameters. If these components are put on top of each other.

2.3.1. Components as the Ni-Based Heterostructure

In general, Ni-based composites are attractive catalysts in water splitting reactions because of the earth-abundance and low cost of Ni, and the components can affect their performance [149,150,151,152,153,154,155,156]. In addition, it has been revealed that Ni-based compounds can be used as catalysts for OER only by controlling counter-ion forms. The integration of the active OER (Ni3N) and HER (NiMoN) catalysts as heterostructures was stated by Fu’s group through a well-designed path for efficient water splitting [156].
Figure 13 demonstrates the protocol for the synthesis of Ni3N-NiMoN heterostructures. Ni-Mo-O nanosheets were synthesized on carbon cloth (CC) by hydrothermal reaction, and then the Ni3N-NiMoN heterostructure was synthesized in the presence of ammonia gas by nitridation with Ni-Mo-O nanosheets. Bonding between Ni3N and NiMoN can be easily formed owing to Ni-Mo-O precursor containing Ni and Mo.
The Ni3N-NiMoN nanosheet structures on CC are well maintained without any apparent collapse (Figure 14a,b). The lattice distances in the HRTEM images are 0.21 and 0.25 nm, which exhibit a good match with the (111) plane of Ni3N and the (100) plane of Ni0.2Mo0.8N, respectively. We have already mentioned that presence of Ni3N and NiMoN can make both the catalysts perform good OER and HER activity. The elemental mapping of Mo (yellow), Ni (cyan), and N (green) demonstrated uniform dispersion on the CC (Figure 14c–f) because Ni-N and Ni-Mo-N were successfully synthesized by controlling the reaction in the presence of ammonia gas.
One of the significant factors in the catalyst is long-term stability. A durability experiment is carried out in 5000 cycles using the Ni3N-NiMoN-5 catalyst for 20 h (Figure 15b). The results indicate that the Ni3N-NiMoN-5 catalyst exhibits excellent stability under alkaline conditions in a long-term electrochemical system (Ni3N-NiMoN-5: 277 mV at 10 mA cm−2, Ni3N-NiMoN-4, Ni3N-NiMoN-6, RuO2, and NiMoO-1 are about 300 mV, 343 mV, 325 mV, and 356 mV at 10 mA cm−2, respectively). These results of the excellent HER activity observed in an alkaline system imply that the catalyst can easily cooperate without most noble metals such as Ru, Pt, and Ir OER catalysts, which typically exhibit good behavior in alkaline media. Figure 15c presents the polarization curves without the iR-corrections. The Ni3N-NiMoN-5 catalyst exhibits the high-performance OER output and needs a modest overpotential of 277 mV to reach the current densities of 10 mA cm−2, whereas those for RuO2 and NiMoO-1 are approximately 325 and 356 mV. Fu’s group has already demonstrated Ni3N-based materials to be active for OER in several earlier studies [152,153,157]. Therefore, earlier reports have suggested that Ni3N is an effective catalyst for OER. Further, the long-term durability of Ni3N-NiMoN-5 for OER was implemented based on its relevance for scientific basis (Figure 15d). The Ni3N-NiMoN-5 catalyst’s polarization curve exhibits a slight change in current density compared to the original. In addition, the time–current density curve indicated that after 20 h of OER activity, the operation degrades only marginally.
In this study, Ni3N-NiMoN could be specifically applied to the anode (HER) and cathode (OER), as the overall water splitting catalysts demonstrate high performance under the same alkaline state.

2.3.2. Co and β-Mo2C NPs Composed of N-Doped CNTs

Recently, it has been reported that complexes of Co-based hybrid materials such as Co-N, Co-Pi, NiCo2O4, NiCo alloy, and others can improve catalytic performance in the OER [158,159,160,161,162]. Moreover, M-Cx, M-Sx, M-Px, and M-Sex (M = transition metal) exhibit optimal catalytic activity for HER [163,164,165,166]. In particular, molybdenum carbide (Mo2C) exhibits brilliant HER activity in both alkaline and acidic systems because the electronic d-band model of molybdenum is similar to that of platinum [167,168,169]. Therefore, it has exhibited effective catalytic activity by combining the secondary phase/structure of Co and Mo-base as catalysts for the overall splitting of water in Ref. [170].
Melamine, which is commonly used as a precursor of nitrogen and carbon, has been used in this study. It was synthesized by mixing Mo and Co salt with facile heat treatment (Figure 16a). Figure 16b,c exhibits the tubular shape of Co/β-Mo2C@CNTs, in which nanotubes were encapsulated with Mo2C and Co. The HRTEM image indicates that the lattice distance of 0.23 and 0.20 nm corresponds to the plane (101) of β-Mo2C NPs and the plane (111) of Co NPs, respectively.
In the XRD patterns of Co/β-Mo2C@N-CNTs, Co, and β-Mo2C peaks were mainly observed at 2θ = 34.3°, 37.7°, 51.9°, 61.5°, 69.1°, 72.4°, and 74.5° well-matched with β-Mo2C (JCPDS No. 35-0787), and peaks at around 2θ = 44.2°, 51.7°, and 75.8° well-matched with Co (JCPDS No. 15-0787), respectively (Figure 17).
Further, in this study, the efficiencies of HER and OER were measured using a standard three-electrode device. Table 2 presents the performance of each catalyst, such as overpotential, Tafel plot, current density, and double-layer capacitance (Cdl). The performance of the three catalysts used in the HER activity was significantly inferior to that of Pt/C [170]. However, the heterostructure that used Co/β-Mo2C@N-CNTs as a catalyst displayed higher activity than others. The Co/β-Mo2C@N-CNTs still exhibited the highest current density when the geometric current density was further normalized to the corresponding Cdl value, thus suggesting that the hybrid is intrinsically most involved. Therefore, the HER performance of Co/β-Mo2C@N-CNTs compares favorably with the actions of many recorded Co- and Mo2C-based representative electrocatalysts. These results indicate that the OER activity is impaired by Co doping. Moreover, the overpotential of IrO2 as a reference catalyst was significantly higher than the Co/β-Mo2@N-CNTs. Both OER and HER activity were tested in 1 M KOH.

3. Redox-Mediated OER and ORR for LOBs with Soluble Redox Catalysts

3.1. Introduction

Although LOBs have attracted significant attention on account of high energy density, they still have certain limits such as uncertainty of electrolyte at high voltage, large overpotentials, and subsequent low cycling stability [72,73,171,172,173,174,175]. Researchers have focused on developing and modifying cathodes and electrocatalysts to promote the ORR and OER to address these problems. However, another problem—the deposition of byproducts such as Li2O2, Li2CO3, and LiOH on the catalysts—impedes the transfer of electrons, causing the overcrowding of electrolyte and oxygen mass transport [176,177,178,179,180,181,182,183,184]. As a solution of these issues, soluble redox catalysts, which enhance the ORR or OER and prevent deposition of byproducts to some degree, have been introduced recently [185,186,187,188,189,190,191,192]. In addition, Wang et al. have designed a new redox-flow lithium battery (RFLB) using redox mediators, which helps the current collector of the electrodes to be changed to separated energy storage tanks [193,194,195,196]. This idea has been successfully applied to address the abovementioned pressing problems with a non-aqueous LOB, resulting in a new device, the redox-flow LOB (RFLOB) [197]. Figure 18a presents the schematic of a couple of ORR and OER electrocatalysts and other components. In fact, the creation and decomposition of byproducts happen in the gas diffusion tank (GDT), spatially isolated from the cathode of the cell, which seamlessly eliminates pore-clogging problems and surface passivation. Furthermore, compared to the traditional enclosed battery, such a decoupled structure as an independent power source and energy storage gives an RFLOB great advantages such as optimization and versatility of operation of redox flow batteries [193,194,195]. In this study, tris{4-[2-(2-methoxyethoxy)ethoxy]phenyl}amine (TMPPA) and 2,5-di-tert-butyl-p-benzoquinone (DTBBQ) are used as OER and ORR catalysts, respectively, to address issues like corrosion, large overpotential owing to oxidation of Li2O2, and instability after long-term cycling (Figure 18b).

3.2. Application of TMPPA and DTBBQ as Redox Mediators to RFLOB

CV analyses were implemented to confirm the redox peak of TMPPA and DTBBQ (Figure 19a). The reversible redox peaks of DTBBQ and TMPPA are depicted at approximately 2.63 and 3.63 V (vs. Li/Li+), respectively, crossing the Li2O2 equilibrium potential at 2.96 V (vs. Li/Li+). Thermodynamically, this indicates that DTBBQ•− may be reduced to O2 and act as an ORR mediator that enhances Li2O2 production during the discharge step, whereas TMPPA•+ oxidizes Li2O2 and works as an OER transmitter substance that promotes Li2O2 decomposition during the charge step. A schematic displaying the redox-targeting reactions of TMPPA and DTBBQ with Li2O2 (Figure 19b) demonstrates the working mechanism of the RFLOB. The discharging potential of the RFLOB is determined by the discharging process, which includes the process through which the reduction of molecules in the catholyte flows into the cathodic section in the cell. After the catholyte flows back to the GDT, the generated DTBBQ•− reduces O2, where the gas is supplied, and produces Li2O2. The reaction of DTBBQ•− with the oxygen molecules in this procedure includes only charge transfer at the tank’s interface between the liquid and gas. The two reactions of discharging process are expressed as follows:
Electrochemical reaction in cell: DTBBQ + e → DTBBQ
Chemical reaction in GDT: 2DTBBQ + 2Li+ + O2 → DTBBQ + Li2O2
Similarly, the charging potential of the cell is determined by the charging process, which implies the oxidation process of both the molecules in the cell. Thereafter, the generated TMPPA•+ oxidizes Li2O2, producing O2 back into the GDT when the catholyte circulates. The detailed reactions of the charging step are explained as follows:
Electrochemical reaction in cell: TMPPA − e → TMPPA•+
Chemical reaction in GDT: 2TMPPA•+ + Li2O2 → 2TMPPA + 2Li+ + O2
Further, as the reaction between redox mediators cannot occur during the charge and discharge steps, the mediators maintain stability with each other.
The OER and ORR reactions of the TMPPA and DTBBQ were analyzed using rotating disc electrode spectrometry (Figure 20a,b, respectively). In the case of the OER performance of TMPPA, despite the presence of Li2O2 in the electrolyte, no current was observed in the OER potential range. However, a steady-stage current of at least 0.26 mA cm−2 emerged at a potential above 3.70 V vs. Li/Li+, equivalent to the TMPPA oxidization, after 2 mM TMPPA was transferred to the above-mentioned electrolyte. This current density is considerably higher than the value of TMPPA alone (~0.16 mA cm−2), suggesting that the reaction of TMPPA•+ with Li2O2 contributes to the increased current. Furthermore, the reaction is resistant to an N2- and O2-saturated solution, suggesting that the TMPPA may be impervious to O2 during the measurement process. As illustrated in Figure 20b, DTBBQ’s ORR efficiency in an O2-saturated electrolyte is significantly greater than that in an N2-saturated electrolyte, with the assumption that DTBBQ•− can effectively minimize oxygen molecules. This phenomenon is similar to the catalytic effect of TMPPA.
To scrutinize the stability of the redox mediators with long-term cycling, galvanostatic tests were conducted by employing TMPPA and DTBBQ as liquid cathodes in half-cells (Figure 20c,d, respectively). The two redox mediators exhibit outstanding electrochemical stabilities under Ar atmosphere. In particular, it is confirmed that the polymerization of molecules is effectively hindered and the development of the TMPPA radical cation and dication is also significantly stabilized because of the alkoxyl group on TMPPA’s phenyl rings. The capacities of TMPPA and DTBBQ were maintained at 87.4% and 89.4%, respectively, after 100 cycles.
In this study, new redox catalysts TMPPA and DTBBQ have been introduced as a solution of pore-clogging issues and the passivation of the cathode surface. This novel device provides several benefits for large and high-density energy storage in terms of flexibility and scalability. However, TMPPA and DTBBQ need further improvement to be powerful enough for the RFLOB. For example, the degradation of mediators, particularly DTBBQ, after prolonged cycling induces a change in the voltage profiles, posing the main hurdle for pragmatic use. Moreover, the redox potentials of Li2O2 are slightly different from those of the redox mediators, which could cause overpotential losses during the discharge and charge processes. Therefore, additional research is needed to address the side reactions of the redox mediators in an O2 atmosphere and the control redox potentials more accurately to alleviate voltage hysteresis.

4. Application of Hydrogen Production via HER and UOR Activities

4.1. Introduction

Among the numerous recently developed energy-saving electrolysis methods, urea electrolysis is extraordinarily interesting, allowing for the treatment of urea-rich sewage by reducing the urea in it to produce a clean, eco-friendly mass of H2 energy [76,199]. In particular, the use of the UOR in ordinary water splitting potentially replaces the OER with a significant thermodynamic potential change from 1.23 V vs. RHE to a mere 0.37 V (Scheme 12). Urea is a suitable and promising storage medium for H2 and CO2 owing to it being highly dense, relatively nontoxic, inexpensive, robust, safe, and easy to carry and store. Nevertheless, the key problem is the existence of UOR, which experiences sluggish kinetics owing to a six-electron transfer cycle composed of gas evolution phases; thus, highly effective and inexpensive electrocatalysts are required to implement this outstanding model of energy conversion [200].

4.2. The Application of Electrocatalyst via Synthesis of Facile Melt-Infiltration

Figure 21 illustrates a simple Ni-Pd alloy NPs preparation strategy integrated into ordered mesoporous carbon (OMC) support with optimized nanostructures through necessary steps. First, Ni and Pd salts were melt-infiltrated on the OMC, and then Ni-Pd bimetallic salts were heat-treated under H2 gas flow.
Consequently, the Ni-Pd salt is formed from the tiny Ni-Pd alloy NPs that are well dispersed on the OMC (Table 3). Further, this study illustrates the metal content, surface area, pore volume, and pore diameter of NiPd/OMC, Ni/OMC, Pd/OMC, and bare OMC.
The operating electrode was constructed by drop-coating the slurry mixture for electrochemical measurements. The arrangement of the electrochemical cells and the method of potential conversion from Ag/AgCl to RHE are the same as described elsewhere (Equation (1)):
ERHE = EAg/AgCl + 0.059 · pH + E°Ag/AgCl
The CV and LSV measurements were performed with 0.33 M urea in N2 saturated 1 M KOH at the scanning frequencies 10 and 5 mV s−1, respectively.

4.3. Results

XRD analysis was used to explore the crystalline structure of the NiPd/OMC, Pd/OMC, and Ni/OMC electrocatalysts (Figure 22a). The NiPd/OMC XRD pattern reflects that the two major characteristic peaks of the bimetallic NiPd alloy’s fcc crystalline structures were clearly displayed and assigned (JCPDS no 01-072-2515). Furthermore, the scale of the bimetallic NiPd alloy NPs was measured to be as small as 1.6 nm using the Scherrer method. TEM and HRTEM produced NiPd alloy NPs found in OMC scale and distribution (Figure 22b,c). Figure 22b reveals the effective loading of flat, widely dispersed, and very low bimetallic NiPd NPs onto the OMC supply. The parallel lattice fringe corresponds to the (111) plane of large crystalline LiPid NPs, with a calculated spacing gap of 0.22 nm, as presented in Figure 22c.
In a tri-electrode system, all the experiments were performed to verify the electrocatalytic behavior of catalytic activity for CV, LSV, and CA in a 1 M KOH solution. The electrocatalytic activities for UOR of electrocatalysts like NiPd/OMC, Ni/OMC, Pd/OMC, and bare OMC were detected and compared with the IrO/C catalyst standard in the existence of 0.33 M urea in N2-saturated 1 M KOH solution at a scanning frequency of 5 mV s−1 (Figure 23a). In addition, their Tafel plots have often compared the UOR kinetics with the prepared electrocatalysts (Figure 23b). Therefore, the Tafel slope displayed the lowest NiPd/OMC of 31 mV dec−1 and Ni/OMC, Pd/OMC, and OMC of 36, 59, and 64 mV dec−1, respectively, in the UOR and little lower than the IrO/C electrocatalyst. This result suggests that a smaller slope value of the Tafel NiPd/OMC electrocatalyst will result in better and faster catalytic kinetics on UOR. The essential factor in the catalytic activity is durability (Figure 23c). The NiPd/OMC electrocatalytic UOR performance displays an overpotential of 1.346 V vs. RHE required to achieve the current density. These remarkably strong catalytic activities, such as high current intensity, good conductivity, low overpotential, low Tafel performance, and good longevity of NiPd/OMC electrocatalysts, may be attributed to the following significant factors:
  • The existence of Ni3+ cationic active material at NiOOH enables effortless electron jumping mechanism and can also induce electrocatalytic reaction transmission.
  • OMC has very large surface area, pore volume, and pore diameters that may indicate higher UOR electroactive interfacial sites and ion/e transfer speeds.
In this review, the as-prepared electrocatalysts were also explored for observations of the HER in the N2-saturated 1 M KOH aqueous solution with and without 0.33 M urea solution. As depicted in Figure 23d,e, the values of NiPd/OMC material electrocatalyst with respect to overpotential (−0.117 V vs. RHE) and Tafel slope suggests the extremely efficient and rapid HER kinetics. The Tafel slope value of 37 mV dec−1 for the NiPd/OMC electrocatalyst is based on familiar mechanisms for the electrocatalytic HER operation and suggests a Volmer–Heyrovsky mechanism [199]. The Tafel slopes of the electrocatalysts depicted in Figure 23e derived from the LSV polarization curves are estimated to be 29, 37, 65, 76, and 210 mV dec−1 for Pt@C, NiPd/OMC, Ni/OMC, Pd/OMC, and bare OMC, respectively. In fact, in the LSV polarization curves, a large decay of its HER behavior was not detected in between 5000 cycles, even after the 5000 period of CV operation (Figure 23f). Apart from the extremely good HER behavior in the N2-saturated 1 M KOH electrolyte with 0.33 M urea solution, the NiPd/OMC electrocatalyst was operated in the empty N2-saturated 1 M KOH electrolyte. Fortunately, no significant shifts in the intensity of cathodic current are seen in Figure 23f, indicating that the NiPd/OMC electrocatalyst has superior catalytic efficiency in both general water and urea electrolysis applications.

5. Conclusions

Herein we have reviewed effective catalysts in terms of their overall reaction to water splitting. In other words, the advantages that dictate the application of these materials as both cathodes and anodes indicate the direction in which hybrid materials are developed. The addition of defect sites, heterostructures, and heteroatom doping, which can be applied to such hybrid materials, indicates their potential for water splitting. Also, transition metal pnictides and chalcogenides using these strategies have displayed positive potentials toward HER and OER because of high conductivity and activity. In HER, heterostructure like Ni and Co hydroxides with 2D nanostructure like MoS2 can show outstanding performance through efficient electron/mass transfer and optimized catalytic sites. In OER, transition metal pnictides and chalcogenides can display excellent catalytic activity with thin oxide layer on the catalysts surface. But there are several things to overcome despite much studies and efforts. For example, it is significant to resist corrosion in electrolytes under high anodic potential without any phase transformation in OER, which is a problem directly related to the durability of the OER catalysts. In HER, even though electrocatalysts-based platinum or platinum alloys exhibit superior activity to Pt/C in alkaline solution, there was no noticeable decrease in Pt loading.
For a fundamental solution of these issues, understanding the mechanism of the reactions must be involved in order to perform synthesis of perfect electrocatalyst toward HER and OER. Catalysts development of in situ methods have many advantages in terms of easy understanding of the reaction mechanism and high stability. Substrates like metal foam such as Ni foam and NiCrAl foam or carbon cloth (CC) can be example of this approaches. Consequently, more methods which display enhanced activity for overall water splitting will be developed to synthesis heterostructure and transition metal complex with doping heteroatoms or defect sites for high conductivity on the substrate. Also, single atom nanocatalysts on substrates can suggest direction of lowering Pt loading toward HER and for minimizing phase transformation in OER, mixing the nanocompounds with other elements or depositing buffer layer can be solution to improve stability of catalysts on substrates. As a recent trend, beyond the limits of only electrochemical water splitting, photoelectrolysis can be considered as one of the environmental methods that exploits solar energy. This strategy will contribute to future energy field because it can present the synergistic effect of water splitting by smoothly moving electrons using solar energy which is unlimited energy. Notably, researches with theoretical calculations or computational results should be performed by processing fabrication electrocatalysts, which can accelerate provision of intrinsic knowledge of HER and OER. Furthermore, UOR, the method of generating electricity by decomposing materials such as urea, can also be implemented to solve environmental problems like air pollution and greenhouse effect. Our previous catalysts, e.g., NiPd/CMK-3 nanocatalyst for the electrocatalyst of UOR and HER activities, also exhibited good catalytic activities. It can be applied to HER because of the efficiency of Mo, Ni, and Co, but we can expect to apply it to different electrochemical reactions like ORR and OER in our next work. We also expect that our approach can be adapted in the near future for other catalytic syntheses.

Author Contributions

K.H.P. provided academic direction, conceptualization, validation, funding acquisition, writing—reviewing and editing. S.J. and K.M. contributed equally on that work. Specially, S.J. collected materials and wrote the introduction and conclusion part. K.M. contributed in results and discussion part and checked English style as well as grammatical errors throughout the manuscript. Y.P. and S.P. contributed to the materials and methods and results and discussion part. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by 2-Year Research Grant of Pusan National University.

Acknowledgments

This work was supported by a 2-Year Research Grant of Pusan National University.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

NPs nanoparticles
OER oxygen evolution reaction
HER hydrogen evolution reaction
ORR oxygen reduction reaction
TMP transition metal phosphide
LOB lithium−oxygen batteries
CVD chemical vapor deposition
ADF-STEM atomic-resolution annular dark field-scanning transmission electron microscope
HAADF high angle annular dark field
EDX energy dispersive X-ray
PXRD powder X-ray diffraction
Pa3 cubic pyrite-type phase
BET Brunauer-Emmett-Teller
PBS phosphate-buffered saline
NF nickel foam
RHE reversible hydrogen electrode
EIS electrochemical impedance spectroscopy
CV cyclic voltammetry
CA chronoamperometry
LSV linear sweep voltammetry
TMD transition metal dichalcogenide
LDH layered double hydroxide
DI deionized
EDA ethylene diamine
CC carbon cloth
HRTEM high-resolution transmission electron microscope
Pi inorganic phosphate
XRD X-ray diffraction
Cdl double-layer capacitance
RFLB rechargeable redox flow lithium battery
RFLOB lithium-oxygen redox flow battery
GDT gas diffusion tank
TMPPA tris{4-[2-(2-methoxyethoxy)ethoxy]phenyl}amine
DTBBQ 2,5-di-tert-butyl-p-benzoquinone
RDE rotating disk electrode
UOR urea electro-oxidation reaction
OMC ordered mesoporous carbon
BJH Barrett-Joyner-Haldenda
fcc face centered cubic

References

  1. Ghosh, T.; Mohammad, A.; Mobin, S.M. Hybrid Cobalt Doped-Cerium Oxide as a Multifunctional Nanocatalyst for Various Organic Transformations. ACS Sustain. Chem. Eng. 2019, 7, 13746–13763. [Google Scholar] [CrossRef]
  2. Díaz, U.; Brunel, D.; Corma, A. Catalysis using Multifunctional Organosiliceous Hybrid Materials. Chem. Soc. Rev. 2013, 42, 4083–4097. [Google Scholar] [CrossRef] [PubMed]
  3. Loy, D.A.; Shea, K.J. Bridged Polysilsesquioxanes. Highly Porous Hybrid Organic-Inorganic Materials. Chem. Rev. 1995, 95, 1431–1442. [Google Scholar] [CrossRef]
  4. Di Carlo, G.; Lualdi, M.; Venezia, A.M.; Boutonnet, M.; Sanchez-Dominguez, M. Design of Cobalt Nanoparticles with Tailored Structural and Morphological Properties via O/W and W/O Microemulsions and Their Deposition onto Silica. Catalysts 2015, 5, 442–459. [Google Scholar] [CrossRef]
  5. Jang, S.; Hira, S.A.; Annas, D.; Song, S.; Yusuf, M.; Park, J.C.; Park, S.; Park, K.H. Recent Novel Hybrid Pd-Fe3O4 Nanoparticles as Catalysts for Various C-C Coupling Reactions. Processes 2019, 7, 422–452. [Google Scholar] [CrossRef] [Green Version]
  6. Tan, L.; Wu, X.; Chen, D.; Liu, H.; Meng, X.; Tang, F. Confining Alloy or Core-Shell Au-Pd Bimetallic Nanocrystals in Silica Nanorattles for Enhanced Catalytic Performance. J. Mater. Chem. A 2013, 1, 10382–10388. [Google Scholar] [CrossRef]
  7. Moussa, S.; Siamaki, A.R.; Gupton, B.F.; El-Shall, M.S. Pd-Partially Reduced Graphene Oxide Catalysts (Pd/PRGO): Laser Synthesis of pd Nanoparticles Supported on PRGO Nanosheets for Carboncarbon Cross Coupling Reactions. ACS Catal. 2012, 2, 145–154. [Google Scholar] [CrossRef]
  8. Nasrollahzadeh, M.; Mohammad Sajadi, S.; Rostami-Vartooni, A.; Khalaj, M. Green Synthesis of Pd/Fe3O4 Nanoparticles Using Euphorbia Condylocarpa M. Bieb Root Extract and Their Catalytic Applications as Magnetically Recoverable and Stable Recyclable Catalysts for the Phosphine-Free Sonogashira and Suzuki Coupling Reactions. J. Mol. Catal. A Chem. 2015, 396, 31–39. [Google Scholar] [CrossRef]
  9. Siamaki, A.R.; Khder, A.E.R.S.; Abdelsayed, V.; El-Shall, M.S.; Gupton, B.F. Microwave-Assisted Synthesis of Palladium Nanoparticles Supported on Graphene: A Highly Active and Recyclable Catalyst for Carbon-Carbon Cross-Coupling Reactions. J. Catal. 2011, 279, 1–11. [Google Scholar] [CrossRef]
  10. Truong, Q.D.; Le, T.S.; Ling, Y.C. Pt Deposited TiO2 Catalyst Fabricated by Thermal Decomposition of Titanium Complex for Solar Hydrogen Production. Solid State Sci. 2014, 38, 18–24. [Google Scholar] [CrossRef]
  11. Woo, H.; Lee, K.; Park, J.C.; Park, K.H. Facile Synthesis of Pd/Fe3O4/Charcoal Bifunctional Catalysts with High Metal Loading for High Product Yields in Suzuki-Miyaura Coupling Reactions. New J. Chem. 2014, 38, 5626–5632. [Google Scholar] [CrossRef]
  12. Tan, Q.; Du, C.; Sun, Y.; Yin, G.; Gao, Y. Pd-around-CeO2-x Hybrid Nanostructure Catalyst: Three-Phase-Transfer Synthesis, Electrocatalytic Properties and Dual Promoting Mechanism. J. Mater. Chem. A 2014, 2, 1429–1435. [Google Scholar] [CrossRef]
  13. Hosseini, S.G.; Abazari, R.; Ghavi, A. Pure CuCr2O4 Nanoparticles: Synthesis, Characterization and Their Morphological and Size Effects on the Catalytic Thermal Decomposition of Ammonium Perchlorate. Solid State Sci. 2014, 37, 72–79. [Google Scholar] [CrossRef]
  14. Veisi, H.; Sedrpoushan, A.; Maleki, B.; Hekmati, M.; Heidari, M.; Hemmati, S. Palladium Immobilized on Amidoxime-Functionalized Magnetic Fe3O4 Nanoparticles: A Highly Stable and Efficient Magnetically Recoverable Nanocatalyst for Sonogashira Coupling Reaction. Appl. Organomet. Chem. 2015, 29, 834–839. [Google Scholar] [CrossRef]
  15. Wang, D.; Li, Y. One-Pot Protocol for Au-Based Hybrid Magnetic Nanostructures via a Noble-Metal-Induced Reduction Process. J. Am. Chem. Soc. 2010, 132, 6280–6281. [Google Scholar] [CrossRef]
  16. Jin, X.; Dang, L.; Lohrman, J.; Subramaniam, B.; Ren, S.; Chaudhari, R.V. Lattice-Matched Bimetallic CuPd-Graphene Nanocatalysts for Facile Conversion of Biomass-Derived Polyols to Chemicals. ACS Nano 2013, 7, 1309–1316. [Google Scholar] [CrossRef]
  17. Dong, B.; Miller, D.L.; Li, C.Y. Polymer Single Crystal as Magnetically Recoverable Support for Nanocatalysts. J. Phys. Chem. Lett. 2012, 3, 1346–1350. [Google Scholar] [CrossRef]
  18. Omar, S.; Abu-Reziq, R. Palladium Nanoparticles Supported on Magnetic Organic-Silica Hybrid Nanoparticles. J. Phys. Chem. C 2014, 118, 30045–30056. [Google Scholar] [CrossRef]
  19. Hu, H.; Xin, J.H.; Hu, H.; Wang, X.; Miao, D.; Liu, Y. Synthesis and Stabilization of Metal Nanocatalysts for Reduction Reactions—A Review. J. Mater. Chem. A 2015, 3, 11157–11182. [Google Scholar] [CrossRef]
  20. Askari, S.; Halladj, R.; Azarhoosh, M.J. Modeling and Optimization of Catalytic Performance of SAPO-34 Nanocatalysts Synthesized Sonochemically using a New Hybrid of Non-Dominated Sorting Genetic Algorithm-II based Artificial Neural Networks (NSGA-II-ANNs). RSC Adv. 2015, 5, 52788–52800. [Google Scholar] [CrossRef]
  21. Zheng, Y.; Chen, C.; Zhan, Y.; Lin, X.; Zheng, Q.; Wei, K.; Zhu, J. Photocatalytic Activity of Ag/ZnO Heterostructure Nanocatalyst: Correlation between Structure and Property. J. Phys. Chem. C 2008, 112, 10773–10777. [Google Scholar] [CrossRef]
  22. Sharma, R.K.; Sharma, S.; Dutta, S.; Zboril, R.; Gawande, M.B. Silica-Nanosphere-Based Organic-Inorganic Hybrid Nanomaterials: Synthesis, Functionalization and Applications in Catalysis. Green Chem. 2015, 17, 3207–3230. [Google Scholar] [CrossRef]
  23. Song, H. Metal Hybrid Nanoparticles for Catalytic Organic and Photochemical Transformations. Acc. Chem. Res. 2015, 48, 491–499. [Google Scholar] [CrossRef] [PubMed]
  24. Guo, L.T.; Cai, Y.Y.; Ge, J.M.; Zhang, Y.N.; Gong, L.H.; Li, X.H.; Wang, K.X.; Ren, Q.Z.; Su, J.; Chen, J.S. Multifunctional Au-Co@CN Nanocatalyst for Highly Efficient Hydrolysis of Ammonia Borane. ACS Catal. 2015, 5, 388–392. [Google Scholar] [CrossRef]
  25. Kim, S.M.; Lee, S.J.; Kim, S.H.; Kwon, S.; Yee, K.J.; Song, H.; Somorjai, G.A.; Park, J.Y. Hot Carrier-Driven Catalytic Reactions on Pt-CdSe-Pt Nanodumbbells and Pt/GaN under Light Irradiation. Nano Lett. 2013, 13, 1352–1358. [Google Scholar] [CrossRef]
  26. Tollefson, J. Fuel of the Future? Automot. Eng. Int. 1977, 85, 48–52. [Google Scholar]
  27. Chu, S.; Majumdar, A. Opportunities and Challenges for a Sustainable Energy Future. Nature 2012, 488, 294–303. [Google Scholar] [CrossRef]
  28. Seh, Z.W.; Kibsgaard, J.; Dickens, C.F.; Chorkendorff, I.; Nørskov, J.K.; Jaramillo, T.F. Combining Theory and Experiment in Electrocatalysis: Insights into Materials Design. Science 2017, 355, 146–157. [Google Scholar] [CrossRef] [Green Version]
  29. Stamenkovic, V.R.; Strmcnik, D.; Lopes, P.P.; Markovic, N.M. Energy and Fuels from Electrochemical Interfaces. Nat. Mater. 2016, 16, 57–69. [Google Scholar] [CrossRef]
  30. Greeley, J.; Jaramillo, T.F.; Bonde, J.; Chorkendorff, I.; Nørskov, J.K. Computational High-Throughput Screening of Electrocatalytic Materials for Hydrogen Evolution. Nat. Mater. 2006, 5, 909–913. [Google Scholar] [CrossRef]
  31. Gao, M.R.; Xu, Y.F.; Jiang, J.; Yu, S.H. Nanostructured Metal Chalcogenides: Synthesis, Modification, and Applications in Energy Conversion and Storage Devices. Chem. Soc. Rev. 2013, 42, 2986–3017. [Google Scholar] [CrossRef] [PubMed]
  32. Gao, M.R.; Liang, J.X.; Zheng, Y.R.; Xu, Y.F.; Jiang, J.; Gao, Q.; Li, J.; Yu, S.H. An Efficient Molybdenum Disulfide/Cobalt Diselenide Hybrid Catalyst for Electrochemical Hydrogen Generation. Nat. Commun. 2015, 6, 1–7. [Google Scholar] [CrossRef] [PubMed]
  33. Walter, M.G.; Warren, E.L.; McKone, J.R.; Boettcher, S.W.; Mi, Q.; Santori, E.A.; Lewis, N.S. Solar Water Splitting Cells. Chem. Rev. 2010, 110, 6446–6473. [Google Scholar] [CrossRef] [PubMed]
  34. Cook, T.R.; Dogutan, D.K.; Reece, S.Y.; Surendranath, Y.; Teets, T.S.; Nocera, D.G. Solar Energy Supply and Storage for the Legacy and Nonlegacy Worlds. Chem. Rev. 2010, 110, 6474–6502. [Google Scholar] [CrossRef] [PubMed]
  35. Chen, W.F.; Muckerman, J.T.; Fujita, E. Recent Developments in Transition metal Carbides and Nitrides as Hydrogen Evolution Electrocatalysts. Chem. Commun. 2013, 49, 8896–8909. [Google Scholar] [CrossRef]
  36. Matsumoto, Y.; Sato, E. Electrocatalytic Properties of Transition Metal Oxides for Oxygen Evolution Reaction. Mater. Chem. Phys. 1986, 14, 397–426. [Google Scholar] [CrossRef]
  37. Jaramillo, T.F.; Ivanovskaya, A.; McFarland, E.W. High-Throughput Screening System for Catalytic Hydrogen-Producing Materials. J. Comb. Chem. 2002, 4, 17–22. [Google Scholar] [CrossRef]
  38. Chen, G.; Delafuente, D.A.; Sarangapani, S.; Mallouk, T.E. Combinatorial Discovery of Bifunctional Oxygen Reduction - Water Oxidation Electrocatalysts for Regenerative Fuel Cells. Catal. Today 2001, 67, 341–355. [Google Scholar] [CrossRef] [Green Version]
  39. Gerken, J.B.; Chen, J.Y.C.; Massé, R.C.; Powell, A.B.; Stahl, S.S. Development of an O2-Sensitive Fluorescence-Quenching Assay for the Combinatorial Discovery of Electrocatalysts for Water Oxidation. Angew. Chem. Int. Ed. 2012, 51, 6676–6680. [Google Scholar] [CrossRef]
  40. Liu, X.; Shen, Y.; Yang, R.; Zou, S.; Ji, X.; Shi, L.; Zhang, Y.; Liu, D.; Xiao, L.; Zheng, X.; et al. Inkjet Printing Assisted Synthesis of Multicomponent Mesoporous Metal Oxides for Ultrafast Catalyst Exploration. Nano Lett. 2012, 12, 5733–5739. [Google Scholar] [CrossRef]
  41. Gregoire, J.M.; Xiang, C.; Mitrovic, S.; Liu, X.; Marcin, M.; Cornell, E.W.; Fan, J.; Jin, J. Combined Catalysis and Optical Screening for High Throughput Discovery of Solar Fuels Catalysts. J. Electrochem. Soc. 2013, 160, F337–F342. [Google Scholar] [CrossRef]
  42. Seley, D.; Ayers, K.; Parkinson, B.A. Combinatorial Search for Improved Metal Oxide Oxygen Evolution Electrocatalysts in Acidic Electrolytes. ACS Comb. Sci. 2013, 15, 82–89. [Google Scholar] [CrossRef] [PubMed]
  43. Berglund, S.P.; Lee, H.C.; Núñez, P.D.; Bard, A.J.; Mullins, C.B. Screening of Transition and Post-Transition Metals to Incorporate into Copper Oxide and Copper Bismuth Oxide for Photoelectrochemical Hydrogen Evolution. Phys. Chem. Chem. Phys. 2013, 15, 4554–4565. [Google Scholar] [CrossRef] [PubMed]
  44. Lu, Q.; Wang, A.L.; Cheng, H.; Gong, Y.; Yun, Q.; Yang, N.; Li, B.; Chen, B.; Zhang, Q.; Zong, Y.; et al. Synthesis of Hierarchical 4H/fcc Ru Nanotubes for Highly Efficient Hydrogen Evolution in Alkaline Media. Small 2018, 14, 1–5. [Google Scholar] [CrossRef]
  45. Tian, H.; Cui, X.; Zeng, L.; Su, L.; Song, Y.; Shi, J. Oxygen Vacancy-Assisted Hydrogen Evolution Reaction of the Pt/WO3 Electrocatalyst. J. Mater. Chem. A 2019, 7, 6285–6293. [Google Scholar] [CrossRef]
  46. Zheng, Y.; Jiao, Y.; Zhu, Y.; Li, L.H.; Han, Y.; Chen, Y.; Jaroniec, M.; Qiao, S.Z. High Electrocatalytic Hydrogen Evolution Activity of an Anomalous Ruthenium Catalyst. J. Am. Chem. Soc. 2016, 138, 16174–16181. [Google Scholar] [CrossRef]
  47. Yang, Y.; Wang, S.; Zhang, J.; Li, H.; Tang, Z.; Wang, X. Nanosheet-Assembled MoSe2 and S-doped MoSe2-x Nanostructures for Superior Lithium Storage Properties and Hydrogen Evolution Reactions. Inorg. Chem. Front. 2015, 2, 931–937. [Google Scholar] [CrossRef]
  48. Peng, S.; Li, L.; Tan, H.; Cai, R.; Shi, W.; Li, C.; Mhaisalkar, S.G.; Srinivasan, M.; Ramakrishna, S.; Yan, Q. MS2 (M = Co and Ni) Hollow Spheres with Tunable Interiors for High-Performance Supercapacitors and Photovoltaics. Adv. Funct. Mater. 2014, 24, 2155–2162. [Google Scholar] [CrossRef]
  49. Lu, X.F.; Yu, L.; Lou, X.W. Highly Crystalline Ni-Doped FeP/Carbon Hollow Nanorods as all-pH Efficient and Durable Hydrogen Evolving Electrocatalysts. Sci. Adv. 2019, 5, 1–10. [Google Scholar] [CrossRef] [Green Version]
  50. Hao, W.; Wu, R.; Zhang, R.; Ha, Y.; Chen, Z.; Wang, L.; Yang, Y.; Ma, X.; Sun, D.; Fang, F.; et al. Electroless Plating of Highly Efficient Bifunctional Boride-Based Electrodes toward Practical Overall Water Splitting. Adv. Energy Mater. 2018, 8, 1–7. [Google Scholar] [CrossRef]
  51. Ha, Y.; Shi, L.; Chen, Z.; Wu, R. Phase-Transited Lysozyme-Driven Formation of Self-Supported Co3O4@C Nanomeshes for Overall Water Splitting. Adv. Sci. 2019, 6, 1–8. [Google Scholar]
  52. Chen, Z.; Ha, Y.; Liu, Y.; Wang, H.; Yang, H.; Xu, H.; Li, Y.; Wu, R. In Situ Formation of Cobalt Nitrides/Graphitic Carbon Composites as Efficient Bifunctional Electrocatalysts for Overall Water Splitting. ACS Appl. Mater. Interfaces 2018, 10, 7134–7144. [Google Scholar] [CrossRef] [PubMed]
  53. Liu, P.; Rodriguez, J.A. Catalysts for Hydrogen Evolution from the [NiFe] Hydrogenase to the Ni2P(001) Surface: The Importance of Ensemble Effect. J. Am. Chem. Soc. 2005, 127, 14871–14878. [Google Scholar] [CrossRef]
  54. Görlin, M.; Chernev, P.; De Araújo, J.F.; Reier, T.; Dresp, S.; Paul, B.; Krähnert, R.; Dau, H.; Strasser, P. Oxygen Evolution Reaction Dynamics, Faradaic Charge Efficiency, and the Active Metal Redox States of Ni-Fe Oxide Water Splitting Electrocatalysts. J. Am. Chem. Soc. 2016, 138, 5603–5614. [Google Scholar] [CrossRef] [PubMed]
  55. Zhu, Y.P.; Liu, Y.P.; Ren, T.Z.; Yuan, Z.Y. Self-Supported Cobalt Phosphide Mesoporous Nanorod Arrays: A Flexible and Bifunctional Electrode for Highly Active Electrocatalytic Water Reduction and Oxidation. Adv. Funct. Mater. 2015, 25, 7337–7347. [Google Scholar] [CrossRef]
  56. Carenco, S.; Portehault, D.; Boissière, C.; Mézailles, N.; Sanchez, C. Nanoscaled Metal Borides and Phosphides: Recent Developments and Perspectives. Chem. Rev. 2013, 113, 7981–8065. [Google Scholar] [CrossRef]
  57. Shi, Y.; Zhang, B. Recent Advances in Transition Metal Phosphide Nanomaterials: Synthesis and Applications in Hydrogen Evolution Reaction. Chem. Soc. Rev. 2016, 45, 1529–1541. [Google Scholar] [CrossRef]
  58. Zhang, R.; Huang, J.; Chen, G.; Chen, W.; Song, C.; Li, C.; Ostrikov, K. In Situ Engineering Bi-Metallic Phospho-Nitride Bi-Functional Electrocatalysts for Overall Water Splitting. Appl. Catal. B Environ. 2019, 254, 414–423. [Google Scholar] [CrossRef]
  59. Wu, R.; Wang, D.P.; Zhou, K.; Srikanth, N.; Wei, J.; Chen, Z. Porous Cobalt Phosphide/Graphitic Carbon Polyhedral Hybrid Composites for Efficient Oxygen Evolution Reactions. J. Mater. Chem. A 2016, 4, 13742–13745. [Google Scholar] [CrossRef]
  60. Lin, Y.; Pan, Y.; Liu, S.; Sun, K.; Cheng, Y.; Liu, M.; Wang, Z.; Li, X.; Zhang, J. Construction of Multi-Dimensional Core/Shell Ni/NiCoP Nano-Heterojunction for Efficient Electrocatalytic Water Splitting. Appl. Catal. B Environ. 2019, 259, 118036–118039. [Google Scholar] [CrossRef]
  61. Li, J.S.; Kong, L.X.; Wu, Z.; Zhang, S.; Yang, X.Y.; Sha, J.Q.; Liu, G.D. Polydopamine-Assisted Construction of Cobalt Phosphide Encapsulated in N-Doped Carbon Porous Polyhedrons for Enhanced Overall Water Splitting. Carbon 2019, 145, 694–700. [Google Scholar] [CrossRef]
  62. Dutta, S.; Indra, A.; Han, H.S.; Song, T. 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]
  63. Hoa, V.H.; Tran, D.T.; Le, H.T.; Kim, N.H.; Lee, J.H. Hierarchically Porous Nickel–Cobalt Phosphide Nanoneedle Arrays Loaded Micro-Carbon Spheres as an Advanced Electrocatalyst for Overall Water Splitting Application. Appl. Catal. B Environ. 2019, 253, 235–245. [Google Scholar] [CrossRef]
  64. Xu, H.; Jia, H.; Fei, B.; Ha, Y.; Li, H.; Guo, Y.; Liu, M.; Wu, R. Charge Transfer Engineering via Multiple Heteroatom Doping in Dual Carbon-Coupled Cobalt Phosphides for Highly Efficient Overall Water Splitting. Appl. Catal. B Environ. 2020, 268, 118404–118412. [Google Scholar] [CrossRef]
  65. Lin, H.; Shi, Z.; He, S.; Yu, X.; Wang, S.; Gao, Q.; Tang, Y. Heteronanowires of MoC-Mo2C as Efficient Electrocatalysts for Hydrogen Evolution Reaction. Chem. Sci. 2016, 7, 3399–3405. [Google Scholar] [CrossRef] [Green Version]
  66. Wang, D.; Li, Q.; Han, C.; Xing, Z.; Yang, X. When NiO@Ni Meets WS2 Nanosheet Array: A Highly Efficient and Ultrastable Electrocatalyst for Overall Water Splitting. ACS Cent. Sci. 2018, 4, 112–119. [Google Scholar] [CrossRef] [Green Version]
  67. Rui, K.; Zhao, G.; Chen, Y.; Lin, Y.; Zhou, Q.; Chen, J.; Zhu, J.; Sun, W.; Huang, W.; Dou, S.X. Hybrid 2D Dual-Metal–Organic Frameworks for Enhanced Water Oxidation Catalysis. Adv. Funct. Mater. 2018, 28, 1–9. [Google Scholar] [CrossRef] [Green Version]
  68. Dou, S.; Wu, J.; Tao, L.; Shen, A.; Huo, J.; Wang, S. Carbon-Coated MoS2 Nanosheets as Highly Efficient Electrocatalysts for the Hydrogen Evolution Reaction. Nanotechnology 2016, 27, 1–6. [Google Scholar] [CrossRef]
  69. Chen, Y.; Zhou, Q.; Zhao, G.; Yu, Z.; Wang, X.; Dou, S.X.; Sun, W. Electrochemically Inert g-C3N4 Promotes Water Oxidation Catalysis. Adv. Funct. Mater. 2018, 28, 1–7. [Google Scholar]
  70. Zhao, G.; Rui, K.; Dou, S.X.; Sun, W. Heterostructures for Electrochemical Hydrogen Evolution Reaction: A Review. Adv. Funct. Mater. 2018, 28, 1–26. [Google Scholar] [CrossRef] [Green Version]
  71. Gupta, S.; Kellogg, W.; Xu, H.; Liu, X.; Cho, J.; Wu, G. Bifunctional Perovskite Oxide Catalysts for Oxygen Reduction and Evolution in Alkaline Media. Chem. Asian J. 2016, 11, 10–21. [Google Scholar] [CrossRef] [PubMed]
  72. Abraham, K.M.; Jiang, Z. ELECTROCHEMICAL SCIENCE AND TECHNOLOGY A Polymer Electrolyte-Based Rechargeable Lithium/Oxygen Battery. J. Electrochem. Soc. 1996, 143, 1–5. [Google Scholar] [CrossRef]
  73. Bruce, P.G.; Freunberger, S.A.; Hardwick, L.J.; Tarascon, J.M. Li-O2 and Li-S Batteries with High Energy Storage. Nat. Mater. 2012, 11, 19–29. [Google Scholar] [CrossRef] [PubMed]
  74. Jin, H.; Guo, C.; Liu, X.; Liu, J.; Vasileff, A.; Jiao, Y.; Zheng, Y.; Qiao, S.Z. Emerging Two-Dimensional Nanomaterials for Electrocatalysis. Chem. Rev. 2018, 118, 6337–6408. [Google Scholar] [CrossRef] [PubMed]
  75. Urbańczyk, E.; Sowa, M.; Simka, W. Urea Removal from Aqueous Solutions—A Review. J. Appl. Electrochem. 2016, 46, 1011–1029. [Google Scholar] [CrossRef] [Green Version]
  76. Xu, W.; Zhang, H.; Li, G.; Wu, Z. Nickel-Cobalt Bimetallic Anode Catalysts for Direct Urea Fuel Cell. Sci. Rep. 2014, 4, 1–6. [Google Scholar] [CrossRef]
  77. Zhu, B.; Liang, Z.; Zou, R. Designing Advanced Catalysts for Energy Conversion Based on Urea Oxidation Reaction. Small 2020, 16, 1906133. [Google Scholar] [CrossRef]
  78. Xie, J.; Yang, X.; Xie, Y. Defect Engineering in Two-Dimensional Electrocatalysts for Hydrogen Evolution. Nanoscale 2020, 12, 4238–4294. [Google Scholar] [CrossRef]
  79. Lee, J.; Kim, C.; Choi, K.S.; Seo, J.; Choi, Y.; Choi, W.; Kim, Y.M.; Jeong, H.Y.; Lee, J.H.; Kim, G.; et al. In-Situ Coalesced Vacancies on MoSe2 Mimicking Noble Metal: Unprecedented TAFEL Reaction in Hydrogen Evolution. Nano Energy 2019, 63, 103844–1038446. [Google Scholar] [CrossRef]
  80. Shu, H.; Zhou, D.; Li, F.; Cao, D.; Chen, X. Defect Engineering in MoSe2 for the Hydrogen Evolution Reaction: From Point Defects to Edges. ACS Appl. Mater. Interfaces 2017, 9, 42688–42698. [Google Scholar] [CrossRef]
  81. He, J.; Zou, Y.; Wang, S. Defect Engineering on Electrocatalysts for Gas-Evolving Reactions. Dalton Trans. 2019, 48, 15–20. [Google Scholar] [CrossRef] [PubMed]
  82. Zhang, H.; Zhang, J.; Li, Y.; Jiang, H.; Jiang, H.; Li, C. Continuous Oxygen Vacancy Engineering of the Co3O4 Layer for an Enhanced Alkaline Electrocatalytic Hydrogen Evolution Reaction. J. Mater. Chem. A 2019, 7, 13506–13510. [Google Scholar] [CrossRef]
  83. Liu, T.; Feng, Z.; Li, Q.; Yang, J.; Li, C.; Dupuis, M. Role of Oxygen Vacancies on Oxygen Evolution Reaction Activity: β-Ga2O3 as a Case Study. Chem. Mater. 2018, 30, 7714–7726. [Google Scholar] [CrossRef]
  84. Hirai, S.; Morita, K.; Yasuoka, K.; Shibuya, T.; Tojo, Y.; Kamihara, Y.; Miura, A.; Suzuki, H.; Ohno, T.; Matsuda, T.; et al. Oxygen Vacancy-Originated Highly Active Electrocatalysts for the Oxygen Evolution Reaction. J. Mater. Chem. A 2018, 6, 15102–15109. [Google Scholar] [CrossRef]
  85. Li, H.; Tsai, C.; Koh, A.L.; Cai, L.; Contryman, A.W.; Fragapane, A.H.; Zhao, J.; Han, H.S.; Manoharan, H.C.; Abild-Pedersen, F.; et al. Activating and Optimizing MoS2 Basal Planes for Hydrogen Evolution through the Formation of Strained Sulphur Vacancies. Nat. Mater. 2016, 15, 48–53. [Google Scholar] [CrossRef]
  86. Ye, G.; Gong, Y.; Lin, J.; Li, B.; He, Y.; Pantelides, S.T.; Zhou, W.; Vajtai, R.; Ajayan, P.M. Defects Engineered Monolayer MoS2 for Improved Hydrogen Evolution Reaction. Nano Lett. 2016, 16, 1097–1103. [Google Scholar] [CrossRef]
  87. Tsai, C.; Li, H.; Park, S.; Park, J.; Han, H.S.; Nørskov, J.K.; Zheng, X.; Abild-Pedersen, F. Electrochemical Generation of Sulfur Vacancies in the Basal Plane of MoS2 for Hydrogen Evolution. Nat. Commun. 2017, 8, 1–8. [Google Scholar] [CrossRef]
  88. Li, B.; Gong, Y.; Hu, Z.; Brunetto, G.; Yang, Y.; Ye, G.; Zhang, Z.; Lei, S.; Jin, Z.; Bianco, E.; et al. Solid–Vapor Reaction Growth of Transition-Metal Dichalcogenide Monolayers. Angew. Chem. Int. Ed. 2016, 55, 1–7. [Google Scholar]
  89. Shaw, J.C.; Zhou, H.; Chen, Y.; Weiss, N.O.; Liu, Y.; Huang, Y.; Duan, X. Chemical Vapor Deposition Growth of Monolayer MoSe2 Nanosheets. Nano Res. 2014, 7, 1–7. [Google Scholar] [CrossRef]
  90. Mahmood, J.; Li, F.; Jung, S.M.; Okyay, M.S.; Ahmad, I.; Kim, S.J.; Park, N.; Jeong, H.Y.; Baek, J.B. An Efficient and pH-Universal Ruthenium-Based Catalyst for the Hydrogen Evolution Reaction. Nat. Nanotechnol. 2017, 12, 441–446. [Google Scholar] [CrossRef]
  91. Wang, D.; Zhang, X.; Bao, S.; Zhang, Z.; Fei, H.; Wu, Z. Phase Engineering of a Multiphasic 1T/2H MoS2 Catalyst for Highly Efficient Hydrogen Evolution. J. Mater. Chem. A 2017, 5, 2681–2688. [Google Scholar] [CrossRef]
  92. Jiang, M.; Zhang, J.; Wu, M.; Jian, W.; Xue, H.; Ng, T.W.; Lee, C.S.; Xu, J. Synthesis of 1T-MoSe2 Ultrathin Nanosheets with an Expanded Interlayer Spacing of 1.17 nm for Efficient Hydrogen Evolution Reaction. J. Mater. Chem. A 2016, 4, 14949–14953. [Google Scholar] [CrossRef]
  93. Xu, C.; Peng, S.; Tan, C.; Ang, H.; Tan, H.; Zhang, H.; Yan, Q. Ultrathin S-Doped MoSe2 Nanosheets for Efficient Hydrogen Evolution. J. Mater. Chem. A 2014, 2, 5597–5601. [Google Scholar] [CrossRef] [Green Version]
  94. Shi, Y.; Zhou, Y.; Yang, D.R.; Xu, W.X.; Wang, C.; Wang, F.B.; Xu, J.J.; Xia, X.H.; Chen, H.Y. Energy Level Engineering of MoS2 by Transition-Metal Doping for Accelerating Hydrogen Evolution Reaction. J. Am. Chem. Soc. 2017, 139, 15479–15485. [Google Scholar] [CrossRef] [PubMed]
  95. Liu, B.; Zhang, L.; Xiong, W.; Ma, M. Cobalt-Nanocrystal-Assembled Hollow Nanoparticles for Electrocatalytic Hydrogen Generation from Neutral-pH Water. Angew. Chem. Int. Ed. 2016, 55, 6725–6729. [Google Scholar] [CrossRef]
  96. Xu, K.; Cheng, H.; Liu, L.; Lv, H.; Wu, X.; Wu, C.; Xie, Y. Promoting Active Species Generation by Electrochemical Activation in Alkaline Media for Efficient Electrocatalytic Oxygen Evolution in Neutral Media. Nano Lett. 2017, 17, 578–583. [Google Scholar] [CrossRef]
  97. Zheng, G.; Peng, Z.; Jia, D.; Al-Enizi, A.M.; Elzatahry, A.A. From Water Oxidation to Reduction: Homologous Ni-Co Based Nanowires as Complementary Water Splitting Electrocatalysts. Adv. Energy Mater. 2015, 5, 1–7. [Google Scholar]
  98. Huang, Z.F.; Song, J.; Li, K.; Tahir, M.; Wang, Y.T.; Pan, L.; Wang, L.; Zhang, X.; Zou, J.J. Hollow Cobalt-Based Bimetallic Sulfide Polyhedra for Efficient All-pH-Value Electrochemical and Photocatalytic Hydrogen Evolution. J. Am. Chem. Soc. 2016, 138, 1359–1365. [Google Scholar] [CrossRef]
  99. Boppella, R.; Tan, J.; Yang, W.; Moon, J. Homologous CoP/NiCoP Heterostructure on N-Doped Carbon for Highly Efficient and pH-Universal Hydrogen Evolution Electrocatalysis. Adv. Funct. Mater. 2019, 29, 1–9. [Google Scholar] [CrossRef]
  100. Zou, X.; Huang, X.; Goswami, A.; Silva, R.; Sathe, B.R.; Mikmeková, E.; Asefa, T. Cobalt-embedded nitrogen-rich carbon nanotubes efficiently catalyze hydrogen evolution reaction at all pH values. Angew. Chem. Int. Ed. 2014, 53, 4372–4376. [Google Scholar] [CrossRef]
  101. Xie, L.; Zhang, R.; Cui, L.; Liu, D.; Hao, S.; Ma, Y.; Du, G.; Asiri, A.M.; Sun, X. High-Performance Electrolytic Oxygen Evolution in Neutral Media Catalyzed by a Cobalt Phosphate Nanoarray. Angew. Chem. Int. Ed. 2017, 56, 1064–1068. [Google Scholar] [CrossRef]
  102. Tian, J.; Liu, Q.; Asiri, A.M.; Sun, X. 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]
  103. Tabassum, H.; Guo, W.; Meng, W.; Mahmood, A.; Zhao, R.; Wang, Q.; Zou, R. Metal–Organic Frameworks Derived Cobalt Phosphide Architecture Encapsulated into B/N Co-Doped Graphene Nanotubes for All pH Value Electrochemical Hydrogen Evolution. Adv. Energy Mater. 2017, 7, 1–7. [Google Scholar] [CrossRef]
  104. Liu, Y.; Xiao, C.; Lyu, M.; Lin, Y.; Cai, W.; Huang, P.; Tong, W.; Zou, Y.; Xie, Y. Ultrathin Co3S4Nanosheets that Synergistically Engineer Spin States and Exposed Polyhedra that Promote Water Oxidation under Neutral Conditions. Angew. Chem. Int. Ed. 2015, 54, 11231–11235. [Google Scholar] [CrossRef] [PubMed]
  105. Anantharaj, S.; Kennedy, J.; Kundu, S. Microwave-Initiated Facile Formation of Ni3Se4 Nanoassemblies for Enhanced and Stable Water Splitting in Neutral and Alkaline Media. ACS Appl. Mater. Interfaces 2017, 9, 8714–8728. [Google Scholar] [CrossRef] [PubMed]
  106. Wang, Y.; Cui, X.; Zhao, J.; Jia, G.; Gu, L.; Zhang, Q.; Meng, L.; Shi, Z.; Zheng, L.; Wang, C.; et al. Rational Design of Fe-N/C Hybrid for Enhanced Nitrogen Reduction Electrocatalysis under Ambient Conditions in Aqueous Solution. ACS Catal. 2019, 9, 336–344. [Google Scholar] [CrossRef]
  107. Zhang, G.; Feng, Y.S.; Lu, W.T.; He, D.; Wang, C.Y.; Li, Y.K.; Wang, X.Y.; Cao, F.F. Enhanced Catalysis of Electrochemical Overall Water Splitting in Alkaline Media by Fe Doping in Ni3S2 Nanosheet Arrays. ACS Catal. 2018, 8, 5431–5441. [Google Scholar] [CrossRef]
  108. Cheng, N.; Liu, Q.; Asiri, A.M.; Xing, W.; Sun, X. A Fe-Doped Ni3S2 Particle Film as a High-Efficiency Robust Oxygen Evolution Electrode with very High Current Density. J. Mater. Chem. A 2015, 3, 23207–23212. [Google Scholar] [CrossRef]
  109. Roger, I.; Shipman, M.A.; Symes, M.D. Earth-Abundant Catalysts for Electrochemical and Photoelectrochemical Water Splitting. Nat. Rev. Chem. 2017, 1, 1–13. [Google Scholar] [CrossRef]
  110. Zhou, H.; Yu, F.; Huang, Y.; Sun, J.; Zhu, Z.; Nielsen, R.J.; He, R.; Bao, J.; Goddard, W.A.; Chen, S.; et al. Efficient Hydrogen Evolution by Ternary Molybdenum Sulfoselenide Particles on Self-Standing Porous Nickel Diselenide Foam. Nat. Commun. 2016, 7, 1–7. [Google Scholar] [CrossRef]
  111. Xie, D.; Tang, W.; Wang, Y.; Xia, X.; Zhong, Y.; Zhou, D.; Wang, D.; Wang, X.; Tu, J. Facile Fabrication of Integrated Three-Dimensional C-MoSe2/Reduced Graphene Oxide Composite with Enhanced Performance for Sodium Storage. Nano Res. 2016, 9, 1618–1629. [Google Scholar] [CrossRef]
  112. Zhou, H.; Yu, F.; Sun, J.; Zhu, H.; Mishra, I.K.; Chen, S.; Ren, Z. Highly Efficient Hydrogen Evolution from Edge-Oriented WS2(1-x)Se2x Particles on Three-Dimensional Porous NiSe2 Foam. Nano Lett. 2016, 16, 7604–7609. [Google Scholar] [CrossRef] [PubMed]
  113. Xu, K.; Wang, F.; Wang, Z.; Zhan, X.; Wang, Q.; Cheng, Z.; Safdar, M.; He, J. Component-Controllable WS2(1-x)Se2x Nanotubes for Efficient Hydrogen Evolution Reaction. ACS Nano 2014, 8, 8468–8476. [Google Scholar] [CrossRef] [PubMed]
  114. Wang, K.; Zhou, C.; Xi, D.; Shi, Z.; He, C.; Xia, H.; Liu, G.; Qiao, G. Component-Controllable Synthesis of Co(SxSe1-x)2 Nanowires Supported by Carbon Fiber Paper as High-Performance Electrode for Hydrogen Evolution Reaction. Nano Energy 2015, 18, 1–11. [Google Scholar] [CrossRef]
  115. Ma, X.; Ma, M.; Liu, D.; Hao, S.; Qu, F.; Du, G.; Asiri, A.M.; Sun, X. Core–Shell-Structured NiS2@Ni-Bi Nanoarray for Efficient Water Oxidation at Near-Neutral pH. ChemCatChem 2017, 9, 3138–3143. [Google Scholar] [CrossRef]
  116. Zeng, L.; Sun, K.; Chen, Y.; Liu, Z.; Chen, Y.; Pan, Y.; Zhao, R.; Liu, Y.; Liu, C. Neutral-pH Overall Water Splitting Catalyzed Efficiently by a Hollow and Porous Structured Ternary Nickel Sulfoselenide Electrocatalyst. J. Mater. Chem. A 2019, 7, 16793–16802. [Google Scholar] [CrossRef]
  117. Anjum, M.A.R.; Lee, J.S. Sulfur and Nitrogen Dual-Doped Molybdenum Phosphide Nanocrystallites as an Active and Stable Hydrogen Evolution Reaction Electrocatalyst in Acidic and Alkaline Media. ACS Catal. 2017, 7, 3030–3038. [Google Scholar] [CrossRef]
  118. Wang, R.; Dong, X.Y.; Du, J.; Zhao, J.Y.; Zang, S.Q. MOF-Derived Bifunctional Cu3P Nanoparticles Coated by a N,P-Codoped Carbon Shell for Hydrogen Evolution and Oxygen Reduction. Adv. Mater. 2018, 30, 1–10. [Google Scholar]
  119. Du, J.; Wang, R.; Lv, Y.R.; Wei, Y.L.; Zang, S.Q. One-Step MOF-Derived Co/Co9S8 Nanoparticles Embedded in Nitrogen, Sulfur and Oxygen Ternary-Doped Porous Carbon: An Efficient Electrocatalyst for Overall Water Splitting. Chem. Commun. 2019, 55, 3203–3206. [Google Scholar] [CrossRef]
  120. Li, S.; Xi, C.; Jin, Y.Z.; Wu, D.; Wang, J.Q.; Liu, T.; Wang, H.B.; Dong, C.K.; Liu, H.; Kulinich, S.A.; et al. Ir-O-V Catalytic Group in Ir-Doped NiV(OH)2 for Overall Water Splitting. ACS Energy Lett. 2019, 4, 1823–1829. [Google Scholar] [CrossRef]
  121. Wu, Z.; Zou, Z.; Huang, J.; Gao, F. NiFe2O4 Nanoparticles/NiFe Layered Double-Hydroxide Nanosheet Heterostructure Array for Efficient Overall Water Splitting at Large Current Densities. ACS Appl. Mater. Interfaces 2018, 10, 26283–26292. [Google Scholar] [CrossRef] [PubMed]
  122. Liu, J.; Wang, J.; Zhang, B.; Ruan, Y.; Lv, L.; Ji, X.; Xu, K.; Miao, L.; Jiang, J. Hierarchical NiCo2S4@NiFe LDH Heterostructures Supported on Nickel Foam for Enhanced Overall-Water-Splitting Activity. ACS Appl. Mater. Interfaces 2017, 9, 15364–15372. [Google Scholar] [CrossRef]
  123. Zhang, Y.; Shao, Q.; Pi, Y.; Guo, J.; Huang, X. A Cost-Efficient Bifunctional Ultrathin Nanosheets Array for Electrochemical Overall Water Splitting. Small 2017, 13, 1–7. [Google Scholar] [CrossRef]
  124. Jiang, J.; Sun, F.; Zhou, S.; Hu, W.; Zhang, H.; Dong, J.; Jiang, Z.; Zhao, J.; Li, J.; Yan, W.; et al. Atomic-Level Insight into Super-Efficient Electrocatalytic Oxygen Evolution on Iron and Vanadium Co-Doped Nickel (oxy)hydroxide. Nat. Commun. 2018, 9, 1–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Fan, K.; Chen, H.; Ji, Y.; Huang, H.; Claesson, P.M.; Daniel, Q.; Philippe, B.; Rensmo, H.; Li, F.; Luo, Y.; et al. Nickel-Vanadium Monolayer Double Hydroxide for Efficient Electrochemical Water Oxidation. Nat. Commun. 2016, 7, 1–9. [Google Scholar] [CrossRef]
  126. Dinh, K.N.; Zheng, P.; Dai, Z.; Zhang, Y.; Dangol, R.; Zheng, Y.; Li, B.; Zong, Y.; Yan, Q. Ultrathin Porous NiFeV Ternary Layer Hydroxide Nanosheets as a Highly Efficient Bifunctional Electrocatalyst for Overall Water Splitting. Small 2018, 14, 1–9. [Google Scholar] [CrossRef] [PubMed]
  127. Stevens, M.B.; Trang, C.D.M.; Enman, L.J.; Deng, J.; Boettcher, S.W. Reactive Fe-Sites in Ni/Fe (Oxy)hydroxide Are Responsible for Exceptional Oxygen Electrocatalysis Activity. J. Am. Chem. Soc. 2017, 139, 11361–11364. [Google Scholar] [CrossRef]
  128. Stevens, M.B.; Enman, L.J.; Batchellor, A.S.; Cosby, M.R.; Vise, A.E.; Trang, C.D.M.; Boettcher, S.W. Measurement Techniques for the Study of Thin Film Heterogeneous Water Oxidation Electrocatalysts. Chem. Mater. 2017, 29, 120–140. [Google Scholar] [CrossRef]
  129. Trotochaud, L.; Young, S.L.; Ranney, J.K.; Boettcher, S.W. Nickel-Iron Oxyhydroxide Oxygen-Evolution Electrocatalysts: The Role of Intentional and Incidental Iron Incorporation. J. Am. Chem. Soc. 2014, 136, 6744–6753. [Google Scholar] [CrossRef]
  130. Zeng, L.; Sun, K.; Yang, Z.; Xie, S.; Chen, Y.; Liu, Z.; Liu, Y.; Zhao, J.; Liu, Y.; Liu, C. Tunable 3D Hierarchical Ni3S2 Superstructures as Efficient and Stable Bifunctional Electrocatalysts for both H2 and O2 Generation. J. Mater. Chem. A 2018, 6, 4485–4493. [Google Scholar] [CrossRef]
  131. Wu, Y.; Li, G.D.; Liu, Y.; Yang, L.; Lian, X.; Asefa, T.; Zou, X. Overall Water Splitting Catalyzed Efficiently by an Ultrathin Nanosheet-Built, Hollow Ni3S2-Based Electrocatalyst. Adv. Funct. Mater. 2016, 26, 4839–4847. [Google Scholar] [CrossRef]
  132. Yang, Y.; Zhang, K.; Lin, H.; Li, X.; Chan, H.C.; Yang, L.; Gao, Q. MoS2-Ni3S2 Heteronanorods as Efficient and Stable Bifunctional Electrocatalysts for Overall Water Splitting. ACS Catal. 2017, 7, 2357–2366. [Google Scholar] [CrossRef]
  133. Yan, Y.; Xia, B.Y.; Zhao, B.; Wang, X. A Review on Noble-Metal-Free Bifunctional Heterogeneous Catalysts for Overall Electrochemical Water Splitting. J. Mater. Chem. A 2016, 4, 17587–17603. [Google Scholar] [CrossRef] [Green Version]
  134. Zhou, W.; Wu, X.J.; Cao, X.; Huang, X.; Tan, C.; Tian, J.; Liu, H.; Wang, J.; Zhang, H. Ni3S2 Nanorods/Ni Foam Composite Electrode with Low Overpotential for Electrocatalytic Oxygen Evolution. Energy Environ. Sci. 2013, 6, 2921–2924. [Google Scholar] [CrossRef]
  135. Feng, J.X.; Xu, H.; Dong, Y.T.; Ye, S.H.; Tong, Y.X.; Li, G.R. FeOOH/Co/FeOOH Hybrid Nanotube Arrays as High-Performance Electrocatalysts for the Oxygen Evolution Reaction. Angew. Chem. Int. Ed. 2016, 55, 3694–3698. [Google Scholar] [CrossRef] [PubMed]
  136. Yu, Z.Y.; Duan, Y.; Gao, M.R.; Lang, C.C.; Zheng, Y.R.; Yu, S.H. A One-Dimensional Porous Carbon-Supported Ni/Mo2C Dual Catalyst for Efficient Water Splitting. Chem. Sci. 2017, 8, 968–973. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Zhuang, Z.; Sheng, W.; Yan, Y. Synthesis of Monodispere Au@Co3O4 Core-Shell Nanocrystals and Their Enhanced Catalytic Activity for Oxygen Evolution Reaction. Adv. Mater. 2014, 26, 3950–3955. [Google Scholar] [CrossRef]
  138. Song, F.; Hu, X. Exfoliation of Layered Double Hydroxides for Enhanced Oxygen Evolution Catalysis. Nat. Commun. 2014, 5, 1–9. [Google Scholar] [CrossRef]
  139. Tian, J.; Cheng, N.; Liu, Q.; Sun, X.; He, Y.; Asiri, A.M. Self-Supported NiMo Hollow Nanorod Array: An Efficient 3D Bifunctional Catalytic Electrode for Overall Water Splitting. J. Mater. Chem. A 2015, 3, 20056–20059. [Google Scholar] [CrossRef]
  140. Zhang, C.; Zhao, J.; Zhou, L.; Li, Z.; Shao, M.; Wei, M. Layer-by-Layer Assembly of Exfoliated Layered Double Hydroxide Nanosheets for Enhanced Electrochemical Oxidation of Water. J. Mater. Chem. A 2016, 4, 11516–11523. [Google Scholar] [CrossRef]
  141. Li, P.; Jin, Z.; Wang, R.; Jin, Y.; Xiao, D. Three-Dimensional Flexible Electrode Derived from Low-Cost Nickel-Phytate with Improved Electrochemical Performance. J. Mater. Chem. A 2016, 4, 9486–9495. [Google Scholar] [CrossRef]
  142. Xia, C.; Jiang, Q.; Zhao, C.; Hedhili, M.N.; Alshareef, H.N. Selenide-Based Electrocatalysts and Scaffolds for Water Oxidation Applications. Adv. Mater. 2016, 28, 77–85. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Xiao, C.; Li, Y.; Lu, X.; Zhao, C. Bifunctional Porous NiFe/NiCo2O4/Ni Foam Electrodes with Triple Hierarchy and Double Synergies for Efficient Whole Cell Water Splitting. Adv. Funct. Mater. 2016, 26, 3515–3523. [Google Scholar] [CrossRef]
  144. Hou, Y.; Lohe, M.R.; Zhang, J.; Liu, S.; Zhuang, X.; Feng, X. Vertically Oriented Cobalt Selenide/NiFe Layered-Double-Hydroxide Nanosheets Supported on Exfoliated Graphene Foil: An Efficient 3D Electrode for Overall Water Splitting. Energy Environ. Sci. 2016, 9, 478–483. [Google Scholar] [CrossRef] [Green Version]
  145. Liu, T.; Asiri, A.M.; Sun, X. Electrodeposited Co-Doped NiSe2 Nanoparticles Film: A Good Electrocatalyst for Efficient Water Splitting. Nanoscale 2016, 8, 3911–3915. [Google Scholar] [CrossRef]
  146. Yan, X.; Li, K.; Lyu, L.; Song, F.; He, J.; Niu, D.; Liu, L.; Hu, X.; Chen, X. From Water Oxidation to Reduction: Transformation from NixCo3-xO4 Nanowires to NiCo/NiCoOx Heterostructures. ACS Appl. Mater. Interfaces 2016, 8, 3208–3214. [Google Scholar] [CrossRef]
  147. Liu, D.; Lu, Q.; Luo, Y.; Sun, X.; Asiri, A.M. NiCo2S4 Nanowires Array as an Efficient Bifunctional Electrocatalyst for Full Water Splitting with Superior Activity. Nanoscale 2015, 7, 15122–15126. [Google Scholar] [CrossRef]
  148. Yu, Y.; Li, P.; Wang, X.; Gao, W.; Shen, Z.; Zhu, Y.; Yang, S.; Song, W.; Ding, K. Vanadium Nanobelts Coated Nickel Foam 3D Bifunctional Electrode with Excellent Catalytic Activity and Stability for Water Electrolysis. Nanoscale 2016, 8, 10731–10738. [Google Scholar] [CrossRef]
  149. Jiang, J.; Gao, M.; Sheng, W.; Yan, Y. Hollow Chevrel-Phase NiMo3S4 for Hydrogen Evolution in Alkaline Electrolytes. Angew. Chem. Int. Ed. 2016, 55, 15240–15245. [Google Scholar] [CrossRef]
  150. Fang, M.; Gao, W.; Dong, G.; Xia, Z.; Yip, S.P.; Qin, Y.; Qu, Y.; Ho, J.C. Hierarchical NiMo-Based 3D Electrocatalysts for Highly-Efficient Hydrogen Evolution in Alkaline Conditions. Nano Energy 2016, 27, 247–254. [Google Scholar] [CrossRef]
  151. Wang, T.; Wang, X.; Liu, Y.; Zheng, J.; Li, X. A Highly Efficient and Stable Biphasic Nanocrystalline Ni-Mo-N Catalyst for Hydrogen Evolution in both Acidic and Alkaline Electrolytes. Nano Energy 2016, 22, 111–119. [Google Scholar] [CrossRef]
  152. Xu, K.; Chen, P.; Li, X.; Tong, Y.; Ding, H.; Wu, X.; Chu, W.; Peng, Z.; Wu, C.; Xie, Y. Metallic Nickel Nitride Nanosheets Realizing Enhanced Electrochemical Water Oxidation. J. Am. Chem. Soc. 2015, 137, 4119–4125. [Google Scholar] [CrossRef] [PubMed]
  153. Nandi, S.; Singh, S.K.; Mullangi, D.; Illathvalappil, R.; George, L.; Vinod, C.P.; Kurungot, S.; Vaidhyanathan, R. Low Band Gap Benzimidazole COF Supported Ni3N as Highly Active OER Catalyst. Adv. Energy Mater. 2016, 6, 1–11. [Google Scholar] [CrossRef]
  154. Masa, J.; Sinev, I.; Mistry, H.; Ventosa, E.; De La Mata, M.; Arbiol, J.; Muhler, M.; Roldan Cuenya, B.; Schuhmann, W. Ultrathin High Surface Area Nickel Boride (NixB) Nanosheets as Highly Efficient Electrocatalyst for Oxygen Evolution. Adv. Energy Mater. 2017, 7, 1–8. [Google Scholar] [CrossRef] [Green Version]
  155. Yang, C.; Gao, M.Y.; Zhang, Q.B.; Zeng, J.R.; Li, X.T.; Abbott, A.P. In-Situ Activation of Self-Supported 3D Hierarchically Porous Ni3S2 Films Grown on Nanoporous Copper as Excellent pH-Universal Electrocatalysts for Hydrogen Evolution Reaction. Nano Energy 2017, 36, 85–94. [Google Scholar] [CrossRef] [Green Version]
  156. Wu, A.; Xie, Y.; Ma, H.; Tian, C.; Gu, Y.; Yan, H.; Zhang, X.; Yang, G.; Fu, H. Integrating the Active OER and HER Components as the Heterostructures for the Efficient Overall Water Splitting. Nano Energy 2018, 44, 353–363. [Google Scholar] [CrossRef]
  157. Ouyang, B.; Zhang, Y.; Zhang, Z.; Fan, H.J.; Rawat, R.S. Nitrogen-Plasma-Activated Hierarchical Nickel Nitride Nanocorals for Energy Applications. Small 2017, 13, 1–10. [Google Scholar] [CrossRef] [PubMed]
  158. Wang, X.R.; Liu, J.Y.; Liu, Z.W.; Wang, W.C.; Luo, J.; Han, X.P.; Du, X.W.; Qiao, S.Z.; Yang, J. Identifying the Key Role of Pyridinic-N–Co Bonding in Synergistic Electrocatalysis for Reversible ORR/OER. Adv. Mater. 2018, 30, 1–10. [Google Scholar] [CrossRef]
  159. Lin, Y.; Yang, L.; Zhang, Y.; Jiang, H.; Xiao, Z.; Wu, C.; Zhang, G.; Jiang, J.; Song, L. Defective Carbon–CoP Nanoparticles Hybrids with Interfacial Charges Polarization for Efficient Bifunctional Oxygen Electrocatalysis. Adv. Energy Mater. 2018, 8, 1–7. [Google Scholar] [CrossRef]
  160. Liu, Z.Q.; Cheng, H.; Li, N.; Ma, T.Y.; Su, Y.Z. ZnCo2O4 Quantum Dots Anchored on Nitrogen-Doped Carbon Nanotubes as Reversible Oxygen Reduction/Evolution Electrocatalysts. Adv. Mater. 2016, 28, 3777–3784. [Google Scholar] [CrossRef]
  161. Fu, Y.; Yu, H.Y.; Jiang, C.; Zhang, T.H.; Zhan, R.; Li, X.; Li, J.F.; Tian, J.H.; Yang, R. NiCo Alloy Nanoparticles Decorated on N-Doped Carbon Nanofibers as Highly Active and Durable Oxygen Electrocatalyst. Adv. Funct. Mater. 2018, 28, 1–10. [Google Scholar] [CrossRef]
  162. Li, F.L.; Shao, Q.; Huang, X.; Lang, J.P. Nanoscale Trimetallic Metal–Organic Frameworks Enable Efficient Oxygen Evolution Electrocatalysis. Angew. Chem. Int. Ed. 2018, 57, 1888–1892. [Google Scholar] [CrossRef] [PubMed]
  163. Wang, J.; Xu, F.; Jin, H.; Chen, Y.; Wang, Y. Non-Noble Metal-based Carbon Composites in Hydrogen Evolution Reaction: Fundamentals to Applications. Adv. Mater. 2017, 29, 1605838–1605872. [Google Scholar] [CrossRef]
  164. Yu, L.; Xia, B.Y.; Wang, X.; Lou, X.W. General Formation of M-MoS3 (M = Co, Ni) Hollow Structures with Enhanced Electrocatalytic Activity for Hydrogen Evolution. Adv. Mater. 2016, 28, 92–97. [Google Scholar] [CrossRef] [PubMed]
  165. Li, G.; Yu, J.; Jia, J.; Yang, L.; Zhao, L.; Zhou, W.; Liu, H. Cobalt–Cobalt Phosphide Nanoparticles@Nitrogen-Phosphorus Doped Carbon/Graphene Derived from Cobalt Ions Adsorbed Saccharomycete Yeasts as an Efficient, Stable, and Large-Current-Density Electrode for Hydrogen Evolution Reactions. Adv. Funct. Mater. 2018, 28, 1–9. [Google Scholar] [CrossRef]
  166. Najafi, L.; Bellani, S.; Oropesa-Nuñez, R.; Ansaldo, A.; Prato, M.; Del Rio Castillo, A.E.; Bonaccorso, F. Engineered MoSe2-Based Heterostructures for Efficient Electrochemical Hydrogen Evolution Reaction. Adv. Energy Mater. 2018, 8, 1–16. [Google Scholar] [CrossRef] [Green Version]
  167. Yun, Q.; Lu, Q.; Zhang, X.; Tan, C.; Zhang, H. Dreidimensionale Architekturen aus Übergangsmetall-Dichalkogenid-Nanomaterialien zur elektrochemischen Energiespeicherung und -umwandlung. Angew. Chem. 2018, 130, 634–655. [Google Scholar] [CrossRef]
  168. Ouyang, T.; Chen, A.N.; He, Z.Z.; Liu, Z.Q.; Tong, Y. Rational Design of Atomically Dispersed Nickel Active Sites in β-Mo2C for the Hydrogen Evolution Reaction at all pH Values. Chem. Commun. 2018, 54, 9901–9904. [Google Scholar] [CrossRef]
  169. Cheng, H.; Ding, L.X.; Chen, G.F.; Zhang, L.; Xue, J.; Wang, H. Molybdenum Carbide Nanodots Enable Efficient Electrocatalytic Nitrogen Fixation under Ambient Conditions. Adv. Mater. 2018, 30, 1–7. [Google Scholar] [CrossRef]
  170. Ouyang, T.; Ye, Y.Q.; Wu, C.Y.; Xiao, K.; Liu, Z.Q. Heterostructures Composed of N-Doped Carbon Nanotubes Encapsulating Cobalt and β-Mo2C Nanoparticles as Bifunctional Electrodes for Water Splitting. Angew. Chem. Int. Ed. 2019, 58, 4923–4928. [Google Scholar] [CrossRef]
  171. Freunberger, S.A.; Chen, Y.; Drewett, N.E.; Hardwick, L.J.; Bardé, F.; Bruce, P.G. The Lithium-Oxygen Battery with Ether-Based Electrolytes. Angew. Chem. Int. Ed. 2011, 50, 8609–8613. [Google Scholar] [CrossRef]
  172. Freunberger, S.A.; Chen, Y.; Peng, Z.; Griffin, J.M.; Hardwick, L.J.; Bardé, F.; Novák, P.; Bruce, P.G. Reactions in the Rechargeable Lithium-O2 Battery with Alkyl Carbonate Electrolytes. J. Am. Chem. Soc. 2011, 133, 8040–8047. [Google Scholar] [CrossRef] [PubMed]
  173. Chen, Y.; Freunberger, S.A.; Peng, Z.; Bardé, F.; Bruce, P.G. Li-O2 Battery with a Dimethylformamide Electrolyte. J. Am. Chem. Soc. 2012, 134, 7952–7957. [Google Scholar] [CrossRef] [PubMed]
  174. McCloskey, B.D.; Speidel, A.; Scheffler, R.; Miller, D.C.; Viswanathan, V.; Hummelshøj, J.S.; Nørskov, J.K.; Luntz, A.C. Twin Problems of Interfacial Carbonate Formation in Nonaqueous Li-O2 Batteries. J. Phys. Chem. Lett. 2012, 3, 997–1001. [Google Scholar] [CrossRef]
  175. Gallant, B.M.; Mitchell, R.R.; Kwabi, D.G.; Zhou, J.; Zuin, L.; Thompson, C.V.; Shao-Horn, Y. Chemical and Morphological Changes of Li-O2 Battery Electrodes upon Cycling. J. Phys. Chem. C 2012, 116, 20800–20805. [Google Scholar] [CrossRef]
  176. Grande, L.; Paillard, E.; Hassoun, J.; Park, J.B.; Lee, Y.J.; Sun, Y.K.; Passerini, S.; Scrosati, B. The Lithium/Air Battery: Still an Emerging System or a Practical Reality? Adv. Mater. 2015, 27, 784–800. [Google Scholar] [CrossRef]
  177. Kim, B.G.; Kim, H.J.; Back, S.; Nam, K.W.; Jung, Y.; Han, Y.K.; Choi, J.W. Improved Reversibility in Lithium-Oxygen Battery: Understanding Elementary Reactions and Surface Charge Engineering of Metal Alloy Catalyst. Sci. Rep. 2015, 4, 18–21. [Google Scholar] [CrossRef]
  178. Sun, B.; Huang, X.; Chen, S.; Munroe, P.; Wang, G. Porous Graphene Nanoarchitectures: An Efficient Catalyst for Low Charge-Overpotential, Long Life, and High Capacity Lithium-Oxygen Batteries. Nano Lett. 2014, 14, 3145–3152. [Google Scholar] [CrossRef]
  179. Ottakam Thotiyl, M.M.; Freunberger, S.A.; Peng, Z.; Chen, Y.; Liu, Z.; Bruce, P.G. A Stable Cathode for the Aprotic Li-O2 Battery. Nat. Mater. 2013, 12, 1050–1056. [Google Scholar] [CrossRef]
  180. Lu, Y.C.; Xu, Z.; Gasteiger, H.A.; Chen, S.; Hamad-Schifferli, K.; Shao-Horn, Y. Platinum-Gold Nanoparticles: A Highly Active Bifunctional Electrocatalyst for Rechargeable Lithium-Air Batteries. J. Am. Chem. Soc. 2010, 132, 12170–12171. [Google Scholar] [CrossRef]
  181. Li, F.; Tang, D.M.; Zhang, T.; Liao, K.; He, P.; Golberg, D.; Yamada, A.; Zhou, H. Superior Performance of a Li-O2 Battery with Metallic RuO2 Hollow Spheres as the Carbon-Free Cathode. Adv. Energy Mater. 2015, 5, 1–6. [Google Scholar] [CrossRef]
  182. Yang, X.H.; He, P.; Xia, Y.Y. Preparation of Mesocellular Carbon Foam and Its Application for Lithium/Oxygen Battery. Electrochem. Commun. 2009, 11, 1127–1130. [Google Scholar] [CrossRef]
  183. Read, J. Characterization of the Lithium/Oxygen Organic Electrolyte Battery. J. Electrochem. Soc. 2002, 149, A1190–A1195. [Google Scholar] [CrossRef]
  184. Zhang, J.G.; Xiao, J.; Xu, W. Primary Lithium Air Batteries. In The Lithium Air Battery: Fundamentals; Springer: New York, NY, USA, 2014; pp. 225–289. [Google Scholar]
  185. Chen, Y.; Freunberger, S.A.; Peng, Z.; Fontaine, O.; Bruce, P.G. Charging a Li-O2 Battery Using a Redox Mediator. Nat. Chem. 2013, 5, 489–494. [Google Scholar] [CrossRef] [PubMed]
  186. Lacey, M.J.; Frith, J.T.; Owen, J.R. A Redox Shuttle to Facilitate Oxygen Reduction in the Lithium Air Battery. Electrochem. Commun. 2013, 26, 74–76. [Google Scholar] [CrossRef]
  187. Bergner, B.J.; Schürmann, A.; Peppler, K.; Garsuch, A.; Janek, J. TEMPO: A Mobile Catalyst for Rechargeable Li-O2 Batteries. J. Am. Chem. Soc. 2014, 136, 15054–15064. [Google Scholar] [CrossRef] [PubMed]
  188. Lim, H.D.; Song, H.; Kim, J.; Gwon, H.; Bae, Y.; Park, K.Y.; Hong, J.; Kim, H.; Kim, T.; Kim, Y.H.; et al. Superior Rechargeability and Efficiency of Lithium-Oxygen Batteries: Hierarchical Air Electrode Architecture Combined with a Soluble Catalyst. Angew. Chem. Int. Ed. 2014, 53, 3926–3931. [Google Scholar] [CrossRef]
  189. Sun, D.; Shen, Y.; Zhang, W.; Yu, L.; Yi, Z.; Yin, W.; Wang, D.; Huang, Y.; Wang, J.; Wang, D.; et al. A Solution-Phase Bifunctional Catalyst for Lithium-Oxygen Batteries. J. Am. Chem. Soc. 2014, 136, 8941–8946. [Google Scholar] [CrossRef]
  190. Bergner, B.J.; Hofmann, C.; Schürmann, A.; Schröder, D.; Peppler, K.; Schreiner, P.R.; Janek, J. Understanding the Fundamentals of Redox Mediators in Li-O2 batteries: A Case Study on Nitroxides. Phys. Chem. Chem. Phys. 2015, 17, 31769–31779. [Google Scholar] [CrossRef] [Green Version]
  191. Kwak, W.J.; Hirshberg, D.; Sharon, D.; Shin, H.J.; Afri, M.; Park, J.B.; Garsuch, A.; Chesneau, F.F.; Frimer, A.A.; Aurbach, D.; et al. Understanding the Behavior of Li-Oxygen Cells Containing LiI. J. Mater. Chem. A 2015, 3, 8855–8864. [Google Scholar] [CrossRef]
  192. Kundu, D.; Black, R.; Adams, B.; Nazar, L.F. A Highly Active Low Voltage Redox Mediator for Enhanced Rechargeability of Lithium-Oxygen Batteries. ACS Cent. Sci. 2015, 1, 510–515. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Pan, F.; Yang, J.; Huang, Q.; Wang, X.; Huang, H.; Wang, Q. Redox Targeting of Anatase TiO2 for Redox Flow Lithium-Ion Batteries. Adv. Energy Mater. 2014, 4, 1–7. [Google Scholar]
  194. Huang, Q.; Li, H.; Grätzel, M.; Wang, Q. Reversible Chemical Delithiation/Lithiation of LiFePO4: Towards a Redox Flow Lithium-Ion Battery. Phys. Chem. Chem. Phys. 2013, 15, 1793–1797. [Google Scholar] [CrossRef]
  195. Huang, Q.; Wang, Q. Next-Generation, High-Energy-Density Redox Flow Batteries. Chempluschem 2015, 80, 312–322. [Google Scholar] [CrossRef]
  196. Jia, C.; Pan, F.; Zhu, Y.G.; Huang, Q.; Lu, L.; Wang, Q. High-Energy Density Nonaqueous all Redox Flow Lithium Battery Enabled with a Polymeric Membrane. Sci. Adv. 2015, 1, 1–7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Zhu, Y.G.; Jia, C.; Yang, J.; Pan, F.; Huang, Q.; Wang, Q. Dual Redox Catalysts for Oxygen Reduction and Evolution Reactions: Towards a Redox Flow Li-O2 Battery. Chem. Commun. 2015, 51, 9451–9454. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. Zhu, Y.G.; Wang, X.; Jia, C.; Yang, J.; Wang, Q. Redox-Mediated ORR and OER Reactions: Redox Flow Lithium Oxygen Batteries Enabled with a Pair of Soluble Redox Catalysts. ACS Catal. 2016, 6, 6191–6197. [Google Scholar] [CrossRef]
  199. Chen, S.; Duan, J.; Vasileff, A.; Qiao, S.Z. Size Fractionation of Two-Dimensional Sub-Nanometer Thin Manganese Dioxide Crystals towards Superior Urea Electrocatalytic Conversion. Angew. Chem. Int. Ed. 2016, 55, 3804–3808. [Google Scholar] [CrossRef]
  200. Muthuchamy, N.; Jang, S.; Park, J.C.; Park, S.; Park, K.H. Bimetallic NiPd Nanoparticle-Incorporated Ordered Mesoporous Carbon as Highly Efficient Electrocatalysts for Hydrogen Production via Overall Urea Electrolysis. ACS Sustain. Chem. Eng. 2019, 7, 15526–15536. [Google Scholar] [CrossRef]
Scheme 1. Schematic diagram of hybrid metal nanoparticles (NPs) from monometallic NPs.
Scheme 1. Schematic diagram of hybrid metal nanoparticles (NPs) from monometallic NPs.
Catalysts 10 00621 sch001
Scheme 2. Illustration of hydrogen evolution reaction (HER) and oxygen evolution reaction (OER) activity in water electrolysis.
Scheme 2. Illustration of hydrogen evolution reaction (HER) and oxygen evolution reaction (OER) activity in water electrolysis.
Catalysts 10 00621 sch002
Scheme 3. Various methods of improve efficiency in water splitting.
Scheme 3. Various methods of improve efficiency in water splitting.
Catalysts 10 00621 sch003
Scheme 4. Schematic image of oxygen reduction reaction (ORR) and OER battery.
Scheme 4. Schematic image of oxygen reduction reaction (ORR) and OER battery.
Catalysts 10 00621 sch004
Scheme 5. Illustration of energy conversion technologies based on urea. Reproduced and adapted from Ref. [77]; Copyright (2020), Wiley.
Scheme 5. Illustration of energy conversion technologies based on urea. Reproduced and adapted from Ref. [77]; Copyright (2020), Wiley.
Catalysts 10 00621 sch005
Scheme 6. Schematic diagram for achieving defect sites of lattice.
Scheme 6. Schematic diagram for achieving defect sites of lattice.
Catalysts 10 00621 sch006
Scheme 7. Schematic illustration for growth and synthesize Se-vacancy controlled MoSe2 with H2 concentration control. H2 plays a dual role in enhancing the reduction of MoO3 (decreased vacancies) and Se etching at the MoSe2 lattice (decreased vacancies). Reproduced and adapted from Ref. [79]; Copyright (2019), Elsevier.
Scheme 7. Schematic illustration for growth and synthesize Se-vacancy controlled MoSe2 with H2 concentration control. H2 plays a dual role in enhancing the reduction of MoO3 (decreased vacancies) and Se etching at the MoSe2 lattice (decreased vacancies). Reproduced and adapted from Ref. [79]; Copyright (2019), Elsevier.
Catalysts 10 00621 sch007
Figure 1. (a) Schematic diagram of lattice structure change in vacancy-MoSe2. Atomic-resolution annular dark-field-scanning transmission electron microscopy (ADF-STEM) position mapping image of (b) MoSe2-30 and (c) MoSe2-50. The bars indicate (b) and (c) 1 nm. Reproduced and adapted from Ref. [79]; Copyright (2019), Elsevier.
Figure 1. (a) Schematic diagram of lattice structure change in vacancy-MoSe2. Atomic-resolution annular dark-field-scanning transmission electron microscopy (ADF-STEM) position mapping image of (b) MoSe2-30 and (c) MoSe2-50. The bars indicate (b) and (c) 1 nm. Reproduced and adapted from Ref. [79]; Copyright (2019), Elsevier.
Catalysts 10 00621 g001
Figure 2. (a) HER polarization curve and (b) corresponding Tafel slopes of MoSe2-10, 20, 30, 40, 50 and Pt/C in 0.5 M H2SO4 aqueous solution. (c) Schematic explanation of HER in AV- and CV-MoSe2, which exhibit process of Tafel reaction in the coalesced Se-vacancy for hydrogen evolution. Reproduced and adapted from Ref. [79]; Copyright (2019), Elsevier.
Figure 2. (a) HER polarization curve and (b) corresponding Tafel slopes of MoSe2-10, 20, 30, 40, 50 and Pt/C in 0.5 M H2SO4 aqueous solution. (c) Schematic explanation of HER in AV- and CV-MoSe2, which exhibit process of Tafel reaction in the coalesced Se-vacancy for hydrogen evolution. Reproduced and adapted from Ref. [79]; Copyright (2019), Elsevier.
Catalysts 10 00621 g002
Scheme 8. Schematic diagram for synthesis of ternary NiS2(1−x)Se2x porous/hollow structure via NiS2. Reproduced and adapted from Ref. [116]; Copyright (2019), The Royal Society of Chemistry.
Scheme 8. Schematic diagram for synthesis of ternary NiS2(1−x)Se2x porous/hollow structure via NiS2. Reproduced and adapted from Ref. [116]; Copyright (2019), The Royal Society of Chemistry.
Catalysts 10 00621 sch008
Figure 3. Transmission electron microscopy (TEM) image of (a) NiS2 precursor and (b) NiS2(1−x)Se2x. The inset in (b) displays the surface of NiS2(1−x)Se2x. (c) Elemental mapping images of NiS2(1−x)Se2x, and (d) high-resolution TEM image of NiS2(1−x)Se2x. The bars indicate (a) 0.5 μm, (b) 50 nm, and (d) 5 nm. Reproduced and adapted from Ref. [116]; Copyright (2019), The Royal Society of Chemistry.
Figure 3. Transmission electron microscopy (TEM) image of (a) NiS2 precursor and (b) NiS2(1−x)Se2x. The inset in (b) displays the surface of NiS2(1−x)Se2x. (c) Elemental mapping images of NiS2(1−x)Se2x, and (d) high-resolution TEM image of NiS2(1−x)Se2x. The bars indicate (a) 0.5 μm, (b) 50 nm, and (d) 5 nm. Reproduced and adapted from Ref. [116]; Copyright (2019), The Royal Society of Chemistry.
Catalysts 10 00621 g003
Figure 4. (a) X-ray diffraction (XRD) pattern of NiS2(1−x)Se2x (red line) and NiS2 (blue line). The inset of (a) is crystal structure. (b) N2 sorption isotherm and measured pore size (the inset of (b)) of NiS2(1−x)Se2x. Reproduced and adapted from Ref. [116]; Copyright (2019), The Royal Society of Chemistry.
Figure 4. (a) X-ray diffraction (XRD) pattern of NiS2(1−x)Se2x (red line) and NiS2 (blue line). The inset of (a) is crystal structure. (b) N2 sorption isotherm and measured pore size (the inset of (b)) of NiS2(1−x)Se2x. Reproduced and adapted from Ref. [116]; Copyright (2019), The Royal Society of Chemistry.
Catalysts 10 00621 g004
Figure 5. HER electrocatalytic performance in neutral condition. (a) Polarization curves and (b) the corresponding Tafel plots for the catalysts. (c) Nyquist plots scanned at −0.25 V (vs. RHE) and (d) time-dependent current density curve of the Ni(S0.5Se0.5)2 at overpotential 125 mV. The inset exhibits linear sweep voltammetry (LSV) curves before and after 2000 cyclic voltammetry (CV) cycles to confirm stability. Reproduced and adapted from Ref. [116]; Copyright (2019), The Royal Society of Chemistry.
Figure 5. HER electrocatalytic performance in neutral condition. (a) Polarization curves and (b) the corresponding Tafel plots for the catalysts. (c) Nyquist plots scanned at −0.25 V (vs. RHE) and (d) time-dependent current density curve of the Ni(S0.5Se0.5)2 at overpotential 125 mV. The inset exhibits linear sweep voltammetry (LSV) curves before and after 2000 cyclic voltammetry (CV) cycles to confirm stability. Reproduced and adapted from Ref. [116]; Copyright (2019), The Royal Society of Chemistry.
Catalysts 10 00621 g005
Figure 6. OER electrocatalytic performance in neutral condition. (a) Polarization curves and (b) the corresponding Tafel plots for the catalysts. (c) Nyquist plots scanned at 1.75 V (vs. RHE) and (d) time-dependent current density curve of the Ni(S0.5Se0.5)2 at 1.73 V. The inset exhibits LSV curves before and after 2000 CV cycles to confirm stability. Reproduced and adapted from Ref. [116]; Copyright (2019), The Royal Society of Chemistry.
Figure 6. OER electrocatalytic performance in neutral condition. (a) Polarization curves and (b) the corresponding Tafel plots for the catalysts. (c) Nyquist plots scanned at 1.75 V (vs. RHE) and (d) time-dependent current density curve of the Ni(S0.5Se0.5)2 at 1.73 V. The inset exhibits LSV curves before and after 2000 CV cycles to confirm stability. Reproduced and adapted from Ref. [116]; Copyright (2019), The Royal Society of Chemistry.
Catalysts 10 00621 g006
Figure 7. Overall water splitting in neutral condition. (a) Polarization curves for an Ni(S0.5Se0.5)2‖Ni(S0.5Se0.5)2 and Pt/C‖IrO2 and (b) long-term durability test to achieve 10 mA cm−2. Reproduced and adapted from Ref. [116]; Copyright (2019), The Royal Society of Chemistry.
Figure 7. Overall water splitting in neutral condition. (a) Polarization curves for an Ni(S0.5Se0.5)2‖Ni(S0.5Se0.5)2 and Pt/C‖IrO2 and (b) long-term durability test to achieve 10 mA cm−2. Reproduced and adapted from Ref. [116]; Copyright (2019), The Royal Society of Chemistry.
Catalysts 10 00621 g007
Scheme 9. Illustration of heteroatoms on the NPs.
Scheme 9. Illustration of heteroatoms on the NPs.
Catalysts 10 00621 sch009
Figure 8. Active sites (top) in the overall water splitting process and differential-charge densities (down) of (a) NiV LDH and (b) NiVIr LDH. The isosurface value is 0.018 Å−3. (c) Scanning electron microscopy (SEM) image of surface morphology of NiVIr catalyst. (d) XRD spectrums of Ni(OH)2, NiV LDH, and NiVIr LDH, respectively. The bars indicate (c) 250 nm. Reproduced and adapted from Ref. [120]; Copyright (2019), American Chemical Society.
Figure 8. Active sites (top) in the overall water splitting process and differential-charge densities (down) of (a) NiV LDH and (b) NiVIr LDH. The isosurface value is 0.018 Å−3. (c) Scanning electron microscopy (SEM) image of surface morphology of NiVIr catalyst. (d) XRD spectrums of Ni(OH)2, NiV LDH, and NiVIr LDH, respectively. The bars indicate (c) 250 nm. Reproduced and adapted from Ref. [120]; Copyright (2019), American Chemical Society.
Catalysts 10 00621 g008
Figure 9. Polarization curves of (a) OER activity and (b) HER activity, respectively. Reproduced and adapted from Ref. [120]; Copyright (2019), American Chemical Society.
Figure 9. Polarization curves of (a) OER activity and (b) HER activity, respectively. Reproduced and adapted from Ref. [120]; Copyright (2019), American Chemical Society.
Catalysts 10 00621 g009
Scheme 10. Design of synthesizing Ni3S2 superstructures on nickel foam (NF). Reproduced and adapted from Ref. [130]; Copyright (2018), The Royal Society of Chemistry.
Scheme 10. Design of synthesizing Ni3S2 superstructures on nickel foam (NF). Reproduced and adapted from Ref. [130]; Copyright (2018), The Royal Society of Chemistry.
Catalysts 10 00621 sch010
Figure 10. (a,b) TEM images of RA-Ni3S2/NF and (c,d) those of NA-Ni3S2/NF. The bars indicate 100 nm (a), 10 nm (b), 500 nm (c), and 100 nm (d). Reproduced and adapted from Ref. [130]; Copyright (2018), The Royal Society of Chemistry.
Figure 10. (a,b) TEM images of RA-Ni3S2/NF and (c,d) those of NA-Ni3S2/NF. The bars indicate 100 nm (a), 10 nm (b), 500 nm (c), and 100 nm (d). Reproduced and adapted from Ref. [130]; Copyright (2018), The Royal Society of Chemistry.
Catalysts 10 00621 g010
Figure 11. (a) HER polarization curves of bare NF, RA-Ni3S2/NF, NF-Ni3S2, Pt/C on NF. (b) Initial HER polarization curves and after 5000 cycles HER polarization curves of RA-Ni3S2/NF, NF-Ni3S2/NF in 1.0 M KOH. (c) OER polarization curves of bare NF, RA-Ni3S2/NF, NF-Ni3S2/NF in 1.0 M KOH. (d) Chronoamperometry of NF-Ni3S2/NF and RA-Ni3S2/NF for OER at 300 mV value of overpotential 1.0 M KOH. Reproduced and adapted from Ref. [130]; Copyright (2018), The Royal Society of Chemistry.
Figure 11. (a) HER polarization curves of bare NF, RA-Ni3S2/NF, NF-Ni3S2, Pt/C on NF. (b) Initial HER polarization curves and after 5000 cycles HER polarization curves of RA-Ni3S2/NF, NF-Ni3S2/NF in 1.0 M KOH. (c) OER polarization curves of bare NF, RA-Ni3S2/NF, NF-Ni3S2/NF in 1.0 M KOH. (d) Chronoamperometry of NF-Ni3S2/NF and RA-Ni3S2/NF for OER at 300 mV value of overpotential 1.0 M KOH. Reproduced and adapted from Ref. [130]; Copyright (2018), The Royal Society of Chemistry.
Catalysts 10 00621 g011
Figure 12. (a) OER polarization curves and (b) CA of RA-Ni3S2/NF, NF-Ni3S2/NF at 1.63 V in 1.0 M KOH. Reproduced and adapted from Ref. [130]; Copyright (2018), The Royal Society of Chemistry.
Figure 12. (a) OER polarization curves and (b) CA of RA-Ni3S2/NF, NF-Ni3S2/NF at 1.63 V in 1.0 M KOH. Reproduced and adapted from Ref. [130]; Copyright (2018), The Royal Society of Chemistry.
Catalysts 10 00621 g012
Scheme 11. Illustration of various layers of heterostructure materials.
Scheme 11. Illustration of various layers of heterostructure materials.
Catalysts 10 00621 sch011
Figure 13. Illustration for the synthesis of Ni3N-NiMoN heterostructure. Reproduced and adapted from Ref. [156]; Copyright (2018), Elsevier.
Figure 13. Illustration for the synthesis of Ni3N-NiMoN heterostructure. Reproduced and adapted from Ref. [156]; Copyright (2018), Elsevier.
Catalysts 10 00621 g013
Figure 14. (a) SEM, (b) high-resolution TEM (HRTEM), (c) HAADF images of Ni3N-NiMoN heterostructure, (df) elemental mapping images of Mo, Ni, and N for Ni3N-NiMoN heterostructure, respectively. The bars indicate (a) 5 µm and (b) 5 nm. Reproduced and adapted from Ref. [156]; Copyright (2018), Elsevier.
Figure 14. (a) SEM, (b) high-resolution TEM (HRTEM), (c) HAADF images of Ni3N-NiMoN heterostructure, (df) elemental mapping images of Mo, Ni, and N for Ni3N-NiMoN heterostructure, respectively. The bars indicate (a) 5 µm and (b) 5 nm. Reproduced and adapted from Ref. [156]; Copyright (2018), Elsevier.
Catalysts 10 00621 g014
Figure 15. HER activity property of (a) polarization curves for Pt/C, various Ni3N-NiMoN-X (X = 4, 5, 6; 400, 500, and 600 °C), (b) durability of Ni3N-NiMoN-5 performed with a scan rate of 100 mV s−1 initial and after 5 K cycles. Inset of (b) current depends on time curves. OER activity property of (c) polarization curves for RuO2, various Ni3N-NiMoN-X, and (d) durability of Ni3N-NiMoN-5 performed with a scan rate of 100 mV s−1 initial and after 5 K cycles. Inset of (d) current depends on time curves. Reproduced and adapted from Ref. [156]; Copyright (2018), Elsevier.
Figure 15. HER activity property of (a) polarization curves for Pt/C, various Ni3N-NiMoN-X (X = 4, 5, 6; 400, 500, and 600 °C), (b) durability of Ni3N-NiMoN-5 performed with a scan rate of 100 mV s−1 initial and after 5 K cycles. Inset of (b) current depends on time curves. OER activity property of (c) polarization curves for RuO2, various Ni3N-NiMoN-X, and (d) durability of Ni3N-NiMoN-5 performed with a scan rate of 100 mV s−1 initial and after 5 K cycles. Inset of (d) current depends on time curves. Reproduced and adapted from Ref. [156]; Copyright (2018), Elsevier.
Catalysts 10 00621 g015
Figure 16. (a) Schematic image of synthesis Co/β-Mo2C@N-CNTs, (b) TEM and (c) HRTEM image of Co/β-Mo2C@N-CNTs. The bars indicate 20 nm (a) and 5 nm (b). Reproduced and adapted from Ref. [170]; Copyright (2019), Wiley.
Figure 16. (a) Schematic image of synthesis Co/β-Mo2C@N-CNTs, (b) TEM and (c) HRTEM image of Co/β-Mo2C@N-CNTs. The bars indicate 20 nm (a) and 5 nm (b). Reproduced and adapted from Ref. [170]; Copyright (2019), Wiley.
Catalysts 10 00621 g016
Figure 17. XRD of the Co/β-Mo2C@N-CNTs, Co, and β-Mo2C the well-defined peaks, respectively. Reproduced and adapted from Ref. [170]; Copyright (2019), Wiley.
Figure 17. XRD of the Co/β-Mo2C@N-CNTs, Co, and β-Mo2C the well-defined peaks, respectively. Reproduced and adapted from Ref. [170]; Copyright (2019), Wiley.
Catalysts 10 00621 g017
Figure 18. (a) Schematic diagram of the redox-flow lithium-oxygen batteries (RFLOB). (b) Molecular structure of the two redox mediator, 2,5-di-tert-butyl-p-benzoquinone (DTBBQ) and tris{4-[2-(2-methoxyethoxy)ethoxy]phenyl}amine (TMPPA). Reproduced and adapted from Ref. [198]; Copyright (2016), American Chemical Society.
Figure 18. (a) Schematic diagram of the redox-flow lithium-oxygen batteries (RFLOB). (b) Molecular structure of the two redox mediator, 2,5-di-tert-butyl-p-benzoquinone (DTBBQ) and tris{4-[2-(2-methoxyethoxy)ethoxy]phenyl}amine (TMPPA). Reproduced and adapted from Ref. [198]; Copyright (2016), American Chemical Society.
Catalysts 10 00621 g018
Figure 19. (a) CV of 10 mM DTBBQ (blue line) and TMPPA (green line) in TEGDME (1 M LiTFSI) electrolyte. (b) Illustration of demonstrating reactions DTBBQ- and TMPPA-catalyzed ORR and OER during the charge and discharge processes of RFLOB. Reproduced and adapted from Ref. [198]; Copyright (2016), American Chemical Society.
Figure 19. (a) CV of 10 mM DTBBQ (blue line) and TMPPA (green line) in TEGDME (1 M LiTFSI) electrolyte. (b) Illustration of demonstrating reactions DTBBQ- and TMPPA-catalyzed ORR and OER during the charge and discharge processes of RFLOB. Reproduced and adapted from Ref. [198]; Copyright (2016), American Chemical Society.
Catalysts 10 00621 g019
Figure 20. (a) RDE test of the OER reaction between 2 mM TMPPA and Li2O2 and (b) ORR reaction between 2 mM DTBBQ and O2 gas in 0.10 M LiTFSI in TEGDME. Galvanostatic tests of (c) 20 mM TMPPA and (d) 20 mM DTBBQ analyzed in static lithium half-cells under an Ar atmosphere. Reproduced and adapted from Ref. [198]; Copyright (2016), American Chemical Society.
Figure 20. (a) RDE test of the OER reaction between 2 mM TMPPA and Li2O2 and (b) ORR reaction between 2 mM DTBBQ and O2 gas in 0.10 M LiTFSI in TEGDME. Galvanostatic tests of (c) 20 mM TMPPA and (d) 20 mM DTBBQ analyzed in static lithium half-cells under an Ar atmosphere. Reproduced and adapted from Ref. [198]; Copyright (2016), American Chemical Society.
Catalysts 10 00621 g020
Scheme 12. Illustration of urea electrolysis by electrochemistry.
Scheme 12. Illustration of urea electrolysis by electrochemistry.
Catalysts 10 00621 sch012
Figure 21. Schematic illustration for preparation of the bimetallic NiPd/OMC catalyst.
Figure 21. Schematic illustration for preparation of the bimetallic NiPd/OMC catalyst.
Catalysts 10 00621 g021
Figure 22. (a) XRD spectra of NiPd/OMC, Ni/OMC, and Pd/OMC. (b) TEM image, and (c) HRTEM image of NiPd/OMC nanocatalyst. The bars indicate 20 nm (b) and 2 nm (c). Reproduced and adapted from Ref. [200]; Copyright (2019), American Chemical Society.
Figure 22. (a) XRD spectra of NiPd/OMC, Ni/OMC, and Pd/OMC. (b) TEM image, and (c) HRTEM image of NiPd/OMC nanocatalyst. The bars indicate 20 nm (b) and 2 nm (c). Reproduced and adapted from Ref. [200]; Copyright (2019), American Chemical Society.
Catalysts 10 00621 g022
Figure 23. Urea oxidation reaction (UOR) electrocatalytic properties of IrO@C, NiPd/OMC, Ni/OMC, Pd/OMC, and OMC. (a) LSV plots, (b) respective Tafel plots derived from (a), (c) durability and OER analysis of NiPd/OMC performed without urea. HER electrocatalytic properties of Pt@C, NiPd/OMC, Ni/OMC, Pd/OMC, and OMC. (d) LSV plots, (e) respective Tafel plots derived from (d), and durability and HER analysis of NiPd/OMC. (f) durability and HER analysis of NiPd/OMC performed without and with 0.33 M urea in N2 saturated 1 M KOH at scan rate = 5 mV s−1. All of reaction performed in N2 saturated 1 M KOH with 0.33 M urea at scan rate = 5 mV s−1. Reproduced and adapted from Ref. [200]; Copyright (2019), American Chemical Society.
Figure 23. Urea oxidation reaction (UOR) electrocatalytic properties of IrO@C, NiPd/OMC, Ni/OMC, Pd/OMC, and OMC. (a) LSV plots, (b) respective Tafel plots derived from (a), (c) durability and OER analysis of NiPd/OMC performed without urea. HER electrocatalytic properties of Pt@C, NiPd/OMC, Ni/OMC, Pd/OMC, and OMC. (d) LSV plots, (e) respective Tafel plots derived from (d), and durability and HER analysis of NiPd/OMC. (f) durability and HER analysis of NiPd/OMC performed without and with 0.33 M urea in N2 saturated 1 M KOH at scan rate = 5 mV s−1. All of reaction performed in N2 saturated 1 M KOH with 0.33 M urea at scan rate = 5 mV s−1. Reproduced and adapted from Ref. [200]; Copyright (2019), American Chemical Society.
Catalysts 10 00621 g023
Table 1. Catalytic performance in OER and HER activity [120].
Table 1. Catalytic performance in OER and HER activity [120].
NiVIr-LDHNiV-LDHNi(OH)2Ir/CPt/C
OEROverpotential (mV)203319337284-
Tafel slope (mA dec−1)55.3119.2214.186.8-
HEROverpotential (mV)41148---
Tafel slope (mA dec−1)35.9119.2--32.8
Both OER and HER activity for overpotential are performed at 10 mA cm−2.
Table 2. HER and OER activity in 1 M KOH electrolyte [170].
Table 2. HER and OER activity in 1 M KOH electrolyte [170].
Co/β-Mo2@N-CNTsCo@N-CNTsMo2C@N-CNTsPt/CIrO2
HERη10 (mV)17027545236-
Tafel plot (mV dec−1)92115171--
Current density (mA cm−2)0.140.0200.053--
Cdl (mF cm−2)2.11.30.7-
OERη10 (mV)356---377
Tafel plot (mV dec−1)6790188--
Cdl (mF cm−2)2.61.20.4--
Table 3. Characterization of each electrocatalysts.
Table 3. Characterization of each electrocatalysts.
NiPd/OMCNi/OMCPd/OMCBare-OMC
Ni (wt%)1010--
Pd (wt%)10-10-
Pore volume (cm3 g−1)0.911.171.261.45
Poer diameter (nm) a5.35.35.25.3
Surface area (m2 g−1) b862124310781434
a The pore diameters were calculated from desorption branches using BJH method. b The surface areas were obtained by the BET method.

Share and Cite

MDPI and ACS Style

Jang, S.; Moon, K.; Park, Y.; Park, S.; Park, K.H. Recent Studies on Multifunctional Electrocatalysts for Fuel Cell by Various Nanomaterials. Catalysts 2020, 10, 621. https://doi.org/10.3390/catal10060621

AMA Style

Jang S, Moon K, Park Y, Park S, Park KH. Recent Studies on Multifunctional Electrocatalysts for Fuel Cell by Various Nanomaterials. Catalysts. 2020; 10(6):621. https://doi.org/10.3390/catal10060621

Chicago/Turabian Style

Jang, Sanha, Kyeongmin Moon, Youchang Park, Sujung Park, and Kang Hyun Park. 2020. "Recent Studies on Multifunctional Electrocatalysts for Fuel Cell by Various Nanomaterials" Catalysts 10, no. 6: 621. https://doi.org/10.3390/catal10060621

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