Next Article in Journal
Phase Transition Field Effect Transistor Observed in an α-(BEDT-TTF)2I3 Single Crystal
Previous Article in Journal
On the RE2TiAl3 (RE = Y, Gd–Tm, Lu) Series—The First Aluminum Representatives of the Rhombohedral Mg2Ni3Si Type Structure
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Iron-Containing Nickel Cobalt Sulfides, Selenides, and Sulfoselenides as Active and Stable Electrocatalysts for the Oxygen Evolution Reaction in an Alkaline Solution

Institut für Anorganische Chemie und Strukturchemie, Heinrich-Heine-Universität Düsseldorf, 40204 Düsseldorf, Germany
*
Author to whom correspondence should be addressed.
Solids 2023, 4(3), 181-200; https://doi.org/10.3390/solids4030012
Submission received: 11 June 2023 / Revised: 4 July 2023 / Accepted: 12 July 2023 / Published: 16 July 2023

Abstract

:
Iron-containing nickel sulfides, selenides, and sulfoselenides were synthesized via a simple two-step hydrothermal reaction (temperature ≤ 160 °C) for their application as electrocatalysts in the oxygen evolution reaction (OER) in an alkaline solution (1 mol L−1 KOH). The study demonstrated that iron-containing nickel cobalt sulfides and selenides exhibit superior OER performance with lower overpotentials compared to iron-free nickel cobalt sulfide and selenide, which highlights the significant role of iron in enhancing OER nickel cobalt electrocatalysts: Fe0.1Ni1.4Co2.9(S0.87O0.13)4, η50 = 318 mV; Fe0.2Ni1.5Co2.8(S0.9O0.1)4, η50 = 310 mV; Fe0.3Ni1.2Co2.5(S0.9O0.1)4, η50 = 294 mV; Fe0.6Ni1.2Co2.5(S0.83O0.17)4, η50 = 294 mV; Fe0.4Ni0.7Co1.6(Se0.81O0.19)4, η50 = 306 mV compared to Ni1.0Co2.1(S0.9O0.1)4, η50 = 346 mV; and Ni0.7Co1.4(Se0.85O0.15)4, η50 = 355 mV (all values at current densities η50 of 50 mA cm−2). Furthermore, the iron-containing nickel cobalt sulfoselenide Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4 displayed exceptional OER performance with η50 = 277 mV, surpassing the benchmark RuO2 electrode with η50 = 299 mV. The superior performance of the sulfoselenide was attributed to its low charge transfer resistance (Rct) of 0.8 Ω at 1.5 V vs. the reversible hydrogen electrode (RHE). Moreover, the sulfoselenide demonstrated remarkable stability, with only a minimal increase in overpotential (η50) from 277 mV to 279 mV after a 20 h chronopotentiometry test. These findings suggest that trimetallic iron, nickel and cobalt sulfide, selenide, and especially sulfoselenide materials hold promise as high-performance, cost-effective, and durable electrocatalysts for sustainable OER reactions. This study provides a valuable approach for the development of efficient electrocatalytic materials, contributing to the advancement of renewable energy technologies.

1. Introduction

According to the US Energy Information Administration (EIA) report, the total world energy consumption will rise to 815 quadrillion Btu (British thermal units) in 2040, a 29% increase compared to 2020 [1], demonstrating the increasing demand for energy in the near future. Using clean and renewable energy is one of the main issues for societies. Green hydrogen (H2), produced with renewable energy, is seen as an alternative fuel and energy storage resource in the future [2]. Water splitting is one of the most studied ways to produce H2. With electrocatalysis, this process includes the hydrogen evolution reaction (HER) and the oxygen evolution reaction (OER) [3]. The anodic reaction (OER) involves a sluggish four-electron/four-proton-coupled transfer reaction. It is the main obstacle to an economic water-splitting process since it requires a much higher potential (1.6–2 V vs. the reversible hydrogen electrode, RHE) than the theoretical equilibrium potential of E° = 1.23 V vs. RHE [4,5,6,7,8].
The well-known benchmark materials for OER are the oxides of iridium and ruthenium (IrO2 and RuO2), while Pt-based materials are used as benchmarks for HER [9]. However, the high cost, scarcity, and low stability of these precious noble metals limit their practical large-scale application for water electrolysis [10,11]. Sustainable water splitting requires utilizing non-precious metals as a catalyst. Using non-precious metals with high abundance, durability, and catalytic activity, especially in OER, can effectively improve the scalability of electrocatalytic hydrogen production. Therefore, in recent years, there has been an impetus to develop high-performance, stable, and low-cost (non-noble) transition metal-based electrocatalysts such as transition metal sulfides [12], hydroxides, oxides, oxide-hydroxides [13,14,15,16], phosphides [17], nitrides [18], perovskites [19,20], and spinels [21] for OER.
Transition metal sulfides possess good conductivity and excellent mechanical and thermal stability, making them promising electrocatalysts for OER and HER and the oxygen reduction reaction (ORR) [22]. Moreover, mixed-metal sulfides, MMSs, show even higher electric conductivity and richer redox reactions through the synergistic effect of multi-transition metal ions, leading to a notable enhancement in electrocatalytic performance compared to monometallic sulfides [23,24].
Bimetallic nickel cobalt sulfide is one of the most studied and promising MMSs, and is extensively studied for electrocatalytic energy conversion and storage devices [25]. For example, the thiospinel NiCo2S4 with Ni(II) on tetrahedral (Td) sites and Co(III) ions on octahedral (Oh) sites receives much attention in many electrocatalysis applications, including water splitting, supercapacitors, and zinc-air batteries [26,27,28,29]. However, the electrocatalytic activity and stability of NiCo2S4 remains lower than IrO2 and RuO2 in OER and lower than Pt-based catalysts for HER [30]. To improve the electrocatalytic performance of nickel cobalt sulfides, researchers have applied many different strategies, including making composites with carbon materials [31,32,33], metal oxides [34,35], oxyhydroxides [15], layer double hydroxide [15], metal sulfides [36,37], incorporating other active metals such as Ru, Ir, and Pt in the structure [38,39] and also partially replacing sulfur ions with nitrogen [40], phosphorous [41] or selenium ions [42].
The improved electrocatalytic performance of nickel cobalt sulfide in the presence of anions such as Se and P can be attributed to effectively altering the surface electron density by modulating the compound’s d-band [42,43,44]. Increased electrocatalytic activity in the presence of other cations is mainly related to improving exposed active sites, reducing the charge transfer resistance, enhancing structural stability, and synergistic interactions between host and guest cations [45,46].
Although replacing anions or cations is widely used to upgrade the water-splitting performance of nickel cobalt sulfides, only a few reports applied a simultaneous cation and anion replacement to evaluate a possible synergistic effect. Recently, Deng et al. synthesized the polymetallic sulfoselenide, Co0.31Ni0.22Ru0.05S0.46Se0.41, which showed an overpotential of η10 = 261 mV (1.491 V vs. RHE) at 10 mA cm−2, while RuO2 needed η10 = 254 mV (1.484 V vs. RHE) under the same conditions [47]. However, ruthenium limits its classification as a completely non-precious metal electrocatalyst.
Here, we establish a facile two-step hydrothermal strategy to synthesize iron-containing nickel cobalt sulfides, selenides, and sulfoselenide as non-precious polymetallic compounds. A series of mixed-metal iron nickel cobalt carbonate hydroxide hydrates, (FexNi1−x)CoCH-(y), were synthesized and used as a precursor for sulfurization, selenization, and sulfoselenization.

2. Materials and Methods

2.1. Materials

All commercial chemicals were used as received without any purification: cobalt chloride hexahydrate CoCl2·6H2O (CAS: 7791-13-1 98% abcr GmbH, Karlsruhe, Germany), nickel chloride hexahydrate NiCl2·6H2O (CAS: 7791-20-0 98%, ACS reagent, Roth, Karlsruhe, Germany), ammonium iron(II) sulfate hexahydrate (NH4)2Fe(SO4)2·6H2O (CAS: 7783-85-9 BioUltra, 99% Sigma–Aldrich, St. Louis, MO, USA), sodium sulfide nonahydrate Na2S·9H2O (CAS: 1313-84-4 98% Acros Organics, New Dehli, India), selenium dioxide SeO2 (CAS: 7446-08-4 98%, Sigma–Aldrich, St. Louis, MO, USA), urea OC(NH2)2 (CAS: 57-13-6 ACS reagent 99.5%, Sigma–Aldrich, St. Louis, MO, USA), potassium hydroxide KOH (CAS: 1310-58-3,1N, Roth, Karlsruhe, Germany), N-methyl-2-pyrrolidone NMP (CAS: 872-32-2 95%, abcr, Karlsruhe, Germany), hydrazine monohydrate N2H4·H2O (CAS: 7803-57-8, Thermo Scientific, Kandel, Germany), polyvinylidene fluoride PVDF (CAS: 24937-79-9, Sigma–Aldrich, Karlsruhe, Germany), and carbon black Vulcan XC-72R (Fuelcellstore, Bryan, TX, USA). Nickel foam NF was purchased from Recemat BV, Cell Material Engineering, The Netherlands. Ultrapure water was produced using the Sartorius Arium mini device. Before using NF, it was cut into 1 × 2 cm2 pieces and cleaned with acetone (CAS: 67-64-1, ACS reagent, ≥99.5%, Sigma–Aldrich, St. Louis, MO, USA), hydrochloric acid (CAS: 7647-01-0, 1 mol L−1, Geel, Belgium), ultrapure water, and ethanol (CAS: 64-17-5, 98%, Th. Geyer, Renningen, Germany) for 15 min in an ultrasonic bath and then dried at 100 °C in a vacuum oven for 15 min.

2.2. Preparation of Nickel Cobalt Carbonate Hydroxide (NiCoCH) and Iron Nickel Cobalt Carbonate Hydroxide (FexNi1−x)CoCH-(y) Precursors

The NiCoCH sample was prepared according to the previously reported method by Chen et al. [20]. The amount of 950 mg (4.00 mmol) CoCl2·H2O, 475 mg (2.00 mmol) NiCl2·6H2O, and 1.1 g (18 mmol) urea was added to 40 mL of ultrapure water and stirred for 10 min. Then, the obtained solution was transferred into a Teflon-lined stainless-steel autoclave and heated to 120 °C for 6 h. The product was washed five times with ultrapure water (50 mL each) and two times with ethanol (25 mL each), then dried in a vacuum oven at 60 °C overnight. Yield = 860 mg. (FexNi1−x)CoCH-(y) precursors were synthesized by the same method except that the amounts of 98, 196, 294, or 392 mg (0.25, 0.50, 0.75, 1.0 mmol) of (NH4)2Fe(SO4)2·6H2O were added. Yields = 875 mg, 900 mg, 917 mg, and 930 mg, respectively.
Thereafter, (FexNi1−x)CoCH-(y) was used to refer to the iron-containing nickel cobalt carbonate hydroxide with y = 0.25, 0.50, 0.75, and 1.0 mmol of (NH4)2Fe(SO4)2·6H2O added to the reaction mixture; (FexNi1−x)CoCH-(y) refers to all samples.

2.3. Preparation of Iron Nickel Cobalt Sulfides, Selenide, and Sulfoselenide

Iron-containing nickel cobalt sulfide samples were prepared by hydrothermal sulfidation of the (FexNi1−x)CoCH-(y) precursors. A chosen amount of iron-containing precursor (125 mg) was dispersed in 40 mL of ultrapure water in an 80 mL Teflon-lined autoclave and stirred for 20 min. After that, 750 mg (3.125 mmol) of Na2S·9H2O was added. The resultant suspension was transferred to the oven, and the temperature was kept at 160 °C for 12 h. The obtained product was washed five times with ultrapure water (50 mL each) and two times with ethanol (20 mL each), then dried in a vacuum oven at 60 °C overnight. Yield was about 90 mg.
The iron-containing nickel cobalt sulfoselenide sample, Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4, was synthesized via sulfidation and selenization of (FexNi1−x)CoCH-(1.0) in one step. (FexNi1−x)CoCH-(1.0) (125 mg) was dispersed in ultrapure water and stirred for 20 min. Then, 680 mg (2.70 mmol) of Na2S·9H2O and 50 mg (0.425 mmol) of SeO2 were added to the suspension. Finally, 10 mL of N2H4·H2O was slowly added to the suspension. The Teflon-lined stainless-steel autoclave was kept at 160 °C for 12 h. The resulting powder was washed five times with ultrapure water (50 mL each time) and two times with ethanol (20 mL each time) and dried at 60 °C overnight. Yield was 100 mg.
Iron-containing nickel cobalt selenide, Fe0.4Ni0.7Co1.6(Se0.81O0.19)4 was synthesized by selenization of (FexNi1−x)CoCH-(1.0). (FexNi1−x)CoCH-(1.0) (125 mg) was dispersed in 30 mL of ultrapure water and stirred for 20 min followed by adding 100 mg (0.9 mmol) SeO2. Then, 10 mL of N2H4 was added to abovementioned suspension and stirred for another 10 min. The resulting suspension was transferred to a stainless autoclave and heated at 160 °C for 12 h. The obtained black powder was washed five times with ultrapure water (50 mL each) and three times with absolute ethanol (20 mL each), then dried at 60 °C in the vacuum oven overnight. Yield was 110 mg.
For comparison, nickel cobalt sulfide, Ni1.0Co2.1(S0.9O0.1)4 and nickel cobalt selenide, Ni0.7Co1.4(Se0.85O0.15)4, and were synthesized by sulfidation or selenization of NiCoCH (Supplementary Materials, Section S1). The schematic illustration of the synthesis is shown in Scheme 1.

2.4. Material Characterization

Powder X-ray diffraction (PXRD) analysis was conducted at ambient temperature on a Rigaku Miniflex 600 powder diffractometer (Rigaku, Tokyo, Japan) using Cu Kα1 radiation with λ = 1.5406 Å (40 kV, 15 mA, 600 W) and a flat silicon low background with a small indent in the range of 2θ = 5°–100°. The obtained PXRD data were evaluated with the Match v3.11 software.
Scanning electron microscopy (SEM) was performed with Jeol JSM-6510LV QSEM (Jeol, Akishima, Japan) advanced electron microscope (LaB6 cathode at 20 kV) equipped with a Bruker Xflash 410 silicon drift detector for energy-dispersive X-ray (EDX) spectroscopy.
Transmission electron microscopy (TEM) with energy-dispersive X-ray spectroscopy (TEM-EDX) was carried out with a FEI Tecnai G2 F20 electron microscope (FEI Co., Hillsboro, OR, USA) operated at 200 kV accelerating voltage.
A Quantachrome Autosorb-6 automatic adsorption analyzer (Quantachrome Gmbh; Odelzhausen, Germany) was used to determine nitrogen sorption isotherms for the BET surface area determination of the samples at 77 K. The samples were degassed at 90 °C in a 10−2 mbar vacuum for 15 h before the gas sorption measurement.
X-ray photoelectron spectroscopy (XPS) data were collected using a ULVAC-PHI VersaProbe II microfocus X-ray photoelectron spectrometer (ULVAC-PHI, Chigasaki, Japan). The spectra were recorded using a polychromatic aluminum Kα X-ray source (1486.8 eV) and referenced to the carbon 1s orbital with a binding energy of 284.8 eV.
Quantification of the metal content of the samples was performed using a Perkin-Elmer PinaAcle 900T atomic absorption spectrometer (Perkin Elmer LAS GmbH, Rodgau-Jügesheim, Germany) (sample preparation described in Supplementary Materials, Section S2).
The sulfur content was determined with a VarioMICRO CHNS microanalyzer (Elementar Analysensysteme, Langenselbold, Germany).

2.5. Electrochemical Measurements

All electrochemical analyses were conducted on a three-electrode cell using an Interface 1010E potentiostat from Gamry Instruments at ambient temperature. The reversible hydrogen electrode, RHE, and Pt foil were used as a reference and counter electrode. Coated nickel foam, NF, was the working electrode. A slurry containing 8:1:1 mass portions of the active materials (mixed-metal sulfide, selenide, or sulfoselenide material), carbon black, Vulcan XC-72R, and polyvinylidene fluoride, PVDF, respectively, in N-methyl-2-pyrrolidone, NMP was prepared and carefully dropped on a 1 cm2 area of the NF surface, and dried at 60 °C for 12 h in a vacuum oven to prepare the coated NF electrode. To prepare the slurry, a weighted mass of 5 mg of the mixed-metal sulfide, selenide, or sulfoselenide material was utilized. Additionally, the weight of the electrode before and after loading the slurry was measured. The slurry was accurately applied to a 1 cm2 area of the electrode surface. Throughout the OER, we ensured the presence of a 1 cm2 electrode immersed in the KOH solution. Before starting the electrochemical analysis, an N2 gas flow was passed through the electrolyte to remove dioxygen from the 1 mol L−1 KOH electrolyte.
Linear sweep voltammetry LSV measurement was applied to determine the catalytic performance of the coated NF electrode at a scan rate of 5 mV s−1. Before starting the LSV measurement, 20 cycles of cyclic voltammetry at a scan rate of 100 mV s−1 were run to reach a stable electrocatalytic performance. The potentials of the LSV polarization curves were corrected by iR compensation. The chronopotentiometry at the current density of 50 mA cm−2 for 20 h was used to evaluate the stability of the selected electrocatalyst in long-term performance. Moreover, to understand the electrode/electrolyte interface behavior, electrochemical impedance spectroscopy, EIS, was performed in the frequency range of 0.1–100 kHz at 1.5 V vs. RHE.

3. Result and Discussion

3.1. Synthesis and Analysis

Nickel cobalt carbonate hydroxide (NiCoCH) and iron-containing nickel cobalt carbonate hydroxides (FexNi1−x)CoCH-(y) were synthesized from CoCl2·6H2O, NiCl2·6H2O, (NH4)2Fe(SO4)2·6H2O and urea as precursors in hydrothermal reactions (Step 1 in Scheme 1). Four different amounts of (NH4)2Fe(SO4)2·6H2O were used to synthesize (FexNi1−x)CoCH-(y), while keeping the NiCl2·6H2O and CoCl2·6H2O amounts constant. The samples were named (FexNi1−x)CoCH-(0.025), (FexNi1−x)CoCH-(0.05), (FexNi1−x)CoCH-(0.075), and (FexNi1−x)CoCH-(0.1), representing the use of 0.025, 0.05, 0.075, and 1.0 mmol of the iron precursor. The sulfidation and selenization of the metal carbonate hydroxide precursors were achieved through a hydrothermal sulfidation process with Na2S·9H2O and a selenization process with SeO2 (Step 2 in Scheme 1). For the sulfoselenide sample, a mixture of Na2S·9H2O and SeO2 was employed in the hydrothermal reaction.
Two different methods were used to determine the chemical formula of the as-prepared samples, a combination of AAS for the metal and CHNS analysis for the sulfur content (method 1) and SEM-EDX (method 2) (Supplementary Materials, Tables S2–S6). Method 1 provides more precise atomic ratios of metal and S content in the samples than EDX. In EDX, the emitted X-rays give a 1–2 µm depth analysis but EDX as an X-ray spectroscopy experiences matrix effects and would need standards of similar composition as the sample for peak identification and accurate quantification. For the sulfoselenide and selenide sample, AAS for the metal content was combined with EDX for the Se content. Moreover, the chemical formulae obtained from AAS + CHNS + EDX were much closer to charge balance than the EDX-derived formulae. Based on the metal-to-sulfur ratios obtained from method 1 and the charge balance calculation of the samples, oxygen should also be incorporated into the structure of the samples. The presence of oxygen in the structure of samples was also proven from the EDX and XPS spectrum. Therefore, the chemical formulae are given here with their estimated oxygen content. The chemical formulas resulting from methods 1 and 2 are provided in Table S6. The measured SEM-EDX data of the samples are provided in Figures S1–S8.
The crystallinity of all sulfide samples was low, as evidenced by broad peaks of low intensity in the powder X-ray diffractograms, PXRDs (Figure 1a). The crystalline phases in the iron-containing nickel cobalt samples were verified as spinels by matching to the known diffractograms of NiCo2S4 (ICDD no. 43-1477) and Co3S4 (ICDD no. 75-1561) (Figure 1). The prominent diffraction peaks located at 26.8°, 31.5°, 38.1°, 50.4°, and 55.2° can be attributed to the (220), (311), (400), (511), and (440) planes of the spinel lattice [24]. In addition, in the PXRD patterns of NiCo2S4 and iron-containing nickel cobalt samples, the diffraction peaks at 29.9° and 52.1° can be attributed to the (311) and (440) planes, respectively, of Co9S8 (ICDD no. 73-1442). It should be mentioned that the presence of an Co9S8 impurity in NiCo2S4 is reported in much of the previous literature, including the work of Chen et al. [24], who first reported the formation of sea-urchin-like NiCo2S4 using metal carbonate hydroxide as a precursor [33,40,47,48,49,50,51].
The content of the crystalline Co9S8 impurity decreases with increasing iron content and is only barely visible in Fe0.3Ni1.2Co2.5(S0.9O0.1)4 and no longer visible in Fe0.6Ni1.2Co2.5(S0.83O0.17)4. The corresponding selenides are of higher crystallinity (Figure 1b). The prominent reflection peaks of NiCo2Se4 and Fe0.4Ni0.7Co1.6(Se0.81O0.19)4 match the simulation for NiCo2Se4 (ICDD no. 04-006-5241), where they correspond to the (002), (311), and (−313) crystal plane located at 33.3°, 44.9°, and 51.4°, respectively [52]. By incorporating both sulfur and selenium in the structure, the crystallinity of Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4 from the PXRD pattern became too low and no clear crystalline phase analysis was possible anymore. Such low crystallinity was also seen in the literature for nickel sulfoselenide, oxygen-containing cobalt sulfide, and nickel sulfide nanoparticles, which were reported with excellent OER properties [53,54,55]. Previous reports have demonstrated that incorporating or doping transition metal sulfides, such as nickel cobalt sulfides with iron, induces lattice strain which results in peak broadening [56]. A small shift towards lower angles of the experimental diffractograms of selenium-containing samples versus the simulated diffractograms of the sulfur-only analogues can be ascribed to a larger lattice spacing which is caused by the larger radius of selenium compared to sulfur [57]. A shift in the peak positions between experimental and simulated diffractograms is also obtained if the sample is not properly aligned with the goniometer axis, e.g., by placing too thick a sample on the sample holder in the Bragg–Brentano geometry [58].
The morphology of the as-prepared samples was studied using scanning electron microscopy, SEM. As shown in Figure 2a, Ni1.0Co2.1(S0.9O0.1)4 consists of needle-like structures combining parts with a sea urchin-like morphology. By increasing the iron content, the morphology became more sea urchin-like (Figure 2b–f). In the selenides and the sulfoselenide Fe0.6Ni1.2Co2.5(S0.83O0.17)4 sample, Figure 2e,f, agglomerations of needle-like primary particles can be seen. The SEM-energy dispersive X-ray (EDX) mappings (Supplementary Materials, Figure S9) support the AAS- and CHNS-based elemental analysis for the chemical formulae, and SEM-EDX was the analysis of choice to determine the selenium content.
To further investigate the sulfoselenide Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4, transition electron microscopy, TEM-EDX, was performed. The TEM images, Figure 3, confirmed the needle-like microstructure. EDX-mapping also proved the uniform presence of Fe, Ni, Co, S, and Se in the sulfoselenide sample (see Supplementary Materials, Table S1 for atom ratios).
Nitrogen adsorption–desorption isotherms, and specific surface area (BET) of the samples are given in Figure S10 and Table S7.
X-ray photoelectron spectroscopy (XPS) was conducted to determine the valence state of the elements in Ni1.0Co2.1(S0.9O0.1)4 and Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4. As shown in Figure S11, X-ray survey spectra indicate the existence of Ni, Co, and S in Ni1.0Co2.1(S0.9O0.1)4 and of Fe, Ni, Co, S, and Se in Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4. The high-resolution spectra of the metal atoms and of Se are shown in Figure 4 and those of S 2p and O1s of Ni1.0Co2.1(S0.9O0.1)4 in Figures S12a and S13a, respectively. The positions of the XPS peaks are summarized in Tables S8 and S9. It should be noted that the analysis depth of XPS is only 0.7–11 nm as the detected photoelectrons can only escape from such a thin surface layer of the sample.
The metal ratios and the ratios between different oxidation states of nickel and cobalt (Table 1) were calculated by integrating the fitted peak area for each metal valence state using the Ni 2p3/2 and Co 2p3/2 regions. In Ni1.0Co2.1(S0.9O0.1)4, Ni2+ and Co3+ are the prevalent oxidation states.
Based on element ratios obtained from AAS and CHNS analysis and also the metal valence states for nickel and cobalt in the Ni1.0Co2.1(S0.9O0.1)4 sample, the chemical formula can be given as ((Ni2+)0.72(Ni3+)0.28)1.0((Co2+)0.27(Co3+)0.72)2.1(S0.9O0.1)4 which is anion-cation charge-balanced within rounding errors.
In contrast to the Ni1.0Co2.1(S0.9O0.1)4 sample, Ni3+ and Co2+ are the dominant valence states in Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4, which might be one of the reasons for the improved OER performance of this catalyst. It is believed that Ni3+ can improve electrophilicity and oxygen adsorption, which can increase the amount of NiOOH active sites during the OER reaction [59].
The Fe 2p spectrum (Figure 4c) for Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4 represents two broad peaks at 712.5 eV and 725.0 eV, which can be assigned to Fe 2p3/2 and Fe 2p1/2, respectively [60]. Furthermore, two satellite peaks were detected at 717.5 and 734.5 eV [61,62]. The 2p3/2 spectrum range is 710 to 720 eV including the satellite peak, while the 2p1/2 spectrum range is 720−735 eV with the satellite peak. For iron, it should be noted that the Fe 2p spectral background is contributed to from the CoLMM and NiLMM Auger peaks, making an unequivocal deconvolution and peak assignment difficult [63,64]. The Se 3d XPS spectra of Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4 (Figure 4d) consist of two peaks centered at 55.0 and 57.0 eV, representing Se 3d5/2, and Se 3d3/2, respectively. The peak at 59.5 eV was attributed to the SeOX forming on the surface due to exposure to air, and to the overlap with the Co 3p signal [42]. The high-resolution spectrum of S 2p and O 1s of Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4 are shown in Figures S12b and S13b (Supplementary Materials), respectively. Based on the elemental ratios obtained from AAS, CHNS, and EDX(Se) analysis, and also metal valence states for iron, nickel, and cobalt in the Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4 sample, the chemical formula can be given as (Fe2+)0.5((Ni2+)0.1(Ni3+)0.9)1.0((Co2+)0.87(Co3+)0.13)2.0(S0.57Se0.25O0.18)4.

3.2. Oxygen Evolution Reaction Performance

The OER activity of the mixed-metal sulfides, selenides, and the sulfoselenide was evaluated by a three-electrode setup in 1.0 mol L−1 KOH solution. The working electrode was prepared according to a method used by Li et al. [33], a slurry containing 8:1:1 mass portions of the active materials (mixed-metal sulfide, selenide, or sulfoselenide material), carbon black, Vulcan XC-72R, and polyvinylidene fluoride, PVDF, respectively, in N-methyl-2-pyrrolidone, NMP (further details in experimental section). As is shown in Figure 5a, the polarization curves of the as-prepared electrodes revealed that the presence of iron could reduce the overpotential of the Ni1.0Co2.1(S0.9O0.1)4 parent compound. In the polarization curves, the peaks around 1.3–1.4 eV are attributed to the oxidation of Ni2+/Ni3+ [65].
The overpotential for Ni1.0Co2.1(S0.9O0.1)4 of 346 mV at 50 mA cm−2 was reduced with increasing iron content in Fe0.1Ni1.4Co2.9(S0.87O0.13)4, Fe0.2Ni1.5Co2.8(S0.9O0.1)4, Fe0.3Ni1.2Co2.5(S0.9O0.1)4, and Fe0.6Ni1.2Co2.5(S0.83O0.17)4 to 318, 310, 294, and 294 mV, respectively (Figure 5a,b). The electronic interaction between Fe, Ni, and Co in the iron-containing samples alters the electronic structure, making Ni2+ oxidation more difficult, resulting in a positive shift in the Ni2+/Ni3+ anodic peak at 1.3–1.4 eV [33,40]. The reduction in the OER overpotential in iron-containing samples can be attributed to reducing the charge transfer resistance through the synergistic electronic interaction between Fe, Co and Ni from a charge redistribution between active sites within the samples. Density functional theory (DFT) calculations in the literature traced the synergy to a decrease in the Gibbs free energy for the formation of a MOOH intermediate, which not only enhanced the intrinsic OER activity, but also significantly improved the intrinsic conductivity of iron-containing samples, greatly facilitating the charge transfer process [66].
In the next step, the effect of the coexistence of sulfur and selenium was investigated. The overpotential of Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4 (green line) of 277 mV at 50 mA cm−2 presents a 17 mV and 22 mV reduction compared to Fe0.6Ni1.2Co2.5(S0.83O0.17)4 (294 mV) and RuO2 (299 mV) (Figure 5c,d), indicating that the insertion of selenium improves the performance of the transition metal sulfide. Moreover, the OER performance of Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4 at 100 mA cm−2 only needed 299 mV while RuO2 required 341 mV overpotential, which demonstrates the excellent electrocatalytic activity of Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4 even under a high current density. The boost of the OER activity of Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4 can be attributed to the effect of selenide incorporation that reduces the energy barrier of the OER reaction, optimizes the electronic structure of active sites via modulating of the d-band of the compound, and also accelerates the kinetics of the reaction [67].
The Tafel slopes of the samples were obtained from LSV polarization curves at a scan rate of 5 mV s−1 using the Tafel Equation (1) [68]:
η = a + b × log(j)
η is the overpotential, b is the Tafel slope, j is the current density, and c is the intercept with the y-axis. The value of the Tafel slope is one of the most useful kinetic parameters and is inversely proportional to the kinetics of the OER reaction. Hence, as demonstrated in Figure 6a,b, Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4, with the lowest value of the Tafel slope, presents the most favorable OER kinetics among the investigated samples. Furthermore, according to Krasil’shchikov’s mechanistic paths (Equations (2)–(5), M = active site, b = Tafel slope) [68,69] with their corresponding Tafel slope values, the rate-determining step of the OER reaction for Ni1.0Co2.1(S0.9O0.1)4 (b = 125 mV dec−1) is metal oxidation with hydroxide formation (reaction (2)). By increasing the iron content in the samples, the Tafel slope decreased and reached 85 mV dec−1 for Fe0.6Ni1.2Co2.5(S0.83O0.17)4, suggesting that metal oxidation with hydroxide formation (2), and deprotonation of a metal hydroxide (3), both present rate-determining steps.
The lowest Tafel slope among the samples belongs to the sulfoselenide Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4 with 82 mV dec−1, which is again evidence for the role of selenium in enhancing the kinetics of the OER reaction. The Tafel value of 82 mV dec−1 of Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4 is close to reaction (3) (b = 60 mV dec−1), being, then, rate-determining in the overall OER process.
M + OH ⇆ MOH + e, b = 120 mV dec−1
MOH + OH ⇆ MO + H2O, b = 60 mV dec−1
MO → MO + e, b = 45 mV dec−1
2MO → 2M + O2, b = 19 mV dec−1
Electrochemical impedance spectroscopy (EIS) was conducted for the electron-transfer kinetics during the OER reaction and to justify the obtained overpotentials [70]. Figure 6c shows Nyquist plots of the samples at the potential of 1.5 V vs. RHE. The semicircle diameter of Nyquist plots is inversely proportional to the charge transfer rate across the electrode and the electrolyte that accelerates reaction kinetics. Hence, a smaller semicircle diameter represents more favorable charge transfer kinetics [27].
The smaller semicircle diameter of the Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4 Nyquist plot indicates that the coexistence of iron and selenium in Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4 can reduce the charge transfer resistance (Rct). Furthermore, a Voigt circuit model was applied to the Nyquist plots to evaluate the specific value for the charge resistance in the OER process. As shown in Table 2, the value of charge resistance for Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4 is the lowest (0.8 Ω) compared to the other investigated samples. The results from the Rct values are in line with the recorded OER overpotentials of the samples and the electrocatalytic performance.
To elucidate the importance of sulfur and selenium in the OER electrocatalysis performance, the (FexNi1−x)CoCH-(1.0) precursor for Fe0.6Ni1.2Co2.5(S0.83O0.17)4 and Fe0.5Ni1.0Co2.0(S0.57 Se0.25O0.18)4 was investigated and found to have an OER LSV polarization curve with a higher overpotential of 330 mV vs. RHE to reach 50 mA cm−2 and a Tafel slope of 98 mV dec−1, higher than the iron-containing sulfide and sulfoselenide (Table 2, Figure S14). Furthermore, the larger Nyquist plot semicircle diameter and higher charge transfer resistance (Rct = 2.5 Ω) of (FexNi1−x)CoCH-(1.0) imply the importance of S and Se in facilitating the charge transfer through the electrode–electrolyte interface in sulfide, selenides, and sulfoselenide samples (Figure S15).
One of the critical parameters to evaluate the performance of electrocatalysts in practical applications is their long-term stability. Hence, a chronopotentiometry test, at a current density of 50 mA cm−2 for 20 h, was conducted to evaluate the long-term stability performance of the Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4 [50].
Figure 6d illustrates that, after 20 h of OER reaction, the overpotential for obtaining 50 mA cm−2 only increased from 277 mV to 279 mV, which is essentially constant and supports the excellent electrocatalyst stability of Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4 for practical applications. Especially since for RuO2 the overpotential increased from 300 mV to 375 mV during the same chronopotentiometry test for 20 h, at 50 mA cm−2. A recent study by Jiang et al. [70,71] proved that a SeOx film on the surface could improve the catalyst stability in the OER reaction, which can be the reason behind the high stability of Fe0.5Ni1.0Co2.0(S0.57 Se0.25O0.18)4 during the OER reaction (the overpotential increased from 277 to only 279 at 50 mA cm−1 after 20 h OER reaction).
The superior OER activity and stability of Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4 material can be attributed to several effects resulting from introducing Fe and Se2− in the structure of the nickel cobalt sulfide base material; introducing Fe sites can enhance OER performance (a) by optimizing bond energies for OER intermediates adsorbing on the electrode surface, which facilitates the OER kinetics [72], (b) by overcoming the metal oxidation step and facilitating O2 evolution [73], or (c) by improving the conductivity of the electrode film [74]. However, it is known that the surface of electrodes containing iron, nickel, and cobalt is oxidized and amorphized in the course of oxygen evolution occurring at high positive electrode potentials [75,76,77].
For comparison, the overpotential values of several high performance electrocatalysts at a current density of 50 mA cm−2 using nickel foam as substrate are presented in Table 3. Notably, the results demonstrate that the OER performance of the sulfoselenide Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4 is on par with the best electrocatalysts reported in the literature, highlighting its comparable effectiveness.
During the course of our studies, we realized the number of parameters and that an orthogonal experimental design would be beneficial [89]. We have suggested a theoretical orthogonal experimental design for future work on iron-containing sulfides, selenides, and sulfoselenides (Section S8, Supplementary Materials).

4. Conclusions

A novel trimetallic sulfoselenide, Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4, electrocatalyst was synthesized via a two-step hydrothermal method. The mixed-metal sulfoselenide possessed higher OER activity than the bimetallic nickel cobalt sulfide Ni1.0Co2.1(S0.9O0.1)4, the bimetallic nickel cobalt selenide Ni0.7Co1.4(Se0.85O0.15)4, the trimetallic iron-containing nickel cobalt sulfides, Fe0.1Ni1.4Co2.9(S0.87O0.13)4, Fe0.2Ni1.5Co2.8(S0.9O0.1)4, Fe0.3Ni1.2Co2.5(S0.9O0.1)4, and Fe0.6Ni1.2Co2.5(S0.83O0.17)4, and the iron-containing nickel cobalt selenide, Fe0.4Ni0.7Co1.6(Se0.81O0.19)4. The trimetallic sulfoselenide required an overpotential of only 277 mV at 50 mV cm−2 and had favorable OER kinetics, manifested by a Tafel slope of 82 mV dec−1. The OER performance of Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4 surpassed the well-known RuO2 benchmark material where the required overpotential (300 mV to generate 50 mA cm−2) was 23 mV higher under the same condition (1 mol L−1 KOH). The 20 h chronopotentiometry analysis revealed that Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4 has remarkable stability during long-term operation in alkaline media (the overpotential increased only from 277 to 279 mV) while, under the same conditions, RuO2 showed a considerable loss in activity (such that the overpotential increasing from 300 to 375 mV during the generation of the 50 mA cm−2 current density).
Furthermore, the incorporation of selenium also significantly affected the OER activity and stability by reducing the energy barrier of the OER reaction, optimizing the electronic structure of active sites by modifying the d-band of the materials. Indeed, a SeOx film on the surface of Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4 can enhance the long-term stability of the catalyst. Considering the improvements in OER performance, this simple two-step hydrothermal synthesis of trimetallic sulfoselenides, from transition metal carbonate hydroxide, (FexNi1−x)CoCH-(y), as precursors can be used as a facile and practical approach to produce the next generation of non-precious polymetallic polychalcogenide materials for the oxygen evolution reaction.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/solids4030012/s1, Section S1: preparation of Ni0.7Co1.4(Se0.85O0.15)4 and Ni1.0Co2.1(S0.9O0.1)4; Section S2: Sample preparation for AAS; Section S3: Scanning electron microscopy and energy dispersive X-ray spectroscopy (SEM/EDX) and TEM/EDX; Section S4: Elemental analysis and atomic spectroscopy measurements; Section S5: Nitrogen sorption measurement; Section S6: X-ray photoelectron spectroscopy; Section S7: Electrochemical characterization; Section S8: Theoretical orthogonal experimental design. References [90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117,118,119,120,121,122,123,124,125,126,127,128] are cited in the Supplementary Materials.

Author Contributions

Conceptualization: S.A. and C.J.; methodology: S.A.; validation: S.A.; formal analysis: S.A., L.R., M.N.A.F. and T.H.Y.B.; writing—original draft preparation: S.A.; writing—review and editing: S.A. and C.J.; supervision: C.J.; project administration: C.J.; funding acquisition: C.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the joint National Natural Science Foundation of China–Deutsche Forschungsgemeinschaft (NSFC-DFG) project (DFG JA466/39-1).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

The authors also thank Annette Ricken for the AAS measurements.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Conti, J.; Holtberg, P.; Diefenderfer, J.; LaRose, A.; Turnure, J.T.; Westfall, L. International Energy Outlook 2016 with Projections to 2040; Report number: DOE/EIA--0484; U.S. Energy Information Administration, U.S. Department of Energy: Washington, DC, USA, 2016; Volume 20585, pp. 7–16.
  2. Wu, D.; Kusada, K.; Yoshioka, S.; Yamamoto, T.; Toriyama, T.; Matsumura, S.; Chen, Y.; Seo, O.; Kim, J.; Song, C.; et al. Efficient Overall Water Splitting in Acid with Anisotropic Metal Nanosheets. Nat. Commun. 2021, 12, 1145. [Google Scholar] [CrossRef]
  3. Huang, W.; Zhang, J.; Liu, D.; Xu, W.; Wang, Y.; Yao, J.; Tan, H.T.; Dinh, K.N.; Wu, C.; Kuang, M.; et al. Tuning the Electronic Structures of Multimetal Oxide Nanoplates to Realize Favorable Adsorption Energies of Oxygenated Intermediates. ACS Nano 2020, 14, 17640–17651. [Google Scholar] [CrossRef] [PubMed]
  4. Jiao, Y.; Zheng, Y.; Jaroniec, M.; Qiao, S.Z. Design of Electrocatalysts for Oxygen- and Hydrogen-Involving Energy Conversion Reactions. Chem. Soc. Rev. 2015, 44, 2060–2086. [Google Scholar] [CrossRef]
  5. Du, P.; Eisenberg, R. Catalysts Made of Earth-Abundant Elements (Co, Ni, Fe) for Water Splitting: Recent Progress and Future Challenges. Energy Environ. Sci. 2012, 5, 6012. [Google Scholar] [CrossRef]
  6. 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] [PubMed]
  7. Sondermann, L.; Jiang, W.; Shviro, M.; Spieß, A.; Woschko, D.; Rademacher, L.; Janiak, C. Nickel-Based Metal-Organic Frameworks as Electrocatalysts for the Oxygen Evolution Reaction (OER). Molecules 2022, 27, 1241. [Google Scholar] [CrossRef]
  8. Öztürk, S.; Moon, G.; Spieß, A.; Budiyanto, E.; Roitsch, S.; Tüysüz, H.; Janiak, C. A Highly-Efficient Oxygen Evolution Electrocatalyst Derived from a Metal-Organic Framework and Ketjenblack Carbon Material. ChemPlusChem 2021, 86, 1106–1115. [Google Scholar] [CrossRef] [PubMed]
  9. Wei, C.; Rao, R.R.; Peng, J.; Huang, B.; Stephens, I.E.L.; Risch, M.; Xu, Z.J.; Shao-Horn, Y. Recommended Practices and Benchmark Activity for Hydrogen and Oxygen Electrocatalysis in Water Splitting and Fuel Cells. Adv. Mater. 2019, 31, 1806296. [Google Scholar] [CrossRef]
  10. Kong, D.; Cha, J.J.; Wang, H.; Lee, H.R.; Cui, Y. First-Row Transition Metal Dichalcogenide Catalysts for Hydrogen Evolution Reaction. Energy Environ. Sci. 2013, 6, 3553. [Google Scholar] [CrossRef]
  11. Qu, M.; Jiang, Y.; Yang, M.; Liu, S.; Guo, Q.; Shen, W.; Li, M.; He, R. Regulating Electron Density of NiFe-P Nanosheets Electrocatalysts by a Trifle of Ru for High-Efficient Overall Water Splitting. Appl. Catal. B 2020, 263, 118324. [Google Scholar] [CrossRef]
  12. Guo, Y.; Park, T.; Yi, J.W.; Henzie, J.; Kim, J.; Wang, Z.; Jiang, B.; Bando, Y.; Sugahara, Y.; Tang, J.; et al. Nanoarchitectonics for Transition-Metal-Sulfide-Based Electrocatalysts for Water Splitting. Adv. Mater. 2019, 31, 1807134. [Google Scholar] [CrossRef] [PubMed]
  13. Jiang, F.; Choy, W.C.H.; Li, X.; Zhang, D.; Cheng, J. Post-Treatment-Free Solution-Processed Non-Stoichiometric NiOx Nanoparticles for Efficient Hole-Transport Layers of Organic Optoelectronic Devices. Adv. Mater. 2015, 27, 2930–2937. [Google Scholar] [CrossRef] [PubMed]
  14. Wang, N.; Cao, Z.; Kong, X.; Liang, J.; Zhang, Q.; Zheng, L.; Wei, C.; Chen, X.; Zhao, Y.; Cavallo, L.; et al. Activity Enhancement via Borate Incorporation into a NiFe (Oxy)Hydroxide Catalyst for Electrocatalytic Oxygen Evolution. J. Mater. Chem. A 2018, 6, 16959–16964. [Google Scholar] [CrossRef] [Green Version]
  15. Li, X.; Kou, Z.; Xi, S.; Zang, W.; Yang, T.; Zhang, L.; Wang, J. Porous NiCo2S4/FeOOH nanowire arrays with rich sulfide/hydroxide interfaces enable high OER activity. Nano Energy 2020, 78, 105230. [Google Scholar] [CrossRef]
  16. Subbaraman, R.; Tripkovic, D.; Chang, K.-C.; Strmcnik, D.; Paulikas, A.P.; Hirunsit, P.; Chan, M.; Greeley, J.; Stamenkovic, V.; Markovic, N.M. Trends in Activity for the Water Electrolyser Reactions on 3d M(Ni,Co,Fe,Mn) Hydr(Oxy)Oxide Catalysts. Nat. Mater. 2012, 11, 550–557. [Google Scholar] [CrossRef]
  17. Wang, C.; Chen, W.; Yuan, D.; Qian, S.; Cai, D.; Jiang, J.; Zhang, S. Tailoring the Nanostructure and Electronic Configuration of Metal Phosphides for Efficient Electrocatalytic Oxygen Evolution Reactions. Nano Energy 2020, 69, 104453. [Google Scholar] [CrossRef]
  18. Denny, S.R.; Tackett, B.M.; Tian, D.; Sasaki, K.; Chen, J.G. Exploring Electrocatalytic Stability and Activity of Unmodified and Platinum-Modified Tungsten and Niobium Nitrides. Int. J. Hydrogen Energy 2020, 45, 22883–22892. [Google Scholar] [CrossRef]
  19. Zhang, H.; Guan, D.; Gao, X.; Yu, J.; Chen, G.; Zhou, W.; Shao, Z. Morphology, crystal structure and electronic state one-step co-tuning strategy towards developing superior perovskite electrocatalysts for water oxidation. J. Mater. Chem. A 2019, 7, 19228–19233. [Google Scholar] [CrossRef]
  20. Zhang, H.; Guan, D.; Hu, Z.; Huang, Y.-C.; Wu, X.; Dai, J.; Dong, C.-L.; Xu, X.; Lin, H.-J.; Chen, C.-T.; et al. Exceptional lattice-oxygen participation on artificially controllable electrochemistry-induced crystalline-amorphous phase to boost oxygen-evolving performance. Appl. Catal. 2021, 297, 120484. [Google Scholar] [CrossRef]
  21. Ma, N.; Chen, G.; Zhu, Y.; Sun, H.; Dai, J.; Chu, H.; Ran, R.; Zhou, W.; Cai, R.; Shao, Z. A Self-Assembled Hetero-Structured Inverse-Spinel and Anti-Perovskite Nanocomposite for Ultrafast Water Oxidation. Small 2020, 16, 2002089. [Google Scholar] [CrossRef]
  22. Ibrahim, K.B.; Tsai, M.C.; Chala, S.A.; Berihun, M.K.; Kahsay, A.W.; Berhe, T.A.; Su, W.-N.; Hwang, B.-J. A review of transition metal-based bifunctional oxygen electrocatalysts. J. Chin. Chem. Soc. 2019, 66, 829–865. [Google Scholar] [CrossRef]
  23. Fu, G.; Lee, J.-M. Ternary Metal Sulfides for Electrocatalytic Energy Conversion. J. Mater. Chem. A 2019, 7, 9386–9405. [Google Scholar] [CrossRef]
  24. Chen, H.; Jiang, J.; Zhang, L.; Wan, H.; Qi, T.; Xia, D. Highly Conductive NiCo2S4 Urchin-like Nanostructures for High-Rate Pseudocapacitors. Nanoscale 2013, 5, 8879–8883. [Google Scholar] [CrossRef] [PubMed]
  25. Shen, L.; Yu, L.; Wu, H.B.; Yu, X.-Y.; Zhang, X.; Lou, X.W. Formation of Nickel Cobalt Sulfide Ball-in-Ball Hollow Spheres with Enhanced Electrochemical Pseudocapacitive Properties. Nat. Commun. 2015, 6, 6694. [Google Scholar] [CrossRef] [Green Version]
  26. Kulkarni, P.; Nataraj, S.K.; Balakrishna, R.G.; Nagaraju, D.H.; Reddy, M.V. Nanostructured Binary and Ternary Metal Sulfides: Synthesis Methods and Their Application in Energy Conversion and Storage Devices. J. Mater. Chem. A 2017, 5, 22040–22094. [Google Scholar] [CrossRef]
  27. Liu, W.; Zhang, J.; Bai, Z.; Jiang, G.; Li, M.; Feng, K.; Yang, L.; Ding, Y.; Yu, T.; Chen, Z.; et al. Controllable Urchin-Like NiCo2S4 Microsphere Synergized with Sulfur-Doped Graphene as Bifunctional Catalyst for Superior Rechargeable Zn-Air Battery. Adv. Funct. Mater. 2018, 28, 1706675. [Google Scholar] [CrossRef]
  28. Zhu, W.; Ren, M.; Hu, N.; Zhang, W.; Luo, Z.; Wang, R.; Wang, J.; Huang, L.; Suo, Y.; Wang, J. Traditional NiCo2S4 Phase with Porous Nanosheets Array Topology on Carbon Cloth: A Flexible, Versatile and Fabulous Electrocatalyst for Overall Water and Urea Electrolysis. ACS Sustain. Chem. Eng. 2018, 6, 5011–5020. [Google Scholar] [CrossRef]
  29. Feng, X.; Jiao, Q.; Liu, T.; Li, Q.; Yin, M.; Zhao, Y.; Li, H.; Feng, C.; Zhou, W. Facile Synthesis of Co9S8 Hollow Spheres as a High-Performance Electrocatalyst for the Oxygen Evolution Reaction. ACS Sustain. Chem. Eng. 2018, 6, 1863–1871. [Google Scholar] [CrossRef]
  30. Chauhan, M.; Reddy, K.P.; Gopinath, C.S.; Deka, S. Copper Cobalt Sulfide Nanosheets Realizing a Promising Electrocatalytic Oxygen Evolution Reaction. ACS Catal. 2017, 7, 5871–5879. [Google Scholar] [CrossRef] [Green Version]
  31. Liu, Q.; Jin, J.; Zhang, J. NiCo2S4@graphene as a Bifunctional Electrocatalyst for Oxygen Reduction and Evolution Reactions. ACS Appl. Mater. Interfaces 2013, 5, 5002–5008. [Google Scholar] [CrossRef]
  32. Xu, X.; Liu, X.; Zhong, W.; Zhang, L.; Liu, G.; Du, Y. Nanostructured NiCo2S4@NiCo2O4-Reduced Graphene Oxide as an Efficient Hydrogen Evolution Electrocatalyst in Alkaline Electrolyte. J. Colloid Interface Sci. 2021, 601, 570–580. [Google Scholar] [CrossRef]
  33. Li, H.; Chen, L.; Jin, P.; Li, Y.; Pang, J.; Hou, J.; Peng, S.; Wang, G.; Shi, Y. NiCo2S4 Microspheres Grown on N, S Co-Doped Reduced Graphene Oxide as an Efficient Bifunctional Electrocatalyst for Overall Water Splitting in Alkaline and Neutral PH. Nano Res. 2022, 15, 950–958. [Google Scholar] [CrossRef]
  34. He, B.; Song, J.-J.; Li, X.-Y.; Xu, C.-Y.; Li, Y.-B.; Tang, Y.-W.; Hao, Q.-L.; Liu, H.-K.; Su, Z. A nitrogen-doped NiCo2S4/CoO hollow multi-layered heterostructure microsphere for efficient oxygen evolution in Zn–air batteries. Nanoscale 2021, 13, 810–818. [Google Scholar] [CrossRef] [PubMed]
  35. Fereja, S.L.; Li, P.; Zhang, Z.; Guo, J.; Fang, Z.; Li, Z.; Chen, W. Construction of NiCo2S4/Fe2O3 Hybrid Nanostructure as a Highly Efficient Electrocatalyst for the Oxygen Evolution Reaction. Electrochim. Acta 2022, 405, 139793. [Google Scholar] [CrossRef]
  36. 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] [PubMed]
  37. Huang, Y.; Ge, S.; Chen, X.; Xiang, Z.; Zhang, X.; Zhang, R.; Cui, Y. Hierarchical FeCo2S4@FeNi2S4 Core/Shell Nanostructures on Ni Foam for High-Performance Supercapacitors. Chem. Eur. J. 2019, 25, 14117–14122. [Google Scholar] [CrossRef]
  38. Su, H.; Song, S.; Gao, Y.; Li, N.; Fu, Y.; Ge, L.; Song, W.; Liu, J.; Ma, T. In Situ Electronic Redistribution Tuning of NiCo2S4 Nanosheets for Enhanced Electrocatalysis. Adv. Funct. Mater. 2022, 32, 2109731. [Google Scholar] [CrossRef]
  39. Zhu, X.; Nguyen, D.C.; Prabhakaran, S.; Kim, D.H.; Kim, N.H.; Lee, J.H. Activating catalytic behavior of binary transition metal sulfide-shelled carbon nanotubes by iridium incorporation toward efficient overall water splitting. Mater. Today Nano 2023, 21, 100296. [Google Scholar] [CrossRef]
  40. Wu, Y.; Liu, X.; Han, D.; Song, X.; Shi, L.; Song, Y.; Niu, S.; Xie, Y.; Cai, J.; Wu, S.; et al. Electron Density Modulation of NiCo2S4 Nanowires by Nitrogen Incorporation for Highly Efficient Hydrogen Evolution Catalysis. Nat. Commun. 2018, 9, 1425. [Google Scholar] [CrossRef] [Green Version]
  41. Min, K.; Yoo, R.; Kim, S.; Kim, H.; Shim, S.E.; Lim, D.; Baeck, S.-H. Facile Synthesis of P-Doped NiCo2S4 Nanoneedles Supported on Ni Foam as Highly Efficient Electrocatalysts for Alkaline Oxygen Evolution Reaction. Electrochim. Acta 2021, 396, 139236. [Google Scholar] [CrossRef]
  42. Liang, T.; Lenus, S.; Liu, Y.; Chen, Y.; Sakthivel, T.; Chen, F.; Ma, F.; Dai, Z. Interface and M3+/M2+ Valence Dual-Engineering on Nickel Cobalt Sulfoselenide/Black Phosphorus Heterostructure for Efficient Water Splitting Electrocatalysis. Energy Environ. Mater. 2023, 6, e12332. [Google Scholar] [CrossRef]
  43. Gu, H.; Fan, W.; Liu, T. Phosphorus-doped NiCo2S4 nanocrystals grown on electrospun carbon nanofibers as ultra-efficient electrocatalysts for the hydrogen evolution reaction. Nanoscale Horiz. 2017, 2, 277–283. [Google Scholar] [CrossRef] [PubMed]
  44. Gong, Q.; Cheng, L.; Liu, C.; Zhang, M.; Feng, Q.; Ye, H.; Zeng, M.; Xie, L.; Liu, Z.; Li, Y. Ultrathin MoS2(1− x)Se2x Alloy Nanoflakes For Electrocatalytic Hydrogen Evolution Reaction. ACS Catal. 2015, 5, 2213–2219. [Google Scholar] [CrossRef]
  45. Zhou, Y.; Wang, Y.; Zhao, H.; Su, J.; Zhang, H.; Wang, Y. Investigation of Anion Doping Effect to Boost Overall Water Splitting. J. Catal. 2020, 381, 84–95. [Google Scholar] [CrossRef]
  46. Liu, R.; Xu, S.; Shao, X.; Wen, Y.; Shi, X.; Huang, L.; Hong, M.; Hu, J.; Yang, Z. Defect-Engineered NiCo-S Composite as a Bifunctional Electrode for High-Performance Supercapacitor and Electrocatalysis. ACS Appl. Mater. Interfaces 2021, 13, 47717–47727. [Google Scholar] [CrossRef]
  47. Deng, W.; Xie, W.; Li, D.; Gai, Y.; Chen, Z.; Yu, J.; Yang, R.; Bao, X.; Jiang, F. Controllable Tuning of Polymetallic Co-Ni-Ru-S-Se Ultrathin Nanosheets to Boost Electrocatalytic Oxygen Evolution. NPG Asia Mater. 2022, 14, 25. [Google Scholar] [CrossRef]
  48. Sarawutanukul, S.; Tomon, C.; Duangdangchote, S.; Phattharasupakun, N.; Sawangphruk, M. Rechargeable Photoactive Zn-Air Batteries Using NiCo2S4 as an Efficient Bifunctional Photocatalyst towards OER/ORR at the Cathode. Batter. Supercaps 2020, 3, 541–547. [Google Scholar] [CrossRef]
  49. Cai, D.; Wang, D.; Wang, C.; Liu, B.; Wang, L.; Liu, Y.; Li, Q.; Wang, T. Construction of Desirable NiCo2S4 Nanotube Arrays on Nickel Foam Substrate for Pseudocapacitors with Enhanced Performance. Electrochim. Acta 2015, 151, 35–41. [Google Scholar] [CrossRef]
  50. Song, Y.; Wang, Z.; Yan, Y.; Zhao, W.; Bakenov, Z. NiCo2S4 Nanoparticles Embedded in Nitrogen-Doped Carbon Nanotubes Networks as Effective Sulfur Carriers for Advanced Lithium–Sulfur Batteries. Microporous Mesoporous Mater. 2021, 316, 110924. [Google Scholar] [CrossRef]
  51. Huang, Z.; He, W.; Shen, H.; Han, G.; Wang, H.; Su, P.; Song, J.; Yang, Y. NiCo2S4 Microflowers as Peroxidase Mimic: A Multi-Functional Platform for Colorimetric Detection of Glucose and Evaluation of Antioxidant Behavior. Talanta 2021, 230, 122337. [Google Scholar] [CrossRef]
  52. Guo, Z.; Diao, Y.; Han, X.; Liu, Z.; Ni, Y.; Zhang, L. Mesoporous NiCo2Se4 Tube as an Efficient Electrode Material with Enhanced Performance for Asymmetric Supercapacitor Applications. CrystEngComm 2021, 23, 2099–2112. [Google Scholar] [CrossRef]
  53. Meng, A.; Yuan, X.; Shen, T.; Zhao, J.; Song, G.; Lin, Y.; Li, Z. Amorphous Nickel Sulfide Nanoparticles Anchored on N-Doped Graphene Nanotubes with Superior Properties for High-Performance Supercapacitors and Efficient Oxygen Evolution Reaction. Nanoscale 2020, 12, 4655–4666. [Google Scholar] [CrossRef] [PubMed]
  54. Cai, P.; Huang, J.; Chen, J.; Wen, Z. Oxygen-Containing Amorphous Cobalt Sulfide Porous Nanocubes as High-Activity Electrocatalysts for the Oxygen Evolution Reaction in an Alkaline/Neutral Medium. Angew. Chem. 2017, 129, 4936–4939. [Google Scholar] [CrossRef]
  55. Chen, N.; Du, Y.-X.; Zhang, G.; Lu, W.-T.; Cao, F.-F. Amorphous Nickel Sulfoselenide for Efficient Electrochemical Urea-Assisted Hydrogen Production in Alkaline Media. Nano Energy 2021, 81, 105605. [Google Scholar] [CrossRef]
  56. Guan, D.; Shi, C.; Xu, H.; Gu, Y.; Zhong, J.; Sha, Y.; Hu, Z.; Ni, M.; Shao, Z. Simultaneously mastering operando strain and reconstruction effects via phase-segregation strategy for enhanced oxygen-evolving electrocatalysis. J. Energy Chem. 2023, 82, 572–580. [Google Scholar] [CrossRef]
  57. Shi, Z.; Qi, X.; Zhang, Z.; Song, Y.; Zhang, J.; Guo, C.; Zhu, Z. Porous Cobalt Sulfide Selenium Nanorods for Electrochemical Hydrogen Evolution. ACS Omega 2021, 6, 23300–23310. [Google Scholar] [CrossRef]
  58. Pecharsky, V.K.; Zavalij, P.Y. Fundamentals of Powder Diffraction and Structural Characterization of Materials, 2nd ed.; ch. 12.1.5; Springer Science & Business Mediaz: New York, NY, USA, 2005; p. 313. Available online: https://link.springer.com/chapter/10.1007/0-387-24567-7_2 (accessed on 2 July 2023).
  59. Wang, H.-Y.; Hsu, Y.-Y.; Chen, R.; Chan, T.-S.; Chen, H.M.; Liu, B. Ni3+-Induced Formation of Active NiOOH on the Spinel Ni-Co Oxide Surface for Efficient Oxygen Evolution Reaction. Adv. Energy Mater. 2015, 5, 1500091. [Google Scholar] [CrossRef]
  60. Wang, Q.; Shang, L.; Shi, R.; Zhang, X.; Zhao, Y.; Waterhouse, G.I.N.; Wu, L.-Z.; Tung, C.-H.; Zhang, T. NiFe Layered Double Hydroxide Nanoparticles on Co,N-Codoped Carbon Nanoframes as Efficient Bifunctional Catalysts for Rechargeable Zinc-Air Batteries. Adv. Energy Mater. 2017, 7, 1700467. [Google Scholar] [CrossRef]
  61. Biesinger, M.C.; Payne, B.P.; Grosvenor, A.P.; Lau, L.W.M.; Gerson, A.R.; Smart, R.S.C. Resolving Surface Chemical States in XPS Analysis of First Row Transition Metals, Oxides and Hydroxides: Cr, Mn, Fe, Co and Ni. Appl. Surf. Sci. 2011, 257, 2717–2730. [Google Scholar] [CrossRef]
  62. Yamashita, T.; Hayes, P. Analysis of XPS Spectra of Fe2+ and Fe3+ Ions in Oxide Materials. Appl. Surf. Sci. 2008, 254, 2441–2449. [Google Scholar] [CrossRef]
  63. Yin, L.I.; Yellin, E.; Adler, I. X-Ray Excited LMM Auger Spectra of Copper, Nickel, and Iron. J. Appl. Phys. 1971, 42, 3595–3600. [Google Scholar] [CrossRef]
  64. He, Y.; Liu, X.; Chen, G.; Pan, J.; Yan, A.; Li, A.; Lu, X.; Tang, D.; Zhang, N.; Qiu, T.; et al. Synthesis of Co(II)-Fe(III) Hydroxide Nanocones with Mixed Octahedral/Tetrahedral Coordination toward Efficient Electrocatalysis. Chem. Mater. 2020, 32, 4232–4240. [Google Scholar] [CrossRef]
  65. Akbari, M.S.A.; Bagheri, R.; Song, Z.; Najafpour, M.M. Oxygen-evolution reaction by nickel/nickel oxide interface in the presence of ferrate(VI). Sci. Rep. 2020, 10, 8757. [Google Scholar] [CrossRef] [PubMed]
  66. Li, J.; Cui, H.; Du, X.; Zhang, X. The Controlled Synthesis of Nitrogen and Iron Co-Doped Ni3S2@NiP2 Heterostructures for the Oxygen Evolution Reaction and Urea Oxidation Reaction. Dalton Trans. 2022, 51, 2444–2451. [Google Scholar] [CrossRef] [PubMed]
  67. Huang, H.; Ning, S.; Xie, Y.; He, Z.; Teng, J.; Chen, Z.; Fan, Y.; Shi, J.; Barboiu, M.; Wang, D.; et al. Synergistic Modulation of Electronic Interaction to Enhance Intrinsic Activity and Conductivity of Fe–Co–Ni Hydroxide Nanotube for Highly Efficient Oxygen Evolution Electrocatalyst. Small 2023, 19, 2302272. [Google Scholar] [CrossRef]
  68. Yang, L.; Qin, H.; Dong, Z.; Wang, T.; Wang, G.; Jiao, L. Metallic S-CoTe with Surface Reconstruction Activated by Electrochemical Oxidation for Oxygen Evolution Catalysis. Small 2021, 17, 2102027. [Google Scholar] [CrossRef] [PubMed]
  69. Shinagawa, T.; Garcia-Esparza, A.T.; Takanabe, K. Insight on Tafel Slopes from a Microkinetic Analysis of Aqueous Electrocatalysis for Energy Conversion. Sci. Rep. 2015, 5, 13801. [Google Scholar] [CrossRef] [Green Version]
  70. Li, G.; Anderson, L.; Chen, Y.; Pan, M.; Abel Chuang, P.-Y. New Insights into Evaluating Catalyst Activity and Stability for Oxygen Evolution Reactions in Alkaline Media. Sustain. Energy Fuels. 2018, 2, 237–251. [Google Scholar] [CrossRef]
  71. Shang, X.; Chen, W.; Jiang, Z.-J.; Song, C.; Jiang, Z. In Situ Growth of SeOx Films on the Surface of Ni–Fe–Selenide Nanosheets as Highly Active and Stable Electrocatalysts for the Oxygen Evolution Reaction. Mater. Adv. 2022, 3, 2546–2557. [Google Scholar] [CrossRef]
  72. Zhang, K.; Zou, R. Advanced Transition Metal-Based OER Electrocatalysts: Current Status, Opportunities, and Challenges. Small 2021, 17, 2100129. [Google Scholar] [CrossRef]
  73. Anantharaj, S.; Kundu, S.; Noda, S. “The Fe Effect”: A Review Unveiling the Critical Roles of Fe in Enhancing OER Activity of Ni and Co Based Catalysts. Nano Energy 2021, 80, 105514. [Google Scholar] [CrossRef]
  74. Kuai, C.; Xi, C.; Hu, A.; Zhang, Y.; Xu, Z.; Nordlund, D.; Sun, C.-J.; Cadigan, C.A.; Richards, R.M.; Li, L.; et al. Revealing the Dynamics and Roles of Iron Incorporation in Nickel Hydroxide Water Oxidation Catalysts. J. Am. Chem. Soc. 2021, 143, 18519–18526. [Google Scholar] [CrossRef] [PubMed]
  75. Lopes, P.P.; Chung, D.Y.; Rui, X.; Zheng, H.; He, H.; Martins, P.F.B.D.; Strmcnik, D.; Stamenkovic, V.R.; Zapol, P.; Mitchell, J.F.; et al. Dynamically stable active sites from surface evolution of perovskite materials during the oxygen evolution reaction. J. Am. Chem. Soc. 2021, 143, 2741–2750. [Google Scholar] [CrossRef]
  76. Friebel, D.; Louie, M.W.; Bajdich, M.; Sanwald, K.E.; Cai, Y.; Wise, A.M.; Cheng, M.-J.; Sokaras, D.; Weng, T.-C.; Alonso-Mori, R.; et al. Identification of Highly Active Fe Sites in (Ni,Fe)OOH for Electrocatalytic Water Splitting. J. Am. Chem. Soc. 2015, 137, 1305–1313. [Google Scholar] [CrossRef] [Green Version]
  77. Sinitsyn, P.A.; Kuznetsov, V.V.; Filatova, E.A.; Levchenko, S.V. Ruddlesden–Popper Oxides LaSrM11−xM2xO4±δ (M1, M2—Fe, Co, Ni) Synthesized by the Spray-Pyrolysis Method as Promising Electrocatalysts for Oxygen Evolution Reaction. Energies 2022, 15, 8315. [Google Scholar] [CrossRef]
  78. Yang, M.; Lu, W.; Jin, R.; Liu, X.-C.; Song, S.; Xing, Y. Superior Oxygen Evolution Reaction Performance of Co3O4/NiCo2O4/Ni Foam Composite with Hierarchical Structure. ACS Sustain. Chem. Eng. 2019, 14, 12214–12221. [Google Scholar] [CrossRef]
  79. Chuah, X.F.; Hsieh, C.-T.; Huang, C.-L.; Raja, D.S.; Lin, H.-W.; Lu, S.-Y. In-Situ Grown, Passivator-Modulated Anodization Derived Synergistically Well-Mixed Ni–Fe Oxides from Ni Foam as High-Performance Oxygen Evolution Reaction Electrocatalyst. ACS Appl. Energy Mater. 2019, 2, 743–753. [Google Scholar] [CrossRef]
  80. Duan, J.-J.; Zhang, R.-L.; Feng, J.-J.; Zhang, L.; Zhang, Q.-L.; Wang, A.-J. Facile Synthesis of Nanoflower-like Phosphorus-Doped Ni3S2/CoFe2O4 Arrays on Nickel Foam as a Superior Electrocatalyst for Efficient Oxygen Evolution Reaction. J. Colloid Interface Sci. 2021, 581, 774–782. [Google Scholar] [CrossRef]
  81. Xue, Z.; Wang, Y.; Yang, M.; Wang, T.; Zhu, H.; Rui, Y.; Wu, S.; An, W. In-Situ Construction of Electrodeposited Polyaniline/Nickel-Iron Oxyhydroxide Stabilized on Nickel Foam for Efficient Oxygen Evolution Reaction at High Current Densities. Int. J. Hydrogen Energy. 2022, 47, 34025–34035. [Google Scholar] [CrossRef]
  82. Zhao, T.; Wang, Y.; Chen, X.; Li, Y.; Su, Z.; Zhao, C. Vertical Growth of Porous Perovskite Nanoarrays on Nickel Foam for Efficient Oxygen Evolution Reaction. ACS Sustain. Chem. Eng. 2020, 8, 4863–4870. [Google Scholar] [CrossRef]
  83. Gao, W.; Ma, F.; Wang, C.; Wen, D. Ce Dopant Significantly Promotes the Catalytic Activity of Ni Foam-Supported Ni3S2 Electrocatalyst for Alkaline Oxygen Evolution Reaction. J. Power Sources 2020, 450, 227654. [Google Scholar] [CrossRef]
  84. Kim, D.Y.; Lee, H.; Choi, S.R.; Choi, S.; An, W.Y.; Cho, H.-S.; Choi, M.; Park, J.-Y. Synthesis of Hierarchically Porous Ni Foam-Supported Heazlewoodite Ni3S2 Nanorod Electrocatalysts for Highly Efficient Oxygen Evolution Reaction. J. Alloys Compd. 2022, 914, 165305. [Google Scholar] [CrossRef]
  85. Tao, K.; Gong, Y.; Zhou, Q.; Lin, J. Nickel Sulfide Wrapped by Porous Cobalt Molybdate Nanosheet Arrays Grown on Ni Foam for Oxygen Evolution Reaction and Supercapacitor. Electrochim. Acta 2018, 286, 65–76. [Google Scholar] [CrossRef]
  86. Sun, Z.; Yuan, M.; Lin, L.; Yang, H.; Li, H.; Sun, G.; Yang, X.; Ma, S. Needle Grass-like Cobalt Hydrogen Phosphate on Ni Foam as an Effective and Stable Electrocatalyst for the Oxygen Evolution Reaction. Chem. Commun. 2019, 55, 9729–9732. [Google Scholar] [CrossRef]
  87. Li, X.; Han, G.-Q.; Liu, Y.-R.; Dong, B.; Hu, W.-H.; Shang, X.; Chai, Y.-M.; Liu, C.-G. NiSe@NiOOH Core–Shell Hyacinth-like Nanostructures on Nickel Foam Synthesized by in Situ Electrochemical Oxidation as an Efficient Electrocatalyst for the Oxygen Evolution Reaction. ACS Appl. Mater. Interfaces 2016, 8, 20057–20066. [Google Scholar] [CrossRef]
  88. Zhang, Y.; Guo, P.; Niu, S.; Wu, J.; Wang, W.; Song, B.; Wang, X.; Jiang, Z.; Xu, P. Magnetic Field Enhanced Electrocatalytic Oxygen Evolution of NiFe-LDH/Co3O4 P-n Heterojunction Supported on Nickel Foam. Small Methods 2022, 6, 2200084. [Google Scholar] [CrossRef]
  89. Ji, L.; Si, Y.; Liu, H.; Song, X.; Zhu, W.; Zhu, A. Application of orthogonal experimental design in synthesis of mesoporous bioactive glass. Microporous Mesoporous Mater. 2014, 184, 122–126. [Google Scholar] [CrossRef]
  90. Mohamed, S.G.; Hussain, I.; Shim, J.-J. One-Step Synthesis of Hollow C-NiCo2S4 Nanostructures for High-Performance Supercapacitor Electrodes. Nanoscale 2018, 10, 6620–6628. [Google Scholar] [CrossRef]
  91. Pu, J.; Cui, F.; Chu, S.; Wang, T.; Sheng, E.; Wang, Z. Preparation and Electrochemical Characterization of Hollow Hexagonal NiCo2S4 Nanoplates as Pseudocapacitor Materials. ACS Sustain. Chem. Eng. 2014, 2, 809–815. [Google Scholar] [CrossRef]
  92. Zou, J.; Xie, D.; Zhao, F.; Wu, H.; Niu, Y.; Li, Z.; Zou, Q.; Deng, F.; Zhang, Q.; Zeng, X. Microwave Rapid Synthesis of Nickel Cobalt Sulfides/CNTs Composites as Superior Cycling Ability Electrode Materials for Supercapacitors. J. Mater. Sci. 2021, 56, 1561–1576. [Google Scholar] [CrossRef]
  93. Bao, Y.; Zhang, W.; Yun, T.; Dai, J.; Li, G.; Mao, W.; Guan, M.; Zhuang, Y. The Application of Transition Metal Sulfide Ni3S4/CNFs in Rechargeable Ni–Zn Batteries. New J. Chem. 2021, 45, 22491–22496. [Google Scholar] [CrossRef]
  94. Kung, C.-W.; Chen, H.-W.; Lin, C.-Y.; Huang, K.-C.; Vittal, R.; Ho, K.-C. CoS Acicular Nanorod Arrays for the Counter Electrode of an Efficient Dye-Sensitized Solar Cell. ACS Nano 2012, 6, 7016–7025. [Google Scholar] [CrossRef]
  95. Khani, H.; Wipf, D.O. Iron Oxide Nanosheets and Pulse-Electrodeposited Ni–Co–S Nanoflake Arrays for High-Performance Charge Storage. ACS Appl. Mater. Interfaces 2017, 9, 6967–6978. [Google Scholar] [CrossRef]
  96. Matoba, M.; Anzai, S.; Fujimori, A. Thermal Expansion, Thermoelectric Power, and XPS Study of the Nonmetal-Metal Transition in Ni1−xS1−ySey. J. Phys. Soc. Jpn. 1991, 60, 4230–4244. [Google Scholar] [CrossRef]
  97. Shi, Z.-T.; Kang, W.; Xu, J.; Sun, L.-L.; Wu, C.; Wang, L.; Yu, Y.-Q.; Yu, D.Y.W.; Zhang, W.; Lee, C.-S. In Situ Carbon-Doped Mo(Se0.85S0.15)2 Hierarchical Nanotubes as Stable Anodes for High-Performance Sodium-Ion Batteries. Small 2015, 11, 5667–5674. [Google Scholar] [CrossRef]
  98. Danilson, M.; Altosaar, M.; Kauk, M.; Katerski, A.; Krustok, J.; Raudoja, J. XPS Study of CZTSSe Monograin Powders. Thin Solid Films 2011, 519, 7407–7411. [Google Scholar] [CrossRef]
  99. Zou, X.; Wu, Y.; Liu, Y.; Liu, D.; Li, W.; Gu, L.; Liu, H.; Wang, P.; Sun, L.; Zhang, Y. In Situ Generation of Bifunctional, Efficient Fe-Based Catalysts from Mackinawite Iron Sulfide for Water Splitting. Chem 2018, 4, 1139–1152. [Google Scholar] [CrossRef] [Green Version]
  100. Sahoo, M.K.; Samantara, A.K.; Behera, J.N. In Situ Transformed Cobalt Metal–Organic Framework Electrocatalysts for the Electrochemical Oxygen Evolution Reaction. Inorg. Chem. 2020, 59, 12252–12262. [Google Scholar] [CrossRef]
  101. Beamson, G.; Briggs, D. High Resolution XPS of Organic Polymers: The Scienta ESCA300 Database; Wiley: New York, NY, USA, 1992. [Google Scholar]
  102. Shen, Y.; Zhu, Y.; Sunarso, J.; Guan, D.; Liu, B.; Liu, H.; Zhou, W.; Shao, Z. New Phosphorus-Doped Perovskite Oxide as an Oxygen Reduction Reaction Electrocatalyst in an Alkaline Solution. Chem. Eur. J. 2018, 24, 6950–6957. [Google Scholar] [CrossRef]
  103. Yang, H.; Li, F.; Zhan, S. Intramolecular hydroxyl nucleophilic attack pathway by a polymeric water oxidation catalyst with single cobalt sites. Nat. Catal. 2022, 5, 414–429. [Google Scholar] [CrossRef]
  104. Feng, C.; Wang, F.; Liu, Z. A self-healing catalyst for electrocatalytic and photoelectrochemical oxygen evolution in highly alkaline conditions. Nat. Commun. 2021, 12, 5980. [Google Scholar] [CrossRef]
  105. Garcés-Pineda, F.A.; Blasco-Ahicart, M.; Nieto-Castro, D. Direct magnetic enhancement of electrocatalytic water oxidation in alkaline media. Nat. Energy 2019, 4, 519–525. [Google Scholar] [CrossRef] [Green Version]
  106. Zhang, S.L.; Guan, B.Y.; Lu, X.F.; Xi, S.; Du, Y.; Lou, X.W.D. Metal Atom-Doped Co3O4 Hierarchical Nanoplates for Electrocatalytic Oxygen Evolution. Adv. Mater. 2020, 32, 2002235. [Google Scholar] [CrossRef] [PubMed]
  107. Sundivich, J.; May, K.J.; Gasteiger, H.A.; Goodenough, G.B.; Horn, Y.S. A Perovskite Oxide Optimized for Oxygen Evolution Catalysis from Molecular Orbital Principles. Science 2011, 334, 1383–1385. [Google Scholar] [CrossRef]
  108. Shan, Z.; Archana, P.S.; Shen, G.; Gupta, A.; Bakker, M.G.; Pan, S. NanoCOT: Low-Cost Nanostructured Electrode Containing Carbon, Oxygen, and Titanium for Efficient Oxygen Evolution Reaction. J. Am. Chem. Soc. 2015, 137, 11996–12005. [Google Scholar] [CrossRef] [PubMed]
  109. Lee, S.; Ji, L.; De Palma, A.C. Scalable highly stable Si-based metal-insulator-semiconductor photoanodes for water oxidation fabricated using thin-film reactions and electrodeposition. Nat. Commun. 2021, 12, 3982. [Google Scholar] [CrossRef] [PubMed]
  110. Chen, R.; Zhang, Z.; Wang, Z.; Wu, W.; Du, S.; Zhu, W.; Lv, H.; Cheng, N. Constructing Air-Stable and Reconstruction-Inhibited Transition Metal Sulfide Catalysts via Tailoring Electron-Deficient Distribution for Water Oxidation. ACS Catal. 2022, 12, 13234–13246. [Google Scholar] [CrossRef]
  111. Shit, S.; Chhetri, S.; Jang, W.; Murmu, N.C.; Koo, H.; Samanta, P.; Kuila, T. Cobalt Sulfide/Nickel Sulfide Heterostructure Directly Grown on Nickel Foam: An Efficient and Durable Electrocatalyst for Overall Water Splitting Application. ACS Appl. Mater. Interfaces 2018, 10, 27712–27722. [Google Scholar] [CrossRef]
  112. Zhang, K.; Min, X.; Zhang, T.; Si, M.; Jiang, J.; Chai, L.; Shi, Y. Biodeposited Nano-CdS Drives the In Situ Growth of Highly Dispersed Sulfide Nanoparticles during Pyrolysis for Enhanced Oxygen Evolution Reaction. ACS Appl. Mater. Interfaces 2020, 12, 54553–54562. [Google Scholar] [CrossRef]
  113. Tiwari, A.; Yoon, Y.; Novak, T.G.; An, K.S.; Jeon, S. Continuous Network of Phase-Tuned Nickel Sulfide Nanostructures for Electrocatalytic Water Splitting. ACS Appl. Nano Mater. 2019, 2, 5061–5070. [Google Scholar] [CrossRef]
  114. Ji, Q.; Kong, Y.; Tan, H.; Duan, H.; Li, N.; Tang, B.; Wang, Y.; Feng, S.; Lv, L.; Wang, C.; et al. Operando Identification of Active Species and Intermediates on Sulfide Interfaced by Fe3O4 for Ultrastable Alkaline Oxygen Evolution at Large Current Density. ACS Catal. 2022, 12, 4318–4326. [Google Scholar] [CrossRef]
  115. Xie, M.; Ai, S.; Yang, J.; Yang, Y.; Chen, Y.; Jin, Y. In-Situ Generation of Oxide Nanowire Arrays from AgCuZn Alloy Sulfide with Enhanced Electrochemical Oxygen-Evolving Performance. ACS Appl. Mater. Interfaces 2015, 7, 17112–17121. [Google Scholar] [CrossRef]
  116. Shit, S.; Bolar, S.; Murmu, N.C.; Kuila, T. Design Principle of Monoclinic NiCo2Se4 and Co3Se4 Nanoparticles with Opposing Intrinsic and Geometric Electrocatalytic Activity toward the OER. ACS Sustain. Chem. Eng. 2019, 7, 18015–18026. [Google Scholar] [CrossRef]
  117. Swathi, S.; Yuvakkumar, R.; Ravi, G.; Thambidurai, M.; Nguyen, H.D.; Velauthapillai, D. Ternary Copper Iron Sulfide Microflowers Anchored on Reduced Graphene Oxide for Water Splitting. ACS Appl. Nano Mater. 2023, 6, 6538–6549. [Google Scholar] [CrossRef]
  118. Waqas, M.; Younis, J.; Awais, M.; Nazar, N.; Hussain, S.; Murtaza, S.; Yasmin, N.; Ashiq, M.N.; Mirza, M.; Safdar, M. CrS2-Modulated Enhanced Catalytic Properties of CdS/MoS2 Heterostructures Toward Photodegradation and Electrochemical OER Kinetics. Energy Fuels 2022, 36, 8391–8401. [Google Scholar] [CrossRef]
  119. Jayaramulu, K.; Masa, J.; Tomanec, O.; Peeters, D.; Ranc, V.; Schneemann, A.; Zboril, R.; Schuhmann, W.; Fischer, R.A. Nanoporous Nitrogen-Doped Graphene Oxide/Nickel Sulfide Composite Sheets Derived from a Metal-Organic Framework as an Efficient Electrocatalyst for Hydrogen and Oxygen Evolution. Adv. Funct. Mater. 2017, 27, 1700451. [Google Scholar] [CrossRef]
  120. Ganesan, P.; Sivananthama, A.; Shanmugam, S. Inexpensive electrochemical synthesis of nickel iron sulphides on nickel foam: Super active and ultra-durable electrocatalysts for alkaline electrolyte membrane water electrolysis. J. Mater. Chem. A 2016, 4, 16394–16402. [Google Scholar] [CrossRef] [Green Version]
  121. Gervas, C.; Khan, M.D.; Zhang, C.; Zhao, C.; Gupta, R.K.; Carleschi, E.; Doyle, B.P.; Revaprasadu, N. Effect of cationic disorder on the energy generation and energy storage applications of NixCo3−xS4 thiospinel. RSC Adv. 2018, 8, 24049–24058. [Google Scholar] [CrossRef]
  122. Wu, H.; Lu, Q.; Zhang, J. Thermal Shock-Activated Spontaneous Growing of Nanosheets for Overall Water Splitting. Nano-Micro Lett. 2020, 12, 162. [Google Scholar] [CrossRef] [PubMed]
  123. Wang, X.; Yang, Y.; Wang, R.; Li, L.; Zhao, X.; Zhang, W. Porous Ni3S2–Co9S8 Carbon Aerogels Derived from Carrageenan/NiCo-MOF Hydrogels as an Efficient Electrocatalyst for Oxygen Evolution in Rechargeable Zn–Air Batteries. Langmuir 2022, 38, 7280–7289. [Google Scholar] [CrossRef] [PubMed]
  124. Li, M.; Xu, Z.; Li, Y.; Wang, J.; Zhong, Q. In situ fabrication of cobalt/nickel sulfides nano hybrid based on various sulfur sources as highly efficient bifunctional electrocatalysts for overall water splitting. Nano Select 2022, 3, 147–156. [Google Scholar] [CrossRef]
  125. Guo, M.-L.; Wu, Z.-Y.; Zhang, M.-M.; Huang, Z.-J.; Zhang, K.-X.; Wang, B.-R.; Tu, J.-C. Coupling interface constructions of FeOOH/NiCo2S4 by microwave-assisted method for efficient oxygen evolution reaction. Rare Met. 2023, 42, 1847–1857. [Google Scholar] [CrossRef]
  126. Jin, C.; Hou, M.; Li, X.; Liu, D.; Qu, D.; Dong, Y.; Xie, Z.; Zhang, C. Rapid electrodeposition of Fe-doped nickel selenides on Ni foam as a bi-functional electrocatalyst for water splitting in alkaline solution. J. Electroanal. 2022, 906, 116014. [Google Scholar] [CrossRef]
  127. Guo, Y.; Zhou, X.; Tang, J.; Tanaka, S.; Kaneti, Y.V.; Na, J.; Jiang, B.; Yamauchi, Y.; Bando, Y.; Sugahara, Y. Multiscale structural optimization: Highly efficient hollow iron-doped metal sulfide heterostructures as bifunctional electrocatalysts for water splitting. Nano Energy 2020, 75, 104913. [Google Scholar] [CrossRef]
  128. Wang, C.-Y.; Dong, W.-D.; Zhou, M.-R.; Wang, L.; Hu, Z.-Y.; Chen, L.; Li, Y.; Su, B.-L. Gradient selenium-doping regulating interfacial charge transfer in zinc sulfide/carbon anode for stable lithium storage. J. Colloid Interface Sci. 2022, 619, 42–50. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Schematic illustration of the preparation process for iron-containing Ni-Co sulfide, sulfoselenide, and selenide. (FexNi1−x)CoCH-(y) represents the precursor iron-containing nickel cobalt carbonate hydroxide hydrate.
Scheme 1. Schematic illustration of the preparation process for iron-containing Ni-Co sulfide, sulfoselenide, and selenide. (FexNi1−x)CoCH-(y) represents the precursor iron-containing nickel cobalt carbonate hydroxide hydrate.
Solids 04 00012 sch001
Figure 1. PXRD patterns of (a) as-prepared sulfide and (b) selenide and sulfoselenide samples compared to simulated (sim.) patterns of different metal sulfides and selenides.
Figure 1. PXRD patterns of (a) as-prepared sulfide and (b) selenide and sulfoselenide samples compared to simulated (sim.) patterns of different metal sulfides and selenides.
Solids 04 00012 g001
Figure 2. SEM images of the as-prepared samples: (a) Ni1.0Co2.1(S0.9O0.1)4, (b) Fe0.1Ni1.4Co2.9(S0.87O0.13)4, (c) Fe0.2Ni1.5Co2.8(S0.9O0.1)4, (d) Fe0.3Ni1.2Co2.5(S0.9O0.1)4, (e) Fe0.6Ni1.2Co2.5(S0.83O0.17)4, (f) Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4, (g) Fe0.4Ni0.7Co1.6(Se0.81O0.19)4, and (h) Ni0.7Co1.4(Se0.85O0.15)4.
Figure 2. SEM images of the as-prepared samples: (a) Ni1.0Co2.1(S0.9O0.1)4, (b) Fe0.1Ni1.4Co2.9(S0.87O0.13)4, (c) Fe0.2Ni1.5Co2.8(S0.9O0.1)4, (d) Fe0.3Ni1.2Co2.5(S0.9O0.1)4, (e) Fe0.6Ni1.2Co2.5(S0.83O0.17)4, (f) Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4, (g) Fe0.4Ni0.7Co1.6(Se0.81O0.19)4, and (h) Ni0.7Co1.4(Se0.85O0.15)4.
Solids 04 00012 g002
Figure 3. TEM images (ae) of Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4 and (fj) HAADF STEM-EDX mapping results of Fe, Ni, Co, S, and Se, recorded from a nano needle section (e) of Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4.
Figure 3. TEM images (ae) of Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4 and (fj) HAADF STEM-EDX mapping results of Fe, Ni, Co, S, and Se, recorded from a nano needle section (e) of Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4.
Solids 04 00012 g003
Figure 4. High-resolution XPS spectra of (a) Ni 2p region and (b) Co 2p region of Ni1.0Co2.1(S0.9O0.1)4 and Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4, (c) Fe 2p region and (d) Se 3d region of Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4.
Figure 4. High-resolution XPS spectra of (a) Ni 2p region and (b) Co 2p region of Ni1.0Co2.1(S0.9O0.1)4 and Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4, (c) Fe 2p region and (d) Se 3d region of Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4.
Solids 04 00012 g004
Figure 5. (a) OER LSV polarization curves and (b) bar chart of corresponding overpotentials at 50 mA cm−2 of iron-containing sulfides. (c) OER LSV polarization curves and (d) bar chart of corresponding overpotentials at 50 mA cm−2 of different iron-containing sulfides, selenides, and sulfoselenide.
Figure 5. (a) OER LSV polarization curves and (b) bar chart of corresponding overpotentials at 50 mA cm−2 of iron-containing sulfides. (c) OER LSV polarization curves and (d) bar chart of corresponding overpotentials at 50 mA cm−2 of different iron-containing sulfides, selenides, and sulfoselenide.
Solids 04 00012 g005
Figure 6. (a) Tafel plots and (b) bar chart of Tafel slopes of the samples. (c) Nyquist plots of selected samples and Voigt circuit model. (d) Chronopotentiometry test of Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4 and RuO2.
Figure 6. (a) Tafel plots and (b) bar chart of Tafel slopes of the samples. (c) Nyquist plots of selected samples and Voigt circuit model. (d) Chronopotentiometry test of Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4 and RuO2.
Solids 04 00012 g006
Table 1. The metal ratios and the ratios between different oxidation states of nickel and cobalt in Ni1.0Co2.1(S0.9O0.1)4 and Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4.
Table 1. The metal ratios and the ratios between different oxidation states of nickel and cobalt in Ni1.0Co2.1(S0.9O0.1)4 and Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4.
Sample At% (a)Ni/Co (a) At% (b)Position
(eV)
M2+/M3+
Ni1.0Co2.1(S0.9O0.1)4 (c)Ni6.901/2Ni2+35.6853.32.59
Ni3+13.7856.0
Co13.90Co2+17.0780.30.37
Co3+45.9778.7
Fe/Ni/Co (a)
Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4 (d)Fe2.301.0/2.5/4.1
Ni5.70Ni2+3.3854.50.10
Ni3+32.9856.5
Co9.40Co2+46.7781.96.65
Co3+7.0779.1
(a) From XPS survey spectrum (Figure S11, Supplementary Materials). (b) Based on the Ni 2p3/2 and Co 2p3/2 regions in XPS. (c) Element composition from AAS-CHNS. (d) Element composition from AAS-CHNS-EDX(Se) (Tables S3–S6, Supplementary Materials).
Table 2. Overpotential at 50 mA cm−2, Tafel slopes, and estimated charge transfer resistance of selected samples at 1.5 V vs. RHE.
Table 2. Overpotential at 50 mA cm−2, Tafel slopes, and estimated charge transfer resistance of selected samples at 1.5 V vs. RHE.
SampleOverpotential
(mV)
Tafel Slope
(mV dec−1)
Charge Transfer Resistance Rct (Ω)
Ni1.0Co2.1(S0.9O0.1)43461251.8
Ni0.7Co1.4(Se0.85O0.15)355971.7
Fe0.6Ni1.2Co2.5(S0.83O0.17)4294852.2
Fe0.4Ni0.7Co1.6(Se0.81O0.19)43061021.4
Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)4277820.8
(FexNi1−x)CoCH-(1.0)330982.5
RuO2299661.2
Table 3. Comparison of nickel- and cobalt-based electrocatalysts for OER.
Table 3. Comparison of nickel- and cobalt-based electrocatalysts for OER.
CatalystOverpotential (mV)Current Density (mA cm−2)Electrode Substrate (a)Ref.
Fe0.5Ni1.0Co2.0(S0.57Se0.25O0.18)427750NFThis work
Co3O4/NiCo2O440750NF[78]
NiO/α-Fe2O324450NF[79]
P-Ni3S2/CoFe2O425450NF[79,80]
PANI (b)/NiFe–OH26050NF[81]
LaCoO342050NF[82]
Ce-doped Ni3S225750NF[83]
Porous Ni3S229150NF[84]
(Co1.2MoO4.21·3H2O)/Ni3S229050NF[85]
CoHPO4·H2O 35050NF[86]
NiSe@NiOOH30050NF[87]
P-containing NiCo2S430050NF[41]
NiFe-LDH (c)/Co3O427450NF[88]
(a) NF = nickel foam; as we used a nickel foam electrode and current densities of 50 mA cm−2, this comparative listing is restricted to the same conditions. (b) PANI = polyaniline. (c) LDH = layered double hydroxides.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Abdpour, S.; Rademacher, L.; Fetzer, M.N.A.; Beglau, T.H.Y.; Janiak, C. Iron-Containing Nickel Cobalt Sulfides, Selenides, and Sulfoselenides as Active and Stable Electrocatalysts for the Oxygen Evolution Reaction in an Alkaline Solution. Solids 2023, 4, 181-200. https://doi.org/10.3390/solids4030012

AMA Style

Abdpour S, Rademacher L, Fetzer MNA, Beglau THY, Janiak C. Iron-Containing Nickel Cobalt Sulfides, Selenides, and Sulfoselenides as Active and Stable Electrocatalysts for the Oxygen Evolution Reaction in an Alkaline Solution. Solids. 2023; 4(3):181-200. https://doi.org/10.3390/solids4030012

Chicago/Turabian Style

Abdpour, Soheil, Lars Rademacher, Marcus N. A. Fetzer, Thi Hai Yen Beglau, and Christoph Janiak. 2023. "Iron-Containing Nickel Cobalt Sulfides, Selenides, and Sulfoselenides as Active and Stable Electrocatalysts for the Oxygen Evolution Reaction in an Alkaline Solution" Solids 4, no. 3: 181-200. https://doi.org/10.3390/solids4030012

Article Metrics

Back to TopTop