Next Article in Journal
Anodized TiO2 Nanotubes Sensitized with Selenium Doped CdS Nanoparticles for Solar Water Splitting
Next Article in Special Issue
Application and Analysis of Liquid Organic Hydrogen Carrier (LOHC) Technology in Practical Projects
Previous Article in Journal
Classification of Highly Imbalanced Supervisory Control and Data Acquisition Data for Fault Detection of Wind Turbine Generators
Previous Article in Special Issue
Density Functional Theory Optimization of Cobalt- and Nitrogen-Doped Graphene Catalysts for Enhanced Oxygen Evolution Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Self-Supporting np-AlFeNiO Bifunctional Electrode Material for Electrochemical Water Splitting Prepared by Electrooxidation

1
School of Materials Science and Engineering, Tianjin University, Tianjin 300350, China
2
Tianjin Key Laboratory of Composite and Functional Materials, Tianjin 300350, China
3
College of Environmental Engineering, Xuzhou Institute of Technology, Xuzhou 221000, China
*
Authors to whom correspondence should be addressed.
Energies 2024, 17(7), 1591; https://doi.org/10.3390/en17071591
Submission received: 21 February 2024 / Revised: 18 March 2024 / Accepted: 22 March 2024 / Published: 26 March 2024
(This article belongs to the Special Issue New Trends and Research in Fuel Cells and Energy Conversion/Storage)

Abstract

:
Hydrogen production through water splitting is a promising path to develop renewable green energy. Effective, stable, and low-cost catalysts are the key to water splitting. In the present work, a series of self-supporting nanoporous alloys are prepared by using a dealloying process followed by electrooxidation. Among them, the np-AlFeNiO-4s sample exhibits remarkable activity (10 mA cm−2 at 32 mV for the HER and 278 mV for the OER) and good long-term stability (100 h) in alkaline conditions for both the HER and the OER. It only requires 1.56 V to reach 10 mA cm−2 current density for total water splitting performance. The very short time of electrooxidation can significantly improve the HER performance. Electrooxidation makes the metal and metal oxide sites on the electrode surface effectively coupled, which greatly enhances the kinetic rate of the Volmer and Heyrovsky steps. Appropriate electrooxidation is a rapid and easy way to improve the activity of the electrocatalyst, which has a broad application prospect in electrochemical water splitting.

Graphical Abstract

1. Introduction

Nowadays, the consumption of huge amounts of traditional fossil fuels has led to the global energy shortage and serious environmental problems [1]. Hydrogen energy is considered one of the most promising energies of the future due to its cleanliness, high efficiency, and renewability [2]. Hydrogen can be achieved by steam methane reforming, gasification, and electrochemical water splitting. Among them, electrochemical water splitting has attracted much attention because of its low pollution and high efficiency in converting energy [3]. Currently, noble metals are the best catalysts for electrochemical water splitting, such as Pt/C for the hydrogen evolution reaction (HER) and RuO2 for the oxygen revolution reaction (OER). However, the low abundance and high cost of noble metals limit their large-scale application [4]. Recently, many researchers have focused on transition metal catalysts for electrochemical water splitting. Due to their valence bonding structure of empty or semi-filled orbitals, transition metal elements are prone to exhibit electrocatalytic synergy with other elements, thus effectively improving their catalytic activity [5,6]. Many transition metallic compounds, such as oxides, sulfides, and phosphides, were developed as water-splitting catalysts. These nanoscale compounds have unique geometrical and electronic properties and further exhibit good electrochemical catalytic properties [7,8,9].
Among them, the transition metal Ni and its oxides exhibit outstanding HER catalytic activity [10]. Ni has a very high ability to adsorb H*, which plays a crucial role in the Volmer step by dissociating H* from water molecules attached to the surface of the catalytic electrode. NiO has a strong adsorption capacity for water, which is very beneficial for the Heyrovsky step. Hence, the combination of Ni and NiO would effectively promote the chain reaction of the HER and increase the reaction rate [11,12,13]. Peng et al. used a process of electrooxidation on porous Ni to prepare Ni/NiO, which exhibits outstanding catalytic performance with an overpotential of 25 mV at the current density of 10 mA cm−2. The synergic effect of NiO and Ni improves the kinetics of Volmer and Heyrovsky steps and the nanoporous structure provides more active sites and facilitates the transfer of substances and charges [14]. This effect can also be extended to other metals and their oxides [15,16,17,18].
The incorporation of Fe into Ni-based catalysts can promote intermetallic synergy. Ni-Fe-based alloys show potential as bifunctional catalysts for both the HER and the OER [19,20,21]. Their oxides also play a great role in improving the water-splitting process. Suryanto et al. synthesized the Janus Ni-Fe nanoparticle with Ni bonded to γ-Fe2O3. The Ni-Fe bimetallic structure surface offers effective active sites for the OER and the HER. Ni provides metallic-type electron conduction and favors electron transport towards the active HER sites, and Fe oxide provides more M-OOH intermediates [22]. In addition, the introduction of Fe can change the surface electronic state of other metal elements and make the oxidation process more difficult, which can facilitate the kinetic steps of the OER. The presence of Fe also can reduce the interfacial charge transfer resistance and promote HER kinetics [23]. Al incorporation into transition metals also enhances the HER activity of transition metals. It can provide electrons to transition metal active sites and promote the reaction process. Meanwhile, Al doping also changes the adsorption energy on the electrode surface, which in turn changes the electrocatalytic reaction energy [24]. Zhou et al. designed a nanoporous NiO/Al3Ni2 catalyst [25]. The presence of Al3Ni2 provided good electrical conductivity for the electrode and also adsorbed more Had. Together with the good synergistic catalytic ability of Ni and NiO, the nanoporous NiO/Al3Ni2 catalyst exhibited excellent HER performance.
Al-Fe-Ni-based materials and their oxides show great potential for electrochemical water splitting. However, more research is still needed on how to integrate metals and their oxides more efficiently and utilize their synergistic effects to construct electrocatalysts with excellent performance. Meanwhile, most of the current catalysts for water splitting are in powder form and must be loaded on conductive support for the reaction, which could damage the catalytic activity and long-term stability and also bring additional costs for large-scale commercial applications [26,27].
Herein, we report a self-supporting nanoporous transition metal bifunctional electrode material for electrochemical water splitting. This high-performance material utilizes a very fast and simple preparation method to effectively couple transition metals and their oxides. Due to the excellent hydrogen adsorption ability of the transition metals and the water adsorption ability of their oxides, the overpotential of the electrode for the HER is greatly reduced. In addition, the transition metal oxides provide a large number of intermediates for the OER to accelerate the OER rate.

2. Materials and Methods

2.1. Material Preparation

In this work, the melt spinning method was used to prepare the precursor alloy ribbons. The synthesis process is shown in Figure 1. The pure metals Al, Ni, and Fe were mixed in the ratio of 70:15:15 under argon atmosphere and melted in a high vacuum arc-melting furnace. The alloy ingot was then remelted in a quartz tube and ejected onto the surface of a rotating copper wheel at a linear speed of 32 m s−1. The thickness and width of the as-formed alloy ribbons were about 30 μm and 1.5 mm, respectively. The nanoporous ribbons were prepared by the potential-static electrochemical dealloying method using a standard three-electrode system with an Ag/AgCl electrode and a Pt mesh electrode as the reference and counter electrodes. At room temperature, the Al70Fe15Ni15 precursor ribbons were dealloyed in 1 M KOH at a potential of −0.9 V (vs. Ag/AgCl) for 4000 s. Al can be corroded in KOH solution. After the dealloying process, the active Al metal was etched, and the material turned into a nanoporous structure. Then, the as-dealloyed ribbons were treated by electrooxidation at the applied potential of 0.77 V (vs. Ag/AgCl) for various amounts of time. Applying a higher potential to the material makes it more susceptible to oxidation, which in turn leads to the effective coupling of the metal and the oxide. The ribbons after electrooxidation were labeled np-AlFeNiO-Xs, where X is the seconds of electrooxidation.
In order to make a comparison, a commercial Pt/C electrode was prepared as a control sample. First, 4 mg of Pt/C powder was mixed with 800 μL of deionized water, 200 μL of ethanol, and 50 μL of Nafion, and then the ink was sonicated for 30 min. A glassy carbon electrode with an area of 0.2826 cm−2 was used to load 50 μL of the ink for the subsequent tests.

2.2. Characterization

The phase composition was characterized by X-ray diffraction (XRD, Bruker D8, Billerica, MA, USA) with Cu Kα as the radiation source. The morphology was determined by scanning electron microscopy (SEM, Hitachi S-4800, Tokyo, Japan) and transmission electron microscopy (TEM, JEOL 2100M, Tokyo, Japan). X-ray photoelectron spectroscopy (XPS, PHI 1600ECSA, Chigasaki, Japan) was used to identify the elemental composition and chemical state on the surface.

2.3. Electrochemical Measurements

All electrochemical tests were performed using a Gamry interface 1000 electrochemical workstation with a classical three-electrode system. The tests were carried out in 1 M KOH solution with the as-prepared catalysts, Pt mesh, and Ag/AgCl electrode as working electrode, counter electrode, and reference electrode, respectively. Linear Sweep Voltammetry (LSV) was performed with a fast sweep of 50 curves at a scan rate of 50 mV s−1 followed by a test at a scan rate of 2 mV s−1. The polarization curve at the 2 mV s−1 scan rate was used to compare the electrocatalytic performance of the as-prepared samples. Each curve was delayed by 5 s to prevent large changes in voltage from affecting the results. The electrode material is activated by fast sweeping, but the high speed of sweeping also makes the reaction speed of the material unable to keep up with the polarization speed, which leads to the deviation of the test results. Therefore, after fast sweeping to make the material activation complete, the test results of the slow sweeping speed were selected for comparison and analysis. All measured potentials in this work were converted to potentials versus reversible hydrogen electrode (RHE) using the following equation:
E RHE = E Ag / AgCl + 0.199 + 0.059 pH
Electrochemical impedance spectra (EIS) measurements were performed at an overpotential of 50 mV with a frequency range of 0.1 Hz to 100 kHz. In addition, the polarization curves were iR-compensated according to the following equation:
E corr = E mea   i R s
where Ecorr, Emea, i, and Rs are the iR-compensated potential, the measured potential, the current density, and the equivalent series resistance measured by EIS. Cyclic voltammetry (CV) measurements were performed at potentials from −0.05 to 0.05 V versus open circuit potential (OCP) at different scan rates from 10 to 200 mV s−1. The electrochemical active area was determined by the double-layer capacitance (Cdl), which was calculated from the linear slope of ΔJ = (Jf − Jr)/2 at 0 V (vs OCP). The ECSA of the catalyst was calculated by the following equation:
ECSA = C dl C S
where CS is the ECSA constant, and the CS of a typical material is 40 µF cm−2 [28]. The turnover frequency (TOF) value was calculated by the following equation [29]:
TOF = jA / F ( 2 n )
where j is the current density at a potential of 100 mV, A is the surface area of the working electrode, F is Faraday’s constant (96,485 c mol−1), and n is the number of moles of active substance loaded on the electrode. Stability was tested under a static current density of 10 mA cm−2 in 1 M KOH for 100 h.

3. Results and Discussion

Figure 2 shows the XRD patterns of the samples. The precursor alloy ribbon is composed of Al5FeNi, Al4Ni3, Fe, Al2O3, and NiO phases. The oxides should be generated due to the exposure of the ribbon in air. The peak intensities of the Al4Ni3 phase and the Al5FeNi phase of the as-dealloyed sample decrease, indicating that these phases would be mainly removed during the dealloying process. The intensity of oxide peak has little change in the electro-oxidized samples, indicating the formation of tiny amount of oxides. The peaks of the Al5FeNi, Fe, and Al4Ni3 phases are weakened, which indicates that they are the components that mainly undergo the oxidation reaction to produce oxides during the electrooxidation process.
Figure 3 shows the SEM images of the samples. Table S1 lists the EDX results of the samples. Figure S1 shows the cross-section of the dealloyed sample. After 4000 s of etching, Al in the samples was basically removed, and the samples were completely corroded. The corroded samples were mainly composed of nanosheets with a thickness of about 20 nm. After oxidation for 4 s, there was no significant change in the sample morphology, and only a small number of tiny spheres appeared on the nanosheets. With the increase in oxidation time, the spheres further increased. When the oxidation time of the sample reached 100 s, the nanosheets were coarsened significantly, and the thickness of the nanosheets grew to about 40 nm. The TEM images of np-AlFeNiO-4s are shown in Figure S2. The sample shows a porous morphology with a pore size of about 20 nm. Figure S2b shows a lattice analysis of this part. The lattice fringes of d(111) = 0.24 nm and d(200) = 0.22 nm with a 54° angle match Fe0.942O. There is no obvious diffraction peak of Fe0.942O in the XRD, which may be due to the low level of oxidation of the sample. It is presumed that after electrooxidation, the spheres grown in the sample nanosheets indicate the oxide spheres of Fe.
Figure 4 shows the XPS spectra of samples with different oxidation times. As shown in Figure 4a–c, the surface of the as-dealloyed sample mainly consists of the oxidized form of the metals, which should result from the charge transfer and oxidation during the dealloying process [25,30]. In Figure 4a, the Al on the surface of the np-AlFeNiO-0s sample mainly exists in the Al3+ state. However, there is still a small amount of Al0 located at 72.8 eV, which would be attributed to the intermetallic compounds AlFeNi and Al4Ni3 [31]. The Al0 peak disappears after electrooxidation, suggesting that the surface Al atoms in the intermetallic compounds should be oxidized. Figure 4b exhibits the 2p orbit peaks of Ni. Ni exists in the Ni2+ form with a small amount of Ni0 [25]. After electrooxidation, the peak of Ni0 at 852.4 eV reduces but does not completely disappear. Figure 4c shows the 2p peaks of Fe. After dealloying, Fe has two valence peaks including an Fe0 peak at 718.3 eV and oxidation state peaks at 710.4 eV and 724.5 eV [32]. After electrooxidation, the peak at 718.3 eV almost disappears, indicating that Fe would be oxidized. Figure 4d shows the O 1s profiles of the samples. For np-AlFeNiO-0s, O in the H2O molecule (M–OH) located at 531.7 eV and divalent lattice O (M–O) located at 531.0 eV can be detected. The O at 531.7 eV represents the H2O molecule attached to the surface of the sample, which would participate in the Volmer step in the electrolysis water reaction process [33,34,35,36]. The ability of the sample to adsorb H2O molecules influences the catalytic performance of the water electrolysis reaction. After electrooxidation, the peak strength of the M–OH bond at 531.7 eV becomes stronger, indicating that the oxide on the surface could improve the adsorption ability for H2O, which would greatly enhance the reaction rate of the Volmer step. In the alkaline electrolytic water reaction, the Volmer step is often a rate-limiting step [37]. Meanwhile, after electrooxidation, the M–OOH bond at 533.1 eV appears on the surface, which is an important intermediate in the OER [31].
Figure 5 shows the results of the HER properties. From Figure 5a, it can be seen that with the increase in electrooxidation time, the overpotential at −10 mA cm−2 shows a trend of first decreasing and then increasing. The overpotential data are listed in Table S2. The np-AlFeNiO-4s sample has the smallest overpotential of 32 mV vs. the RHE at 10 mA cm−2, which shows a significant reduction by about 100 mV compared with the unoxidized sample. Figure 5b shows the Nyquist plots. The impedance data are fitted and analyzed by an equivalent electrical circuit (inset of Figure 5b), which consists of an electrolyte solution resistance (Rs), interfacial capacitance (Cint), and charge transfer impedance (Rct). With an increase in electrooxidation time, the Rct of the nanoporous samples shows a tendency of first decreasing and then increasing. A certain degree of oxidation could increase the reaction sites and promote the reaction rate. However, due to the poor conductivity of the oxides, too many oxides would reduce the charge transfer rate, further damaging the catalytic performance. Figure 5c shows the Tafel slopes of the samples with different oxidation times. The Tafel slope reflects the hindering of the polarization process, which is determined by the rate-limiting step of the catalytic reaction [38]. The Tafel slope values for the catalysts at each oxidation time are listed in Table S2. The Tafel slope of the unoxidized sample is 141 mV dec−1, which is close to 120 mV dec−1, indicating that its reaction rate-limiting step in the HER should be the Volmer step. It is suggested that the adsorption capacity of water on the surface of the catalyst would significantly influence the HER rate. When the Tafel slope is between 40 mV dec−1 and 120 mV dec−1, the HER rate-limiting step is the Heyrovsky step [39]. After oxidation, the Tafel slope decreases to below 120 mV dec−1. The strong adsorption ability of the oxides to water would accelerate the Volmer step, resulting in a decrease in the Tafel slope, and the rate-limiting step would turn into the Heyrovsky step. With a further increase in oxidation time, the Tafel slope starts to increase. The oxides on the surface cover the original metal sites. The metal atoms have a strong adsorption ability for Had, so the number of metal sites has a great influence on the reaction rate of the Heyrovsky step [14]. Even though the generation of oxides promotes the occurrence of the Volmer step, an excess of oxides can hinder the progress of the Heyrovsky step. Hence, the HER catalytic performance of the samples shows a tendency to increase first and then decrease with the increase in oxidation time. Figure 5d shows the CV curves of np-AlFeNiO-4s, and Figure 5e shows the double-layer capacitances (Cdl), which were obtained from Figure 5d. With the increase in electrooxidation time, the double-layer capacitance Cdl shows a trend of increasing and then decreasing. TOF values were calculated for each sample at η = 100 mV, as shown in Table S2 [29]. With the increase in oxidation time, the TOF value shows a trend of increasing and then decreasing. These results indicate that the formation of a certain proportion of oxide spheres would increase the electrochemically active area, but too much oxide would reduce the number of active sites on the electrode surface, resulting in a decrease in the intrinsic catalytic activity. Table S3 and Figure 6 show the data of overpotentials and Tafel slopes at 10 mA cm−2 from the present work and some recently reported papers. np-AlFeNiO-4s has a great competitive advantage compared with most of the other catalysts.
Figure 7 shows the HER stability test results for np-AlFeNiO-4s. The stability of the sample was tested by the constant current method. It can be seen that the potential under the constant current of 10 mA cm−2 has no significant change after the 100 h test (as shown in Figure 7a). Figure 7b shows the SEM image of the sample after 100 h of the stability test. The surface of the sample subjected to the stability test still maintains the nanosheet morphology with certain attached oxide spheres. Figure 7c–f show the comparison of XPS results of np-AlFeNiO-4s before and after the stability test. After a long period of work, both the Al and Ni contents decrease and a few oxides appear, but there is no significant change in the valence state. The above results indicate that np-AlFeNiO-4s would have outstanding long-term stability.
Figure 8a shows the OER test results of the samples. Table S4 shows the overpotentials of the OER of the catalysts at 10 mA cm−2 and 100 mA cm−2 for each oxidation time. The amount of oxidation time does not have a significant effect on the OER catalytic performance of the electrode. Nonetheless, the catalysts still exhibit good OER performance, especially at the high current density. The overpotentials of np-AlFeNiO-4s are 279 mV at 10 mA cm−2 and 318 mV at 100 mA cm−2. Figure 8b shows the Tafel curves of the samples with different oxidation times, and the Tafel slopes are listed in Table S4. The Tafel slopes of all samples are around 40 mV dec−1, indicating the oxidation process would have little influence on the OER performance. Figure 9 and Table S5 compare the overpotentials and Tafel slopes of some OER catalysts. Based on the fact that np-AlFeNiO-4s exhibits excellent OER performance at high currents, we also compared it with other catalysts reported in the literature at 100 mA cm−2. The catalysts in this work exhibit highly competitive compared with the other reported high-activity alkaline OER catalysts. Figure 10 shows the results of the OER stability tests on np-AlFeNiO-4s. The stability of the samples was tested by the constant current method. It can be seen that the potential is stable during the constant current (10 mA cm−2) test of 100 h, which indicates that np-AlFeNiO-4s would have good OER stability.
Based on the good HER and OER performance, the overall water-splitting electrolyzer was built using nanoporous ribbons as both the anode and cathode in an alkaline aqueous solution. As shown in Figure 11a, np-AlFeNiO-4s exhibits excellent total water-splitting performance. Only 1.56 V of voltage is required to reach 10 mA cm−2 of current density, while the unoxidized sample requires a voltage of 1.78 V. For the long-time oxidized samples, such as np-AlFeNiO-100s, although the required voltage at a small current (10 mA cm−2) is similar to that of np-AlFeNiO-4s, greater voltages are required to drive them at high currents because of the poorer conductivity of the oxides. The LSV curves of the np-AlFeNiO-0s, -4s, and -100s two-electrode configuration systems in 1 M KOH with iR correction are shown in Figure S3. Figure 11b shows the overall water-splitting performance stability of the np-AlFeNiO-4s catalyst. The water-splitting voltage remains constant after a constant current test of 10 mA cm−2 for 100 h, indicating the excellent stability of np-AlFeNiO-4s. Figure 12 and Table S6 list the total water-splitting performance of the catalysts in this work and other reports in the literature. It can be seen that the catalysts in this work have excellent total water-splitting performance.

4. Conclusions

In the present work, electrooxidation was carried out to improve the water-splitting performance of the nanoporous AlFiNi alloy with a self-supporting structure. A very short time of electrooxidation can significantly enhance the HER activity of the nanoporous AlFeNi alloy. The np-AlFeNiO-4s electrode exhibits remarkable HER activity with the overpotential of 32 mV at 10 mA cm−2. Electrooxidation makes the metal and metal oxide sites on the electrode surface effectively coupled, which greatly enhances the kinetic rate of the Volmer and Heyrovsky steps. The np-AlFeNiO-4s electrode also shows good OER activity with an overpotential of 279 mV at 10 mA cm−2. Only 1.56 V of voltage is required to drive 10 mA cm−2 of current density for the whole water electrolysis reaction. Electrooxidation improves HER performance by enhancing water adsorption abilities. The np-AlFeNiO-4s electrode exhibits outstanding long-term stability with little change after a 100 h water splitting test. Overall, np-AlFeNiO-4s has a promising application in water electrolysis reactions.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/en17071591/s1, Figure S1: The SEM of a cross-section of the dealloyed sample; Figure S2: (a) TEM and (b) HRTEM images of the np-AlFeNiO-4s sample; Figure S3: The LSV curves of the np-AlFeNiO-0s, -4s, and -100s two-electrode configuration systems in 1 M KOH with iR correction; Table S1: The EDX results of the samples; Table S2: The overpotentials at 10 mA cm−2, Tafel slope, Cdl, ECSA, and TOF of np-AlFeNiO for the HER; Table S3: The overpotentials at 10 mA cm−2, Tafel slope for HER electrocatalysts in the recently reported literature; Table S4: The overpotentials at 10 mA cm−2 and 100 mA cm−2, the Tafel slope of np-AlFeNiO for the OER; Table S5: The overpotentials at 10 mA cm−2 and 100 mA cm−2, the Tafel slope for OER electrocatalysts in the recently reported literature; Table S6: Comparison of the overall water-splitting performance for electrocatalysts in the recently reported literature.

Author Contributions

Conceptualization, Z.M.; methodology, Z.M., Y.L. and H.J.; validation, H.Z. and S.Z.; formal analysis, Z.L. and Z.C.; investigation, Z.M., W.X. and Z.G.; resources, S.Z.; data curation, Z.M.; writing—original draft preparation, Z.M.; writing—review and editing, Z.G. and S.Z.; supervision, W.X. and S.Z.; project administration, S.Z.; funding acquisition, S.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China, grant numbers 52271152 and 52371161, and the Tianjin Natural Science Foundation, grant number 22JCQNJC00670.

Data Availability Statement

The data are contained within this article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Xia, X.Y.; Wang, L.J.; Sui, N.; Colvin, V.L.; Yu, W.W. Recent progress in transition metal selenide electrocatalysts for water splitting. Nanoscale 2020, 12, 12249–12262. [Google Scholar] [CrossRef] [PubMed]
  2. Anantharaj, S.; Noda, S. Amorphous Catalysts and Electrochemical Water Splitting: An Untold Story of Harmony. Small 2020, 16, 1905779. [Google Scholar] [CrossRef] [PubMed]
  3. Wang, W.; Xu, X.M.; Zhou, W.; Shao, Z.P. Recent Progress in Metal-Organic Frameworks for Applications in Electrocatalytic and Photocatalytic Water Splitting. Adv. Sci. 2017, 4, 1600371. [Google Scholar] [CrossRef] [PubMed]
  4. Jiang, R.; Cui, Z.D.; Xu, W.; Zhu, S.L.; Liang, Y.Q.; Li, Z.Y.; Wu, S.L.; Chang, C.T.; Inoue, A. Highly efficient amorphous np-PdFePC catalyst for hydrogen evolution reaction. Electrochim. Acta 2019, 328, 135082. [Google Scholar] [CrossRef]
  5. Gao, R.; Zhu, J.; Yan, D.P. Transition metal-based layered double hydroxides for photo(electro)chemical water splitting: A mini review. Nanoscale 2021, 13, 13593–13603. [Google Scholar] [CrossRef] [PubMed]
  6. Sanati, S.; Morsali, A.; Garcia, H. First-row transition metal-based materials derived from bimetallic metal-organic frameworks as highly efficient electrocatalysts for electrochemical water splitting. Energy Environ. Sci. 2022, 15, 3119–3151. [Google Scholar] [CrossRef]
  7. Wang, J.J.; Yue, X.Y.; Yang, Y.Y.; Sirisomboonchai, S.; Wang, P.F.; Ma, X.L.; Abudula, A.; Guan, G.Q. Earth-abundant transition-metal-based bifunctional catalysts for overall electrochemical water splitting: A review. J. Alloys Compd. 2020, 819, 153346. [Google Scholar] [CrossRef]
  8. Al-Naggar, A.H.; Shinde, N.M.; Kim, J.S.; Mane, R.S. Water splitting performance of metal and non-metal-doped transition metal oxide electrocatalysts. Coord. Chem. Rev. 2023, 474, 214864. [Google Scholar] [CrossRef]
  9. Yang, F.; Xiong, T.Z.; Huang, P.; Zhou, S.H.; Tan, Q.R.; Yang, H.; Huang, Y.C.; Balogun, M.S. Nanostructured transition metal compounds coated 3D porous core-shell carbon fiber as monolith water splitting electrocatalysts: A general strategy. Chem. Eng. J. 2021, 423, 11. [Google Scholar] [CrossRef]
  10. Vij, V.; Sultan, S.; Harzandi, A.M.; Meena, A.; Tiwari, J.N.; Lee, W.G.; Yoon, T.; Kim, K.S. Nickel-Based Electrocatalysts for Energy-Related Applications: Oxygen Reduction, Oxygen Evolution, and Hydrogen Evolution Reactions. ACS Catal. 2017, 7, 7196–7225. [Google Scholar] [CrossRef]
  11. Oshchepkov, A.G.; Bonnefont, A.; Saveleva, V.A.; Papaefthimiou, V.; Zafeiratos, S.; Pronkin, S.N.; Parmon, V.N.; Savinova, E.R. Exploring the Influence of the Nickel Oxide Species on the Kinetics of Hydrogen Electrode Reactions in Alkaline Media. Top. Catal. 2016, 59, 1319–1331. [Google Scholar] [CrossRef]
  12. Danilovic, N.; Subbaraman, R.; Strmcnik, D.; Chang, K.C.; Paulikas, A.P.; Stamenkovic, V.R.; Markovic, N.M. Enhancing the Alkaline Hydrogen Evolution Reaction Activity through the Bifunctionality of Ni(OH)2/Metal Catalysts. Angew. Chem.—Int. Ed. 2012, 51, 12495–12498. [Google Scholar] [CrossRef] [PubMed]
  13. Gong, M.; Zhou, W.; Tsai, M.C.; Zhou, J.; Guan, M.; Lin, M.C.; Zhang, B.; Hu, Y.; Wang, D.Y.; Yang, J.; et al. Nanoscale nickel oxide/nickel heterostructures for active hydrogen evolution electrocatalysis. Nat. Commun. 2014, 5, 4695. [Google Scholar] [CrossRef] [PubMed]
  14. Peng, L.X.; Liang, Y.Q.; Wu, S.L.; Li, Z.Y.; Sun, H.J.; Jiang, H.; Zhu, S.L.; Cui, Z.D.; Li, L. Nanoporous Ni/NiO catalyst for efficient hydrogen evolution reaction prepared by partial electro-oxidation after dealloying. J. Alloys Compd. 2022, 911, 165061. [Google Scholar] [CrossRef]
  15. Alsabban, M.M.; Eswaran, M.K.; Peramaiah, K.; Wahyudi, W.; Yang, X.L.; Ramalingam, V.; Hedhili, M.N.; Miao, X.H.; Schwingenschlögl, U.; Li, L.J.; et al. Unusual Activity of Rationally Designed Cobalt Phosphide/Oxide Heterostructure Composite for Hydrogen Production in Alkaline Medium. ACS Nano 2022, 16, 3906–3916. [Google Scholar] [CrossRef] [PubMed]
  16. Ullah, N.; Zhao, W.T.; Lu, X.Q.; Oluigbo, C.J.; Shah, S.A.; Zhang, M.M.; Xie, J.M.; Xu, Y.G. In situ growth of M-MO (M = Ni, Co) in 3D graphene as a competent bifunctional electrocatalyst for OER and HER. Electrochim. Acta 2019, 298, 163–171. [Google Scholar] [CrossRef]
  17. Liu, J.X.; Fan, X.Y.; Ning, G.Y.; Zheng, M.; Shi, K.; Sun, Y.; Gao, Y.J.; Zhang, Y.F.; Wang, H. The high-efficiency electrochemical catalysis of nitrogen-doped carbon nanotubes materials modified with Cu-Fe oxide alloy nanoparticles for HER and ORR. Int. J. Hydrogen Energy 2022, 47, 34090–34101. [Google Scholar] [CrossRef]
  18. NavakoteswaraRao, V.; Shankar, M.V.; Yang, B.; Ahn, C.W.; Yang, J. Effective excitons separation in starfish Bi2S3/TiO2 nanostructures for enhanced hydrogen production. Mater. Today Chem. 2022, 26, 101096. [Google Scholar] [CrossRef]
  19. Zhao, X.J.; Pachfule, P.; Li, S.; Simke, J.R.J.; Schmidt, J.; Thomas, A. Bifunctional Electrocatalysts for Overall Water Splitting from an Iron/Nickel-Based Bimetallic Metal-Organic Framework/Dicyandiamide Composite. Angew. Chem.—Int. Ed. 2018, 57, 8921–8926. [Google Scholar] [CrossRef]
  20. Su, M.Y.; Zhu, S.L.; Cui, Z.D.; Li, Z.Y.; Wu, S.L.; Guo, M.Q.; Jiang, H.; Liang, Y.Q. A self-supported FeNi layered double hydroxide anode with high activity and long-term stability for efficient oxygen evolution reaction. Sustain. Energy Fuels 2021, 5, 3205–3212. [Google Scholar] [CrossRef]
  21. 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]
  22. Suryanto, B.H.R.; Wang, Y.; Hocking, R.K.; Adamson, W.; Zhao, C. Overall electrochemical splitting of water at the heterogeneous interface of nickel and iron oxide. Nat. Commun. 2019, 10, 5599. [Google Scholar] [CrossRef] [PubMed]
  23. Zhang, Q.; Bedford, N.M.; Pan, J.; Lu, X.; Amal, R. A Fully Reversible Water Electrolyzer Cell Made Up from FeCoNi (Oxy)hydroxide Atomic Layers. Adv. Energy Mater. 2019, 9, 1901312. [Google Scholar] [CrossRef]
  24. Marakatti, V.S.; Peter, S.C. Synthetically tuned electronic and geometrical properties of intermetallic compounds as effective heterogeneous catalysts. Prog. Solid State Chem. 2018, 52, 1–30. [Google Scholar] [CrossRef]
  25. Zhou, Y.Y.; Liu, H.P.; Zhu, S.L.; Liang, Y.Q.; Wu, S.L.; Li, Z.Y.; Cui, Z.D.; Chang, C.T.; Yang, X.J.; Inoue, A. Highly Efficient and Self-Standing Nanoporous NiO/Al3Ni2 Electrocatalyst for Hydrogen Evolution Reaction. ACS Appl. Energy Mater. 2019, 2, 7913–7922. [Google Scholar] [CrossRef]
  26. Yu, M.Q.; Budiyanto, E.; Tuysuz, H. Principles of Water Electrolysis and Recent Progress in Cobalt-, Nickel-, and Iron-Based Oxides for the Oxygen Evolution Reaction. Angew. Chem.—Int. Ed. 2022, 61, e202103824. [Google Scholar] [CrossRef]
  27. Xue, B.W.; Zhang, C.H.; Wang, Y.Z.; Xie, W.W.; Li, N.W.; Yu, L. Recent progress of Ni-Fe layered double hydroxide and beyond towards electrochemical water splitting. Nanoscale Adv. 2020, 2, 5555–5566. [Google Scholar] [CrossRef]
  28. McCrory, C.C.L.; Jung, S.; Ferrer, I.M.; Chatman, S.M.; Peters, J.C.; Jaramillo, T.F. Benchmarking Hydrogen Evolving Reaction and Oxygen Evolving Reaction Electrocatalysts for Solar Water Splitting Devices. J. Am. Chem. Soc. 2015, 137, 4347–4357. [Google Scholar] [CrossRef]
  29. Kibsgaard, J.; Tsai, C.; Chan, K.; Benck, J.D.; Norskov, J.K.; Abild-Pedersen, F.; Jaramillo, T.F. Designing an improved transition metal phosphide catalyst for hydrogen evolution using experimental and theoretical trends. Energy Environ. Sci. 2015, 8, 3022–3029. [Google Scholar] [CrossRef]
  30. Sun, Q.Q.; Dong, Y.J.; Wang, Z.L.; Yin, S.W.; Zhao, C. Synergistic Nanotubular Copper-Doped Nickel Catalysts for Hydrogen Evolution Reactions. Small 2018, 14, 1704137. [Google Scholar] [CrossRef]
  31. Cui, X.D.; Zhang, B.L.; Zeng, C.Y.; Wen, H.; Guo, S.M. Monolithic nanoporous Ni-Fe alloy by dealloying laser processed Ni-Fe-Al as electrocatalyst toward oxygen evolution reaction. Int. J. Hydrogen Energy 2018, 43, 15234–15244. [Google Scholar] [CrossRef]
  32. Dong, C.; Kou, T.; Gao, H.; Peng, Z.; Zhang, Z. Eutectic-Derived Mesoporous Ni-Fe-O Nanowire Network Catalyzing Oxygen Evolution and Overall Water Splitting. Adv. Energy Mater. 2018, 8, 1701347. [Google Scholar] [CrossRef]
  33. Feng, S.J.; Yang, W.; Wang, Z.B. Synthesis of porous NiFe2O4 microparticles and its catalytic properties for methane combustion. Mater. Sci. Eng. B-Adv. Funct. Solid-State Mater. 2011, 176, 1509–1512. [Google Scholar] [CrossRef]
  34. Zhao, L.; Zhang, Y.; Zhao, Z.L.; Zhang, Q.H.; Huang, L.B.; Gu, L.; Lu, G.; Hu, J.S.; Wan, L.J. Steering elementary steps towards efficient alkaline hydrogen evolution via size-dependent Ni/NiO nanoscale heterosurfaces. Natl. Sci. Rev. 2020, 7, 27–36. [Google Scholar] [CrossRef] [PubMed]
  35. Zhang, B.; Zhu, C.; Wu, Z.S.; Stavitski, E.; Lui, Y.H.; Kim, T.H.; Liu, H.; Huang, L.; Luan, X.C.; Zhou, L.; et al. Integrating Rh Species with NiFe-Layered Double Hydroxide for Overall Water Splitting. Nano Lett. 2020, 20, 136–144. [Google Scholar] [CrossRef] [PubMed]
  36. Kumar, A.; Rao, V.N.; Kumar, A.; Mushtaq, A.; Sharma, L.; Halder, A.; Pal, S.K.; Shankar, M.V.; Krishnan, V. Three-Dimensional Carbonaceous Aerogels Embedded with Rh-SrTiO3 for Enhanced Hydrogen Evolution Triggered by Efficient Charge Transfer and Light Absorption. ACS Appl. Energy Mater. 2020, 3, 12134–12147. [Google Scholar] [CrossRef]
  37. Duan, J.J.; Chen, S.; Zhao, C. Ultrathin metal-organic framework array for efficient electrocatalytic water splitting. Nat. Commun. 2017, 8, 15341. [Google Scholar] [CrossRef] [PubMed]
  38. Niu, S.; Jiang, W.J.; Tang, T.; Zhang, Y.; Li, J.H.; Hu, J.S. Facile and Scalable Synthesis of Robust Ni(OH)2 Nanoplate Arrays on NiAl Foil as Hierarchical Active Scaffold for Highly Efficient Overall Water Splitting. Adv. Sci. 2017, 4, 1700084. [Google Scholar] [CrossRef] [PubMed]
  39. Pehlivan, I.B.; Arvizu, M.A.; Qiu, Z.; Niklasson, G.A.; Edvinsson, T. Impedance Spectroscopy Modeling of Nickel-Molybdenum Alloys on Porous and Flat Substrates for Applications in Water Splitting. J. Phys. Chem. C 2019, 123, 23890–23897. [Google Scholar] [CrossRef]
  40. Liu, Y.K.; Jiang, S.; Li, S.J.; Zhou, L.; Li, Z.H.; Li, J.M.; Shao, M.F. Interface engineering of (Ni, Fe)S2@MoS2 heterostructures for synergetic electrochemical water splitting. Appl. Catal. B-Environ. 2019, 247, 107–114. [Google Scholar] [CrossRef]
  41. Zhang, B.; Zhang, X.M.; Wei, Y.; Xia, L.; Pi, C.R.; Song, H.; Zheng, Y.; Gao, B.; Fu, J.J.; Chu, P.K. General synthesis of NiCo alloy nanochain arrays with thin oxide coating: A highly efficient bifunctional electrocatalyst for overall water splitting. J. Alloys Compd. 2019, 797, 1216–1223. [Google Scholar] [CrossRef]
  42. Zhang, X.; Xue, Y.Q.; Yin, X.Z.; Shen, L.; Zhu, K.; Huang, X.M.; Cao, D.X.; Yao, J.X.; Wang, G.L.; Yan, Q. Defect-rich MnxOy complex Fe-Ni sulfide heterogeneous electrocatalyst for a highly efficient hydrogen evolution reaction. J. Power Sources 2022, 540, 231664. [Google Scholar] [CrossRef]
  43. Velayutham, R.; Raj, C.J.; Jang, H.M.; Cho, W.J.; Palanisamy, K.; Kaya, C.; Kim, B.C. Surface-oriented heterostructure of iron metal organic framework confined molybdenum disulfides as an efficient bifunctional electrocatalyst for overall water splitting. Mater. Today Nano 2023, 24, 100387. [Google Scholar] [CrossRef]
  44. Xu, H.; Zhang, D.; Liu, M.; Ye, D.; Huo, S.; Chen, W.; Zhang, J. Self-supporting hierarchical Co3O4-nanowires@NiO-nanosheets core-shell nanostructure on carbon foam to form efficient bifunctional electrocatalyst for overall water splitting. J. Colloid Interface Sci. 2023, 654, 1293–1302. [Google Scholar] [CrossRef] [PubMed]
  45. Liu, X.Y.; Lu, H.L.; Zhu, S.L.; Cui, Z.D.; Li, Z.Y.; Wu, S.L.; Xu, W.C.; Liang, Y.Q.; Long, G.K.; Jiang, H. Alloying-Triggered Phase Engineering of NiFe System via Laser-Assisted Al Incorporation for Full Water Splitting. Angew. Chem.—Int. Ed. 2023, 62, e202300800. [Google Scholar] [CrossRef]
  46. He, Y.J.; Shen, J.; Li, Q.; Zheng, X.Z.; Wang, Z.X.; Cui, L.; Xu, J.T.; Liu, J.Q. In-situ growth of VS4 nanorods on Ni-Fe sulfides nanoplate array towards achieving a highly efficient and bifunctional electrocatalyst for total water splitting. Chem. Eng. J. 2023, 474, 145461. [Google Scholar] [CrossRef]
  47. Lim, D.; Kim, S.; Kim, N.; Oh, E.; Shim, S.E.; Baeck, S.H. Strongly Coupled Ni/Ni(OH)2 Hybrid Nanocomposites as Highly Active Bifunctional Electrocatalysts for Overall Water Splitting. ACS Sustain. Chem. Eng. 2020, 8, 4431–4439. [Google Scholar] [CrossRef]
  48. Huang, C.L.; Chuah, X.F.; Hsieh, C.T.; Lu, S.Y. NiFe Alloy Nanotube Arrays as Highly Efficient Bifunctional Electrocatalysts for Overall Water Splitting at High Current Densities. ACS Appl. Mater. Interfaces 2019, 11, 24096–24106. [Google Scholar] [CrossRef] [PubMed]
  49. Karuppasamy, L.; Gurusamy, L.; Ananan, S.; Barton, S.C.; Liu, C.H.; Wu, J.J. Metal-organic frameworks derived interfacing Fe2O3/ZnCo2O4 multimetal oxides as a bifunctional electrocatalyst for overall water splitting. Electrochim. Acta 2023, 449, 142242. [Google Scholar] [CrossRef]
  50. Chauhan, P.; Siraj, S.; Joseph, K.S.; Dabhi, S.; Bhadu, G.R.; Sahatiya, P.; Sumesh, C.K. Synergistically Driven CoCr-LDH@VNiS2 as a Bifunctional Electrocatalyst for Overall Water Splitting and Flexible Supercapacitors. ACS Appl. Mater. Interfaces 2023, 15, 32515–32524. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic illustration of the formation of the samples.
Figure 1. Schematic illustration of the formation of the samples.
Energies 17 01591 g001
Figure 2. XRD patterns of the precursor ribbon, the ribbon after dealloying, and the ribbons oxidized for 4 s and 100 s.
Figure 2. XRD patterns of the precursor ribbon, the ribbon after dealloying, and the ribbons oxidized for 4 s and 100 s.
Energies 17 01591 g002
Figure 3. SEM images of the (a) precursor ribbon, (b) the dealloyed ribbon, and ribbons oxidized for (c) 4 s and (d) 100 s.
Figure 3. SEM images of the (a) precursor ribbon, (b) the dealloyed ribbon, and ribbons oxidized for (c) 4 s and (d) 100 s.
Energies 17 01591 g003
Figure 4. The XPS spectra of (a) Al 2p, (b)Ni 2p, (c) Fe 2p, and (d) O 1s.
Figure 4. The XPS spectra of (a) Al 2p, (b)Ni 2p, (c) Fe 2p, and (d) O 1s.
Energies 17 01591 g004
Figure 5. (a) LSV curves of the np-AlFeNiO catalyst and Pt/C for the HER. (b) Tafel plots derived from corresponding LSV curves. (c) Nyquist plots of the np-AlFeNiO catalysts. (d) The CV curve of np-AlFeNiO-4s. (e) Linear fitting of the capacitive currents of the plotted against the scan rates.
Figure 5. (a) LSV curves of the np-AlFeNiO catalyst and Pt/C for the HER. (b) Tafel plots derived from corresponding LSV curves. (c) Nyquist plots of the np-AlFeNiO catalysts. (d) The CV curve of np-AlFeNiO-4s. (e) Linear fitting of the capacitive currents of the plotted against the scan rates.
Energies 17 01591 g005
Figure 6. Comparison of the Tafel slope and η10 for the reported HER catalysts [4,9,17,19,25,35,40,41,42,43,44,45].
Figure 6. Comparison of the Tafel slope and η10 for the reported HER catalysts [4,9,17,19,25,35,40,41,42,43,44,45].
Energies 17 01591 g006
Figure 7. (a) Time-dependent potential profile of np-AlFeNiO-4s for the 100 h HER test, (b) the SEM image after the stability test, and (cf) XPS results before and after the stability test.
Figure 7. (a) Time-dependent potential profile of np-AlFeNiO-4s for the 100 h HER test, (b) the SEM image after the stability test, and (cf) XPS results before and after the stability test.
Energies 17 01591 g007
Figure 8. (a) LSV curves of np-AlFeNiO catalysts for the OER and (b) Tafel plots derived from the corresponding LSV curves.
Figure 8. (a) LSV curves of np-AlFeNiO catalysts for the OER and (b) Tafel plots derived from the corresponding LSV curves.
Energies 17 01591 g008
Figure 9. Comparison of (a) the Tafel slope and η10 and (b) the Tafel slope and η100 for the reported OER catalysts [31,35,40,41,43,45,46,47,48,49].
Figure 9. Comparison of (a) the Tafel slope and η10 and (b) the Tafel slope and η100 for the reported OER catalysts [31,35,40,41,43,45,46,47,48,49].
Energies 17 01591 g009
Figure 10. Time-dependent potential profile of np-AlFeNiO-4s for the 100 h OER test.
Figure 10. Time-dependent potential profile of np-AlFeNiO-4s for the 100 h OER test.
Energies 17 01591 g010
Figure 11. (a) LSV curves of the np-AlFeNiO-0s, -4s, and -100s two-electrode configuration systems in 1 M KOH without iR correction and (b) the time-dependent potential profile of np-AlFeNiO-4s serving as both the cathode and anode under a current density of 10 mA cm−2 for 100 h in 1 M KOH solution.
Figure 11. (a) LSV curves of the np-AlFeNiO-0s, -4s, and -100s two-electrode configuration systems in 1 M KOH without iR correction and (b) the time-dependent potential profile of np-AlFeNiO-4s serving as both the cathode and anode under a current density of 10 mA cm−2 for 100 h in 1 M KOH solution.
Energies 17 01591 g011
Figure 12. Comparison of the overall water-splitting performance at the current density of 10 mA cm−2 for electrocatalysts in the recently reported literature [35,40,41,43,45,46,47,48,49,50].
Figure 12. Comparison of the overall water-splitting performance at the current density of 10 mA cm−2 for electrocatalysts in the recently reported literature [35,40,41,43,45,46,47,48,49,50].
Energies 17 01591 g012
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

Ma, Z.; Xu, W.; Gao, Z.; Liang, Y.; Jiang, H.; Li, Z.; Cui, Z.; Zhang, H.; Zhu, S. Self-Supporting np-AlFeNiO Bifunctional Electrode Material for Electrochemical Water Splitting Prepared by Electrooxidation. Energies 2024, 17, 1591. https://doi.org/10.3390/en17071591

AMA Style

Ma Z, Xu W, Gao Z, Liang Y, Jiang H, Li Z, Cui Z, Zhang H, Zhu S. Self-Supporting np-AlFeNiO Bifunctional Electrode Material for Electrochemical Water Splitting Prepared by Electrooxidation. Energies. 2024; 17(7):1591. https://doi.org/10.3390/en17071591

Chicago/Turabian Style

Ma, Zhihui, Wence Xu, Zhonghui Gao, Yanqin Liang, Hui Jiang, Zhaoyang Li, Zhenduo Cui, Huifang Zhang, and Shengli Zhu. 2024. "Self-Supporting np-AlFeNiO Bifunctional Electrode Material for Electrochemical Water Splitting Prepared by Electrooxidation" Energies 17, no. 7: 1591. https://doi.org/10.3390/en17071591

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