Next Article in Journal
Hybrid Method for Endpoint Prediction in a Basic Oxygen Furnace
Next Article in Special Issue
Study on the Effect of A/B Site Co-Doping on the Oxygen Evolution Reaction Performance of Strontium Cobaltite
Previous Article in Journal
Effect of Prior Cold Reduction of C–Si–Mn Hot-Rolled Sheet on Microstructures and Mechanical Properties after Quenching and Partitioning Treatment
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Magnetic-Field-Induced Strain Enhances Electrocatalysis of FeCo Alloys on Anode Catalysts for Water Splitting

School of Materials Science and Engineering, Hebei University of Technology, Tianjin 300130, China
*
Author to whom correspondence should be addressed.
Metals 2022, 12(5), 800; https://doi.org/10.3390/met12050800
Submission received: 8 April 2022 / Revised: 28 April 2022 / Accepted: 30 April 2022 / Published: 5 May 2022

Abstract

:
In water splitting, the oxygen evolution reaction (OER) performance of transition metal alloy catalysts needs to be further improved. To solve this problem, the method of an external magnetic field was used to improve the OER catalytic performance of the alloy catalyst. In this paper, FeCo alloys with different composition ratios were prepared by an arc melting method, and OER catalysts with different compositions were obtained by annealing treatment. Under the action of a magnetic field, all three groups of catalysts showed a better catalytic performance than those without a magnetic field. The overpotentials of Fe35Co65, Fe22Co78 and Fe15Co85 at a current density of 20 mA cm−2 were reduced by 12 mV, 6 mV and 2 mV, respectively. It is found that, due to the magnetostrictive properties of FeCo alloys, the catalyst itself will generate strain under the action of a magnetic field, and the existence of strain may be the main reason for the enhanced OER performance of the magnetic field. Therefore, this work provides a new idea for the development of magnetic material catalysts and a magnetic field to improve the performance of catalysts.

1. Introduction

In recent years, with the development of industrialization, the burning of fossil fuels has caused serious environmental pollution and climate change, leading to an increasing demand for clean energy for global sustainable development. Due to its high energy density and carbon neutrality, hydrogen is a clean energy carrier and is considered a promising alternative to fossil fuels [1]. Among various hydrogen production methods, electrocatalytic water splitting is the most attractive one [2]. Compared with traditional steam reforming or pyrolysis routes that rely on fossil fuel consumption, electrocatalytic water splitting is theoretically clean, and zero carbon emissions can be achieved by using intermittent renewable electricity [3]. The electrolysis of water includes a cathode hydrogen evolution reaction (HER) and an anode oxygen evolution reaction (OER) [4,5]. The slow multi-step electron and proton transfer process of OER leads to high overpotential, and the corresponding high energy consumption is the bottleneck for improving electrochemical energy conversion efficiency [4]. Hence, researchers have made considerable efforts to explore highly active OER catalysts [6]. At present, Ru/Ir-based catalysts are the benchmarks of OER. However, their large-scale application is severely limited by their scarcity and high cost [7]. In recent years, transition metals, such as Fe, Co and Ni, with their abundant element reserves, environmentally friendly characteristic and low-cost, have received widespread attention due to their outstanding OER performance, and are widely regarded as outstanding alternatives to expensive ruthenium/iridium-based catalysts [5,8,9,10,11]. Compared with single-metal-based catalysts, metal alloys show better OER activity, which benefits from strain effects and synergistic effects, which have an impact on crystal structure changes and electronic modulation [12]. Binary alloys such as NiFe [10,13] and FeCo [14,15,16] have been reported as effective OER catalysts, but the performance improvement of these catalysts has reached a bottleneck. It is still a great challenge to obtain more efficient and stable catalysts.
In addition to the development of efficient and stable catalysts, the proposal of an external field has also aroused the interest of researchers, such as a gravity field [17], light field [18], ultrasonic field [19] and electric field [20], which can effectively improve the mass transfer on the electrode surface and change the reaction kinetics. Interestingly, the latest research confirms that coupling the magnetic field and electrochemistry is a promising new strategy to enhance the electrochemical reaction [21,22,23]. For example, Elias et al. prepared a Ni-W alloy as the HER electrode material, and, under the action of a magnetic field of 0.1–0.4 mT, the hydrogen production efficiency was significantly improved. Magnetohydrodynamic driving (MHD)-induced convection and the rapid detachment of H2 bubbles lead to a reduced polarization resistance and exposure of active sites, which, in turn, accelerate the catalytic reaction [24]. Westsson et al. confirmed the action of the magnetic field on the hydrogen adsorption of Pt-based catalysts with a core–shell structure [25]. Galán-Mascarós et al. showed that highly magnetic electrocatalysts, such as mixed Ni-Fe-Zn-based oxides, exhibit higher oxygen evolution reaction activity when a magnetic field is applied to the anode. After the introduction of a magnetic field, the ferromagnetic domain of oxygen atoms is increased to suppress the antiferromagnetic domain, which is beneficial to the formation of active sites in parallel configuration, and the catalytic reaction activity is improved by the evolution of singlet spin to triplet spin [21]. Li et al. put the electrolytic cell between permanent magnets with a medium magnetic field and proved the influence of the magnetic field on the Co3O4 catalyst. Under the action of the magnetic field, the OER performance was significantly improved [26]. Wei et al. further studied the possible improvement of the catalytic activity of the same material for the oxygen reduction reaction (ORR). By applying an external magnetic field, a small increase in the selectivity of the 4e-pathway was achieved. The magnetic field induces the degeneracy of the unpaired electron spins in Co3O4, resulting in enhanced electronic energy states that contribute to the activation energy of the electron transfer reaction, thereby promoting the redox reaction rate. This work minimizes the generation of H2O2 and improves energy utilization [27]. These recent studies have shown the influence of magnetic fields on the catalytic performance of transition metals and their oxides.
It is well known that the FeCo alloy is not only a catalyst with an excellent performance that plays an important role in the field of catalysis, but also a magnetostrictive material with a wide range of applications. Recently, it has been shown that magnetostriction enhancement occurs at the junction of the FeCo alloy bcc/(fcc + bcc) phases, and the enhancement is due to the precipitation of nanoprecipitated phases in the bcc matrix. Under the action of a magnetic field, greater strains can be generated [28,29]. The engineered strain effect in solid-solution alloys enhances catalytic activity by shifting the d-band center and changing the intermediate adsorption energy [30,31], which may facilitate a reduction in the reaction potential of the OER. Magnetic fields promote electrocatalysis in many ways, but, for now, the effect on metal alloy catalysts has rarely been reported.
Herein, an attempt was made to explore the changes in the OER electrocatalytic properties of FeCo alloys under magnetic fields. We prepared FeCo alloys as OER catalysts by a mechanosynthesis method, and the alloys were properly annealed. The electrochemical properties were tested in the presence and absence of a magnetic field and compared. The results indicate that, in the presence of a magnetic field, strain is generated within the FeCo alloy, resulting in a significant reduction in the OER overpotential and exhibiting a better OER performance than in the absence of a magnetic field. Therefore, this experiment provides a new idea to study the performance enhancement of magnetostrictive materials and catalysts, such as magnetic shape memory alloys with a magnetic field.

2. Experimental Process

2.1. Materials Preparations

The alloy compositions selected in the experiment were Fe35Co65, Fe22Co78 and Fe15Co85, respectively. The raw materials for preparing the samples were all Fe and Co elements with a purity higher than 99.95% (mass fraction), purchased from Zhongnuo New Materials Technology Co., Ltd. (Beijing, China) The test samples were first smelted in a vacuum arc furnace under the protection of an argon atmosphere. In order to ensure the uniform composition, the alloys were repeatedly turned and remelted four times, and, after cooling, a button-shaped ingot was formed. The samples were annealed in an argon atmosphere to make the alloy composition uniform and the atoms arranged in a highly orderly manner. The annealing temperature was 700 °C, kept for 24 h and quenched in salt water to room temperature. The cooled block alloys were polished into particles with 800-mesh sandpaper, separated with permanent magnets and washed with ethanol several times.

2.2. Material Characterizations

The phase structure of the sample was analyzed by X-ray diffraction (XRD) using the German Bruker D8 Discover Cu-Kα radiation X-ray diffractometer (Bruker, Rheinstetten, Germany). The morphological characteristics of the prepared samples were characterized by field emission environment scanning electron microscope (SEM, Quanta 450 FEG, Thermo Fisher Scientific, Waltham, MA, USA). An energy spectrometer (EDS) was used to determine the content of chemical elements in the product. Using Al-Kα excitation source, X-ray photoelectron spectroscopy (XPS, Escalab 250Xi, Thermo Fisher Scientific, Waltham, MA, USA) was used to study the surface states of samples with different Co contents. At room temperature, a vibrating sample magnetometer (VSM, Lakeshore 7307, Lake Shore Cryotronics, Inc., Westerville, OH, USA) was used to measure the magnetostriction of the sample under a maximum magnetic field of 900 mT.

2.3. Electrochemical Measurements

The electrochemical performances of the catalysts were studied in a standard three-electrode system using CHI750E electrochemical workstation. In particular, 10 mg catalyst and 5 mg carbon black (acid-treated XC-72) were placed into a 4mL centrifuge tube, and 600 μL deionized water and 400 μL isopropanol were added into the centrifuge tube, respectively. Finally, 30 μL Nafion concentration of 5% was added, and ultrasound was performed for 40 min to prepare electrode ink. The ink was dropped on the foamed copper (1 × 1 cm2) substrate as the working electrode, and the carbon rod and the saturated calomel (saturated potassium chloride as the filling liquid) electrode were used as the counter electrode and the reference electrode, respectively. The tests were all carried out in an alkaline environment and at ambient temperature. The electrochemical test was performed in a 1 mol/L KOH solution saturated with oxygen. Before the measurement, the working electrode was activated by cyclic voltammetry (CV) at a scan rate of 10 mV/s, and then linear sweep voltammetry (LSV) was performed at a scan rate of 5 mV/s to describe the OER performance. All of the potentials versus Hg/HgO electrode were calibrated to the reversible hydrogen electrode (RHE) according to the Nernst equation E ( RHE ) = E ( SCE ) + 0.0591 × pH + 0.24 V.

3. Results and Discussion

A brief schematic diagram of the experimental setup is shown in Figure S1. The three-electrode system was placed in a magnetic field as shown. The magnetic field is provided by the VSM system, and the strength and direction of the magnetic field can be easily adjusted. It is important that the magnetic field strength remains constant during each step of the test. The magnetic fields applied in the experiments are all provided by the VSM system. The XRD patterns of the three samples Fe35Co65, Fe22Co78 and Fe15Co85 are shown in Figure 1. It can be clearly found that the three samples have different phase structures due to the different percentages of Co. When the Co composition is 65 at.%, the three typical peaks at 2θ are 45.2°, 65.7° and 83.3°, corresponding to (110), (200) and (211) crystal planes, respectively, and are the bcc-FeCo phase structure (PDF#50-0795). When the Co content is 78 at.%, in addition to the three peaks mentioned above, 2θ peaks of 44.2°, 51.5° and 75.8° appear in the spectrum, corresponding to (111), (200) and (211) crystals, respectively. This is a typical fcc-Co (PDF#15-0806) phase structure, which shows that the sample Fe22Co78 has a two-phase structure of bcc + fcc. When the Co content is 85 at.%, a single-phase structure of fcc is obtained, which is consistent with the results reported in the literature [30]. It is well known that the bcc-FeCo phase behaves as a replacement solid solution. Since the atomic radius of the Co element (0.126 nm) is smaller than that of the Fe element (0.127 nm), the substitution of Fe by more Co in the FeCo phase may cause the lattice shrinkage of FeCo products [32,33,34].
The microscopic morphology and particle size of the samples were analyzed by SEM. The SE IMAGES and EDS spectra of Fe35Co65, Fe22Co78 and Fe15Co85 are shown in Figure 2. The SEM in Figure 2a–c shows that the alloys powder obtained by sandpaper grinding are irregular micro-particles with fine dispersion and a particle size of 1–10 μm; the smaller particles will affect the magnetic properties of ferromagnetic materials [35]. The EDS in Figure 2d–f confirms that the composition ratios of Fe35Co65, Fe22Co78 and Fe15Co85 are in line with the stoichiometric ratio of the initial ingredients, which also shows that the surface of the alloy particles is slightly oxidized.
XPS was used to measure the electronic state of the elements of Fe and Co atoms in samples with different Co contents. Figure 3 shows the XPS narrow scan spectra of Fe 2p and Co 2p of the Fe35Co65 sample. The Co 2p spectrum (Figure 3a) clearly shows two spin-orbit double peaks at 780.2 eV and 796.2 eV, corresponding to Co2+ 2p3/2 and Co2+ 2p1/2, respectively, as oxidation peaks. The two oscillating satellites are at 786.2 eV and 802.3 eV (denoted as “Sat”). In addition, the zero-valent Co peaks are concentrated at 777.8 eV and 793 eV, indicating the presence of cobalt derived from FeCo alloys [36]. Similarly, the Fe 2p spectrum (Figure 3b) shows two spin-orbit double peaks for Fe2+ 2p3/2 (711.8 eV) and Fe2+ 2p1/2 (724.6 eV) and an oscillating peak (719.3 eV). In addition, the peak at 707.1 eV corresponds to zero-valent iron [37,38], which originates from the FeCo alloy.
Figure 4a–c shows that the FeCo samples exhibited a typical hysteresis loop at room temperature, indicating that the samples had ferromagnetic characteristics at room temperature, and their saturation magnetization was above 180 emu g−1. The hysteresis loop in Figure 4d shows the anti-magnetic properties of the copper foam substrate. Additionally, it was shown that the catalysts were nearly saturated and magnetized under the action of a permanent magnetic field (~700 mT), which suggests that the impact of a permanent magnetic field on the magnetic catalyst was effective.
The magnetostriction of products with different initial Co contents was studied at room temperature using a VSM with a maximum magnetic field of 900 mT at room temperature. The magnetostriction coefficient test method is as follows: bond the strain resistance sheet to the material to be tested. Under the action of the magnetic field, the size of the magnet will change, and the stress will be applied to the strain resistance, thereby changing the resistance value of the strain resistance. The proportional relationship is transformed into the strain value of the specimen and displayed, which is the magnetostriction coefficient λ. The magnitude of magnetostriction is described by the magnetostriction coefficient λs (λs = λ − λ), as shown in Figure 5. In the Fe100−xCox alloy system, the magnetostriction first gradually becomes larger with the increase in Co content, and the maximum magnetostriction occurs at the interface (x~65) where the bcc phase and the (bcc + fcc) phase intersect. The reason is that the precipitation of nano-precipitates in the bcc matrix will cause the tetragonal distortion of the bcc matrix, thereby increasing the large magnetostriction. However, with the continuous increase in Co content and the growth of nano-precipitated phases, large-sized fcc-Co phases gradually appear, which leads to a sharp decrease in magnetostriction [28,29].
In order to visualize the influence of the magnetic field on the OER performance, LSVs were carried out in a KOH electrolyte as shown in Figure 6a–c. It can be clearly seen that the OER performance of all three catalysts is enhanced under the action of the magnetic field. Interestingly, under the same magnetic field strength, the change in the sample Fe35Co65 is the most obvious, where the LSV curve is shifted to the left more and the overpotential decreases the most. For Fe22Co78, the LSV curve also shows a more pronounced leftward shift, which indicates that its OER properties also improve to some extent under the magnetic field, while the change is smaller compared to Fe35Co65. For Fe15Co85, on the other hand, the LSV curve shows only a small deviation in the presence or absence of a magnetic field. Intriguingly, all three samples appeared to perform better than before under the action of a magnetic field. The amount of variation in the OER performance of these catalysts under a magnetic field shows a clear agreement with the strain generated under the magnetic field. Therefore, a preliminary conclusion can be drawn that the applied magnetic field changes the OER performance of the catalyst by affecting the strain of the material. In addition, the variation in the OER properties of blank copper foam in the presence or absence of a magnetic field was also tested. Copper is the anti-magnetic substance from Figure 4 that exhibits weak magnetic properties, so the response to the magnetic field is relatively small. As shown in Figure S2, the LSV curve does not change significantly, but showed a slight deviation similar to Fe15Co85, which could be the influence of the magnetic field on the mass transfer of the solution [24,39,40]. Figure 6d shows the corresponding relationship between the magnitude of magnetostriction and the change in overpotential. It can be clearly seen that the greater the strain caused by the magnetic field, the more the overpotential decreases, which means the more obvious the effect of the magnetic field on the catalyst. Meanwhile, Figure S3 shows the impedance change with and without the presence of a magnetic field. From the figure, we can see that, under the action of a magnetic field, the catalysts all exhibit smaller impedance, which is consistent with the corresponding electrocatalytic OER activity.
In addition, the capacitance (Cdl) of the electrochemical double layer (EDL), which is related to the electrochemically active surface area (ECSA) of the catalysts [41,42], was determined using the CV curves in the potential ranges of 0.1−0.15 V. As shown in Figure S4, it can be seen clearly that the Cdl values of the three samples under a magnetic field were greater than those under no magnetic field. This further confirms that the external magnetic field increases the active area and improves the catalytic performance of the material.
To further verify the effect of magnetic-field-induced strain on the OER performance, the OER performance of sample Fe35Co65 was tested at different magnetic field strengths. The LSV curves measured under the gradient of a magnetic field strength of 0, 300 and 700 mT are shown in Figure 7 and Table S1. An interesting phenomenon can be seen: as the magnetic field strength increases, the LSV curve without the magnetic field (black line) starts to gradually shift to the left; when the magnetic field strength reached 300 mT (red line), a significant shift started to appear, and as the magnetic field strength increased to 700 mT (green line), the curve continued to shift toward a lower voltage and the current density became larger. It can be concluded that there is a correlation between the magnitude of the magnetic field strength and the degree of improvement in the OER performance. In fact, the magnetostriction of the FeCo catalyst has reached saturation at approximately 500 mT, and the overpotential change in the sample at this magnetic field strength has also reached the maximum. A further increase in the magnetic field will not further improve the performance of the catalyst (Figure S5), but the maximum magnetic field of 700 mT was finally chosen to maintain the consistency of the experiment. Table S2 summarizes examples of magnetic fields affecting electrochemical catalytic reactions in recent years [21,24,25,27,43,44,45,46,47,48].
Based on the above results, it can be explained in conjunction with the magnetostriction curve of Figure 5. As the magnetic field strength increases, the strain generated within the alloy increases and the improvement in OER properties becomes greater and greater (Figure 8). However, when the magnetic field increases to a certain strength, the performance no longer changes (Figure 8b). The reason is that the magnetostriction of the FeCo alloy reaches saturation and the strain no longer occurs as the magnetic field continues to increase, so the OER performance also does not change anymore. Under the action of a magnetic field, magnetostriction occurs, and the strain caused by magnetostriction causes the metal atoms to have less d-orbital overlap, which reduces the bandwidth and increases the d-band center energy, thereby driving more antibonding states above the Fermi level and allowing the adsorbates to interact with them, increasing the binding energy of the adsorbates, which, in turn, enhances the OER performance [30,31,49,50].

4. Conclusions

Fe100−xCox alloys with different composition ratios were prepared by a mechanical alloying method, and three OER catalysts with different phase structures were obtained after annealing. After applying an external magnetic field, the three catalysts showed different enhancements in performance due to the different strains generated, with Fe35Co65 increasing the most significantly and Fe22Co78 also showing more pronounced changes, whereas Fe15Co85 showed minimal changes, which corresponded well to their magnetic strains. The relationship between the strain magnitude and enhanced catalytic performance is also well illustrated by comparing the performance of different magnetic field gradients. Therefore, in this study, the magnetic field induced the strain of the catalyst, and the strain effect may be the main reason for the enhanced OER performance.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/met12050800/s1, Figures S1–S5; Tables S1 and S2. Refs. [21,22,24,27,43,44,45,46,47,48] are cited in Supplementary Materials.

Author Contributions

H.L.: writing—review and editing. Y.R.: conceptualization, methodology, software, investigation, writing—original draft. K.W.: validation, formal analysis, visualization. X.M.: validation, formal analysis, visualization. S.S.: resources, supervision, data curation. J.G.: resources, supervision, data curation. X.Y.: writing—review and editing. Z.L.: writing—review and editing. 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 (Nos. 52171176, 51871088).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shi, Y.; Zhang, B. Recent advances in transition metal phosphide nanomaterials: Synthesis and applications in hydrogen evolution reaction. Chem. Soc. Rev. 2016, 45, 1529–1541. [Google Scholar] [CrossRef]
  2. Wang, J.; Cui, W.; Liu, Q.; Xing, Z.; Asiri, A.M.; Sun, X. Recent progress in cobalt-based heterogeneous catalysts for electrochemical water splitting. Adv. Mater. 2016, 28, 215–230. [Google Scholar] [CrossRef]
  3. Zou, X.; Zhang, Y. Noble metal-free hydrogen evolution catalysts for water splitting. Chem. Soc. Rev. 2015, 44, 5148–5180. [Google Scholar] [CrossRef] [PubMed]
  4. Suen, N.T.; Hung, S.F.; Quan, Q.; Zhang, N.; Xu, Y.J.; Chen, H.M. Electrocatalysis for the oxygen evolution reaction: Recent development and future perspectives. Chem. Soc. Rev. 2017, 46, 337–365. [Google Scholar] [CrossRef] [PubMed]
  5. Li, P.; Zhao, R.; Chen, H.; Wang, H.; Wei, P.; Huang, H.; Sun, X. Recent advances in the development of water oxidation electrocatalysts at mild pH. Small 2019, 15, 1805103. [Google Scholar] [CrossRef] [PubMed]
  6. Seh, Z.W.; Kibsgaard, J.; Dickens, C.F.; Chorkendorff, I.B.; Nørskov, J.K.; Jaramillo, T.F. Combining theory and experiment in electrocatalysis: Insights into materials design. Science 2017, 355, eaad4998. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Shi, Q.; Zhu, C.; Du, D.; Lin, Y. Robust noble metal-based electrocatalysts for oxygen evolution reaction. Chem. Soc. Rev. 2019, 48, 3181–3192. [Google Scholar] [CrossRef]
  8. Han, L.; Dong, S.; Wang, E. Transition-metal (Co, Ni, and Fe)-based electrocatalysts for the water oxidation reaction. Adv. Mater. 2016, 28, 9266–9291. [Google Scholar] [CrossRef]
  9. Burke, M.S.; Kast, M.G.; Trotochaud, L.; Smith, A.M.; Boettcher, S.W. Cobalt–iron (oxy) hydroxide oxygen evolution electrocatalysts: The role of structure and composition on activity, stability, and mechanism. J. Am. Chem. Soc. 2015, 137, 3638–3648. [Google Scholar] [CrossRef] [Green Version]
  10. Gong, M.; Dai, H. A mini review of NiFe-based materials as highly active oxygen evolution reaction electrocatalysts. Nano Res. 2015, 8, 23–39. [Google Scholar] [CrossRef] [Green Version]
  11. Lyu, F.; Wang, Q.; Choi, S.M.; Yin, Y. Noble-metal-free electrocatalysts for oxygen evolution. Small 2019, 15, 1804201. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Bo, X.; Dastafkan, K.; Zhao, C. Design of multi-metallic-based electrocatalysts for enhanced water oxidation. ChemPhysChem 2019, 20, 2936–2945. [Google Scholar] [CrossRef] [PubMed]
  13. Chen, G.; Zhu, Y.; Chen, H.M.; Hu, Z.; Hung, S.F.; Ma, N.; Shao, Z. An amorphous nickel–iron-based electrocatalyst with unusual local structures for ultrafast oxygen evolution reaction. Adv. Mater. 2019, 31, 1900883. [Google Scholar] [CrossRef]
  14. Zhu, W.; Zhu, G.; Yao, C.; Chen, H.; Hu, J.; Zhu, Y.; Liang, W. Porous amorphous FeCo alloys as pre-catalysts for promoting the oxygen evolution reaction. J. Alloys Compd. 2020, 828, 154465. [Google Scholar] [CrossRef]
  15. Liu, H.; Yang, D.H.; Wang, X.Y.; Zhang, J.; Han, B.H. N-doped graphitic carbon shell-encapsulated FeCo alloy derived from metal–polyphenol network and melamine sponge for oxygen reduction, oxygen evolution, and hydrogen evolution reactions in alkaline media. J. Colloid Interface Sci. 2021, 581, 362–373. [Google Scholar] [CrossRef]
  16. Mattioli, G.; Giannozzi, P.; Amore Bonapasta, A.; Guidoni, L. Reaction pathways for oxygen evolution promoted by cobalt catalyst. J. Am. Chem. Soc. 2013, 135, 15353–15363. [Google Scholar] [CrossRef] [Green Version]
  17. Wang, M.; Wang, Z.; Guo, Z. Water electrolysis enhanced by super gravity field for hydrogen production. Int. J. Hydrog. Energy 2010, 35, 3198–3205. [Google Scholar] [CrossRef]
  18. Liu, G.; Li, P.; Zhao, G.; Wang, X.; Kong, J.; Liu, H.; Ye, J. Promoting active species generation by plasmon-induced hot-electron excitation for efficient electrocatalytic oxygen evolution. J. Am. Chem. Soc. 2016, 138, 9128–9136. [Google Scholar] [CrossRef]
  19. Su, R.; Hsain, H.A.; Wu, M.; Zhang, D.; Hu, X.; Wang, Z.; Pennycook, S.J. Nano-Ferroelectric for High Efficiency Overall Water Splitting under Ultrasonic Vibration. Angew. Chem. Int. Ed. 2019, 58, 15076–15081. [Google Scholar] [CrossRef]
  20. Wang, J.; Yan, M.; Zhao, K.; Liao, X.; Wang, P.; Pan, X.; Mai, L. Field effect enhanced hydrogen evolution reaction of MoS2 nanosheets. Adv. Mater. 2017, 29, 1604464. [Google Scholar] [CrossRef]
  21. Garcés-Pineda, F.A.; Blasco-Ahicart, M.; Nieto-Castro, D.; López, N.; Galán-Mascarós, J.R. Direct magnetic enhancement of electrocatalytic water oxidation in alkaline media. Nat. Energy 2019, 4, 519–525. [Google Scholar] [CrossRef]
  22. Niether, C.; Faure, S.; Bordet, A.; Deseure, J.; Chatenet, M.; Carrey, J.; Rouet, A. Improved water electrolysis using magnetic heating of FeC–Ni core–shell nanoparticles. Nat. Energy 2018, 3, 476–483. [Google Scholar] [CrossRef]
  23. Gatard, V.; Deseure, J.; Chatenet, M. Use of magnetic fields in electrochemistry: A selected review. Curr. Opin. Electrochem. 2020, 23, 96–105. [Google Scholar] [CrossRef]
  24. Elias, L.; Hegde, A.C. Effect of magnetic field on HER of water electrolysis on Ni–W alloy. Electrocatalysis 2017, 8, 375–382. [Google Scholar] [CrossRef]
  25. Westsson, E.; Picken, S.; Koper, G. The Effect of Magnetic Field on Catalytic Properties in Core-Shell Type Particles. Front. Chem. 2020, 8, 163. [Google Scholar] [CrossRef]
  26. Li, Y.; Zhang, L.; Peng, J.; Zhang, W.; Peng, K. Magnetic field enhancing electrocatalysis of Co3O4/NF for oxygen evolution reaction. J. Power Sources 2019, 433, 226704. [Google Scholar] [CrossRef]
  27. Zeng, Z.; Zhang, T.; Liu, Y.; Zhang, W.; Yin, Z.; Ji, Z.; Wei, J. Magnetic Field Enhanced 4-electron Pathway of the Well-aligned Co3O4/ECNFs in the Oxygen Reduction Reaction. ChemSusChem 2018, 11, 580–588. [Google Scholar] [CrossRef]
  28. Hunter, D.; Osborn, W.; Wang, K.; Wang, K.; Kazantseva, N.; Hattrick-Simpers, J.; Suchoski, R.; Takeuchi, I. Giant magnetostriction in annealed Co1−xFex thin-films. Nat. Commun. 2011, 2, 1–7. [Google Scholar] [CrossRef] [Green Version]
  29. Han, Y.; Wang, H.; Zhang, T.; He, Y.; Coey, J.M.D.; Jiang, C. Tailoring the heterogeneous magnetostriction in Fe-Co alloys. J. Alloys Compd. 2017, 699, 200–209. [Google Scholar] [CrossRef]
  30. Xia, Z.; Guo, S. Strain engineering of metal-based nanomaterials for energy electrocatalysis. Chem. Soc. Rev. 2019, 48, 3265–3278. [Google Scholar] [CrossRef]
  31. Luo, M.; Guo, S. Strain-controlled electrocatalysis on multimetallic nanomaterials. Nat. Rev. Mater. 2017, 2, 1–13. [Google Scholar] [CrossRef]
  32. Kodama, D.; Shinoda, K.; Sato, K.; Konno, Y.; Joseyphus, R.J.; Motomiya, K.; Jeyadevan, B. Chemical synthesis of sub-micrometer-to nanometer-sized magnetic FeCo dice. Adv. Mater. 2006, 18, 3154–3159. [Google Scholar] [CrossRef]
  33. Szumiata, T.; Gzik-Szumiata, M.; Brzózka, K.; Górka, B.; Gawroński, M.; Finkel, A.C.; Morley, N.A. Interface influence on the properties of Co90Fe10 films on soft magnetic underlayers–Magnetostrictive and Mössbauer spectrometry studies. J. Magn. Magn. Mater. 2016, 401, 943–948. [Google Scholar] [CrossRef]
  34. Brzózka, K.; Szumiata, T.; Olekšáková, D.; Kollár, P. Structural and Magnetic Properties of Mechanically Alloyed Fe50Co50 Systems. Acta Phys. Pol. A. 2020, 137, 726–729. [Google Scholar] [CrossRef]
  35. Wu, L.Z.; Ding, J.; Jiang, H.B.; Chen, L.F.; Ong, C.K. Particle size influence to the microwave properties of iron based magnetic particulate composites. J. Magn. Magn. Mater. 2005, 285, 233–239. [Google Scholar] [CrossRef]
  36. Song, J.; Lv, X.; Jiao, Y.; Wang, P.; Xu, M.; Li, T.; Zhang, Z. Catalyst nanoarchitecturing via functionally implanted cobalt nanoparticles in nitrogen doped carbon host for aprotic lithium-oxygen batteries. J. Power Sources 2018, 394, 122–130. [Google Scholar] [CrossRef]
  37. Palaniselvam, T.; Kashyap, V.; Bhange, S.N.; Baek, J.B.; Kurungot, S. Nanoporous graphene enriched with Fe/Co-N active sites as a promising oxygen reduction electrocatalyst for anion exchange membrane fuel cells. Adv. Funct. Mater. 2016, 26, 2150–2162. [Google Scholar] [CrossRef]
  38. Xiao, X.; Li, X.; Yu, G.; Wang, J.; Yan, G.; Wang, Z.; Guo, H. FeCox alloy nanoparticles encapsulated in three-dimensionally N-doped porous carbon/multiwalled carbon nanotubes composites as bifunctional electrocatalyst for zinc-air battery. J. Power Sources 2019, 438, 227019. [Google Scholar] [CrossRef]
  39. Liu, Y.; Pan, L.; Liu, H.; Chen, T.; Yin, S.; Liu, M. Effects of magnetic field on water electrolysis using foam electrodes. Int. J. Hydrog. Energy 2019, 44, 1352–1358. [Google Scholar] [CrossRef]
  40. Iida, T.; Matsushima, H.; Fukunaka, Y. Water electrolysis under a magnetic field. J. Electrochem. Soc. 2007, 154, E112. [Google Scholar] [CrossRef]
  41. Ma, T.Y.; Cao, J.L.; Jaroniec, M.; Qiao, S.Z. Interacting carbon nitride and titanium carbide nanosheets for high-performance oxygen evolution. Angew. Chem. 2016, 128, 1150–1154. [Google Scholar] [CrossRef]
  42. Ahsan, M.A.; Puente Santiago, A.R.; Hong, Y.; Zhang, N.; Cano, M.; Rodriguez-Castellon, E.; Noveron, J.C. Tuning of trifunctional NiCu bimetallic nanoparticles confined in a porous carbon network with surface composition and local structural distortions for the electrocatalytic oxygen reduction, oxygen and hydrogen evolution reactions. J. Am. Chem. Soc. 2020, 142, 14688–14701. [Google Scholar] [CrossRef] [PubMed]
  43. Zhou, W.; Chen, M.; Guo, M.; Hong, A.; Yu, T.; Luo, X.; Wang, S. Magnetic enhancement for hydrogen evolution reaction on ferromagnetic MoS2 catalyst. Nano Lett. 2020, 20, 2923–2930. [Google Scholar] [CrossRef] [PubMed]
  44. Gupta, U.; Rajamathi, C.R.; Kumar, N.; Li, G.; Sun, Y.; Shekhar, C.; Rao, C.N.R. Effect of magnetic field on the hydrogen evolution activity using non-magnetic Weyl semimetal catalysts. Dalton Trans. 2020, 49, 3398–3402. [Google Scholar] [CrossRef] [PubMed]
  45. Ren, X.; Wu, T.; Sun, Y.; Li, Y.; Xian, G.; Liu, X.; Xu, Z.J. Spin-polarized oxygen evolution reaction under magnetic field. Nat. Commun. 2021, 12, 1–12. [Google Scholar] [CrossRef] [PubMed]
  46. Yan, J.; Wang, Y.; Zhang, Y.; Xia, S.; Yu, J.; Ding, B. Direct magnetic reinforcement of electrocatalytic ORR/OER with electromagnetic induction of magnetic catalysts. Adv. Mater. 2021, 33, 2007525. [Google Scholar] [CrossRef]
  47. Wang, L.; Yang, H.; Yang, J.; Yang, Y.; Wang, R.; Li, S.; Ji, S. The effect of the internal magnetism of ferromagnetic catalysts on their catalytic activity toward oxygen reduction reaction under an external magnetic field. Ionics 2016, 22, 2195–2202. [Google Scholar] [CrossRef] [Green Version]
  48. Monzon, L.M.A.; Rode, K.; Venkatesan, M.; Coey, J.M.D. Electrosynthesis of iron, cobalt, and zinc microcrystals and magnetic enhancement of the oxygen reduction reaction. Chem. Mater. 2012, 24, 3878–3885. [Google Scholar] [CrossRef]
  49. Zhou, D.; Wang, S.; Jia, Y.; Xiong, X.; Yang, H.; Liu, S.; Liu, B. NiFe hydroxide lattice tensile strain: Enhancement of adsorption of oxygenated intermediates for efficient water oxidation catalysis. Angew. Chem. Int. Ed. 2019, 58, 736–740. [Google Scholar] [CrossRef]
  50. Han, M.; Li, S.; Li, C.; Wu, J.; Han, J.; Wang, N.; Liang, H. Strain-modulated Ni3Al alloy promotes oxygen evolution reaction. J. Alloys Compd. 2020, 844, 156094. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of samples Fe35Co65, Fe22Co78, Fe15Co85.
Figure 1. XRD patterns of samples Fe35Co65, Fe22Co78, Fe15Co85.
Metals 12 00800 g001
Figure 2. SEM image and EDS spectrum of Fe35Co65 (a,d), Fe22Co78 (b,e), Fe15Co85 (c,f).
Figure 2. SEM image and EDS spectrum of Fe35Co65 (a,d), Fe22Co78 (b,e), Fe15Co85 (c,f).
Metals 12 00800 g002
Figure 3. XPS spectra of (a) Co 2p and (b) Fe 2p of Fe35Co65 sample.
Figure 3. XPS spectra of (a) Co 2p and (b) Fe 2p of Fe35Co65 sample.
Metals 12 00800 g003
Figure 4. Hysteresis loop of Fe35Co65 (a), Fe22Co78 (b), Fe15Co85 (c) and Cu foam (d).
Figure 4. Hysteresis loop of Fe35Co65 (a), Fe22Co78 (b), Fe15Co85 (c) and Cu foam (d).
Metals 12 00800 g004
Figure 5. Magnetostriction curves of Fe35Co65, Fe22Co78, Fe15Co85.
Figure 5. Magnetostriction curves of Fe35Co65, Fe22Co78, Fe15Co85.
Metals 12 00800 g005
Figure 6. LSV curves of Fe35Co65 (a), Fe22Co78 (b) and Fe15Co85 (c) in the presence and absence of magnetic field; correspondence between strain and overpotential change (d).
Figure 6. LSV curves of Fe35Co65 (a), Fe22Co78 (b) and Fe15Co85 (c) in the presence and absence of magnetic field; correspondence between strain and overpotential change (d).
Metals 12 00800 g006
Figure 7. LSV curve of sample Fe35Co65 under different magnetic field gradients.
Figure 7. LSV curve of sample Fe35Co65 under different magnetic field gradients.
Metals 12 00800 g007
Figure 8. (a) Variation diagram of strain and overpotential of Fe100−xCox alloys under a magnetic field of 700 mT; (b) variation diagram of strain and overpotential of Fe35Co65 alloy under different magnetic field intensities.
Figure 8. (a) Variation diagram of strain and overpotential of Fe100−xCox alloys under a magnetic field of 700 mT; (b) variation diagram of strain and overpotential of Fe35Co65 alloy under different magnetic field intensities.
Metals 12 00800 g008
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, H.; Ren, Y.; Wang, K.; Mu, X.; Song, S.; Guo, J.; Yang, X.; Lu, Z. Magnetic-Field-Induced Strain Enhances Electrocatalysis of FeCo Alloys on Anode Catalysts for Water Splitting. Metals 2022, 12, 800. https://doi.org/10.3390/met12050800

AMA Style

Liu H, Ren Y, Wang K, Mu X, Song S, Guo J, Yang X, Lu Z. Magnetic-Field-Induced Strain Enhances Electrocatalysis of FeCo Alloys on Anode Catalysts for Water Splitting. Metals. 2022; 12(5):800. https://doi.org/10.3390/met12050800

Chicago/Turabian Style

Liu, Heyan, Yanwei Ren, Kai Wang, Xiaoming Mu, Shihao Song, Jia Guo, Xiaojing Yang, and Zunming Lu. 2022. "Magnetic-Field-Induced Strain Enhances Electrocatalysis of FeCo Alloys on Anode Catalysts for Water Splitting" Metals 12, no. 5: 800. https://doi.org/10.3390/met12050800

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