In-situ selective oxidation created Cr2O3 assisting CrMnFeCoNi for ultrahigh power density zinc-air batteries
Abstract
Highly efficient and stable bifunctional catalysts toward sluggish oxygen reduction reaction (ORR) and oxygen evolution reaction (OER) are critical for practical applications of rechargeable zinc-air batteries (ZAB). Heterostructure engineering is effective in boosting catalytic performances of the bifunctional catalysts. Here,
Keywords
INTRODUCTION
Rechargeable zinc-air batteries (ZABs) are considered a pivotal advancement in electrochemical energy storage technologies because of their high energy densities, environmental friendliness, and renewability. Nevertheless, the advancement of rechargeable ZABs is currently constrained by several technical challenges, with low catalytic activities and stability toward catalysis of sluggish oxygen reduction reaction (ORR) at discharging and oxygen evolution reaction (OER) at charging being the most critical. Although noble metal-based catalysts, for example, Pt/C for ORR and RuO2 for OER, exhibit superior performances, their high cost, extremely limited availability, and mono-functionality toward catalysis of oxygen reactions, restrict their practical applications for rechargeable ZABs. Consequently, development of highly efficient, durable, and cost-effective bifunctional catalysts is essential for large scale applications of rechargeable ZABs.
Among the many previously explored electrocatalysts, transition metals, such as Mn, Fe, Co, and Ni, based materials are recognized for their low cost and abundant availability. When supported on N-doped carbons, these materials exhibited promising catalytic activities and stability toward both ORR and OER[1,2]. N-doped carbons not only offer necessary electrical conductivities to facilitate relevant electrochemical processes, but also effectively tackle several detrimental issues, including catalyst corrosion, aggregation, and detachment under severe reaction conditions[3-5].
For catalyst design, interface engineering is an effective strategy to enhance activities and stability of the catalysts. One common approach in interface engineering is the formation of heterostructures, where two or more different materials are combined to form interfaces of a microscopic scale[6]. Charge transfer, occurring across the interface, helps regulate charge distribution near active sites and adsorption energies of reaction intermediates, thereby enhancing catalytic activities. Furthermore, formation of heterostructures can create a large number of defect sites, through lattice mismatch-induced local strains, on and around the interfaces, thus increasing the number of active sites on the catalyst surface[7,8], beneficial for activity enhancement.
Metal-metal oxide heterostructures have been shown to modulate binding strength between intermediates and catalysts because of the strong electronic interactions occurring at the interface[9]. Zhang et al.[10] synthesized Co-CoO heterojunctions embedded in N-doped carbon nanospheres as the air electrode catalyst for rechargeable ZABs, achieving a potential gap (ΔE) of 0.83 V. Li et al.[11] utilized a thermal decomposition method with dopamine precursors to prepare Co-MnO heterostructures for catalysis of ORR. It was found that MnO, upon forming a heterojunction with Co, exhibited a Mott-Schottky effect at the interface, modulating the charge distribution and thereby enhancing ORR activities. Luo et al.[12] fabricated a multi-element heterostructure, FeNiCo-MnGaOx, for applications in rechargeable ZABs. The heterostructured catalyst demonstrated excellent electrochemical performances with a half-wave
Here, a multi-element alloy-based alloy-oxide heterostructured catalyst, CrMnFeCoNi-CrOx, was developed to serve as a highly active and stable bifunctional catalyst for the air electrode of a high-performance rechargeable ZAB. The multi-element material system of CrMnFeCoNi has been explored as bifunctional oxygen catalysts for ZABs[13] and as electrocatalysts for OER[14], exhibiting promising electrocatalytic performances. Introduction of heterostructure to the CrMnFeCoNi system is a rational extension to gain further improvements. The CrMnFeCoNi-CrOx heterostructure was created through thermal reduction of a multi-element Prussian blue analogue [PBA, M1xM2y(CN)6] precursor, CrMnFeCoNi PBA, in a reducing atmosphere, 5%H2/95%Ar, in the presence of a dispersing support, carbon black (CB). CB, coated on skeleton surfaces of a conductive porous substrate, nickel foam (NF), was introduced to disperse the growth of the PBA and thus the final PBA-derived N-doped carbon supported multi-element alloy catalyst. The presence of CB also enabled in-situ selective oxidation during the subsequent thermal reduction treatment. Upon thermal treatment in a reducing atmosphere, metal ions and CN groups of the PBA were reduced to form N-doped carbon supported multi-element alloy nanoparticles dispersed on the CB layer. The
RESULTS AND DISCUSSION
Materials characterizations
Scheme 1 illustrates the fabrication of the N-doped carbon supported heterostructured catalyst, CrMnFeCoNi-(CrOx)n/NC. Detailed experimental procedures, including precursor formulas for syntheses of CrMnFeCoNi-(CrOx)n/NC [Supplementary Table 1] and mass loading of the sample catalysts
Scheme 1. Schematic illustration of fabrication of CrMnFeCoNi-(CrOx)n/NC. PBA: Prussian blue analogue.
Figure 1A and Supplementary Figure 2 show the XRD patterns of the three heterostructured catalysts, CrMnFeCoNi-(CrOx)n/NC (n=0.5, 1, 3), and the two control sample catalysts, MnFeCoNi/NC and
Figure 1. (A) XRD patterns of MnFeCoNi/NC and CrMnFeCoNi-(CrOx)1/NC; (B and C) SEM images of CrMnFeCoNi-(CrOx)1/NC on CB@NF; (D) HR-TEM image and (E) HAADF-STEM image of CrMnFeCoNi-(CrOx)1/NC nanoparticle; (F) TEM-EDS elemental mapping of CrMnFeCoNi-(CrOx)1/NC nanoparticle; (G and H) TEM elemental line scan of CrMnFeCoNi-(CrOx)1/NC nanoparticle. XRD: X-ray diffraction; HR-TEM: high-resolution-transmission electron microscopy; HAADF-STEM: high-angle annular dark-field scanning transmission electron microscopy; TEM-EDS: transmission electron microscopy-energy dispersive X-ray spectroscopy; SEM: scanning electron microscope.
The morphologies of these sample catalysts were next investigated with scanning electron microscope (SEM) imaging. Figure 1B and C shows the SEM images of the main catalyst, CrMnFeCoNi-(CrOx)1/NC, loaded on the substrate CB@NF, at increasing magnifications. At a low magnification [Figure 1B], a segment of the skeleton of the NF is observed, which is uniformly coated with a CB layer, and on top of the CB layer numerous nanoparticles reside, appearing as tiny white dots in Figure 1B. When zoomed in to a high magnification [Figure 1C], these nanoparticles, with an average size of 530±130 nm, appear to be decorated with even smaller-sized nanoparticles (6.0±2.2 nm), designated by yellow arrows. There are also some sheet structures observed around these nanoparticles. These sheet structures are later identified as
TEM imaging analyses were conducted on CrMnFeCoNi-(CrOx)1/NC [Figure 1D and E, Supplementary Figure 4A and B] and CrOx/NC [Supplementary Figure 4C] nanoparticles to examine their detailed compositions. Figure 1D and E shows high-resolution (HR)-TEM and high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) images of a CrMnFeCoNi-(CrOx)1/NC composite nanoparticle, revealing attachment of a smaller-size nanoparticle, highlighted with a red dashed square, to the surface of a larger nanoparticle [Figure 1D] and the existence of an interface between the two nanoparticles [Figure 1E], respectively. Distinct lattice fringes were observed on the two sides of the interface. An interlayer distance of 0.259 nm, in good agreement with the d-spacing of the (110) planes of Cr2O3, was determined for the attachment domain, indicating the attached smaller-size nanoparticle to be a Cr2O3 nanoparticle. In addition, interlayer distances of 0.176 nm and 0.203 nm, in good agreement with the d-spacing of the (200) and (111) planes of an FCC phase, respectively, were determined for the larger size nanoparticle domain. The two crystalline planes have a dihedral angle of 54.6°, closely matching the theoretical value of 54.7° for the dihedral angle between the (200) and (111) planes of an FCC phase, further confirming the nanoparticle to be a single FCC phase multi-element alloy nanoparticle. In addition to Cr2O3 decoration, the CrMnFeCoNi alloy nanoparticle is also coated with a thin layer of poorly crystallized
The composition of CrMnFeCoNi-(CrOx)1 was further examined with XAS, which offers information on coordination environments of atoms to reveal constituent components of the material. For comparison purposes, XAS spectra of MnFeCoNi/NC were also collected. Figure 2A-E shows the K-edge extended
Figure 2. K-edge EXAFS spectra of MnFeCoNi/NC and CrMnFeCoNi-(CrOx)1/NC in R space: (A) Cr, (B) Mn, (C) Fe, (D) Co, and (E) Ni; (F) Cr pre-edge and (G) EXAFS spectra of CrMnFeCoNi-(CrOx)1/NC, Cr foil, and CrO3 in R space; (H) WT-EXAFS k-R contour plots of CrMnFeCoNi-(CrOx)1/NC, Cr foil, and Cr2O3. EXAFS: Extended X-ray absorption fine structure; WT-EXAFS: wavelet transform-EXAFS.
The distinct coordination environment of Cr in CrMnFeCoNi-(CrOx)1 is further discussed with a more detailed analysis of the XAS spectra. Supplementary Figure 5 shows the K-edge X-ray absorption near edge structure (XANES) spectrum of Cr in CrMnFeCoNi-(CrOx)1/NC, along with those of Cr metal foil and
To further elucidate the coordination state of Cr in CrMnFeCoNi-(CrOx)1/NC, Cr K-edge EXAFS spectra of CrMnFeCoNi-(CrOx)1/NC, Cr foil, Cr2O3 standard, and CrO3 standard were compared as shown in Figure 2G. CrMnFeCoNi-(CrOx)1/NC shows two major coordination shells. The first coordination shell is located at 1.5 Å, contributed by the Cr-O coordination, same as the first coordination shell of
Recall that, in the TEM elemental line scan, a small amount of Mn was detected in the CrOx
Surface chemical states of CrMnFeCoNi-(CrOx)1/NC and MnFeCoNi/NC were also examined with
Electrochemical characterizations
Electrochemical properties of the five sample catalysts, two control samples, and three heterostructured samples, along with the benchmark catalysts, Pt/C for ORR and RuO2 for OER, toward ORR and OER were characterized in 0.1 M KOH. Supplementary Figure 7 shows the cyclic voltammetry (CV) curves recorded for the five sample catalysts under the conditions of both O2 and N2 saturation to examine their capabilities toward catalysis of ORR. Evidently, reduction peaks appear only in O2-saturated electrolytes, implying occurrence of ORR and confirming ORR-activeness of these catalysts. Among all characterized samples, CrMnFeCoNi-(CrOx)1/NC shows the highest oxygen reduction peak potential of 0.816 V, thus the lowest ORR overpotential, indicating its superior efficiency in reducing oxygen.
The ORR activities of these sample catalysts, along with the benchmark ORR catalyst, Pt/C, were characterized by half-wave potentials (E1/2) derived from the polarization curves recorded in 0.1 M KOH
Figure 4. Electrochemical performances of CrOx/NC, MnFeCoNi/NC, CrMnFeCoNi-(CrOx)1/NC, and benchmark catalysts in 0.1 M KOH: (A) ORR LSV curves recorded at 1600 rpm with scan rate of 5 mV s-1; (B) ORR Tafel slopes; (C) ORR electron transfer numbers (n) and hydrogen peroxide yields (H2O2%); (D) OER LSV curves at scan rate of 1 mV s-1; (E) OER Tafel slopes; (F) potential gap (ΔE) between η10 of OER and E1/2 of ORR. ORR: Oxygen reduction reaction; LSV: linear sweep voltammetry; OER: oxygen evolution reaction.
Figure 4D and E, Supplementary Figure 11 display the OER polarization curves and corresponding Tafel slopes. It is to be noted that the LSV for OER [Figure 4D] was conducted in a negative scan mode to avoid possible interference from pre-OER oxidation toward determination of overpotentials at 10 mA cm-2 (η10). For metallic alloy samples, the surface oxide layer of high-valence metals formed at high applied potentials is reduced to oxides of low-valence metals at low applied potentials, leading to the showing of the reduction peak observed. Among the tested catalysts, CrOx/NC shows the poorest performance. The formation of the CrMnFeCoNi-CrOx heterostructure significantly reduces the OER overpotential, with CrMnFeCoNi-(CrOx)1/NC performing the best, achieving the lowest overpotential (273 mV) at 10 mA cm-2 (η10) and Tafel slope (59.4 mV dec-1), even lower than those of the benchmark catalyst, RuO2 (284 mV and 80.1 mV dec-1). The superiority of the heterostructured catalyst becomes more pronounced with increasing applied potentials [Figure 4D], beneficial for practical operations at high current densities.
Electric double layer capacitances (Cdl’s), a direct measure of electrochemical surface areas (ECSAs), were also determined to normalize the LSV curves for examination of the intrinsic activities of the catalysts. To determine Cdl, CV curves [Supplementary Figure 12] recorded within a non-Faradaic potential window at increasing scan rates were collected, from which Cdl’s were calculated as halves of the slopes of the current density difference versus scan rate plots, obtained through linear regression. Interestingly, all multi-element alloy-containing samples exhibit much higher Cdl’s (> 12 mF cm-2) than CrOx/NC (6.2 mF cm-2), correlating well with their morphological characteristics. Recall that CrOx/NC has a morphology of sparsely distributed small-size nanoparticles decorated on CB nanoparticles, whereas the multi-element alloy-containing samples have a morphology of densely populated nanoparticles, with/without decoration of small-size nanoparticles, uniformly distributed on top of the CB layer, giving higher amounts of exposed surface areas of the catalysts. These Cdl’s were used to normalize the OER LSV curves [Supplementary Figure 13] to exclude the effect of electrochemical surface areas, thus quantity of active sites, from the LSV curves to reveal the intrinsic activities of the catalysts. Evidently, the main catalyst, CrMnFeCoNi-(CrOx)1/NC, exhibits the lowest overpotentials, indicating its highest intrinsic activities. Furthermore, the charge transfer resistances involved in OER catalyzed by the product catalysts were examined with EIS [Supplementary Figure 10B]. As evident from Supplementary Figure 10B, CrMnFeCoNi-CrOx/NC exhibits a significantly lower Rct of 8.0 Ω than that of MnFeCoNi/NC (12.2 Ω), confirming the merit of the alloy-oxide heterostructure in accelerating the OER kinetics.
Efficiency of the bifunctionality of the sample catalysts toward catalysis of ORR and OER was
Catalyst robustness, critical to practical applications of ZABs, was also characterized by a 20 h chronoamperometric test as shown in Supplementary Figure 14. For both ORR and OER, CrMnFeCoNi-(CrOx)1/NC exhibits only 4.3 and 5.0% decay, respectively. As for the two benchmark catalysts, Pt/C and RuO2 experience severer decay of 7.3% for ORR and 17.1% for OER, respectively. This indicates that CrMnFeCoNi-(CrOx)1/NC possesses not only superior bifunctional catalytic activities, but also superior operational stability.
Zinc-air battery characterizations
Electrochemical characterizations on ZABs, constructed by using the present developed catalysts as the catalysts for the air electrode, were conducted to further show the merits of the alloy-oxide heterostructured catalysts on ZABs. Figure 5A and B, Supplementary Figure 15A and B show the discharge-charge curves and corresponding power density curves of the two control and the three heterostructured sample
Figure 5. Electrochemical characterizations of CrOx/NC, MnFeCoNi/NC, CrMnFeCoNi-(CrOx)1/NC, and (Pt/C+RuO2)-based ZABs: (A) discharge-charge polarization curves and (B) corresponding power density curves at current ramping rate of 1 mA s-1; (C) ORR LSV curves, (D) OER LSV curves, and (E) potential gaps (△E10 and △E100) of CrMnFeCoNi-(CrOx)1/NC and (Pt/C+RuO2)-based air electrodes in 6.0 M KOH at scan rate of 1 mV s-1; (F) Step voltage diagrams of CrOx/NC, MnFeCoNi/NC, CrMnFeCoNi-(CrOx)1/NC, and (Pt/C+RuO2)-based ZABs working at varying charge/discharge current densities from 10 to 50 and then back to 10 mA cm-2; (G) Specific capacities achieved by CrOx/NC, MnFeCoNi/NC, CrMnFeCoNi-(CrOx)1/NC, and (Pt/C+RuO2)-based ZABs at 50 mA cm-2; (H) Galvanostatic cycling discharge-charge stability tests of CrMnFeCoNi-(CrOx)1/NC and (Pt/C+RuO2)-based ZABs at 10 mA cm-2 with duration of 20 min per cycle (10 min discharge + 10 min charge). ORR: Oxygen reduction reaction; LSV: linear sweep voltammetry; OER: oxygen evolution reaction; ZABs: zinc-air batteries.
In this study, the air electrode was fabricated by pressing firmly a gas diffusion carbon paper layer with a porous catalyst-loaded NF layer to form a composite binder-free air electrode. Unlike the traditional binder assisted catalyst ink drop-casting process, the catalyst in this work was grown in-situ on the skeleton surface of NF, thus being binder-free. To examine the merit of the binder-free approach in loading catalysts, the discharge-charge polarization curve and the corresponding discharge power density curve of the drop-cast CrMnFeCoNi-(CrOx)1/NC-based ZAB were collected for comparison with those achieved by the CrMnFeCoNi-(CrOx)1/NC and (Pt/C+RuO2)-based ZABs [Supplementary Figure 15C and D]. For the
Electrochemical performances of the sample catalysts toward ORR and OER were further characterized
Figure 5F and Supplementary Figure 17A present the step voltage diagrams for each sample ZAB working at varying charge/discharge current densities from 10 to 50 and then back to 10 mA cm-2. The CrMnFeCoNi-(CrOx)1/NC-based ZAB consistently exhibits superior charge/discharge platforms and narrower voltage gaps, as compared to those of the MnFeCoNi/NC, CrOx/NC, and even (Pt/C+RuO2)-based ones, with the superiority turning more pronounced at increasing working current densities. The voltage recovery rates of the tested ZABs at the end of the tests are summarized in Supplementary Table 6 for comparison. The CrMnFeCoNi-(CrOx)1/NC-based ZAB exhibits the highest voltage recovery rates, 99.99% for the discharging and 99.95% for the charging. The above results indicate that the CrMnFeCoNi-(CrOx)1/NC-based ZAB achieved not only outstanding charge/discharge efficiency but also outstanding voltage recovery characteristics, relevant to enhanced stability, high reversibility, and sustained catalytic performances. Furthermore, at a working current density of 50 mA cm-2, the CrMnFeCoNi-(CrOx)1/NC-based ZAB achieves a specific capacity of 814 mAh gZn-1 and an energy density of 918 Wh kgZn-1, surpassing
Finally, the durability of the ZABs was assessed with long-term cycling operations. Figure 5H shows the cycling discharge-charge profiles of the CrMnFeCoNi-(CrOx)1/NC and (Pt/C+RuO2)-based ZABs at a current density of 10 mA cm-2 with a duration of 20 min per cycle. Initially, the CrMnFeCoNi-(CrOx)1/NC-based ZAB exhibits charging and discharging plateaus of 1.98 and 1.20 V, respectively, giving a voltage gap of 0.78 V. In comparison, the (Pt/C+RuO2)-based ZAB shows slightly inferior initial charging and discharging plateaus of 2.00 and 1.18 V, respectively with a voltage gap of
In-situ spectroscopic analyses
In-situ Raman spectroscopy and in-situ XAS were conducted to gain insights on the mechanisms of ORR and OER catalyzed by CrMnFeCoNi-(CrOx)1/NC. Figure 6 shows the in-situ XANES and EXAFS spectra and in-situ Raman spectra collected for CrMnFeCoNi-(CrOx)1/NC toward catalysis of ORR. Figure 6A-E presents the in-situ XANES spectra of constituent metal elements of CrMnFeCoNi-(CrOx)1/NC at
Figure 6. In-situ XANES spectra for ORR catalyzed by CrMnFeCoNi-(CrOx)1/NC at (A) Cr K-edge, (B) Mn K-edge, (C) Fe K-edge, (D) Co K-edge, and (E) Ni K-edge; Corresponding EXAFS spectra in R space (k=2) for (F) Cr K-edge, (G) Mn K-edge, (H) Fe K-edge, (I) Co K-edge, and (J) Ni K-edge; (K) In-situ Raman spectra of ORR catalyzed by CrMnFeCoNi-(CrOx)1/NC. XANES: X-ray absorption near edge structure; ORR: oxygen reduction reaction; EXAFS: extended X-ray absorption fine structure.
Figure 6K presents in-situ Raman spectra of CrMnFeCoNi-(CrOx)1/NC under varying applied potentials. Initially, no signals other than the D and G bands of the N-doped carbon supports are observed for the dry samples. Upon contact with the electrolyte, the spectra remain unchanged at OCP and the applied potential of 0.9 V (vs. RHE). Further lowering the applied potential to 0.8 V, corresponding to an increase in reducing powers, a pronounced peak appears at 1,056 cm-1, attributable to vibrations of superoxo (O2-) intermediates[25,26], indicating proceeding of the ORR. The intensity of this characteristic peak increases with further decreasing applied potentials (i.e., increasing reducing powers), reaching its maximum at 0.6 V, indicating increasing extent of ORR. The intensity of this characteristic peak slightly decreases with further decreasing applied potentials from 0.6 V to 0.4 V and drops significantly upon returning to the OCP. The observed variations in peak intensity can be correlated to the coverage of the superoxo intermediate on the catalyst surface. At high applied potentials (i.e., low reducing powers), the catalyst surface is only partially covered by the superoxo intermediate. As the applied potential decreases (i.e., increasing reducing powers), more superoxo intermediates adsorb onto the catalyst surface, resulting in an increase in peak intensity. Subsequently, at even lower applied potentials (i.e., even higher reducing powers), part of the superoxo intermediates are successfully reduced to hydroxide ions, leading to a slight decrease in peak intensity. Upon ceasing application of the potential, the superoxo intermediates largely desorb from the catalyst surface, resulting in a sharp drop in intensity of the characteristic peak.
Figure 7 presents the in-situ XANES and EXAFS spectra and in-situ Raman spectra collected
Figure 7. In-situ XANES spectra for OER catalyzed by CrMnFeCoNi-(CrOx)1/NC at (A) Cr K-edge, (B) Mn K-edge, (C) Fe K-edge, (D) Co K-edge, and (E) Ni K-edge. Corresponding EXAFS spectra in R space (k=2): (F) Cr K-edge, (G) Mn K-edge, (H) Fe K-edge, (I) Co
Furthermore, the K-edge XANES spectra of Mn, Co, and Ni (Figure 7B, D and E, respectively) show significant shifts in absorption edge positions to higher energies under anodic conditions, indicating slight oxidation of Mn, Co, and Ni of CrMnFeCoNi-(CrOx)1/NC during OER. Figure 7F-J depicts the K-edge EXAFS spectra in R-space for constituent metal elements of CrMnFeCoNi-(CrOx)1/NC under anodic conditions. Under the applied anodic bias, both the first and second coordination shells of Cr exhibit a decrease in peak intensity. The intensity reduction of the first shell located at 1.5 Å is attributed to the transformation of part of the octahedral Cr2O3 to tetrahedral CrO3 during OER. For CrO3, the number of oxygen atoms coordinating with Cr (4 Cr-O) is less than that in Cr2O3 (6 Cr-O), thereby reducing the intensity of the first coordination shell. Furthermore, in the tetrahedral crystalline structure (CrO3), Cr and O are more loosely packed than in the octahedral crystalline structure (Cr2O3), and thus partial conversion of Cr2O3 to CrO3 results in a decrease in intensity of the second coordination shell (Cr-O-Cr). The above observations conclude that partial conversion of Cr2O3 to CrO3 occurs during OER[27]. In addition, it has been proposed that oxidation of Cr2O3 protects β-NiOOH, a highly OER-active intermediate, from being overoxidized into γ-NiOOH, a less OER-active species, during the OER process[28,29]. For Mn, Co, and Ni, their K-edge EXAFS spectra in R-space exhibit intensity reductions for the metal-metal coordination shell located at 2.2 Å upon application of the anodic potential. For Mn, its Mn-O coordination shell located at
In-situ Raman spectra recorded for CrMnFeCoNi-(CrOx)1/NC at varying applied anodic potentials were used to investigate potential intermediates involved in the OER process [Figure 7K]. In the dry state and at low anodic potentials, no characteristic peaks other than the D and G bands of the N-doped carbon support are observed. Upon elevating the anodic potential to 1.43 V (vs. RHE), weak characteristic peaks emerge at 455 and 549 cm-1, which turn more pronounced with increasing anodic potentials. These peaks are attributed to Ni/Co oxyhydroxides (NiOOH and CoOOH), which are well-recognized as favorable intermediates for OER[30-32]. As the anodic potential increases to 1.58 V, an additional broad weak peak appears in the range of 870-1,180 cm-1, attributable to the vibration of superoxide intermediates (M-O2) interacting laterally with metal sites. These superoxide intermediates, formed from oxyhydroxides, can only be observed at high enough anodic potentials[30,33]. Upon ceasing application of the anodic potential, characteristic peaks for Ni/CoOOH and corresponding superoxides almost disappear, whereas peaks at around 570 and 798 cm-1, attributable to MnO2[30,34], emerge. The driving force for formation of Ni/CoOOH and corresponding superoxides disappears upon returning the bias to the OCP, leading to desorption of oxygen-containing species from the catalyst surface and thus disappearance of the associated peaks. As to the emergence of characteristic peaks associated with MnO2, it can be explained as follows. The concentration of Mn in CrMnFeCoNi-(CrOx)1/NC is quite low, and thus the concentration of MnO2 derived from Mn is even lower. Consequently, the peaks associated with MnO2 cannot be observed under the relatively strong showing of the peaks associated with Ni/CoOOH. These peaks however become pronounced upon disappearance of the Ni/CoOOH peaks. The findings concluded from the in-situ Raman study align well with those from the in-situ XAS investigation, revealing that Mn, Co, and Ni serve as the primary active sites for OER.
CONCLUSION
A non-precious quinary alloy-metal oxide heterostructured catalyst, CrMnFeCoNi-(CrOx)1/NC, was developed as an outstanding bifunctional catalyst for a binder-free composite air electrode, based on which a high-performance ZAB was fabricated. The heterostructured catalyst showed outstanding catalytic efficiency toward catalysis of ORR and OER, achieving a high E1/2 of 0.834 V for ORR and a low η10 of
In-situ Raman and XAS spectroscopy studies revealed that Mn, Fe, and Co served as the primary ORR active sites, and Mn, Co, and Ni worked as the main active sites for catalysis of OER. The presence of Cr2O3 helped boost the ORR and OER performances by offering abundant oxygen vacancies to enhance oxygen adsorption for ORR and by protecting OER-active intermediates from over-oxidation through its
DECLARATIONS
Authors’ contributions
Data curation, formal analysis, investigation, methodology, software, validation, writing - original draft: Yen, F. Y.
Data curation, formal analysis, investigation: Chang, S. I.; Ting, Y. C.; Chang, C. W.; Lee, K. A.
Conceptualization, funding acquisition, project administration, supervision, validation, visualization, writing - review and editing: Lu, S. Y.
Availability of data and materials
The detailed materials and methods in the experiment are available within the Supplementary Materials. Further data are available from the corresponding author upon reasonable request.
Financial support and sponsorship
This research was financially supported by the National Science and Technology Council (NSTC) of Taiwan, ROC, under grant number MOST 111-2221-E-007-008-MY3 (SYL). The authors express their deep gratitude for the beamtimes of TLS 17C1, TLS 16A1, and TLS 01C1 provided by the National Synchrotron Radiation Research Center (NSRRC) of Taiwan, ROC. The authors also sincerely acknowledge the use of the spherical-aberration corrected field emission TEM (JEM-ARM200FTH, JEOL Ltd.) and HR-XPS (PHI Quantera II, ULVAC-PHI Inc.) facilities at the Instrument Center of National Tsing Hua University, Taiwan, ROC.
Conflicts of interest
All authors declared that there are no conflicts of interest.
Ethical approval and consent to participate
Not applicable.
Consent for publication
Not applicable.
Copyright
© The Author(s) 2025.
Supplementary Materials
REFERENCES
1. Niu, Y.; Teng, X.; Gong, S.; Xu, M.; Sun, S. G.; Chen, Z. Engineering two-phase bifunctional oxygen electrocatalysts with tunable and synergetic components for flexible Zn-air batteries. Nano-Micro. Lett. 2021, 13, 126.
2. Bin, D.; Yang, B.; Li, C.; et al. In situ growth of nife alloy nanoparticles embedded into N-doped bamboo-like carbon nanotubes as a bifunctional electrocatalyst for Zn-air batteries. ACS. Appl. Mater. Interfaces. 2018, 10, 26178-87.
3. Niu, H. J.; Zhang, L.; Feng, J. J.; Zhang, Q. L.; Huang, H.; Wang, A. J. Graphene-encapsulated cobalt nanoparticles embedded in porous nitrogen-doped graphitic carbon nanosheets as efficient electrocatalysts for oxygen reduction reaction. J. Colloid. Interface. Sci. 2019, 552, 744-51.
4. Gebremariam, T. T.; Chen, F.; Jin, Y.; Wang, Q.; Wang, J.; Wang, J. Bimetallic NiCo/CNF encapsulated in a N-doped carbon shell as an electrocatalyst for Zn-air batteries and water splitting. Catal. Sci. Technol. 2019, 9, 2532-42.
5. Li, C.; Wu, M.; Liu, R. High-performance bifunctional oxygen electrocatalysts for zinc-air batteries over mesoporous Fe/Co-N-C nanofibers with embedding FeCo alloy nanoparticles. Appl. Catal. B. Environ. 2019, 244, 150-8.
6. Wu, X.; Yan, Q.; Wang, H.; et al. Heterostructured catalytic materials as advanced electrocatalysts: classification, synthesis, characterization, and application. Adv. Funct. Mater. 2024, 34, 2404535.
7. Liu, Y.; Zhu, Q.; Zhang, L.; Xu, Q.; Li, X.; Hu, G. Nickel-induced charge transfer in semicoherent Co-Ni/Co6Mo6C heterostructures for reversible oxygen electrocatalysis. J. Colloid. Interface. Sci. 2024, 674, 361-9.
8. Yang, L.; He, R.; Botifoll, M.; et al. Enhanced oxygen evolution and zinc-air battery performance via electronic spin modulation in heterostructured catalysts. Adv. Mater. 2024, 36, 2400572.
9. Zhu, J.; Xiao, M.; Li, G.; et al. A Triphasic bifunctional oxygen electrocatalyst with tunable and synergetic interfacial structure for rechargeable Zn‐air batteries. Adv. Energy. Mater. 2020, 10, 1903003.
10. Zhang, F.; Chen, L.; Zhang, Y.; et al. Engineering Co/CoO heterojunctions stitched in mulberry-like open-carbon nanocages via a metal-organic frameworks In-situ sacrificial strategy for performance-enhanced zinc-air batteries. Chem. Eng. J. 2022, 447, 137490.
11. Li, K.; Cheng, R.; Xue, Q.; Zhao, T.; Wang, F.; Fu, C. Construction of a Co/MnO mott-schottky heterostructure to achieve interfacial synergy in the oxygen reduction reaction for aluminum-air batteries. ACS. Appl. Mater. Interfaces. , 2023, 9150-9.
12. Luo, L.; Liu, Y.; Chen, S.; et al. FeNiCo|MnGaOx heterostructure nanoparticles as bifunctional electrocatalyst for Zn-air batteries. Small 2024, 20, 2308756.
13. He, R.; Yang, L.; Zhang, Y.; et al. A CrMnFeCoNi high entropy alloy boosting oxygen evolution/reduction reactions and zinc-air battery performance. Energy. Storage. Mater. 2023, 58, 287-98.
14. Xiao, T.; Sun, C.; Wang, R. Electrodeposited CrMnFeCoNi oxy-carbide film and effect of selective dissolution of Cr on oxygen evolution reaction. J. Mater. Sci. Technol. 2024, 200, 176-84.
15. Hoffman, A. S.; Greaney, M.; Finzel, J.; et al. Elucidation of puzzling questions regarding the CrOx/Al2O3 catalyst I. X-ray absorption spectroscopy aided identification of the nature of the chromium oxide species in the CrOx/Al2O3 dehydrogenation catalyst system. Appl. Catal. A. Gen. 2023, 660, 119187.
16. Peterson, M. L.; Brown, G. E.; Parks, G. A.; Stein, C. L. Differential redox and sorption of Cr (III/VI) on natural silicate and oxide minerals: EXAFS and XANES results. Geochim. Cosmochim. Acta. 1997, 61, 3399-412.
17. Song, X. Z.; Zhao, Y. H.; Zhang, F.; et al. Coupling plant polyphenol coordination assembly with Co(OH)2 to enhance electrocatalytic performance towards oxygen evolution reaction. Nanomaterials 2022, 12, 3972.
18. Bumajdad, A.; Al-Ghareeb, S.; Madkour, M.; Sagheer, F. A. Non-noble, efficient catalyst of unsupported α-Cr2O3 nanoparticles for low temperature CO oxidation. Sci. Rep. 2017, 7, 14788.
19. Wang, Z.; Xi, L.; Yang, Y.; et al. Spin-dependent transport properties of CrO2 micro rod. Nano-Micro. Lett. 2014, 6, 365-71.
20. Menezes, P. W.; Indra, A.; Gutkin, V.; Driess, M. Boosting electrochemical water oxidation through replacement of Oh Co sites in cobalt oxide spinel with manganese. Chem. Commun. 2017, 53, 8018-21.
21. Zhao, X.; Lu, M.; Zhang, G.; et al. Boosting ORR activity via bidirectional regulation of the electronic structure by coupling MnO/Mn3O4 composite materials with N-doped carbon. ACS. Sustain. Chem. Eng. 2024, 12, 8425-35.
22. Biesinger, M. C.; Payne, B. P.; Grosvenor, A. P.; Lau, L. W.; Gerson, A. R.; Smart, R. S. 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, 2717-30.
23. Cheng, M.; Fan, H.; Song, Y.; Cui, Y.; Wang, R. Interconnected hierarchical NiCo2O4 microspheres as high-performance electrode materials for supercapacitors. Dalton. Trans. 2017, 46, 9201-9.
24. Grosvenor, A. P.; Biesinger, M. C.; Smart, R. S. C.; Gerson, A. R. The influence of final-state effects on XPS spectra from first-row transition-metals. Hard. X-ray. Photoelectron. Spectroscopy. , Woicik J, editor; Springer Series in Surface Sciences, Vol.59; Cham, Springer International Publishing; 2016. pp. 217-62.
25. Chen, M.; Liu, D.; Qiao, L.; et al.
26. Wei, J.; Xia, D.; Wei, Y.; Zhu, X.; Li, J.; Gan, L. Probing the oxygen reduction reaction intermediates and dynamic active site structures of molecular and pyrolyzed Fe-N-C electrocatalysts by in situ raman spectroscopy. ACS. Catal. 2022, 12, 7811-20.
27. Balasubramanian, M.; Mcbreen, J.; Davidson, I. J.; Whitfield, P. S.; Kargina, I. In situ X-ray absorption study of a layered manganese-chromium oxide-based cathode material. J. Electrochem. Soc. 2002, 149, A176.
28. Li, L.; Cao, X.; Huo, J.; et al. High valence metals engineering strategies of Fe/Co/Ni-based catalysts for boosted OER electrocatalysis. J. Energy. Chem. 2023, 76, 195-213.
29. Bo, X.; Hocking, R. K.; Zhou, S.; et al. Capturing the active sites of multimetallic (oxy)hydroxides for the oxygen evolution reaction. Energy. Environ. Sci. 2020, 13, 4225-37.
30. Radinger, H. Operando Raman spectroscopy of transition metal oxide catalysts in regard to the oxygen evolution reaction. MS thesis,Technische Universität Darmstadt, 2023.
31. Chang, S.; Cheng, C.; Cheng, P.; Huang, C.; Lu, S. Pulse electrodeposited FeCoNiMnW high entropy alloys as efficient and stable bifunctional electrocatalysts for acidic water splitting. Chem. Eng. J. 2022, 446, 137452.
32. Huang, C.; Lin, Y.; Chiang, C.; et al. Atomic scale synergistic interactions lead to breakthrough catalysts for electrocatalytic water splitting. Appl. Catal. B. Environ. 2023, 320, 122016.
33. Moysiadou, A.; Lee, S.; Hsu, C. S.; Chen, H. M.; Hu, X. Mechanism of oxygen evolution catalyzed by cobalt oxyhydroxide: cobalt superoxide species as a key intermediate and dioxygen release as a rate-determining step. J. Am. Chem. Soc. 2020, 142, 11901-14.
Cite This Article

How to Cite
Download Citation
Export Citation File:
Type of Import
Tips on Downloading Citation
Citation Manager File Format
Type of Import
Direct Import: When the Direct Import option is selected (the default state), a dialogue box will give you the option to Save or Open the downloaded citation data. Choosing Open will either launch your citation manager or give you a choice of applications with which to use the metadata. The Save option saves the file locally for later use.
Indirect Import: When the Indirect Import option is selected, the metadata is displayed and may be copied and pasted as needed.
About This Article
Copyright
Data & Comments
Data

Comments
Comments must be written in English. Spam, offensive content, impersonation, and private information will not be permitted. If any comment is reported and identified as inappropriate content by OAE staff, the comment will be removed without notice. If you have any queries or need any help, please contact us at [email protected].