HEO impurity model. The initial studies of unary, binary and ternary mixtures of Cr, Mn, Fe, Co and Ni are described in the Methods Section,. In the case of more complex metal oxides, rather than constructing the fully periodic bulk model of an arbitrary mixture, it is more computationally feasible to study the HEO effect via local mixing near the active site. As schematically illustrated in Fig. 1a, we select a central metal site and allow neighboring metals (or ligands) to be interchanged. Such an approach can be extended to the next-neighbor shells and beyond while also being site-specific and computationally tractable. This approach can be used to develop both bulk and surface impurity models. The surface model is applied to extract the mixing enthalpy and study the catalytic properties, such as OER, as discussed later.
For the equimolar HEO surface systems, we purposely choose Co3O4(N) as it forms a stable normal spinel structure with high cubic symmetry (see Table S1). Arguably, some other systems or volume-averaged spinel models could also have been used. Here, we choose the (100) facet terminated with the octahedral site (100)-A type surface of Co3O4 from the literature.52 (see Methods Section). In this setting, each active site at the top-most layer is surrounded by five nearest elements (< 3.5 Å) (A-E, Fig. 1a), and these five neighboring sites can be occupied by Cr, Co, Mn, Ni, and Fe in any arrangements, which leads to 120 distinct permutations per active site. Similar to approach by Rossmeisl et al. for high entropy alloys,45 this strategy allows the local, tractable treatment of the HEO catalytic effects at the surface. Including the next-neighbor shells (within 6.6 Å) in our model would result in a 100-fold increase of distinct permutations. However, based on few cases, we determined that this effect is less than 0.3 eV per adsorption energy. Moreover, we limit our study to Co, Cr, and Fe as active sites because both Co and Cr pure spinel oxides show the best activity in pure form for OER, while Fe active sites represent weak binding limits (Fe, Mn, Ni) (discussed later). The likelihood that Fe-site behaves similarly in Fe-doped NiOOHx62 and related HEO systems12 is also relevant factor.
Stability of HEO. The feasibility of successful HEO formation is estimated thermodynamically by calculating the Gibbs free energy of mixing (\({\varDelta G}_{mix}\)) using Eq. (1) as a more negative value of \({\varDelta G}_{mix}\) indicates a more stable and homogenous mixing.
$${\varDelta G}_{mix}={\varDelta H}_{mix}-{T\varDelta S}_{mix} \left(1\right)$$
We evaluated the theoretical configurational entropy of mixing \({\varDelta S}_{mix}\) for an equimolar 5-metal system by using the following formula as
$${\varDelta S}_{mix}=-R{\sum }_{i=1}^{M}{x}_{i} ln \left({x}_{i}\right)$$
2
where R is the ideal gas constant (0.0000862 eV K− 1 metal− 1); M is the number of metal cations in HEOs;\({x}_{i}\)is the molar fraction of each metal cation. For a 5-component (M = 5) equimolar system, \({\varDelta S}_{mix}=1.6R\) and the calculated configuration mixing entropy at the flame annealing temperature of 1000 K is \({-T\varDelta S}_{mix}\)~ -0.14 eV/metal. We used our surface HEO impurity model to evaluate the enthalpy of mixing, \({\varDelta H}_{mix}\) per metal referenced to pure Co3O4 (100) surface simply as
$${\varDelta H}_{mix}=\left[{E}_{HEO\left(100\right)}^{slab}-{E}_{C{o}_{3}{O}_{4}\left(100\right)}^{slab}-{\sum }_{i=1}^{M}{E}_{i}^{bulk}\left({M}_{3}^{i}{O}_{4}\right)+M{\bullet E}^{bulk}\left({Co}_{3}{O}_{4}\right)\right]/M$$
3
,
where \({E}_{HEO\left(100\right)}^{slab}\) and \({E}_{C{o}_{3}{O}_{4}\left(100\right)}^{slab}\) are the DFT total energies for the HEO and Co3O4 (100) slabs, respectively. In Eq. (3), the spinel bulk energies for individual mixing metals (M = Co, Fe, Ni, Cr, Mn) are labeled as \({E}_{i}^{bulk}\left({M}_{3}^{i}{O}_{4}\right)\). Because surface energies of the HEO (100) model and Co3O4 (100) approximately cancel out, the mixing enthalpy of this surface model is also a reasonable estimation of the bulk HEO mixing.
Figure 1b plots the distribution of the resulting mixing enthalpies for the Co, Cr, and Fe active sites. The distribution of \({\varDelta H}_{mix}\) per metal for all three active sites is rather symmetric around the mean value and spans up to 0.5 eV, which is much larger than that of the rated HEA study (~ 0.06 eV)53 by considering an equimolar five elements system. Moreover, over 85% of combinations have a negative \({\varDelta H}_{mix}\). Since the \({-T\varDelta S}_{mix}\) is also negative (-0.14 eV/metal), \({\varDelta G}_{mix}\) will be mostly negative, indicating the formation of HEOs is thermodynamically favorable.
Synthesis and characterization of HEOs. Figure 2a illustrates the sol-flame process48,49 of synthesizing spinel HEO (Co, Fe, Ni, Cr, Mn) as described in the Experimental section. It entails precursor mixing in the liquid phase, followed by spin coating and rapid high-temperature flame annealing and quenching. The high-temperature annealing and subsequent quenching process are crucial to preventing the formation of intermetallic phases.15,54 The X-ray diffraction (XRD) of the resulting HEO in Fig. 2b shows a pure spinel Fd-3m (227) crystal structure with ICDD 00-01-0325. The in-situ XRD results (Fig. S5) show that the pure spinel structure starts to appear at 400 ˚C and is prominent at the flame treatment temperature of 1000 ˚C, indicating that the sol-flame temperature is high enough to synthesize the spinel structure of HEO. Moreover, the transmission electron microscopy (TEM) (Fig. 2c) and scanning electron microscopy (SEM) (Fig. S6) images show that the as-synthesized HEO consists of interconnected nanoparticles (tens of nm in diameter) that are uniformly coated on the FTO/glass substrate. In addition, the high-resolution TEM image (Fig. 2d) and the selected area electron diffraction (SAED) pattern (Fig. 2e) reveal that those nanoparticles are the polycrystalline phase of spinel oxide, mainly indexed to (311), (220), and (222), corresponding to the d-spacings of 0.252, 0.240, and 0.290 nm, respectively. The d-spacing values slightly deviate from the reference NiFe2O4 (00-010-0325), 0.251, 0.241, and 0.294 nm for (311), (220), and (222), which is probably attributed to the atoms with different ionic radius in the lattice. The fast Fourier transform (FFT) image (Fig. 2d, inset) further confirms that the lattice fringes agree well with the XRD and scanning transmission electron microscopy (STEM) results. Figure 2f shows the dark-field STEM and energy dispersive X-ray spectroscopy (EDS) elemental mapping images, verifying the homogeneous mixing of each element (Co, Fe, Ni, Cr, Mn, and O).
The X-ray photoelectron spectroscopy (XPS) (Fig. S7) quantification analysis shows that the HEO surface has a near-equimolar concentration of each metal cation. Detailed elemental XPS analysis in Fig. S8 shows that Cr only exists as Cr3+ and other elements (Mn, Fe, Co, and Ni) have both divalent and trivalent oxidation states. The O 1s spectra were also deconvoluted to three different peaks; M-O at 529.8 eV, M-OH at 531.2 eV, and adsorbed H2O at 532.5 eV, suggesting that the synthesized HEO particle surface contains hydroxide or oxyhydroxide species as well. For binary oxides to HEO, we deconvoluted the detailed Fe 2p3/2 and Co 2p3/2 XPS spectra to see the different ratios of divalent and trivalent oxidation states, as shown in Fig. S9. They exhibit different Fe2+/Fe3+ and Co2+/Co3+ at the surface where HEO shows the lower Fe2+/Fe3+ and higher Co2+/Co3+ than the lower entropy oxides. This implies that HEO has more Fe3+ and Co2+ at the surface, elaborated by electron energy loss spectroscopy (EELS) later. Thus, the different ratios of divalent and trivalent states in Co and Fe 2p3/2 influence the OER activity. Also, the concentration of trivalent and divalent states of all elements for different spinel oxides is summarized in Table S2.
Furthermore, the X-ray absorption spectroscopy (XAS) measurements (Fig. S10) show that the bulk valence states of Co and Fe in HEO contain divalent and trivalent states. In detail, the slope of X-ray absorption near-edge structure (XANES) profiles for both Co and Fe K-edge show that HEO exhibits the higher bulk Co3+ than Co2+ and lower Fe2+ than Fe3+ overall throughout the particles, compared to those of the lower entropy oxides, which is consistent with the XPS results. A detailed description of XANES is also shown in Supplementary information (caption in Fig. S10).
The electron energy loss spectroscopy (EELS) analysis (Fig. 2g to 2k, 80 nm distance, spacing roughly 16 nm) was performed on the HEO particle to quantify its chemical composition. The EEL spectra (Fig. 2h) show the presence of the L2,3-edges of all five metal elements (Cr, Mn, Fe, Co, Ni) at six positions in Fig. 2g. Moreover, Figs. 2i and 2j plot the L3/L2 white-line intensity ratio for Co and Fe at those six positions. As the L3/L2 white-line intensity ratio was reported to be closely associated with the orbital occupancy of the 3d transition metal elements and their oxidation states,55–58 the results suggest that the HEO has more Fe3+ and Co2+ on the surface and more Fe2+ and Co3+ in the center of the HEO particle (80 nm). Figure 2k shows the energy separation between O K-edge pre-peak to the main peak, and its larger values near the surface indicate the higher oxidation states of overall elements on the surface.59
Electrochemical performance of high-entropy spinel oxides. Comparisons of the polarization curves of different spinel oxides containing one to five metal cations in a 1 M KOH electrolyte are shown in Fig. 3a. Among them, the HEO (CoFeNiCrMn) has the lowest overpotential of 309 mV at 10 mA cm− 2, in comparison to the other spinel oxides: CoFeNiCr (327 mV) < CoFeNiMn (379 mV) ~ CoFeNi (351 mV) < CoFe (418 mV) ~ Co (433 mV), and < Fe (522 mV) (Fig. S11). Starting from Co, adding Fe is negligible for the activity improvement, while adding Ni lowers the overpotential. Combining Mn to CoFeNi degrades the performance, while adding Cr is hugely beneficial. Interestingly, HEO (CoFeNiCrMn) shows the highest performance even though it contains Mn, which probably stems from the tuned adsorption energies discussed later. The HEO also has the smallest Tafel slope (30 mV dec− 1) than other spinel oxides (Fig. 3b). In addition, the electrochemical impedance spectroscopy (EIS) measurement shows that HEO has both the lowest charge transfer resistance inside the bulk material and across the interface between the catalyst and electrolyte (Fig. S12 and Table S3). It suggests that HEO has higher electrical conductivity than the lower entropy oxides. For HEOs, we also fixed the four metal cations of CoFeNiMn and varied the fifth one from Cr to V, Cu, Zn, and W (Fig. S13). Among them, the inclusion of Cr exhibits the lowest onset potential, which is consistent with the report that Cr (oxide or oxyhydroxide) is active for OER.60 Fig. 3c summarizes the activity trend for all the overpotential values of the different spinel oxides, showing that the OER activity is improved with increasing the number of the metal cations. Also, the HEO (CoFeNiCrMn) displays nearly 95% of Faradaic efficiency (FE) at 1.54 V vs. RHE, demonstrating great selectivity (Fig. S14).
The stability of HEO (CoFeNiCrMn) was further evaluated by chronopotentiometry and cyclic voltammetry tests. The chronopotentiometry measurement under 10 mA cm− 2 shows that the overpotential for HEO increases by ~ 12% for 168 hours (7 days) (Fig. 3d). It should be noted that part of the charge is related to surface coverage by generated oxygen bubbles that were not entirely removed during the measurement. In comparison, spinel oxides with three and four metal cations show a pronounced 27% and 22% increase in overpotentials within 2 hours (Fig. 3d, inset). The trend demonstrates that the stability is improved with the larger degree on mixing. Furthermore, the cyclic voltammetry test of HEO shows that the oxidation peak at approximately 1.4 V vs. RHE gradually increases during the 1000 cycles (scan rate of 100 mV s− 1) (Fig. S15), suggesting that the surface oxidation states were altered during the OER. The OER activity and stability values for different HEMs, binary spinel oxides, and binary layered double hydroxides (LDH) from other literature are also summarized for comparison in Table S4. Here, we conducted the durability test for the longest time among the catalysts in the table, and HEO shows comparable stability and activity compared to the reported materials, even though we simply synthesized it as a thin film, not including complex synthesis and engineering. In addition, time-dependent ex-situ XPS analysis was performed to analyze the surface state change of HEO. Figure 3e shows that the ratio of trivalent state in Fe and Ni and divalent state in Mn increase within a couple of hours of the test. The O 1s XPS spectra in Fig. 3f show that the surface OH out of (OH + O) became dominant with time. Detailed XPS analysis in Fig. S16 indicates that the Fe3+ and Ni3+ and OH concentrations were increased at the surface for 5 hours. This result demonstrates that the surface Fe-NiOOH was formed during the EC measurement, which influenced the slight change of overpotential (Fig. 3e). Although it is well known that the surface oxyhydroxide is also a promising OER catalyst,60 HEO still plays an important role as support contributing to the great stability. Thus, we also evaluated the surface oxidation OER activity on oxyhydroxide by the DFT calculation that will be elaborated on in the later section.
Theoretical investigation of OER activity. After experimental synthesis and OER activity/stability measurements, we found that the spinel HEO shows lower overpotential than lower entropy oxides. To interpret the experimental activity, we calculated the O* and OH* binding energies for three different active sites using the HEO impurity model (see Fig. 1), along with the pure spinel systems. The results are summarized in a 2D volcano (O*-OH*, OH* descriptors) overpotential heat map (Fig. 4a), developed previously10,47,60,61 and based on scaling in Fig. S1. By considering only the octahedral (+ 3) active sites at the (100) facets, our calculations show that the Co3O4 spinel has the smallest overpotential of 0.63 V among all pure spinel oxides. The trends of overpotential for various pure spinel oxides are: Co3O4 (0.63 V) < Cr3O4 (0.64 V) < Mn3O4 (0.91 V) < Ni3O4 (1.25 V) < Fe3O4 (1.76 V). The binding energy for all adsorbates and their corresponding overpotentials are tabulated in Table S5 and plotted in Fig. 4a. Next, we introduce the O* and OH* binding energies from the HEO impurity model (see Fig. 1) for Cr, Co, and Fe active sites. The predicted more than one thousand binding energies and their corresponding overpotentials were grouped for each active site (Fig. 4a). The distributions of overpotentials (and their corresponding descriptor values) for Co (light blue) and Cr (orange) sites in HEO are densely packed, while Fe (light green) sites is widespread. The arrows represent the changes in the activity trends compared to their respective pure spinel overpotentials. In all three tested active metal types, DGO* - DGOH* OER descriptor values are moved from right to left side towards region of stronger adsorption, and Cr shows the smallest while Fe shows the largest shift.
The overall OER activity for HEO depends strongly on the active sites. The Co active site in HEO shows a minimum overpotential of 0.29 V, while both Cr and Fe result in a minimum overpotential of 0.34 V for the OER activity. Figure 4b shows the OER activity distribution as a function of overpotential, extracted from the 2D volcano plot (Fig. 4a). It shows that less than 5% of the active sites are activated below 0.4 V applied overpotential and the majority contribution to activity comes from the Co active sites. Nearly 25% of the sites are active when the overpotential is less than 0.5 V, and Co sites continues to dominate in terms of activity. Next, several Cr active sites were activated within the region of intermediate overpotential of 0.5–0.8 V. More Cr active sites are activated as the overpotential increases. Unlike Co and Cr, only a few Fe active sites are found bellow 0.7 V overpotential. We also compared the OER activity on the Fe-NiOOH oxyhydroxide system calculated previosly62 to the spinel systems. The theoretical activity Fe-NiOOH (0.37 V) is slightly worse than the most active Fe site in HEO (0.34 V) and overall active site Co site in HEO (0.29 V).
To analyze the activity trend results obtained in the experiment, we compared the OER activity of HEO with that of other spinel oxides, such as ternary and quaternary with various elemental combinations, along with Fe-NiOOH (see Fig. 4c). We predicted that the combination of 5 elements in HEO has the lowest calculated overpotential of 0.29 V. The calculated overpotential trends for these oxide systems are as follows: HEO (CoFeNiCrMn) (0.29 V) < Fe-NiOOH (0.37 V) < CoFeNiCr (0.49 V) < CoFeNi (0.52 V) < CoFeNiMn (0.69 V), which shows good overall consistency with experimental finding, justifying the applicability of the HEO impurity model. Finally, we also analyzed the evolution of board adsorption energy distributions for O* and OH* descriptors, as discussed in detail next.
The role of local strain on HEO activities. The HEO DGO* - DGOH* OER descriptor values are generally shifted towards the region of stronger adsorption (Fig. 4a and S17). However, the individual distributions of O* and OH* binding energies are more complex. The Cr active sites move to the left (stronger ads.), while Co and Fe move to the right (weaker ads.) (Fig. S18). We attempted to investigate the possible reason for these changes considering two limiting cases having either strong or weak adsorption energies relative to the pure system using the Co active site (Figs. S19 and S20). The major finding here is that the local expansion (contraction) to equatorial Co-O bonds causes the stronger (weaker) O* and OH* binding. The presence of Fe as a surface neighbor strongly attracts oxygen, causing an expansion of these lateral Co-O bonds. In contrast, when the surface neighbor is Mn, and the sub-surface neighbor is Cr, these Co-O bonds contract, resulting in a weaker O* and OH* adsorption. The charge density difference analysis clearly shows that such expansion of Co-O bonds creates electron depletion along Co sites causing strong O* and OH* adsorption compared to pure and contraction cases. Finally, the equatorial expansion-contraction is also compensated by the axial sub-surface oxygen M-O bond to shorten-elongate, respectively (Figs. S21 and S22).