We present an exploration of a family of compositionally complex cubic spinel ferrites featuring combinations of Mg, Fe, Co, Ni, Cu, Mn, and Zn cations, systematically investigating the average and local atomic structures, chemical short-range order, magnetic spin configurations, and magnetic properties. All compositions result in ferrimagnetic average structures with extremely similar local bonding environments; however, the samples display varying degrees of cation inversion and, therefore, differing apparent bulk magnetization. Additionally, first-order reversal curve analysis of the magnetic reversal behavior indicates varying degrees of magnetic ordering and interactions, including potentially local frustration. Finally, reverse Monte Carlo modeling of the spin orientation demonstrates a relationship between the degree of cation inversion and the spin collinearity. Collectively, these observations correlate with differences in synthesis procedures. This work provides a framework for understanding magnetic behavior reported for “high-entropy spinels,” revealing many are likely compositionally complex oxides with differing degrees of chemical short-range order—not meeting the community established criteria for high or medium entropy compounds. Moreover, this work highlights the importance of reporting complete sample processing histories and investigating local to long-range atomic arrangements when evaluating potential entropic mixing effects and assumed property correlations in high entropy materials.

High entropy oxides (HEOs) are being increasingly reported in various oxide crystal structures, including in the perovskite, fluorite, spinel, and pyrochlore structure motifs.1–18 HEOs are generally described as single phase solid solution systems exhibiting five or more homogeneously distributed cations in equimolar or near-equimolar ratios on a single crystallographic site. When a positive entropy of formation (ΔSf) overcomes a positive enthalpy of formation (ΔHf), the Gibbs free energy of formation (ΔGf = ΔHfTΔSf) will be negative, making the system an entropy stabilized oxide (ESO). The ESO, (Mg0.2Ni0.2Co0.2Cu0.2Zn0.2)O, a five component oxide exhibiting a single rock salt phase, was first introduced by Rost et al. in 2015.19 HEOs/ESOs draw interest due to inherent opportunities to tailor and combine materials properties. The extent to which their enhanced compositional and configurational complexity impart additional stability, synergistic effects, and emergent/tunable properties to HEO archetypes is an active area of exploration. The magnetic properties of HEOs represent one of the fastest growing research directions in this scientific community. From the perspective of magnetic behavior, the complexity of the local and long-range nuclear and spin structures found in HEOs gives rise to an unusually large number of metal–oxygen–metal interactions.20 The inherent variation in coordination geometry, valence state, spin state, number of cations, and metal cation type each HEO lattice can accommodate presents a rich pallet for exploring the tunability of the structure–property effects in the class. However, relatively few compositions have been studied in detail,20 no doubt due in part to the complex cation and spin order likely to be present.

Spinels crystallizing in cubic space group Fd3̄m, with nominal formula AB2O4, feature a close-packed oxygen substructure with two thirds of cations occupying octahedral interstices and one third occupying tetrahedral ones. The spinel compounds in this work involve A and B cations in the cation charge state combination of +2 and +3 (charge-balanced spinel oxide families with +4 and +2, and +6 and +1 cation combinations are also known to exist). Spinels are typically classified according to which cations occupy the respective coordination environments, with a degree of inversion, γ, denoting the fraction of A ions occupying the octahedral sites, [A1−γBγ]tet[AγB2−γ]octO4. In the “normal” spinel structure (γ = 0, [A]tet[B2]octO4), B3+ fully occupies the spinel octahedral site, while A2+ fully occupies the tetrahedral site. In the “inverse” spinel structure (γ = 1, [B]tet[AB]octO4), half of the B3+ cations fully occupy the tetrahedral site, leaving an equal amount of A2+ and B3+ atoms occupying the octahedral site. In a “random” spinel structure (γ = 2/3, [A1/3B2/3]tet[A2/3B4/3]octO4), A2+ and B3+ atoms take up residence on both sites according to their overall molar ratio, with 1/3 and 2/3 distribution, respectively. Arbitrary intermediate states between the normal and inverse spinels are possible (γ value between 0 and 1). Figure 1 provides a depiction of high entropy spinel AB2O4 (with A = 5 equimolar cations and B = a single metal cation), shown for the case of the inverse spinel ferrite structure.

FIG. 1.

Schematic view of (a) the inverse spinel unit cell with blue tetrahedra (A-sites) and orange octahedra (B-sites). Cations on the octahedral site are drawn half blue and half orange to represent site mixing. Oxygen atoms sitting on polyhedral vertices are omitted for clarity. (b) The [0 0 1] projection of the cation substructure in the unit cell (black square) overlaid with the unit cell of the magnetic structure (blue frame). (c) The room-temperature zero-field magnetic spin structure composed of magnetic moments on the tetrahedral site (blue vectors pointing down) and octahedral site (orange vectors pointing up).

FIG. 1.

Schematic view of (a) the inverse spinel unit cell with blue tetrahedra (A-sites) and orange octahedra (B-sites). Cations on the octahedral site are drawn half blue and half orange to represent site mixing. Oxygen atoms sitting on polyhedral vertices are omitted for clarity. (b) The [0 0 1] projection of the cation substructure in the unit cell (black square) overlaid with the unit cell of the magnetic structure (blue frame). (c) The room-temperature zero-field magnetic spin structure composed of magnetic moments on the tetrahedral site (blue vectors pointing down) and octahedral site (orange vectors pointing up).

Close modal

Specific (not random) cation site preferences of the crystal chemistry in spinel oxides are well documented in the literature,21,22 going back to early reports23,24 linking cation site distribution and magnetic properties in spinel ferrites. There has been ample research into the magnetic properties of spinel AFe2O4 since. AFe2O4 (A = Fe2+,25,26 Ni2+,27 Cu2+,28 Mg2+,29 or Co2+26,30,31) spinels have a variety of important applications in magnetic technologies.24,32,33 They all have an inverse (either partially or completely) spinel structure. Previous works have shown that, in spinel ferrites, Zn2+ prefers to occupy the tetrahedral site while Mg2+, Mn2+, Mn3+, Fe2+, Fe3+, Ni2+, Cu2+, and Co2+ tend to occupy the octahedral site in spinel ferrites.24 It remains to be seen whether high entropy spinels will follow the trends in site preference observed among the ternary spinel ferrite systems or if the potential for increased configurational entropy will lead to more random cation distributions.

High entropy spinels have been synthesized by solid state,34–49 co-precipitation,50,51 solution combustion synthesis (SCS),52–59 sol–gel method,60,61 solvothermal synthesis,62 reverse co-precipitation,63 flame spray pyrolysis,64 hydrothermal method,65 polymerized complex method,66 and glycine-combustion method67 among other techniques. Examples of reported high-entropy ferrite powders are summarized in Table I.

TABLE I.

A summary of powder sample high entropy spinel ferrite compositions appearing in the literature, including year reported (Year), chemical composition (Chem.), magnetic ground state (GS) (where FIM denotes ferrimagnetic), and the reference (Ref.).

YearChem.GSReferences
Solid state method + sintering 
2019 (Mg0.2M0.2Co0.2Ni0.2Cu0.2)Fe2O4, M = Zn, Mn, Fe FIM 35  
2019 (Mn0.2Fe0.2Co0.2Ni0.2Cu0.2)Fe2O4 FIM 35  
2020 (Co0.2Mn0.2Ni0.2Fe0.2Zn0.2)Fe2O4 (no sintering) N/A 46  
2022 (Mg0.2Fe0.2Co0.2Ni0.2M0.2)Fe2O4, M = Mn, Cu, Zn FIM 49  
Co-precipitation + annealing 
2019 (Zn0.2Mg0.2Ni0.2Fe0.2Cd0.2)Fe2O4 N/A 50  
Sol–gel method 
2020 (Co0.2Cr0.2Fe0.2Mn0.2Ni0.2)Fe2O4 FIM 60  
Solution combustion synthesis (SCS) 
2021 (MnNiCuZn)1−xCoxFe2O4 (x = 0.05, 0.1, 0.2, 0.3) FIM 58  
Glycine-combustion method 
2022 (M0.2Zn0.2Co0.2Ni0.2Cu0.2)Fe2O4, M = Mn, Fe N/A 67  
2022 (Mn0.2Fe0.2Co0.2Ni0.2Cu0.2)Fe2O4 N/A 67  
YearChem.GSReferences
Solid state method + sintering 
2019 (Mg0.2M0.2Co0.2Ni0.2Cu0.2)Fe2O4, M = Zn, Mn, Fe FIM 35  
2019 (Mn0.2Fe0.2Co0.2Ni0.2Cu0.2)Fe2O4 FIM 35  
2020 (Co0.2Mn0.2Ni0.2Fe0.2Zn0.2)Fe2O4 (no sintering) N/A 46  
2022 (Mg0.2Fe0.2Co0.2Ni0.2M0.2)Fe2O4, M = Mn, Cu, Zn FIM 49  
Co-precipitation + annealing 
2019 (Zn0.2Mg0.2Ni0.2Fe0.2Cd0.2)Fe2O4 N/A 50  
Sol–gel method 
2020 (Co0.2Cr0.2Fe0.2Mn0.2Ni0.2)Fe2O4 FIM 60  
Solution combustion synthesis (SCS) 
2021 (MnNiCuZn)1−xCoxFe2O4 (x = 0.05, 0.1, 0.2, 0.3) FIM 58  
Glycine-combustion method 
2022 (M0.2Zn0.2Co0.2Ni0.2Cu0.2)Fe2O4, M = Mn, Fe N/A 67  
2022 (Mn0.2Fe0.2Co0.2Ni0.2Cu0.2)Fe2O4 N/A 67  

Research into the magnetic properties and cation site ordering among high entropy spinels is still in its infancy. There is one recent report68 demonstrating that the site ordering (Co0.6Fe0.4)(Cr0.3Fe0.1Mn0.3Ni0.3)2O4 in A3O4 spinel is governed by maximization of the crystal field stabilization energy (CFSE) of the constituent ions through site selectivity rather than by maximization of the configurational entropy (Sconfig) through random and homogeneous cation disorder.68 It is not known whether this generalizes to other high entropy spinel compositions, including compositional variations of the technologically important spinel ferrite (AFe2O4). The magnetic properties of spinel ferrites are highly sensitive to the distribution of cations (the degree of inversion, γ) across the tetrahedral and octahedral sublattices.24,69 For example, this phenomena has been well explored by mixing of Zn-ferrite (a normal spinel) with Co-ferrite (an inverse spinel), resulting in non-magnetic Zn2+ swapping for magnetic Fe3+ on octahedral sites and increasing the overall magnetization.70,71 Similar effects may arise in compositionally complex spinel samples, potentially with greater sublattice disorder and/or expanded substitution limits for specific cations.

In 2019, Musicó et al.35 reported on the magnetic properties and trends among a series of high entropy spinel ferrites (compositions AFe2O4) and chromates (compositions ACr2O4). All reported high entropy spinel ferrites were shown to be room temperature ferrimagnets and, like several other HEO magnetic phases, were reported to generally follow trends found in the majority of the simpler ternary systems involving groups of the constituent cations. For example, the high entropy system, including Zn and Mg (the two non-magnetic cations), had the largest magnetic moment and lowest Néel temperature (TN) relative to other high entropy spinels. On the other hand, several other HEO ferrites [(Mg0.2Fe0.2Co0.2Ni0.2Cu0.2)Fe2O4 and (Mg0.2Mn0.2Co0.2Ni0.2Cu0.2)Fe2O4] were noted to show peculiar magnetic behavior. For example, the low temperature field warming magnetization dips below the field cooling magnetization and for intermediate temperatures indicates an increase in magnetization as temperature is raised. A possible explanation for this peculiar thermal hysteresis is the antiferromagnetic coupling between neighboring elemental components in the cation sublattice. This leads to a key question of whether the reported high entropy spinels contain cation short-range order and small locally ordered magnetic clusters or chemically homogeneous distributions with more dilute clusters, and whether the substructures can be tuned.

The extent to which high entropy spinels will exhibit cation distributions different from bulk ternary spinels that influence their properties remains to be determined. In this work, we explore the influence of local cation ordering upon magnetic properties in a family of AFe2O4 high entropy spinels using magnetometry, neutron powder diffraction (NPD), neutron atomic pair distribution function (PDF) studies, and reverse Monte Carlo (RMC) modeling of spin correlations. We draw several connections between observed magnetic properties, the average and local atomic and magnetic structures, and the influence of material processing history in the series.

In this work, the four phase-pure spinel ferrites originally reported by Musicó et al.35 were synthesized by solid state reaction using identical starting materials and processing conditions. We have elected to use the same naming convention as in the original work, with samples labeled F1, F3, F4, and F5, as noted in Table II. Briefly, all the samples were prepared in 5-g batches using mixtures of stoichiometric proportions of dried precursor oxide powders (MgO, NiO, CoO, CuO, ZnO, MnO2, and Fe2O3). The mixtures were ball milled, pressed into pellets, and heat treated according to the details in Table II. Mixtures for samples F1 and F3 were initially heated at 1250 °C for 10 h. Samples F4 and F5 (both containing Mn) were obtained by first heating the mixed powder at 950 °C in order to convert the MnO2 precursor to Mn3O4. After the first heating step, all samples were ground, repressed, and sintered a second time at 1250 °C for 10 h. The heat treatment protocols are the same used for the compositions reported in Musicó et al.,35 though the heat treatment temperatures and times were not included in the original work.

TABLE II.

High entropy spinels in this study, with firing temperatures and number of hours noted. Sample names correspond with those reported by Musicó et al.35 

Heat treatment
NameCompositionNo. 1No. 2
F1 (Mg0.2Fe0.2Co0.2Ni0.2Cu0.2)Fe2O4 1250 °C, 10 h 1250 °C, 10 h 
F3 (Mg0.2Co0.2Ni0.2Cu0.2Zn0.2)Fe2O4 1250 °C, 10 h 1250 °C, 10 h 
F4 (Mg0.2Mn0.2Co0.2Ni0.2Cu0.2)Fe2O4 950 °C, 10 h 1250 °C, 10 h 
F5 (Mn0.2Fe0.2Co0.2Ni0.2Cu0.2)Fe2O4 950 °C, 10 h 1250 °C, 10 h 
Heat treatment
NameCompositionNo. 1No. 2
F1 (Mg0.2Fe0.2Co0.2Ni0.2Cu0.2)Fe2O4 1250 °C, 10 h 1250 °C, 10 h 
F3 (Mg0.2Co0.2Ni0.2Cu0.2Zn0.2)Fe2O4 1250 °C, 10 h 1250 °C, 10 h 
F4 (Mg0.2Mn0.2Co0.2Ni0.2Cu0.2)Fe2O4 950 °C, 10 h 1250 °C, 10 h 
F5 (Mn0.2Fe0.2Co0.2Ni0.2Cu0.2)Fe2O4 950 °C, 10 h 1250 °C, 10 h 

Magnetometry measurements were performed using a vibrating sample magnetometer (VSM) as part of a Quantum Design PPMS DynaCool system. Temperature-dependent magnetometry (M vs T) was performed using field cooled (FC) and zero field cooling (ZFC) methods. For each sample, a ZFC measurement was first performed by cooling the powder sample to 2 K in the absence of a magnetic field, then applying a 10 mT field and measuring the magnetic moment as the temperature was increased from 2 to 400 K. After the ZFC measurement, the applied field was kept at 10 mT while measuring the magnetization as the temperature was reduced from 400 to 2 K for the FC curve. Magnetic hysteresis (MH) loops between ±2 T were also measured at temperatures ranging from 2 K and in 25 K steps up to 400 K. All measurements have been normalized to the sample mass.

First order reversal curve (FORC) measurements were performed to evaluate the distribution of magnetic phases and interactions within the samples.72–74 Following previously published procedures, FORC measurements were performed by first positively saturating the sample, then reducing the magnetic field to a reversal field (HR). Next, the magnetization is measured as the applied field (H) is increased from HR back to positive saturation. This sequence was repeated at HR between the positive and negative saturation. The FORC distribution (ρ) is calculated by applying a mixed second order derivative to the dataset,

(1)

where MS is the saturation magnetization of the sample. FORC measurements were completed at 2 and 300 K for each sample. While H and HR separately probe up and down-switching events, respectively, FORC distributions are commonly presented with a change of coordinates from (H, HR) to

(2)

where HC and HB represent the local coercive and bias (interaction) fields, respectively. As all samples show features prominent along axes in both coordinate systems, FORC distributions are presented in (H, HR) while also indicating the (HC, HB) axes. The FORC distributions also include significant reversible components, which are captured using a constant extension mechanism.75 

Time of flight (TOF) neutron powder diffraction (NPD) data were collected at 300 K on the Nanoscale Ordered Materials Diffractometer (NOMAD) at the Spallation Neutron Source (SNS) at Oak Ridge National Laboratory (ORNL).76 For each sample, ∼0.3 g of powder was loaded into a 3 mm diameter quartz capillary, placed in a temperature controlled Ar cryostream, and data were collected for 1 h. Data reduction was completed with the Advanced Diffraction Environment (ADDIE) suite.77 The instrument background and sample container contributions were subtracted, data were corrected for detector deadtime, absorption, and multiple scattering effects, and were normalized by the incident flux and total sample scattering cross section. The maximum value of Q vector used for generation of the PDF was 40 Å−1. The coherent neutron scattering lengths vary considerably among the cations in the sample series, with (in order from smallest to largest) bMn = −3.73 fm, bCo = 2.49 fm, bCr = 3.635 fm, bMg = 5.375 fm, bZn = 5.68 fm, bCu = 7.718 fm, bFe = 9.45 fm, and bNi = 10.3 fm. The neutron scattering length of oxygen is bO = 5.803 fm.

The experimental diffraction data (31° 2θ detector bank) were analyzed with the Rietveld method as embodied in the TOPAS-V778 program. Analysis was completed for each dataset fixing the cation site occupancy according to a normal ([A]tet[Fe2]octO4), an inverse ([Fe]tet[A,Fe]octO4) [shown in Fig. 1(a)], and a random ([A1/3Fe2/3]tet[A2/3Fe4/3]octO4) spinel structure model (Fd3̄m). In all fits, the magnetic scattering was modelled using a magnetic structure model with an I41/amd′ structure featuring one constrained magnetic moment for all tetrahedral sites and one constrained magnetic moment oriented in opposite direction for all octahedral sites [shown in Fig. 1(c)]. Background parameters, scale factors, cubic lattice parameters anuc, isotropic atomic displacement parameters (ADPs) (one for oxygen atoms, one for all cations on the tetrahedral site, and one for all atoms on the octahedral site), instrument profile parameters, and the oxygen atom position (u) were refined for the nuclear structure models. For the refinement of magnetic structure, the isotropic ADPs were constrained to be equivalent to those in the nuclear structure, while the lattice parameters amag = bmag were constrained to equal 2anuc/2 and the lattice parameter cmag was constrained to equal anuc [according to geometry, Fig. 1(b)]. The scale factor of the magnetic phase was constrained to one-half that of the nuclear phase (since Vmag/Vnuc = 1/2). Magnetic moments were refined along the z direction for the magnetic elements sitting on the respective tetrahedral and octahedral sites (constrained as one unique value per lattice site).

PDF refinements and model simulations were completed using normal, spinel inverse spinel and the random solution models in the PDFgui79 program, and included a scale factor, lattice parameters, quadratic peak sharpening parameters for correlated atomic motion, and isotropic ADPs. Instrument parameters for real-space dampening and broadening were fixed to values determined through refinement of a silicon crystalline standard reference material.

RMC modeling involving both total scattering and Bragg data was conducted using the RMCProfile package.80 Fits for both neutron F(Q) and neutron Bragg data were completed. F(Q) is defined by

(3)

where Q = kk′ is the scattering vector of length 4π sin θ/λ for a neutron of wavelength λ scattered at an angle 2θ, and k and k′ are the initial and final wavevectors of the scattered neutron, respectively. S(Q) is the normalized total-scattering structure factor. b̄i is the coherent bound neutron scattering length of species i, averaged over the different isotopes and nuclear spin states of i. ci is the portion of species i in the material.81 A 10 × 10 × 10 supercell was first built, based on the unit cell model obtained via Rietveld refinement for each sample. The supercell structural configuration was then optimized in a data-driven manner following the Metropolis approach. Two sets of RMC modeling was completed for each sample, assuming either inverse or random site occupancy (the normal model was excluded because of the poor agreement demonstrated by Rietveld refinements). No atomic position swapping was allowed between tetrahedral and octahedral sites. In the original report by Musicó et al., multiple valence states for Fe and Mn were detected among high entropy spinels via x-ray absorption spectroscopy. Thus, RMC trials were performed for several samples assuming various valence states for Fe and Mn atoms (since valence determines the magnetic form factors assigned). Here, it should be pointed out that the magnetic moment vectors in RMC configurations are unit vectors with magnetic moment magnitude specified (fixed) according to each cation involved. The bond valence sum (BVS) constraint in RMC fitting is implemented in such a way that it effectively constrains the net valence of the whole system to be 0. The findings discussed in this report were found to be independent of the valence settings in RMCProfile. The final configurations were initiated assuming a +2 valence state for all A cations in the AFe2O4 compositional series and a +3 valence state for the remaining two Fe cations.

Neutron total scattering structure factor data, F(Q), and Bragg data were used for RMC modeling, since the calculation for magnetic diffuse scattering is performed only in Q-space in RMCProfile. For the inclusion of Bragg data in RMC modeling, peak profiles and background were extracted from Topas refinement and tabulated for RMCProfile to read in—hence the RMC engine only calculates peak intensities. In addition to the experimental data, a minimum distance for each single pair of atoms and an overall bond valence sum were added as model constraints. Given multiple constraints exerted upon the structural model, an automatic weight assigning scheme82 was used.

A kernel density estimation (KDE) and the Henze–Zirkler (HZ) statistic were utilized to interpret variation in the distribution of atom sites and magnetic spin correlations resulting from RMC model fits to neutron total scattering data. KDE is a commonly used method for estimating the underlying probability distribution function in a collection of data points with a non-parameterized approach. Here, the standard normal function is used as the kernel and, in practice, that means one would have a standard normal distribution function with the center at each of the data points. The overall distribution function is then composed of all those kernels, while being normalized to give the overall distribution probability of 1. The HZ parameter83–86 is a parameter characterizing the normality of a distribution and the closer its value gets to 1, the closer that underlying distribution is to a Gaussian distribution. The following equation defines the HZ parameter:

(4)

where Di is the squared Mahalanobis distance of the ith observation to the centroid and Dij gives the squared Mahalanobis distance between ith and jth observations

(5)
(6)

β is given by

(7)

which includes the particle sample size np and sample dimensions d.

The cation distribution in high entropy spinels will be impacted by a number of thermodynamic and crystal-chemical factors. The CFSE of the oxygen ligand field and the Sconfig resulting from the number of participating cations are two important factors. The electron configuration energy difference between ligand field and isotropic field is defined as CFSE,

(8)

The CFSE can be straightforwardly calculated for each participating cation assuming a specific charge state and either an octahedral or tetrahedral geometry, and the net CFSE can be compared for specific distributions of cations to evaluate potential site preferences.

The CFSE for the four compositions in the AFe2O4 series are shown in Fig. 2(a), assuming a +2 valence state for all five cations on the A site, and a +3 valence state for the remaining two Fe cations. Regardless of the oxidation state used in calculations, the CFSE of the inverse structure model is the lowest among the model spinel structures while the CFSE of the normal structure model is the highest.

FIG. 2.

(a) The CFSE calculations for F1 (□), F3 (★), F4 (×), and F5 (♢) with normal spinel, random solution, and inverse spinel structure models assumed (with a +2 valence state for all A cations in the AFe2O4 compositional series, and a +3 valence state for the remaining two Fe cations). (b) The calculated Sconfig of AFe2O4 high entropy spinels with the degree of inversion (γ) changing from 0 to 1 and the assumption that there are N (from 1 to 8, Fe is not included) cations occupying the A site (the γ values 0, 2/3, and 1 were labeled according to the corresponding normal, random solution, and inverse spinel structure configurations).

FIG. 2.

(a) The CFSE calculations for F1 (□), F3 (★), F4 (×), and F5 (♢) with normal spinel, random solution, and inverse spinel structure models assumed (with a +2 valence state for all A cations in the AFe2O4 compositional series, and a +3 valence state for the remaining two Fe cations). (b) The calculated Sconfig of AFe2O4 high entropy spinels with the degree of inversion (γ) changing from 0 to 1 and the assumption that there are N (from 1 to 8, Fe is not included) cations occupying the A site (the γ values 0, 2/3, and 1 were labeled according to the corresponding normal, random solution, and inverse spinel structure configurations).

Close modal

The entropy contribution made by the compositional mixture on the tetrahedral and octahedral lattice sites will differ for normal, inverse, and random spinels. The ideal Sconfig per cation for HEOs with multiple cations occupying different sublattice sites can be calculated68 by

(9)

where R is the universal gas constant, ax is the number of sites on the x sublattice, fix is the fraction of elemental species occupied on the respective sublattice, and N is the number of elements in a given sublattice. For example, for the normal spinel AFe2O4, Eq. (9) can be evaluated as

(10)

As shown in Fig. 2(b) the random spinel offers the highest entropy contribution, regardless of the number N of cations on the A sublattice site of AFe2O4 (from 1 to 8, with Fe not included), followed by the inverse and normal spinel configurations. The Sconfig calculations for N from 1 to 8 with Fe from the B site included are also completed in Fig. S1 in the supplementary material. Since F3 and F4 do not have Fe on the A site their Sconfigs are larger than those of F1 and F5 when random and inverse spinel models are considered, respectively. It is worth noting that the Sconfigs for the compositions in the series are all smaller than 1.5R, meaning the spinels do not meet the classic definition of a high entropy oxide (or even medium entropy oxide). In thermodynamic terms, these compositions are more accurately classified as compositionally complex oxides (CCOs).

To summarize, the CFSE and Sconfig terms for AFe2O4 spinels provide competing drivers for cation site selectivity, favoring inverse vs random cation configurations, respectively. Observed cation distributions could very well be influenced by the sample-specific processing procedures.

Results of temperature- and field-dependent magnetometry for the samples in our study are shown in Figs. 3 and 4(a), respectively. For all of the samples, the black dashed ZFC curve shows a continuous increase with increasing temperature, indicating the enhanced role of thermal fluctuations by reorienting the magnetization with the magnetic field. Also, the ZFC curve is well separated from the FC curve, with no apparent inflection, and the FC curve is smooth and persistently non-zero, indicating the system does not have a blocking temperature and is not superparamagnetic in this temperature range. This is consistent with the previous work by Musicó et al.,35 which identified the Néel temperature (TN) of >400 K; the Néel temperature of other spinel ferrites includes 858 K for Fe3O4 and 725 K for CoFe2O4. These conclusions are also consistent with the MH loops, shown in Fig. 4(a), which show that every sample has a very small, but non-zero, coercivity (≈2–5 mT) at 300 K.

FIG. 3.

The comparison of field-cooled (FC), zero-field-cooled (ZFC), and zero-field-warming (FCW) for all ferrite samples measured from 2 to 400 K with an applied field of 10 mT: (a) F1, (b) F3, (c) F4, and (d) F5. The legend in (c) applies to all subplots. The insets are displayed in order to emphasize the cross-over and hysteresis in the FC and FCW curves.

FIG. 3.

The comparison of field-cooled (FC), zero-field-cooled (ZFC), and zero-field-warming (FCW) for all ferrite samples measured from 2 to 400 K with an applied field of 10 mT: (a) F1, (b) F3, (c) F4, and (d) F5. The legend in (c) applies to all subplots. The insets are displayed in order to emphasize the cross-over and hysteresis in the FC and FCW curves.

Close modal
FIG. 4.

The comparison of magnetic hysteresis loop data for ferrite compositions (F1, F3, F4, F5) measured at 300 K with a ±0.5 T range for (a) materials in this paper, measured with a 10 mT cooling; and (b) materials studied by Musicó et al., measured with a 100 mT cooling. The legend in (b) applies to both subplots.

FIG. 4.

The comparison of magnetic hysteresis loop data for ferrite compositions (F1, F3, F4, F5) measured at 300 K with a ±0.5 T range for (a) materials in this paper, measured with a 10 mT cooling; and (b) materials studied by Musicó et al., measured with a 100 mT cooling. The legend in (b) applies to both subplots.

Close modal

The field-dependent magnetometry measurements of F1, F3, F4, and F5 samples in this study are shown in Fig. 4(a), the measurements completed by Musicó et al.35 are reproduced in Fig. 4(b) for comparison. This comparison shows significant differences in the MH loops, with the samples in this work having a nearly constant (compositionally independent) magnetic saturation (MS) and an exceedingly small coercivity in all of the samples. By comparison, the samples from Musicó et al.35 show a variable MS between the different samples, with F3 being particularly large, while the others are similar to the current work. Also the coercivity in Ref. 35 is generally larger than in the current work, especially for F5. The two batches of samples were synthesized in the same laboratory, by the same researcher, using the same batch size, grinding and sintering procedure, and precursors from the same supplier. The only known change in sample preparation involved a replaced heating element for the furnace used. Additionally, the magnetometry measurements of samples prepared for Musicó et al.35 were completed soon after synthesis, whereas and magnetometry measurements of samples prepared in the present work was completed two years after synthesis. The observed difference in magnetic behavior among samples of the same composition suggests a potential high sensitivity to synthesis and processing conditions, and/or ageing, in high entropy spinels, and deserves further investigation. For consistency, it is important to note that the MH loops in this work were field cooled in 10 mT, while Ref. 35 uses 100 mT. This difference would manifest if the system has an antiferromagnetic phase coupled to the ferromagnetic phase or high anisotropy phase with a saturation field larger than the 2 T applied here. In these cases, the hysteresis loop would be vertically and horizontally shifted in Fig. 4 due to the exchange bias effect. The absence of these features suggests that the cooling field is not responsible for the observed differences.

The differences in the MH plots maybe the result of a strong sensitivity to local compositional distributions and microstructure. This sensitivity would manifest in the magnetic reversal behavior of the sample, which can be probed using the FORC technique. The calculated FORC distributions are shown in Fig. 5 at 2 K [(a), (c), (e), and (g)] and 300 K [(b), (d), (f), and (h)] for all samples.

FIG. 5.

The comparison of FORC measurements at 2 and 300 K for sample F1 [(a) and (b)], F3 [(c) and (d)], F4 [(e) and (f)], and F5 [(g) and (h)], respectively. The directions of HB and HC are plotted for clarification, and the scale bar indicates the rate at which the magnetic reversal speed changes for different HR.

FIG. 5.

The comparison of FORC measurements at 2 and 300 K for sample F1 [(a) and (b)], F3 [(c) and (d)], F4 [(e) and (f)], and F5 [(g) and (h)], respectively. The directions of HB and HC are plotted for clarification, and the scale bar indicates the rate at which the magnetic reversal speed changes for different HR.

Close modal

The FORC distributions all show their primary feature aligned along the HB axis with a small feature in the HC direction that is more prominent at low temperatures. A small bump in the HR direction is also seen at 2 K but is absent at room temperature. While these features are present for all of the samples, they vary in intensity and prominence, with the attributes being lowest for sample F3.

The FORC distribution encodes a map of the switching events within a particular hysteretic system. The bias–field axis identifies the shift from H = 0 of a particular set of switching events. For a system without exchange bias, this shift can be attributed to interactions within the sample. The FORC distribution shown here has a wide spread in HB, indicating strong demagnetizing interactions, e.g., interactions that drive the system away from a saturated configuration.74,87 These interactions can be the result of dipolar coupling or exchange-motivated frustration. The presence of long-range ferrimagnetic magnetic ordering without a blocking temperature, and with only a very small coercivity, suggests the interactions are likely due to frustration. At low temperatures, the protrusion along the HC axis suggests the intrinsic anisotropy has increased compared to the room temperature measurements. The protrusion along the H/HR axes indicate some of the sample reverses by a domain nucleation/propagation mechanism.88 The appearance of the domain reversal mechanism indicates that large areas of the sample have become magnetically coupled and can support domain growth through their boundaries. Complementary to this, the higher temperature samples may consist of magnetic islands or regions that are coupled by non-magnetic or antiferromagnetic phases.

The neutron diffraction patterns of all four samples appear visually similar [Fig. 6(a)] and can be indexed to a cubic spinel Fd3̄m structure model with corresponding I41/amd′ ferrimagnetic ordering model, with no detected secondary phases. Distinguishing cation order is challenging with Bragg diffraction data. The scattering length of involved elements, the degree of cation inversion, and the magnetic order on the cation sublattice may all contribute to observed variation in peak intensities. As a demonstration, the calculated neutron diffraction patterns for nuclear and magnetic scattering assuming inverse, random, and normal spinel structure models with assumed ferrimagnetic spin order for each sample are shown in Fig. S2 in the supplementary material. It can be seen in Fig. 6(a) that the nuclear Bragg peak intensities are higher for F1 and F3 samples relative to F4 and F5 samples. The intensities of dominant magnetic peaks [for example, the (331) and (111) reflections] appear similar for all compositions.

FIG. 6.

(a) The comparison of neutron Bragg data (31° detector bank) from F1 (orange line), F3 (purple line), F4 (red line), and F5 (green line) samples. Several hkl peaks featuring ferrimagnetic contributions are labeled. Best-fit Reitveld refinement of (a) F1 data using inverse spinel and ferrimagnetic spin order model and (c) F4 data using random solution spinel and ferrimagnetic spin order model. In (b) and (c) data are shown as black “x”s, fits are shown as solid orange lines, and difference curves fall below the data and fits in black lines. The blue and green solid curves displayed below the refinement results display the magnetic and nuclear contributions to the fits, respectively.

FIG. 6.

(a) The comparison of neutron Bragg data (31° detector bank) from F1 (orange line), F3 (purple line), F4 (red line), and F5 (green line) samples. Several hkl peaks featuring ferrimagnetic contributions are labeled. Best-fit Reitveld refinement of (a) F1 data using inverse spinel and ferrimagnetic spin order model and (c) F4 data using random solution spinel and ferrimagnetic spin order model. In (b) and (c) data are shown as black “x”s, fits are shown as solid orange lines, and difference curves fall below the data and fits in black lines. The blue and green solid curves displayed below the refinement results display the magnetic and nuclear contributions to the fits, respectively.

Close modal

Separate Rietveld refinements assuming normal, random, and inverse spinel models were completed for each sample (supplementary material, Table S1 for F1, Table S2 for F3, Table S3 for F4, Table S4 for F5), each incorporating the ferrimagnetic structure shown in Fig. 1(c). A summary of Rietveld refinement results is presented in Table III. The Rietveld refinement results reveal distinct preference in cation site ordering. For all samples, the normal spinel structure model resulted in poor agreement and nonsensical magnetic moments. The combination of inverse spinel and ferrimagnetic model fits the data better than those of random and normal spinel models for F1 and F3 samples [the F1 result is shown in Fig. 6(b)], while the combination of random solution spinel and ferrimagnetic model fits the data better than normal and inverse models for F4 and F5 samples [the F4 result is shown in Fig. 6(c)]. The ferrimagnetic scattering contributes 100% to the hkl peak (331), a large portion to the peaks (531), (222), (111), and some percentage to the (422), (400), and (220) peaks. The intensity of peaks with combination contributions from nuclear and magnetic structures are not precisely fit with either random or inverse structure; the average structures may lie somewhere in between. Since the nuclear Bragg peaks are overlapped with the magnetic Bragg peaks and there are a high number of participating cations, Bragg diffraction data alone is not enough to determine the detailed structure and cation site preferences of the high entropy spinels.

TABLE III.

Parameters for the best fitting model results among the normal, inverse and random solution Fd3̄m spinel models applied for F1, F3, F4, and F5 Rietveld refinements. The table includes the refined magnetic moment (M), lattice parameter (a), oxygen position (u), isotropic atomic displacement parameters (Uiso) for the 8b, 16c, and 32e sites, and the resulting refinement goodness-of-fit, Rwp.

SampleF1F3F4F5
Best model Inverse Inverse Random Random 
M (μB/Å) 0.406 0.215 1.162 0.532 
a (Å) 8.3796(1) 8.3793(1) 8.3894(1) 8.3839(1) 
u (frac.) 0.2440(0) 0.2438(0) 0.2433(0) 0.2436(0) 
Uiso,8b20.0057(1) 0.0066(1) 0.0054(1) 0.0069(1) 
Uiso,16c20.0040(1) 0.0035(1) 0.0047(1) 0.0069(1) 
Uiso,32e20.0079(1) 0.0088(1) 0.0085(1) 0.0084(1) 
Rwp(%) 5.19 5.28 5.14 4.54 
SampleF1F3F4F5
Best model Inverse Inverse Random Random 
M (μB/Å) 0.406 0.215 1.162 0.532 
a (Å) 8.3796(1) 8.3793(1) 8.3894(1) 8.3839(1) 
u (frac.) 0.2440(0) 0.2438(0) 0.2433(0) 0.2436(0) 
Uiso,8b20.0057(1) 0.0066(1) 0.0054(1) 0.0069(1) 
Uiso,16c20.0040(1) 0.0035(1) 0.0047(1) 0.0069(1) 
Uiso,32e20.0079(1) 0.0088(1) 0.0085(1) 0.0084(1) 
Rwp(%) 5.19 5.28 5.14 4.54 

A comparison of the neutron PDFs for the high entropy spinels is provided in Fig. 7(a), displayed between 1.7 and 10 Å. All pair correlations have been numbered and identified by category with a color key to aid discussion. First, it is immediately apparent that the local atomic configurations of the high entropy spinels in the present study are similar; in particular, the location and peak shapes of nearest-neighbor and next-nearest neighbor pair correlations are coincident across the four compositions. This immediately rules out significant local distortion of specific A site cations. Next, there are specific sets of peaks in the data that vary systematically from sample to sample at low r: the peaks highlighted by magenta (peak 1, 5, 9, 12, 15, 17), green (peak 2, 6, 10, 13), red (peak 4), and pink (peak 8) dashed lines representing A–O, Aoct–Aoct/O–O, O–O, and Aoct–Aoct/Atet–Atet/A–O correlations, respectively, are matched in amplitude; the peaks highlighted by sky blue (peak 3, 11, 16) and orange (peak 7, 14), representing Aoct–Atet/Atet–Atet/A–O and Aoct–Atet/Atet–O correlations, respectively, differ in amplitude. Inspection reveals that samples F1 and F3 exhibit similar and higher amplitudes of specific correlations relative to samples F4 and F5. PDF data comparisons for every 10 Å across the whole range (1.7–50 Å) are shown in supplementary material, Fig. S3.

FIG. 7.

(a) The PDF data comparison of F1 (orange line), F3 (purple line), F4 (red line), and F5 (green line). The six peaks (1, 5, 9, 12, 15, 17) highlighted with magenta dashed lines correspond to A–O pairs. The four peaks (2, 6, 10, 13) indicated with green dashed lines arise from Aoct–Aoct and O–O pair correlations. The three peaks (3, 11, 16) contributed by Aoct–Atet, Atet–Atet and A–O are identified in sky blue dashed lines. The peak (4) around 4 Å highlighted in a red dashed line corresponds to an O–O correlation. The combination of Aoct–Atet and Atet–O pair correlations contribute the two peaks (7, 14) indicated by orange dashed lines. The peak (8) highlighted by a pink dashed line arises from Aoct–Aoct, Atet–Atet and A–O pair correlations. A comparison of the PDF refinements for (b) F1 sample data and (c) F4 sample data using different structure models (random, inverse, and normal spinel ferrite models). PDF data for F1 and F4 samples are shown in orange and red dots, respectively. The blue, lemon green, and sky blue solid curves present fits to the data completed with the random solution spinel structure, inverse spinel structure, and normal spinel structure model, respectively. The fit difference curves displayed below the data and fits are represented as black solid curves.

FIG. 7.

(a) The PDF data comparison of F1 (orange line), F3 (purple line), F4 (red line), and F5 (green line). The six peaks (1, 5, 9, 12, 15, 17) highlighted with magenta dashed lines correspond to A–O pairs. The four peaks (2, 6, 10, 13) indicated with green dashed lines arise from Aoct–Aoct and O–O pair correlations. The three peaks (3, 11, 16) contributed by Aoct–Atet, Atet–Atet and A–O are identified in sky blue dashed lines. The peak (4) around 4 Å highlighted in a red dashed line corresponds to an O–O correlation. The combination of Aoct–Atet and Atet–O pair correlations contribute the two peaks (7, 14) indicated by orange dashed lines. The peak (8) highlighted by a pink dashed line arises from Aoct–Aoct, Atet–Atet and A–O pair correlations. A comparison of the PDF refinements for (b) F1 sample data and (c) F4 sample data using different structure models (random, inverse, and normal spinel ferrite models). PDF data for F1 and F4 samples are shown in orange and red dots, respectively. The blue, lemon green, and sky blue solid curves present fits to the data completed with the random solution spinel structure, inverse spinel structure, and normal spinel structure model, respectively. The fit difference curves displayed below the data and fits are represented as black solid curves.

Close modal

Neutron PDF refinements in Figs. 7(b) and 7(c) show the effect of changing the cation ordering model (inverse, normal, random configuration models) for F1 and F4 samples, respectively. Fits for F3 and F5 are shown in the supplementary material, Fig. S4, and provide similar results to F1 and F4, respectively. It is highly likely the unfit residual is in part due to magnetic spin correlations that are unaccounted for in the models.89,90 At present, there are no PDF data refinement programs that allow for shared site occupancy and magnetic PDF refinement; thus, we have elected to present the nuclear structure fits alone here and address magnetic spin correlations separately. The 15th pair amplitude (an A–O correlation at ∼8.75 Å) is shown to be sensitive to the degree of inversion in spinel ferrite models. Overall, for F1 and F3, the random model and inverse model provide a very similar fit quality and cannot be easily distinguished with PDF refinement. For sample F4 and F5, the fit quality for a random model is superior to that for an inverse model. These results correlate well to conclusions drawn from Rietveld refinement. However, no further conclusion can be made from PDF refinement due to the high numbers of cations present per site and the limitation of existing software for exploring local nuclear and magnetic cation ordering together.

As noted above, quantitative real-space refinement of the series of data is complicated by the presence of magnetic spin correlations and the large number of cations involved on the spinel tetrahedral and octahedral sublattices. Large box modeling of PDF data with nuclear and magnetic spin correlations was pursued to provide additional insight into structure–property characteristics in the series. Combined fits of F(Q) (supplementary material, Figs. S5 and S6) and Bragg (supplementary material, Figs. S7 and S8) intensities were completed for all samples assuming an inverse spinel model and a random solution spinel model (trials with a normal spinel model were not tried as they were ruled out conclusively by Rietveld analysis and small box PDF modeling). The agreement of fits to experimental data is high for both model types for all four samples. Table S5 (for F1 and F3) and Table S6 (for F4 and F5) in the supplementary material list the resulting anisotropic atomic displacement parameters (ADPs) on cation tetrahedral sites, cation octahedral sites, and oxygen sites of each RMC refinement for F1, F3, F4, and F5. For F4 and F5 samples, resulting ADP values for the tetrahedral sites and for oxygen positions in the random solution spinel model fits are found to be nearly half those found for inverse spinel model fits. Such an anomaly, combined with Rietveld refinement results, indicate a strong preference for a random model over an inverse one for F4 and F5 samples. Thus, the following discussion is based on analysis of the model configurations obtained from an inverse spinel model for samples F1 and F3 and a random solution spinel model for samples F4 and F5.

With the magnetic spin configuration obtained via RMC modeling for all samples, all magnetic moment vectors contained in the supercell were mapped into a single unit cell and then projected onto the XY plane. Results are shown visually in the supplementary material Figs. S9 and S10, and the magnetic spin distribution of F1 and F3 samples is more dispersed on both tetrahedral and octahedral sites than F4 and F5. In all samples, the magnetic spin distributions of octahedral sites are more disperse than those found on tetrahedral sites. Overall, less collinear spin order and stronger displacements from atom sites are observed for F1 and F3 samples. Two types of statistical analyses were completed to quantify the findings. First, a KDE analysis was conducted, capturing the proportion of magnetic moments found within a specified radius (with the variation ranging from 0 to 1). In principle, the faster such a proportion approaches 1, the stronger the collinearity in magnetic ordering. The results are presented in Fig. 8(a) for the tetrahedral site magnetic vector distribution and (b) for the octahedral site magnetic vector distribution, indicating a significantly more collinear spin arrangement in F4 and F5 samples than in F1 and F3 samples. This result is consistent with the smaller net magnetic moment in F1 and F3 samples—see Table S7 in the supplementary material. Additionally, in all samples, tetrahedral site spin correlations are found more collinear than octahedral site spin correlations.

FIG. 8.

The proportion of magnetic moment vectors with projection on the XY plane for samples F1, F3, F4, and F5 falling into specified radii, relative to atoms of (a) tetrahedral (Tetra) and (b) octahedral (Octa) sites that the corresponding magnetic moments belong to. Results presented correspond to an inverse spinel structure model for F1 and F3 sample data and a random solution spinel model for F4 and F5 sample data. It is noted that similar results are obtained when the alternate models are used (inverse spinel model for F1 and F3 sample data and random solution spinel model for F4 and F5 samples).

FIG. 8.

The proportion of magnetic moment vectors with projection on the XY plane for samples F1, F3, F4, and F5 falling into specified radii, relative to atoms of (a) tetrahedral (Tetra) and (b) octahedral (Octa) sites that the corresponding magnetic moments belong to. Results presented correspond to an inverse spinel structure model for F1 and F3 sample data and a random solution spinel model for F4 and F5 sample data. It is noted that similar results are obtained when the alternate models are used (inverse spinel model for F1 and F3 sample data and random solution spinel model for F4 and F5 samples).

Close modal

Second, the HZ statistic was calculated to quantify the non-Gaussian behavior of the oxygen coordinate distributions and magnetic moment vectors within the XY plane. Different starting cation distribution configurations are found to influence the HZ statistic values for F1 and F3 samples; however, values determined for F1 and F3 are generally larger than those found for F4 and F5. Overall, in Figs. 9(a) and 9(c), it can be observed that F1 and F3 samples exhibit a more non-Gaussian behavior in oxygen atom distributions and in magnetic moment vector distributions on the octahedral site than F4 and F5 samples. (HZ parameters for F4 and F5 samples are much closer to 1, i.e., no obvious non-Gaussian distortion could be observed.) Meanwhile, if focusing on magnetic moment vector distributions on the tetrahedral site [Fig. 9(b)], all four samples show HZ parameter values close to 1, indicating a nearly Gaussian distribution. These comprehensive results here suggest strong coupling between non-collinear magnetic correlations originating primarily from octahedral site spins and oxygen site displacements, especially in F1 and F3 samples. Variation in the level of spin–lattice coupling among high entropy spinels of similar composition demonstrates a potentially tunable interaction in the more extended family.

FIG. 9.

The HZ statistic for the distribution of (a) oxygen atom (O) coordinates and magnetic moment vectors (M) on (b) tetrahedral sites (Tetra) and (c) octahedral sites (Octa), for F1, F3, F4, and F5 samples. The horizontal dashed lines indicate a Gaussian distribution.

FIG. 9.

The HZ statistic for the distribution of (a) oxygen atom (O) coordinates and magnetic moment vectors (M) on (b) tetrahedral sites (Tetra) and (c) octahedral sites (Octa), for F1, F3, F4, and F5 samples. The horizontal dashed lines indicate a Gaussian distribution.

Close modal

We have determined the average ferrimagnetic structure in a family of AFe2O4 high entropy spinels with neutron Bragg diffraction analysis and found evidence for a variation in the tetrahedral and octahedral site occupancies of A site cations with local structure analysis of neutron total scattering data. These high entropy spinels are not found to have strictly random cation configuration structures (highest Sconfig), but rather, F1/F3 samples in this study were found to have configurations closer to an inverse spinel. The distinct populations of Fe and other magnetic cations on tetrahedral and octahedral sites in the samples manifests in variation in the prevalence and propagation of local spin clusters, as revealed by FORC analysis of magnetometry data. RMC modeling reveals spin non-collinearity in all four samples, enhanced in cases presenting a larger degree of cation inversion.

While identifying the cause for the variation in cation disorder among similar compositions falls outside the scope of the present study, we bring attention to a potential competition between crystal field stabilization effects (which promotes inverse spinel ferrite configurations for the compositions studied) and the entropy stabilization present in truly random spinel ferrite configurations. How much the relative effects of these phenomena are influenced by sample synthesis conditions or other factors, potentially creating samples in one of many different metastable states, remains unknown. Overall, this work highlights that it may prove extremely challenging to achieve a true solid solution with ideal/homogeneous cation mixing in compositionally complex ceramics with complex energy landscapes. We suggest great care be taken in this community to precisely describe reaction conditions, etc. when reporting HEO properties and drawing conclusions regarding the effects of configurational entropy on resultant products. Additionally, experimental characterization of the local to long-range crystal–chemical structure in HEOs may be key to determining and understanding their physical properties. The limitations of the characterization methods explored herein indicate future efforts are needed toward combined experimental probes and advanced data modeling. Hard and soft x-ray absorption near edge spectroscopy (XANES), soft x-ray magnetic circular dichroism (XMCD), and 57Fe Mössbaruer spectroscopy (with and without external magnetic field) are examples of techniques that may provide additional insights.

See the supplementary material for additional configurational entropy calculations, average structure results from Rietveld refinement, and local structure results from small-box PDF refinement and large box RMC modeling.

K.P. acknowledges financial support from the National Science Foundation (Grant No. DMR-2145174). This research used beamline 1B (NOMAD) of the Spallation Neutron Source, a DOE Office of Science User Facility operated by the Oak Ridge National Laboratory. We acknowledge Dr. Jue Liu for helpful discussion regarding Rietveld refinements and Namila Liyanage, Nan Tang, and Tianyu Li for helpful discussion regarding magnetometry results. D.A.G. and C.K. were supported by the U.S. Department of Energy (Grant No. DE-SC0021344) Partial funding for open access to this research was provided by University of Tennessee′s Open Publishing Support Fund.

The authors have no conflicts to disclose.

Xin Wang: Conceptualization (equal); Formal analysis (equal); Investigation (lead); Methodology (equal); Writing – original draft (lead); Writing – review & editing (equal). Brianna Musicó: Conceptualization (equal); Investigation (equal); Methodology (equal); Writing – review & editing (equal). Corisa Kons: Conceptualization (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Writing – original draft (equal); Writing – review & editing (equal). Peter Metz: Formal analysis (equal); Methodology (equal); Writing – review & editing (equal). Veerle Keppens: Conceptualization (equal); Methodology (equal); Supervision (equal); Writing – review & editing (equal). Dustin A. Gilbert: Conceptualization (equal); Funding acquisition (equal); Methodology (equal); Supervision (equal); Writing – review & editing (equal). Yuanpeng Zhang: Formal analysis (equal); Investigation (equal); Methodology (equal); Writing – original draft (equal); Writing – review & editing (equal). Katharine Page: Conceptualization (lead); Funding acquisition (equal); Investigation (equal); Methodology (equal); Supervision (equal); Writing – original draft (equal); Writing – review & editing (lead).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
S.
Jiang
,
T.
Hu
,
J.
Gild
,
N.
Zhou
,
J.
Nie
,
M.
Qin
,
T.
Harrington
,
K.
Vecchio
, and
J.
Luo
, “
A new class of high-entropy perovskite oxides
,”
Scr. Mater.
142
,
116
120
(
2018
).
2.
R.
Witte
,
A.
Sarkar
,
R.
Kruk
,
B.
Eggert
,
R. A.
Brand
,
H.
Wende
, and
H.
Hahn
, “
High-entropy oxides: An emerging prospect for magnetic rare-earth transition metal perovskites
,”
Phys. Rev. Mater.
3
,
034406
(
2019
).
3.
R.
Witte
,
A.
Sarkar
,
L.
Velasco
,
R.
Kruk
,
R. A.
Brand
,
B.
Eggert
,
K.
Ollefs
,
E.
Weschke
,
H.
Wende
, and
H.
Hahn
, “
Magnetic properties of rare-earth and transition metal based perovskite type high entropy oxides
,”
J. Appl. Phys.
127
,
185109
(
2020
).
4.
D. A.
Vinnik
,
E. A.
Trofimov
,
V. E.
Zhivulin
,
S. A.
Gudkova
,
O. V.
Zaitseva
,
D. A.
Zherebtsov
,
A. Y.
Starikov
,
D. P.
Sherstyuk
,
A. A.
Amirov
,
A. V.
Kalgin
 et al., “
High entropy oxide phases with perovskite structure
,”
Nanomaterials
10
,
268
(
2020
).
5.
Z.
Teng
,
L.
Zhu
,
Y.
Tan
,
S.
Zeng
,
Y.
Xia
,
Y.
Wang
, and
H.
Zhang
, “
Synthesis and structures of high-entropy pyrochlore oxides
,”
J. Eur. Ceram. Soc.
40
,
1639
1643
(
2020
).
6.
A.
Sarkar
,
Q.
Wang
,
A.
Schiele
,
M. R.
Chellali
,
S. S.
Bhattacharya
,
D.
Wang
,
T.
Brezesinski
,
H.
Hahn
,
L.
Velasco
, and
B.
Breitung
, “
High-entropy oxides: Fundamental aspects and electrochemical properties
,”
Adv. Mater.
31
,
1806236
(
2019
).
7.
H.
Li
,
Y.
Zhou
,
Z.
Liang
,
H.
Ning
,
X.
Fu
,
Z.
Xu
,
T.
Qiu
,
W.
Xu
,
R.
Yao
, and
J.
Peng
, “
High-entropy oxides: Advanced research on electrical properties
,”
Coatings
11
,
628
(
2021
).
8.
R.-Z.
Zhang
and
M. J.
Reece
, “
Review of high entropy ceramics: Design, synthesis, structure and properties
,”
J. Mater. Chem. A
7
,
22148
22162
(
2019
).
9.
J.
Liu
,
G.
Shao
,
D.
Liu
,
K.
Chen
,
K.
Wang
,
B.
Ma
,
K.
Ren
, and
Y.
Wang
, “
Design and synthesis of chemically complex ceramics from the perspective of entropy
,”
Mater. Today Adv.
8
,
100114
(
2020
).
10.
B. L.
Musicó
,
D.
Gilbert
,
T. Z.
Ward
,
K.
Page
,
E.
George
,
J.
Yan
,
D.
Mandrus
, and
V.
Keppens
, “
The emergent field of high entropy oxides: Design, prospects, challenges, and opportunities for tailoring material properties
,”
APL Mater.
8
,
040912
(
2020
).
11.
C.
Oses
,
C.
Toher
, and
S.
Curtarolo
, “
High-entropy ceramics
,”
Nat. Rev. Mater.
5
,
295
309
(
2020
).
12.
S. H.
Albedwawi
,
A.
AlJaberi
,
G. N.
Haidemenopoulos
, and
K.
Polychronopoulou
, “
High entropy oxides-exploring a paradigm of promising catalysts: A review
,”
Mater. Des.
202
,
109534
(
2021
).
13.
Y.
Sun
and
S.
Dai
, “
High-entropy materials for catalysis: A new frontier
,”
Sci. Adv.
7
,
eabg1600
(
2021
).
14.
H.
Xiang
,
Y.
Xing
,
F.-z.
Dai
,
H.
Wang
,
L.
Su
,
L.
Miao
,
G.
Zhang
,
Y.
Wang
,
X.
Qi
,
L.
Yao
 et al., “
High-entropy ceramics: Present status, challenges, and a look forward
,”
J. Adv. Ceram.
10
,
385
(
2021
).
15.
A.
Salian
and
S.
Mandal
, “
Entropy stabilized multicomponent oxides with diverse functionality—A review
,”
Crit. Rev. Solid State Mater. Sci.
47
,
142
(
2021
).
16.
A.
Sarkar
,
B.
Breitung
, and
H.
Hahn
, “
High entropy oxides: The role of entropy, enthalpy and synergy
,”
Scr. Mater.
187
,
43
48
(
2020
).
17.
M.
Fu
,
X.
Ma
,
K.
Zhao
,
X.
Li
, and
D.
Su
, “
High-entropy materials for energy-related applications
,”
iScience
24
,
102177
(
2021
).
18.
S. J.
McCormack
and
A.
Navrotsky
, “
Thermodynamics of high entropy oxides
,”
Acta Mater.
202
,
1
21
(
2021
).
19.
C. M.
Rost
,
E.
Sachet
,
T.
Borman
,
A.
Moballegh
,
E. C.
Dickey
,
D.
Hou
,
J. L.
Jones
,
S.
Curtarolo
, and
J. P.
Maria
, “
Entropy-stabilized oxides
,”
Nat. Commun.
6
,
8485
(
2015
).
20.
A.
Sarkar
,
R.
Kruk
, and
H.
Hahn
, “
Magnetic properties of high entropy oxides
,”
Dalton Trans.
50
,
1973
1982
(
2021
).
21.
A.
Navrotsky
and
O. J.
Kleppa
, “
Thermodynamics of formation of simple spinels
,”
J. Inorg. Nucl. Chem.
30
,
479
498
(
1968
).
22.
A.
Navrotsky
and
O. J.
Kleppa
, “
The thermodynamics of cation distributions in simple spinels
,”
J. Inorg. Nucl. Chem.
29
,
2701
2714
(
1967
).
23.
E. J. W.
Verwey
and
E. L.
Heilmann
, “
Physical properties and cation arrangement of oxides with spinel structures I. Cation arrangement in spinels
,”
J. Chem. Phys.
15
,
174
180
(
1947
).
24.
E. W.
Gorter
, “
Magnetization in ferrites: Saturation magnetization of ferrites with spinel structure
,”
Nature
165
,
798
800
(
1950
).
25.
A.
El Maazouzi
,
R.
Masrour
, and
A.
Jabar
, “
Thickness-dependent magnetic properties of inverse spinel Fe3O4
,”
Phase Transitions
93
,
733
740
(
2020
).
26.
Nitika
,
R.
Anu
, and
K.
Vinod
, “
Study of magnetic and optical transitions in MFe2O4 (M=Co, Zn, Fe, Mn) with spinel structure
,”
Nanosyst.: Phys., Chem., Math.
12
,
481
491
(
2021
).
27.
K.
Bouferrache
,
Z.
Charifi
,
H.
Baaziz
,
A. M.
Alsaad
, and
A.
Telfah
, “
Electronic structure, magnetic and optic properties of spinel compound NiFe2O4
,”
Semicond. Sci. Technol.
35
,
095013
(
2020
).
28.
S. S.
Selima
,
M.
Khairy
, and
M. A.
Mousa
, “
Comparative studies on the impact of synthesis methods on structural, optical, magnetic and catalytic properties of CuFe2O4
,”
Ceram. Int.
45
,
6535
6540
(
2019
).
29.
A.
Blößer
,
H.
Kurz
,
J.
Timm
,
F.
Wittkamp
,
C.
Simon
,
S.
Hayama
,
B.
Weber
,
U.-P.
Apfel
, and
R.
Marschall
, “
Tailoring the size, inversion parameter, and absorption of phase-pure magnetic MgFe2O4 nanoparticles for photocatalytic degradations
,”
ACS Appl. Nano Mater.
3
,
11587
11599
(
2020
).
30.
F.
Sharifianjazi
,
M.
Moradi
,
N.
Parvin
,
A.
Nemati
,
A.
Jafari Rad
,
N.
Sheysi
,
A.
Abouchenari
,
A.
Mohammadi
,
S.
Karbasi
,
Z.
Ahmadi
 et al., “
Magnetic CoFe2O4 nanoparticles doped with metal ions: A review
,”
Ceram. Int.
46
,
18391
18412
(
2020
).
31.
J.
Venturini
,
A. M.
Tonelli
,
T. B.
Wermuth
,
R. Y. S.
Zampiva
,
S.
Arcaro
,
A.
Da Cas Viegas
, and
C. P.
Bergmann
, “
Excess of cations in the sol-gel synthesis of cobalt ferrite (CoFe2O4): A pathway to switching the inversion degree of spinels
,”
J. Magn. Magn. Mater.
482
,
1
8
(
2019
).
32.
M.
Sugimoto
, “
The past, present, and future of ferrites
,”
J. Am. Ceram. Soc.
82
,
269
280
(
1999
).
33.
C.-W.
Nan
,
M.
Bichurin
,
S.
Dong
,
D.
Viehland
, and
G.
Srinivasan
, “
Multiferroic magnetoelectric composites: Historical perspective, status, and future directions
,”
J. Appl. Phys.
103
,
031101
(
2008
).
34.
J.
Dąbrowa
,
M.
Stygar
,
A.
Mikuła
,
A.
Knapik
,
K.
Mroczka
,
W.
Tejchman
,
M.
Danielewski
, and
M.
Martin
, “
Synthesis and microstructure of the (Co,Cr,Fe,Mn,Ni)3O4 high entropy oxide characterized by spinel structure
,”
Mater. Lett.
216
,
32
36
(
2018
).
35.
B.
Musicó
,
Q.
Wright
,
T. Z.
Ward
,
A.
Grutter
,
E.
Arenholz
,
D.
Gilbert
,
D.
Mandrus
, and
V.
Keppens
, “
Tunable magnetic ordering through cation selection in entropic spinel oxides
,”
Phys. Rev. Mater.
3
,
104416
(
2019
).
36.
Z.
Grzesik
,
G.
Smoła
,
M.
Miszczak
,
M.
Stygar
,
J.
Dąbrowa
,
M.
Zajusz
,
K.
Świerczek
, and
M.
Danielewski
, “
Defect structure and transport properties of (Co,Cr,Fe,Mn,Ni)3O4 spinel-structured high entropy oxide
,”
J. Eur. Ceram. Soc.
40
,
835
839
(
2020
).
37.
D.
Wang
,
S.
Jiang
,
C.
Duan
,
J.
Mao
,
Y.
Dong
,
K.
Dong
,
Z.
Wang
,
S.
Luo
,
Y.
Liu
, and
X.
Qi
, “
Spinel-structured high entropy oxide (FeCoNiCrMn)3O4 as anode towards superior lithium storage performance
,”
J. Alloys Compd.
844
,
156158
(
2020
).
38.
J.
Cieslak
,
M.
Reissner
,
K.
Berent
,
J.
Dabrowa
,
M.
Stygar
,
M.
Mozdzierz
, and
M.
Zajusz
, “
Magnetic properties and ionic distribution in high entropy spinels studied by Mössbauer and ab initio methods
,”
Acta Mater.
206
,
116600
(
2021
).
39.
C.-C.
Lin
,
C.-W.
Chang
,
C.-C.
Kaun
, and
Y.-H.
Su
, “
Stepwise evolution of photocatalytic spinel-structured (Co,Cr,Fe,Mn,Ni)3O4 high entropy oxides from first-principles calculations to machine learning
,”
Crystals
11
,
1035
(
2021
).
40.
S.
Marik
,
D.
Singh
,
B.
Gonano
,
F.
Veillon
,
D.
Pelloquin
, and
Y.
Bréard
, “
Enhanced magnetic frustration in a new high entropy diamond lattice spinel oxide
,”
Scr. Mater.
186
,
366
369
(
2020
).
41.
S.
Marik
,
D.
Singh
,
B.
Gonano
,
F.
Veillon
,
D.
Pelloquin
, and
Y.
Bréard
, “
Long range magnetic ordering and magneto-(di) electric effect in a new class of high entropy spinel oxide
,”
Scr. Mater.
183
,
107
110
(
2020
).
42.
M.
Stygar
,
J.
Dąbrowa
,
M.
Moździerz
,
M.
Zajusz
,
W.
Skubida
,
K.
Mroczka
,
K.
Berent
,
K.
Świerczek
, and
M.
Danielewski
, “
Formation and properties of high entropy oxides in Co-Cr-Fe-Mg-Mn-Ni-O system: Novel (Cr,Fe,Mg,Mn,Ni)3O4 and (Co,Cr,Fe,Mg,Mn)3O4 high entropy spinels
,”
J. Eur. Ceram. Soc.
40
,
1644
1650
(
2020
).
43.
M.
Fracchia
,
M.
Manzoli
,
U.
Anselmi-Tamburini
, and
P.
Ghigna
, “
A new eight-cation inverse high entropy spinel with large configurational entropy in both tetrahedral and octahedral sites: Synthesis and cation distribution by X-ray absorption spectroscopy
,”
Scr. Mater.
188
,
26
31
(
2020
).
44.
B.
Liang
,
Y.
Ai
,
Y.
Wang
,
C.
Liu
,
S.
Ouyang
, and
M.
Liu
, “
Spinel-type (FeCoCrMnZn)3O4 high-entropy oxide: Facile preparation and supercapacitor performance
,”
Materials
13
,
5798
(
2020
).
45.
D. A.
Vinnik
,
E. A.
Trofimov
, and
V. E.
Zhivulin
, “
The conditions and results of ferrite based on multicomponent crystal formation in high entropic oxide systems
,”
Solid State Phenom.
299
,
246
251
(
2020
).
46.
Y.
Zhang
,
T.
Lu
,
Y.
Ye
,
W.
Dai
,
Y.
Zhu
, and
Y.
Pan
, “
Stabilizing oxygen vacancy in entropy-engineered CoFe2O4-type catalysts for co-prosperity of efficiency and stability in an oxygen evolution reaction
,”
ACS Appl. Mater. Interfaces
12
,
32548
32555
(
2020
).
47.
T.-Y.
Chen
,
S.-Y.
Wang
,
C.-H.
Kuo
,
S.-C.
Huang
,
M.-H.
Lin
,
C.-H.
Li
,
H.-Y. T.
Chen
,
C.-C.
Wang
,
Y.-F.
Liao
,
C.-C.
Lin
 et al., “
In operando synchrotron X-ray studies of a novel spinel (Ni0.2Co0.2Mn0.2Fe0.2Ti0.2)3O4 high-entropy oxide for energy storage applications
,”
J. Mater. Chem. A
8
,
21756
21770
(
2020
).
48.
H.
Chen
,
N.
Qiu
,
B.
Wu
,
Z.
Yang
,
S.
Sun
, and
Y.
Wang
, “
A new spinel high-entropy oxide (Mg0.2Ti0.2Zn0.2Cu0.2Fe0.2)3O4 with fast reaction kinetics and excellent stability as an anode material for lithium ion batteries
,”
RSC Adv.
10
,
9736
9744
(
2020
).
49.
J.
Ma
,
B.
Zhao
,
H.
Xiang
,
F.-Z.
Dai
,
Y.
Liu
,
R.
Zhang
, and
Y.
Zhou
, “
High-entropy spinel ferrites MFe2O4 (M = Mg, Mn, Fe, Co, Ni, Cu, Zn) with tunable electromagnetic properties and strong microwave absorption
,”
J. Adv. Ceram.
11
,
754
768
(
2022
).
50.
A.
Radoń
,
Ł.
Hawełek
,
D.
Łukowiec
,
J.
Kubacki
, and
P.
Włodarczyk
, “
Dielectric and electromagnetic interference shielding properties of high entropy (Zn,Fe,Ni,Mg,Cd)Fe2O4 ferrite
,”
Sci. Rep.
9
,
20078
(
2019
).
51.
S. K.
Shaw
,
A.
Gangwar
,
A.
Sharma
,
S. K.
Alla
,
S.
Kavita
,
M.
Vasundhara
,
S. S.
Meena
,
P.
Maiti
, and
N. K.
Prasad
, “
Structural and magnetic properties of nanocrystalline equi-atomic spinel high-entropy oxide (AlCoFeMnNi)3O4 synthesised by microwave assisted co-precipitation technique
,”
J. Alloys Compd.
878
,
160269
(
2021
).
52.
A.
Mao
,
F.
Quan
,
H.-Z.
Xiang
,
Z.-G.
Zhang
,
K.
Kuramoto
, and
A.-L.
Xia
, “
Facile synthesis and ferrimagnetic property of spinel (CoCrFeMnNi)3O4 high-entropy oxide nanocrystalline powder
,”
J. Mol. Struct.
1194
,
11
18
(
2019
).
53.
S.
Dai
,
M.
Li
,
X.
Wang
,
H.
Zhu
,
Y.
Zhao
, and
Z.
Wu
, “
Fabrication and magnetic property of novel (Co,Zn,Fe,Mn,Ni)3O4 high-entropy spinel oxide
,”
J. Magn. Magn. Mater.
536
,
168123
(
2021
).
54.
A.
Mao
,
H.-Z.
Xiang
,
Z.-G.
Zhang
,
K.
Kuramoto
,
H.
Zhang
, and
Y.
Jia
, “
A new class of spinel high-entropy oxides with controllable magnetic properties
,”
J. Magn. Magn. Mater.
497
,
165884
(
2020
).
55.
A.
Mao
,
H.-X.
Xie
,
H.-Z.
Xiang
,
Z.-G.
Zhang
,
H.
Zhang
, and
S.
Ran
, “
A novel six-component spinel-structure high-entropy oxide with ferrimagnetic property
,”
J. Magn. Magn. Mater.
503
,
166594
(
2020
).
56.
H.-Z.
Xiang
,
H.-X.
Xie
,
A.
Mao
,
Y.-G.
Jia
, and
T.-Z.
Si
, “
Facile preparation of single phase high-entropy oxide nanocrystalline powders by solution combustion synthesis
,”
Int. J. Mater. Res.
111
,
246
249
(
2020
).
57.
H.
Zhu
,
H.
Xie
,
Y.
Zhao
,
S.
Dai
,
M.
Li
, and
X.
Wang
, “
Structure and magnetic properties of a class of spinel high-entropy oxides
,”
J. Magn. Magn. Mater.
535
,
168063
(
2021
).
58.
F.
Hosseini Mohammadabadi
,
S. M.
Masoudpanah
,
S.
Alamolhoda
, and
H. R.
Koohdar
, “
Electromagnetic microwave absorption properties of high entropy spinel ferrite ((MnNiCuZn)1−xCoxFe2O4)/graphene nanocomposites
,”
J. Mater. Res. Technol.
14
,
1099
1111
(
2021
).
59.
H.-Z.
Xiang
,
H.-X.
Xie
,
Y.-X.
Chen
,
H.
Zhang
,
A.
Mao
, and
C.-H.
Zheng
, “
Porous spinel-type (Al0.2CoCrFeMnNi)0.58O4δ high-entropy oxide as a novel high-performance anode material for lithium-ion batteries
,”
J. Mater. Sci.
56
,
8127
8142
(
2021
).
60.
T.
Parida
,
A.
Karati
,
K.
Guruvidyathri
,
B. S.
Murty
, and
G.
Markandeyulu
, “
Novel rare-earth and transition metal-based entropy stabilized oxides with spinel structure
,”
Scr. Mater.
178
,
513
517
(
2020
).
61.
G.
Wang
,
J.
Qin
,
Y.
Feng
,
B.
Feng
,
S.
Yang
,
Z.
Wang
,
Y.
Zhao
, and
J.
Wei
, “
Sol–gel synthesis of spherical mesoporous high-entropy oxides
,”
ACS Appl. Mater. Interfaces
12
,
45155
45164
(
2020
).
62.
D.
Wang
,
Z.
Liu
,
S.
Du
,
Y.
Zhang
,
H.
Li
,
Z.
Xiao
,
W.
Chen
,
R.
Chen
,
Y.
Wang
,
Y.
Zou
, and
S.
Wang
, “
Low-temperature synthesis of small-sized high-entropy oxides for water oxidation
,”
J. Mater. Chem. A
7
,
24211
24216
(
2019
).
63.
N. J.
Usharani
,
H.
Sanghavi
, and
S. S.
Bhattacharya
, “
Factors influencing phase formation and band gap studies of a novel multicomponent high entropy (Co,Cu,Mg,Ni,Zn)2TiO4 orthotitanate spinel
,”
J. Alloys Compd.
888
,
161390
(
2021
).
64.
A. H.
Phakatkar
,
M. T.
Saray
,
M. G.
Rasul
,
L. V.
Sorokina
,
T. G.
Ritter
,
T.
Shokuhfar
, and
R.
Shahbazian-Yassar
, “
Ultrafast synthesis of high entropy oxide nanoparticles by flame spray pyrolysis
,”
Langmuir
37
,
9059
9068
(
2021
).
65.
T. X.
Nguyen
,
J.
Patra
,
J.-K.
Chang
, and
J.-M.
Ting
, “
High entropy spinel oxide nanoparticles for superior lithiation–delithiation performance
,”
J. Mater. Chem. A
8
,
18963
18973
(
2020
).
66.
T.
Hirakawa
,
Y.
Shimokawa
,
W.
Tokuzumi
,
T.
Sato
,
M.
Tsushida
,
H.
Yoshida
,
S.
Hinokuma
,
J.
Ohyama
, and
M.
Machida
, “
Multicomponent spinel oxide solid solutions: A possible alternative to platinum group metal three-way catalysts
,”
ACS Catal.
9
,
11763
11773
(
2019
).
67.
L.
Danyang
,
S.
Liping
,
L.
Qiang
,
X.
Tian
,
H.
Lihua
, and
Z.
Hui
, “
High-entropy oxide (Fe0.2Zn0.2Co0.2Ni0.2Cu0.2)Fe2O4: An efficient and stable spinel-type electrocatalyst for H2O2 production in alkaline media
,”
J. Alloys Compd.
913
,
165148
(
2022
).
68.
A.
Sarkar
,
B.
Eggert
,
R.
Witte
,
J.
Lill
,
L.
Velasco
,
Q.
Wang
,
J.
Sonar
,
K.
Ollefs
,
S. S.
Bhattacharya
,
R. A.
Brand
 et al., “
Comprehensive investigation of crystallographic, spin-electronic and magnetic structure of (Co0.2Cr0.2Fe0.2Mn0.2Ni0.2)3O4: Unraveling the suppression of configuration entropy in high entropy oxides
,”
Acta Mater.
226
,
117581
(
2021
).
69.
K. R.
Sanchez-Lievanos
,
J. L.
Stair
, and
K. E.
Knowles
, “
Cation distribution in spinel ferrite nanocrystals: Characterization, impact on their physical properties, and opportunities for synthetic control
,”
Inorg. Chem.
60
,
4291
4305
(
2021
).
70.
L.
Cervera-Gabalda
,
A.
Zielińska-Jurek
, and
C.
Gómez-Polo
, “
Tuning the photocatalytic performance through magnetization in Co–Zn ferrite nanoparticles
,”
J. Magn. Magn. Mater.
560
,
169617
(
2022
).
71.
G. C.
Lavorato
,
R.
Das
,
Y.
Xing
,
J.
Robles
,
F. J.
Litterst
,
E.
Baggio-Saitovitch
,
M.-H.
Phan
, and
H.
Srikanth
, “
Origin and shell-driven optimization of the heating power in core/shell bimagnetic nanoparticles
,”
ACS Appl. Nano Mater.
3
,
1755
1765
(
2020
).
72.
C. R.
Pike
,
A. P.
Roberts
, and
K. L.
Verosub
, “
Characterizing interactions in fine magnetic particle systems using first order reversal curves
,”
J. Appl. Phys.
85
,
6660
6667
(
1999
).
73.
A. P.
Roberts
,
C. R.
Pike
, and
K. L.
Verosub
, “
First-order reversal curve diagrams: A new tool for characterizing the magnetic properties of natural samples
,”
J. Geophys. Res.: Solid Earth
105
,
28461
28475
, (
2000
).
74.
D. A.
Gilbert
,
G. T.
Zimanyi
,
R. K.
Dumas
,
M.
Winklhofer
,
A.
Gomez
,
N.
Eibagi
,
J. L.
Vicent
, and
K.
Liu
, “
Quantitative decoding of interactions in tunable nanomagnet arrays using first order reversal curves
,”
Sci. Rep.
4
,
4204
(
2014
).
75.
D. A.
Gilbert
,
P. D.
Murray
,
J.
De Rojas
,
R. K.
Dumas
,
J. E.
Davies
, and
K.
Liu
, “
Reconstructing phase-resolved hysteresis loops from first-order reversal curves
,”
Sci. Rep.
11
,
4018
(
2021
).
76.
J.
Neuefeind
,
M.
Feygenson
,
J.
Carruth
,
R.
Hoffmann
, and
K. K.
Chipley
, “
The nanoscale ordered materials diffractometer NOMAD at the spallation neutron source SNS
,”
Nucl. Instrum. Methods Phys. Res., Sect. B
287
,
68
75
(
2012
).
77.
M. T.
McDonnell
,
D. P.
Olds
,
K. L.
Page
,
J. C.
Neufeind
,
M. G.
Tucker
,
J. C.
Bilheux
,
W.
Zhou
, and
P. F.
Peterson
, “
ADDIE: Advanced diffraction environment—A software environment for analyzing neutron diffraction data
,”
Acta Crystallogr., Sect. A: Found. Adv.
73
,
a377
(
2017
).
78.
A. A.
Coelho
, “
TOPAS and TOPAS-academic: An optimization program integrating computer algebra and crystallographic objects written in C++
,”
J. Appl. Crystallogr.
51
,
210
218
(
2018
).
79.
C. L.
Farrow
,
P.
Juhas
,
J. W.
Liu
,
D.
Bryndin
,
E. S.
Božin
,
J.
Bloch
,
T.
Proffen
, and
S. J. L.
Billinge
, “
PDFfit2 and PDFgui: Computer programs for studying nanostructure in crystals
,”
J. Phys.: Condens. Matter
19
,
335219
(
2007
).
80.
M. G.
Tucker
,
D. A.
Keen
,
M. T.
Dove
,
A. L.
Goodwin
, and
Q.
Hui
, “
RMCProfile: Reverse Monte Carlo for polycrystalline materials
,”
J. Phys.: Condens. Matter
19
,
335218
(
2007
).
81.
D. A.
Keen
, “
A comparison of various commonly used correlation functions for describing total scattering
,”
J. Appl. Crystallogr.
34
,
172
177
(
2001
).
82.
Y.
Zhang
,
M.
Eremenko
,
V.
Krayzman
,
M. G.
Tucker
, and
I.
Levin
, “
New capabilities for enhancement of RMCProfile: Instrumental profiles with arbitrary peak shapes for structural refinements using the reverse Monte Carlo method
,”
J. Appl. Crystallogr.
53
,
1509
1518
(
2020
).
83.
N.
Henze
and
B.
Zirkler
, “
A class of invariant consistent tests for multivariate normality
,”
Commun. Stat.
19
,
3595
3617
(
1990
).
84.
S.
Flegel
,
M.
Möckel
, and
J.
Bennett
, “
Covariance size and the breakdown of Gaussianity in GEO uncertainty predictions
,” in
Proceedings of the 7th European Space Debris Conference
,
2017
.
85.
S. K.
Flegel
and
J. C.
Bennett
, “
State uncertainty normality detection
,”
J. Astronaut. Sci.
67
,
1044
1062
(
2020
).
86.
A.
Trujillo-Ortiz
,
R.
Hernandez-Walls
,
K.
Barba-Rojo
, and
L.
Cupul-Magana
, “
HZMVNTEST: Henze-Zirkler’s multivariate normality test
,” A MATLAB file,
2007
, pp.
81
83
.
87.
C. R.
Pike
,
C. A.
Ross
,
R. T.
Scalettar
, and
G.
Zimanyi
, “
First-order reversal curve diagram analysis of a perpendicular nickel nanopillar array
,”
Phys. Rev. B
71
,
134407
(
2005
).
88.
J. E.
Davies
,
O.
Hellwig
,
E. E.
Fullerton
,
G.
Denbeaux
,
J. B.
Kortright
, and
K.
Liu
, “
Magnetization reversal of Co/Pt multilayers: Microscopic origin of high-field magnetic irreversibility
,”
Phys. Rev. B
70
,
224434
(
2004
).
89.
B. A.
Frandsen
,
X.
Yang
, and
S. J. L.
Billinge
, “
Magnetic pair distribution function analysis of local magnetic correlations
,”
Acta Crystallogr., Sect. A: Found. Adv.
70
,
3
11
(
2014
).
90.
B. A.
Frandsen
and
S. J. L.
Billinge
, “
Magnetic structure determination from the magnetic pair distribution function (mPDF): Ground state of MnO
,”
Acta Crystallogr., Sect. A: Found. Adv.
71
,
325
334
(
2015
).

Supplementary Material