Fei
Li
*a,
Shi-Kuan
Sun
b,
Yinjuan
Chen
c,
Takashi
Naka
d,
Takeshi
Hashishin
e,
Jun
Maruyama
f and
Hiroya
Abe
*a
aJoining and Welding Research Institute, Osaka University, Osaka 5670047, Japan. E-mail: feili@jwri.osaka-u.ac.jp; h-abe@jwri.osaka-u.ac.jp
bSchool of Material Science and Energy Engineering, Foshan University, Foshan, 528000, China
cKey Laboratory of Precise Synthesis of Functional Molecules of Zhejiang Province, School of Science, Instrumentation and Service Center for Molecular Sciences, Westlake University, Hangzhou 310024, China
dNational Institute for Materials Science, Ibaraki 3050047, Japan
eFaculty of Advanced Science and Technology, Kumamoto University, Kumamoto, 8608555, Japan
fOsaka Research Institute of Industrial Science and Technology, Osaka 5368553, Japan
First published on 21st April 2022
Low-dimensional high-entropy materials, such as nanoparticles and two-dimensional (2D) layers, have great potential for catalysis and energy applications. However, it is still challenging to synthesize 2D layered high-entropy materials through a bottom-up soft chemistry method, due to the difficulty of mixing and assembling multiple elements in 2D layers. Here, we report a simple polyol process for the synthesis of a series of 2D layered high-entropy transition metal (Co, Cr, Fe, Mn, Ni, and Zn) hydroxides (HEHs), involving the hydrolysis and inorganic polymerization of metal-containing species in ethylene glycol media. The as-synthesized HEHs demonstrate 2D layered structures with interlayer distances ranging from 0.860 to 0.987 nm and homogeneous elemental distribution of designed equimolar stoichiometry in the layers. These 2D HEHs exhibit a low overpotential of 275 mV at 10 mA cm−2 in a 0.1 M KOH electrolyte for the oxygen evolution reaction. Superparamagnetic spinel-type high-entropy nanoparticles can also be obtained by annealing these HEHs. Our polyol approach creates opportunities for synthesizing low-dimensional high-entropy materials with promising properties and applications.
Two-dimensional (2D) materials have received increasing attention since the discovery of graphene.37–39 Great efforts have been dedicated to exploring new 2D materials due to their unique set of properties and broad applications.40,41 Unlike the large numbers of research studies on zero-dimensional (nanoparticle) and bulk HEMs, little attention is paid to 2D HEMs. Only limited studies have been reported on 2D HEMs, including high-entropy hydroxides (HEHs),42 HE oxides,43 HE carbides,44,45 HE dichalcogenides,46 HE halides,46 and HE phosphorus trisulfides.46,47 Like the synthesis of other 2D materials, the methods for making 2D HEMs can be divided into two categories, i.e., top-down and bottom-up.39 The top-down process starts with the synthesis of bulk HEMs as precursors, followed by exfoliation or etching for thinning them. Hosono et al. developed a mechanical exfoliation method to synthesize a series of high-entropy van der Waals 2D materials (dichalcogenides, halides, and phosphorus trisulfides) with an intriguing set of properties.46 Wang et al. developed high-entropy layered double hydroxides (HE-LDH) (FeAlCrCoNiZnCu)–OH intercalated with CO32− by a hydrothermal method and exfoliated them to ultrathin defective nanosheets with a diameter of 50 nm with argon plasma.42 Wang et al. also reported the synthesis of spinel-type high-entropy oxide (FeCrCoNiCu)3O4 nanosheets with diameters up to 100 nm by oxygen plasma exfoliation of HE-LDH.43 These plasma-exfoliated HE-LDH nanosheets demonstrated efficient electrochemical catalytic activity, such as in the oxygen evolution reaction (OER). Very recently, high-entropy 2D transition metal carbides (MXenes) were achieved through selective etching of Al layers from high-entropy MAX phases by Anasori et al. and almost at the same time by Yang et al.44,45 These high-entropy MXenes indicate promising applications including energy storage and catalysis related areas. It is worth noting that the stoichiometry of 2D HEMs may deviate from their original design due to the post-chemical exfoliation and etching.45 The bottom-up process starts with atomic ingredients and they were assembled together to form 2D HEMs. Miura et al. reported a hydrothermal synthesis of CO32− intercalated HE layered hydroxide (MgAlCoNiZn)–OH using water as solvent, showing aggregated hexagonal platelet crystals of 300 nm in size.48 Ting et al. reported HE (CoCrFeMnNi)-glycerate for a high performance OER by a solvothermal technique using glycerol as solvent.49 The HE glycerate has a layered lattice structure and yet a spherical morphology.49 Due to the difficulty in mixing multiple elements at the nanoscale and almost at the same time forming and maintaining the 2D morphology, it is significantly challenging to synthesize 2D HEMs through a bottom-up process.
In this work, we present an attempt to synthesize 2D layered HEHs through a bottom-up polyol approach. The polyol process is widely recognized as a unique soft chemistry method for the preparation of various low-dimensional materials.50,51 Six transition metal cations (Co, Cr, Fe, Mn, Ni, and Zn) were selected to give out six five-element and one six-element equimolar high-entropy compositions. A solvothermal treatment using ethylene glycol as solvent was carried out at 200 °C for only 2 h, involving the complexation of polyol with metal cations, hydrolysis and inorganic polymerization. 2D layered HEHs with a large area, high compositional homogeneity, and designed stoichiometry were successfully synthesized. Binary transition metal hydroxides (e.g. NiFe, CoNi, and ZnCo) are widely studied as electrocatalysts for water oxidation.52,53 Doping with additional metal elements in these binary hydroxides, which could tune their surface chemistry and electronic structure, has been used to tailor their OER performance. The combination of multiple 3d transition metal ions in the 2D layered HEHs would possibly tune the activity for the OER. Upon post-annealing at 200 °C, these 2D HEHs could be converted into spinel-type oxide nanoparticles. The corresponding HE spinels contain multiple 3d magnetic ions. It would be of interest if the HE multicomponent spinels have emergent phenomena by an order-by-disorder mechanism or entropy stabilization. The magnetic properties and the OER performance of these 2D HEHs and the corresponding oxide nanoparticles were preliminarily studied. A reduced overpotential for the OER and superparamagnetism were found in these 2D HEHs and their corresponding oxide nanoparticles, respectively.
Label | Metal elements |
---|---|
HEH-1# | Co, Cr, Fe, Mn, Ni |
HEH-2# | Co, Cr, Fe, Mn, Zn |
HEH-3# | Co, Cr, Fe, Ni, Zn |
HEH-4# | Co, Cr, Mn, Ni, Zn |
HEH-5# | Co, Fe, Mn, Ni, Zn |
HEH-6# | Cr, Fe, Mn, Ni, Zn |
HEH-7# | Co, Cr, Fe, Mn, Ni, Zn |
X-ray diffraction (XRD; Cu-Kα, 30 kV and 10 mA; Bruker D2, Germany) was used to determine the phase assemblage of the samples. The refinement of the powder-XRD patterns was performed using the GSAS software package. XRD patterns used for refinement were collected at 25 °C with 2θ ranging from 5° to 120° and at a scan rate of 0.002° s−1. The metal contents of the samples after 600 °C annealing were measured using an inductively coupled plasma mass spectrometer (ICP-MS, Ultima-2, HORIBA, Japan). The specific surface area of the samples was determined by using a nitrogen adsorption analyzer (Macsorb 1201, MOUNTECH, Japan). The microstructures and elemental distribution of the specimens were analyzed by using a scanning electron microscope (SEM, SU-70, HITACHI, Japan), transmission electron microscope (TEM, JEM 2100F, JEOL, Japan) equipped with an energy-dispersive spectrometer (EDS, Oxford), and atomic force microscope (AFM, NanoNavi S-image, SII, Japan). The pyrolysis behavior of the samples was monitored using a NETZSCH STA 2500 (Germany) in flowing air at temperatures ranging from 40 °C to 800 °C with a heating rate of 10 °C min−1. Fourier transform infrared spectra (FT-IR) were collected using an IRPrestige-21 (Shimadzu, Japan). The pH values of the solutions were measured by using a pH meter (F-52, Horiba, Japan). Magnetic properties were measured by using a superconducting quantum interference device magnetometer (MPMS-XL, Quantum Design, USA).
Electrochemical characterization was carried out using a three-electrode system at 25 °C using an Automatic Polarization System (HZ-3000, Hokuto Denko, Japan). The electrolyte was 0.1 M KOH, the counter electrode was carbon cloth, and the reference electrode was Hg/HgO/0.1 M KOH. 5 mg of sample and 1 mg of carbon black (TKB, TOKABLACK #3855) were dispersed in a mixture of 0.5 mL H2O, 0.4 mL ethanol, and 0.1 mL Nafion solution (Sigma Aldrich) [equivalent weight (molar mass/mol of ion-exchange sites) = 1100, 5 wt% in a mixture of lower aliphatic alcohols and 15–20% water]. The suspension was ultrasonicated for 30 min to get a homogeneous slurry. The working electrode was prepared by depositing 6 μL slurry onto a glassy carbon electrode with a diameter of 5 mm. Prior to recording cyclic voltammetry data, the working electrode was repeatedly scanned between −0.9 and 0.3 V at 50 mV s−1 in an argon atmosphere in order to obtain a stable voltammogram. The linear sweep voltammetry (LSV) curve for the oxygen evolution reaction was recorded in an O2 atmosphere at 10 mV s−1 from 0.2 to 0.8 V. The stability of the catalyst was evaluated by cyclic voltammetry up to 100 cycles. For calculating the current density, the active mass loading was normalized to 1 mg cm−2. The electrode potential (EHg/HgO) was converted to the RHE potential (ERHE) using E(RHE) = E(Hg/HgO) + 0.177 V + 0.059 × pH.
Samples | Chemical composition (at%) | pH | Interlayer distance (nm) | |||||
---|---|---|---|---|---|---|---|---|
Co | Cr | Fe | Mn | Ni | Zn | |||
HEH-1# EG-K | 20.3 | 23.1 | 24.5 | 9.3 | 22.8 | — | 7.9 | 0.839 |
HEH-2# EG-K | 21.8 | 22.8 | 25.2 | 7.6 | — | 22.5 | 8.0 | 0.982 |
HEH-3# EG-K | 18.1 | 21.0 | 22.5 | — | 20.2 | 18.2 | 8.2 | 0.985 |
HEH-4# EG-K | 21.3 | 24.3 | — | 8.2 | 24.1 | 22.0 | 8.5 | 0.963 |
HEH-5# EG-K | 21.4 | — | 24.6 | 10.0 | 22.2 | 21.8 | 8.4 | 0.846 |
HEH-6# EG-K | — | 22.9 | 24.5 | 8.6 | 24.6 | 19.5 | 8.4 | 0.850 |
HEH-1# EG-K/N | 20.0 | 20.4 | 19.3 | 20.2 | 20.1 | — | 11.5 | 0.860 |
HEH-1# H2O–K/N | 19.8 | 20.5 | 19.5 | 19.9 | 20.3 | — | 12.2 | — |
ICP-MS was used to determine the chemical compositions for these HEHs. The normalized elemental compositions are gathered in Table 2. ICP-MS results confirm the existence of Co, Cr, Fe, Ni, and Zn elements in a near equimolar ratio for EG-K products. However, the Mn content is much lower than designed, which is probably due to the high solubility of Mn-containing species (Table S1†). The pH of the supernatant after solvothermal treatment is about 8.0 (Table 2), and seldom changes with increasing solvothermal time. The basicity of KOAC is insufficient to precipitate Mn cations completely from multi-component-containing EG solutions, though it has been used to synthesize ternary Mn-containing spinels via a similar polyol process.55 NaOH was used to control the solution pH. Under EG-K/N conditions, whose supernatant pH is 11.5–11.9 (Table 2 and S2†), all HEH-1#–7# EG-K/N as-synthesized products possess similar diffraction patterns compared with EG-K ones, as shown in Fig. 1 and S1.† The calculated interlayer distances of EG-K/N as-synthesized products are slightly larger than the EG-K ones (Table S2†). Taking HEH-1# as an example, an almost perfect equimolar composition can be achieved under these EG-K/N conditions (Table 2). In order to investigate the influence of solvent on the features of the samples, water is used as solvent in HEH-1# cases. In the case of H2O–K/N, whose supernatant pH is 12.2, a precipitate with five metal species in an equimolar ratio can also be obtained after hydrothermal treatment (Table 2). The H2O–K/N as-synthesized powders are well crystallized and can be indexed to a spinel structure (Fig. 1a), with the absence of layered structures. The main factor that governs the composition of HEHs in our polyol process at 200 °C is the solution pH. In order to obtain the designed high-entropy composition, it is necessary to carry out the polyol process at a moderate pH that can precipitate the most soluble metal species. It is Mn in our cases. However, higher pH (>13) and/or temperature (>230 °C) will result in the reduction of metal cations during the polyol process, as shown in Fig. S2.†
Under EG-K/N conditions, in the presence of a moderate amount of OH− in EG solution, the as-synthesized products present a typical layered morphology, agreeing well with the XRD result, as shown in Fig. 2d and S4a–c.†Fig. 2e shows the HRTEM image of the layered materials, indicating their amorphous structures. All five metal elements are uniformly distributed in the layered structures. No obvious elemental segregation can be observed in the EG-K/N sample at both the submicron and the nanoscale, as shown in Fig. 2f and S4d.† A higher OH− concentration (i.e., pH) in EG solution will improve the elemental homogeneity compared with the HEH-1# EG-K sample. The AFM image of the HEH-1# EG-K/N as-synthesized sample shows that a thickness of one piece in the sample is about 100 nm (Fig. S4e†). Some irregularly shaped particles are also observed on its surface.
When using water as solvent, the HEH-1# H2O–K/N as-synthesized powders demonstrate a complicated scenario (Fig. 2g and S5a–c†). Spherical and faceted particles along with large numbers of irregularly shaped particles can be observed. The particle sizes differ greatly from each other. Clear lattice fringes are observed from HRTEM images (Fig. 2h and S5c†), indicating that both the shaped and irregular particles are well crystallized. Elemental mapping results for H2O–K/N as-synthesized powders (Fig. 2i and S5d†) demonstrate that all five metal element segregation occurs at both the submicron and the nanoscale. By comparing HEH-1# EG-K/N and H2O–K/N samples, it can be concluded that EG solvent plays a key role in inhibiting elemental segregation.
STEM-EDS quantitative analyses of the HEH-1# as-synthesized products are provided in Fig. 3. In the case of EG-K, both Fe- and Ni-rich clusters are confirmed, as shown in Fig. 3a. The average content values for the five metal elements in the selected area are well consistent with the ICP-MS data. The relative contents of the five metal elements are close to each other in different particles for Fe- and Ni-rich clusters, respectively. The normalized metal composition is Cr20.7±0.7Mn9.9±0.5Fe36.2±1.7Co22.2±0.7Ni11.0±0.8 and Cr20.9±0.4Mn7.4±0.1Fe7.6±0.3Co23.0±0.2Ni41.1±0.5 for Fe-rich clusters and Ni-rich clusters, respectively. The Ni-rich cluster is probably a compositionally complex single-phase compound, which explains why all five metal elements are homogeneously distributed at the nanoscale (Fig. S3c†). However, the Fe-rich cluster is probably a compositionally complex mixture at the nanoscale, as supported by the obvious Ni segregation shown in Fig. S3e.† In the case of EG-K/N, the EDS quantitative analysis (Fig. 3b) reveals the excellent compositional homogeneity of the layered materials. The five metal cations are in an equimolar ratio at different positions in the layered structures. For H2O–K/N as-synthesized products, the composition of the particles differs greatly from each other, as shown in Fig. 3c. Each particles contain five metal elements with different stoichiometry, which precipitate separately from solution. The particles share a near equimolar composition on average, agreeing well with the ICP-MS result (Table 2). As aforementioned, two compounds with different stoichiometry (i.e., Fe- and Ni-rich), one compound with the five metal elements in an equimolar ratio, and numbers of spinel-type compounds with various stoichiometry are formed for HEH-1# composition after solvothermal treatment under EG-K, EG-K/N, and H2O–K/N conditions, respectively. This assumption could facilitate understanding the elemental segregation in the as-synthesized products. Elements should be homogeneously distributed inside one single compound, while elemental separation occurs in compounds with different stoichiometry.
Fig. 3 TEM images of the HEH-1# as-synthesized products under various synthesis conditions and corresponding chemical compositions. (a) EG-K. (b) EG-K/N. (c) H2O–K/N. |
Fig. 4 and S6† compare the microstructure and elemental distribution of the HEH-2#–7# as-synthesized products under EG-K/N and EG-K conditions, respectively. The morphological changes for HEH-2#–7# samples show a similar trend to the HEH-1# ones when increasing the solution pH. For EG-K, the HEH-2#–7#, except for HEH-5#, as-synthesized products generally are aggregates of spherical clusters with elemental segregation at the submicron scale. The most obvious elemental segregation occurs for Ni, Zn, and Fe, as compared in Fig. S6,† probably due to the differences in standard electrode potentials and solubility product constants, or a relatively high solvothermal temperature (200 °C).51 It is worth noting that, for HEH-5# EG-K, the as-synthesized products are aggregates of crumpled nano-sheets with uniform elemental distribution (Fig. S6d†). Increasing the solution pH, the HEH-2#–7# as-synthesized products appear to be layered structures with sizes up to several hundreds of nanometers, as depicted in Fig. 4. These layered materials demonstrate improved elemental homogeneity compared with their corresponding EG-K samples. Taking HEH-5# EG-K/N for an example, all five metal elements are homogeneously distributed in these thin flat sheets, as shown in Fig. 4d and S7.† According to the aforementioned results and discussion, our polyol process is proved to be a simple and effective method for synthesizing layered HEHs with high compositional homogeneity and desired stoichiometry. By simply tuning the solution pH, the morphology of the as-synthesized products can be transformed from the spherical clusters into 2D layered sheets. Some irregularly shaped particles can also be observed in Fig. 4, which warrants further optimization of the synthesis conditions.
Fig. 4 TEM images and elemental maps for HEH-2#–7# EG-K/N as-synthesized products. Scale bars in the elemental maps are 400 nm. |
Fig. 5 FT-IR spectra of the HEH-1# as-synthesized products under EG-K, EG-K/N, and H2O–K/N conditions. |
A relatively high solution pH makes it possible to precipitate all metal cations completely with targeted composition (on average) from solution. Meanwhile, the polyol solvent plays a key role in inhibiting elemental segregation possibly through forming polyol-based metal hydroxides.50,51 It has been reported that reduction and hydrolysis are the main chemical reactions during the polyol process.51 The existence of OH− and acetate ions can facilitate the complexation of polyol with metal cations through deprotonating the hydroxyl group.51 A higher pH results in increased OH− and H2O contents in the as-synthesized HEHs. More OH− and H2O in the compound may facilitate stabilizing the layers composed of edge-sharing metal–oxygen octahedra and thus benefit the formation of 2D layered sheets through the polyol process.54
The phase evolution and morphological change of the annealed particles are investigated by XRD and TEM-EDS. Fig. 7 shows the XRD patterns of the HEH-1# EG-K/N as-synthesized products before and after annealing at 200–600 °C. The low-angle diffraction peak disappears after annealing at 200 °C, confirming the decomposition of the metal hydroxide. For HEH-2#–7# EG-K/N as-synthesized products, low-angle reflections disappear upon annealing, indicating the collapse of M–OH lamellar structures, as shown in Fig. S8a.† The crystallinity of the nanoparticles increases with increasing annealing temperature. It demonstrates a well crystallized spinel-type structure after annealing at 600 °C, as compared in Fig. 7 and S8b.† Rietveld fitting of 600 °C-annealed HEH-1# EG-K/N nanoparticles (Fig. S9†) gives a unit cell lattice parameter of 8.3128 Å and volume of 574.457 Å3. The crystallite size calculated by using the Scherrer equation is 12.6 nm for HEH-1# EG-K/N.58
Fig. 7 XRD patterns of HEH-1# EG-K/N as-synthesized products before and after annealing in air at 200–600 °C for 2 h, respectively. |
Fig. 8 shows the TEM images and elemental maps for HEH-1# EG-K/N nanoparticles after annealed in air at 200 and 600 °C. The 2D layered morphology is well maintained after 200 °C annealing. Thin, flat and large-area layers of HEMs composed of nanoparticles can be clearly observed in Fig. 8a and b. The HRTEM image in Fig. 8c shows that the crystallites are about 5 nm in size and have high crystallinity. They possess the same lattice plane that can be indexed to the (311) plane of the spinel structure. Fig. 8d clearly shows that all metal elements are uniformly distributed without any obvious segregation. The specific surface area of the 200 °C-annealed nanoparticles is 242.0 m2 g−1. These results manifest that 2D layered high-entropy spinel oxides can be successfully synthesized through our polyol process, followed by post-annealing treatment at 200 °C, a low synthesis temperature compared with other methods.33,36,59–61 The flat nanosheets become a particulate when annealed at 600 °C. The size of the nanoparticles increases (Fig. 8e–g), and the specific surface area decreases to 49.5 m2 g−1. In Fig. 8g, clear lattice fringes can be observed, which can also be indexed to the (311) plane of the spinel structure. All five metal elements are well distributed on the nanoparticles. For HEH-1# EG-K and H2O–K/N as-synthesized products, their phase evolution upon heating is similar to the EG-K/N sample. Increasing the annealing temperature only leads to the increased crystallinity of the nanoparticles (Fig. S10†). After annealing at 600 °C, their XRD patterns can be indexed to a single-phase spinel structure of high crystallinity. However, they should be indeed mixtures of spinel particles with various compositions, as evidenced by the TEM-EDS analyses in Fig. S11.†
Fig. 8 TEM images with different magnifications and elemental maps for the HEH-1# EG-K/N sample after annealed at (a)–(d) 200 °C and (e)–(h) 600 °C. |
The FC curve of the as-synthesized sample shows a steady decrease to near zero in the whole measuring temperature range, while a weak anomaly is indicated where the ZFC curve shows the maximum, indicating that the magnetic transition is around 5 K. The corresponding M–H curve shows a ferromagnetic hysteresis loop at 2 K. However, it is not magnetically saturated even at 50 × 103 Oe, suggesting that ferromagnetism and paramagnetism coexist. On the other hand, at high temperatures, the temperature dependence of magnetization (Fig. 9a) and the M–H curve (Fig. 10) indicate that it is paramagnetic. In fact, above about 100 K, the magnetization obeyed the Curie–Weiss law with negative Weiss temperature (θ) (Fig. 9b). The negative Weiss temperature indicates antiferromagnetic interaction between magnetic ions. The magnetic properties observed in the as-synthesized samples at low temperatures may be due to the presence of multiple magnetic ions coupled antiferromagnetically with neighbouring magnetic ones having different magnetic moments.62
The ZFC and FC curves for the 600 °C-annealed nanoparticles are more magnetized than those for the as-synthesized sample. They are also similar to those of ZFC-FC observed for superparamagnetic nanoparticles;63 the ZFC curve exhibits a peak, indicating a transition from a magnetically blocked state (low temperature) to a superparamagnetic state (high temperature). As can be seen from Fig. 9a, the blocking temperature (TB) for 600 °C-annealed nanoparticles is estimated to be 245 K. For particles annealed at 1000 °C, the ZFC curve indicates a fast increase at 2–75 K and increase with a relatively slow rate above 75 K, and eventually overlaps the FC curve at temperature above ∼360 K. Assuming the overlapping point as the maximum of the ZFC curve, the TB would be ∼360 K. The crystallite size of the 1000 °C-annealed particles is 40.5 nm calculated by using the Scherrer equation, which is larger than that of the 600 °C-annealed one. The blocking temperature increases as the volume of the particle increases, agreeing well with previous publications.63 The shapes of the ZFC curves could be explained by a magnetic anisotropy of magnetic domains and the inter-site exchange interactions among different metal cations at different sublattices.64,65 It requires further study to get a better understanding of the composition-metal species-magnetic property relationship, which is beyond the scope of this work.
The M–H curves recorded at 300 K represent superparamagnetic behavior, agreeing well with the ZFC-FC plots, with inappreciable coercivity and residual magnetization, as compared in Table 3 and Fig. S12.† The saturation magnetization (Ms) increases as the annealing temperature (i.e., the crystallite size) increases. It becomes comparable to that of the control sample prepared by solid state reactions when annealed at 1000 °C. The M–H plots recorded at 2 K demonstrate hysteresis loops for both the as-synthesized and annealed nanoparticles. It is interesting to note that the coercivity decreases dramatically from 2339.0 Oe to 284.3 Oe, while the residual magnetization and saturation magnetization decreases from 17.5 emu g−1 to 9.5 emu g−1 and increases from 30.8 emu g−1 to 44.5 emu g−1 for 600 °C- and 1000 °C-annealed nanoparticles, respectively. The coercivity for the 1000 °C-annealed sample at 300 K and 2 K does not change so much, while that of the other samples increases dramatically with the reduction of temperature. The ZFC-FC and M–H plots confirm a very strong dependence of the magnetic properties on the annealing temperature, which is correlated with the increased crystallite size and crystallinity of the samples.
Sample | Temperature (K) | Coercivity Hc (Oe) | Residual magnetization Mr (emu g−1) | Saturation magnetization Ms (emu g−1) |
---|---|---|---|---|
a Note: Control sample was prepared by solid state reactions using Co, Cr, Fe, Mn, and Ni binary oxides as raw materials at 1000 °C for 1 h. | ||||
As syn | 2 | 961.0 | 3.7 | 17.6 |
200 °C ann | 2 | 2532.5 | 1.4 | 4.9 |
600 °C ann | 2 | 2339.0 | 17.5 | 30.8 |
1000 °C ann | 2 | 284.3 | 9.5 | 44.5 |
As syn | 300 | 0.0 | 0.0 | 0.2 |
200 °C ann | 300 | 38.5 | 0.0 | 0.5 |
600 °C ann | 300 | 15.4 | 0.3 | 7.7 |
1000 °C ann | 300 | 104.5 | 0.6 | 16.0 |
Controla | 2 | 565.0 | 16.0 | 43.1 |
Controla | 300 | 70.4 | 3.3 | 16.3 |
Footnote |
† Electronic supplementary information (ESI) available. See https://doi.org/10.1039/d1na00871d |
This journal is © The Royal Society of Chemistry 2022 |