Next Article in Journal
Effect of Short-Chain Polymer Binders on the Mechanical and Electrochemical Performance of Silicon Anodes
Previous Article in Journal
A Comprehensive Review of Equalization Techniques for Reconfigured Second-Life Battery Systems
Previous Article in Special Issue
Review of Quasi-Solid Aqueous Zinc Batteries: A Bibliometric Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Stable Manganese-Based High-Entropy Prussian Blue for Enhanced Sodium-Ion Storage

1
Joint International Laboratory on Environmental and Energy Frontier Materials, School of Environmental and Chemical Engineering, Shanghai University, Shanghai 200444, China
2
Shenzhen Institutes of Advanced Technology, Chinese Academy of Sciences, Shenzhen 518055, China
3
Centre for Clean Energy Technology, University of Technology Sydney, Broadway, Sydney, NSW 2007, Australia
*
Authors to whom correspondence should be addressed.
Batteries 2025, 11(9), 328; https://doi.org/10.3390/batteries11090328
Submission received: 21 July 2025 / Revised: 23 August 2025 / Accepted: 25 August 2025 / Published: 1 September 2025
(This article belongs to the Special Issue Battery Interface: Analysis & Design)

Abstract

Prussian blue (PB) and its analogs (PBAs) are considered ideal cathode materials for sodium-ion batteries (SIBs) due to the following merits, including high redox potential, simple synthesis methods, and excellent structural stability. Herein, we synthesized a high-entropy PB cathode material, Na1.20Mn0.38Fe0.15Ni0.14Co0.15Cu0.16[Fe(CN)6]0.820.18·0.38H2O (HE-HCF), through a facile co-precipitation method. The five transition metals in HE-HCF have similar atomic sizes and electronegativity, collectively occupying the high-spin Fe-HS sites. The manganese-based system design reduces the preparation cost, and the high-entropy doping approach further decreases the content of crystalline water in the structure. Benefiting from the synergistic effects of the multiple component elements, HE-HCF demonstrates a capacity retention rate of 72.7% at 0.1 A g−1. Moreover, it even maintains 85.3% of its initial capacity after 1000 cycles at 1 A g−1. Electrochemical impedance spectroscopy (EIS) and galvanostatic intermittent titration technique (GITT) analyses further confirm that HE-HCF exhibits low charge transfer resistance and a small reaction activation energy.

1. Introduction

Sodium-ion batteries (SIBs) represent a next-generation energy storage technology, widely positioned as the prime substitution to lithium-ion batteries (LIBs) because of their superior crustal abundance, inherent safety advantages, and remarkable cost-effectiveness [1,2,3,4]. Among the critical factors governing the electrochemical performance of SIBs, the structural design and performance optimization of cathode materials are of pivotal importance [5]. PB and PBAs are deemed as ideal cathode materials for the commercialization of SIBs, owing to their simple preparation process, low raw material costs, and large crystal frameworks that facilitate the transport of sodium ions. The introduction of nano-scale gap interfaces has brought practical flexibility to the field of ions that facilitate the transport of sodium ions [6,7]. Among the various PBAs, manganese-based hexacyanoferrates (NaxMn[Fe(CN)6·nH2O, abbreviated as Mn-HCF) stand out as the most promising candidates on account of their abundant reserves, low extraction costs, and high average voltage, making them highly suitable for the commercial application of SIBs [8,9]. However, PBAs undergo irreversible phase transitions and lattice distortions during cycling, which detrimentally affects the material’s electrochemical cycling stability. Additionally, the unavoidable formation of coordination water during the synthesis process can react with the electrolyte, leading to harmful side reactions [8]. These factors significantly hinder the practical commercialization of SIBs [10,11,12].
PBAs follow the general formula AxM[Fe(CN)6]y·zH2O, in which A denotes alkaline or multivalent cations (e.g., Na+, K+, and Mg2+), while M stands for transition metals (e.g., Fe, Mn, Cu, and Ni), y indicates the occupancy of [Fe(CN)6] units (0 < y < 1), with 0 < x < 2 and z < 4 [13,14,15]. Different elements have different influences on the kinetic reactions during the charging and discharging processes [14]. Vacancies at [Fe (CN)6] sites are commonly present, affecting both structural and electrochemical properties. The structure comprises alternating low-spin Fe-C6 and high-spin M-N6 octahedra cyanide-bridged into an open 3D double-perovskite architecture with abundant redox-active sites. Intrinsic [Fe (CN)6] vacancies further tune both structural stability and electrochemical behavior [16,17].
In Mn-hexacyanoferrates (Mn-HCFs), pronounced Mn3+ induced Jahn-Teller distortion within Mn-N6 octahedra triggers severe phase transition that degrades cycling reversibility and accelerates capacity decay [18]. Improving the practical viability of PBAs and tackling their inherent issues requires optimization through methods like metal substitution, surface modification, defect tuning, and forming PBA composites [19]. Raising the configurational entropy through equimolar incorporation of five or more transition-metal cations (>1.5R) has recently emerged as an effective remedy: the resulting high-entropy lattice suppresses distortion and irreversible phase change while enhancing cycling stability [13,20,21,22]. The high-entropy strategy shows great potential in improving the general electrochemical performance of materials.
The benefits are exemplified by Brezesinski’s quasi-zero-strain PBA cathode containing five transition metals [23]. These elements share the same nitrogen coordination sites in equimolar ratios, increasing the system’s configurational entropy to over 1.5R. This innovation achieved a quasi-zero-strain working mechanism. Chou et al. fabricated a disordered cubic high-entropy framework structure containing five transition metals, exhibiting an ultra-long lifespan exceeding 50,000 cycles alongside outstanding rate performance [24]. Additionally, Ting et al. discovered that the constituent elements in a high-entropy PBA (HE-PBA) cathode comprising five metals play distinct roles [25]. For instance, Cu doping effectively enhances electrical conductivity [26], while Zn doping regulates vacancy formation. Their co-doping even produces a synergistic effect, adjusting particle size distribution and specific surface area. These works highlight the synergistic “mixing effect” intrinsic to high-entropy chemistry.
Building on this concept, we synthesized a manganese-based (high-entropy hexacyanoferrate) HE-HCF using a facile one-step co-precipitation method, introducing Fe, Co, Ni and Cu—transition metals with similar ionic radii and electronegativities—into the Mn lattice. The single-phase, high-entropy structure offers multiple active redox sites, dampens the Jahn–Teller effect and lowers production cost. The material delivers 72.7% capacity retention at 0.1 A g−1 and 85.3% following 1000 cycles at 1 A g−1, together with eminent rate properties, underscoring the efficacy of the high-entropy strategy for Mn-based PBA cathodes.

2. Materials and Methods

2.1. Material Synthesis

2.1.1. Synthesis of FeMn-PBA

The sodium-rich monoclinic FeMn-HCF was synthesized via a facile co-precipitation maneuver. Solution A was fabricated through dispersing 4 mmol of MnCl2·4H2O in 150 mL of deionized water and magnetically stirring for 6 h at ambient temperature. Solution B contained 6 mmol Na4Fe(CN)6·10H2O, 20 mmol C6H5Na3O7, and 20 mmol NaCl, dissolved in 150 mL deionized water. Afterwards, solution A was slowly added to solution B with continuous stirring (this step is carried out under a nitrogen atmosphere). Then the resultant mixture was allowed to age under ambient conditions for 24 h followed by triple washing with deionized water and ethanol. Finally, the precipitate was vacuum-dried at 120 °C for 12 h to yield the FeMn-HCF sample.

2.1.2. Synthesis of HE-PBA

HE-HCF (40% Mn) was prepared via an ordinary ambient-temperature co-precipitation strategy. 10 mmol of Na4Fe(CN)6·10H2O, 20 mmol of C6H5Na3O7, and 20 mmol of NaCl were dissolved in 300 mL of deionized water (denoted as solution A). 2 mmol of MnCl2·4H2O and 3 mmol of other precursors (NiCl2·6H2O, FeCl2·4H2O, CuCl2·2H2O, CoCl2·6H2O, each 0.75 mmol) were dissolved in 300 mL of deionized water (denoted as solution B). Solution A was slowly introduced into solution B, followed by continuous magnetic stirring for 6 h (this step is carried out under a nitrogen atmosphere). The mixture aged quiescently for 24 h followed by centrifugation and triple washing with deionized water and ethanol. The precipitate was vacuum-dried at 60 °C for 12 h, mechanically pulverized, and finally annealed at 120 °C in vacuum oven for 12 h to obtain the HE-HCF sample.

2.2. Materials Characterizations

Structural and morphological characterization employed a ZEISS Gemini-SEM 300 (15 kV, Oberkochen, Baden-Württemberg, Germany) for scanning electron microscopy while high-resolution transmission electron microscopy (HR-TEM) analyses were performed using a JEOL-2100F (200 kV, Akiruno City, Tokyo Metropolis, Japan) instrument for atomic-resolution characterization. Chemical states of constituent elements were probed via X-ray photoelectron spectroscopy (XPS) using an ESCALAB 250Xi system (Thermo Fisher Scientific, Waltham, MA, USA). Crystal structures were characterized through X-ray diffraction (XRD, Rigaku D/MAX 2200 V, Akiruno City, Tokyo Metropolis, Japan) with Cu Ka radiation, collected over a 2θ range of 10–70° at a scanning speed of 8° min−1.

2.3. Electrochemical Measurements

Electrochemical evaluation of Mn-HCF and HE-HCF cathodes was employed by assembling CR2032 coin cells inside an argon-filled glove box. Electrodes contained 70:20:10 wt% active material/Ketjen Black/polyvinylidene fluoride binder (PVDF) binder, using N-methyl pyrrolidone (NMP) as dispersing solvent [27]. The active material mass loadings spanned 1.0–2.0 mg cm−2. A glass fiber GF/D separator was used, and pure sodium metal was employed as the counter electrode. Each coin cell contained 75 μL of electrolyte, which was a solution of 1 mol L−1 NaClO4 in a 1:1 mixture of ethylene carbonate (EC) and diethylene carbonate (DEC), with 5% fluoroethylene carbonate (FEC) as an additive. The electrochemical tests were carried out on the NEWARE battery testing system.

3. Results

Employing an ordinary co-precipitation approach, the Mn-PBA framework was engineered with four equimolar transition metals (Fe, Ni, Co, Cu) to synthesize HE-HCF. The particle size of HE-HCF has reached the nanometer level, presenting an irregular cubic shape, with a particle size of approximately 100 nm (Figure 1a,b). Transmission electron microscopy (TEM) images of FeMn-HCF at a magnification of 1 μm reveal well-defined stacked cubic structures. Corresponding selected area electron diffraction (SAED) patterns unambiguously confirm its monocrystalline character (Figure S2). High-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) (Figure S3c) in conjunction with elemental mapping (Figure S3d–h) demonstrates the uniform distribution of component elements throughout the material. For HE-HCF, TEM images recorded at a scale of 200 nm show nanoscale particles with irregular morphology, consistent with a single-crystalline structure (Figure 1c). The SAED pattern (Figure 1d) exhibits distinct diffraction rings corresponding to the (200), (220), (400), (420), (440), and (620) planes. Elemental mapping further verifies the homogeneous spatial distribution of constituent metallic species across the HE-HCF particles (Figure 1f–m).
X-ray diffraction (XRD) analysis reveals that FeMn-HCF exhibits typical monoclinic phase characteristics, with noticeable splitting of characteristic peaks at (220), (420), (440), and (620) planes (Figure 2a). In contrast, HE-HCF shows no such peak splitting. The peak splitting in FeMn-HCF is attributed to its high sodium content [28]. The single-phase structure of HE-HCF indicates that all five transition metal elements are well-incorporated without forming other impurity phases. The Rietveld refined XRD patterns of both samples reveal the complete phase purity and crystal structure integrity of FeMn-HCF and HE-PBA (Figure 2c,f). The crystallographic parameters from the refined XRD patterns are summarized in Tables S1 and S2. It can be seen that FeMn-HCF and HE-HCF belong to the monoclinic phases with P21/n space group. It can be observed that the unit cell parameters and lattice volume of HE-HCF are slightly smaller than those of FeMn-HCF due to the introduction of other cationic ions, which can well explain the redshift of its Raman peaks. Additionally, from Table S2, it can be seen that in HE-HCF, Fe occupies the 2a (Fe1, high-spin state, Fe-HS) and 2d (Fe2, low-spin state, Fe-LS) sites. The other four transition metal elements occupy the same site as Fe1, indicating that they are codoped. In PB and PBAs, the adsorbed water and interstitial water are relatively easy to remove, while the crystalline water tends to form strong coordination chemical bonds within the crystals and is difficult to remove by common drying and heat treatment methods [29]. Thermogravimetric analysis (TGA) shows that HE-HCF experiences a mass loss of 9.53% before 180 °C, attributed to desorption of adsorbed and interstitial water. Between 180 °C and 210 °C, there is a 4.77% mass loss, associated with the elimination of coordinated water occupying vacancy sites [12]. For FeMn-HCF, a mass loss of 1.93% occurs up to 180 °C, and between 180 °C and 270 °C, a 10.57% mass loss is observed, corresponding to the removal of coordinated water. These TGA results suggest that high-entropy doping with transition metals can reduce the coordinated water content in the crystal structure. The subsequent mass loss beyond 270 °C for FeMn-HCF and beyond 210 °C for HE-HCF is due to the collapse and decomposition of their crystal structures under high-temperature conditions (Figure 2b). Raman spectroscopy indicates that the peak at 2117 cm−1 in HE-HCF is attributed to the stretching vibration of Fe2+-CN-M2+(M2+ = Fe2+, Mn2+, Ni2+, Co2+, Cu2+) bonds, while the feature at 2143 cm−1 assigns the stretching vibration of Fe2+-CN-M3+ bonds (Figure 2d). Compared to FeMn-HCF, both peaks in HE-HCF exhibit a red shift, suggesting that the stretching vibration of Fe2+-CN-M2+ is stronger than that of Fe2+-CN-Mn2+. This implies that the interatomic distances between M2+ ions and coordinated C and N atoms are shorter than those between Mn2+ ions and coordination atoms, leading to slight distortions in the crystal structure due to changes in lattice parameters. In addition to TGA, Fourier-transform infrared spectroscopy (FTIR) further confirms the crystalline water incorporation. Within the wavenumber range of 500 to 4000 cm−1, distinct stretching vibration peaks are located at 595 cm−1 (Fe-CN), 1619 cm−1 (H-O-H), 2080 cm−1 (-C≡N), and 3461 cm−1 (O-H) (Figure 2e). Notably, compared to FeMn-HCF [5,30], the peaks associated with O-H and H-O-H functional groups in HE-HCF are significantly reduced, nearly disappearing. This indicates a substantial reduction in coordinated water within the sample material, which is beneficial for structural stability during electrochemical cycling. The decrease in the Fe-CN peak intensity is attributed to the substitution of high-spin Fe2+ (Fe-HS) by other doping atoms. Both TGA and FTIR results demonstrate that HE-HCF contains less coordinated water than FeMn-HCF, suggesting that high-entropy doping is an effective method for reducing coordinated water content in the crystal. Typically, high-spin Fe-HS sites may be occupied by vacancies or coordinated water; however, during the doping process with five transition metal elements, a competitive relationship forms, wherein these metals vie with coordinated water for Fe-HS sites, thereby reducing the coordinated water content within the crystal.
Inductively Coupled Plasma Optical Emission Spectroscopy (ICP-OES) analysis (Table S3) indicates that the chemical formula of HE-HCF is Na1.20Mn0.38Fe0.15Ni0.14Co0.15Cu0.16[Fe(CN)6]0.820.18 (□denotes vacancies within the lattice), with elemental molar ratios closely aligning with the intended stoichiometric ratios. Additionally, the chemical formula for FeMn-HCF is determined to be Na1.74Mn [Fe(CN)6]0.820.18·1.77 H2O, indicating a sodium-rich monoclinic phase, in good accordance with the XRD characterization data. Utilizing the configurational entropy formula:
S c o n f = i = 1 n x i l n x i
where S c o n f represents the configurational entropy, and xi denotes the molar fraction. The calculated ∆Sconf value for HE-HCF is 1.5R, confirming that the synthesized HE-HCF is a high-entropy material.
To further investigate the elemental composition and oxidation states of FeMn-HCF and HE-HCF, X-ray photoelectron spectroscopy (XPS) analyses were carried out. The survey spectrum of FeMn-HCF confirms the presence of Na, Fe, and Mn in the sample (Figure 3a). In contrast, the survey spectrum of HE-HCF demonstrates that, in addition to these three elements, Ni, Co, and Cu are also successfully incorporated (Figure 3b). The Na 1s and Fe 2p peaks in FeMn-HCF (Figure S4a,b) show no significant differences compared to those in HE-HCF (Figure 3c,d). Judging from the peak position, it is Fe2+ [31]. However, the satellite peak at 646.28 eV observed in the Mn 2p spectrum of HE-HCF is not present in the FeMn-HCF, indicating that Mn exists solely in the Mn3+ oxidation state in FeMn-HCF (Figure S4c) [30]. For HE-HCF, the Na 1s spectrum exhibits a solitary peak at 1071.9 eV (Figure 3c). Meanwhile, the Fe 2p peaks appear at 708.6 eV (Fe 2p3/2) and 721.5 eV (Fe 2p1/2), with a spin–orbit splitting energy of 12.9 eV, indicating the presence of Fe2+ (Figure 3d) [32,33,34]. In the Mn 2p spectrum, peaks at 642.2 eV and 654.1 eV are assigned to Mn2+ 2p3/2 and Mn2+ 2p1/2, respectively (Figure 3e) [35]. The Ni 2p spectrum exhibits characteristic Ni2+ signatures including a spin–orbit doublet at 856.5 eV (2p3/2) and 874.0 eV (2p1/2), accompanied by satellite features at higher binding energies (Figure 3f) [5]. Additionally, small peaks at 858.6 eV and 876.2 eV are attributed to Ni3+. Similarly, the Co 2p and Cu 2p spectra display satellite peaks at 786.2 eV and 802.4 eV, as well as at 944.3 eV and 964.7 eV, respectively, confirming the presence of Co2+ [36] and Cu2+ (Figure 3g,h) [33,34]. Moreover, the Cu 2p spectrum displays characteristic Cu+ signatures at 933.0 eV (2p3/2) and 952.9 eV (2p1/2) [37,38,39]. This observation indicates the occurrence of an internal redox reaction during the synthesis process, leading to the reduction of Cu2+ to Cu+, accompanied by partial oxidation of Ni2+ to Ni3+. The existence of Cu+ can decrease the distortion, contributing to ameliorated structural integrity and suppressed irreversible phase transformation [39]. Furthermore, the average chemical state of Fe coordinated with carbon (Fe1) and metal ions coordinated with nitrogen (Fe2, Mn, Ni, Co, and Cu) is determined to be +2.

4. Discussion

Figure 4a,b presents the initial three cycles of charge–discharge curves of FeMn-HCF at 0.1 A g−1 and 1.0 A g−1, respectively. Two distinct electron-transfer processes in FeMn-HCF are resolved by CV measurement, closely related to high-spin Mn2+/Mn3+ redox at 3.72/3.52 V and low-spin Fe2+/Fe3+ couple at 3.50/3.08 V (Figure S5) [40]. The CV curves of HE-HCF show a single redox couple at 3.53 V/3.14 V (Figure 4c). This is due to the high-entropy doping, which makes the five transition metal elements in Fe-HS sites have a similar coordination environment, resulting in the normalization of the charge–discharge voltage platform. This cooperative redox behavior manifests as a singular redox couple in the CV curve, which is also called the “cocktail” effect [41]. Furthermore, the overlapping CV curves of HE-HCF across three cycles demonstrate better reversibility and a more stable structure. Figure 4d,e shows the charge–discharge profiles of FeMn-HCF at varying current densities. In contrast, the specific capacity of HE-HCF remained relatively stable over the first three cycles. Moreover, HE-PBA exhibits a much higher specific capacity than FeMn-HCF when the current density is larger than 0.2 A g−1, and the specific capacity loss upon returning to the initial current is relatively small (Figure 4f). At 0.1 A g−1, FeMn-HCF delivers a discharge capacity of 122.4 m Ah g−1 at the initial cycle versus 97.6 m Ah g−1 for HE-HCF. However, after 13 cycles, the capacity of FeMn-HCF begins to decline sharply, dropping to 47.9 m Ah g−1 after cycling 300 times (39.1% retention). In striking contrast, HE-HCF maintains 71.0 m Ah g−1 after 300 cycles, retaining 72.7% of its initial discharge capacity. HE-PBA exhibits a much higher specific capacity than FeMn-HCF under various discharge current densities, and the specific capacity loss upon returning to the initial current is relatively small. At 1 A g−1, HE-HCF exhibits an initial discharge capacity of 72.3 m Ah g−1, which decreases to 61.7 m Ah g−1 after 1000 cycles, maintaining an impressive capacity retention of 85.3%, significantly outperforming FeMn-HCF in cycling stability (Figure 4h). The pronounced degradation of FeMn-HCF is primarily attributed to the Jahn-Teller effect inherent in manganese-based PBAs, where Mn2+ dissolution from the lattice structure during cycling leads to severe lattice collapse and structural degradation. However, through the high-entropy doping strategy of transition metals, the electrochemical cycling stability of HE-HCF is significantly enhanced, effectively suppressing Mn2+ dissolution, mitigating lattice distortion, and reducing internal stress.
EIS measurements were conducted at different temperatures to investigate the impedance characteristics of FeMn-HCF and HE-HCF in depth. By fitting and analyzing the EIS data, the activation energy ( E a ) values of FeMn-HCF and HE-HCF were further derived. The Nyquist plots feature a high-frequency semicircular arc and low-frequency linear region, representing the charge transfer resistance (Rct) of Na+ diffusion and Warburg impedance, respectively. EIS measurements were performed at different temperatures for both FeMn-HCF and HE-HCF. As shown in Figure 5a, the Rct values of HE-HCF at these temperatures are 147.8, 93.7, 75.0, 62.3, and 46.6 Ω. In contrast, the Rct values of FeMn-HCF (Figure S6) are significantly higher across all temperatures, indicating a greater internal charge transfer resistance, which is detrimental to Na+ transport. Conversely, HE-HCF exhibits enhanced Na+ diffusion capability and lower lattice diffusion resistance. Furthermore, both FeMn-HCF and HE-HCF show a decreasing trend in Rct values as the temperature increases, suggesting that elevated temperatures facilitate Na+ transport and enhance the Na+ diffusion rate. The Arrhenius equation quantitatively governs the exponential correlation between thermal energy and reaction kinetics, serving as the fundamental framework for analyzing temperature-dependent behavior. The Arrhenius equation can be expressed as:
k = A · e E a R T
in which k denotes the reaction rate constant, while A stands for the pre-exponential factor, manifesting the reaction occurrences per second under ideal conditions. Additionally, E a is the activation energy, which denotes the minimum energy required to overcome reaction barriers, R represents the ideal gas constant, and T refers to the absolute temperature (in Kelvin). This equation shows that the reaction rate constant k increases as temperature rises. By fitting the linear relationship between the ionic conductivity ln(100σT) (S K cm−1) and 1000/T (K−1), the calculated slope values for FeMn-HCF and HE-HCF are −6.555 and −2.475, respectively (Figure 5b). Consequently, based on the Arrhenius equation, the activation energy E a of FeMn-HCF is determined to be 0.56 eV, whereas that of HE-HCF is only 0.21 eV (Figure 5c). This significant reduction in E a due to the high-entropy doping effect indicates a substantial decrease in the reaction energy barrier, which further facilitates the kinetic processes.
To gain deeper insights into the Na+ diffusion kinetics during electrochemical cycling, Galvanostatic Intermittent Titration Technique (GITT) measurements were performed on FeMn-HCF and HE-HCF. The Na+ diffusion coefficient (DNa+) is derived from transient voltage relaxation profiles using the following equation [42]:
D N a + = 4   π τ m B V M M B A 2 Δ E S Δ E τ 2
where mB, MB, and VM represent the mass, molar mass, and molar volume of the active material, respectively. Since VM remains constant throughout the electrochemical process, and A denotes the electrode area, the voltage-time profiles of FeMn-HCF and HE-HCF are used to extract the required parameters (Figure S7a,b). Based on the calculations, the DNa+ value for FeMn-HCF and HE-HCF are 3.421 × 10−11 cm2 s−1 and 2.047 × 10−10 cm2 s−1, respectively (Figure S7c,d). The DNa+ value of HE-HCF is approximately six times higher than that of FeMn-HCF, aligning with its superior rate properties and lower impedance values. This enhancement is primarily attributed to the smaller particle size of HE-HCF, which effectively shortens the Na+ diffusion path and accelerates ionic transport within the electrode.
To elucidate the charge storage mechanism of the HE-HCF electrode, cyclic voltammetry (CV) tests were conducted at varying scan rates (0.1, 0.2, 0.4, 0.6, and 0.8 mV s−1). As shown in Figure 5d, the CV curves at different scan rates display well-defined and consistent redox peaks, although a gradual increase in polarization is observed with increasing scan rate. The contribution from capacitive processes was quantified by analyzing the linear relationship between the peak current (i) and the square root of the scan rate (ν), based on the equation.
i = a v b
The b-value, reflecting the dominant charge storage mechanism, was derived through fitting the linear relationship between ln(v) and ln(i). CV analysis of HE-HCF (Figure 5e) yielded Tafel slopes of 0.667 and 0.644 for oxidation (O1) and reduction (R1) peaks, respectively. These b-values suggest that the charge storage behavior is mainly diffusion-controlled. To further quantify the capacitive contribution at each scan rate, the following equation was applied [43]:
i   ( V )   = k 1 ν + k 2 ν 1 / 2
capacitive (k1ν) and diffusion-controlled (k2ν1/2) contributions were quantified [44]. The constants k1 and k2 were determined via analysis of the linear relationship between ν1/2 and i/ν1/2 (Figure S8). The capacitive contribution ratio, calculated at different scan rates, indicates that the proportion of capacitive contribution increases progressively as the scan rate rises. Specifically, the capacitive contribution in the HE-HCF electrode accounts for 19% at 0.1 mV s−1. When the scan rate increases to 0.8 mV s−1, this proportion rises to 40% (Figure 5f).

5. Conclusions

In summary, a manganese-based high-entropy Prussian blue cathode material, Na1.20Mn0.38Fe0.15Ni0.14Co0.15Cu0.16[Fe(CN)6]0.820.18·0.38H2O (HE-HCF), was successfully synthesized via a co-precipitation method under a N2 atmosphere. The HE-HCF exhibits a nanoscale irregular particle morphology, with five transition metal elements uniformly distributed within the structure. The choice of a manganese-based system effectively reduces the material’s synthesis cost. The high-entropy doping strategy decreases the crystallized water content in the material and suppresses the Jahn–Teller effect associated with Mn2+ dissolution. Benefiting from the incorporation of five transition metals, HE-HCF demonstrates prominent crystal stability and rate properties in contrast to FeMn-HCF. Specifically, even after cycling 1000 times at 1 A g−1, the HE-HCF can still maintain 85.3% of its initial capacity, whereas FeMn-HCF retains only 72.7%. These results collectively demonstrate that HE-HCF possesses superior electrochemical performance and reaction kinetics.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/batteries11090328/s1. Figure S1. Diagram of the preparation process of HE-HCF; Figures S2 and S3: SEM, TEM image, selected electron diffraction pattern, HAADF diagram. EDS mapping images of FeMn-HCF; Figure S4: High-resolution spectra of FeMn-HCF; Figures S5 and S6: CV curves and EIS spectra of FeMn-HCF; Figure S7: GITT profiles and the corresponding Na+ diffusion coefficients of FeMn-HCF and HE-HCF; Figure S8. Plot of ν1/2i/ν1/2 used for calculating k1 and k2; Table S1: Crystallographic parameters of FeMn-HCF by Rietveld method; Table S2: Crystallographic parameters of HE-HCF by Rietveld method; Table S3: ICP-OES results of FeMn-HCF and HE-HCF.

Author Contributions

Conceptualization, H.L., H.G., Y.Z., Y.X. and C.L.; methodology, Y.X. and C.L.; validation, C.L., Y.X., D.Z. and X.Y.; formal analysis, C.L., Y.X., D.Z. and X.Y.; investigation, C.L., Y.X., D.Z. and X.Y.; resources, H.G. and H.L.; data curation, Y.X. and C.L.; writing—original draft preparation, C.L., Y.X., D.Z. and X.Y.; writing—review and editing, H.G. and C.L.; visualization, H.G., C.L., Y.X., D.Z. and X.Y.; supervision, H.G., J.X. and H.L.; project administration, H.G. and H.L.; funding acquisition, H.G. and H.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Science and Technology Commission of Shanghai Municipality (22010500400), “Joint International Laboratory on Environmental and Energy Frontier Materials”, “Innovation Research Team of High-Level Local Universities in Shanghai” in Shanghai University, and “Shanghai Pujiang Program” (23PJ1402800).

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Acknowledgments

The authors are grateful to all colleagues who contributed to this work, and their contributions are reflected in this article.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
PBPrussian blue
PBAsPrussian blue Analogs
SIBsSodium-ion batteries
EISElectrochemical impedance spectroscopy
GITTGalvanostatic intermittent titration technique
XRDX-ray diffraction
TGAThermogravimetric analysis
CVConstant voltage

References

  1. Yang, C.; Xin, S.; Mai, L.; You, Y. Materials Design for High-Safety Sodium-Ion Battery. Adv. Energy Mater. 2021, 11, 2000974. [Google Scholar] [CrossRef]
  2. Qi, S.; Wu, D.; Dong, Y.; Liao, J.; Foster, C.W.; O’Dwyer, C.; Feng, Y.; Liu, C.; Ma, J. Cobalt-Based Electrode Materials for odium-Ion Batteries. Chem. Eng. J. 2019, 370, 185–207. [Google Scholar] [CrossRef]
  3. Jin, T.; Ji, X.; Wang, P.; Zhu, K.; Zhang, J.; Cao, L.; Chen, L.; Cui, C.; Deng, T.; Liu, S. High-Energy Aqueous Sodium-Ion Batteries. Angew. Chem. Int. Ed. 2021, 60, 11943–11948. [Google Scholar] [CrossRef]
  4. Jiang, M.; Hou, Z.; Ren, L.; Zhang, Y.; Wang, J.-G. Prussian Blue and Its Analogues for Aqueous Energy Storage: From Fundamentals to Advanced Devices. Energy Storage Mater. 2022, 50, 618–640. [Google Scholar] [CrossRef]
  5. Huang, Y.; Zhang, X.; Ji, L.; Wang, L.; Xu, B.B.; Shahzad, M.W.; Tang, Y.; Zhu, Y.; Yan, M.; Sun, G. Boosting the Sodium Storage Performance of Prussian Blue Analogs by Single-Crystal and High-Entropy Approach. Energy Storage Mater. 2023, 58, 1–8. [Google Scholar] [CrossRef]
  6. Maier, J. Thermodynamic aspects and morphology of nano-structured ion conductors: Aspects of nano-ionics part I. Solid State Ion. 2002, 154–155, 291–301. [Google Scholar] [CrossRef]
  7. Zhang, G.; Gao, H.; Zhang, D.; Xiao, J.; Sun, L.; Li, J.; Li, C.; Sun, Y.; Yuan, X.; Huang, P. Transformative catalytic carbon conversion enabling superior graphitization and nanopore engineering in hard carbon anodes for sodium-ion batteries. Carbon Energy 2025, 7, e713. [Google Scholar] [CrossRef]
  8. Joos, M.; Kang, X.; Merkle, R.; Maier, J. Water uptake of solids and its impact on ion transport. Nat. Mater. 2025, 24, 821–834. [Google Scholar] [CrossRef]
  9. Wu, F.; Wu, S.; Ye, X.; Ren, Y.; Wei, P. Research Progress of High-Entropy Cathode Materials for Sodium-Ion Batteries. Chin. Chem. Lett. 2025, 36, 109851. [Google Scholar] [CrossRef]
  10. Wu, R.; Ren, B.; Wang, X.; Lin, J.; Li, X.; Zheng, J.; Yang, H.Y.; Shang, Y. Vacancy Remediation in Prussian Blue Analogs for High-Performance Sodium and Potassium Ion Batteries. Adv. Funct. Mater. 2024, 35, 2418018. [Google Scholar] [CrossRef]
  11. Dai, J.; Tan, S.; Wang, L.; Ling, F.; Duan, F.; Ma, M.; Shao, Y.; Rui, X.; Yao, Y.; Hu, E. High-Voltage Potassium Hexacyanoferrate Cathode via High-Entropy and Potassium Incorporation for Stable Sodium-Ion Batteries. ACS Nano 2023, 17, 20949–20961. [Google Scholar] [CrossRef] [PubMed]
  12. Ma, Y.; Ma, Y.; Dreyer, S.L.; Wang, Q.; Wang, K.; Goonetilleke, D.; Omar, A.; Mikhailova, D.; Hahn, H.; Breitung, B. High-Entropy Metal–Organic Frameworks for Highly Reversible Sodium Storage. Adv. Mater. 2021, 33, 2101342. [Google Scholar] [CrossRef]
  13. Wang, Y.; Zhou, S.; Li, N.; Han, J.; Zhang, S.; Zhuang, Z.; Sun, Z.; Wang, X.; Wu, X.; Chen, Z.; et al. Long-Durable Potassium Ion Batteries Enabled by Medium-Entropy Lattice Engineering on Prussian Blue Analogues Cathodes. Adv. Energy Mater. 2024, 15, 2405007. [Google Scholar] [CrossRef]
  14. Maier, J. Utility of Simple Rate Equations for Solid State Reactions. Z. Anorg. Allg. Chem. 2005, 631, 433. [Google Scholar] [CrossRef]
  15. Chen, M.; Li, X.; Yan, Y.; Yang, Y.; Xu, Q.; Liu, H.; Xia, Y. Polypyrrole-Coated K2Mn[Fe(CN)6] Stabilizing Its Interfaces and Inhibiting Irreversible Phase Transition during the Zinc Storage Process in Aqueous Batteries. ACS Appl. Mater. Interfaces J. 2022, 14, 1092–1101. [Google Scholar] [CrossRef]
  16. Deng, L.; Qu, J.; Niu, X.; Liu, J.; Zhang, J.; Hong, Y.; Feng, M.; Wang, J.; Hu, M.; Zeng, L. Defect-Free Potassium Manganese Hexacyanoferrate Cathode Material for High-Performance Potassium-Ion Batteries. Nat. Commun. 2021, 12, 2167. [Google Scholar] [CrossRef]
  17. Peng, J.; Zhang, W.; Liu, Q.; Wang, J.; Chou, S.; Liu, H.; Dou, S. Prussian Blue Analogues for Sodium-Ion Batteries: Past, Present, and Future. Adv. Mater. 2022, 34, 2108384. [Google Scholar] [CrossRef]
  18. Qian, Z.; Luo, R.-J.; Ma, C.; Du, C.-Y.; Zeng, J.; Xu, X.; Mei, Z.; Zhou, Z.-T.; Fu, Z.-W.; Zhou, Y.-N. Six Element High-Entropy Prussian Blue Analogue Cathode Enabling High Cycle Stability for Sodium-Ion Batteries. Chem. Eng. J. 2024, 500, 156767. [Google Scholar] [CrossRef]
  19. Hong, P.; Xu, C.; Yan, C.; Dong, Y.; Zhao, H.; Lei, Y. Prussian blue and its analogues for commercializing fast-charging sodium/potassium-ion batteries. ACS Energy Lett. 2025, 10, 750–778. [Google Scholar] [CrossRef]
  20. Du, M.; Geng, P.; Pei, C.; Jiang, X.; Shan, Y.; Hu, W.; Ni, L.; Pang, H. High-Entropy Prussian Blue Analogues and Their Oxide Family as Sulfur Hosts for Lithium-Sulfur Batteries. Angew. Chemie Int. Ed. 2022, 61, e202209350. [Google Scholar] [CrossRef]
  21. Jiang, W.; Wang, T.; Chen, H.; Suo, X.; Liang, J.; Zhu, W.; Li, H.; Dai, S. Room Temperature Synthesis of High-Entropy Prussian Blue Analogues. Nano Energy 2021, 79, 105464. [Google Scholar] [CrossRef]
  22. Lei, Y.; Wang, S.; Zhao, L.; Li, C.; Wang, G.; Qiu, J. Entropy Engineering Constrain Phase Transitions Enable Ultralong-life Prussian Blue Analogs Cathodes. Adv. Sci. 2024, 11, 2402340. [Google Scholar] [CrossRef]
  23. Chou, C.; Tuan, H. Whispers of Entropic Distortion Elevating Voltage and Electrochemical Depth in Potassium Hexacyanoferrate Cathodes. Adv. Funct. Mater. 2024, 35, 2418680. [Google Scholar] [CrossRef]
  24. Peng, J.; Zhang, B.; Hua, W.; Liang, Y.; Zhang, W.; Du, Y.; Peleckis, G.; Indris, S.; Gu, Q.; Cheng, Z. A Disordered Rubik’s Cube-Inspired Framework for Sodium-Ion Batteries with Ultralong Cycle Lifespan. Angew. Chem. Int. Ed. 2023, 62, e202215865. [Google Scholar] [CrossRef]
  25. Nguyen, V.T.; Sari, F.N.I.; Saputro, B.W.; Ting, J.-M. Structural and Defect Modulations of Co-Precipitation Synthesized High-Entropy Prussian Blue Analogue Nanocubes via Cu/Zn Co-Doping for Enhanced Electrochemical Performance. J. Mater. Chem. A 2023, 11, 19483–19495. [Google Scholar] [CrossRef]
  26. Xiao, J.; Gao, H.; Xiao, Y.; Wang, S.; Gong, C.; Huang, Z.; Sun, B.; Dong, C.-L.; Guo, X.; Liu, H. A hydro-stable and phase-transition-free P2-type cathode with superior cycling stability for high-voltage sodium-ion batteries. Chem. Eng. J. 2025, 506, 160010. [Google Scholar] [CrossRef]
  27. Lu, X.; Zhong, H.-Y.; Huang, P.-W.; Wu, Y.-N.; Xu, S.-Y.; Tan, X.-Y.; Wu, X.-H. Constructing FeSe2 nanorods supported on ketjen black with superior cyclability for potassium-ion batteries. J. Mater. Chem. A 2024, 12, 19995–20005. [Google Scholar] [CrossRef]
  28. Hurlbutt, K.; Giustino, F.; Volonakis, G.; Pasta, M. Origin of the High Specific Capacity in Sodium Manganese Hexacyanomanganate. Chem. Mater. J. 2022, 34, 4336–4343. [Google Scholar] [CrossRef]
  29. Ge, L.; Song, Y.; Niu, P.; Li, B.; Zhou, L.; Feng, W.; Ma, C.; Li, X.; Kong, D.; Yan, Z. Elaborating the Crystal Water of Prussian Blue for Outstanding Performance of Sodium Ion Batteries. ACS Nano 2024, 18, 3542–3552. [Google Scholar] [CrossRef] [PubMed]
  30. Peng, J.; Gao, Y.; Zhang, H.; Liu, Z.; Zhang, W.; Li, L.; Qiao, Y.; Yang, W.; Wang, J.; Dou, S. Ball Milling Solid-State Synthesis of Highly Crystalline Prussian Blue Analogue Na2−x MnFe(CN)6 Cathodes for All-Climate Sodium-Ion Batteries. Angew. Chem. Int. Ed. 2022, 61, e202205867. [Google Scholar] [CrossRef] [PubMed]
  31. Ma, Y.; Hu, Y.; Pramudya, Y.; Diemant, T.; Wang, Q.; Goonetilleke, D.; Tang, Y.; Zhou, B.; Hahn, H.; Wenzel, W. Resolving the Role of Configurational Entropy in Improving Cycling Performance of Multicomponent Hexacyanoferrate Cathodes for Sodium-Ion Batteries. Adv. Funct. Mater. 2022, 32, 2202372. [Google Scholar] [CrossRef]
  32. Huang, Y.; Xie, M.; Zhang, J.; Wang, Z.; Jiang, Y.; Xiao, G.; Li, S.; Li, L.; Wu, F.; Chen, R. A Novel Border-Rich Prussian Blue Synthetized by Inhibitor Control as Cathode for Sodium Ion Batteries. Nano Energy 2017, 39, 273–283. [Google Scholar] [CrossRef]
  33. Tang, Y.; Li, W.; Feng, P.; Zhou, M.; Wang, K.; Wang, Y.; Zaghib, K.; Jiang, K. High-Performance Manganese Hexacyanoferrate with Cubic Structure as Superior Cathode Material for Sodium-Ion Batteries. Adv. Funct. Mater. 2020, 30, 1908754. [Google Scholar] [CrossRef]
  34. Bie, X.; Kubota, K.; Hosaka, T.; Chihara, K.; Komaba, S. Synthesis and Electrochemical Properties of Na-Rich Prussian Blue Analogues Containing Mn, Fe, Co, and Fe for Na-Ion Batteries. J. Power Sources 2018, 378, 322–330. [Google Scholar] [CrossRef]
  35. Liu, Y.; He, D.; Han, R.; Wei, G.; Qiao, Y. Nanostructured Potassium and Sodium Ion Incorporated Prussian Blue Frameworks as Cathode Materials for Sodium-Ion Batteries. Commun. Chem. 2017, 53, 5569–5572. [Google Scholar] [CrossRef] [PubMed]
  36. Kang, Y.; Wang, S.; Hui, K.S.; Li, H.-F.; Liang, F.; Wu, X.-L.; Zhang, Q.; Zhou, W.; Chen, L.; Chen, F. [Fe(CN)6] Vacancy-Boosting Oxygen Evolution Activity of Co-Based Prussian Blue Analogues for Hybrid Sodium-Air Battery. Mater. Today Energy 2021, 20, 100572. [Google Scholar] [CrossRef]
  37. Zhang, X.; Xia, M.; Liu, T.; Peng, N.; Yu, H.; Zheng, R.; Zhang, L.; Shui, M.; Shu, J. Copper Hexacyanoferrate as Ultra-High Rate Host for Aqueous Ammonium Ion Storage. Chem. Eng. J. 2021, 421, 127767. [Google Scholar] [CrossRef]
  38. Wang, Q.; Sarkar, A.; Wang, D.; Velasco, L.; Azmi, R.; Bhattacharya, S.S.; Bergfeldt, T.; Düvel, A.; Heitjans, P.; Brezesinski, T. Multi-Anionic and -Cationic Compounds: New High Entropy Materials for Advanced Li-Ion Batteries. Energy Environ. Sci. 2019, 12, 2433–2442. [Google Scholar] [CrossRef]
  39. Zhou, Z.; Dong, Y.; Ma, Y.; Zhang, H.; Meng, F.; Ma, Y.; Wu, Y. Innovative high-entropy strategy extending traditional metal substitution for optimizing Prussian blue analogues in rechargeable batteries. Susmat 2025, 5, e265. [Google Scholar] [CrossRef]
  40. Gu, Y.; Lu, Y.; Dai, P.; Cao, X.; Zhou, Y.; Tang, Y.; Fang, Z.; Wu, P. Self-assembled high-entropy Prussian blue analogue nanosheets enabling efficient sodium storage. J. Colloid Interface Sci. 2025, 677, 307–313. [Google Scholar] [CrossRef]
  41. Kumar, K.; Kundu, R. Empowering Energy Storage Technology: Recent Breakthroughs and Advancement in Sodium-Ion Batteries. ACS Appl. Energy Mater. J. 2024, 7, 3523–3539. [Google Scholar] [CrossRef]
  42. Zhou, P.; Liu, X.; Weng, J.; Wang, L.; Wu, X.; Miao, Z.; Zhao, J.; Zhou, J.; Zhuo, S. Synthesis, Structure, and Electrochemical Properties of O′3-Type Monoclinic NaNi0.8Co0.15Al0.05O2 Cathode Materials for Sodium-Ion Batteries. J. Mater. Chem. A 2019, 7, 657–663. [Google Scholar] [CrossRef]
  43. Liu, Y.; Shen, Q.; Zhao, X.; Zhang, J.; Liu, X.; Wang, T.; Zhang, N.; Jiao, L.; Chen, J.; Fan, L. Hierarchical Engineering of Porous P2-Na2/3Ni1/3Mn2/3O2 Nanofibers Assembled by Nanoparticles Enables Superior Sodium-Ion Storage Cathodes. Adv. Funct. Mater. 2020, 30, 1907837. [Google Scholar] [CrossRef]
  44. Zhang, Z.; Liu, Y.; Liu, Z.; Li, H.; Huang, Y.; Liu, W.; Ruan, D.; Cai, X.; Yu, X. Dual-Strategy of Cu-Doping and O3 Biphasic Structure Enables Fe/Mn-Based Layered Oxide for High-Performance Sodium-Ion Batteries Cathode. J. Power Sources 2023, 567, 232930. [Google Scholar] [CrossRef]
Figure 1. (a,b) SEM images; (c) TEM of HE-HCF; (d) selected electron diffraction pattern of HE-HCF; (e) HAADF diagram of HE-HCF; (fm) the corresponding elemental mappings of HE-HCF.
Figure 1. (a,b) SEM images; (c) TEM of HE-HCF; (d) selected electron diffraction pattern of HE-HCF; (e) HAADF diagram of HE-HCF; (fm) the corresponding elemental mappings of HE-HCF.
Batteries 11 00328 g001
Figure 2. (a) XRD patterns of FeMn-HCF and HE-HCF; (b) TGA patterns of FeMn-HCF and HE-HCF; (c) Rietveld refined XRD pattern of FeMn-HCF; (d) Raman patterns of FeMn-HCF and HE-HCF; (e) FTIR patterns of FeMn-HCF and HE-HCF; (f) Rietveld refined XRD pattern of HE-HCF.
Figure 2. (a) XRD patterns of FeMn-HCF and HE-HCF; (b) TGA patterns of FeMn-HCF and HE-HCF; (c) Rietveld refined XRD pattern of FeMn-HCF; (d) Raman patterns of FeMn-HCF and HE-HCF; (e) FTIR patterns of FeMn-HCF and HE-HCF; (f) Rietveld refined XRD pattern of HE-HCF.
Batteries 11 00328 g002
Figure 3. (a) Full XPS spectrum of FeMn-HCF; (b) Full XPS spectrum of HE-HCF; high-resolution spectra of HE-HCF: (c) Na 1s; (d) Fe 2p; (e) Mn 2p; (f) Ni 2p; (g) Co 2p; (h) Cu 2p.
Figure 3. (a) Full XPS spectrum of FeMn-HCF; (b) Full XPS spectrum of HE-HCF; high-resolution spectra of HE-HCF: (c) Na 1s; (d) Fe 2p; (e) Mn 2p; (f) Ni 2p; (g) Co 2p; (h) Cu 2p.
Batteries 11 00328 g003
Figure 4. (a,b) Charge and discharge curves of HE-HCF in the initial three cycles at 0.1 A g−1 and 1.0 A g−1; (c) CV curves of HE-HCF at a scan rate of 0.1 mV s−1. (d,e) Charge and discharge curves of FeMn-HCF in the initial three cycles at 0.1 A g−1 and 1.0 A g−1; (f) Rate capability of FeMn-HCF and HE-HCF at different current densities; (g) Cyclic performance of FeMn-HCF and HE-HCF at a current density of 0.1 A g−1; (h) Cyclic performance test diagram of FeMn-HCF and HE-HCF at a current density of 1.0 A g−1.
Figure 4. (a,b) Charge and discharge curves of HE-HCF in the initial three cycles at 0.1 A g−1 and 1.0 A g−1; (c) CV curves of HE-HCF at a scan rate of 0.1 mV s−1. (d,e) Charge and discharge curves of FeMn-HCF in the initial three cycles at 0.1 A g−1 and 1.0 A g−1; (f) Rate capability of FeMn-HCF and HE-HCF at different current densities; (g) Cyclic performance of FeMn-HCF and HE-HCF at a current density of 0.1 A g−1; (h) Cyclic performance test diagram of FeMn-HCF and HE-HCF at a current density of 1.0 A g−1.
Batteries 11 00328 g004
Figure 5. (a) EIS spectra of HE-HCF at different temperatures of 25, 35, 45, 55 and 65 °C; (b) Linear relationships of 1000T−1(K−1)-ln100σT(S K cm−1) of FeMn-HCF and HE-HCF. (c) Calculated activation energy of FeMn-HCF and HE-HCF; (d) CV curves of HE-HCF at different scan rates ranging from 0.1 to 0.8 mV s−1; (e) Plot of ln(v) versus ln(i); (f) Contribution rates of capacitance and diffusion-controlled behavior of HE-HCF at different scan rates.
Figure 5. (a) EIS spectra of HE-HCF at different temperatures of 25, 35, 45, 55 and 65 °C; (b) Linear relationships of 1000T−1(K−1)-ln100σT(S K cm−1) of FeMn-HCF and HE-HCF. (c) Calculated activation energy of FeMn-HCF and HE-HCF; (d) CV curves of HE-HCF at different scan rates ranging from 0.1 to 0.8 mV s−1; (e) Plot of ln(v) versus ln(i); (f) Contribution rates of capacitance and diffusion-controlled behavior of HE-HCF at different scan rates.
Batteries 11 00328 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Li, C.; Xiao, Y.; Zhang, D.; Yuan, X.; Xiao, J.; Zhao, Y.; Gao, H.; Liu, H. Stable Manganese-Based High-Entropy Prussian Blue for Enhanced Sodium-Ion Storage. Batteries 2025, 11, 328. https://doi.org/10.3390/batteries11090328

AMA Style

Li C, Xiao Y, Zhang D, Yuan X, Xiao J, Zhao Y, Gao H, Liu H. Stable Manganese-Based High-Entropy Prussian Blue for Enhanced Sodium-Ion Storage. Batteries. 2025; 11(9):328. https://doi.org/10.3390/batteries11090328

Chicago/Turabian Style

Li, Congcong, Yang Xiao, Dingyi Zhang, Xinyao Yuan, Jun Xiao, Yufei Zhao, Hong Gao, and Hao Liu. 2025. "Stable Manganese-Based High-Entropy Prussian Blue for Enhanced Sodium-Ion Storage" Batteries 11, no. 9: 328. https://doi.org/10.3390/batteries11090328

APA Style

Li, C., Xiao, Y., Zhang, D., Yuan, X., Xiao, J., Zhao, Y., Gao, H., & Liu, H. (2025). Stable Manganese-Based High-Entropy Prussian Blue for Enhanced Sodium-Ion Storage. Batteries, 11(9), 328. https://doi.org/10.3390/batteries11090328

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop