Zhuozheng
Hong
abc,
Zhuang-Chun
Jian
ac,
Yan-Fang
Zhu
*ac,
Yan-Jiang
Li
ace,
Qi-Cong
Ling
ac,
Hanshen
Xin
*d,
Didi
Wang
b,
Chao
Wu
*b and
Yao
Xiao
*ac
aCollege of Chemistry and Materials Engineering, Wenzhou University, Wenzhou, Zhejiang 325035, P. R. China. E-mail: xiaoyao@wzu.edu.cn; yanfangzhu@wzu.edu.cn
bInstitute of Energy Materials Science (IEMS), University of Shanghai for Science and Technology, Shanghai, 200093, P. R. China. E-mail: chaowu@usst.edu.cn
cZhejiang Provincial Key Laboratory of Advanced Battery Materials and Technology, Wenzhou University Technology Innovation Institute for Carbon Neutralization, Wenzhou, Zhejiang 325035, China
dSchool of Microelectronics, Shanghai University, Shanghai, 201800, P. R. China. E-mail: xinhanshen@shu.edu.cn
eKey Laboratory of Spin Electron and Nanomaterials of Anhui Higher Education Institutes, Suzhou University, Suzhou 234000, China
First published on 8th September 2025
Sodium-ion batteries (SIBs) are promising alternatives to lithium-ion batteries (LIBs) owing to abundant resources and cost-effectiveness. However, cathode materials face persistent challenges in structural stability, ion kinetics, and cycle life. This review highlights the transformative potential of high-entropy (HE) strategies that leveraging multi-principal element synergies to address these limitations via entropy-driven mechanisms. By establishing thermodynamic criteria for high-entropy materials (HEMs), we elucidate the universal principles whereby configurational entropy mitigates lattice distortion, suppresses phase transitions, and enhances Na+ diffusion kinetics via multi-element interactions. HE design demonstrates unique advantages for layered oxides, Prussian blue analogues (PBAs) and polyanionic cathode systems: it alleviates Jahn–Teller distortion through dopant synergy to stabilize layered structures; optimises ion migration channels by tuning exposed crystal facets; suppresses irreversible phase changes and mechanical strain to enable reversible structural evolution; and enhances redox reversibility via multi-site charge compensation among transition metals. Furthermore, reasonable design principles for the HE strategy in cathode materials for SIBs were proposed, along with the future expansion of theoretical calculations and the application of the HE strategy in the future. At the same time, potential challenges that may occur during this process and the current viewpoints and methods for solving these problems were emphasized. Overall, this review provides valuable guidance for the further exploration of the HE strategy in the field of SIBs.
![]() | ||
Fig. 1 (a) Distribution of global lithium resources. Reprinted with permission from ref. 2. Copyright 2024 John Wiley and Sons. (b) Research hotspots in the HE cathode of SIBs. |
Elemental doping usually refers to the incorporation of a small amount of secondary elements into the main material, which is an effective way to endow SIB materials with good performance.43,44 However, this simple elemental doping method has certain limitations in terms of exploration and performance optimization. To significantly expand the available composition space and unlock new research directions, the HE approach has emerged as a research hotspot in cathode design for SIBs, as evidenced by the keyword clustering analysis in Fig. 1b. The HE approach combines multiple principal elements at equiatomic or near-equiatomic ratios. Unlike conventional doping, this strategy enhances compositional flexibility in compositional and structural design, resulting in a wider range of crystal and electronic structures. In addition, the incorporation of multiple elements can cause local disorder within the lattice, thereby increasing the system's disorder and enhancing its tolerance to structural evolution during electrochemical processes. At the same time, specific defects may be introduced, which optimize ion migration channels and indirectly facilitate electron migration by modulating the electronic structure.45
The growing interest in this strategy is further quantified by the sharp rise in both publications and citations of HE-based SIB cathodes over recent years.46 The cathode materials utilized in SIBs can be categorized into three primary groups: layered oxides,47–49 polyanionic,50–52 and Prussian blue analogues (PBAs).53–55 Layered oxide materials exhibit high energy density, excellent rate performance, and structural diversity. However, their practical applications are impeded by their susceptibility to atmospheric moisture and rapid capacity degradation. Polyanionic cathodes show remarkable stability and superior long-cycle performance. Nevertheless, they are subject to inherent limitations such as poor electrical conductivity and the need for material modification using carbon coating, doping, etc. These constraints pose significant challenges and will result in reduced energy density. PBA emerges as a promising sodium storage material owing to its distinctive metal–organic framework structure, which presents exceptional energy density. Nevertheless, for existing Prussian blue materials, removing crystalline water presents a substantial challenge, which dramatically determines their electrochemical performances. HEMs show remarkable advantages compared to traditional materials. The HE effect boosts the compatibility among elements, allowing for the formation of stable single-phase solid solutions at relatively lower temperatures.56 The differences in atomic radius and electronegativity of multiple elements will cause local fluctuation in lattice potential energy. This mainly inhibits the continuous diffusion of transition metal (TM) atoms through vacancies, thus reducing their ability for disordered migration.57,58 Such a particular effect not only helps maintain the structural stability but also restrains atomic migration phenomena that could potentially undermine the cycle stability during the battery cycling process. Simultaneously, the HE structure can more effectively address the changes in local interaction forces brought about by the Na+ intercalation/deintercalation and the redox reactions of TM elements. Due to this adaptability, HEMs can effectively prevent irreversible phase transitions during charging and discharging, significantly enhancing the structural stability and long-cycle performance. The rapid development of these different cathode materials highlights the transformative potential of HE engineering in the development of high-performance SIBs. Therefore, given the broad design prospects offered by HE strategies, the development of high-performance HEMs is undoubtedly within reach. Fig. 2 summarizes the landmark research achievements in chronological order, demonstrating the rapid development of HE strategies in the cathodes for SIB.59,60
![]() | ||
Fig. 2 Timeline of representative works on HEMs for sodium-ion batteries. All the illustrations are from the literature. Reprinted with permission from ref. 92. Copyright 2020 Wiley VCH. Reprinted with permission from ref. 59. Copyright 2021 Springer Nature. Reprinted with permission from ref. 107. Copyright 2022 John Wiley and Sons. Reprinted with permission from ref. 121. Copyright 2023 Elsevier. Reprinted with permission from ref. 115. Copyright 2024 Wiley VCH. Reprinted with permission from ref. 60. Copyright 2025 RSC. |
In 2020, Zhao et al. officially incorporated the HE concept into the design of SIBs. They synthesized the layered oxide cathode materials that contained as many as nine TM elements. Compared with traditional doped materials, these newly developed materials demonstrated outstanding long-term cycling stability. Specifically, after 500 cycles, their capacity retention could reach up to 83%. These materials also showed highly reversible O3–P3 phase transformation characteristics. Notably, more than 60% of the total capacity was stored in the layered O3 region. Additionally, the research team put forward the mechanism of the HE structure. These findings laid the foundation for the development of HE and further advanced the application of the HE strategy in SIBs.61 Although the HE strategy presents numerous possibilities for SIB materials, current research in this area remains in its nascent stage, and the fundamental concept of HEMs lacks precision.62–67 Initially, it was believed that when there were more than five metallic elements, the material was considered a HEM. Later, configurational entropy was introduced to determine whether it was a HEM, but this method could only be applied to some materials. This is the severe challenge faced by the implementing of the HE strategy in battery materials. Therefore, it is imperative to deeply investigate the working mechanism of multi-component structures in battery materials, provide a clear definition of HE, and elucidate various action mechanisms of different HEMs. In this review, we provide an overview of the most recent studies on the application of HEMs in various components of SIBs. We emphatically examine its influences on the phase structures and electrochemical characteristics of electrode materials, and propose to integrate the HE strategy into the overall design principles of battery materials as well as the future development direction.
To evaluate the different mixing entropies, the configurational entropy change (ΔSconfig) of the compositions are generally used according to the following formula:
Sconfig = −k![]() ![]() | (1) |
![]() | (2) |
Sconfig = R![]() ![]() | (3) |
ΔGmix = ΔHmix − TΔSmix | (4) |
In this context, k represents the Boltzmann constant (1.38 × 10−23 J K−1),81w denotes the degeneracy factor, R is the molar gas constant [8.314 J (mol−1 K−1)], N signifies the number of components, xi represents the molar fraction of cations, and xj indicates the molar fraction of anions.82 ΔHmix represents the mixing enthalpy, ΔSmix refers to the mixing entropy, and T denotes the absolute temperature. For single-phase HEMs where the ΔSconfig is greater than or equal to 1.5 R, the product of temperature and mixing entropy (TΔSmix) exceeds the ΔHmix. Therefore, the overall Gibbs free energy change (ΔGmix) becomes negative, achieving entropy-stabilized conditions (Fig. 3a and b). Materials exhibiting a ΔSconfig of 1.5 R or greater are classified as HE systems. Those with a ΔSconfig between 1 R and 1.5 R (including 1 R but excluding 1.5 R) are categorized as medium-entropy, and materials with a ΔSconfig less than 1 R are designated as low-entropy systems.83 Elements of varying sizes can induce the formation of disordered structures, leading to an atomic size mismatch effect. Consequently, certain materials, despite fulfilling the criteria for HEAs, exhibit inherent instability, with some displaying multiple phase states.84 It is important to note that the definition of HEAs and their derivatives remains an evolving area of research. Currently, a majority of research focuses on boosting ΔSconfig by incorporating multiple elements, aiming to explore the structural stability and enhanced properties of materials attributed to the HE effect.85,86
![]() | ||
Fig. 3 Thermodynamic analyses of the HE mixing considering both (a) entropy and (b) enthalpy, which are mainly determined by the composition of the HEMs. Reprinted with permission from ref. 83. Copyright 2024 Elsevier. |
The calculation of ΔSconfig is further extended to include oxides with multiple cation sites,87 such as perovskite oxides ABO3, where A and B represent cations occupying 12- and 6-coordination sites, respectively.88,89
![]() | (5) |
The rational design and preparation of HE cathode materials are pivotal for the advancement of practical applications. However, the intricate compositional complexity poses challenges in synthesis, such as reducing particle size and enhancing uniform mixing. In fact, it is the homogeneous blending of diverse elements that leads to substantial variations in chemical and physical properties, such as melting points and electrochemical behavior. Therefore, a precise approach to design and prepare HE cathodes can improve both the structure and battery performance.96,97
![]() | ||
Fig. 5 (a) The conventional O3-type Na-based layered oxides contain three different TM elements. (b) The proposed HEO cathodes have multiple TM elements (TM2 representes the redox elements and is marked in blue). Reprinted with permission from ref. 92. Copyright 2020 Wiley VCH. |
Layered oxides are regarded as one of the most promising cathode materials for SIBs due to their high energy density and straightforward synthesis process. However, when redox reactions occur at TM sites, the lattice stress caused by volume changes will weaken the bonding force between TMs and oxygen. Under high-voltage conditions, oxygen release is prone to occur, resulting in irreversible structural alterations during cycling. Additionally, TM dissolution and lattice cracking during cycling also affect the stability of the oxide cathode. Considering the benefits of HEAs, researchers have developed HEOs by modifying the ratio and type of TMs. These materials demonstrate excellent structural stability, enhanced fracture toughness, superior performance at both high and low temperatures, and outstanding electrochemical properties.
![]() | ||
Fig. 6 HE affects structural transformation. (a) XRD of Na0.7Mn0.4Ni0.3Cu0.1Fe0.1Ti0.1O1.95F0.1 at different temperatures during calcination. (b) The ratios of P2/O3 phases in NaMnNiCuFeTiOF sintered at various temperatures. (c) Schematic illustration of the structural change mechanisms of Na0.7Mn0.4Ni0.3Cu0.1Fe0.1Ti0.1O1.95F0.1. Reprinted with permission from ref. 106. Copyright 2023, Elsevier. (d) Schematic diagram of the charging and discharging behavior of Na2/3Li1/6Fe1/6Co1/6Ni1/6Mn1/3O2. (e) The bond length changes of TM–O (where TM represented Ni, Fe, Co, and Mn) varies during the charging process from 2.0 to 4.5 V. Reprinted with permission from ref. 107. Copyright 2022 John Wiley and Sons. Schematic diagram of the DOS for (f) LMM and (g) LMCNM. Reprinted with permission from ref. 108. Copyright 2025 American Chemical Society. |
Yao et al. successfully prepared Na2/3Li1/6Fe1/6Co1/6Ni1/6Mn1/3O2 HE cathodes, achieving an ultra-stable superlattice structure with Li/TM ordering (Fig. 6d).107 X-ray absorption spectroscopy (XAS) analysis showed that the redox of Ni2+/Ni3+/Ni4+ and Fe3+/Fe4+ provided a high coulombic efficiency during charge/discharge processes. The Mn4+ and Co3+ significantly improved the stability of the material during Na+ ion intercalation, suppressing Jahn–Teller distortion. Most importantly, the superlattice structure remained intact after long-term charge/discharge cycles. The reversible substitution between Li+ and Na+ and their interactions with different TMs dramatically enhanced the diffusion kinetics of Na+, stabilized the material's structure, and significantly reduced the irreversible phase transition. Extended X-ray absorption fine structure (EXAFS) analysis presented that in the TM–O coordination shell layer of the material, the atomic spacing at the K-edge of Ni and Fe was markedly shortened due to the oxidation of Ni and Fe, while the changes in Co and Mn were less pronounced. The K-edge transformation of Ni, Co, Fe, and Mn resulted from the reversible O3/P3 phase transition in the low-voltage region of the superlattice structure. The small deviation in the Fourier transform of the TM–O and TM–TM shell layers during the initial charge transfer more clearly indicated the stability of this superlattice structure with multiple element interactions (Fig. 6e). Generally speaking, by applying this HE strategy, it was possible to realize a rapid and reversible O3–P3 phase transition in the low-voltage range. Moreover, in the high-voltage region, the phase transition and oxygen redox reaction can be effectively suppressed. As a result, even under the conditions of high cut-off charging voltage and high cycling rate, high reversible capacity and outstanding stability could be achieved. The integration of HE and superlattice stabilization provided more opportunities for regulating the structure and phase transition, as well as the reversibility and activity of oxygen redox reactions via Li–TM interactions in layered cathode materials.
Due to the wide availability of P2-type Mn-based layered oxide, they are expected to be widely employed in large-scale electrochemical energy storage. However, the trivalent manganese in the MnO6 octahedron has a tendency to experience significant phase transformation and lattice oxygen loss at high working voltages. This issue severely impairs the capacity and cycling stability of these materials. Li et al. putted forward a ΔSconfig tuning approach to optimize the P2-type Na0.8Li0.17Mg0.18Mn0.66O2 (LMM) cathode. The synthesized cathode material, Na0.8Li0.17Ca0.025Mg0.12Ni0.05Mn0.66O2 (LMCNM), adhered to the standard P63/mmc crystal phase. Notably, this material demonstrated a capacity retention of 92% after 100 cycles at a 0.4C rate (where 1C is defined as 125 mA h g−1). Moreover, during charge–discharge cycles at a relatively high working voltage range of 2.0–4.3 V, it showed a minimal volume change of only 0.94%. In situ XRD, ex situ XPS, and computational analysis together suggested that there was no obvious Jahn–Teller distortion upon cycling for LMCNM (as shown in Fig. 6f and g). Instead, there was clear evidence of charge compensation from Mn3+ to Mn4+. Additionally, by co-doping with Ca and Ni, partial reversible anionic redox was achieved, which helped to balance the requirements of stability and high capacity.
Due to the Jahn–Teller effect of high-spin Mn3+ and Ni3+, Ni and/or Mn are dominant in layered oxides, which leads to structural distortion and electrode instability. Increasing ΔSconfig through HE strategies can significantly enhance the disorder of the system, uniformly disperse local stress, and effectively suppress the stretching of bond lengths of elements such as Mn3+, thereby improving the Jahn–Teller effect. This design not only improves the lifespan and rate performance of the cathode materials, but also provides a new idea for the development of high-energy-density and high-safety SIBs.
![]() | ||
Fig. 7 Some effects of HE on the crystal surface. (a) Schematic diagram of the morphological structure of the HEM compared with the normal material. (b) HAADF-STEM images of HEO424. (c) The HEM possesses a variety of cations that provide highly stable structures and diffusion channels, making more (010) active crystalline surfaces for ion transport. Reprinted with permission from ref. 109. Copyright 2022 American Chemical Society. (d) The six (010) active crystalline planes of HEM. (e) Variation of (010) active crystalline facet content with the change of ΔSconfig. Reprinted with permission from ref. 111. Copyright 2022 Springer Nature. (f) In situ XRD patterns and corresponding lattice parameter changes of TMO5. Reprinted with permission from ref. 112. Copyright 2025 Elsevier. |
Fu et al. used Na0.62Mn0.67Ni0.37O2 as a base for multi-element doping to design the HEM Na0.62Mn0.67Ni0.23Cu0.05Mg0.07Ti0.01O2 (CuMgTi-571). Besides structural stability, such HE material with abundant active crystal surfaces provided rapid Na+ transport.110,111 The CuMgTi-571 microparticles exhibited a hexagonal single-crystal structure. The active crystal planes displayed an open atomic arrangement structure (Fig. 7d), which was more favorable for Na+ transfer. As illustrated in Fig. 7e, by examining various materials with differing entropy values, it was observed that the proportion of active crystalline planes was improved as the increase of ΔSconfig. Furthermore, for materials with straightforward compositions and lower entropy, their surface energy is predominantly affected by temperature. In contrast, the surface energy of complex multicomponent systems with higher entropy is chiefly influenced by ΔSconfig. Therefore, the aforementioned P2 materials were mainly ruled by ΔSconfig, which modulated the growth rates of various crystallographic facets, resulting in high-performance HEO cathodes.
O3-type layered oxides are regarded as promising cathode materials for SIBs because of their high theoretical capacity. However, they frequently encounter issues such as structural instability and poor Na+ diffusion, which result in rapid capacity fading. Wang et al. proposed a HE strategy combined with a synergistic multi-metal effect to address these limitations. This was accomplished by improving the structural stability and reaction kinetics. A novel O3-type layered HE cathode material, Na0.9Fe0.258Co0.129Ni0.258Mn0.258Ti0.097O2 (TMO5),112 was synthesized via a simple solid-state method. The integration of experimental analysis and in situ/non-in situ characterizations verified that the HE metal ion mixture is advantageous for enhancing the reversibility of redox reactions and the O3–P3–O3 phase transition process (Fig. 7f), as well as increasing the Na+ diffusion rate. Due to its structural and compositional advantages, TMO5 presented a higher initial specific capacity of 159.6 mA h g−1. After 100 cycles at 2C, it still exibited a specific capacity of 110.1 mA h g−1, with a capacity retention rate as high as 85.6%. This study indicated that HE design is a promising approach to develop robust and high-performance O3-type layered oxide cathodes as well.
The HE strategy introduces multiple metal elements in nearly equimolar ratios to form layered oxide materials. The unique HE effect significantly promotes the Na+ transfer kinetics from multiple aspects. The random distribution of multiple principal elements causes significant local lattice distortions. Such distortions are not disordered but form a rich variety of local microenvironments with different energies within the framework of the average crystal structure. This complex local environment can soften the energy barriers along the Na+ diffusion paths. Some local sites may form depressions with lower energy or wider bottlenecks, providing more low-energy migration paths or lower migration activation energy for Na+, thereby accelerating the two-dimensional diffusion of Na+ between TM layers.
As reported by Hu et al., the TM sites in the HE cathode material NaNi0.12Cu0.12Mg0.12Fe0.15Co0.15Mn0.1Ti0.1Sn0.1Sb0.04O2 were occupied by nine different TM ions, resulting in an unexpectedly super-stable structure and exhibiting excellent cycling stability and high sodium storage capacity. Such HE cathode presented a capacity retention of 83% after 500 cycles and approximately 80% at 5C. Upon cycling, the material underwent a highly reversible O3 to P3 phase transformation, with more than 60% of its total capacity residing in the O3-type structure. The phase transition behavior of the samples was illustrated in Fig. 8a. In comparison with traditional O3-type layered oxide, the TM elements within the multi-component HEO cathode exhibited obvious interaction characteristics, which leaded to more distinct local interactions between Na in NaO2 and the elements in the TM layer during the charging and discharging process. Seven TM elements participated in charge compensation, resulting in intricate variations in their oxidation states. Concurrently, the sodium content adjusted accordingly due to Na+ intercalation and deintercalation processes, inducing structural localization and phase transitions (Fig. 8b).92 For conventional O3 phase cathodes, the elements involved in redox reactions are generally uniformly distributed in the lattice to minimize energy. As their ionic size and oxidation state change, the phase transitions of the overall material structure were easily visualized. However, for HEMs with multiple elements acting together, the HE composition leads to a stochastic distribution of the redox elements. Local fluctuations were accommodated by an increased diversity of local structural characteristics, effectively delaying the phase transition.
![]() | ||
Fig. 8 Suppression of phase transition through HE design. (a) In situ XRD patterns of NaNi0.12Cu0.12Mg0.12Fe0.15Co0.15Mn0.1Ti0.1Sn0.1Sb0.04O2 within 2.0–3.9 V at 0.1C. (b) Crystal structure evolution of the HE layered oxide cathode. Reprinted with permission from ref. 92. Copyright 2020 Wiley VCH. (c) In situ XRD patterns of NaMn0.2Fe0.2Co0.2Ni0.2Ti0.2O2 during charge/discharge cycling. (d) Variations in the content of O3 and P2 phases during the first charge/discharge cycle. (e) Graphical visualization of O3- and P3-type crystal structures. Reprinted with permission from ref. 115. Copyright 2022 Elsevier. |
The work reported by Molenda et al. developed a HE NaMn0.2Fe0.2Co0.2Ni0.2Ti0.2O2, which featured a dual-phase structure comprising O3(1) and O3(2) phases (Fig. 8c).115 As shown in Fig. 8d, the structural development exhibited a highly reversible phase transitions, characterized by the conversion process from O3 to P3 and returned to O3. With the assistance of HE doping, the P3 phase was able to be maintained in high voltage states. The precise phase composition of NaMn0.2Fe0.2Co0.2Ni0.2Ti0.2O2 during desodiation and sodiation was illustrated in Fig. 8e. The highly reversible O3–P3 phase transition mitigated the impact of oxygen vacancies, resulting in this ultrastable HEM cathode.
Wang et al. have created a new six-component HEO O3–Na(Fe0.2Co0.2Ni0.2Ti0.2Sn0.1Li0.1)O2.116 This material enabled highly reversible electrochemical reactions and phase transitions. After 100 cycles at a current density of 0.5C, such HEO cathode could still maintain 81% of its initial capacity, proving outstanding stability. Moreover, it also demonstrated outstanding rate performance. This phenomenon was ascribed to a relatively high Na+ diffusion coefficient, which was greater than 5.75 × 10−11 cm2 s−1, outperforming many previously reported O3-type cathodes. In addition, the HE cathode manifested excellent compatibility with the hard carbon anode, achieving a specific capacity of 90.4 mA h g−1 (with an energy density of 267 Wh kg−1). In situ XRD and XAS of iron revealed that charge compensation was facilitated by redox reactions involving Ni2+/Ni3+, Co3+/Co4+, and Fe3+/Fe2+ pairs. The disordered arrangement of multi-component TMs in the HEO restricted the ordering of charge and sodium vacancies. Consequently, the interlayer slipping and phase transitions were prevented.
The random distribution of the material results in the formation of a highly disordered TM layer, which in turn destroys the long-range order of the lattice. This, in turn, serves to reduce the driving force of phase transition and thereby inhibit the phase transition caused by Na+ deposition fundamentally. Concurrently, the disparity in atomic size and chemical properties, the local stress caused by the release of dispersed Na+ will also inhibit the structural instability. This design not only solves the problem of cycle stability of the layered oxide cathode but also provides a new way for developing high-voltage and high-capacity SIBs. Further studies are required to achieve the collaborative optimization of phase transition inhibition and ion dynamics by entropy regulating in the future.
![]() | ||
Fig. 9 Inhibition of grain surface strain by HE. (a) Schematic diagram of designing HE O3-type layer cathode material. (b) Crystal structure optimization diagram of HE cathode materials, showing the (001) crystal facet andELF. (c) The curves of Na+ ion formation energy and ΔSconfig during extraction and intercalation in HEMs. (d) Schematic diagram of reversible phase transition of O3-type HE cathode materials. Reprinted with permission from ref. 117. Copyright 2024 Wiley VCH. (e) SEM images of HEO and NMO before and after 200 cycles. (f) Calculation of hardness and Young's modulus for both materials. Reprinted with permission from ref. 110. Copyright 2023 Wiley VCH. (g) Structural model of O3 phase and schematic representation of the correlation between strain distribution and ionic displacement for NCFMT and NCFMS viewed in perpendicular and parallel to the TMO2 layers. Reprinted with permission from ref. 118. Copyright 2024 Springer Nature. |
Xiao et al. incorporated eight TM elements into the same lattice site of O3–NaNi0.5Mn0.5O2 (NMMO) and designed a HEO cathode material NaNi0.1Mn0.15Co0.2Cu0.1Fe0.1Li0.1Ti0.15Sn0.1O2, achieving ultra-strong mechanical properties at the microscale.110 The integrity of the crystal structure could be maintained even after long cycling. After 50, 100, and 200 cycles, the cross-sections of NMMO and HEO particles were shown in Fig. 9e. Obviously, cracks were gradually formed and there were a large number of dispersed cracks from the outside to the inside of the particles. Compared with NMMO, the surface morphology of HEO remained almost unchanged after cycling. The cracks penetrating the particles destroy the integrity of the cathode structure, leading to a series of side reactions and causing the TM ion migration and dissolution. Meanwhile, Young's modulus nanoindentation tests were conducted on HEO and NMMO to evaluate the structural damage during the cycling process. As seen, the Young's modulus of the pristine HEO was significantly greater than that of the NMMO (Fig. 9f).
Ding et al. synthesized O3-type HE NaNi0.3Cu0.1Fe0.2Mn0.3Ti0.1O2 (NCFMT) and NaNi0.3Cu0.1Fe0.2Mn0.3Sn0.1O2 (NCFMS) by substituting Ti4+ with Sn4+.118 The research found that NCFMT exhibited excellent coulombic efficiency and outstanding cycling stability, while NCFMS performed poorly because of structural instability. It has been revealed that the structural integrity of layered cathode materials is influenced by the compatibility of elements in the TM oxide layer. The planar strain in NCFMS was caused by metal ion displacement, leading to element separation and crack formation during cycling. In contrast, NCFMT demonstrated a stable structural framework due to the high mechanical-chemical compatibility among its constituent elements, which was conducive to the stable Na+ storage (Fig. 9g).
Upon cycling, the repeated intercalation and deintercalation of Na+ can cause significant lattice strain (volume change and lattice parameter fluctuation). This strain accumulation leads to microcracks in particles, an increased irreversibility of phase transitions, and thus seriously impairs cycle stability and rate performance. The HE strategy achieved fine control of lattice strain at the atomic and lattice scales. The core mechanism is reflected in the following aspects: (1) during deep desodiation process, the instability of oxygen ions (oxygen evolution) and the resulting TM migration and structural collapse cause significant lattice strain and capacity decay. The synergistic effect of multiple TMs can adjust the overall band structure and the position of the oxygen p-band center, enhancing the stability of oxygen. Strong lattice distortion and powerful inter-element interaction can anchor oxygen ions, inhibiting their migration and precipitation. (2) During the redox reactions triggered by Na+ intercalation and deintercalation, the variations in bond lengths corresponding to different elements (such as the Jahn–Teller effect of Mn3+ and the significant volume contraction of Ni2+/Ni4+) show differences or even counteract each other at the local scale. This cocktail effect produced by the synergy of multiple elements effectively buffers the drastic volume fluctuation in local regions, disperses and minimizes the overall lattice strain, as well as prominently reduces the local lattice stress concentration. (3) The synergistic interactions among multiple elements enhance the mechanical stability of the lattice by distributing local stress uniformly, thereby preventing microcrack formation and improving long-term cycling performance.
![]() | ||
Fig. 10 Ex situ XAS spectra of TMs in NaMn0.2Fe0.2Co0.2Ni0.2Ti0.2O2 in both TEY (a) and PFY (b) modes during charging and discharging. Reprinted with permission from ref. 115. Copyright 2022 Wiley VCH. (c) Normalized in situ XANES spectra and the corresponding voltage curve of the NCNFMT at Ti, Mn, Fe, Ni, and Cu K-edges in various states. (d) K-edge XANES spectra of Ti, Mn, Fe, Ni, and Cu and the layered structure scheme where M–O and M–M bond lengths were optimized in NCNFMT. Reprinted with permission from ref. 119. Copyright 2022 Wiley VCH. |
Wang et al. performed charge compensation analysis on O3–Na(Fe0.2Co0.2Ni0.2Ti0.2Sn0.1Li0.1)O2 HEMs.116 The XAS analysis showed that Ti and Sn were non-electrochemically active elements, and the redox potentials of Ti3+/Ti4+ and Sn4+/Sn5+ were lower than 2 V. The XANES spectra of Fe, Co, and Ni at various charge/discharge states were analyzed. During the initial charging process, the K-edge of Ni shifts toward higher energies, suggesting that Ni was oxidized to a higher oxidation state. Since the K-edge energy of Ni at 4.1 V was higher than that of Ni3+, it could be inferred that it was oxidized to approximately +3.5 at 4.1 V, providing a total capacity of about 72 mA h g−1. Additionally, Fe and Co also experienced shifts to higher energy and were oxidized to approximately +3.5, both contributing partial capacity. During the first discharge, Ni3.5+, Co3.5+, and Fe3.5+ were reduced. The K-edges of all three elements returned to lower absorption energies. After discharge to 3 V, the Ni K-edge continued to move toward lower energy levels, while Co and Fe showed no significant changes. This case indicated that the reduction of the Ni, Fe, and Co provided charge compensation in the voltage range of 4.1–3 V. From 3 V to 2 V, the Fe or Co was not involved in the redox reaction, while Ni was continuously reduced to +2 valence to provide charge compensation. They further analyzed the charge compensation mechanism of Fe by using Mössbauer spectroscopy. After charging to 4.1 V, an asymmetric double peak before charging revealed partial oxidation of Fe3+ to Fe4+. It was further shown that Fe3+ in the HEO material exhibited electrochemical activity based on Fe3+/Fe4+ redox couple. The spectra after complete discharge showed that there was no prominent high-spin Fe4+ in the HEO electrode. However, there were still noticeable differences from the original spectra, indicating that the structure of Fe was not completely restored after the first complete cycle, possibly due to a small amount of Fe4+ was not fully reduced to Fe3+ after the first cycle, and also that the HEO underwent slight structural distortion during the first cycle. In summary, the redox processes involved Ni2+/Ni3.5+, Co3+/Co3.5+, and some Fe3+/Fe4+ redox processes were crucial for charge balance, enabling this HEO's high reversible capacity.
Chen et al. designed a multi-element HE O3 cathode NaCu0.1Ni0.3Fe0.2Mn0.2Ti0.2O2. The charge compensation mechanism involving the Cu, Ni, Fe, Mn, and Ti K edges had be examined through XAS analysis as shown in Fig. 10c.119 The HE electrodes in different states were known for their oxidation states of doping precursors such as Ti(IV)O2, Mn(IV)O2, Fe(III)2O3, Ni(II)O, and Cu(II)O. During the initial charge/discharge cycle, no notable shift in absorption was observed in the Ti and Mn spectra. However, a subtle change in intensity was still detectable, suggesting that there was a slight alteration in the chemical environment surrounding Ti and Mn, but the valence state remained unchanged. This was due to the increase in structural disorder around Mn caused by lattice distortion of Na+ deinsertion and insertion, which also reflected the important role of entropy in stabilizing the structure. For the K-edge spectra of Fe and Cu, there was a small edge shift to higher energy during charging, but it reverted to the position of the original edge as discharge proceeded. Thus, the Fe3+/Fe4+ and Cu2+/Cu3+ were reversible redox couples, calculations indicated that the Fe and Cu occured redox reactions exclusively in the high-voltage range. For Ni's K-edge spectrum, the edge energy during the charged state exceeded that of Fe2O3. This behavior suggested the occurrence of Ni2+/Ni4+ redox coupling, where the average oxidation state of Ni attained +4 upon full charging (Fig. 10d). The above observations suggested that among the inactive elements in the stable structure, Fe and Cu contributed less to the capacity while Ni supplied more. The excellent performances of HEMs were attributed to the mutiple element symbiotic effect.
Liu et al. prepared the P2–Na0.67Mn0.6Cu0.08Ni0.09Fe0.18Ti0.05O2 (MCNFT) cathode material, which exhibited enhanced stability in a deeply desodiated state.113 This improvement facilitated internal Na+ migration, thereby enhancing both charge capacity and coulombic efficiency. The novel MCNFT demonstrated promising results by extracting 0.61 Na and exhibiting superior reversibility compared to NMO. External XAS and XPS analyses showed that the charge capacity improved from 92.8 mA h g−1 to 158.1 mA h g−1 upon charging to 4.5 V, while the coulombic efficiency increased from 57.7% to 98.2%. These enhancements could be attributed to the synergistic effects of cationic and anionic redox reactions. Furthermore, in situ XRD analysis proved that the crystal structure maintained its stability even when subjected to high sodium removal conditions. These experimental observations were supported by density functional theory (DFT) calculations.
Due to the disordered TM distribution and the regulation of the electronic structure, uniform charge compensation is achieved. The multi-element electronegativity difference and orbital hybridization can reduce the charge compensation barrier energy, thereby comprehensively improving the efficiency, reversibility, and capacity of charge compensation. Future studies can further explore the redox matching of element combinations and the microscopic mechanism of anion/cation co-compensation. Therefore, the research progress of HE layered oxides was discussed in detail and their main components and electrochemical properties were summarized in Table 1.
Cathode materials | Voltage range | Initial capacity (mA h g−1) | Cycle retention | Rate capacity (mA h g−1) | Ref. |
---|---|---|---|---|---|
NaNi0.12Cu0.12Mg0.12Fe0.15Co0.15Mn0.1Ti0.1Sn0.1Sb0.04O2 | 2–3.9 V | 110 | 83% (500 cycles 3C) | 86 (5C) | 92 |
Na0.7Mn0.4Ni0.3Cu0.1Fe0.1Ti0.1O1.95F0.1 | 2–4.3 V | 118.4 | 88.9% (200 cycles 2C) | 108.5 (2C) | 106 |
Na2/3Li1/6Fe1/6Co1/6Ni1/6Mn1/3O2 | 2–4.5 V | 171.2 | 63.7% (300 cycles 5C) | 78.2 (10C) | 107 |
Na0.8Li0.17Ca0.025Mg0.12Ni0.05Mn0.66O2 | 2–4.3 V | 163.6 | 80.43% (300 cycles 4C) | 81 (4C) | 108 |
NaNi0.25Mg0.05Cu0.1Fe0.2Mn0.2Ti0.1Sn0.1O2 | 2–4 V | 130.8 | 75% (500 cycles 1C) | 108 (5C) | 109 |
NaNi0.1Mn0.15Co0.2Cu0.1Fe0.1Li0.1Ti0.15Sn0.1O2 | 2–4.1 V | 115 | 82.7% (1000 cycles 1C) | 100 (8C) | 110 |
Na0.62Mn0.67Ni0.23Cu0.05Mg0.07Ti0.01O2 | 2–4.3 V | 148.2 | 87% (500 cycles 1C) | 59.3 (10C) | 111 |
Na0.9Fe0.258Co0.129Ni0.258Mn0.258Ti0.097O2 | 2–4.2 V | 159.6 | 85.6% (100 cycles 2C) | 110.1 (2C) | 112 |
NaCu0.1Ni0.3Fe0.2Mn0.2Ti0.2O2 | 2–3.9 V | 130 | 87.2% (100 cycles 0.1C) | 85 (5C) | 119 |
NaMn0.2Fe0.2Co0.2Ni0.2Ti0.2O2 | 1.5–4.2 V | 180 | 97% (100 cycles 0.1C) | 95 (2C) | 115 |
Na(Fe0.2Co0.2Ni0.2Ti0.2Sn0.1Li0.1)O2 | 2–4.1 V | 90.4 | 81% (100 cycles 0.5C) | 81 (2C) | 116 |
NaNi0.2Fe0.2Mn0.35Cu0.05Zn0.1Sn0.1O2 | 2–4 V | 128 | 87% (500 cycles 3C) | — | 117 |
NaNi0.3Cu0.1Fe0.2Mn0.3Ti0.1O2 | 2–4 V | 141.5 | 85% (500 cycles 1C) | 120 (1C) | 118 |
Ma et al. designed a high-entropy HE-PBA featuring outstanding performance and a strain-free structure by incorporating five cations—Fe, Mn, Ni, Cu, and Co—at the nitrogen-coordinated M site. This material exhibited a ΔSconfig value of 1.61 R as shown in Fig. 11a and b.95 They performed the first in situ probing of the gas production behavior of HE-PBA using differential electrochemical mass spectrometry (DEMS). Hydrogen gas evolution during the second cycle may originate from the reduction of crystalline water and the electrolyte at the anode, while the formation of CO2 was related to the carbonate electrolyte at the cathode. If in a fully reversible reaction, the gas production of the active material generally didn't change as the reaction proceeds. However, the gas production in the first cycle differed from that in the second cycle. This discrepancy was caused by the side-reaction of hydrogen hexacyanide, which was presented in most PBAs. By setting the voltage range between 2.5 and 4.1 V, the generation of gas can be effectively suppressed by the suppression of previously mentioned side reactions through HE, which was still an issue in conventional PBAs. To better investigate the structural transformation during Na deintercalation and intercalation, in situ XRD analysis was performed in the first two cycles (Fig. 11c and d).123,124 It could be found that the deintercalation process was fully reversible, and no phase transitions emerged throughout the reaction. Hence, the HE-PBA exhibited superior structural stability compared to traditional PBAs, providing additional evidence of the efficacy of the HE approach.
![]() | ||
Fig. 11 (a) Schematic diagram of the crystal structure of multi-element doping. (b) In situ DEMS test of the HEM during the second charge/discharge. (c) In situ XRD patterns during charging and discharging. Reprinted with permission from ref. 95. Copyright 2021 Wiley VCH. (d) The structures of PBAs with three different entropies were simulated by using DFT calculations. Reprinted with permission from ref. 69. Copyright 2022 Wiley VCH. |
Ma et al. synthesized NaxMn0.4Fe0.15Ni0.15Cu0.15Co0.15[Fe(CN)6] (HE-MnPBAs). To understand the effect of entropy increase on the structure, they performed DFT calculations for several materials by using VASP and found that the formation energy of HEM-HCF HEMs was always the lowest. With the increase of ΔSconfig, the formation energy decreased, leading to more stable materials. Compared with medium and low entropy samples, HEM-HCF exhibited a more stable structure during the reaction (Fig. 12a). The Mn K-edge analysis of HEM-HCF and Mn-HCF revealed that the half-path lengths of the MN, M
C, and M
M shells in both materials showed striking similarities. These findings indicated that the 4b (M) site (occupied by Fe, Mn, Co, Ni, and Cu) in each material shares an identical coordination environment with the Mn site, specifically –Fe(4a)–C
N–M(4b)–N
C–Fe(4a)–. Fe K-edge data for Mn-HCFs showed a comparable distribution of half-path lengths. Conversely, Fe-HCF presented a distinct first-shell layer distribution, characterized by discernible metal–ligand spacings, which highlighted the stability of its HE structure (Fig. 12b).69
![]() | ||
Fig. 12 (a) Comparison of XAS data for HEM and Mn-HCF. Comparison of XAS data for HEM, Mn-HCF, and Fe-HCF. Reprinted with permission from ref. 69. Copyright 2022 Wiley VCH. (b) Schematic diagram of the structural simulation of SC-HEPBA during the cycling process. (c) Schematic representation of the unit volume change of SC-HEPBA during charging and discharging. (d) The TM contents in the electrolyte after more than 100 cycles. Reprinted with permission from ref. 121. Copyright 2023 Elsevier. (e) Schematic diagram for the chemical etching process of HEPBA-Etched-0. Reprinted with permission from ref. 128. Copyright 2024 American Chemical Society. |
As reported by Huang et al., a HE monoclinic single-crystal Prussian blue analogue (SC-HEPBA) was prepared. The unique HE characteristic brought about rapid Na+ diffusion and effectively suppressed metal dissolution. Additionally, the micron-scale single-crystal morphology enhanced the electrochemical performance and minimized structural deterioration during charge–discharge cycles.125,126 As illustrated in Fig. 12c, the structural transformation of SC-HEPBA exhibited high reversibility during cycling. Fig. 12d further quantified the minimal volume change (3.3% during charge and 2.6% during discharge), which was attributed to the entropy-stabilized framework.121 It could be inferred that this reverse three-phase transformation was caused by the structure formed by using the single crystal strategy, and the relatively small volume change was attributed to the structure of HEPBAs. This structure greatly enhanced the Na+ transport rate and ensured cycle stability of the material. The entropy-mediated capabilities of structural decay and phase transition were in the following order: low-entropy manganese-based PBAs were less than medium-entropy manganese-based PBAs, which were less than HE manganese-based PBA.
PBAs are regarded as competitive cathode materials for SIBs because of their low cost and ease of synthesis. Nevertheless, the practical application of PBA is restricted by its intrinsically poor electronic conductivity and structural degradation. Wang et al. integrated the benefits of HE and the 3D confinement effect provided by carbon wrapping (CW). For the first time, they designed and synthesized a distinctive three-dimensional carbon-wrapped HE-PBA (HE-PBA@C).127 The HE strategy guaranteed the inherent stability of PBA and facilitated Na+ diffusion. Meanwhile, the CW technique had a dual function. It enhanced electronic conductivity through the 3D confinement effect and alleviated lattice distortion. Consequently, the HE-PBA@C cathode showed remarkable electrochemical performance. This included a high specific capacity (120.2 mA h g−1 at 10 mA g−1), excellent rate performance (73.0 mA h g−1 at 4 A g−1), outstanding capacity retention (80.1% after 4000 cycles at 2 A g−1), and unprecedented stability under ambient conditions. Furthermore, in full cells, the HE-PBA@C cathode exhibited excellent compatibility with hard carbon (95.4% after 700 cycles) and NaTi2(PO4)3 (98.1% after 1000 cycles) anodes. This highlighted its extensive potential for large-scale energy storage applications.
Introducing the HE concept into the PBAs structure further enhances its multiple performances. However, the severe agglomeration of HEPBA particles still limits its fast charging capacity. Zhang et al. prepared a HEPBA (Nax(FeMnCoNiCu)[Fe(CN)6]y□1−y·nH2O) with a hollow stepped spherical structure by chemically etching the traditional cubic structure of HEPBA.128 Electrochemical characterization, kinetic analysis, and COMSOL Multiphysics simulation showed that the HE nature and the hollow stepped spherical structure could greatly improve the diffusion behavior of Na+. Moreover, the hollow structure effectively alleviated the volume change of HEPBA, ultimately extending its service life (Fig. 12e). Therefore, the prepared HEPBA cathode exhibited excellent rate performance (126.5 and 76.4 mA h g−1 at 0.1 and 4.0 A g−1, respectively) and stable long-term performance (retaining 75.6% of its capacity after 1000 cycles), which was attributed to its unique structure. Additionally, the waste generated during the etching process could be easily recycled to prepare more HEPBA products. This processing method brings great hope for the design of advanced HE PBAs nanostructures for SIBs.
PBAs are promising SIB cathodes due to ease of synthesis and high Na+ diffusivity. Nevertheless, irreversible structural alterations resulted from Jahn–Taller distortion cause performance decline, particularly in Fe/Mn-based PBAs. Additionally, disproportionation reactions lead to ion dissolution in acidic electrolytes. The HE approach alleviates rapid capacity fading caused by irreversible structural changes during cycling via the cocktail effect from multi-metal coordination and structural stabilization. The application of HE strategies in PBAs in recent years is summarized, and the significant role of HE in PBAs is demonstrated through the composition and electrochemical performance shown in Table 2.
Cathode materials | Voltage range | Initial capacity (mA h g−1) | Cycle retention | Rate capacity (mA h g−1) | Ref. |
---|---|---|---|---|---|
Na1.26Mn0.4Fe0.15Ni0.15Cu0.15Co0.15[Fe(CN)6]0.81□0.19·1.12H2O | 2–4.1 V | 117 | 90% (200 cycles 0.1 A g−1) | 112 (0.1 A g−1) | 69 |
Na1.19(Fe0.2Mn0.2Ni0.2Cu0.2Co0.2)[Fe(CN)6]0.79□0.21·1.16H2O | 2–4.2 V | 120 | 94% (150 cycles 0.1 A g−1) | 96 (0.1 A g−1) | 95 |
Na1.70Fe0.2Mn0.2Co0.2Ni0.2Cu0.2[Fe(CN)6]0.98□0.02·2.35H2O | 2–4 V | 109.4 | 77.8% (2000 cycles 50 mA g−1) | 96.5 (10C) | 123 |
Na1.41Mn0.32Fe0.11Co0.28Ni0.32Cu0.32[Fe(CN)6]·2.89H2O | 2–4.2 V | 105.1 | 85.8% (200 cycles 1.5C) | 92.6 (1.5C) | 124 |
Na1.36Ni0.4Fe0.15Co0.15Mn0.15Cu0.15[Fe(CN)6]0.91□0.09·1.18H2O | 0–1.2 V | 118.6 | 81.2% (1800 cycles 100 mA g−1) | 90 (5C) | 125 |
Na1.65Mn0.4Fe0.12Ni0.12Cu0.12Co0.12Cd0.12[Fe(CN)6]0.92□0.08·1.09H2O | 1.5–4.2 V | 129 | 98.7% (200 cycles 0.5 A g−1) | 73 (0.5 A g−1) | 126 |
Na1.63Mn0.40Co0.22Cu0.25Ni0.06Fe0.06[Fe(CN)6]0.92·3.62H2O | 2–4 V | 120.2 | 80.1% (4000 cycles 2 A g−1) | 75.1 (0.5 A g−1) | 127 |
Na1.816Fe0.179Mn0.215Co0.215Ni0.195Cu0.196[Fe(CN)6]0.964□0.036·2.443H2O | 2–4.2 V | 126.5 | 75.6% (1000 cycles 1 A g−1) | 124 (0.1 A g−1) | 128 |
As shown in Fig. 13a, HE-NVPF exhibited excellent crystallinity, demonstrating a strong correlation with tetragonal structures belonging to the space group P42/mnm. The six cations were randomly dispersed among the 8j sites. Fig. 13b showed the schematic structures of Na3V2(PO4)2F3 (p-NVPF) and Na3V1.9(Ca, Mg, Al, Cr, Mn)0.1(PO4)2F3 (HE-NVPF), which exhibited an open architecture.78 As shown in Fig. 13c and d, the HE effect significantly enhanced the operating voltage, which in turn boosted the discharge capacity of the NVPF cathode and mitigated adverse effects during low-voltage discharge. Within the NVPF structure, there were two types of sodium sites, Na (1) and Na (2). The Na (1) sites demonstrated minimal electrochemical activity and didn't participate in redox reactions, while the Na (2) sites were highly electrochemically active, facilitating Na+ intercalation/deintercalation. In the HE structure, the proportion of Na (2) sites increased from 50% to 60.2%. This implied that the diffusion rate of Na+ was enhanced. Due to the substantial accumulation of Na+ at the Na (2) sites, the Na+ concentration at the Na (1) sites decreased. The Na (2) sites could more efficiently facilitate the rearrangement of Na+, thereby effectively suppressing phase transition. As shown visually in Fig. 13e, it depicted the migration path of Na+ moving from the Na (2) site to the Na (1) site in both p-NVPF and HE-NVPF. In these diagrams, IS represented the initial state, TS standed for the transition state, and FS indicated the final state. The computed energy barriers for sodium migration were 1.984 eV for p-NVPF and 0.963 eV for HE-NVPF. Notably, HE-NVPF exhibited significantly lower migration barriers than p-NVPF, which enhanced the diffusion of Na+. Furthermore, the final state energy of HE-NVPF (0.5370 eV) was considerably lower than that of p-NVPF (1.6750 eV). This suggested that Na+ in the HEM exhibited a greater tendency to escape, thereby reducing the likelihood of Na site rearrangement and inhibiting reactive phase transitions in the low-voltage region.133–136 The feasibility of the HE approach was further verified. The HE structure significantly enhanced the low-voltage discharge plateau of the material, providing a higher operating voltage and increased energy density. Meanwhile, the highly stable structure achieved through structural regulation delivered good cycling stability (2000 cycles at 2C with 80% capacity retention), suppressed the irreversible phase transition, and induced the escape of Na+ from the Na (2) site. This provided valuable inspiration for the design of SIB cathodes.
![]() | ||
Fig. 13 (a) The XRD pattern of HE-NVPF. (b) The structures of p-NVPF and HE-NVPF. (c) GCD curves within the potential window of 2.0–4.3 V. (d) Cs values and capacity contributions during discharge. (e) Na+ migration pathways in p-NVPF. Reprinted with permission from ref. 78. Copyright 2022 Wiley VCH. COMSOL simulation and postmortem analysis. (f) Stress field analysis at different depths of discharge (DOD) states based on the COMSOL platform. Reprinted with permission from ref. 128. Copyright 2025 RSC. |
For the advancement of high-performance NASICON-type cathode materials for SIBs, it is of great significance to prevent severe structural deformation and irreversible phase transition, as well as to achieve stable multi-electron redox reactions.133,134 Du et al. synthesized a high-entropy Na3.45V0.4Fe0.4Ti0.4Mn0.45Cr0.35(PO4)3 (HE-Na3.45TMP) cathode material via ultrafast high-temperature shock treatment.137 This method effectively reduced the likelihood of phase separation and enabled reversible and stable multi-electron transfer of 2.4/2.8 e− with respect to Na+/Na within the voltage ranges of 2.0–4.45/1.5–4.45 V, corresponding to capacities of 137.2/162.0 mA h g−1. Constant current charge–discharge experiments and in situ X-ray diffraction tests demonstrated the continuous redox reaction and approximate solid solution phase transition process of HE-Na3.45TMP. DFT calculations were carried out to analyse the migration pathways and energy barriers, further validating the superior reaction kinetics of HE-Na3.45TMP. Consequently, HE-Na3.45TMP showed outstanding wide-temperature adaptability and could operate stably in the temperature range from −50 to 60 °C. After 400 cycles at −40 °C, the capacity retention rate reached up to 92.8%, and it could still maintain a capacity of 73.7 mA h g−1 at −50 °C. The assembled hard carbon//HE-Na3.45TMP full cell offered an energy density of approximately 301 Wh kg−1, which verified the application feasibility of HE-Na3.45TMP. This work provided an innovative and ultrafast approach for the rational fabrication of high-performance cathodes for SIBs.
Dong et al. prepared spherical HE Na4Fe2.95(MgCaAlCrMn)0.01(PO4)2P2O7 (HE-NFPP) cathode materials via spray drying in combination with solid-state reaction.138 By adopting the HE approach, the co-doping of TMs in NFPP could effectively mitigate volume variations. This was achieved by decreasing the band gap between the conduction band and the valence band, which in turn improved its electronic conductivity. Additionally, through the HE strategy, the co-doping of TM ions in NFPP creates three-dimensional network Na+ ion diffusion channels. These channels were beneficial for Na+ ion diffusion, thus enhancing the electrode reaction kinetics at low temperatures.
NASICON-type compounds possess the advantages of rapid Na+ diffusion and structural stability. Nevertheless, they still suffer from several drawbacks, such as a low specific capacity (less than 120 mA h g−1) and significant volume changes. HE engineering is a method aimed at enhancing structural and functional stability by using multiple dopants. This strategy has demonstrated great potential in layered oxides. However, applying it to NASICON is inherently difficult because of the rigid poly-anionic framework, dopant incompatibility, and redox mismatch. Zhang et al. presented an entropy-stabilized NASICON cathode, namely Na3.2V1.5Cr0.1Mn0.1Fe0.1Al0.1Mg0.1(PO4)3 (HE-V).139 It was synthesized via complex doping at the 12c Wyckoff site. The HE method enabled the fabrication of single-phase materials with enhanced redox flexibility and decreased lattice strain. The HE-V material was able to reach a high reversible capacity of 170 mA h g−1 by means of multi-electron V5+/V4+/V3+/V2+ redox reactions. Additionally, it exhibited outstanding rate performance and cycling stability, retaining its performance after over 10000 cycles at 50C. Both in situ and ex situ characterizations demonstrated near-zero strain structures, low defect generation, and strengthened local bonding (Fig. 13f). Theoretical computations further confirmed a narrowed bandgap and improved charge transport. This study had created a new type of HE polyanionic cathodes, overcoming long-standing constraints in NASICON chemistry and providing a practical path for the development of long-lasting and high-energy SIBs.
Polyanion cathodes, characterized by three-dimensional framework architectures and distinctive tetrahedral PO4 units, exhibit remarkable structural stability and a high operating voltage. However, the relatively low electronic conductivity of these polyanion cathode materials severely undermines their electrochemical performances. For instance, they display a low specific capacity and inferior rate performance. The HE approach enhances the energy density and extends the cycle life of the polyanionic cathode through multi-scale coordination, structural stabilization, and dynamic optimization. In the future, the relationship between element ratios and synthesis processes can be further explored, and the optimal HE combination can be predicted using machine learning. In addition, interface-volume phase multi-scale entropy regulation and the development of wide-voltage-window electrolytes will be the key directions to improve the performance of the entire battery. Based on the electrochemical performance of HE polyanion cathodes reported so far, which is summarized in Table 3, the HE doping strategy plays a significant role in enhancing the specific capacity and cycling stability of polyanion cathodes. However, research in this field is still in its early stages and requires further efforts.
Cathode materials | Voltage range | Initial capacity (mA h g−1) | Cycle retention | Rate capacity (mA h g−1) | Ref. |
---|---|---|---|---|---|
Na3V1.9(Ca,Mg,Al,Cr,Mn)0.1(PO4)2F3 | 2–4.3 V | 118.5 | 80.4% (2000 cycles 20C) | 71.4 (50C) | 78 |
Na4Fe2.85(Ni,Co,Mn,Cu,Mg)0.03(PO4)2P2O7 | 1.5–4.2 V | 122 | 82.3% (1500 cycles 10C) | 85 (50C) | 132 |
Na4Fe2.95(NiCoMnMgZn)0.01(PO4)2P2O7 | 1.5–4.2 V | 124.6 | 93% (5000 cycles 20C) | 82.7 (1C) | 133 |
Na4Fe2.5Mn0.1Mg0.1Co0.1Ni0.1Cu0.1(PO4)2(P2O7) | 1.4–4.0 V | 108.9 | 99.2% (200 cycles 1 A g−1) | 55.0 (10 A g−1) | 134 |
Na3.12MnTi0.9(VFeMgCrZr)0.02(PO4)3 | 1.5–4.3 V | 169.6 | 81.7% (1000 cycles 5C) | 132.3 (1C) | 135 |
Na3.6VMn0.4Fe0.4Ti0.1Zr0.1(PO4)3 | 1.5–4.3 V | 110 | 80.6% (10![]() |
78.5 (20C) | 136 |
Na3.45V0.4Fe0.4Ti0.4 Mn0.45Cr0.35(PO4)3 | 1.5–4.45 V | 162 | 83.7% (2000 cycles 5C) | 73.7 (5C) | 137 |
Na4Fe2.95(NiCoMnMgZn)0.01(PO4)2P2O7 | 1.7–4.3 V | 106.5 | 92.0% (1000 cycles 1C) | 67.1 (50C) | 138 |
Na3.2V1.5Cr0.1Mn0.1Fe0.1Al0.1Mg0.1(PO4)3 | 1.5–4.1 V | 170 | 93.3% (10![]() |
160 (0.2C) | 139 |
In conclusion, HE approach offers a revolutionary method for developing advanced materials for SIBs, which dramatically improves the ΔSconfig and effectively contributes to the design of battery materials. This review summarizes the underlying mechanisms and highlights six advantages of HEMs. HEMs can be seen as extensions of four effects observed in HEAs: (1) reduced volume strain, (2) inhibited adverse phase transitions, (3) facilitated conversion processes, (4) improved ion transport, (5) enhanced disordered ion diffusion, and (6) cocktail-induced synergistic mechanisms. These advantages are not mutually exclusive, instead, multiple research advantages may coexist within a single HEBM. Furthermore, this review establishes a composition–structure–dynamics framework that integrates HE design principles with electrochemical performance across SIB cathodes (Fig. 15). Departing from static configurations, we emphasize operationally adjustable entropy as a dynamic lever to optimize material behavior. By synthesizing recent advances in layered oxides, PBAs, and polyanionic systems, demonstrate how HEMs mitigate lattice distortion, suppress phase transitions, and enhance Na+ diffusion kinetics through multi-element synergies. This study further proposed the design standards for entropy-driven cathode engineering, outlined the current challenges and the future research directions for HE cathodes. This framework provides actionable insights for developing next-generation SIBs.
![]() | ||
Fig. 15 The application of HE strategies in the field of SIBs is summarised on the left, while the future research directions that require to be explored are showed on the right. |
This journal is © The Royal Society of Chemistry 2025 |