Emerging multimetal LMFP-based cathodes for lithium-ion batteries: a review

Josué M. Gonçalves *a, Syra Mubarac b, Gustavo T. M. Silva ac, Bruno Freitas b, Benedito G. Aguiar Neto a and Hudson Zanin *b
aMackenzie Institute for Research in Graphene and Nanotechnologies (MackGraphe), Mackenzie Presbyterian Institute, São Paulo, SP, Brazil. E-mail: josue.goncalves@mackenzie.br
bAdvanced Energy Storage Division, Center for Innovation on New Energies, School of Electrical and Computer Engineering, University of Campinas, Av Albert Einstein 400, Campinas, SP 13083-852, Brazil. E-mail: hzanin@unicamp.br
cCenter for Studies in Energy and Oil, University of Campinas, Rua Cora Coralina 350, SP, Campinas, 13083-896, Brazil

Received 22nd April 2025 , Accepted 16th September 2025

First published on 23rd September 2025


Abstract

The increasing demand for smart portable electronics and electric vehicles is driving advancements in high-energy-density lithium-ion batteries (LIBs). Among various cathode materials, lithium manganese iron phosphate (LiMnyFe1−yPO4 – LMFP) stands out due to its cost-effectiveness, excellent safety profile, long cycle life, high operating voltage, robust thermal stability, and competitive energy density. However, despite notable progress, LMFP still encounters key challenges such as limited electronic conductivity, sluggish Li-ion diffusion, manganese dissolution affecting cycling stability, and low tap density. To address these issues, significant efforts have been devoted to developing multimetal LMFP and other multimetal olivine-based cathodes with enhanced electrochemical properties. Nevertheless, a comprehensive review of these recent advancements remains lacking. This article aims to bridge this gap by examining emerging strategies and recent developments in multimetal LMFP cathodes for LIBs. It systematically compares various optimization approaches, including (i) doping with single or multiple elements at Li-, M-, or Li & M-sites, (ii) integrating additional primary constituents or structural components, and (iii) increasing the configurational entropy of electrode materials. The advantages and limitations of these strategies are critically assessed. Furthermore, key synthesis methodologies and processing techniques used to enhance LMFP cathode performance are discussed. Finally, the review provides insights into the benefits and challenges associated with these materials, highlighting future perspectives and potential research directions.


image file: d5ta03194j-p1.tif

Josué M. Gonçalves

Josué Martins Gonçalves is a Researcher at the Mackenzie Institute for Research in Graphene and Nanotechnologies (MackGraphe) within the Mackenzie Presbyterian Institute (IPM), Brazil. He graduated in Chemistry at the University Vale do Acaraú (UVA) in 2014 and received his PhD from the Institute of Chemistry at the University of São Paulo (USP) in 2019. He developed cutting-edge research during his postdoctoral internships at USP (2019–2021), the University of Illinois-Chicago (UIC, 2021–2022), and the University of Campinas (Unicamp, 2023–2024). His current research interests include applications of high-entropy materials for energy conversion and storage.

image file: d5ta03194j-p2.tif

Syra Mubarac

Syra Mubarac is a PhD candidate at the State University of Campinas (UNICAMP) and a member of the Advanced Energy Storage Division of the Center for Innovation on New Energies, Brazil. She obtained her bachelor's degree in chemical engineering from the State University of Amazonas (UEA) and her MSc from the School of Electrical Engineering at the UNICAMP in 2024 with emphasis on Electrical Energy. She has worked on the development of lithium-ion batteries, from synthesis to manufacturing. Her current research interests include the study and synthesis of multi-metallic and high-entropy materials applied to intercalation energy storage systems.

image file: d5ta03194j-p3.tif

Gustavo T. M. Silva

Gustavo Thalmer M. Silva obtained his undergraduate degree in Chemistry from the State University of Rio Grande do Norte (2012) and a Master’s degree in Chemistry from the Federal University of Rio Grande do Norte (2014). He received his PhD at the Institute of Chemistry of the University of São Paulo (IQ-USP) in 2019 under the supervision of Frank H. Quina and is currently conducting postdoctoral research at the same institution. His research interests include energy sources, colloid and materials chemistry, the chemistry and photochemistry of natural pigments and their synthetic analogues, and quantum chemical methodology for electronically excited states.

image file: d5ta03194j-p4.tif

Bruno Freitas

Bruno Guilherme Aguiar Freitas is a Postdoctoral Researcher at the Center for Innovation in New Energies, affiliated with the University of Campinas (UNICAMP), Brazil. He holds a B.Sc. (2014) and an M.Sc. (2017) in Chemistry from the Federal Fluminense University (UFF) and a Ph.D. in Electrical Engineering (2023) from UNICAMP. He specializes in energy storage materials, using the Brazilian synchrotron, Sirius, and international facilities like SLAC (USA) and MAX IV (Sweden) for in situ/operando XRD studies of supercapacitors and batteries. His current work focuses on synthesizing and characterizing electrodes for batteries and supercapacitors using electrochemical and synchrotron-based techniques.

image file: d5ta03194j-p5.tif

Benedito G. Aguiar Neto

Benedito Guimarães Aguiar Neto holds a degree in Electrical Engineering from the Federal University of Paraíba (1977), a Master's degree in Electrical Engineering from the Federal University of Paraíba (1982), Brazil, and a PhD in Electrical Engineering from the Technische Universität Berlin, Germany (1987). Postdoctoral fellow at the University of Washington, USA (2017–2018) and at the University of Minho, Portugal (2021–2022). He was a Professor at the Federal University of Campina Grande, Paraíba, Brazil, until 2011, and since 2011, he has been at the Mackenzie Presbyterian University, São Paulo, Brazil, where he was Rector until 2020. Since December 2022, he has been General Director of the Mackenzie Institute for Research in Graphene and Nanotechnologies (MackGraphe).

image file: d5ta03194j-p6.tif

Hudson Zanin

Hudson Zanin is a Professor at the School of Electrical and Computer Engineering of the University of Campinas, Brazil, and a Researcher in the Advanced Energy Storage Division of the Center for Innovation on New Energies. His research focuses on developing functional (nano)materials and their application in Energy Storage and Conversion Devices. He has experience in supercapacitors, batteries, and fuel cells' materials and devices. He is the founder of the first South American lab-scale batteries and supercapacitors manufacturing. Dr Zanin has a PhD degree in Electrical Engineering from the University of Campinas (2012).


1. Introduction

The growing demand for advanced portable electronics and electric vehicles is driving innovations in lithium-ion batteries (LIBs),1,2 with enhanced energy density, ushering in a new era in the energy transition. In the pursuit of next-generation energy storage systems, significant efforts have been focused on the development of next-generation cathode materials, with a strong emphasis on designing nickel-rich layered oxides,3,4 especially super nickel-rich (cobalt-low) compositions. This research particularly centers on compounds such as LiNixCoyMnzO2 (NCM), LiNixCoyAlzO2 (NCA), and LiNixCoyMnzAl1−xyzO2 (NCMA), where x ≥ 0.85.5 While nickel-rich cathodes hold significant promise, these compositions alone are unlikely to meet future demands, particularly given the growing emphasis on safety in energy storage devices.6

Regarded as a strong alternative to nickel-rich cathodes, lithium manganese iron phosphate (LiMnyFe1−yPO4, LMFP) has attracted considerable interest as a potential cathode material for LIBs, owing to its low cost, excellent safety characteristics, prolonged cycle stability, elevated operating voltage, reliable performance under high-temperature conditions, and competitive energy density.7 In fact, LMFP is viewed as an advanced version of LiFePO4 (LFP), offering a slightly higher operating voltage (∼3.5 V vs. Li/Li+ for the Fe2+/Fe3+ redox couple) than single-phase LFP (3.4 V vs. Li/Li+). It is compatible with existing LFP production lines, making it a cost-effective choice for industrial-scale production, aiming to accelerate its commercialization.8

Despite substantial progress over recent decades, LMFP still faces major challenges, including limited electronic conductivity, sluggish lithium-ion diffusion, manganese dissolution impacting cycling stability, and low tap density. These drawbacks greatly affect its energy storage capacity, rate capability, and cycling stability, limiting its broader adoption in high-performance LIBs.7 To overcome these limitations, various modification strategies, such as morphology control, elemental doping, applying conductive surface coating, micronizing primary particle size, and carefully managing defect concentrations, are widely used to boost conductivity, facilitate lithium-ion transport, and optimize both kinetics and electrochemical performance.9

More recently, the design of multielemental cathode materials has been an essential strategy to improve stability and overall electrochemical performance, especially in electrode materials for energy storage. Notably, these enhancements can be achieved by (i) doping with one or multiple elements, (ii) adding one or more principal constituents or structural components, and/or (iii) increasing the entropy of the electrode materials. Doping with various elements has been investigated as a strategy to improve both electronic and ionic conductivity. Dopants, which can be either cations or anions, may substitute at Li-sites, Mn/Fe-sites, or anionic (PO4) sites.10 On the other hand, when the dopant concentration exceeds 5% by mass or atomic percentage, it should be regarded as a primary constituent or structural component of the electrode material. These primary constituents must be carefully selected, as higher concentrations can improve stability but may lead to reduced specific capacity, particularly when dealing with electrochemically inactive elements. Similar to the addition of primary constituents, when these elements are in near-equimolar proportions, entropy stabilization, and/or synergistic effects may enhance electrochemical properties. For instance, incorporating a third or fourth element can yield a medium-entropy material (MEM), whereas the presence of five or more elements results in a high-entropy material (HEM).11

Recent review articles have spotlighted advancements and innovative strategies in the design of LMFP cathode materials. For instance, Xu et al.8 reviewed recent advancements and challenges in optimizing strategies to enhance the electrochemical performance of LMFP. Their work highlights approaches such as fine-tuning the Mn/Fe ratio, compositing with conductive materials, elemental doping, and morphology control. Particular emphasis was placed on regulating the additional discharge plateau, a critical factor for preventing energy density loss while ensuring the consistency and reliability of LMFP-based batteries. Similarly, Zhang and colleagues7 reviewed the reaction mechanisms, synthesis approaches, and electrochemical behavior of LMFP, offering insights that provided guidance for better material development. They outlined key challenges, proposing strategies like multi-scale particle tailoring, heteroatom doping, surface coating, and structural morphology engineering to improve kinetics and overall electrochemical properties. Furthermore, Akhmatova and collaborators10 emphasized recent advancements in multi-element doping of LFP cathodes as an effective strategy to enhance their electrochemical performance. However, to our knowledge, no reviews have comprehensively addressed the promising outcomes, comparisons, and distinctions associated with elemental doping, including principal constituents or structural components, and/or entropy enhancement in LMFP materials. While there has been a noticeable increase in publications on this subject in recent years, showcasing intriguing performances of multimental LMFP cathodes, a detailed analysis remains absent.

In that regard, this review aims to fill this gap, highlighting recent advancements, trends, and emerging strategies in the development of multimental LMFP-based cathodes for LIBs (Scheme 1). It provides a comprehensive comparison of approaches, including (i) doping with single or multiple elements, (ii) incorporating additional primary constituents or structural components, and (iii) enhancing the entropy of electrode materials. The advantages and limitations of these strategies are thoroughly analyzed. Additionally, the review outlines key synthesis protocols and techniques employed to optimize the electrochemical properties of LMFP cathodes. Finally, the benefits and drawbacks of these materials are critically evaluated, and insights into future directions and potential advancements are presented.


image file: d5ta03194j-s1.tif
Scheme 1 Key optimization strategies and substitution sites employed in the design of multimetal LMFP cathodes. At the center, the primary metal ions utilized in the development of advanced LMFP and other multimetal olivine cathodes. The LMFP structural model shown at the core of the scheme has been adapted with permission from ref. 12. Copyright © 2025 Wiley-VCH GmbH.

2. Principles and challenges in LIBs

Since their commercial introduction by Sony Corporation in 1991, lithium-ion batteries (LIBs) have attracted significant attention as a key technology for energy storage.5,13–17 Their relatively high energy density, lightweight design, fast recharge capability, long cycle life, and environmental friendliness have established LIBs as the dominant energy source for both portable electronic devices, electric vehicles (EVs), and stationary energy storage devices.5,13–17 In the context of EVs, there is a growing demand for higher battery capacity, improved performance, and greater stability to effectively replace internal combustion engines and accelerate the global adoption of EVs. Therefore, LIBs have shown great promise in meeting these demands due to their qualities.18

Fig. 1 shows the composition of a LIB. In LIBs, the charge–discharge process is reversible, requiring both oxidation and reduction to take place at the same electrode; nevertheless, the negative and positive electrodes are conventionally referred to as the anode and cathode, respectively. During the discharging process, positively charged lithium ions move from the anode (where oxidation occurs) to the cathode (where reduction takes place). These ions are transported through the electrolyte, which facilitates ion movement, and pass through the separator, which prevents the free flow of electrons and avoids short circuits between the electrodes.5,13–19 As the ions reach the electrodes, electrons flow through the external circuit. The current collectors serve as a connection (providing the pathway) for the flow of electrons between the electrodes and the external circuit. On the other hand, during the charging process, the positive lithium ions are released from the cathode to the anode.5,13–20


image file: d5ta03194j-f1.tif
Fig. 1 Schematic representation of the composition and discharge/charge cycle for LIBs.

In order to find solutions to mitigate CO2 emissions, governments have been facilitating the global production of EVs.21–23 However, the next generation of EVs requires LIBs with enhanced structural stability, higher energy density, and improved performance to meet the demands of long driving ranges between charges.24,25 Since electrode conductivity plays a decisive role in the electrochemical response of LIBs, advances in anode and cathode design are crucial. Research in this field has fostered the development of alternative active materials intended to reduce production costs and increase safety, while also enhancing stability and key charge–discharge parameters such as voltage, reversibility, and capacity.26–29 The use of graphite as anode material has been well stablished, with moderate specific capacities of c.a. 372 mAh g−1 and low cost.30 Therefore, the cathode material is a determining factor not only in the cost but also in the energy density and power density of LIBs.

Among conventional cathode materials, LiCoO2 (LCO), LiNi1−xyCoxMnyO2 (NCM), and LiFePO4 (LFP) are some of the most widely used. Although LCO presents a high theoretical specific capacity (274 mAh g−1) and energy density, the actual specific capacity is only about half of the theoretical specific capacity under a cutoff voltage of 4.2 V.31 For a higher cutoff voltage, lattice distortion and deterioration during cycling can occur. In addition, safety problems have occurred due to the dissolution of Co in the electrolyte and the decomposition of electrolyte.7 NCM, which tends to replace the determined fraction of Co by Ni and Mn, has been an alternative for improving specific capacity and energy density,4 mainly when increasing the Ni-content.5 However, problems caused by Li/Ni cation mixing, microcrack formation, and interfacial side reactions have hampered commercial applications.5 On the other hand, LFP stands out as a highly attractive option for energy storage. Composed of abundant, low-cost, and environmentally friendly raw materials, it offers a safer, more economical, and less toxic alternative to the LCO and their cobalt-containing derivatives, which face availability constraints.19,32 In addition, it also exhibits superior thermal stability and longer cycle life.33,34 However, LFP exhibits unsatisfactory performance in low-temperature environments and has a relatively low energy density32 compared to other LIB cathode materials, which limits its use in applications with high energy demand, such as EVs.34 Besides, the intrinsically sluggish lithium-ion diffusion and limited electronic conductivity significantly hinder its rate capability.35 Fortunately, considering the strengths and limitations of the aforementioned cathode materials, each chemistry can play a complementary role in battery technology. Rather than aiming for full substitution, their optimal value lies in strategic integration (i.e., blended cathode design)36 or application-specific deployment, leveraging the unique advantages of each material for tailored energy storage solutions.

On the other hand, the addition of manganese, replacing partially iron, forms lithium manganese iron phosphate (LiMnyFe1−yPO4, LMFP), an innovative and environmentally friendly cathode with higher energy density and good cycle stability, which combines the safety and stability of LFP with the higher voltage characteristics of lithium manganese phosphate (LMP).33 LMFP can offer enhanced energy density due to the higher operating voltage of the Mn2+/3+ redox couple (∼4.0 V) compared to that of Fe2+/3+ (∼3.4 V), and its theoretical energy density can reach ca. 21% more than that of LFP.37 Indeed, LMFP offers several advantages, including low cost, high safety, long cycling life, elevated voltage, good high-temperature performance, and improved energy density.7,38,39 However, LMFP also faces several challenges that affect the overall performance and limit its applications.

Despite the numerous advantages of LMFP, there are still structural limitations that compromise its electrochemical performance and hinder its widespread application. Olivine-type structure phosphates exhibit intrinsic limitations, such as low electrical conductivity and slow lithium-ion kinetics.40 In fact, some structural challenges in LMFP cathode materials are associated with the Jahn–Teller effect triggered by Mn3+ ions. The Jahn–Teller effect of the Mn3+ ion causes kinetic limitations of Mn2+/3+ redox, resulting in lattice strain, poor lithium-ion diffusion, and low electronic conductivity.33,37,41 This leads to significant polarization under high currents during charge–discharge cycles, affecting its performance and capacity.33 Additionally, low-temperature environments exacerbate these issues by reducing conductivity and increasing ion migration impedance.33,41 The oxidation of Mn2+ to Mn3+ during LMFP delithiation induces significant MnO6 octahedral distortion, increasing the activation energy for ionic diffusion. This structural deformation manifests through the selective elongation of two equatorial Mn–O bonds (edge-sharing with PO4 tetrahedra) and the contraction of the remaining Mn–O bonds, differing fundamentally from classical Jahn–Teller distortion.42–48 Normally, the Jahn–Teller distortion in Mn3+O6 octahedra manifests as axial bond elongation (c/a > 1) with equatorial bond contraction (Fig. 4a).

However, LiMnPO4 exhibits a distinct distortion pattern due to structural constraints imposed by (PO4)3− tetrahedra. During delithiation, this pre-existing structural configuration directs the Mn3+ Jahn–Teller distortion toward an unusual c/a < 1 geometry, characterized by the shortened axial Mn–O1/Mn–O2 bonds and the elongated equatorial Mn–O3/Mn–O3′ bonds (Fig. 4b). This inverse distortion results from the combined effects of Mn–P Coulomb repulsion and strong (PO4)3− bonding, which preferentially stabilizes the 3z2r2 orbital configuration while driving Mn3+ displacement toward O3–O3 edges (Fig. 4c). The bond length changes confirm this atypical distortion behavior, demonstrating how polyhedral connectivity in olivine phosphates can fundamentally alter conventional Jahn–Teller expectations.43 The structural distortion induced by the formation of Mn3+ during the charged state of the MnO6 octahedra increases the obstruction of Li+ migration. This distortion reduces Li+ diffusivity between the initial (Mn2+-rich) and fully charged (Mn3+-rich) states. This kinetic limitation originates from the lattice strain generated by the Mn2+/Mn3+ redox couple, which distorts the crystal lattice and obstructs Li+ transport channels. The strong correlation between Mn3+ concentration and decreased Li+ mobility confirms that mitigating this Jahn-Teller distortion is critical for developing LMFP cathode materials with enhanced rate capability.44,45 This deformation significantly increases internal resistance within the LMFP lattice, resulting in an extremely low ion diffusion coefficient, which hampers the efficiency and performance of these materials in electrochemical systems due to the kinetic barrier at the interfacial diffusion in the resulting biphasic interface.49

In order to solve the aforementioned challenges and enhance the performance of LMFP, various modification strategies have been implemented. Key approaches include surface coating, elemental doping, morphology control, material nanosizing, and electrolyte system regulation.41 Among these, elemental doping in the LMFP crystal lattice stands out for its potential to enhance the material's intrinsic conductivity. For instance, cation doping can cause changes in lattice vacancies or atomic bond lengths, facilitating lithium movement within the lattice and improving electrochemical performance. Thus, based on the growing body of literature and several reviews that broadly outline advances in strategies for enhancing LMFP cathodes, this review specifically emphasizes (with greater depth) the promising outcomes, key comparisons, and notable distinctions related to elemental doping, incorporation of principal constituents or structural components, and/or entropy enhancement in LMFP materials. Although a marked increase in publications has been observed in recent years (many reporting impressive performances of multimetal LMFP cathode), a comprehensive and critical analysis of these developments is still lacking, as will be discussed later.

3. LMFP: structural features and electrochemical aspects

Lithium iron and manganese phosphate-based cathodes (LiMnyFe1−yPO4, LMFP) hold great prospects for large-scale production in advanced lithium-ion batteries. In fact, several companies (mainly in China) have announced pilot-scale LMFP cathodes and cell prototypes, aiming to use them in electric vehicles (EVs) and energy storage systems as a mid-point technology between LFP and NMC. The combination of high performance at higher operating voltages provided by lithium manganese phosphate (LiMnPO4, LMP) with the high safety of lithium iron phosphate (LiFePO4, LFP) makes this material a strong competitor for the commercial batteries for EVs and high-power energy storage systems, due to its promise of environmental compatibility and high energy density compared to LFP.7,10 Indeed, partially substituting iron with manganese in the LFP structure can enhance the energy density by approximately 15–20%.50

The manganese (Mn2+) inserted into the crystal structure of LFP is primarily responsible for giving LMFP the characteristic of operating within a high voltage working window when applied in electrochemical devices. Moreover, LMFP materials exhibit an asymmetric charge–discharge voltage electrochemical signature, primarily due to their intricate phase transition processes.51 With an increased operational voltage range, the system tends to exhibit a higher upper limit of specific energy density, as this energy density is the result of the product of voltage and specific capacity.7 LMFP solid solutions, with varying Fe/Mn ratios,52 are isostructural with their LFP and LMP endmembers.53 These mixed-metal phosphates adopt the same olivine-type lattice, maintaining structural consistency across different compositions.53–55 However, the distinct d-electron configuration of Mn introduces additional complexity in its electronic rearrangement and lattice evolution mechanisms.51

The olivine structure exhibited by LMFP is orthorhombic (space group Pnma), consisting of PO4 tetrahedra with both Fe and Mn ions occupying octahedral sites where Fe2+ and Mn2+ ions share corners (4c sites), phosphorus ions are located in tetrahedral sites, and Li+ ions occupy the edge-shared octahedral site (4a sites), with the latter running parallel to the b-axis (Fig. 3a and b).53,54 In a more detailed study, Kope et al.56 described LMFP as a stable solid solution due to slight lattice expansion caused by the small difference in ionic radius between high-spin Fe2+ and Mn2+ ions. They also demonstrated that the Mn-rich phase (y > 0.8) in LixMnyFe1−yPO4 was unsuitable for high-specific-capacity cathode materials due to intrinsic limitations arising from strong electron–lattice interactions (Mn3+:3d4, Jahn–Teller effect) in the charged state (Mn3+/Fe3+) that induces lattice distortions.54,56

Computational studies by Gardiner & Islam53 employing three distinct Li+ migration pathways were studied in LiFe0.5Mn0.5PO4: (1) Path α along [010] (b-axis, 2.9–3.0 Å), (2) Path β along (001) (4.6–4.7 Å), and (3) Path γ along (101) (5.6–5.8 Å) (Fig. 3c). The analysis revealed that the lowest energy barrier (Emig = 0.59 eV) corresponds to one-dimensional migration along the (010) crystallographic direction (Path α), characterized by a 2.9–3.0 Å jump distance between adjacent Li sites (Fig. 3d and e). This finding agrees remarkably well with experimental activation energies (0.63 eV) reported for LiFe0.45Mn0.55PO4. In contrast, alternative pathways along (001) (Path β) and (101) (Path γ) directions exhibited prohibitively high energy barriers (>2.2 eV), effectively restricting Li+ mobility to the b-axis channels and confirming the strongly anisotropic nature of ionic diffusion in olivine-type materials. Detailed analysis of the migration trajectory revealed a curved pathway between adjacent lithium sites (Fig. 2d3e), with a maximum deviation of 0.44 Å from linearity, in agreement with neutron diffraction studies. Therefore, Li+ tends to move along the b-axis (010) (Fig. 3c) in one-dimensional channels, and diffusion transport in only one-dimension results in limited Li+ kinetics due to potential lattice defects. This severely restricts the rate at which lithium ions can move during electrochemical charge–discharge cycles.49


image file: d5ta03194j-f2.tif
Fig. 2 (a) Classical Jahn–Teller effect. Reproduced with permission from ref. 47. Copyright © 2013 American Chemical Society. (b) The distorted interstitial octahedron in the LiMnPO4 to MnPO4 framework. Reproduced with permission from ref. 43. Copyright © 2012 Elsevier B.V. All rights reserved. (c) Effects of (PO4)3− Coulomb repulsion and strong bonding on framework distortion. Adapted with permission from ref. 44. Copyright © 2016 Elsevier Ltd all rights reserved.

image file: d5ta03194j-f3.tif
Fig. 3 Olivine-type structure of LMFP illustrating arrangement of FeO6 and MnO6 octahedra; (a) regular and (b) segregated planes. Reproduced with permission from ref. 53. Copyright © 2009 American Chemical Society. (c) Illustrates the material LiMPO4 (M = Mn, Fe) for Li+ migration in olivine structure. The dotted circles denote the most probable path for Li+ migration in the (d) a and b plane view and (e) a–c plane view. Reproduced with permission from ref. 49. Copyright © 2025 by the Korean Electrochemical Society.

In addition to the aforementioned Li+ migration studies, extensive investigations have characterized the phase transition behavior of olivine LMFP cathodes during charge/discharge processes, revealing both “two-phase transition” and “single-phase solid-solution” mechanisms to explain lithiation and delithiation processes.52,57–62 Moreover, it is widely recognized that the Mn/Fe ratio plays a pivotal role in determining the material's phase transition behavior (see Ling et al.12 for a detailed review).

The phase transition mechanism of Li+ extraction/insertion involves the emergence of a new phase during the cycle and its disappearance after the cycle completion.57 This two-phase transition exhibits characteristic reversibility, with a single-phase transition at the beginning and the end of charge (x ≤ 0.42 and x ≥ 0.91, respectively), while a two-phase coexistence region (olivine and heterosite) appears at intermediate states (0.42 < x ≤ 0.91) in Li(1−x)Mn0.7Fe0.3PO4 (0 ≤ x ≤ 1.0). The resulting phase interface significantly impacts kinetic behavior, where lattice mismatch at the boundary increases electronic transport barriers and reduces conductivity.62 In contrast to the two-phase transition model, the solid-solution transformation model proposes continuous and homogeneous phase evolution with lithium content variation. Particularly during Li+ insertion, the phase transition from LxMFP to LMFP occurs through a continuous solid-solution pathway, where the LxMFP composition progressively evolves until reaching stoichiometric LMFP (Fig. 4). This coherent solid-solution behavior significantly enhances the rate capability during charge/discharge cycles.63,64 Fortunately, other recent reviews51,65,66 have summarized and outlined the phase transition mechanisms of LMFP during charge/discharge, describing three reaction regions linked to Fe2+/Fe3+ and Mn2+/Mn3+ redox processes, highlighting how factors like Mn/Fe ratio, particle size, and synthesis conditions influence structural evolution and electrochemical performance (see for more details12,65,66).


image file: d5ta03194j-f4.tif
Fig. 4 The coherent transformation model shows hysteresis during the first charge (Li extraction) and discharge (Li insertion) cycles as LMFP transforms to LxMFP. Adapted with permission from ref. 63. Copyright © 2014 American Chemical Society.

Interestingly, doping strategies, whether with single or multiple metals, profoundly influence these phase transition pathways, often shifting the reaction kinetics from a dominant two-phase model towards a more favorable solid-solution behavior, thereby enhancing structural stability and electrochemical performance.67–69 The beneficial impact of doping is further highlighted by comparing pristine and doped materials. While Zn doping did not radically alter the sequence of phase transitions-both Zn-LMFP/C and undoped LMFP/C underwent a two-phase reaction during the Mn3+/Mn2+ redox process, followed by a solid-solution reaction during the Fe3+/Fe2+ redox process-it significantly improved the reaction kinetics.69 This suggests that the primary role of certain dopants is not to change the fundamental thermodynamic pathway but to facilitate faster Li+ diffusion, thereby reducing the kinetic hindrances that exacerbate phase separation and polarization.

In situ X-ray diffraction (XRD) studies have been pivotal in elucidating how cation doping alters the de/intercalation mechanisms. For instance, Nb5+ doping was shown to induce a hybrid mechanism. During charging, the extraction of Li+ from LiMn0.6Fe0.4PO4 leads to a two-phase transition, evidenced by the disappearance of lithiated-phase peaks and the emergence of new peaks corresponding to the delithiated Mn0.6Fe0.4Nb0.02PO4 phase. Crucially, at higher voltages (>4.2 V), the diffraction peaks of this delithiated phase shift continuously towards higher angles, indicating a solid-solution process within the Mn-based redox couple.67 A similar phenomenon was observed with Cu2+ doping, where the reversible transformation between LiMn0.6Fe0.39Cu0.01PO4 and Mn0.6Fe0.039Cu0.01PO4 was accompanied by continuous peak shifts, confirming that the doping element stabilizes the structure without disrupting the olivine framework and promotes a more continuous reaction.68 Besides, and as highlighted in section 5, the advances in in situ studies have shown that aliovalent metal doping into LMFP can significantly affect the structural evolution of LMFP during lithiation and delithiation, especially when combined with strategies such as carbon coating and Fe-vacancy engineering. These approaches help narrow the two-phase reaction region, leading to more uniform phase transitions, improved reaction stability, and enhanced electrochemical performance.70,71

The optimization of LMFP cathodes through single or multi-metal doping is profoundly effective due to its ability to modify the fundamental phase transition mechanisms. The overarching trend is a doping-induced shift from a classical, sluggish two-phase reaction towards a faster, more homogeneous solid-solution-type process. This modulation mitigates structural strain and reduces polarization. More importantly, the primary benefit often lies in the dramatic enhancement of Li+ diffusion kinetics, which addresses the kinetic origins of voltage decay by ensuring faster transport and minimizing non-equilibrium concentration gradients. Therefore, the ideal doping strategy should aim not only to stabilize the crystal structure but also to create a favorable ionic transport landscape, ensuring that the phase transitions proceed in a more reversible and kinetically unhindered manner.67–69

In the next sections, the main features related to the synthesis, structure, and commercialization of LMFP are discussed, including potential advantages, performance characteristics, and strategies for optimizing its properties to meet the demands of high-energy applications.

4. General synthesis protocols of multimetal LMFPs

The synthesis process is crucial for optimizing the performance of multimetal LMFP cathodes, with each method offering unique benefits and factors to consider for effective production. Fig. 5 shows the general synthesis methods for multimetal LMFP cathodes. A well-controlled synthesis is essential to enhance ion conductivity, improve structural stability, and achieve better cycling performance and energy density. In addition, factors such as precursor selection, synthesis temperature, and applied methodology significantly influence the crystallinity, morphology, and electrochemical properties of the prepared cathode materials.
image file: d5ta03194j-f5.tif
Fig. 5 Overview of synthesis routes widely reported in the literature for the multimetal LMFP-based cathodes.

The solid-state method is currently the most widely used for the obtention of multimetal LMFP materials (Fig. 5). This method is favored for its simplicity, high yield, and suitability for large-scale industrial production. However, it requires high temperatures and extended processing times, and challenges remain in achieving uniform particle size, distribution, and morphology.7

Fig. 6 shows an illustration that represents a simple process of solid-state reaction. The process begins with the mixing and milling of solid compounds containing lithium, manganese, iron, and phosphorus, i.e., the raw materials (in a solid powder form), in order to obtain atomic-level uniformity. Then, the designed precursor material undergoes two stages of a high-temperature heating treatment in a controlled inert atmosphere. The precursor is pre-sintered at temperatures between 300 and 500 °C to decompose organic matter and remove structural water and then sintered at higher temperatures (600–850 °C) to promote the growth of olivine-type crystals and enhance their crystallinity.7 Typically, to facilitate the milling process, the reagents are suspended in a liquid medium, such as ethanol, which is subsequently evaporated.72,73 In this case, this first step of the process may be considered as a mechanochemical liquid-phase process, in which the mechanical activation is performed in a liquid medium. This method involves the mixture of solid reactants with a small amount of liquid, subjecting the mixture to mechanical forces, such as milling or grinding.74 The use of mechanical forces in a liquid medium facilitates interactions and introduces better control of the reaction conditions. In addition, the incorporation of organic carbon additives, such as sucrose, tends to improve the conductivity of LMFP cathodes. Mechanical or high-energy ball milling ensures the effective comminution and uniform dispersion of the carbon precursor, enabling the formation of a continuous, conductive carbon network on particle surfaces during high-temperature carbonization.12 This carbon coating can boost electrochemical performance by mitigating particle agglomeration during milling, refining microstructural homogeneity, and facilitating more efficient charge transport.75 V3+-doped LiMn0.8Fe0.2PO4@C were synthesized through a solid-state high-temperature carbothermal reduction reaction 76. In the procedure, the raw materials were weighed according to the stoichiometric ratio and then uniformly mixed through mechanical activation, with sucrose as a carbon source. Acetic acid was added during the activation to promote the oxidation of Fe and Mn to their +2 valence states. The mixture was ball-milled with a moderate amount of H2O in a planetary ball mill using zirconia balls as a mill medium. After milling, the wet mixture was dried, resulting in a precursor, which was then calcined in an argon atmosphere, first at 450 °C for 8 hours, then heated to 650 °C for 8 hours, and finally cooled to room temperature. A resulting composite showed discharge capacities of 115.9 mAh g−1 at 5.0C and high-capacity retention of 93.4% after 100 cycles at 1.0C.76 The enhanced rate performance of the composites was attributed to the combined benefits of nanocrystalline particle incorporation and V3+-doping. In addition, the mechanical liquid-phase activation approach accelerated the reaction between LiH2PO4 and metal powders, while efficiently generating nano-sized, highly reactive precursors.


image file: d5ta03194j-f6.tif
Fig. 6 Illustration of the solid-state reaction process.

Recently, the ball-milling technique combined with high-temperature solid-state reaction has been successfully employed to synthesize LiMn0.6Fe0.4−xNbxPO4/C cathode materials.71 LiH2PO4, Li2CO3, MnCO3, FePO4, Nb2O5, and 3 wt% sucrose were dispersed in anhydrous ethanol and subjected to high-speed ball-milling (600 rpm, 8 h) to induce liquid-assisted mechanochemical activation, ensuring intimate mixing, particle size reduction, and enhanced precursor reactivity. The resulting mixture was vacuum-dried and preheated at 350 °C for 3 h in Ar. Subsequently, the precursor was combined with an additional 2 wt% sucrose, ball-milled at 300 rpm for 8 h to achieve a uniform secondary carbon coating, vacuum-dried, and calcined at 650 °C for 10 h in Ar (5 °C min−1), yielding fine, well-dispersed carbon-coated particles. This dual modification strategy, combining Nb5+ doping with a second-time carbon coating, significantly enhanced conductivity, structural stability, and electrochemical performance.71

The solvothermal process has emerged as the second most commonly used synthesis technique for LMFPs. In this process, a solution of the raw materials in the determined solvent is put in a stainless-steel autoclave reactor and kept under high pressure and temperature for a period of time. As a result, the morphology, particle size, and purity can be more precisely controlled.77 Indeed, this process enables the production of materials with reduced particle sizes, uniform distributions, well-developed crystallinity, and reduced impurities. However, materials obtained through this approach frequently present limited electrochemical performance as a result of atomic disorder within the crystal lattice, which requires high-temperature annealing to restore structural alignment and improve their electrochemical performance. In addition, the solvothermal method faces challenges such as low reproducibility and the high cost of specialized reactors, limiting their scalability and feasibility for large-scale industrial production.7

A solvothermal synthesis was employed to produce Nb- and Sb-doped LMFP/C cathode materials,78,79 where stoichiometric amounts of MnSO4·4H2O, FeSO4·7H2O, and the respective dopant (C10H5NbO20 for Nb or C6H9O6Sb for Sb) were dissolved in water, mixed with ethylene glycol, and combined with LiOH·H2O and H2PO4. The mixtures were then heated at 180 °C for 5 hours in a Teflon-lined autoclave, followed by washing, drying, milling with glucose, and calcination at 700 °C in an argon atmosphere to obtain the final Nb- and Sb-doped Li(Fe0.6Mn0.4)1−xPO4/C cathode materials with varying dopant concentrations (x = 0, 0.01, 0.02, 0.03, 0.04, 0.05).78,79 The Nb- and Sb-doped samples exhibited enhanced electrochemical performance compared to undoped LMFP/C, with the LMFP/C–Nb0.02 and LMFP/C–Sb0.04 samples showing superior cycling stability and rate capability.78,79 In addition, doping improved lithium-ion diffusion and crystal structure stability, with Nb doping enhancing the metal–oxygen framework and Sb doping mitigating the Jahn–Teller effect.79Fig. 7 shows a schematic representation of the synthesis used to obtain the Nb-doped LMFP/C cathode materials, effectively illustrating the key steps of a general solvothermal synthesis.


image file: d5ta03194j-f7.tif
Fig. 7 Schematic representation of the synthesis of Nb-doped LMFP/C cathode materials. Reproduced with permission from ref. 78. Copyright © 2024 Published by Elsevier Ltd.

The hydrothermal method is similar to the solvothermal method. The basic difference between solvothermal and hydrothermal is due to the type of solvent used: organic solvent for solvothermal and water for hydrothermal. It is evident from the literature that the solvothermal method is preferred over the hydrothermal method for LMFP synthesis. This preference can be attributed to the greater flexibility offered by solvothermal processes, which utilize a range of organic solvents. These solvents allow for better control over reaction conditions, enhancing material properties and improving solubility. In contrast, hydrothermal methods are limited to water, thereby restricting the types of materials that can be synthesized. Furthermore, parameters such as temperature and pressure control, along with pH adjustment, are likely considered in the choice of synthesis method.

LiMn0.79Fe0.2Mg0.01PO4/C composite materials were synthesized via a hydrothermal method, where metal salts (MnSO4·H2O, FeSO4·7H2O, and MgSO4·7H2O) were dissolved in deionized water and purged with Ar to remove dissolved oxygen, then mixed with a colloidal solution of Li3PO4.80 This mixture was transferred to a high-pressure hydrothermal reactor, where it was heated at 180 °C for 5 hours, followed by filtration, washing, and drying. The dried precursor was ball-milled with glucose and ascorbic acid in ethanol, then heated at 720 °C in an Ar atmosphere to yield the carbon-coated LiMn0.79Fe0.2Mg0.01PO4/C material.80 The hydrothermal process described above is shown in Fig. 8. In this study, the influence of different sources of Li3PO4 on the performance of synthesized Mg trace LMFP cathodes was investigated, demonstrating that the choice of Li3PO4 precursor and the synthesis conditions directly dictated the particle morphology, size, and electrochemical performance. Li3PO4 with smaller crystal grains dissolves more readily due to its higher surface area, promoting nucleation of Mg-doped LMFP and resulting in smaller, more uniform particles. Conversely, larger Li3PO4 grains limit further LMFP formation, leading to residual Li3PO4 impurities. Thus, Mg-doped LMFP, prepared from well-dispersed, low-aggregation Li3PO4, exhibited the smallest particle size and uniform distribution. During co-sintering with glucose, the LMFP precursor enabled the formation of a uniform carbon coating and a larger specific surface area, which enhanced electron transport and lithium-ion diffusion. Consequently, this material exhibited superior lithium-ion kinetics, high-rate capability, and excellent cycling stability. These results clearly illustrate that controlling the hydrothermal synthesis, particularly the precursor characteristics, is essential for tailoring the structural and electrochemical properties of LMFP cathodes.80


image file: d5ta03194j-f8.tif
Fig. 8 Schematic representation of the hydrothermal synthesis process Reproduced with permission from ref. 80. Copyright © 2024 Elsevier Ltd and Techna Group S.r.l. All rights reserved.

These effects on morphology and size, and consequently on electrochemical properties, are indeed expected, but they are not solely attributable to the intrinsic nature of the precursor properties. Dai et al.81 explored different synthesis variables, such as the anionic species of the Fe precursor (FeCl2 (Cl) and FeSO4 (SO42−)) and the sequence of precursor addition: P + M + L sequence (LiOH solution dropwise introduced into the mixture of Li3PO4 and transition metal salts) or P + L + M sequence (Mn2+ and Fe2+ mixture slowly added into the mixture of Li3PO4 and LiOH). The proposed solvothermal synthesis proved highly effective in controlling the morphology of LMFP nanoparticles by adjusting the anion species and the precursor feeding sequence. The P + M + L sequence favored the formation of spindle-like nanostructures with reduced size and lower agglomeration, particularly when FeSO4 was used as the iron source, leading to more uniform particles compared to FeCl2. In contrast, the P + L + M sequence predominantly yielded nanoplates oriented along the ac plane, with lengths below 100 nm and thicknesses of about 20–30 nm. These morphological differences had a direct impact on electrochemical performance, as the particles obtained via the P + M + L sequence exhibited a reversible capacity of 129.7 mAh g−1 at 0.1C and excellent rate, retaining 77% of their capacity at 1C.

Sol–gel and co-precipitation methods have been less commonly used for the synthesis of multielemental LMFPs. The sol–gel process82,83 begins with the preparation of a colloidal solution (known as a sol) by dissolving metal salts or metal alkoxides in a given solvent, in which phosphoric acid (H3PO4) is added as the source of phosphate. The metal precursors in solution undergo hydrolysis, forming hydroxyl groups on the metal centers, and then they interact with the H3PO4 to form a metal-phosphate network through a condensation reaction. As the metal ions and phosphate ions interact, they form a gel-like structure, which is essentially a three-dimensional network of metal-phosphate bonds. To aid in this process, chelating agents such as citric acid are often used, which help stabilize the metal ions, provide a carbon source, and control gel formation.84

As condensation progresses, the sol transitions into a gel. The gel is then aged to further strengthen its metal-phosphate framework and dried to remove excess solvent. The resultant material is calcined to remove organic residues and enhance the crystallinity and stability of the metal-phosphate structure. This method requires low synthesis temperature and provides excellent control over the reaction process, as well as the composition and structure of the final product.85 Some studies12,38 have highlighted that sol–gel processing allows precise tailoring of LMFP structure and morphology, directly linking synthesis conditions to electrochemical performance. By adjusting parameters such as precursor composition, carbon sources, and gelation conditions, several works have obtained nanoscale particles with uniform morphology, porous structures, and consistent carbon coatings, enhancing lithium-ion transport and electronic conductivity. Additives such as glucose, citric acid, or acetylene black further improve particle dispersion and form conductive networks, resulting in higher discharge capacity, rate performance, and cycling stability. However, according to Zhang et al.,7 the sol–gel method remains predominantly limited to laboratory research due to extended drying and processing times, complex procedural steps, high raw material costs, and potential hazards to human health and the environment, which hinder its applicability for large-scale industrial production. Fig. 9 provides a schematic representation of the steps involved in the sol–gel synthesis procedure. Qiao et al.86 synthesized Li(Fe0.65Mn0.35)0.98Mg0.02PO4/C, Li0.98Na0.02Fe0.65Mn0.35PO4/C, and Li0.98Na0.02(Fe0.65Mn0.35)1−xMgxPO4/C (x = 0.01, 0.02, 0.03, 0.05) composites via a sol–gel process modified by Na+ and Mg2+ doping. In this approach, stoichiometric amounts of manganese source (Mn(CH3COO)2·4H2O), carbon source (citric acid), phosphorus source (NH4H2PO4), iron source (FeCl2·4H2O), and lithium source (CH3COOLi·2H2O) were sequentially dissolved in deionized water. The solution was then supplemented with sodium (CH3COONa·3H2O) and magnesium (Mg(CH3COO)2·4H2O) sources, followed by gel formation, drying, preheating, and calcination under a nitrogen atmosphere. The study demonstrated that doping with Na+ and Mg2+ enhanced electrochemical performance by reducing impedance, improving specific capacity, and increasing cycle stability.86


image file: d5ta03194j-f9.tif
Fig. 9 Schematic illustration of sol–gel synthesis.

The co-precipitation method involves preparing a solution of metal salts, to which a precipitating agent is added to initiate the precipitation of the desired compounds. The resulting precipitate is then filtered, washed, dried at moderate temperatures, and calcined at high temperatures to promote the formation of a crystalline material with the desired composition and structure. The co-precipitation method is simple and scalable, allowing precise control over the composition, phase, and morphology of active particles, which is crucial for optimizing electrochemical performance. However, achieving monodisperse particles with regular morphology requires careful control of reaction conditions, including pH, reagent concentrations, temperature, stirring rate, and the use of chelating agents. Furthermore, understanding the nucleation and growth processes is essential to predict and control the synthesis of crystalline precursor particles.87 Indeed, the importance of carefully selecting and controlling the synthesis route to optimize particle size, morphology, and porosity, thereby directly enhancing the electrochemical properties of the cathodes. Li et al.88 showed by preparing three Ni, Fe co-doped LiMnPO4/C composite materials (LMP) using co-precipitation and solvothermal methods that the synthesis method critically influenced the morphology, microstructure, and electrochemical performance of the LMP. Co-precipitation produced irregular, large, and agglomerated particles of LMP, whereas solvothermal synthesis yielded more uniform rod-like particles. Glucose-assisted solvothermal synthesis further improved particle uniformity, producing small, spindle-shaped, porous structures, which shortened Li+ diffusion paths and enhanced electrolyte contact, resulting in higher discharge capacities (121.4 mAh g−1 at 0.1C), superior rate performance (53.7 mAh g−1 at 2C), and improved cycling stability (91% retention after 100 cycles at 1C).

For instance, LiMn1/2Fe1/4Ni1/4PO4 precursor was synthesized via a co-precipitation method, with Fe and Ni co-doping to enhance the electrochemical properties.89 The schematic representation of the co-precipitation synthesis is shown in Fig. 10. The synthesis of the precursor involved dissolving MnSO4·H2O, FeSO4·7H2O, and NiSO4·6H2O in water according to a 2[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1 molar ratio. The Fe2+ ions were oxidized to Fe3+ with hydrogen peroxide to facilitate phosphate nucleation, while Mn2+ and Ni2+ remained in their divalent states. The resulting mixture was then added dropwise, together with ammonia, into a combined solution of H3PO4 and NaOH, precipitating the phosphate precursor at pH 4. The precipitate was filtered, washed, dried, and then dispersed in a small amount of anhydrous ethanol, together with Li2CO3, rock sugar, and NH4H2PO4, to synthesize the cathode material. During the calcination at 650 °C under N2 with rock sugar, Fe3+ was reduced back to Fe2+, enabling its incorporation into the structure of LiMn1/2Fe1/4Ni1/4PO4C. The resulting LiMn1/2Fe1/4Ni1/4PO4C material exhibited a significantly higher initial charge–discharge capacity and excellent cyclic stability compared to LiMnPO4/C, demonstrating the effectiveness of Fe–Ni doping in improving the electrochemical performance of the material.89


image file: d5ta03194j-f10.tif
Fig. 10 Co-precipitation method for synthesis of the LiMn1/2Fe1/4Ni1/4PO4 precursor, followed by milling and calcination for obtention of the LiMn1/2Fe1/4Ni1/4PO4 cathode material. Reproduced with permission from ref. 89. Copyright © 2021, The Author(s), under exclusive license to Springer-Verlag GmbH Germany, part of Springer Nature.

Techniques based on physical processes, such as freeze-drying and spray-drying, are commonly used to complement the methods described above.40,90–95 These techniques play a crucial role in material synthesis, particularly in controlling particle size, homogeneity, and processing conditions. Freeze-drying is a method that removes water or other solvents from a sample by sublimation.96,97 When pressure and temperature are adjusted to the triple point, even a slight increase in temperature will cause the solvent to sublimate directly.77,96,97 This process avoids the need for high temperatures and ensures that delicate structures remain intact. In addition, it offers several advantages, including the ability to achieve lower calcination temperatures and ensure homogeneous distribution of reactants.77,96,97 In the spray-drying process, solutions of raw materials (commonly metal salts) are atomized into fine droplets and rapidly dried to obtain dry powder known as a precursor.98,99 After obtaining the precursor, it is typically subjected to an annealing or calcination step at higher temperatures to form the final product.99 As well as freeze-drying, spray-drying provides specific advantages in terms of control over particle size, uniformity, and temperature, making it useful in various applications in material synthesis and processing.98–100

It is important to emphasize that all of the methods described here use calcination as the final step. This high-temperature process facilitates the development of the desired material properties, including crystallinity, particle size, and phase formation. In addition, the different parameters of the synthesis, such as precursors, sintering temperature, solvent, and time, usually have significant effects on the morphology, chemical composition, and electrochemical performance of the cathode materials. Therefore, the choice of the process and, consequently, its parameters must be carefully considered to produce adequate cathode materials.

5. Multimetal LMFPs-based electrodes for LIBs

As previously described, the design of multimetal LMFP cathodes encompasses different approaches, including (i) doping with single or multiple elements, (ii) incorporating additional primary constituents or structural components, and/or (iii) enhancing the entropy of electrode materials. These approaches are particularly categorized based on the concept of doping, where the dopant concentration is typically limited to 5% by mass or atomic percentage. Beyond this threshold, the added metal ion functions as a structural component rather than a dopant. In fact, many authors in the literature mistakenly use the term doping when referring to the incorporation of a new metal ion at concentrations as high as 10 mol% or mass%. However, under these conditions, the introduced element should be regarded as a structural component of the composition rather than a dopant. Besides, incorporating large amounts of new elements into the LMFP structure is expected to induce major structural disorder. In fact, when incorporating new elements via isomorphic substitution, the presence of ions with different ionic radii can induce lattice distortion and increase structural disorder. Additionally, recent advancements in multimetal materials reveal that the design of medium- and high-entropy materials (MEMs and HEMs) can significantly improve their electrochemical and electrocatalytic properties.101,102 Indeed, the high-entropy concept has emerged as a promising strategy and a hot topic in the design of novel active materials for lithium batteries.103 According to statistical thermodynamics, the link between entropy and the composition of a system can be obtained from Boltzmann's equation. The molar configurational entropy of mixing (ΔSconfig) for a crystalline phase containing multiple elements can be determined using eqn (1), where R is the universal gas constant (8.314 J K−1 mol−1) and x represents the molar fraction of each element.104
 
image file: d5ta03194j-t1.tif(1)

Based on calculations obtained by the aforementioned equation, materials are classified as “low-entropy” (LEMs) when their Sconfig is below 1.0 R, as “medium-entropy” when Sconfig lies between 1.0 R and 1.5 R, and as “high-entropy” when the value is greater than 1.5 R.104–106 Accordingly, the library of MEMs and HEMs has increasingly expanded to include olivine-type lithium metal phosphates,107 a subject that will be examined in the following subsections.

This section explores the design strategies for multimetal LMFP electrode materials, focusing on the incorporation of foreign elements into different sites and the specific types of metal ions introduced. Interestingly, the progress shows that significant experimental advances highlight the use of isovalent ions (M2+), more specifically, Mg2+ and Ni2+. These two cations were reported by approximately 40% of the prepared multimetal LMFPs. On the other hand, among the aliovalent ions, Nb5+ and V3+ were the most studied (Fig. 11). In fact, the incorporation of at least one of these elements represents more than 50% of the compositions reported in the literature.


image file: d5ta03194j-f11.tif
Fig. 11 Percentage of the main metal ions incorporated in multimetal LMFP cathode materials reported in the literature, with some materials incorporating one or two elements within the same active cathode structure. The LMFP structural model shown at the core of the scheme has been adapted with permission from ref. 12. Copyright © 2025 Wiley-VCH GmbH.

As outlined in the subsequent sections, all multielemetal compositions achieving capacity retention of 90% or higher after 200 or more cycles incorporate some form of carbon-based material (Fig. 12). In the carbon-coating approach, LMFP particles are surrounded by a carbon layer, which improves electrical conductivity, rate capability, and cycling stability, ultimately enhancing their overall electrochemical performance.9,108 In fact, surface carbon coating is one of the most widely employed modification techniques for LMFP (referred to as LMFP/C), addressing key challenges such as low electronic conductivity, Mn dissolution, structural instability, and effectively creating a conductive network that enhances electron transport efficiency.9,40 The following highlights the strategies, trends, and key findings, with a focus on the synergistic effects arising from incorporating various isovalent and/or aliovalent metal ions into the structure of carbon-coating-based LMFP cathodes, wherein the overall performance exceeds the sum of the individual contributions.


image file: d5ta03194j-f12.tif
Fig. 12 Multimetal LMFP cathode materials with capacity retention ≥90% after 200 or more charge–discharge cycles. Maximum capacity discharge at 0.1C and or 0.2C. LiMn0.59Fe0.4Na0.005Co0.005PO4/C ref. 90; LiFe0.491Mn0.499Nb0.01PO4/C ref. 94; LiMn0.59Fe0.4Mg0.005Ni0.005PO4/C ref. 40; LiMn0.6Fe0.3850□0.0075Ti0.0075PO4/C ref. 70; LiMn0.6Fe0.38Nb0.02PO4/C ref. 71; LiMn0.8Fe0.155V0.03□0.015PO4/C ref. 109; LiMn0.6Fe0.395La0.005PO4/C ref. 110; Li(Mn0.5Fe0.5)0.98V0.02PO4/MXene + C ref. 111; Li0.98Mn0.6Fe0.4Nb0.02PO4/C ref. 67; Li0.98Fe0.6Mn0.4Nb0.02PO4/C ref. 78; Ti/V–LiMn0.6Fe0.4PO4/C ref. 112; 1Nb3Mg–LiMn0.5Fe0.5PO4/C ref. 92; LiMn0.6Fe0.39Mg0.01PO4/C ref. 113; Li(Fe0.6Mn0.4)0.96Sb0.04PO4/C ref. 79; LiMn0.79Fe0.2Mg0.01PO4/C ref. 80; LiNa0.03Mn0.8Fe0.2PO4/C ref. 114; LiMn0.8Fe0.19Ni0.01PO4/C ref. 115; Li0.97Mg0.015Mn0.8Fe0.2PO4/C ref. 116; LiMn0.6Fe0.39Cu0.01PO4/C ref. 68; Li(Fe0.5Mn0.5)1−3xMoxPO4/(x = 0.01)/C ref. 117; LiFe0.47Mn0.5Ca0.03PO4/C ref. 118.

5.1. Elemental doping strategy

The incorporation of elements by doping has been explored as an effective strategy to improve both electronic and ionic conductivity.10 In fact, elemental doping is essential for boosting the internal conductivity of LMFP-based electrode materials.8 In general, multi-element doping in different sites for LFP or LMFP, given their structural similarities, follows two main groups: single-site doping (Li-site or M-site) and dual-site doping (Li-site and M-site).40,49,119 Besides, choosing the optimal doping ratio is vital to balance the benefits of Fe and Mn elements while preventing potential deviations or degradation in electrochemical performance, particularly from an excess of Mn.8
5.1.1. Cation doping with M-site substitution into LMFP. Fig. 11 highlights materials demonstrating capacity retention of 90% or higher after 200 or more cycles. Analysis of these results indicates that cation doping through M-site substitution in LMFP emerges as a key strategy for developing multi-elemental LMFP with enhanced stability during charge–discharge cycling and improved capacity at high rates. Interestingly, among the primary inactive metal ion dopants, Mg2+ and Nb5+ have been the most widely utilized, yielding significant performance improvements. Besides, Ni2+-doping has been explored in various studies, showcasing significant enhancements in performance.

Among divalent ions, the electrochemically inactive isovalent Mg2+ emerges as a promising candidate due to its cost-effectiveness, environmental advantages, and its ability to promote solid-solution behavior. By substituting at Fe2+/Mn2+ sites, Mg2+ can mitigate the effects of Jahn–Teller distortion and enhance structural stability in Mn-rich compositions. Notably, among cathode materials exhibiting capacity retention of 90% or higher after 200 or more cycles, 33% are Mg-containing LMFP compositions. The ionic radius of Mg2+ incorporated into the LMFP crystal lattice is smaller than that of Mn2+ and Fe2+, leading to two main structural modifications and reaction mechanisms. Firstly, there is a shortening of the bond length within the olivine MO6 (M = Mn, Fe, Mg) octahedral structure. Conversely, the Li–O bond length in the LiO6 octahedron increases. This elongation of the Li–O bond facilitates lithium-ion migration by widening the diffusion channel, thereby enhancing electrochemical reactions with improved rate capability and efficiency in multicomponent olivine structures.7,113,120

Conversely, the formation of carbon-coated LMFP is a critical approach for achieving high-performance multimetal LMFP-based cathodes. As illustrated in Fig. 12, 100% of the materials are integrated with some form of carbon-based compound.

A typical example of the use of these approaches was employed by Duan et al.,121 who reported that an eco-efficient mechano-chemical liquid-phase activation method was employed to synthesize Mn-rich Li(Mn0.9Fe0.1)1−xMgxPO4/C nanocomposites, optimizing synthesis temperature and Mg2+-doping ratio to enhance electrochemical performance. This method successfully produced nano-sized precursors, which were converted into Li(Mn0.9Fe0.1)1−xMgxPO4/C composites via one-step calcination. Among the samples, Li(Mn0.9Fe0.1)0.95Mg0.05PO4/C, synthesized at 700 °C with ∼3 wt% carbon additive, exhibited the best properties, delivering a specific discharge capacity of 153 mAh g−1 at 0.1C and excellent cycling stability with 100% capacity retention after 100 cycles. Mg-doping reduced the lattice parameters, mitigating structural strain during lithium insertion/extraction and improving conductivity. However, excessive Mg-doping (e.g., 7%) led to unfavorable results, such as reduced electrochemical activity and higher resistance. Thus, 5% Mg2+ doping was determined to be optimal, balancing structural stability and electrochemical performance. These findings highlight the potential of Li(Mn0.9Fe0.1)1−xMgxPO4/C as a competitive cathode material for advanced LIBs. Similarly, Hu and colleagues122 developed an Mn-rich LiMn0.9Fe0.09Mg0.01PO4/C material that exhibits impressive electrochemical performance, including a high discharge capacity of 160 mAh g−1 at 0.1C and remarkable cycling stability at elevated temperatures. Even under a high discharge rate of 1C at 50 °C, the material achieves a capacity of approximately 135 mAh g−1 with no capacity degradation after 120 cycles. This performance highlights the effective suppression of Mn dissolution in the electrolyte, ensuring enhanced structural and electrochemical stability of the electrode material.

Although the aforementioned studies demonstrate promising results, particularly in terms of high-capacity retention, further investigations into long-term cycling stability are essential. Extensive testing over a greater number of charge–discharge cycles is required to fully validate the advantages of Mg-doping, especially in the context of large-scale prototype applications. From this perspective, other Mg-doped LMFP cathodes have been developed, including Mn-rich compositions such as LiMn0.79Fe0.2Mg0.01PO4/C80 and LiMn0.6Fe0.39Mg0.01PO4/C.113 These materials demonstrated impressive cycling stability, achieving capacity retentions of 97.07% at 1C after 200 cycles and 96.2% at 1C after 300 cycles, respectively.

Still considering 1% Mg-doping in Mn-rich LMFP, Kisu et al.123 developed spherical LiMn0.792Fe0.198Mg0.010PO4 nanocrystals (10–40 nm) embedded within the interstices of supergrowth single-walled carbon nanotubes (SGCNTs) using an innovative in situ processing method called “ultra-centrifuging treatment” followed by brief heat treatment (Fig. 13a and b). Transmission electron microscopy (TEM) images revealed a direct attachment of these single-crystalline nanospheres to the highly conductive SGCNTs, forming a well-dispersed composite (Fig. 13c and d). This approach resulted in a 200% enhancement in Li+ diffusivity and a 50% reduction in charge transfer resistance, enabling ultrafast Li+ migration. These “nano–nano” LiMn0.792Fe0.198Mg0.010PO4/SGCNT composites overcame the limitations of one-dimensional diffusion. It delivered exceptional electrochemical performance, achieving a capacity of ∼54 mAh g−1 per composite (equivalent to 77 mAh g−1 per pure LiMn0.792Fe0.198Mg0.010PO4) at a high rate of 50 C (Fig. 13e). Moreover, the material demonstrated outstanding cycling stability, retaining 84% of its initial capacity after 3000 cycles (Fig. 13f). Interestingly, while the capacity retention is below 90%, the ability to sustain performance over 3000 cycles and operate at high power densities presents a promising opportunity for advancing high-power density devices.


image file: d5ta03194j-f13.tif
Fig. 13 (a) TEM and (b) HRTEM images of LiMn0.792Fe0.198Mg0.010PO4/SGCNT composites and (c) their corresponding schematic illustration, with (d) 3D image of a LiMn0.792Fe0.198Mg0.010PO4 nanocrystal with the direction of Li+ diffusion. (e) Discharge capacity as a function of current density for (A, blue) LiMn0.8Fe0.2PO4/SGCNTs and (B, red) LiMn0.792Fe0.198Mg0.010PO4/SGCNTs. Dashed lines represent the capacity of the composites, while solid lines indicate the capacity of the pure LiMn0.8(1−z)Fe0.2(1−z)MgzPO4 (z = 0 or 0.01), excluding the contribution from SGCNTs in the composite. (f) Cycle-life testing of the LiMn0.792Fe0.198Mg0.010PO4/SGCNT composite over 3000 cycles. The inset highlights the charge–discharge profiles at the 1st, 100th, 500th, 1000th, 2000th, and 3000th cycles. Reproduced with permission from ref. 123. Copyrights © The Royal Society of Chemistry 2014.

Conversely, other isovalent ions have also been reported. Among them, Ni2+-doping has been explored in various studies, showcasing significant enhancements in performance. In fact, various doping percentages have been investigated, particularly in Mn-rich LMFP compositions (Table 1). One such study utilized a Ni-doping strategy to optimize the morphology of Mn-rich LiMn0.8Fe0.2PO4 nanoparticles, promoting a preferred growth orientation that enhanced electrical and ionic conductivity as well as electrode kinetics.124 In further detail, Ni-D0 (LiMn0.8Fe0.2PO4 pristine) exhibits a mix of flake and ellipsoidal shapes with larger particles (∼53 × 120 nm) (Fig. 14a), while Ni-D5 particles (5 mol% Ni, LiMn0.8Fe0.15Ni0.05PO4) are ellipsoidal with an average size of 58 nm (Fig. 14c). Ni-D5 features a thin, uniform graphitized carbon coating (∼1 nm, Fig. 14d), whereas Ni-D0 has a thicker, amorphous layer (∼3 nm, Fig. 14b). In more detail, the Ni doping promotes a highly graphitized carbon layer, forming a conductive network that enhances electrical conductivity and Li+ transport, thereby improving electrochemical performance (Fig. 14e). In fact, compared to the pristine sample, the LiMn0.8Fe0.15Ni0.05PO4@C sample shows significant improvements, with electrical conductivity of 9.57 × 10−2 S cm−1 (Fig. 14f). Additionally, Ni doping induces the grains to grow along the (010) plane of LiMnPO4, shortening Li+ diffusion paths and facilitating efficient ion transport. Furthermore, incorporating 5 mol% Ni into LiMn0.8Fe0.2PO4@C reveals a typical two-phase transition characteristic of the olivine LiMnPO4 structure with a solid-solution reaction, which ensures exceptional structural stability and reversibility.124

Table 1 Electrochemical performance of multimetal LMFP-based electrodes for LIBs
Material Coating material Metal doping Synthesis methods Capacity retention (%) First discharge capacity (mAh g−1) WVW (V vs. Li/Li+) Ref.
a Values were taken from images or graphs. b (SGCNTs) = interstices of supergrowth (single-walled) carbon nanotubes.
LiMn0.9Fe0.09Mg0.01PO4 C Mg Solid-state 141.0@0.1C 2.0–4.5 130
LiMn0.8Fe0.19Mg0.01PO4 C Mg Solid-state 138.5@1C 2.0–4.5 131
Li(Mn0.85 Fe0.15)0.92Ti0.08PO4 C Ti Solid-state 100% after 50 cycles@1C 144.4@1C 2.0–4.5 132
LiNa0.03Mn0.8Fe0.2PO4 C Na Solvothermal 96.65% after 200 cycles@0.5C 138.0@0.1Ca 2.5–4.5 114
Li0.98Na0.02(Fe0.65Mn0.35)0.97Mg0.03PO4 C Na/Mg Sol–gel 96.2% after 40 cycles@0.1C 147.7@0.1C 2.5–4.5 86
LiFe0.4Mn0.595Cr0.005PO4 C Cr Solid-state 99.2% after 50 cycles@0.1C 164.0@1C 3.5–4.1 133
Li(Mn0.9Fe0.10)0.95Mg0.05PO4 C Mg Mechanochemical liquid-phase activation technique 98.1% after 100 cycles@0.1C 153.0@0.1C 2.5–4.1 121
LiMn0.8Fe0.19Ni0.01PO4 C Ni Solvothermal 94.1% after 200 cycles@0.5C 163.0@0.1Ca 2.0–4.5 115
LiMn0.792Fe0.198Mg0.01PO4 SGCNTb Mg Sol–gel 84.0% after 3000 cycles@1C 119.0@1C 2.5–4.5 123
LiFe0.95Ni0.02Mn0.03PO4 C Ni Solid-state 98.6% after 100 cycles@1C 145.4@1C 2.3–4.2 134
LiMn0.9(FeZn)0.05PO4 C Zn Solid-state 96.7% after 100 cycles@1C 153.0@0.1C 2.5–4.5 135
LiMn0.8Fe0.15Ni0.05PO4 C Ni Solvothermal 89.45% 100 cycles@0.1C 144.0@0.1C 2.5–4.5 124
Li0.97Mg0.015Mn0.8Fe0.2PO4 C Mg Solvothermal 93.5% after 200 cycles@0.5C 156.9@0.1C 2.0–4.5 116
(1−x)Li0.8Mn0.8Fe0.2PO4xLi3V2(PO4)3 LiVPO4 + C V Solid-state 93.4% after 100 cycles@1C 155.4@0.1C 2.5–4.5 76
LiMn0.79Fe0.2Mg0.01PO4 C Mg Hydrothermal 97.07% after 200 cycles@1C 134.7@0.1C 2.5–4.5 80
Li(Fe0.5Mn0.5)1−3xMoxPO4/(x = 0.01) C Mo Solvothermal 91.2% after 200 cycles@2C 153.2@0.1C 2.0–4.5 117
LiMn0.59Fe0.4Na0.005Co0.005PO4 C Co/Na Solvothermal/spray-dryer 97.1% after 1000 cycles@1C 150.0@0.1C 2.5–4.5 90
1Nb3Mg–LiMn0.5Fe0.5PO4 LiNbO3 + C Mg/Nb Co-precipitation/spray-dryer 99.0% after 300 cycles@0.5C ∼140@0.1C 2.0–4.4 92
Li0.98Fe0.6Mn0.4Nb0.02PO4 C Nb Solvothermal 94.15% after 500 cycles@5C 156.9@0.2C 2.0–4.5 78
LiMnxFey(V)1−xyPO4 (x = 1/3, y = 1/3) C V Solid-state 98.6% after 50 cycles@0.5C 161@0.05C 2.0–4.5 73
LiMn0.48Fe0.5Mg0.02PO4 C Mg Solid-state 97.7% after 100 cycles@1C 160.7@1C 2.0–4.3 136
LiMn0.8Fe0.155V0.03□0.015PO4 C V Solid-state 98.9% after 600 cycles@0.2C 155.0@0.1C 2.0–4.5 109
LiMn0.9Fe0.05Co0.05PO4 C Co Solid-state 145.0@0.05C 2.5–4.8 137
LiMn0.9Fe0.05Mg0.05PO4 C Mg Solid-state 99.0% after 30 cycles@0.2C 140.0@0.1C 2.5–4.5 138
LiMn0.9Fe0.09Mg0.01PO4 C Mg Solid-state 95.0% after 120 cycles@1C 160.0@0.1C 2.5–4.6 122
LiFe0.47Mn0.5Ca0.03PO4 C Ca Solid-state 91.0% after 200 cycles@1C 156.1@0.1C 2.0–4.5 118
LiFe0.4Mn0.595Cr0.005PO4 C Cr Solid-state 99.2% after 50 cycles@0.1C 164.0 @0.1C 2.2–4.5 133
LiMn0.8Fe0.1Ni0.1PO4 C Ni Solid-state/freeze-drying ∼100% after 100 cycles@C/25 110@C/25 2.5–4.5 93
LiMn0.8Fe0.1Co0.1PO4 C Co Solid-state/freeze-drying ∼100% after 100 cycles@C/25 80@C/25 2.5–4.5 93
LiMn0.8Fe0.1Cu0.1PO4 C Cu Solid-state/freeze-drying ∼100% after 100 cycles@C/25 57@C/25 2.5–4.5 93
LiFe0.491Mn0.499Nb0.01PO4 C Nb Hydrothermal/spray-dryer 95.4% after 1000 cycles@1C 152@0.1C 2.5–4.3 94
Li(Fe0.50Mn0.50)0.88V0.08PO4 V Solid-state ∼98.0@0.05C 2.0–4.7 139
LiMn0.9(FeZn)0.05PO4 C Zn Solid-state 96.7% after 100 cycles@1C 151.3@ 0.1C 2.0–4.5 135
LiFe0.4Mn0.595Cr0.005PO4 C Cr Solid-state 98.84% after 100 cycles@1C 163.6@0.1C 2.2–4.5 140
LiMn0.8Fe0.19Mg0.01PO4 C Mg Hydrothermal 137.0@0.1C 2.0–4.5 141
LiMn0.6Fe0.38Ni0.02PO4 C Ni Solid state/spray-dyer 98.3% after 100 cycles@1C 159.3@0.2C 2.5–4.5 95
LiMn0.6Fe0.39Mg0.01PO4 C Mg Solid-state 96.2% after 300 cycles@1C 159.6@0.2C 2.5–4.5 113
LiMn0.59Fe0.4Mg0.005Ni0.005PO4 C Mg/Ni Solvothermal/spray-dryer 95.0% after 1000@1C 152.0@0.1C 2.5–4.3 40
LiNa0.02Mn0.6Fe0.4PO4 C Na Carbothermal/spray-dryer 96.7% after 100 cycles@1C 153.1@0.2C 2.5–4.5 91
LiK0.03Mn0.05Fe0.95PO4 C K Solvothermal 97.0% after 100 cycles@1C 159.2@0.1C 2.5–4.2 127
Li(Fe0.6Mn0.4)0.96Sb0.04PO4 C Sb Solvothermal 93.2% after 300 cycles@5C 166.6@0.2C 2.0–4.5 79
LiFe0.48Mn0.48Mg0.04PO4 Mg 152.2@0.1C 2.7–4.4 49
LiMn0.8Fe0.19Mg0.01PO4 C Mg Solid-state 96.55% after 50 cycles 0.1C 145.0@0.1C 2.0–4.5 142
LiCo0.33Mn0.33Fe0.33PO4 Co Solid-state 140.0@0.05C 2.0–5.0 72
Li(Mn0.5Fe0.5)0.98V0.02PO4/MXene C V Solvothermal 96.19% after 500 cycles@1C 156.62@0.1C 2.5–4.5 111
Li0.995Nb0.005Mn0.85Fe0.15PO4 C Nb Solid-state 100% after 50 cycles@1C 166.0@0.12C 2.0–4.5 143
LiMn0.6Fe0.385□0.0075Ti0.0075PO4 C Ti Hydrothermal 92.55% after 1000 cycles@1C 139.7@0.2C 2.0–4.5 70
LiMn0.6Fe0.38Nb0.02PO4 C Nb High-temperature solid-phase 93.2% after 700 cycles@1C 158.3@0.1C 2.0–4.5 71
Li0.97Na0.03Mn0.8Fe0.2PO4 C Na Solvothermal 96.65% after 200 cycles@0.5C 141.7@0.05C 2.5–4.8a 38
LiMn0.6Fe0.395La0.005PO4 C La High-temperature solid-phase 96.45% after 600 cycles@1C 158.61@0.1C 2.0–4.5 110
LiMn0.5Fe0.49Zn0.01PO4 C Zn Solvothermal 92.4% after 100 cycles@1C 141.75 @0.2C 2.0–4.5 144
LiFe0.5Mn0.49Y0.01PO4 C Y Carbothermal reduction of solid-state 87.74% after 300 cycles@1C 160.0@0.2C 2.0–4.5 145
Ti/V–LiMn0.6Fe0.4PO4 C V/Ti Co-precipitation/solid-state 93.81% after 500 cycles@1C 161.9@0.1C 2.0–4.5 112
Li0.98Mn0.6Fe0.4Nb0.02PO4 C Nb Co-precipitation/solid-state 95.07% after cycles 500@1C 155.63@0.1C 2.0–4.5 67
LiMn0.84Fe0.15Mg0.01PO4 C Mg Solid-state 80.8% after 500 cycles@1C 152.1@0.1C 2.5–4.5 146
LiMn0.6Fe0.39Cu0.01PO4 C Cu Solvothermal/solid-state 92.5% after 200 cycles@1C 160.3@0.1C 2.5–4 0.3 68
Li(Mn0.35Fe0.35Co0.1Mg0.1Ca0.1)PO4 Co/Mg/Ca Mechanochemistry 78.0% after 20 cycles 0.1C 104.0@0.1C 2.0–4.9 107
Li(Mn0.2Fe0.2Co0.2Ni0.2Mg0.2)PO4 Co/Ni/Mg Mechanochemistry 60.0@0.1C 2.0–4.9 107



image file: d5ta03194j-f14.tif
Fig. 14 SEM and HRTEM images (and FFT patterns) of (a and b) Ni-D0 (pristine LMFP) and (c and d) Ni-D5. (e) Rate capability and (f) electrical conductivity measurements of the samples. Reproduced with permission from ref. 124 Copyrights © 2023 Elsevier B.V. All rights reserved.

Similarly, Tian et al.95 reported that lattice distortion induced by an optimal level of Ni-doping effectively reduces the surface energy of LiMn0.6Fe0.4PO4 crystals, limiting crystal growth and ensuring uniform particle size and shape. The resulting spherical LiMn0.6Fe0.4PO4 cathode materials exhibit a smooth surface, consistent sphericity, ideal lattice volume, and high phase purity, contributing to outstanding cycling stability and rate performance. After 100 cycles at 1C, the LiMn0.6Fe0.38Ni0.02PO4 cathode retains 98.3% of its initial capacity, delivering 147.3 mAh g−1. At higher rates of 10C and 15C, it achieves discharge capacities of 125.1 and 115.4 mAh g−1, respectively.

Despite significant progress in the development of isovalent Mg2+- or Ni2+-doped LMFP, the dual doping (or co-doping) with Mg/Ni combined with carbon coating emerges as a highly promising strategy for achieving multimetal LMFP materials with excellent capacity retention, even after thousands of charge–discharge cycles. Using this strategy, a Mg/Ni dual-doped and carbon-coated LMFP micro-sized secondary sphere was developed (Fig. 15a), exhibiting significantly enhanced reaction kinetics and exceptional structural stability.40 Ni2+ contributes to charge compensation during cycling, reducing the average valence state of Mn ions to mitigate Jahn–Teller distortion caused by Mn3+ while enhancing electronic conductivity in synergy with the carbon coating (Fig. 15c). Meanwhile, Mg2+, with its smaller ionic radius, expands the Li-ion transport channels by elongating the Li–O bonds. As a result, MgNi-LMFP/C achieves an impressive reversible capacity of 152 mAh g−1 at 0.1C, surpassing its counterpart, which delivers 148 mAh g−1 under the same conditions (Fig. 15b). Besides, this optimized LMFP cathode achieves a reversible specific capacity of 115 mAh g−1 at 5C, approximately 2.4 times higher than the pristine LMFP sample (Fig. 15d). Extended cycling tests conducted at a current density of 1C (Fig. 15e) demonstrate that MgNi-LMFP/C retains 95% of its capacity after 1000 cycles, significantly outperforming LMFP/C, which achieves only 88% retention. In a pouch-type full cell, it retains 92% of its initial capacity after 2000 cycles at 1C (Fig. 15f), underscoring its promising potential for high-power and long-life LIB applications.


image file: d5ta03194j-f15.tif
Fig. 15 (a) TEM image and EDS elemental mapping of the MgNi-LMFP/C sample. (b) Initial charge/discharge profiles at 0.1C. (c) Schematic representation of the crystal structures for MgNi-LMFP/C and LMFP/C in their charged states. (d) The rate capabilities of the four samples. (e) Cycling performance over 1000 cycles at 1C for MgNi-LMFP/C and LMFP/C. (f) Capacity retention after 2000 cycles at 1C for MgNi-LMFP/C and LMFP/C in pouch-type full cells. Reproduced with permission from ref. 40. Copyrights © The Royal Society of Chemistry.

Conversely, as previously emphasized, the Nb5+ cation stands out as a highly promising aliovalent dopant. Its charge disparity introduces vacancies into the crystal lattice, effectively increasing carrier concentration. This adjustment not only enhances the structural stability of LMFP by mitigating the Jahn–Teller effect but also significantly boosts its intrinsic conductivity.8 In fact, Nb-doping not only delivers a high initial specific capacity but also ensures capacity retention exceeding 90% over charge–discharge cycles as high as 500–1000. Similarly, Kong et al.78 demonstrated that Nb-doping facilitates the growth [001] crystal orientation of the material, thereby enhancing lithium-ion diffusion, reducing electrode polarization, and lowering charge transfer resistance, all contributing to its superior rate performance (Fig. 16a and b). Thus, a carbon-coated Li(Fe0.6Mn0.4)1−xNbxPO4 cathode material with x = 0.02 (named LFMP/C–Nb0.02), exhibited notable improvements in rate performance. The LFMP/C–Nb0.02 cathode achieved a high specific capacity of 156.9 mAh g−1 at 0.2 C and 121.4 mAh g−1 at 5C. Additionally, Nb doping increases the M–O bond energy, stabilizing the LFMP/C metal–oxygen framework and reinforcing the resilience of the crystal structure during lithium-ion intercalation and deintercalation cycles. A capacity retention rate of 94.15%, even after 500 cycles, significantly surpasses the original LFMP/C, which maintained only 61.47% of its capacity after 300 cycles.


image file: d5ta03194j-f16.tif
Fig. 16 (a) Cycling stability of all samples tested at 5C, and (b) discharge capacity retention comparison between LFMP/C and LFMP/C–Nb0.02 across varying rates. Reproduced with permission from ref. 78. Copyrights © 2024 Published by Elsevier Ltd (c) SEM micrograph, (d) high-magnification view, (e) high-resolution TEM image, and (f) TEM-EDS elemental mapping of LMFP/C–1%Nb. The inset in (e) displays the corresponding FFT pattern. (g) Mn discharge capacity retention across various rates, and (h) cycling performance at 1C for pristine LMFP/C and LMFP/C–Nb samples with varying Nb-doping levels. Reproduced with permission from ref. 94. Copyright © 2023 American Chemical Society.

Analogously, Jin and collaborators94 synthesized a 1% Nb-doped carbon-coated LiFe0.5Mn0.5PO4 (LMFP/C–1%Nb) cathode material using a combination of hydrothermal and spray-drying techniques. Microscopy techniques revealed the microstructure of LMFP/C–1%Nb samples, showing secondary microspheres (1–4 μm) composed of tightly packed nanoparticles (50–150 nm) (Fig. 16c and d). High-resolution TEM confirmed favorable crystallinity with lattice fringes matching LMFP planes and a 1.5 nm graphitized C-coating layer (Fig. 16e). Elemental mapping verified the uniform distribution of Mn, Fe, P, O, and Nb, unaffected by Nb-doping levels (Fig. 16f). The mechanism of action for the two strategies in LMFP operates through different pathways. The carbon coating is thought to create a highly conductive network between the primary particles while also preventing secondary reactions by shielding the material from direct electrolyte exposure. The strong Nb–O coordination exhibits higher bond energy compared to Mn–O and Fe–O, thus improving the stability of the crystal structure in LMFP materials.94 This helps suppress the Jahn–Teller effect of Mn3+, while simultaneously accelerating ion diffusion and electron transport within the nanoparticles. Thus, the dual modification strategy significantly enhanced both the lithium-ion diffusion coefficient and the electronic conductivity of LMFP, resulting in faster reaction kinetics. As a result, the optimized LMFP/C–1%Nb cathode exhibited outstanding electrochemical performance, delivering a lithium storage capacity of 115 mAh g−1 at 5 C and achieving exceptional capacity retention of 95.4% after 1000 cycles at 1 C (Fig. 16g and h).

Recently, advanced studies using in situ techniques have revealed fundamental insights into understanding the contributions of aliovalent ion doping (and/or carbon coating).71 For instance, in situ XRD analyses were carried out on carbon-coated and Nb-doped LMFP composites (named LMFP/DC-2Nb) and LMFP/C to gain deeper insights into the structural changes of LMFP throughout the lithiation and delithiation processes (Fig. 17a). Notably, LMFP/DC-2Nb exhibits a markedly reduced two-phase region compared to non-doped LMFP/C, with its fraction decreasing from 32% to 27%, indicating a more uniform phase transition and enhanced reaction consistency (Fig. 17b and c). DFT calculations further demonstrate that introducing an optimal amount of Nb5+ strengthens M–O bond energy, enhances structural stability, and boosts the material's intrinsic conductivity by narrowing its band gap. Still by in situ XRD, Gao et al.70 demonstrated a similar behavior in a Ti4+-doped LMFP (designated 0.75% Ti-LMFP/C) with high-rate performance, where the two-phase reaction region (LxMFTP → MFTP) becomes narrower through the combined effects of aliovalent ion doping and Fe vacancies (Fig. 17d–f).


image file: d5ta03194j-f17.tif
Fig. 17 (a) Structural evolution of LMFP/DC-2Nb throughout the charge–discharge cycle. (b) The initial charge–discharge profile at 0.2C over the 2.0–4.5 V range, along with representative in situ XRD patterns. (c) Rate performance at 1C for LMFP/DC-xNb samples (x = 0, 0.5, 1, 2, and 3) compared with LMFP/C. Reproduced with permission from ref. 71. Copyright © 2025 Elsevier Ltd All rights are reserved, including those for text and data mining, AI training, and similar technologies. In situ XRD patterns for (d) LMFP/C and (e) 0.75% Ti-LMFP/C electrodes, and (f) the corresponding phase transition schematic during the charge–discharge process of 0.75% Ti-LMFP/C. Reproduced with permission from ref. 70. Copyright © 2025 Elsevier Ltd All rights are reserved, including those for text and data mining, AI training, and similar technologies.

On the other hand, aliovalent Nb5+ ions have also been incorporated into coating materials to enhance the stabilization of multimetal LMFP structures. In fact, the use of high-valence ions such as Nb, Ti, V, Zr, and Si has been shown to form oxide-based layers, which can influence the lithium-ion diffusion kinetics and improve stability over cycles by increasing the reactivity of the surface between the cathode and electrolyte.94,125 In this context, P. Vanaphuti & A. Manthiram92 reported that LiNbO3 predominantly resides on the particle surface, while isovalent Mg2+ doping occurs within the bulk structure. The LiNbO3 surface coating enhances ionic conductivity, leading to a synergistic effect that provides higher specific capacity compared to pristine, Mg-doped, or Nb-doped LMFP. Apart from the higher specific capacity observed in the doubly modified sample, all samples demonstrate outstanding cycling stability over 200 cycles at a C/3 rate, maintaining a capacity retention of ≥98%. This remarkable performance is attributed to improved ionic/electronic transport from the LiNbO3 coating and accelerated Mn2+/3+ redox kinetics due to Mg bulk doping. Besides, Mg-containing LMFP with Nb-based coating can promote the formation of Fe2P, which enhances the material's conductivity and facilitates electronic transport in the cathode. Overall, the study highlights the critical role of synergistic dopants and coating materials in enhancing the capacity and cycling stability of LMFP cathodes.

5.1.2. Cation doping with Li-site substitution into LMFP. As previously mentioned, Li-site doping in LMFP cathodes is another possible strategy, resulting in a multi-elemental compound with generally improved electrochemical properties, especially when the newly introduced metal is present at low doping concentrations. In fact, doping at the Li-site significantly affects the diffusion of Li+, increasing the available one-dimensional pathway for its movement.90,114,126 Interestingly, and in a manner similar to results previously reported for LFP studies, the Li-site can be doped with various metal ions, including alkali metals, alkaline earth metals, and transition metals. For instance, different isovalent and aliovalent metal ions have been used in LMFP Li-sites, such as K+, Na+, Mg2+, V3+, Nb5+, etc. (Table 1).

The isovalent alkaline metal ions (Na+ and K+) have the same valence state as Li+, which does not alter the charge state of the Li-site in the olivine structure.10 For instance, Xu et al.91 highlighted the significant impact of Na+ doping on the electrochemical performance of Li1–xNaxMn0.6Fe0.4PO4/C (x = 0, 0.01, 0.02, 0.03) materials. A combination of experimental analysis and density functional theory (DFT) calculations demonstrated that Na+ doping enhances the crystallinity, reduces Li–Fe antisite defects, decreases primary particle size, and homogenizes particle size distribution without altering the olivine structure. The inferior rate and cycling performance of undoped LMFP were attributed to the sluggish Li+ diffusion kinetics associated with Mn redox processes. Na+-doping effectively increases the Li–O bond length, broadens Li+ diffusion channels, and lowers diffusion energy barriers, leading to accelerated Li+ diffusion and improved Mn redox kinetics. Additionally, the structural role of Na+ as a stabilizing pillar reduces Mn3+ content and weakens Mn–Mn interactions, mitigating the Jahn–Teller effect. These combined effects significantly enhance the high-rate capability and cycling stability of the Na+-doped materials. Among the tested compositions, Li0.98Na0.02Mn0.6Fe0.4PO4/C (x = 0.02) exhibited the best performance, delivering a specific discharge capacity of 125.0 mAh g−1 at 5C and retaining 96.7% of its capacity after 100 cycles at 1C.

Still for Na+-doping, Li et al.114 designed Li1−xNaxMn0.8Fe0.2PO4/C nanocapsule synthesized by a solvothermal method. Interestingly, it was observed that the doping of Na+ can effectively inhibit the dissolution of Mn2+ and Fe2+ in the electrolyte, a critical issue in maintaining cathode stability. The doping process, confirmed by EDS and XPS analyses, did not induce cation rearrangement in the structure, ensuring the structural integrity of the material. The regular diffraction lattice observed in Li0.97Na0.03Mn0.8Fe0.2PO4/C confirms the well-developed crystal structure, while ex situ XRD patterns after 200 cycles demonstrate remarkable structural stability. Importantly, Na+ doping does not alter the electrochemical reaction potential, ensuring compatibility with the inherent properties of the LMFP framework. In fact, electrochemical studies reveal that a 3% Na+ doping level achieves optimal performance, delivering specific capacities of 141.7, 125.0, and 89.5 mAh g−1 at 0.05C, 1C, and 5C rates, respectively. This improvement is attributed to the enhanced lithium-ion diffusion coefficient and the structural and chemical stability imparted by Na+-doping. Additionally, the capacity retention after 200 cycles at 0.5C remains at an impressive 96.65%, indicating excellent cyclic stability. The nanocapsule morphology further facilitates fast lithium-ion diffusion, contributing to the material's superior rate performance. In summary, Na+-doping in strategy effectively inhibits the dissolution of Mn2+ and Fe2+ in the electrolyte while maintaining the structural integrity of the cathode, making it a promising approach for next-generation LIBs.

Analogously to Na+-, K+-doping provides structural stability by mitigating volume changes during cycling and suppressing the Jahn–Teller effect of Mn, thereby enhancing the long-term durability of cathode material.127 This stabilization is evidenced by an increased intensity of the diffraction related to the (111) plane, attributed to lattice growth induced by K+. Due to its larger ionic radius, K+ induces significant lattice distortion when substituted at Li-sites, widening ion diffusion pathways and acting as a structural pillar. Furthermore, the introduction of point defects at the Li-site lowers the activation barrier for Li+ migration, thereby facilitating its transport and improving electrochemical performance. In fact, thanks to the widened ion channels and enhanced structural stability induced by doping, the KLMFP cathode delivers an impressive capacity of 145 mAh g−1 at 5C and maintains 96% of its capacity after 400 cycles. However, excessive K+ incorporation can disrupt the lattice structure, degrade cycling performance, and partially occupy Li+ sites, negatively impacting electrochemical behavior.127 On the other hand, it is important to highlight that the K+-containing LMFP reported by Geng et al.127 can be considered a dual-doped material, as it also incorporates Mn ions at the M-site with an atomic percentage of 5%.

Although significant advancements have been made through Na+- and K+-doping, experimental results indicate that only limited amounts of these dopants positively impact electrochemical performance.10 When their concentration exceeds a critical threshold, both discharge capacity and cycling stability deteriorate. This decline can be attributed to the substantially larger ionic radii of Na+ (1.02 Å) and K+ (1.31 Å) compared to Li+ (0.68 Å), leading to structural distortions that compromise the crystal integrity.10,90

Beyond isovalent doping, a recent approach involves incorporating aliovalent ions, with supervalent doping emerging as a growing trend in certain cases. In fact, aliovalent doping can serve as a strategic approach to introducing controlled atomic disorder within the ordered olivine structure, ultimately enhancing battery performance.128 A noteworthy case that still requires thorough investigation is Li-site doping with Mg2+ ions. For instance, while most studies report Mg2+-doping in LMFP at the M-site, research has also demonstrated the potential for Mg2+ incorporation at the Li-site.128 In a study by Hu et al.,116 it was observed that with increasing Mg-doping, the LMFP-based samples exhibited a slight shift toward lower angles (Fig. 18a). This shift occurs because the incorporation of Mg2+ at the Li-site extends the Li–O and O–O bond lengths, a consequence of lithium vacancy formation under the charge compensation mechanism, ultimately causing expansion of the olivine unit cells (Fig. 18b). This expansion increases the cell volume and broadens the one-dimensional Li+ diffusion channels, thereby enhancing the electrochemical performance of the material. Among the synthesized samples, Li0.97Mg0.015Mn0.8Fe0.2PO4 (LMFP-2) exhibits the highest reversible capacity (Fig. 18c) and cycle stability (Fig. 18d), with a discharge capacity of 149.8 mAh g−1 at 0.5C, which remains at 140.1 mAh g−1 after 200 cycles, corresponding to a retention rate of 93.5%. These findings suggest that Mg2+ substitution at the Li+-site effectively enhances electronic conductivity and Li+ mobility (as confirmed by EIS, Fig. 18e), thereby improving overall electrochemical performance. Furthermore, the superior rate capability highlights its potential for high-performance LIBs. The solvothermal synthesis of Mg-doped LMFP thus offers a promising pathway for the sustainable development and optimization of olivine-structured LFP materials.


image file: d5ta03194j-f18.tif
Fig. 18 (a) The diffractograms of four LMFP-based samples. (b) Crystal structure of the LMFP-0 (top) and LMFP-2 (bottom). (c) Rate capability, (d) cyclability at 1C, and (e) Nyquist plots of pristine LMFP and Mg-doped LMFP cathode materials. Reproduced with permission from ref. 116. Copyright © 2023 Elsevier Ltd all rights reserved.

Interestingly, studies suggest that supervalent dopants smaller than the host Li+ and Fe2+ tend to occupy the Li-site preferentially rather than replacing ions at the M-site in LFP-based materials.128,129 In fact, the olivine lattice can accommodate aliovalent cations ranging from 2+ to 5+, with charge compensation occurring through Li vacancies as one of the possible defect mechanisms.128 Although progress in the design of supervalent ion doping into the Li-site of LFP has been reported, the progress for doping in LMFP still needs to be studied, demonstrating that there is still plenty of room for further development.

5.1.3. Cation co-doping with Li-site & M-site substitution into LMFP. The investigation of multimetal incorporation in olivine-type materials has gained attention in recent studies due to its potential to improve the electrochemical performance of LMFP cathodes. In fact, the rational design in the selection of dopant ions can lead to synergistic effects, contributing to improvements, particularly in electronic conductivity, which consequently enhances the power density of devices, as well as the stability of LMFP-based cathodes during charge–discharge cycles.

To date, only two studies have reported a dual-site doping approach, indicating that the design of multimetal LMFP by this strategy is still in its infancy. For instance, recently, Wang et al.90 employed a dual-site doping design to enhance both ionic and electronic conductivities by simultaneously incorporating Na and Co ions into LMFP-electrode material. As shown in Fig. 19a, the Na and Co ions in LMFP act synergistically within the olivine structure. The large ionic radius of Na+ (1.02 Å) enhances the Li+ transport dynamics by expanding the Li+ channel through its placement at the Li-site. Meanwhile, Co2+ ions positioned at the M-site reduce the band gap, improving electronic conductivity, and mitigating the increase in the b-axis parameter, thereby shortening the Li+ transport path.90 Interestingly, the LMFP-Co/Na material exhibited a decrease in particle size from 91.8 to 75.6 nm compared to the undoped LMFP. This reduction in primary particle size can lead to superior kinetic performance, as the Li+ diffusion path is shortened. In Fig. 19b, the spherical shape and the homogeneous dispersion of Fe, Mn, Co, and Na elements in the material can be observed through TEM-EDS analysis. The LMFP modified by double doping exhibits excellent specific capacity values at a rate of 0.1C and a coulombic efficiency of 95.5% (Fig. 19c). In addition to the improved specific capacity at low rates, the material also demonstrates good capacity retention even at higher rates, whereas pure LMFP, when tested at a high rate such as 1C, showed a significant decrease in specific capacity (Fig. 19d). The excellent cycling stability and specific capacity retention of 97.1% after 1000 cycles at a rate of 1C demonstrate the effectiveness of the modification in the LMFP structure in mitigating the challenges of the Li+ diffusion path within the crystal lattice of the cathode (Fig. 19e). Thus, the synergistic effect of the Na/Co double doping led to a notable improvement in the electrochemical performance of the olivine cathode. This opens a promising new pathway for the implementation of novel multi-element co-doping combinations at multiple sites in LMFP.90


image file: d5ta03194j-f19.tif
Fig. 19 (a) Schematic diagram of LMFP modified by Na/Co dual-site doping approach; (b) HRTEM images of LMFP-Co/Na and EDS mapping images of LMFP-Co/Na; (c) first charge/discharge curves, (d) rate capability, and (e) ionic conductivity of LMFP, LMFP-Co, LMFP-Na, and LMFP-Co/Na. Reproduced with permission from.90 Copyright © 2024, the Author(s), under exclusive license to Springer-Verlag GmbH Germany, part of Springer Nature.

Similarly, Qiao et al.86 synthesized Li0.98Na0.02(Fe0.65Mn0.35)1−xMgxPO4/C (x = 0.01, 0.02, 0.03, 0.05) using the sol–gel method, incorporating Na+ and Mg2+ through a dual-site doping strategy. Their findings indicate that when x = 0.03, the material exhibits the most favorable electrochemical performance. The initial discharge capacity at 0.1C reaches an impressive 147.7 mAh g−1. Even after 40 cycles, when the current returns to 0.1C, the discharge capacity remains as high as 142.1 mAh g−1. These results suggest that optimized Na+ and Mg2+ doping enhances electrochemical properties, primarily by improving the Li+ diffusion coefficient and reducing charge transfer resistance, facilitating ion transport, and promoting electrochemical reactions.

5.2. Elemental incorporating structural components and/or entropy strategy

As highlighted previously, when the dopant concentration exceeds 5% by mass or atomic percentage, it should be regarded as a primary constituent or structural component of the electrode material. Thus, at high concentrations of the third metal ion, the term “doping” becomes inaccurate and may lead to confusion or misunderstanding. Therefore, this subtopic will mainly discuss advances in the incorporation of other ions at concentrations exceeding 5 mol% into LMFP cathodes. Based on these observations, LMFP cathodes incorporating a third structural component should be classified as ternary electrode materials.

Among the reported ternary electrode materials, metal ions such as Ni2+, Co2+, Cu2+, V3+, Ti4+, and Nb5+ have been incorporated into the olivine structure of LMFP (Table 1). It is essential to mention that even though such materials are not highlighted in the figure of merit, many of these ternary electrode materials have demonstrated high-capacity retention after 50–100 charge–discharge cycles.

In one of these interesting studies, Iturrondobeitia et al.93 designed three different ternary Mn-rich LiMn0.8Fe0.1M0.1PO4/C (M = Co2+, Ni2+, and Cu2+, isovalent metal ions) cathode materials via a freeze-drying method. The morphological characterization of ternary LiMn0.8Fe0.1M0.1PO4/C composites revealed that nanoparticles are embedded within an amorphous carbon matrix. This structural feature plays a crucial role in enhancing the electrochemical performance of these materials. Among the synthesized compositions, LiMn0.8Fe0.1Ni0.1PO4/C emerged as the most promising cathode, demonstrating superior electrochemical properties (Fig. 20a–c). For instance, the partial substitution of 0.1 mol of Fe2+ with Ni2+ led to an increase in specific capacity, reaching 110 mAh g−1 (Fig. 20a), while the incorporation of Co2+ resulted in a moderate capacity improvement, reaching 80 mAh g−1. These enhancements are attributed to several synergistic factors: the favorable morphological characteristics, the conductive network formed by the in situ generated carbon, and the mitigation of structural stress within the olivine framework due to the presence of Ni2+ and Fe2+, which contribute to improved stability and charge transport. In contrast, the incorporation of Cu2+ led to a marked reduction in specific capacity, reaching only 57 mAh g−1. This decline is likely attributed to the instability of the LiCuPO4 olivine phase, which impairs lithium-ion diffusion and structural integrity. The rate capability of LiMn0.8Fe0.1Ni0.1PO4/C was 95% of its initial specific capacity at 1C (Fig. 20c). Similarly, LiMn0.8Fe0.1Co0.1PO4/C exhibited commendable rate performance, maintaining 93% capacity retention at 1C. Conversely, LiMn0.8Fe0.1Cu0.1PO4/C displayed the lowest capacity retention of 85% at 1C, corroborating its higher polarization and inferior energy storage capability. These insights underscore the importance of compositional tuning in optimizing olivine-based cathode materials for high-performance LIBs.


image file: d5ta03194j-f20.tif
Fig. 20 (a) Initial charge–discharge curves and (b) cycling stability of the four evaluated LiMn0.8Fe0.1M0.1PO4/C (M = Fe, Co, Ni, Cu) composites at a C/25. (c) Corresponding modified Peukert plot. Reproduced with permission from ref. 93. Copyright © 2015 American Chemical Society (d) Initial charge/discharge profiles at 1C, (e) cycling stability at 1C, and (f) rate capability at varying C-rates for pristine and aliovalent cation-doped LMFP. Reproduced with permission from ref. 132. Copyright © 2016 Elsevier Ltd and Techna Group S.r.l. All rights reserved.

In addition to isovalent ions, alliovalent structural components have also been incorporated into the M-site in LMFP cathodes. For instance, the incorporation of Ti4+ at the M-site significantly enhances the electrochemical performance of LMFP due to the synergistic interaction between Ti4+ and Fe2+ (ref. 132) (Fig. 20d–f). This interaction contributes to improved structural stability of the olivine framework, facilitates lithium-ion diffusion, and enhances electrochemical kinetics. When compared to LiMn0.85Fe0.15PO4/C and LiMn0.92Ti0.08PO4/C, the Li(Mn0.85Fe0.15)0.92Ti0.08PO4/C composition demonstrates a significantly higher discharge capacity and superior rate capability. It achieves a capacity of 144.4 mAh g−1 with an exceptional retention rate of 99.9% after 50 cycles at 1C, highlighting its excellent cycling stability and rate performance. These findings confirm that the dual incorporation of an aliovalent cation at the M-site is an effective strategy for enhancing the electrochemical properties of LMFP-based cathodes.132

Even higher proportions of the third metal ion have been incorporated into LMFP cathodes, with reported concentrations reaching close to equimolarity at the M-site. Compounds with these characteristics are common among layered metal oxides for LIBs, such as LiNixCoyMnzO2 with x = y = z = 0.33 (NCM111). In fact, the increase in entropy has also extended to other types of materials for cathodes or catalysts, being a recent trend in sodium-ion11 and Li–S batteries.102 Thus, the design of electrode materials with equimolar proportions of metal ions is an important strategy, especially in the design of electrode materials with higher electrochemical stability, thereby configuring desirable medium- and high-entropy materials (ME- and HEMs).11,102 For instance, a medium-entropy lithium multi-transition metal phosphate, ternary LiCo1/3Mn1/3Fe1/3PO4, has been successfully synthesized via a solid-state reaction, exhibiting an olivine-type crystal structure.72 This material demonstrates a high specific capacity of 140 mAh g−1 and an operating voltage of 3.72 V at a 0.05C discharge rate, attributed to the redox activity of the Mn2+/Mn3+ and Co2+/Co3+ couples. In situ XRD analysis confirms that medium-entropy LiCo1/3Mn1/3Fe1/3PO4 maintains structural stability even at 5 V during charge/discharge cycles. While all three cations (Co2+, Mn2+, and Fe2+) occupy the same 4c site, the Co2+/Co3+ redox couple contributes a relatively lower capacity of approximately 20 mAh g−1. This reduced contribution is attributed to the significantly lower lithium-ion diffusivity associated with the Co redox process compared to Fe2+/Fe3+ and Mn2+/Mn3+. It is also important to emphasize that, while increased entropy can enhance structural stability, it may come at the cost of reduced specific capacity when compared with low-entropy LFP or LMFP. Therefore, achieving an optimal balance between stability and overall electrochemical performance is crucial.

On the other hand, Fan et al.107 employed the concept of entropy enhancement to highlight the potential for increasing the capacities of other olivine-type lithium metal phosphates, such as LiMnPO4 (LMP) and LiNiPO4 (LNP), despite their inherently insulating nature. They adopted an approach centered on high-entropy cathode engineering, capitalizing on the excellent conductivity imparted by both the maximized entropy and the distinctive antisite disorder in the framework. In more detail, a high-entropy variant of conventional olivine-based cathodes was produced using a mechanochemically assisted annealing method. Owing to the configurational entropy effect, the quinary Li(Mn0.2Fe0.2Co0.2Ni0.2Mg0.2)PO4 delivered higher specific capacities compared with unary LiMnPO4. Besides, adjusting the cation distribution and introducing Ca as a structural component expanded the crystal lattice in Li(Mn0.35Fe0.35Co0.1Mg0.1Ca0.1)PO4, thereby enhancing Li-ion diffusion. Thus, when employed as a high-voltage cathode, this quinary compound achieved an initial specific capacity of 104 mAh g−1, with three well-defined voltage plateaus attributed to the Fe, Mn, and Co redox processes between 2.0–4.9 V vs. Li+/Li, significantly outperforming single-metal LMP analogs (15 mAh g−1).

6. Conclusions and prospects

Lithium manganese iron phosphate (LiMnyFe1−yPO4, LMFP) has garnered significant attention as a cathode material due to its affordability, excellent safety, prolonged cycle life, high operating voltage, superior thermal stability, and remarkable energy density. However, despite notable advancements over the past decades, LMFP continues to encounter significant challenges. Its low electronic conductivity and sluggish Li-ion diffusion significantly restrict high-rate performance, while manganese dissolution compromises long-term cycling stability. Additionally, its low tap density poses further obstacles to practical applications. To facilitate the large-scale commercialization of LMFP in power batteries, continued research and innovative strategies are essential to overcome these technical hurdles.

A key emerging strategy for enhancing the stability and overall electrochemical performance of LMFP cathode materials, particularly in energy storage applications, is the development of multielemental compositions. These improvements can be achieved through various approaches, including (i) doping with one or more elements, (ii) incorporating additional primary constituents or structural components, and (iii) increasing the entropy of the electrode materials. Such design principles have proven effective in optimizing LMFP properties, ultimately advancing the performance and durability of next-generation energy storage systems. Table 1 and Fig. 12 provide a comprehensive overview of all multimetal LMFP cathodes, including key electrochemical performance metrics. Based on the analyzed strategies, several important conclusions can be derived, as highlighted below.

The synthesis process plays a pivotal role in optimizing the performance of multimetal LMFP cathodes since factors such as precursor selection, synthesis temperature, and methodology applied significantly influence the crystallinity, morphology, ion conductivity, structural stability, cycling performance, and energy density of the cathode materials. Among the various synthesis techniques, the solid-state reaction is the most widely used, offering advantages such as simplicity, cost-effectiveness, and the ability to achieve high crystallinity and uniformity through high-temperature treatments. The solvothermal method tends to be advantageous compared to the hydrothermal method, while more expensive and less reproducible, provides reasonable control over morphology, particle size, and crystallinity, making it useful for producing high-quality materials despite its higher cost and scalability limitations. Although mainly confined to laboratory use due to its lengthy procedure and high material costs, the sol–gel process allows for precise control over the composition and structure of the final product, which is valuable for research purposes. The co-precipitation method, with its low cost, short processing times, and simplicity, offers advantages in terms of the availability of raw materials and ease of execution. However, it faces challenges in controlling crystallization and ion behavior. In addition, physical processes such as freeze-drying and spray-drying enhance particle size control, uniformity, and overall processing efficiency. Each synthesis method, when optimized, plays a critical role in tailoring the properties of the LMFP cathodes, thereby impacting their overall electrochemical performance and advancing the development of high-performance battery technologies.

Incorporating multi-metals into LMFP-based electrodes follows distinct approaches, including elemental doping, incorporation of structural modification, and/or entropy-driven strategies, as highlighted previously. While doping typically should be restricted to a maximum of 5% by mass or atomic fraction, higher concentrations lead to structural integration of the introduced metal. Unfortunately, numerous authors in the literature incorrectly use the term doping to describe the incorporation of a new metal ion at concentrations as high as 10 mol% or mass%. However, at such levels, the added element should be considered a structural constituent of the composition rather than a dopant. Medium-entropy LMFP-based materials or high-entropy olivine structures have emerged as promising candidates, offering enhanced electrochemical properties in equimolar metal concentrations. Interestingly, among the investigated strategies, doping with one or more elements remains the most effective in fine-tuning electrochemical performance by optimizing lithium-ion diffusion, enhancing electronic conductivity, and stabilizing structural integrity. The incorporation of additional primary constituents or structural components can further improve electrode performance, particularly by mitigating volume expansion and facilitating charge transfer, but its success heavily depends on the compatibility and synergy of the added element. Increasing the entropy of electrode materials via equimolar multi-component design presents a promising avenue for long-term stability and structural robustness. However, its effectiveness in achieving high specific capacity and rate capability is still limited by sluggish kinetics and the challenge of balancing entropy-driven stabilization with the electrochemical activity of added metal ions. Therefore, a synergistic approach that employs metal or multimetal doping is the most promising path for achieving high-performance electrode materials. In fact, beyond enhancing electrochemical performance, doping also allows the incorporation of costlier elements than Fe and Mn without substantially increasing material costs. This is because only small concentrations of the added element are required (at% ≤3%), unlike the larger amounts needed when the third metal acts as a structural component or in entropy-enhanced electrode materials.

Among multimetal LMFPs, isovalent ions such as Mg2+ and Ni2+ dominate, whereas aliovalent metals like Nb5+ and V3+ are extensively studied. Transition-metal dopants like divalent Co2+ and Ni2+ generally enhance electronic conductivity and facilitate faster charge-transfer kinetics by introducing additional redox couples, whereas Mg2+ and Zn2+ tend to stabilize the olivine lattice and improve cycling stability at high rates. Aliovalent dopants such as Ti4+, Al3+, and particularly Nb5+ play a crucial role in suppressing antisite defects and reinforcing the crystal framework; Nb5+ can additionally promote lattice effects that enhance Li+ mobility, leading to notable improvements in both rate capability and long-term cycling. Comparative studies indicate that while transition-metal doping often provides higher initial capacity and rate performance, lattice-stabilizing dopants like Nb5+ yield superior structural integrity and capacity retention over prolonged cycling. Besides, recent in situ XRD studies further reveal that aliovalent ion doping (e.g., Nb5+ and Ti4+) markedly reduces the two-phase reaction region compared to non-doped LMFP/C (especially when associated with other strategies such as carbon coating or the presence of Fe vacancies), yielding a more uniform phase transition and enhanced reaction consistency, as well as narrowing the band gap and boosting intrinsic conductivity. Thus, a synergistic doping strategy, combining elements that enhance both conductivity and structural stability, appears to be the most promising route for optimizing LMFP performance for large-scale applications.

Conversely, Nb5+ aliovalent ions have also been integrated into coating materials to improve the stability of multimetal LMFP structures. Additionally, carbon-based materials, particularly LMFP/C composites, play a crucial role in optimizing electrochemical performance by establishing conductive networks. These advancements highlight the potential of multimetal LMFPs for next-generation lithium-ion batteries, balancing structural stability, conductivity, and electrochemical activity.

Despite considerable research advancements in recent years toward designing multimetal LMFP cathodes, substantial gaps remain unexplored. In this context, the following future perspectives and research directions should be considered for further advancements, as outlined below.

(i) Few research papers report the preparation of medium-entropy olivine structures, demonstrating that there is still much research to be done in developing such systems. In fact, while few medium-entropy cathodes have been explored, the design of high-entropy olivine electrodes remains unreported. In other types of cathode active materials, the development of medium- and high-entropy compounds has not led to significant improvements in specific capacity. However, their stability over charge–discharge cycles has shown promising results, particularly when incorporating electrochemically inactive metal ions. Therefore, future efforts should focus on this area, particularly targeting the development of cathode materials with enhanced energy density and charge retention, ideally under high-power operation. Furthermore, in the context of developing medium- and high-entropy olivine structures, it is essential to consider the total configurational entropy contribution from both the Li-site and M-site, despite the challenges in determining the specific site occupancy of certain metal elements.

(ii) The use of low-, medium-, and high-entropy coordination compounds as precursor materials for cathode-active materials remains largely unexplored. From this perspective, there is still significant research potential in designing multimetal olivine cathodes derived from metal–organic frameworks (MOFs), Prussian blue analogues (PBAs), as well as metal-polyphenol coordination polymers, and other typical coordination compounds. In fact, this strategy has been employed in many other research areas to prepare new materials for energy conversion and storage.147

(iii) So far, a few studies have explored a dual-site doping strategy, suggesting that the development of multimetal LMFP using this approach remains in its early stages. Future studies could focus on optimizing elemental selection and distribution at both sites to enhance electrochemical performance. Additionally, a deeper understanding of the structural and electronic interactions resulting from dual-site modifications could pave the way for improved cycling stability, rate capability, and overall energy storage efficiency. Expanding this strategy may also contribute to the design of next-generation LMFP cathodes with superior electrochemical properties.

(iv) Similar to the case of the Mg2+ ion, while most studies report its incorporation into LMFP at the M-site, some research has also suggested the possibility of its integration at the Li-site. This indicates that Mg2+ and other metal ions can potentially occupy both sites. However, accurately determining the exact location of the incorporated metal ion remains challenging. Therefore, it is crucial to develop studies utilizing methodologies capable of unambiguously identifying and quantifying the atomic occupancy of each site. To achieve this, advanced characterization techniques should be employed, particularly those leveraging synchrotron radiation. These include Extended X-ray Absorption Fine Structure (EXAFS), X-ray Absorption Near-Edge Structure (XANES), Near-Edge X-ray Absorption Fine Structure (NEXAFS), and X-ray Photoelectron Spectroscopy (XPS), which provide valuable insights into the structural and electronic environment of the incorporated elements.

(v) Despite the fact that nearly all high-performance multimetal LMFP cathodes incorporate a carbon coating approach, other carbonaceous materials have been scarcely reported in multimetal cathodes. In this regard, the use of alternative carbon-based materials, such as graphene/reduced graphene oxide, carbon nanotubes, and carbon nanofibers, could serve as promising platforms to enhance the power density of energy storage devices.

In summary, recent advancements in developing multimetal LMFP cathodes for LIBs demonstrate promising progress; however, significant challenges remain to fully meet the demands of our energy-intensive society. Given the latest findings, emerging trends, and future research directions, we hope this review contributes to the recognition and establishment of multimetal LMFP cathodes as next-generation electrode materials for LIBs. Furthermore, continued efforts in battery materials research are essential to developing advanced systems that can drive future breakthroughs in energy storage technologies.

Conflicts of interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this minireview.

Data availability

No primary research results, software, or code have been included, and no new data were generated or analysed as part of this review.

Acknowledgements

The authors wish to convey their profound gratitude for the substantial financial backing rendered by several distinguished Brazilian funding agencies, namely, the São Paulo Research Foundation (FAPESP 2023/17560-0 and 2017/11986-5), the Brazilian Innovation Agency (FINEP), the Foundation for Research Development (FUNDEP), and the National Council for Scientific and Technological Development (CNPq). This indispensable support has played an instrumental role in facilitating the successful execution of our research endeavors. B. G. A. N. and J. M. G acknowledge funding through MACKPESQUISA (Fundo Mackenzie de Pesquisa e Inovação; grant #231020 and #251035). Furthermore, we wish to extend our heartfelt appreciation to Shell for their unwavering commitment and valuable support throughout this study. In addition, we would like to acknowledge the strategic significance of the unwavering assistance provided by the Brazilian National Oil, Natural Gas, and Biofuels Agency (ANP) through the Research and Development (R&D) levy regulation.

References

  1. M. M. Hasan, R. Haque, M. I. Jahirul, M. G. Rasul, I. M. R. Fattah, N. M. S. Hassan and M. Mofijur, J. Energy Storage, 2025, 120, 116511 CrossRef.
  2. F. M. N. U. Khan, M. G. Rasul, A. S. M. Sayem and N. Mandal, Energy Rep., 2023, 9, 11–21 CrossRef.
  3. J. M. Gonçalves, G. T. M. Silva and H. Zanin, J. Mater. Chem. A, 2024, 12, 17756–17770 RSC.
  4. H. Wang, Q. Shi, J. Dong, M. Wang, Y. Lu, Y. Liu, J. Liu, N. Li, Q. Huang, Y. Su, F. Wu and L. Chen, Adv. Funct. Mater., 2025, 35, 2422806 CrossRef CAS.
  5. S. Mubarac, M. N. T. Silva, G. T. M. Silva, B. Freitas, J. M. Gonçalves and H. Zanin, J. Energy Storage, 2024, 96, 112612 CrossRef.
  6. J. R. Croy, A. Gutierrez, M. He, B. T. Yonemoto, E. Lee and M. M. Thackeray, J. Power Sources, 2019, 434, 226706 CrossRef CAS.
  7. B. Zhang, X. Wang, S. Wang, Y. Li, L. Chen, H. Jiao, Z. Yu and J. Tu, J. Energy Chem., 2025, 100, 1–17 CrossRef CAS.
  8. E. Xu, X. Sun, W. Lyv, F. Li, R. Li, W. Cai, H. Wu, K. Wu and Y. Zhang, Ind. Eng. Chem. Res., 2024, 63, 9631–9660 CrossRef CAS.
  9. Z. Qiu, Q. Man, Y. Mu, H. Gu, Z. Zou, M. Han and L. Zeng, Chem.: Methods, 2025, 5, 2400065 Search PubMed.
  10. K. Akhmetova, F. Sultanov, A. Mentbayeva, N. Umirov, Z. Bakenov and B. Tatykayev, J. Power Sources, 2024, 624, 235531 CrossRef CAS.
  11. N. G. Garcia, J. M. Gonçalves, C. Real, B. Freitas, J. G. Ruiz-Montoya and H. Zanin, Energy Storage Mater., 2024, 67, 103213 CrossRef.
  12. W. Ling, S. Fang, W. Zhou, C. Ye, L. Wang, N. Zhou and X. He, Adv. Energy Mater., 2025, 2502720 CrossRef CAS.
  13. D. Bresser, E. Paillard and S. Passerini, in Advances in Batteries for Medium and Large-Scale Energy Storage, Elsevier, 2015, pp. 125–211 Search PubMed.
  14. M. Li, J. Lu, Z. Chen and K. Amine, Adv. Mater., 2018, 30, 1800561 CrossRef PubMed.
  15. Y. Liu, R. Zhang, J. Wang and Y. Wang, iScience, 2021, 24, 102332 CrossRef CAS PubMed.
  16. A. A. Kebede, T. Kalogiannis, J. Van Mierlo and M. Berecibar, Renewable Sustainable Energy Rev., 2022, 159, 112213 CrossRef CAS.
  17. Q. Dong, S. Liang, J. Li, H. C. Kim, W. Shen and T. J. Wallington, iScience, 2023, 26, 107195 CrossRef CAS PubMed.
  18. Z. M. Ali, M. Calasan, F. H. Gandoman, F. Jurado and S. H. E. Abdel Aleem, Ain Shams Eng. J., 2024, 15, 102442 CrossRef.
  19. Z. W. B. Iton, S. S. Kim, E. S. Patheria, M. D. Qian, S. D. Ware and K. A. See, in Comprehensive Inorganic Chemistry III, Elsevier, 2023, pp. 308–363 Search PubMed.
  20. P. Zhu, D. Gastol, J. Marshall, R. Sommerville, V. Goodship and E. Kendrick, J. Power Sources, 2021, 485, 229321 CrossRef CAS.
  21. S. Li, X. Zhu, Y. Ma, F. Zhang and H. Zhou, J. Policy Anal. Manage., 2022, 41, 450–485 CrossRef.
  22. Y. Li, C. Liang, F. Ye and X. Zhao, Transp. Res. A: Policy Pract., 2023, 167, 103558 CrossRef.
  23. S. Abdul Qadir, F. Ahmad, A. Mohsin, A. B. Al-Wahedi, A. Iqbal and A. Ali, Energ. Strateg. Rev., 2024, 53, 101379 CrossRef.
  24. S. Yin, W. Deng, J. Chen, X. Gao, G. Zou, H. Hou and X. Ji, Nano Energy, 2021, 83, 105854 CrossRef CAS.
  25. G. Zhou, H. Chen and Y. Cui, Nat. Energy, 2022, 7, 312–319 CrossRef CAS.
  26. S. Zhao, Z. Guo, K. Yan, S. Wan, F. He, B. Sun and G. Wang, Energy Storage Mater., 2021, 34, 716–734 CrossRef.
  27. Y. Zou, G. Liu, K. Zhou, J. Zhang, T. Jiao, X. Zhang, Y. Yang and J. Zheng, ACS Appl. Energy Mater., 2021, 4, 11051–11061 CrossRef CAS.
  28. M. Zybert, H. Ronduda, W. Raróg-Pilecka and W. Wieczorek, Front Energy Res., 2023, 11, 1248641 CrossRef.
  29. J. Shi, D. Xiao, M. Ge, X. Yu, Y. Chu, X. Huang, X. Zhang, Y. Yin, X. Yang, Y. Guo, L. Gu and L. Wan, Adv. Mater., 2018, 30, 1705575 CrossRef PubMed.
  30. P. U. Nzereogu, A. D. Omah, F. I. Ezema, E. I. Iwuoha and A. C. Nwanya, Appl. Surf. Sci. Adv., 2022, 9, 100233 CrossRef.
  31. J. Qian, L. Liu, J. Yang, S. Li, X. Wang, H. L. Zhuang and Y. Lu, Nat. Commun., 2018, 9, 4918 CrossRef PubMed.
  32. A. Mauger and C. M. Julien, Batteries, 2018, 4, 39 CrossRef.
  33. X. Yao, D. Li, L. Guo, M. Kallel, S. D. Alahmari, J. Ren, I. Seok, G. Roymahapatra and C. Wang, Adv. Compos. Hybrid Mater., 2024, 7, 63 CrossRef CAS.
  34. S.-P. Chen, D. Lv, J. Chen, Y.-H. Zhang and F.-N. Shi, Energy Fuels, 2022, 36, 1232–1251 CrossRef CAS.
  35. R. Tang, J. Dong, C. Wang, Y. Guan, A. Yin, K. Yan, Y. Lu, N. Li, G. Zhao, B. Li, W. Shen, F. Wu, Y. Su and L. Chen, Adv. Funct. Mater., 2025, 35, 2421284 CrossRef CAS.
  36. S. B. Chikkannanavar, D. M. Bernardi and L. Liu, J. Power Sources, 2014, 248, 91–100 CrossRef CAS.
  37. P. Vanaphuti, K. Scanlan and A. Manthiram, RSC Sustain., 2024, 2, 1969–1978 RSC.
  38. N. Du, X. Lei, Q. Yu, C. Wang and S. Luo, Mater. Today Energy, 2025, 52, 101965 CrossRef CAS.
  39. G. Bree, J. Zhao, V. Majherova, D. Proprentner, G. J. Paez Fajardo and L. F. J. Piper, Energy Fuels, 2025, 39, 3683–3689 CrossRef CAS.
  40. H. Yu, E. Zhang, J. Yu, S. Yu, Y. Fang, L. Chen, H. Jiang and C. Li, J. Mater. Chem. A, 2024, 12, 26076–26082 RSC.
  41. K. Zhang, Z.-X. Li, X. Li, X.-Y. Chen, H.-Q. Tang, X.-H. Liu, C.-Y. Wang and J.-M. Ma, Rare Met., 2023, 42, 740–750 CrossRef CAS.
  42. M. D. Johannes, K. Hoang, J. L. Allen and K. Gaskell, Phys. Rev. B: Condens. Matter Mater. Phys., 2012, 85, 115106 CrossRef.
  43. Y. Dong, L. Wang, S. Zhang, Y. Zhao, J. Zhou, H. Xie and J. B. Goodenough, J. Power Sources, 2012, 215, 116–121 CrossRef CAS.
  44. S. Wi, J. Park, S. Lee, J. Kang, T. Hwang, K. S. Lee, H. K. Lee, S. Nam, C. Kim, Y. E. Sung and B. Park, Nano Energy, 2017, 31, 495–503 CrossRef CAS.
  45. S. Wi, J. Park, S. Lee, J. Kim, B. Gil, A. J. Yun, Y. E. Sung, B. Park and C. Kim, Nano Energy, 2017, 39, 371–379 CrossRef CAS.
  46. Y. Asari, Y. Suwa and T. Hamada, Phys. Rev. B: Condens. Matter Mater. Phys., 2011, 84, 134113 CrossRef.
  47. L. F. J. Piper, N. F. Quackenbush, S. Sallis, D. O. Scanlon, G. W. Watson, K. W. Nam, X. Q. Yang, K. E. Smith, F. Omenya, N. A. Chernova and M. S. Whittingham, J. Phys. Chem. C, 2013, 117, 10383–10396 CrossRef CAS.
  48. Y. Mishima, T. Hojo, T. Nishio, H. Sadamura, N. Oyama, C. Moriyoshi and Y. Kuroiwa, J. Phys. Chem. C, 2013, 117, 2608–2615 CrossRef CAS.
  49. D. Jang, K. Palanisamy, Y. Kim and W.-S. Yoon, J. Electrochem. Sci. Technol., 2013, 4, 102–107 CrossRef CAS.
  50. Y. Lu and T. Zhu, MRS Commun., 2024, 14, 888–899 CrossRef CAS.
  51. D. Kwon and D. Kim, Adv. Energy Mater., 2025, 15, 2501444 CrossRef CAS.
  52. A. K. Padhi, K. S. Nanjundaswamy and J. B. Goodenough, J. Electrochem. Soc., 1997, 144, 1188–1194 CrossRef CAS.
  53. G. R. Gardiner and M. S. Islam, Chem. Mater., 2010, 22, 1242–1248 CrossRef CAS.
  54. Y. Deng, C. Yang, K. Zou, X. Qin, Z. Zhao and G. Chen, Adv. Energy Mater., 2017, 7, 1601958 CrossRef.
  55. S. Li, H. Zhang, Y. Liu, L. Wang and X. He, Adv. Funct. Mater., 2024, 34, 2310057 CrossRef CAS.
  56. M. Kope¢, A. Yamada, G. Kobayashi, S. Nishimura, R. Kanno, A. Mauger, F. Gendron and C. M. Julien, J. Power Sources, 2009, 189, 1154–1163 CrossRef.
  57. I. Bezza, M. Kaus, R. Heinzmann, M. Yavuz, M. Knapp, S. Mangold, S. Doyle, C. P. Grey, H. Ehrenberg, S. Indris and I. Saadoune, J. Phys. Chem. C, 2015, 119, 9016–9024 CrossRef CAS.
  58. A. Yamada, Y. Kudo and K. Y. Liu, J. Electrochem. Soc., 2001, 148, A747–A754 CrossRef CAS.
  59. C. C. Yang and W. H. Chen, Mater. Chem. Phys., 2016, 173, 482–490 CrossRef CAS.
  60. A. Paolella, G. Bertoni, E. Dilena, S. Marras, A. Ansaldo, L. Manna and C. George, Nano Lett., 2014, 14, 1477–1483 CrossRef CAS PubMed.
  61. A. Yamada, Y. Kudo and K. Y. Liu, J. Electrochem. Soc., 2001, 148, A1153–A1158 CrossRef CAS.
  62. N. D. Trinh, Z. W. Ai, G. Liang and S. B. Schougaard, Solid State Ionics, 2018, 324, 33–39 CrossRef CAS.
  63. D. B. Ravnsbæk, K. Xiang, W. Xing, O. J. Borkiewicz, K. M. Wiaderek, P. Gionet, K. W. Chapman, P. J. Chupas and Y. M. Chiang, Nano Lett., 2014, 14, 1484–1491 CrossRef PubMed.
  64. L. Yang, W. Deng, W. Xu, Y. Tian, A. Wang, B. Wang, G. Zou, H. Hou, W. Deng and X. Ji, J. Mater. Chem. A, 2021, 9, 14214–14232 RSC.
  65. Y. Li, C. Yu, X. Wang, Y. Zhong, Z. Wu, X. Wang and B. Zhong, Ind. Eng. Chem. Res., 2025, 64, 13920–13949 CrossRef CAS.
  66. Z. Peng, X. Xia, Y. Xiao and T. Huang, J. Alloys Compd., 2025, 1016, 178901 CrossRef CAS.
  67. S. Xin, H. Zhang, Z. Hu, P. Zhao, C. Zhou, H. He, P. Liu, J. Zhang and J. Zhou, ACS Appl. Mater. Interfaces, 2025, 17, 33783–33795 CrossRef CAS PubMed.
  68. J. Han, J. Zhu, X. He, M. Yang, C. Yan, D. Ma, L. Zhang and P. Zhang, ACS Appl. Mater. Interfaces, 2025, 17, 32381–32391 CrossRef CAS PubMed.
  69. J. Huang, X. Tang, Y. Zhou, T. Wang, F. Zhao, W. Wang, Y. Meng, W. Cen and Y. Zhang, J. Mater. Chem. A, 2025, 13, 10550–10560 RSC.
  70. X. Gao, S. Zheng, Y. Song, C. Chen, X. Long, Z. Qiu, W. Li, P. Liu, X. Li, Y. Cui and W. Xing, J. Energy Storage, 2025, 131, 117670 CrossRef.
  71. C. Ye, W. Ling, X. Huang, Y. Wu, N. Zhou and X. Yang, Electrochim. Acta, 2025, 536, 146789 CrossRef CAS.
  72. Y.-C. Chen, J.-M. Chen, C.-H. Hsu, J.-J. Lee, T.-C. Lin, J.-W. Yeh and H. C. Shih, J. Power Sources, 2010, 195, 6867–6872 CrossRef CAS.
  73. L. Qin, Y. Xia, H. Cao, L. Yang and Z. Liu, Electrochim. Acta, 2016, 222, 1660–1667 CrossRef CAS.
  74. E. Boldyreva, Chem. Soc. Rev., 2013, 42, 7719 RSC.
  75. M. Ratynski, M. Winkowska-Struzik, D. Buchberger, B. Hamankiewicz, M. Krajewski and A. Czerwinski, ACS Appl. Energy Mater., 2025, 8, 11053–11067 CrossRef CAS.
  76. Y. Cao, J. He, W. Tu, R. Li, W. Luo, X. Zhang, J. Duan, D. Wang, X. Wang, P. Dong and Y. Zhang, Solid State Ionics, 2024, 411, 116576 CrossRef CAS.
  77. M.-J. Uddin, P. K. Alaboina and S.-J. Cho, Mater. Today Energy, 2017, 5, 138–157 CrossRef.
  78. D. Kong, H. Chen, F. Wu, R. Zhang, J. Li, Y. Mai, Y. Wei, J. Wang and X. Dai, Electrochim. Acta, 2024, 506, 145060 CrossRef CAS.
  79. D. Kong, H. Chen, F. Wu, R. Zhang, J. Li, Y. Mai, Y. Du, J. wang and X. Dai, Ceram. Int., 2024, 50, 47308–47316 CrossRef CAS.
  80. Q. Deng, T. Li, J. Wang, S. Zhang, H. Yang, C. Xu and M. Wu, Ceram. Int., 2024, 50, 13702–13710 CrossRef CAS.
  81. Z. Dai, L. Wang, X. He, F. Ye, C. Huang, J. Li, J. Gao, J. Wang, G. Tian and M. Ouyang, Electrochim. Acta, 2013, 112, 144–148 CrossRef CAS.
  82. D. Bokov, A. Turki Jalil, S. Chupradit, W. Suksatan, M. Javed Ansari, I. H. Shewael, G. H. Valiev and E. Kianfar, Adv. Mater. Sci. Eng., 2021, 2021, 5102014 CrossRef.
  83. L. L. Hench and J. K. West, Chem. Rev., 1990, 90, 33–72 CrossRef CAS.
  84. A. E. Danks, S. R. Hall and Z. Schnepp, Mater. Horiz., 2016, 3, 91–112 RSC.
  85. H. Liu, Y. P. Wu, E. Rahm, R. Holze and H. Q. Wu, J. Solid State Electrochem., 2004, 8, 450–466 CrossRef CAS.
  86. S. Qiao, L. Zhu, E. Han, L. Li, C. Du and Y. He, Int. J. Electrochem. Sci., 2019, 14, 10616–10629 CrossRef.
  87. H. Dong and G. M. Koenig, CrystEngComm, 2020, 22, 1514–1530 RSC.
  88. J. Li, Z. Liu and J. Ma, Sci. Rep., 2025, 15, 28004 CrossRef CAS PubMed.
  89. X. Bi, L. Chang, S. Luo, S. Cao, A. Wei, W. Yang, J. Liu and F. Zhang, Ionics, 2022, 28, 577–591 CrossRef CAS.
  90. P. Wang, E. Zhang, Y. Fang, Y. Chen, H. Yu, Y. Zhang, Q. Cheng and H. Jiang, J. Solid State Electrochem., 2024, 28, 4303–4310 CrossRef CAS.
  91. J. Xu, K. Hou, X. Li, Y. Bian, Y. Wang, L. Wang and G. Liang, ACS Appl. Energy Mater., 2024, 7, 8694–8704 CrossRef CAS.
  92. P. Vanaphuti and A. Manthiram, Small, 2024, 20, 2404878 CrossRef CAS PubMed.
  93. A. Iturrondobeitia, A. Goñi, I. Gil de Muro, L. Lezama, C. Kim, M. Doeff, J. Cabana and T. Rojo, Inorg. Chem., 2015, 54, 2671–2678 CrossRef CAS PubMed.
  94. H. Jin, J. Zhang, L. Qin, Y. Hu, H. Jiang and C. Li, Ind. Eng. Chem. Res., 2023, 62, 1029–1034 CrossRef CAS.
  95. S. Tian, K. Zhang, J. Cao, H. Guo, R. Liu and G. Liang, Ionics, 2021, 27, 2877–2887 CrossRef CAS.
  96. S. Mi, Z. Liu, C. Luo, L. Cai, Z. Zhang and L. Li, Drying Technol., 2020, 38, 1544–1564 CrossRef.
  97. W. Abdelwahed, G. Degobert, S. Stainmesse and H. Fessi, Adv. Drug Deliv. Rev., 2006, 58, 1688–1713 CrossRef CAS PubMed.
  98. B. Vertruyen, N. Eshraghi, C. Piffet, J. Bodart, A. Mahmoud and F. Boschini, Materials, 2018, 11, 1076 CrossRef PubMed.
  99. C. Arpagaus, A. Collenberg and D. Rütti, Drying Technol., 2019, 37, 426–434 CrossRef CAS.
  100. A. B. D. Nandiyanto and K. Okuyama, Adv. Powder Technol., 2011, 22, 1–19 CrossRef CAS.
  101. R. Palm, A. M. Baena-Moncada and J. M. Gonçalves, J. Mater. Chem. A, 2024, 12, 29402–29431 RSC.
  102. J. M. Gonçalves, É. A. Santos, P. R. Martins, C. G. Silva and H. Zanin, Energy Storage Mater., 2023, 63, 102999 CrossRef.
  103. T. G. Ritter, S. Pappu and R. Shahbazian-Yassar, Batteries, 2024, 10, 96 CrossRef CAS.
  104. S. Schweidler, M. Botros, F. Strauss, Q. Wang, Y. Ma, L. Velasco, G. Cadilha Marques, A. Sarkar, C. Kübel, H. Hahn, J. Aghassi-Hagmann, T. Brezesinski and B. Breitung, Nat. Rev. Mater., 2024, 9, 266–281 CrossRef.
  105. S. Baúti, J. V. T. Feyh, F. C. Antunes, T. Dias, G. Doubek, J. M. Gonçalves, B. B. N. S. Brandão and H. Zanin, J. Energy Chem., 2025, 111, 513–544 CrossRef.
  106. J. M. Gonçalves, É. A. Santos, M. M. Amaral, E. Nossol, R. O. Figueiredo and H. Zanin, ACS Omega, 2025, 10(35), 39309–39333 CrossRef PubMed.
  107. J. Fan, T. Wang, Y. Yuan, A. Borisevich, C.-L. Do-Thanh, Z. Yang and S. Dai, Appl. Phys. Lett., 2024, 124, 171909 CrossRef CAS.
  108. X. Zhang, M. Hou, A. G. Tamirate, H. Zhu, C. Wang and Y. Xia, J. Power Sources, 2020, 448, 227438 CrossRef CAS.
  109. T. Wu, J. Liu, L. Sun, L. Cong, H. Xie, A. Abdel-Ghany, A. Mauger and C. M. Julien, J. Power Sources, 2018, 383, 133–143 CrossRef CAS.
  110. J. Yu, J. Ma, H. Zhang, J. Hang, W. Tian, X. Tian and H. Yang, Ceram. Int., 2025, 51, 26443–26451 CrossRef CAS.
  111. Z. Li, Y. You, Z. Zhu, L. Wang, S. Ou, J. Xu and M. Yuan, J. Energy Storage, 2025, 110, 115111 CrossRef.
  112. M. Zhao, Y. Zhou, Y. Chen, X. Liang, J. Zeng, K. Bai, X. Xu, H. Wang, X. Jiang and H. He, Adv. Funct. Mater., 2025, 2509461 CrossRef.
  113. K. Zhang, J. Cao, S. Tian, H. Guo, R. Liu, X. Ren, L. Wen and G. Liang, Ionics, 2021, 27, 4629–4637 CrossRef CAS.
  114. R. Li, C. Fan, W. Zhang, M. Tan, T. Zeng and S. Han, Ceram. Int., 2019, 45, 10501–10510 CrossRef CAS.
  115. Y. Wang, H. Yang, C.-Y. Wu and J.-G. Duh, J. Mater. Chem. A, 2017, 5, 18674–18683 RSC.
  116. H. Hu, H. Li, Y. Lei, J. Liu, X. Liu, R. Wang, J. Peng and X. Wang, J. Energy Storage, 2023, 73, 109006 CrossRef.
  117. D. Kim, S. Lee and W. Choi, J. Alloys Compd., 2024, 989, 174396 CrossRef CAS.
  118. W. Liu, X. Liu, R. Hao, Z. Yang, B. Ouyang, M. Zhang, M. Pan and K. Liu, J. Electroanal. Chem., 2023, 929, 117117 CrossRef CAS.
  119. J. Ni, Y. Zhao, J. Chen, L. Gao and L. Lu, Electrochem. Commun., 2014, 44, 4–7 CrossRef CAS.
  120. G. K. P. Dathar, D. Sheppard, K. J. Stevenson and G. Henkelman, Chem. Mater., 2011, 23, 4032–4037 CrossRef CAS.
  121. J. Duan, G. Hu, Y. Cao, K. Du and Z. Peng, Ionics, 2016, 22, 609–619 CrossRef CAS.
  122. C. Hu, B. Wang, Huihua, J. Zhang, Y. Hu and J. Li, Int. J. Electrochem. Sci., 2018, 13, 5824–5831 CrossRef CAS.
  123. K. Kisu, E. Iwama, W. Onishi, S. Nakashima, W. Naoi and K. Naoi, J. Mater. Chem. A, 2014, 2, 20789–20798 RSC.
  124. T. Zeng, P. Gao, Z. Zhou, C. Fan, Z. Liu, F. Zhang, J. Liu and J. Liu, Energy Storage Mater., 2024, 65, 103125 CrossRef.
  125. L. Li, L. Wu, F. Wu, S. Song, X. Zhang, C. Fu, D. Yuan and Y. Xiang, J. Electrochem. Soc., 2017, 164, A2138–A2150 CrossRef CAS.
  126. P. Su, H. Zhang, L. Yang, C. Xing, S. Pan, W. Lu and S. Zhang, Chem. Eng. J., 2022, 433, 133203 CrossRef CAS.
  127. J. Geng, Z. Zou, T. Wang, S. Zhang, S. Zhong, W. Ling, X. Peng and X. Hu, J. Electroanal. Chem., 2023, 933, 117275 CrossRef CAS.
  128. N. Meethong, Y. Kao, S. A. Speakman and Y. Chiang, Adv. Funct. Mater., 2009, 19, 1060–1070 CrossRef CAS.
  129. S.-Y. Chung, J. T. Bloking and Y.-M. Chiang, Nat. Mater., 2002, 1, 123–128 CrossRef CAS PubMed.
  130. H. Yi, C. Hu, H. Fang, B. Yang, Y. Yao, W. Ma and Y. Dai, Electrochim. Acta, 2011, 56, 4052–4057 CrossRef CAS.
  131. S. Liu, H. Fang, E. Dai, B. Yang, Y. Yao, W. Ma and Y. Dai, Electrochim. Acta, 2014, 116, 97–102 CrossRef CAS.
  132. Q.-Y. Huang, Z. Wu, J. Su, Y.-F. Long, X.-Y. Lv and Y.-X. Wen, Ceram. Int., 2016, 42, 11348–11354 CrossRef CAS.
  133. P. Xiao, Y. Cai, X. Chen, Z. Sheng and C. Chang, RSC Adv., 2017, 7, 31558–31566 RSC.
  134. H. Shu, X. Wang, Q. Wu, B. Hu, X. Yang, Q. Wei, Q. Liang, Y. Bai, M. Zhou, C. Wu, M. Chen, A. Wang and L. Jiang, J. Power Sources, 2013, 237, 149–155 CrossRef CAS.
  135. H. Yi, C. Hu, X. He and H. Xu, Ionics, 2015, 21, 667–671 CrossRef CAS.
  136. B.-K. Zou, Y. Shao, Z.-Y. Qiang, J.-Y. Liao, Z.-F. Tang and C.-H. Chen, J. Power Sources, 2016, 336, 231–239 CrossRef CAS.
  137. W. Xiang, Y. Zhong, Y. Tang, H. Shen, E. Wang, H. Liu, B. Zhong and X. Guo, J. Alloys Compd., 2015, 635, 180–187 CrossRef CAS.
  138. C. Hu, H. Yi, H. Fang, B. Yang, Y. Yao, W. Ma and Y. Dai, Electrochem. Commun., 2010, 12, 1784–1787 CrossRef CAS.
  139. J. E. Frąckiewicz and T. K. Pietrzak, Materials, 2021, 14, 6434 CrossRef PubMed.
  140. Y. Cai, D. Zhang, C. Chang, Z. Sheng and K. Huang, Ionics, 2016, 22, 1011–1019 CrossRef CAS.
  141. X. Chu, L. Li, W. Chen and H. Fang, Ionics, 2021, 27, 2927–2935 CrossRef CAS.
  142. H. Fang, E. Dai, B. Yang, Y. Yao and W. Ma, J. Power Sources, 2012, 204, 193–196 CrossRef CAS.
  143. X.-Y. Lv, Q.-Y. Huang, Z. Wu, J. Su, Y.-F. Long and Y.-X. Wen, J. Solid State Electrochem., 2017, 21, 1499–1507 CrossRef CAS.
  144. B. J. Jeong, J. Y. Sung, F. Jiang, S. P. Jung and C. W. Lee, J. Energy Storage, 2024, 96, 112552 CrossRef.
  145. J. Zheng, J. Yang, J. Wu, S. Li, M. Wang, B. Huang, Y. Li, S. Xiao and Q. Zhu, J. Alloys Compd., 2023, 960, 170610 CrossRef CAS.
  146. Y. Niu, S. Wang, R. Chang, P. Yang, H. Xing and Y. Xu, ACS Appl. Mater. Interfaces, 2025, 17, 18831–18843 CrossRef CAS PubMed.
  147. J. M. Gonçalves and J. G. Ruiz-Montoya, J. Mater. Chem. A, 2023, 11, 20872–20885 RSC.

Footnote

Equal contributors.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.