Next Article in Journal
Electrochemical and Optical Carbon Dots and Glassy Carbon Biosensors: A Review on Their Development and Applications in Early Cancer Detection
Previous Article in Journal
High-Precision Tunneling Magnetoresistance (TMR) Current Sensor for Weak Current Measurement in Smart Grid Applications
Previous Article in Special Issue
Preparation and Performance Study of MXene-Regulated Ethylene Glycol-Induced WO3 Film
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Review of Layered Transition Metal Oxide Materials for Cathodes in Sodium-Ion Batteries

Department of Chemical and Materials Engineering, New Mexico State University, Las Cruces, NM 88003, USA
*
Authors to whom correspondence should be addressed.
Micromachines 2025, 16(2), 137; https://doi.org/10.3390/mi16020137
Submission received: 7 January 2025 / Revised: 21 January 2025 / Accepted: 22 January 2025 / Published: 24 January 2025
(This article belongs to the Special Issue Energy Conversion Materials/Devices and Their Applications)

Abstract

:
The growing interest in sodium-ion batteries (SIBs) is driven by scarcity and the rising costs of lithium, coupled with the urgent need for scalable and sustainable energy storage solutions. Among various cathode materials, layered transition metal oxides have emerged as promising candidates due to their structural similarity to lithium-ion battery (LIB) counterparts and their potential to deliver high energy density at reduced costs. However, significant challenges remain, including limited capacity at high charge/discharge rates and structural instability during extended cycling. Addressing these issues is critical for advancing SIB technology toward industrial applications, particularly for large-scale energy storage systems. This review provides a comprehensive analysis of layered sodium transition metal oxides, focusing on their structural properties, electrochemical performance, and degradation mechanisms. Special attention is given to the intrinsic and extrinsic factors contributing to their instability, such as structural phase transitions, and cationic/anionic redox behavior. Additionally, recent advancements in material design strategies, including doping, surface modifications, and composite formation, are discussed to highlight the progress toward enhancing the stability and performance of these materials. This work aims to bridge the knowledge gaps and inspire further innovations in the development of high-performance cathodes for sodium-ion batteries.

1. Introduction

The urgent global need for clean, affordable, and sustainable energy storage systems has become increasingly critical due to the escalating effects of climate change, the rapid depletion of fossil fuel resources, and rising environmental pollution. To address these challenges, the adoption of advanced technologies for energy storage and conversion has gained immense importance. Over the past few decades, renewable energy sources such as wind, solar, nuclear, and wave energy have emerged as promising alternatives. However, the intermittent nature of these energy sources, influenced by factors like time, weather, and location, has limited their widespread application, underscoring the need for efficient energy storage systems [1,2,3,4].
Electrochemical energy storage devices, including batteries, fuel cells, and supercapacitors, are among the most effective technologies for storing electricity and powering modern devices. Among these, rechargeable (secondary) batteries stand out for their superior energy and power density compared to fuel cells and supercapacitors. Since their commercialization by Sony in the 1990s, lithium-ion batteries (LIBs) have become a cornerstone of modern society, transforming how we interact with technology by enabling wireless and portable solutions. LIBs have dominated battery technology for decades, powering devices such as laptops, smartphones, and even electric and hybrid vehicles, significantly reducing dependence on fossil fuels [5,6,7,8].
Lithium’s unique properties, including high voltage, energy density, and specific power compared to other elements (Figure 1a) make it an ideal material for battery applications, particularly in the automotive industry. However, the surging demand for LIBs has led to a steep rise in the cost of lithium-based precursors like lithium carbonate (Li2CO3) over the past decade (Figure 1b). Additionally, the geographically constrained lithium reserves, predominantly located in Latin America and Australia, have raised concerns about supply security and cost, especially for large-scale applications. These challenges highlight the need for alternative energy storage solutions that are cost-effective, widely available, and sustainable (Figure 1c) [9,10,11].
Sodium-ion batteries (SIBs) share a similar architecture and working principles with LIBs (Figure 2), making them a promising alternative. They are environmentally benign, cost-effective, and abundant, addressing several limitations associated with LIBs. Research into SIBs began in the early 1980s, but the successful commercialization of LIBs in the 1990s shifted focus away from sodium-based systems [12]. Table 1 provides a general comparison between sodium and lithium batteries, highlighting their key characteristics and differences.
In addition to SIBs, solid-state sodium batteries are emerging as a promising candidate for the future of large-scale energy storage applications. The organic electrolytes currently used in SIBs—such as propylene carbonate, ethylene carbonate, and dimethyl carbonate—struggle to meet the performance requirements of both cathodes and anodes, particularly in achieving high electrochemical performance. Moreover, the flammability of these liquid electrolytes raises significant safety concerns, as seen in lithium-ion batteries, particularly in large battery systems for vehicles and grid applications. Solid electrolytes offer a compelling solution by effectively mitigating the risks of short circuits caused by sodium dendrites and eliminating the hazards associated with flammable organic electrolytes. The use of solid-state batteries, which leverage the inherent non-flammability of solid electrolytes, is anticipated to address these safety issues comprehensively [13,14].
Several types of solid electrolytes have been explored for rechargeable sodium batteries, including Na-ion conductive ceramics, polymers, and ceramic–polymer composites. Among these, organic polymer electrolytes—created by dissolving sodium salts into a polymer matrix—have garnered significant attention. These materials offer high flexibility, excellent thermal stability, and the ability to form intimate contact with both electrodes and inorganic solid electrolytes. With moderate ionic conductivity (>10−4 S cm−1 at room temperature), polymer-based solid electrolytes remain one of the most widely investigated options for solid-state sodium batteries [15,16,17,18].
Sodium is one of the most abundant elements in the Earth’s crust, with a Clarke number of 2.64 (Figure 3a). It exists in various forms such as sodium carbonate (Na2CO3), sodium chloride (NaCl), and sodium sulfate (Na2SO4), commonly found in minerals and brines. Sodium also exhibits an electrochemical redox potential close to that of lithium (−2.71 V for sodium versus −3.04 V for lithium, both measured against the standard hydrogen electrode), making it suitable for battery applications [19,20].
The abundance of sodium resources makes it a highly accessible material, with significantly lower costs compared to lithium. For example, the cost of trona, a key source of sodium carbonate, ranges between USD 135 and USD 165 per ton, whereas lithium carbonate, a precursor material for LIBs, cost approximately USD 5000 per ton in 2010. This stark contrast in resource availability and cost highlights sodium as a viable alternative for scalable and cost-effective energy storage systems (Figure 3b,c). Therefore, the cost per kilowatt-hour (USD/kWh) for SIBs can be significantly reduced due to the absence of expensive transition metals like cobalt and nickel in the cathode materials.
Additionally, aluminum can replace copper as the current collector for the anode because sodium does not react with aluminum. This allows the aluminum current collector to remain stable in the electrolyte at 0 V, unlike in LIBs, where lithium and copper form a binary alloy at low potentials (below 0.1 V vs. Li/Li+).
However, sodium’s larger ionic radius (1.02 A° compared to 0.76 A° for lithium) and heavier atomic mass (23 g mol−1 vs. 7 g mol−1 for lithium) contribute to slower ion kinetics during the (de)sodiation process and lower gravimetric capacity (1165 mAh g−1 compared to 3829 mAh g−1). This leads to structural instability in the host materials, increased heat generation from exothermic reactions, and a lower theoretical energy density. While high energy density is not a primary requirement for stationary, in large-scale energy storage systems—such as those paired with solar panels where cost efficiency is prioritized—these issues still negatively affect the performance of SIBs [23,24,25].
Moreover, the solid electrolyte interphase (SEI) formed on the electrodes in SIBs is less stable due to sodium’s lower Lewis acidity, leading to a high solubility of the SEI components. This results in incomplete coverage of the electrode surface, promoting side reactions and excessive heat generation. These factors highlight that the electrochemical behavior of SIBs cannot be directly extrapolated from LIBs. To overcome these challenges and improve performance, comprehensive studies are essential to understanding the unique electrochemical processes and underlying mechanisms of SIBs.
Cathode materials reported for SIBs include layered oxides, polyanions, organic compounds, and Prussian blue and its analogues. Despite their potential, these materials face persistent challenges such as low ionic conductivity and thermal instability, which limit their performance. Nonaqueous organic liquid electrolytes are predominantly used in SIBs due to their wide electrochemical stability window, high ionic conductivity, and efficient mass transfer at the electrolyte–electrode interface. However, they also present safety concerns, including flammability and potential for thermal runaway.
Identifying a cathode material with superior electrochemical properties—such as high capacity, long cycle life, and optimal operating voltage—remains a top priority for advancing SIB technology. Currently, cathode materials in SIBs typically exhibit specific capacities ranging from 120 to 200 mA h g−1, with operating potentials between 2.6 and 3.2 V. These values are considerably lower than the oxidation potential of electrolytes, which exceeds 4.0 V. In contrast, key anode materials offer specific capacities of 300–600 mA h g−1, with operating potentials between 0.1 and 0.6 V, which are close to the potential for sodium plating. Consequently, cathode materials not only represent a larger mass fraction in SIBs, but they also offer more potential for enhancing the overall voltage of the system. Therefore, cathodes play a critical role in determining the cost, safety, specific energy, cycle life, and specific power of SIBs. Unfortunately, the fundamental differences in intercalation chemistry between lithium and sodium often result in the underperformance of sodium-based analogues of lithium-based compounds. These differences underscore the need for tailored materials and innovative design strategies to optimize the performance of SIB cathodes [26,27,28,29,30].
The α-NaFeO2 (R-3m) structured materials hold a pivotal role in the development of LIBs and serve as a foundation for similar materials in SIBs. These materials feature a layered structure composed of slabs of edge-sharing MO6 octahedra, with alkali ions intercalated between the layers [31]. Similar to lithium-based transition metal oxides, such as LiCoO2 and Li(Ni, Mn)O2, sodium transition metal oxides (NaxTMO2, where TM = transition metals and 0.67 ≤ x ≤ 1) are commonly used as positive electrodes in SIBs [32,33,34].
Typical sodium layered oxides can be synthesized in O3-, P2-, and P3-phases, which are determined by the surrounding Na environment and the number of unique oxide layer packings. The notation O3, P3, and P2 describe the coordination environment and stacking order: O indicates that sodium ions occupy octahedral (O) sites, while P signifies sodium ions in trigonal prismatic (P) sites. The numbers 3 and 2 represent the number of edge-sharing MO6 octahedra in the oxygen stacking sequence, arranged in ABCABC or ABBA patterns, respectively [35]. The O3 phase typically occurs when 0.7 ≤ x ≤ 1. The oxide layer stacking follows the ABCABC pattern, where all Na ions share one edge and one face with adjacent octahedra. When 0.6 < x < 0.7, the P2 phase is known to occur, which features oxide layer stacking that follows the ABBA pattern, with Na ions sharing either entirely edge or entirely face positions with adjacent octahedra. P3 typically occurs when x ~ 0.5, the oxide layer stacking in this phase follows the ABBCCA pattern, where Na ions share one face with a transition metal octahedron (MO6) and three edges with three adjacent MO6 octahedra, as illustrated in Figure 4 [36]. The O3 phases belong to the R-3m space group, whereas P2 phases crystallize in the P63/mmc space group. Structural distortions in these phases, such as monoclinic (C2/m) or orthorhombic (Cmcm) modifications, are denoted by a prime symbol (for example, O3′ or O′3) [37,38].
The extraction of alkali ions during battery operation creates vacancies in the Li or Na layers, which can facilitate the migration of TMs into these sites. This migration leads to irreversible structural degradation and poor electrochemical performance in layered oxides. The phenomenon of TM migration has been extensively studied in LIBs cathode materials. For instance, Wang et al. [39] observed the migration of Ni and Co ions in LiNi1/3Mn1/3Co1/3O2 and attributed it to structural distortions and the formation of irreversible phases.
Both experimental observations and theoretical calculations indicate that distortions in the MO6 octahedron play a critical role in promoting TM migration. These distortions may arise from TM oxidation or oxygen oxidation during cycling [40,41]. A key characteristic of these oxidation processes is the alteration in TM-O bond lengths, which induces strain within the MO6 octahedra or the O-TM-O slabs. This strain serves as a driving force for TM migration, ultimately impacting the structural integrity and electrochemical stability of the cathode material.
The cycling life of cathode materials in SIBs is primarily influenced by the structural stability of the materials, as well as the electrochemical stability of the compatible electrolytes. Electrochemical behavior is significantly influenced by the structure of the phase, not only because of the amount of Na in the pristine state but also due to the stability of each layer. The kinetics of sodium insertion and extraction are affected by the surrounding environment of Na within the layered structure, which impacts the overall performance of the material. The stability of each layer plays a crucial role in maintaining structural integrity during cycling, while the arrangement and interaction of Na ions within the structure influence the efficiency of ion transport and the electrochemical response. This review focuses on the phase transitions, structural stability, and electrochemical characteristics of transition metal layered oxide materials, as well as various strategies reported to enhance the stability of these materials for SIBs.
Figure 4. Schematic illustration of the typical crystal structures of NaxTMO2, showing (a) P2, (b) O2, (c) OP4, (d) O3, and (e) P3 phases. Purple spheres represent transition metals, yellow spheres denote Na+ ions, and red spheres indicate O2− ions. Reprinted from Ref. [42] with permission from Wiley.
Figure 4. Schematic illustration of the typical crystal structures of NaxTMO2, showing (a) P2, (b) O2, (c) OP4, (d) O3, and (e) P3 phases. Purple spheres represent transition metals, yellow spheres denote Na+ ions, and red spheres indicate O2− ions. Reprinted from Ref. [42] with permission from Wiley.
Micromachines 16 00137 g004

2. Materials Classification

2.1. NaCrO2

Among NaxMO2 compounds (0 < x ≤ 1, M = Ni, Co, Mn, Fe, V, and Cr), O3-NaCrO2 has garnered significant attention as a promising cathode material for sodium-ion batteries. It delivers a discharge capacity of 120 mAh g−1 with a desirable voltage plateau at ~3 V (vs. Na/Na+) and exhibits excellent thermal stability compared to LiCoO2 and LiFePO4. This stability is attributed to the robust chemical stability of Cr4+ within CrO2 slabs at octahedral sites, combined with high-rate capability and excellent coulombic efficiency [43,44,45].
Interestingly, while LiCrO2, which shares a layered rock salt structure with LiCoO2 and LiNiO2, is electrochemically inactive, NaCrO2 demonstrates high capacity. This discrepancy stems from differences in interstitial spaces between the two compounds. In LiCrO2, the irreversible migration of Cr ions into tetrahedral sites during lithium de-intercalation renders the material inactive after the first charge. In contrast, NaCrO2 avoids this migration when x ≤ 0.5 due to its larger tetrahedral sites, although further sodium removal (x > 0.5) leads to gradual inactivation [46,47]. Structurally, both LiCrO2 and NaCrO2 feature Cr in the trivalent state with a d3 electronic configuration (t2g3, eg0) under octahedral coordination. Upon oxidation, Cr(III) forms the unstable Cr(IV) (d2), which disproportionate into Cr(III) (d3) and Cr(VI) (d0) due to their higher stability. Cr(VI) typically adopts a tetrahedral coordination with four oxygen atoms, as seen in Cr(VI)O42− of Li2CrO4·2H2O. The dimensional match between the interstitial tetrahedron and Cr(VI)O42− plays a crucial role in structural stability [48].
For LiCrO2, the interstitial tetrahedron with an O-O bond length of 2.89 A° and a height of 2.56 A° closely aligns with the Cr(VI)O42− tetrahedron. However, in NaCrO2, the interstitial tetrahedron within the NaO2 slab exhibits an O-O bond length of 2.97 A° and a height of 3.15 A°, better accommodating the Cr(VI)O4 structure. Consequently, the bond length mismatch in NaCrO2 prevents irreversible Cr migration, unlike in LiCrO2 (Figure 5). Magnetic measurements of chemically deintercalated Na1-xCrO2 confirm the presence of Cr(IV) during sodium deintercalation.
In NaCrO2, oxygen atoms at the 6c sites are stacked along the c-axis, with Na and Cr ions alternately distributed on both sides of the oxygen layers at the 3a and 3b octahedral sites. This layered structure enables two-dimensional Na+ ion transport between CrO2 slabs, enhancing ionic mobility. However, consecutive phase transitions during cycling—from O3 to O′3 and eventually P3—pose significant challenges.
To date, most NaCrO2 materials have been synthesized via solid-state reactions, which suffer from capacity decay and poor power performance due to irreversible phase transitions, low Na+ diffusivity, and side reactions at the electrode–electrolyte interface [7]. Addressing these issues is critical for realizing the full potential of NaCrO2 as a cathode material in sodium-ion batteries.
The initial charge and discharge curves for NaCrO2 and LiCrO2 are shown in Figure 6a,b, respectively. LiCrO2, upon electrochemical oxidation to 4.5 V vs. Li, demonstrates a capacity of 55 mAh g−1, corresponding to the formation of Li0.81CrO2. This capacity is consistent with the charge compensation mechanism driven by the increased valence of Cr ions during lithium extraction. A slight reversible capacity is observed in the potential range of 3.0–4.5 V; however, subsequent cycles reveal a drastic drop in capacity to less than 10 mAh g−1, signifying the absence of reversible lithium intercalation. This behavior is attributed to the so-called ‘freeze effect’ of Cr6+ ions migrating into tetrahedral sites post-delithiation, as suggested by Komaba et al. [49,50,51].
In the case of NaCrO2 synthesized by traditional solid-state reaction of stoichiometric Cr2O3 and Na2CO3 at 900 °C in Ar environment, sodium deintercalation occurs electrochemically, exhibiting multiple plateaus (Figure 6b) corresponding to phase transitions: hexagonal O3 → monoclinic O3 → monoclinic P3 [52]. However, when cycled to a high cutoff voltage of 4.5 V, the intercalation process becomes nearly irreversible, resulting in a limited reversible capacity of 26 mAh g−1 (Figure 6c). This suggests that complete sodium deintercalation induces an irreversible structural change due to the destabilization of the deintercalated phase, which lacks the sodium pillars necessary to maintain the structural integrity of CrO2 slabs.
At a lower cutoff voltage of 3.6 V, the charge–discharge curves reveal highly reversible behavior with a capacity of 104 mAh g−1. This highlights that extracting more than 50% of sodium ions from NaCrO2 disrupts crystallinity, transitioning the material into a highly disordered and irreversible phase. The irreversible order-to-disorder transition during deep charging is attributed to Cr migration into vacant Na layers, driven by the disproportionation reaction of Cr(IV) at high potentials (3Cr4+ → 2 Cr3+ + Cr6+). The smaller ionic radius of Cr6+ and its preference for tetrahedral coordination facilitate this migration. Importantly, this phenomenon occurs exclusively in O-like phases, as P-like phases lack the necessary tetrahedral sites in their sodium-ion layers.
The disordered bulk structure is believed to comprise a mixture of O3-like NaδCrO2 (δ ~ 0) and rock salt CrO2, as suggested by prior investigations [43]. These structural transformations underline the importance of optimizing charge cutoff voltages to mitigate irreversible phase changes and enhance the cycling stability of NaCrO2 cathodes.
Figure 7a,b depict the cycling performance of a Na/NaCrO2 cell. The distinct and repeatable plateaus observed in the charge and discharge curves demonstrate the reversible nature of Na+ intercalation within the NaCrO2 electrode [48].
Zhou et al. [51] provided a comprehensive analysis of the reaction mechanism and real-time structural evolution of the layered NaCrO2 cathode material during Na extraction. Utilizing synchrotron-based in situ XRD (Figure 8a), the structural changes in NaCrO2 were examined under a current density of C/12 and a cutoff voltage of 3.6 V.
At the onset of charging, the (003) diffraction peak shifted progressively to lower two-theta angles, indicative of a solid solution reaction accompanied by expansion along the c-axis. When x = 0.08, a new (003) peak emerged to the left of the original peak, and its intensity increased with continued charging. This indicated the formation of a distinct phase. In the range 0.08 < x < 0.2, a two-phase coexistence region was observed, with the intensity of the secondary (003) peak increasing at the expense of the first, while its position continued to shift to lower angles. By x = 0.25, the second phase dominated completely, with the original phase disappearing.
As the charge progressed to x = 0.3, another (003) peak appeared, and the second phase’s peak broadened, marking the formation of a third phase. By x = 0.4, only the third phase remained, and its (003) peak shifted further, albeit to a lesser extent than the second phase. This sequence of transitions demonstrated the systematic structural evolution during Na extraction.
Additional insights were obtained by analyzing the (110) peak, initially located at 62.4°, which reflected changes in the a-b lattice plane. As charging advanced, the sharp (110) peak diminished in intensity and broadened, indicating lattice distortion in the a-b plane. Concurrently, its shift to higher two-theta angles suggested a contraction in the a and b lattice parameters.
Based on the observed Bragg peak variations, the structural evolution of NaCrO2 involved phase transitions among three solid-solution phases, interspersed with two two-phase coexistence regions. Initially, the material adopted an O3-type rhombohedral phase (O3R). After x = 0.08, the second phase emerged, as evidenced by the splitting of the rhombohedral (012) peak, signaling a transition to a monoclinic phase. By x = 0.25, the second phase, identified as the monoclinic Na0.74CrO2 phase (O3M), dominated while maintaining O3 symmetry, albeit with distorted a-b planes and ABCABC oxygen atom stacking.
As x increased from 0.3 to 0.48, the third phase grew to dominate, and its Bragg reflections corresponded to the Na0.52CrO2 phase, identified as a P3-type monoclinic phase (P3M) with ABBCCA oxygen atom stacking. These findings highlight the complexity of the structural transformations occurring in NaCrO2 during electrochemical charging and their implications for its electrochemical performance.
The evolution of the a, b, and c lattice parameters during the initial charge of the NaCrO2 electrode, as derived from in situ XRD analysis, is presented in Figure 8b. During the charging process, the c-axis parameter exhibited an increase, which is attributed to the enhanced electrostatic repulsion between the MO2 layers as sodium ions were removed. These sodium ions act as a structural “glue” between the MO2 layers, and their extraction leads to an expansion of the interlayer spacing.
In contrast, the a and b lattice parameters decreased, reflecting a contraction of the CrO6 octahedra. This contraction is a direct consequence of the oxidation of Cr3+ to higher oxidation states during charging, resulting in shorter bond lengths and a more compact structure within the CrO2 slabs. These lattice parameter changes collectively highlight the structural adjustments occurring in NaCrO2 during sodium extraction.
The sodium extraction mechanism for the NaCrO2 electrode during the charging process, as elucidated through the authors’ experiments, can be summarized as follows:
Open Circuit Voltage (OCV) to 3.05 V:
NaCrO2 (O3R) → Na0.75CrO2 (O3M) + 0.25 Na+ + 0.25 e
3.05 V to 3.60 V:
Na0.75CrO2 (O3M) → Na0.5CrO2 (P3M) + 0.25 Na+ + 0.25 e
From the OCV to approximately 3.05 V, the O3R-type NaCrO2 undergoes a two-phase transformation into O3M-type Na0.75CrO2 (Reaction 1), facilitated by transitions through two solid-solution regions. Within each region, the ‘a’ and ‘b’ lattice parameters decrease, reflecting the contraction of the CrO6 octahedra, while the ‘c’ lattice parameter increases due to the reduced electrostatic interactions between the MO2 layers as sodium ions are extracted. This structural evolution is accompanied by a lattice distortion, transitioning from a rhombohedral to a monoclinic phase. During this process, Cr ions remain confined to octahedral sites, while Na ions migrate from octahedral to pseudo-tetrahedral sites.
Between 3.05 V and 3.60 V, the O3M-type Na0.75CrO2 undergoes another two-phase transformation into the P3M-type Na0.5CrO2 (Reaction 2), also involving two solid-solution regions. In this range, the Cr ions retain their octahedral coordination but decrease in size due to further oxidation. Simultaneously, Na ions transition from pseudo-tetrahedral to prismatic sites. These phase transitions are characterized by the gliding of oxygen planes, which is directly associated with the progressive removal of sodium ions.
Bo et al. [53] explored the structural evolution of NaCrO2 under a current density of C/50 up to 4 V, corroborating the structural changes observed by Zhou et al. [51] up to 3.6 V. As illustrated in Figure 9a, further sodium extraction beyond 3.8 V, corresponding to a Na0.38CrO2 composition, marks the onset of irreversible electrochemical behavior. This is evidenced by the appearance of diffraction peaks associated with new phase(s), notably a broad peak around 8°. In Figure 9b, these unidentified phases are denoted with an ‘X’, highlighting their ambiguous nature.
The transition from the P′3 phase to the unidentified ‘X’ phase progresses as the cutoff voltage of 4 V is approached, resulting in a system containing approximately 0.04 Na, as inferred from the capacity measurements. The phase transition near Na0.4CrO2 (around 3.7 V) and the significant decrease in discharge capacity observed beyond 3.8 V imply that this structural transformation is a key contributor to the capacity fading of NaCrO2.
In addition to phase transition characterization techniques, cyclic voltammetry (CV) was employed to investigate the structural changes occurring at both reversible and irreversible voltage ranges. Jakobsen et al. [54] conducted CV experiments in the reversible potential range of 2.5–3.7 V, as depicted in Figure 10a. During the initial charge, two distinct oxidation peaks were observed at 3.07 V (with a shoulder at 3.13 V) and 3.33 V, with corresponding reduction peaks at 2.83 V and 3.28 V during the first discharge. With repeated cycling, the potential hysteresis of the lowest potential event diminished, and after five cycles, the peaks shifted to 3.04 V and 2.87 V. Additionally, the reduction peak revealed a shoulder in the range of 2.9–2.95 V, analogous to the shoulder at 3.13 V during the charging process. Importantly, this shoulder is not associated with irreversibility, as this potential range is characterized by excellent cycling reversibility.
The authors attribute these features to several crystalline-to-crystalline structural transitions, summarized in Figure 10b. Initially, monoclinic distortions result in the formation of the O′3-NaxCrO2 phase through a first-order two-phase transition. This phase subsequently transforms into another monoclinic phase, O′3-E-NaxCrO2 (where ‘E’ denotes an expanded interlayer distance), before transitioning to the monoclinic P′3-NaxCrO2 phase. Despite the asymmetry between charge and discharge processes, the overall O3 ⇋ P′3 phase transitions are reversible. Furthermore, O3-NaCrO2 maintains good cycling performance when less than 0.5 Na ions are extracted.
When charged beyond 3.5 V, Jakobsen et al. [43] identified an oxidation peak at 3.67 V in the CV profile. If the charge does not exceed this peak (Figure 10c), the subsequent cycles closely replicate the initial cycle, demonstrating stability in the electrochemical response. Conversely, when the charge surpasses 3.67 V (Figure 10d), subsequent cycles deviate significantly from the initial one, exhibiting broader peaks indicative of increased solid-solution behavior. Despite these changes, Na intercalation remains reversible.
At voltages approaching 4.0 V, corresponding to the complete removal of Na ions, an additional oxidation peak emerges at approximately 3.85 V. As shown in Figure 10e, deep charging to this voltage induces irreversibility in NaxCrO2, as evidenced by the absence of reduction peaks during the discharge process. This suggests that the extraction of the final Na ions disrupts the structural integrity of the material, rendering the intercalation mechanism non-reversible.
Yabuuchi et al. [44] investigated the thermal stability of the NaCrO2 cathode material. Their results showed that NaCrO2 exhibited no phase transitions during heating in the temperature range of 27–527 °C, with only thermal expansion observed. The thermal expansion of NaCrO2 was significantly lower than that of LiNi0.5Mn0.5O2. Specifically, the linear thermal expansion coefficients for NaCrO2 were 0.90 × 10−5 °C−1 (ahex) and 1.48 × 10−5 °C−1 (chex), compared to 1.68 × 10−5 °C−1 (ahex) and 2.21 × 105 °C−1 (chex) for LiNi0.5Mn0.5O2. The thermal expansion of NaCrO2 was anisotropic, with the ratio of ahex/chex increasing as the temperature rose. Furthermore, oxygen loss in P3-NaCrO2 occurred at significantly higher temperatures (above 500 °C) compared to Ni-based cathode materials in LIBs, which experience oxygen loss between 240 and 260 °C.
In order to improve structural stability of NaCrO2 cathode materials, elemental doping has been an interesting strategy, for example, Ti [55], Sn [56], Ru [57], Nb [58], Mn [59], Al [60], and Ca [46,61]. Ma et al. [62] explored the effect of Sb substitution on the structural stability of Na1-2xCr1-xSbxO2 (x = 0, 0.05, 0.1, and 0.2) as a cathode material for SIBs. The Sb5+ ion, a semiconducting metal ion with an ionic radius (0.0600 nm) comparable to Cr (0.0615 nm), enhances structural stability due to its high valence state and the introduction of electron holes, which modify the local electronic environment. Their findings indicated that substituting a suitable amount of Sb, specifically in Na0.90Cr0.95Sb0.05O2 (NCS-0.05), significantly improved the material’s cycling and rate performance.
NCS-0.05 demonstrated exceptional cycling stability, retaining 70.9% of its initial capacity after 1000 cycles, in contrast to the 46.8% and 47.8% capacity retention observed for NaCrO2 and Na0.8Cr0.9Sb0.1O2 (NCS-0.1), respectively (Figure 11a). At a low current density of 0.1 C, NCS-0.05 exhibited a reversible capacity of approximately 120 mAh g−1. While the capacity decreased with increasing Sb substitution in the initial cycles, as the current density increased from 0.1 C to 32 C, the capacities of NaCrO2 (NC), NCS-0.05, and NCS-0.1 declined to 31.3, 96.1, and 69.3 mAh g−1, respectively. Upon reverting the current density to 0.1 C, their capacities recovered near their initial values. This comparison highlights that NCS-0.05 exhibited the best rate capability among the samples (Figure 11b). Based on the characterization results, the increase in the c/a ratio due to Sb substitution expands the Na+ diffusion channel, facilitating enhanced sodium diffusion. This structural adjustment is advantageous for improving the reversibility of Na+ intercalation and de-intercalation processes.
Ti doping has been reported for both P2- and O3-type NaxCrO2 cathode materials in SIBs [55,63,64]. Li et al. [64] investigated the electrochemical performance of O3-Na1-xCr1-xTixO2 (with x = 0.03, 0.05, and 0.10). Although Ti doping reduces the initial capacity of the material, it significantly enhances the initial coulombic efficiency due to improved structural stabilization and the expansion of the NaO2 interlayer spacing, attributed to an increase in the c/a lattice ratio. Long-term cycling performance at a rate of 1 C revealed that as x increases, the initial specific discharge capacities of the samples decrease (103.7, 100.6, 96.7, and 88.8 mAh g−1 for x = 0, 0.03, 0.05, and 0.10, respectively). However, the corresponding capacity retentions improve substantially (50.0%, 71.6%, 80.1%, and 85.8% after 800 cycles, respectively, Figure 11c). This demonstrates that Ti doping greatly enhances the cycling performance of NaCrO2, further validating the improved structural stability with increasing Ti content. Additionally, the Ti-doped samples exhibit superior rate performance. When tested at a high rate of 30 C (3 A g−1), Na0.95Cr0.95Ti0.05O2 delivers a reversible capacity of 67 mAh g−1, significantly outperforming the 42 mAh g−1 capacity of undoped NaCrO2 (Figure 11d). The improved stability of Ti-doped samples is attributed to the higher bond dissociation energy of the Ti-O bond (661.9 kJ mol−1) compared to the Cr-O bond (427 kJ mol−1) at 298 K. This stronger bond suppresses TMO2 layer gliding during cycling, delaying the phase transition from the hexagonal O3 structure to the monoclinic P3 structure, thereby enhancing the material’s structural integrity and electrochemical performance.
Figure 11. (a) Rate capability of NC, NCS-0.05, and NCS-0.1 at rates ranging from 0.1 C to 32 C, highlighting the superior performance of NCS-0.05 under high current densities, (b) cycling performance at 5 C for NC, NCS-0.05, and NCS-0.1, with NCS-0.05 exhibiting the highest capacity retention over prolonged cycles. Reprinted from Ref. [62] with permission from Elsevier, (c) cycling performance of Na1-xCr1-xTixO2 at a rate of 1 C, indicating the enhanced capacity retention of Ti-doped samples over prolonged cycling, and (d) rate capability of Na1−xCr1−xTixO2 samples, tested at rates ranging from 0.1 C to 30 C, highlighting the superior rate performance of Ti-doped materials compared to undoped NaCrO2. Reprinted from Ref. [64] with permission from Elsevier. Ex situ HR-TEM images of (e) NaCrO2, (f) Na0.97Cr0.97Ti0.03O2, (g) Na0.93CrCa0.035O2, and (h) Na0.9Ca0.035Cr0.97Ti0.03O2 after 1000 cycling at 10 C. SAED pattern of (i) NaCrO2, (j) Na0.97Cr0.97Ti0.03O2, and (k) Na0.93CrCa0.035O2 along the P′3 [001] zone axis, and (l) Na0.9Ca0.035Cr0.97Ti0.03O2 along the O3 [001] zone axis, respectively. Reprinted from Ref. [61] with permission from Elsevier.
Figure 11. (a) Rate capability of NC, NCS-0.05, and NCS-0.1 at rates ranging from 0.1 C to 32 C, highlighting the superior performance of NCS-0.05 under high current densities, (b) cycling performance at 5 C for NC, NCS-0.05, and NCS-0.1, with NCS-0.05 exhibiting the highest capacity retention over prolonged cycles. Reprinted from Ref. [62] with permission from Elsevier, (c) cycling performance of Na1-xCr1-xTixO2 at a rate of 1 C, indicating the enhanced capacity retention of Ti-doped samples over prolonged cycling, and (d) rate capability of Na1−xCr1−xTixO2 samples, tested at rates ranging from 0.1 C to 30 C, highlighting the superior rate performance of Ti-doped materials compared to undoped NaCrO2. Reprinted from Ref. [64] with permission from Elsevier. Ex situ HR-TEM images of (e) NaCrO2, (f) Na0.97Cr0.97Ti0.03O2, (g) Na0.93CrCa0.035O2, and (h) Na0.9Ca0.035Cr0.97Ti0.03O2 after 1000 cycling at 10 C. SAED pattern of (i) NaCrO2, (j) Na0.97Cr0.97Ti0.03O2, and (k) Na0.93CrCa0.035O2 along the P′3 [001] zone axis, and (l) Na0.9Ca0.035Cr0.97Ti0.03O2 along the O3 [001] zone axis, respectively. Reprinted from Ref. [61] with permission from Elsevier.
Micromachines 16 00137 g011
Lee et al. [61] investigated the effects of Ca and Ti co-substitution in the NaCrO2 structure and found that it significantly improved cycling performance at a high current discharge capacity of 10 C. Based on ex situ TEM images (Figure 11e–h), structural damage and deformation were observed in pure NaCrO2, as well as in samples with single-element substitution (Ca or Ti). In contrast, the co-substituted sample exhibited a smooth surface, indicating minimal distortion in the layered structure, even at the particle edges, which are typically the most sensitive locations.
The structural damage in the other samples was attributed to large variations in unit cell volume and irreversible phase transitions during cycling. Selected area electron diffraction (SAED) analysis (Figure 11i–l) further confirmed the structural stability of Na0.9Ca0.035Cr0.97Ti0.03O2. The angle between the (010) and (1 –10) planes for this sample remained close to that of the O3-NaCrO2 structure (119.7° compared to 120°), demonstrating excellent structural retention over cycling. In contrast, the corresponding angles for cycled NaCrO2, Na0.93CrCa0.035O2, and Na0.97Cr0.97Ti0.03O2 were 116.9°, 117.9°, and 118.1°, respectively, values closer to the P3’-Na0.4CrO2 structure, indicating less structural stability.
Surface modification has been widely recognized as an effective approach to enhance the structural stability of NaCrO2 cathode materials. Techniques such as Cr2O3 [60,65] and carbon coatings [66,67,68] have been frequently employed in various studies. Ding et al. [69] investigated the effects of carbon coating on NaCrO2 cathode materials to improve their stability.
The charge and discharge curves for both uncoated (naked) and carbon-coated NaCrO2 electrodes at different current rates are shown in Figure 12a. While the voltage profiles of the carbon-coated and naked electrodes are similar, the carbon-coated electrode exhibits a lower voltage polarization between the charge and discharge profiles. At a current rate of 5 mA g−1, the carbon-coated NaCrO2 delivers higher charge and discharge capacities of 135 mAh g−1 and 116 mAh g−1, respectively, compared to 125 mAh g−1 and 106 mAh g−1 for the naked electrode. However, an increase in current rate leads to higher polarization and rapid capacity fading for both electrodes.
Figure 12b presents the cycling performance at a current rate of 5 mA g−1. Both the carbon-coated and naked electrodes initially show an increase in discharge capacity, reaching maximum values of 118 mAh g−1 and 110 mAh g−1 in the third and fourth cycles, respectively. The naked NaCrO2, however, exhibits faster capacity fading during cycling. After 40 cycles, the carbon-coated NaCrO2 retains a discharge capacity of 110 mAh g−1, which is notably higher than that of the uncoated electrode. This highlights the effectiveness of carbon coating in enhancing the cycling stability of NaCrO2 cathode materials.
Yu et al. [19] explored the effect of carbon coating on NaCrO2 cathodes and observed a significant enhancement in conductivity, increasing from 4.4 × 10−3 S cm−1 for the bare sample to 4.7 × 10−1 S cm−1 for a 14.6 wt% carbon-coated sample (Figure 12c). The initial charge–discharge profiles of bare NaCrO2 and carbon-coated NaCrO2 (C-NaCrO2, 3.4 wt%) were similar, except for the discharge capacities, which were 112 mAh g−1 and 121 mAh g−1, respectively. The carbon coating facilitated greater capacity delivery above 3.07 V during charging, attributed to a reversible phase transformation from an O3 to a P3 structure (Figure 12d).
Bare NaCrO2 exhibited significant capacity fading during cycling, while C-NaCrO2 electrodes demonstrated excellent stability, retaining over 96% of their capacity after 50 cycles for coatings ranging from 1.4 to 14.2 wt%. Among these, the 3.4 wt% C-NaCrO2 electrode displayed the best performance (Figure 12e).
The rate capability of C-NaCrO2 was particularly remarkable (Figure 12f). While bare NaCrO2 delivered only 3 mAh g−1 at a high rate of 50 C, the 3.4 wt% and 7.3 wt% C-NaCrO2 electrodes delivered capacities of 106 mAh g−1 at 50 C (5.5 A g−1), retaining 87% of the capacity at 0.5 C (55 mA g−1). The performance extended to an extremely high rate of 150 C (16.5 A g−1), with a 27 s discharge delivering 99 mAh g−1, demonstrating that the capacity was solely linked to Na+ insertion, as evidenced by the flat discharge profile (Figure 12g).
Table 2 compares the electrochemical performance of Cr-based cathode materials in SIBs to provide better clarification of their behavior and capabilities.

2.2. NaFeO2

Minor-metal-free LiFeO2 is unsuitable as a cathode material for LIBs due to its electrochemical inactivity. The oxidation of Fe3+ to Fe4+ is challenging because oxygen evolution precedes Fe3+ oxidation, unlike the Fe2+ to Fe3+ oxidation observed in LiFePO4 [71]. Additionally, the similar ionic radius of Fe3+ and Li+ promote cation mixing. An empirical rule suggests that layered AMO2 compounds are stabilized when the ionic radius ratio of the alkali metal (rA) to the transition metal (rM), rA/rM, exceeds 1.23. For LiFeO2 (rLi/rFe = 1.18), this criterion is unmet, leading to structural instability [72]. In contrast, NaFeO2 (rNa/rFe = 1.58) avoids such issues due to the larger Na+ ion. NaFeO2 exists as orthorhombic β-NaFeO2, which is electrochemically inactive, and rhombohedral α-NaFeO2, a cathode with O3-type layered structure per Delmas’ notation [73,74,75,76].
Okada and co-workers [73,77] were the first to report the reversible sodium cycling of α-NaFeO2, synthesized by solid-state reaction of equal molar ratio Fe2O3 and Na2O2 at 650 °C in air, in sodium-ion cells using a sodium metal counter electrode. Charging up to 3.6 V vs. Na yielded an initial charge capacity of 103 mAh g−1, corresponding to the formation of Na0.58FeO2, assuming no parasitic reactions. The cell discharged reversibly to 1.5 V vs. Na with an initial discharge capacity of 85 mAh g−1 and exhibited a flat operating voltage plateau at ~3.4 V. Sodium iron oxide has garnered interest as a cathode material for SIBs due to its high operating voltage (~3.3 V) based on the Fe3+/Fe4+ redox reaction and a capacity of ~80 mAh g−1. Iron’s abundance makes it appealing for SIBs; however, challenges arise at high potentials, including structural instability of the cathode and electrolyte decomposition. These issues cause rapid performance degradation due to irreversible structural transitions, particularly the formation of monoclinic Na0.5FeO2 during desodiation beyond x > 0.5 in Na1-xFeO2 [38,73,78,79].
Tetravalent iron (Fe4+) is rare but has been observed in certain perovskite oxides synthesized under high oxygen pressure. Mössbauer spectroscopy of α-NaFeO2 and its charged states reveals that it operates via the Fe3+/Fe4+ redox couple. The initial NaFeO2 exhibits high-spin Fe3+ characterized by an isomer shift (IS) of 0.36 mm s−1 and quadrupole splitting (QS) of 0.50 mm s−1, indicating a uniform Fe3+ valence state. Upon charging, an increase in QS suggests distortion of FeO6 octahedra. Signals for Fe4+ (IS = 0.05 mm s−1, QS = 0.55 mm s−1) were observed, with 18% of the iron identified as Fe4+ at 3.6 V, demonstrating the feasibility of sodium-ion deintercalation and Fe3+ oxidation (Figure 13a) [73,80]. Takeda et al. [81] corroborated these findings through Mössbauer analysis of NaFeO2, confirming Fe4+ formation during Na extraction.
The O3-NaFeO2 structure comprises Na and Fe octahedrally coordinated with oxygen, forming Na layers between O-Fe-O slabs (Figure 13b). During the initial charge to 4.4 V, the voltage profile exhibits plateaus at ~3.3 V and ~3.8 V, corresponding to Fe3+ oxidation and possible oxygen redox, respectively. Approximately 64% of Na-ion extraction (150 mAh g−1) occurs during the 3.3 V plateau. However, significant polarization (≥1.0 V) and low coulombic efficiency (~42%) were observed during discharge, with voltage plateaus shifting to 2.5 V and 2.3 V. Subsequent cycles show diminished capacity and increased polarization, with the Fe3+/Fe4+ redox reaction restricted to the first cycle (Figure 13c). This degradation is attributed to severe structural changes, including Fe migration, highlighting the challenges of using NaFeO2 as a cathode material.
Yabuuchi et al. [82] investigated the influence of cutoff voltage on the structural and electrochemical performance of the α-NaFeO2 cathode in sodium-ion cells. Their findings indicate that the reversibility of the electrode material is significantly affected by the cutoff voltage during charging (sodium extraction), as illustrated in Figure 14a. While a higher cutoff voltage increased the sodium extraction capacity, it simultaneously reduced the reversible capacity. Optimal performance, with minimal polarization, was achieved at a cutoff voltage of 3.4 V. Beyond this value, charging to 4.0 V or higher significantly increased polarization and reduced reversible capacity, rendering the material nearly inactive as an electrode at higher potentials. Approximately 70% of sodium can be extracted from O3-type NaFeO2 by oxidation up to 4.5 V, assuming no side reactions. However, this degree of sodium extraction compromises the electrode’s electrochemical activity and structural integrity (Figure 14a). The cyclability of the O3-type NaFeO2 at various cutoff voltages was further examined (Figure 14b,c). Cells charged to different voltages between 3.4 and 4.5 V, then discharged to 2.5 V at a current rate of 12.1 mA g−1, demonstrated varying degrees of performance. A cutoff voltage of 3.4 V maintained good capacity retention, with an initial reversible capacity of 80 mAh g−1 and approximately 75% capacity retention after 30 cycles (Figure 14c). Increasing the cutoff voltage to 3.5 V enhanced the first-cycle reversible capacity to 100 mAh g−1 but slightly reduced long-term capacity retention compared to the 3.4 V cycle. Further increases in the cutoff voltage to 4.0 V or higher led to significant capacity fade and poor overall cyclability.
Ex situ XRD analysis was conducted to elucidate the mechanism underlying the poor reversibility observed at higher cutoff voltages (Figure 14d). At a cutoff voltage of 3.4 V, sodium ions were reversibly extracted in a topotactic manner. An increase in the intensity of the (003)hex diffraction line relative to the (104)hex line after cycling between 3.4 and 2.5 V indicates the formation of sodium-ion vacancies at 3b octahedral sites. In O3-type sodium-layered compounds, phase transitions from O3 to P3 are commonly observed during sodium extraction when x > 0.1 in Na1-xMeO2. These transitions, caused by sodium/vacancy ordering at 3b sites, lead to stepwise voltage changes during galvanostatic oxidation/reduction. However, in Na1-xFeO2, the cell voltage increases monotonically with sodium extraction.
Irreversible structural changes occur when the charging voltage exceeds 4.0 V (x > 0.5 in Na1-xFeO2). At 4.5 V, the irreversible capacity increases significantly, likely due to the formation of additional vacancies at 3b sites. This is accompanied by a reduction in the intensity of the (003)hex diffraction line relative to the (101)hex and (104)hex lines. Such changes suggest possible iron migration from 3a to 6c (face-shared tetrahedral to octahedral sites) or to 3b sites, assuming no oxygen release from the lattice. This migration of iron ions may block sodium conduction pathways, contributing to the observed decline in electrode performance.
The chemical instability of Fe4+ in NaFeO2 cathode materials further exacerbates performance issues. Lee et al. [83] demonstrated that the electrochemical performance of NaFeO2 significantly deteriorates when stored in a charged state. In their study, a coin cell charged to 3.6 V, stored for 24 h, and then discharged to 2.0 V exhibited a much lower specific capacity (62.7 mAh g−1) and average cell voltage compared to a cell discharged immediately after charging (80 mAh g−1). This decline is attributed to the chemical instability of Fe4+, which spontaneously reduces Fe3+ in the charged state, accompanied by electrolyte decomposition.
Given that batteries are often stored in fully or partially charged states, this instability leads to significant self-discharge and rapid performance degradation. Addressing the chemical instability of Fe4+ is thus critical to unlocking the potential of Fe3+/Fe4+-based redox cathodes.
The thermal stability of NaFeO2 is promising compared to cathodes used in LIBs. When NaFeO2 was heated to 450 °C, no exothermic reactions were observed. In contrast, exothermic reactions were reported for Li0.5CoO2—a widely used cathode material in portable devices—at temperatures of 190 °C and 395 °C. However, it is noteworthy that heat generation occurred in the case of charged NaFeO2 (Na0.58FeO2) above 350 °C, as observed in differential scanning calorimetry (DSC) measurements, according to reaction 3 [73]:
Na0.58FeO2 → 0.58 NaFeO2 + 0.21 Fe2O3 + 0.105 O2
Elemental doping has proven to be an effective strategy for enhancing the performance of NaFeO2 cathode materials [84]. Xu et al. [85] incorporated Ru as a doping element into NaFeO2, demonstrating that Ru-doped NaFeO2 (Na4FeRuO6) delivered a high reversible capacity of 120 mA h g−1 and achieved approximately 80% capacity retention after 100 cycles when charged up to 4 V. Furthermore, Ru doping significantly reduced potential polarization in NaFeO2.
The typical charge–discharge cycles of Na4FeRuO6 and NaFeO2 are presented in Figure 15a,b. NaFeO2 delivered about 85 mAh g−1 reversible capacity at 0.2 C during the initial charge–discharge cycle, accompanied by a low coulombic efficiency of 81.7%. In contrast, Na4FeRuO6 exhibited a reversible capacity of 120 mA h g−1 within a voltage range of 2–4 V, with a much higher coulombic efficiency of 96.7%, indicating improved reversibility and reduced polarization.
Na4FeRuO6 demonstrated superior rate capability, delivering capacities of 120 mAh g−1 at 0.2 C, 96 mAh g−1 at 0.5 C, 88 mAh g−1 at 1 C, 75 mAh g−1 at 2 C, and 55 mAh g−1 at 5 C. In contrast, NaFeO2 showed poor rate performance, with negligible capacity at 2 C (Figure 15c,d). The improved performance of Ru-doped NaFeO2 is attributed to the suppression of Fe-ion migration, which reduces internal strain and mitigates particle cracking associated with the structural distortions caused by Fe-ion migration.
Co-doping in NaFeO2 has been shown to significantly enhance its electrochemical performance, as demonstrated by Yoshida et al. [86]. The study revealed that although NaFe0.5Co0.5O2 is considered an intermediate solid solution between NaFeO2 and NaCoO2, it delivers the highest reversible capacity of approximately 160 mAh g−1 among the three materials tested. Notably, the average operating voltage of NaFe0.5Co0.5O2 (3.14 V) in Na cells exceeds that of NaCoO2 (2.96 V) (Figure 15e).
Additionally, both NaFe0.5Co0.5O2 and NaCoO2 exhibit relatively good capacity retention over cycling, as illustrated in Figure 15f. A remarkable finding is the superior rate capability of NaFe0.5Co0.5O2, which significantly outperforms NaCoO2. Even at a high rate of 30 C (7.23 A g−1), micrometer-sized NaFe0.5Co0.5O2 achieves a discharge capacity exceeding 100 mAh g−1, highlighting its excellent fast-charging and high-rate performance.
Shevchenko et al. [87] investigated the effect of Mn and Ni substitution on the performance of NaFeO2-based cathodes, synthesizing NaFe1/3Mn1/3Ni1/3O2 to reduce Fe ion migration into the Na layers. The electrochemical performance of this cathode material was evaluated by charging to 4.0 V and 4.2 V and discharging to 1.9 V. At a C/10 rate, the material delivered specific capacities of approximately 130 mAh g−1 and 160 mAh g−1, respectively. However, charging up to 4.2 V introduced a plateau around 4.1 V, which led to capacity degradation over cycling due to structural instability.

2.3. NaCuO2

Copper is abundant, relatively inexpensive, and among the less toxic transition metals. Despite its favorable properties, there have been no reported electrochemical studies on LiCuO2. Efforts to synthesize LiCuO2 via solid-state calcination have been unsuccessful, as the reaction results in the formation of Li2CuO2 and CuO, as shown below:
2 CuCO3 + Li2CO3 → Li2CuO2 + CuO + 3 CO2
In contrast, NaCuO2 has been successfully synthesized using the solid-state calcination of Cuo and Na2O2 at 450 °C in oxygen environment. This behavior parallels that of nickelates, where NaNiO2 can be readily prepared via solid-state calcination, while stoichiometric LiNiO2 is notably challenging to obtain. Structural studies reveal that NaCuO2 adopts a monoclinic structure with the same space group and elemental arrangement as LiCuO2, as illustrated in Figure 16a. This structural similarity indicates that NaCuO2 may hold promise as a positive electrode material for sodium-ion batteries.
Figure 16b presents the discharge–charge profiles of a Na/NaCuO2 cell from the first to the fifth cycle within the voltage range of 0.75 to 3.0 V [88].
Theoretically, sodium-ion insertion and extraction in NaCuO2 correspond to the reduction of Cu3+ to Cu2+ during discharge and the oxidation of Cu3+ to Cu4+ during charge, respectively.
NaCu3+O2 + x Na+ + x e → Na1+xCu2+O2
NaCu3+O2 → Na1-xCu4+O2 + x Na+ + x e
The subsequent charge capacity of NaCuO2 reached 142 mAh g−1, nearly equivalent to its initial discharge capacity. Inductively Coupled Plasma Atomic Emission Spectroscopy (ICP-AES) analysis indicated that sodium-ion insertion and extraction during the first discharge and charge processes corresponded to 0.8 mol and 0.5 mol, respectively. In the case of LiCuO2, lithium-ion insertion occurs at approximately 2.0 V (vs. Li), leading to the formation of Li1.7CuO2.
Despite stable discharge–charge plateaus over the first five cycles, a noticeable decline in discharge capacity was observed with continued cycling. This significant capacity fading suggests that structural changes occur in NaCuO2 during repeated electrochemical cycling, ultimately impacting its performance. Ono et al. [89] investigated the poor cycling performance of NaCuO2 in the voltage range of 1–3 V and identified the underlying mechanism. Sodium-ion insertion into NaCuO2 during the first discharge results in an irreversible conversion of NaCuO2 to an amorphous phase, Na2CuO2. Subsequent charge and discharge reactions proceed between these amorphous phases. However, the discharge product, Na2CuO2 (more accurately Na1.6CuO2), is unstable and undergoes decomposition, forming copper oxide (CuO) and sodium oxide (Na2O):
NaCuO2 + Na+ + e → Na2CuO2
Na2CuO2 → Na2O + CuO
subsequently, CuO reacts with sodium ions, producing Cu2O and additional Na2O:
CuO + Na + + e 1 2 Cu 2 O + 1 2 Na 2 O
These irreversible reactions contribute to the significant capacity fading observed in NaCuO2 during repeated electrochemical cycling.

2.4. NaNiO2

The structure of NaNiO2 was first identified in 1954 using film methods with a precession camera and later refined through Rietveld analysis with X-ray and neutron scattering in 1977. Subsequent studies in 2000 and 2005 employed neutron and X-ray diffraction, respectively, to further understand its structure [90,91,92,93]. NaNiO2 exhibits two polymorphs: a low-temperature monoclinic O3 layered structure and a high-temperature rhombohedral phase. The low-temperature phase features NiO6 octahedra that are edge-sharing and elongated due to the Jahn–Teller distortion of the Ni3+ ions. Sodium ions occupy interlayer positions, forming distorted octahedral coordination with oxygen atoms, and this phase is isostructural with NaMnO2. At approximately 450 K, thermal motion overcomes the static Jahn–Teller distortion, causing a phase transition to the rhombohedral form, which is isostructural with α-NaFeO2 and LiNiO2. The monoclinic NaNiO2 polymorph was synthesized by solid-state reaction of excess amounts of Na2O and NiO at 650 °C under O2 flow and tested as a cathode material, exhibiting a stepwise charge–discharge profile with an initial open-circuit voltage of 2.47 V. Vassilaras et al. [94] demonstrated, in contrast to previous studies (which reported during the first electrochemical cycle, only 0.2 sodium ions could be extracted from NaNiO2 [94,95]), a charge capacity of 199 mAh g−1 (x = 0.85 in NaxNiO2) and a discharge capacity of 147 mAh g−1 (x = 0.62 in NaxNiO2) within the voltage range of 2.0–4.5 V. Galvanostatic charge and discharge measurements of NaNiO2 at C/10 (1 C = 235 mAh g−1) were conducted over multiple cycles in two different voltage ranges: 1.25–3.75 V and 2.0–4.5 V. The results are illustrated in Figure 17a,b.
The charge and discharge curves of NaNiO2 exhibit the expected step-like features, indicating distinct phase transitions during sodium extraction and insertion. Except for the first charge cycle, the similarity in the plateaus suggests consistent phase transitions throughout subsequent cycling. In the voltage range of 1.25–3.75 V (Figure 17a), the charge and discharge capacities for the first, second, tenth, and twentieth cycles are 147/123, 121/119, 119/118, and 117/116 mAh g−1, respectively, with corresponding coulombic efficiencies of 83.8%, 98.4%, and 99.2% for the first, second, and subsequent cycles [94].
When the voltage range is extended to 2.0–4.5 V (Figure 17b), the charge and discharge capacities for the first, second, tenth, and twentieth cycles are 199/147, 136/128, 127/107, and 113/98 mAh g−1, respectively. The coulombic efficiencies under this range are 73.7%, 93.7%, 84.6%, and 86.3% for the first, second, tenth, and twentieth cycles, respectively. A significant decline in capacity and coulombic efficiency is observed at higher voltages, indicative of side reactions and structural instability.
The first charge curve deviates from subsequent charge cycles, particularly below 3.0 V. However, above 3.0 V, the charge profiles align with those of later cycles, suggesting that excess capacity in the first cycle is associated with reactions at lower voltages. Cyclic voltammetry analysis confirms the capacity decline and differences between charge and discharge profiles over extended cycling at higher voltages (Figure 18a). The CV curves reveal four reversible oxidation and reduction peaks at 2.7/2.3, 3.1/2.9, 3.4/3.2, and 3.5/3.4 V. Beyond the first cycle, subsequent scans show consistent reversibility within the 2.0–4.0 V range.
The oxidation peaks between 2.5 and 3.0 V during the first cycle differ from those in later cycles, corroborating the galvanostatic cycling data. Above 4.0 V, oxidation processes appear partially reversible, as no corresponding reduction peaks are observed near 4.0 V. These observations suggest that side reactions at higher voltages contribute to poor cycling performance. Both galvanostatic and CV results highlight distinct reactions occurring in the 2.5–3.0 V range during the initial charge cycle, which are absent in subsequent cycles.
To further understand structural changes after the first cycle, XRD analysis was performed, and the diffraction patterns of three samples are compared in Figure 18b.
The bottom XRD pattern (black) corresponds to the pristine NaNiO2 sample. The middle pattern (red) represents the XRD of the as-prepared electrode, which consists of NaNiO2 mixed with PTFE and carbon black. In addition to the pristine NaNiO2 phase, a secondary phase, identified as Na0.91NiO2, is observed. This indicates a small loss of sodium during electrode preparation, likely caused by air exposure or interactions with the binder and carbon black. The top pattern (blue) shows the XRD of the electrode after discharge following cycling between 2.0 and 3.75 V. Only the Na0.91NiO2 phase is present, with no evidence of the original NaNiO2 phase, suggesting that the discharge process does not fully restore the sodiated NaNiO2 phase.
Further analysis was conducted at higher voltages to investigate the observed capacity reduction. XRD patterns recorded after charging NaNiO2 to 3.75 V and 4.5 V are displayed in Figure 18c. While the majority of the patterns remain similar, the highly charged sample exhibits two additional peaks at approximately 23.7° and 35.0°. These peaks indicate the formation of a new phase, although its structure remains unidentified [94].
Han et al. [96] revisited the phase transitions of NaNiO2 during charge–discharge cycling within a voltage range of 1.5–4.0 V at a rate of C/10 (Figure 19a). During the first charge, four distinct plateaus were observed, reflecting multiple phase transitions. These transitions are attributed to gliding of the oxide layers, variations in structural distortion due to sodium deintercalation, and rearrangement of sodium ions within the interlayer space. The OCVs associated with these plateaus are 2.59 V, 3.04 V, 3.38 V, and 3.53 V, respectively.
During discharge, four stepwise plateaus similar to those during charge were observed, indicating that the phase transitions are highly reversible throughout the electrochemical process. However, the first discharge capacity is notably lower than the first charge capacity, a well-documented phenomenon attributed to incomplete re-sodiation during the first discharge. This results in significant capacity loss.
Additionally, the lengths of the plateaus during charge and discharge are approximately identical, except for the first plateau in both processes. A pronounced voltage drop below 2.3 V suggests that a highly unfavorable phase transition is necessary to achieve the fully sodiated phase.
Real-time in situ XRD measurements of a NaNiO2 powder electrode were performed within the 2θ ranges of 15–18° and 31–38°, as shown in Figure 19b. The initial scan clearly identifies the pristine material as pure NaNiO2. The initial scan identifies the pristine material as pure NaNiO2. However, during the charging process, the fully sodiated phase diminishes rapidly, and partially desodiated phases emerge, evidenced by a shift of the (001) reflection toward lower 2θ angles. This shift corresponds to an increase in the c-spacing as sodium ions are deintercalated. The removal of positively charged sodium ions enhances electrostatic repulsion between the oxide layers, leading to structural expansion. Concurrently, the (−201) reflection shifts toward a smaller d-spacing, indicating contraction along the a- and b-axes. This shrinkage is attributed to the increasing concentration of smaller Ni4+ ions during oxidation.
As the charge progresses, new biphasic regions appear. Between sodium contents of ~1.0 and 0.74 per formula unit, O′3, P′3, and O’’’’3 phases emerge. The transition between these phases is accompanied by increased distortion, evidenced by changes in lattice parameters. Upon reaching a sodium concentration of 0.55 at 3.26 V, most O′3 and P′3 phases disappear, and the P″3 phase dominates. The a/b ratio increases from 1.73 to 1.76, reflecting enhanced structural distortion, though the β angle remains stable at 121.5°, preserving hexagonal symmetry.
Further oxidation at 3.38 V (sodium content ~0.55 to 0.42) reveals a new plateau corresponding to the coexistence of P″3, P‴3, and O″3 phases. During this stage, the P″3 phase transitions into a more distorted P3-like structure, before the formation of the O″3 phase. The newly identified P‴3 phase exhibits greater distortion than P″3, though it does not conform to the O″3 structural model. As the P‴3 phase transitions to O″3, the a/b ratio remains relatively stable, but the β angle shifts significantly from 119.75° to 106.0°, marking a transition from hexagonal to monoclinic symmetry. This phase evolution, driven by increased interlayer repulsion due to extensive sodium removal, exemplifies the transition from P-type to O-type phases.
A small biphasic region is observed between O″3 and O‴3 when the sodium content decreases from 0.42 to 0.40. This quickly transitions to a solid solution of O‴3. By 3.74 V in the relaxed in situ XRD scans, peaks corresponding to the O′3, P′3, P″3, and O″3 phases are nearly absent, leaving only the O‴3 phase detectable at sodium concentrations as low as 0.37. The lattice parameters ‘a’ and ‘c’ represent the diagonal and edge lengths of the MO6 octahedron, respectively, and relate to the hexagonal lattice parameters as follows:
a (monoclinic) = √3 × a (hexagonal)
b (monoclinic) = a (hexagonal)       
Pristine NaNiO2 exhibits an a/b ratio of 1.88 and a β angle of 110.45°, indicative of a significant Jahn–Teller distortion caused by the low-spin t62g e1g electronic configuration of Ni3+ [92]. This distorted monoclinic phase is stable at room temperature, while the ideal rhombohedral phase stabilizes only above 450 K, where thermal energy overcomes the Jahn–Teller effect.
During discharge, the phase transitions and voltage steps mirror those observed during charging. However, as the discharge voltage approaches 2.3 V, polarization increases, indicating that sodium re-intercalation into the host structure becomes kinetically unfavorable near full sodiation. Similar observations were made by Wang et al. [97], who reported irreversible structural changes in NaNiO2 during charge–discharge cycling in the 2–4.5 V range. Two-dimensional phase mapping revealed that these structural changes significantly hinder the electrochemical reversibility of the material.
Notably, a steep voltage drop below 2.3 V and the emergence of an O’’’’3 phase (marked with an asterisk in Figure 19b) during discharge highlight the difficulty in achieving full sodiation. This behavior underscores the structural and kinetic limitations of NaNiO2 during prolonged cycling.
Kim et al. [98] investigated the thermal stability of Na0.91NiO2 and Na0.5NiO2 in the presence of an electrolyte within the temperature range of 300–600 °C, as shown in Figure 20a–d. Two phase transitions were observed in both samples during heating: the first transition occurred between 247 and 352 °C, and the second between 528 and 600 °C (Figure 20a,c). Both samples exhibited the formation of a Ni metal phase and a structural transition from a layered structure to rock salt (Figure 20b,d).
However, a key difference was identified between the two samples. The onset temperature for phase decomposition was lower for the desodiated sample (Na0.5NiO2), indicating that increased Na extraction from the structure reduces the decomposition temperature.
A substantial number of studies have explored the use of Ni as a primary transition metal in SIB cathode materials, often combined with reinforcing elements such as Ti, Mn, and Co. Various compositions have been successfully synthesized and reported, including: NaFe0.2Ni0.4Ti0.4O2, NaFe0.4Ni0.3Ti0.3O2 [99], NaNi0.5Mn0.5O2 [100], NaNi0.5Mn0.3Ti0.2O2 [101], NaNi0.8Co0.15Al0.05O2 [102], NaNi0.815Co0.15Al0.035O2 [103], NaNixFeyMnzO2 [104], NaNixMn1-xO2 [105], NaFe0.5Ni0.5O2 [106,107], Na0.8[Ni0.3Co0.2Ti0.5]O2 [108], NaNi0.5Ti0.5O2 [107,109], Na(Ni2/3Sb1/3)O2 [110], NaNi0.9Ti0.1O2 [111], and NaxNi0.6Co0.4O2 [112] have been successfully reported as cathode materials in SIBs.
NaNi0.5Mn0.5O2 has been widely studied as a promising cathode material for sodium-ion batteries (SIBs) [113,114,115,116,117]. However, the O3-NaNi0.5Mn0.5O2 phase undergoes multiple complex phase transitions (O3-O3′-P3-P3′) and exhibits multiple voltage platforms during the charge and discharge processes. These transitions result in significant unit cell volume changes, structural collapse, and sluggish kinetic performance [100].
Ti substitution for Mn has been shown to increase interslab distances and suppress phase transformations in the high-voltage region [101,118]. Yuan et al. [117] reported that La doping in the structure of NaNi0.5Mn0.5O2 improves stability and rate capability by reducing particle size, stabilizing the crystal structure, and enhancing Na-ion diffusion. This enhancement is attributed to the formation of strong La-O bonds, which expand the Na interlayer spacing.
Liang et al. [114] developed a core–shell structure to improve the cycling and rate performance of NaNi0.5Mn0.5O2. In this design, a highly stable P2-Na2/3MnO2 coating was deposited on the high-capacity NaNi0.5Mn0.5O2 shell. The reinforced structure demonstrated significant performance improvements, retaining 85.3% of its capacity after 150 cycles at 1C (125 mAh g−1), compared to 74.8% for the bare sample. Additionally, the core–shell structure exhibited excellent high-rate capability, delivering 103 mAh g−1 at 15C, which is notably higher than the 82 mAh g−1 achieved by the uncoated sample.
Meghnani et al. [119] developed a Na3PO4 coating on NaNi0.815Co0.15Al0.035O2 particles to enhance air stability and mitigate the formation of insulating compounds such as NaOH and Na2CO3. This modification improved the material’s viability as a cathode for SIBs. The Na3PO4 coating contributed to a highly stable structure, improved sodium intercalation/deintercalation ability, and facilitated reversible phase transformations during the discharge process. These factors collectively resulted in promising electrochemical performance of Na3PO4@NaNi0.815Co0.15Al0.035O2 over extended cycling, as demonstrated in Figure 21a,b.
Li et al. [120] synthesized O3-type Ni-rich NaNi2/3Mn1/6Fe1/6O2 (NNMF) at various calcination temperatures (800 °C, 830 °C, 850 °C, and 880 °C) as a promising cathode material for SIBs. The first-cycle charge–discharge curves of NNMF samples, tested within a voltage range of 1.5–4.2 V at 0.2 C (1 C = 160 mA g−1), are shown in Figure 21c. The first discharge-specific capacities for NNMF-800, NNMF-830, NNMF-850, and NNMF-880 were 113, 117, 226, and 58 mAh g−1, respectively.
The cycling performance after 100 cycles, as shown in Figure 21d, revealed significant differences among the samples. The NNMF-850 electrode exhibited a gradual decrease in discharge capacity from 226 mAh g−1 to 158 mAh g−1, corresponding to a capacity retention rate of ~70%, demonstrating excellent cycling stability. In contrast, NNMF-800, NNMF-830, and NNMF-880 showed capacity retention rates of 40%, 46%, and 41.7%, respectively, indicating poor cycling stability.
Figure 21e illustrates the discharge performance of the electrodes at various current densities. The NNMF-850 electrode displayed superior rate performance compared to the other samples. For NNMF-850, the discharge capacities at 0.2 C, 0.5 C, 1 C, and 2 C were 225, 193, 168, and 145 mAh g−1, respectively, and the capacity recovered to 193 mAh g−1 upon returning to 0.2 C, demonstrating excellent reversibility. On the other hand, the discharge capacities of NNMF-800, NNMF-830, and NNMF-880 decreased more significantly as the current density increased, indicating inferior rate capabilities.
The superior performance of NNMF-850 is attributed to the optimization of its crystallinity and morphology, which enhances its structural stability and Na+ diffusion, ensuring excellent cycling stability and rate capability.

2.5. NaCoO2

Investigations into NaxCoO2 (0.5 ≤ x ≤ 1) have identified four distinct phases: O3 (0.9 ≤ x ≤ 1, α phase), O′3 (x = 0.75, α′ phase), P3 (0.55 ≤ x ≤ 0.6, β phase), and P2 (0.64 ≤ x ≤ 0.74, γ phase). These phases depend on non-stoichiometry and synthesis conditions (The most common synthesis method is solid-state reaction of (non)stoichiometric Co3O4 and Na2CO3 at different temperatures, based on phase type). Notably, the P2 phase demonstrates superior reversibility as a cathode material and enhanced thermoelectric power [121,122,123,124].
Among these, the O3 phase demonstrates superior energy density at voltages below 3.5 V [125,126,127]. However, O3 materials are strongly hygroscopic, which poses challenges for processing in air.
P2 compounds, on the other hand, have garnered significant attention due to their higher sodium mobility compared to O3 compounds, as demonstrated in computational and experimental studies. The P2 structure forms in the stoichiometry NaxMO2 when 0.6 < x < 0.8, while the O3 phase is more stable at full sodiation (x = 1.0). Consequently, synthesizing P2 compounds requires a higher average transition metal oxidation state (M3.2+/M3.4+) than for O3. As a result, only three single-metal P2 compounds (M = V, Mn, and Co) have been synthesized to date, while others are obtained in binary or ternary systems incorporating high-valent ions. However, these high-valence elements often contribute minimally to the electrochemical performance, either being inactive or adding only to the low-voltage capacity [37]. The charge–discharge curves of O3, P3, and P2 sodium cobalt oxide are shown in Figure 22a. Despite differences in the stacking sequences of CoO2 layers, NaxCoO2 phases exhibit multiple voltage plateaus. Large reversible capacities are achieved for all three phases: 140 mAh g−1 for the O3 phase, 130 mAh g−1 for the P3 phase, and 120 mAh g−1 for the P2 phase.
Layered NaCoO2 stands out due to its excellent structural stability and superior cyclability compared to LiCoO2. The sodium-ion diffusion coefficient of NaxCoO2 (0.5–1.5 × 10−10 cm2 s−1) is comparable to that of lithium-ion diffusion in LiCoO2 (1 × 10−11 cm2 s−1) [128]. In LiCoO2, excessive lithium-ion removal can lead to a structural transformation from hexagonal to tetragonal, causing capacity degradation. In contrast, NaCoO2 benefits from sodium’s larger ionic radius, preventing occupation of tetrahedral sites, which avoids structural deformation and lowers the potential barrier for ion migration.
Among the various phases of NaxCoO2 (0.5 ≤ x ≤ 1), the P2 phase (0.6 ≤ x ≤ 0.74) exhibits superior reversibility as a cathode material due to its large trigonal prismatic sites that enhance sodium-ion intercalation and improve cycle life. However, its complex phase transitions—caused by the rotation of MO6 octahedra and M-O bond breaking—can also enhance its cyclic performance, making the P2 phase a focus of recent research [129].
Figure 22. (a) Comparison of typical charge–discharge curves of Na cells with layered polymorphs: O3-, P′3-, and P2-type NaxCoO2. The curves show similar profiles across the three polymorphs, except in the region with high sodium content (x > 0.7), where differences become apparent. Reprinted from Ref. [130] with permission from the American Chemical Society, (b) galvanostatic charge curves and structural evolution of Na1-xCoO2, where detailed structural changes have not been reported in the monoclinic P′3 phase above 3.5 V. Reprinted from Ref. [38] with permission from Wiley, (c) rate capability of NaCoO2, indicating high capacity at low rates but diminished performance at higher current densities due to phase boundary movements and Na+/vacancy ordering. Reprinted from Ref. [86] with permission from Elsevier, and (d) evolution of the chex parameter as a function of sodium content in P2-NaxCoO2, showing its correlation with the electrochemical potential (red curve) and structural parameters in both single-phase (filled blue circles) and biphasic domains (open circles). In single-phase regions, the chex parameter shows an approximately linear evolution with sodium content (red dashed line), though deviations occur near phase composition boundaries, especially close to biphasic domains. Key single-phase compositions, corresponding to specific electrochemical transitions (drops no. 3, 7–9), are highlighted as red circles. Reprinted from Ref. [127] with permission from Nature.
Figure 22. (a) Comparison of typical charge–discharge curves of Na cells with layered polymorphs: O3-, P′3-, and P2-type NaxCoO2. The curves show similar profiles across the three polymorphs, except in the region with high sodium content (x > 0.7), where differences become apparent. Reprinted from Ref. [130] with permission from the American Chemical Society, (b) galvanostatic charge curves and structural evolution of Na1-xCoO2, where detailed structural changes have not been reported in the monoclinic P′3 phase above 3.5 V. Reprinted from Ref. [38] with permission from Wiley, (c) rate capability of NaCoO2, indicating high capacity at low rates but diminished performance at higher current densities due to phase boundary movements and Na+/vacancy ordering. Reprinted from Ref. [86] with permission from Elsevier, and (d) evolution of the chex parameter as a function of sodium content in P2-NaxCoO2, showing its correlation with the electrochemical potential (red curve) and structural parameters in both single-phase (filled blue circles) and biphasic domains (open circles). In single-phase regions, the chex parameter shows an approximately linear evolution with sodium content (red dashed line), though deviations occur near phase composition boundaries, especially close to biphasic domains. Key single-phase compositions, corresponding to specific electrochemical transitions (drops no. 3, 7–9), are highlighted as red circles. Reprinted from Ref. [127] with permission from Nature.
Micromachines 16 00137 g022
Yoshida et al. [86] reported that O3-NaCoO2 delivers a discharge capacity of over 140 mAh g−1 within a voltage range of 2.5–4.0 V, featuring a stepwise voltage profile. This profile is attributed to multiple reversible phase transitions involving CoO2 slab gliding and in-plane vacancy-Na ordering (Figure 22b). The capacity loss during cycling in O3-NaCoO2 is likely due to the degradation of the electrolyte solution, exacerbated by the non-passivated surface of metallic sodium used as the counter electrode. Furthermore, the voltage profile of NaxCoO2 suggests a limited rate capability, attributed to phase boundary movements in two-phase regions and Na/vacancy ordering coupled with charge ordering (Figure 22c).
Unlike LiCoO2, the P2-type NaxCoO2 exhibits a highly intricate electrochemical charge–discharge profile characterized by several voltage plateaus. These plateaus are attributed to Na+/vacancy ordering, which arises from in-plane patterning between Na+ and vacancy sites, a phenomenon absent in the layered structure cathodes of LIBs. Na+/vacancy ordering results from energy minimization involving three types of cation interactions: Na+-Na+, Na+-Co3+/4+, and Co3+/4+-Co3+/4+, occurring below a critical temperature. This ordering is highly sensitive to the sodium content in the host material and is accompanied by a first-order phase transition at specific sodium concentrations. Such transitions introduce additional activation energy barriers for Na+ accommodation. Berthelot et al. [127] investigated the structural changes in P2-NaxCoO2 cathode materials and observed that the chex parameter decreases during discharge. This decrease is attributed to enhanced structural cohesiveness upon sodium intercalation. Simultaneously, a slight increase in the ahex parameter-representing the M-M (metal-metal) distance within the slabs-is linked to the reduction of M4+ ions (Figure 22d). It can be observed that the voltage plateaus in the electrochemical curve correspond to biphasic transitions, during which the chex parameter remains constant for each phase. Conversely, in single-phase regions, the chex parameter evolves continuously with changes in sodium composition.
Similar to other sodium transition metal oxides, the doping of electrochemically (in)active elements has been explored as an effective strategy to enhance the performance of sodium cobalt oxide cathode materials. This approach aims to stabilize the structure, improve cycling stability, and optimize the electrochemical properties by mitigating issues such as phase transitions and structural degradation during sodium intercalation and deintercalation [37,131,132,133]. Kang et al. [134] investigated the effects of Ti doping in P2-type Na0.67Co1-xTixO2 (x < 0.2) cathodes under an extended potential range exceeding 4.4 V (Figure 23a–c). Their findings (Figure 23d) demonstrated that Na0.67Co0.90Ti0.10O2 exhibited excellent capacity retention of 115 mAh g−1 even after 100 cycles, while Na0.67CoO2 showed negligible capacity retention (<10 mAh g−1) under a 4.5 V cutoff voltage. Additionally, Na0.67Co0.90Ti0.10O2 displayed exceptional rate capabilities (Figure 23e), achieving 108 mAh g−1 at a current density of 1000 mA g−1 (7.4 C). These results suggest that the enhanced rate capability is primarily due to reduced kinetic resistance from Ti substitution, rather than the elimination of active sites for Na ions at high potential conditions above 4.4 V.
Ca doping has been identified as an effective way to enhance the cycling stability of P2-type NaxCoO2 cathode materials [135,136]. Han et al. [136] explored the cycling and rate performance of NaxCayCoO2 (0.45 ≤ x ≤ 0.64, 0.02 ≤ y ≤ 0.10) to understand the impact of Ca doping on cathode materials. The ionic radius of Ca2+ (1.00 A°) is considerably larger than that of Co3+ (0.61 A° for high spin) or Co4+ (0.53 A° for high spin), making it unlikely to replace Co directly within the CoO6 octahedra. Instead, Ca2+ ions are positioned between CoO6 slabs, where they strengthen the electrostatic interactions between these layers. While Ca doping reduces the initial discharge capacity, it significantly enhances the structural stability of Na0.73CoO2. Among the compositions studied, the Na0.60Ca0.07CoO2 cathode exhibited the highest capacity retention over extended cycling (Figure 23f) and superior rate performance (Figure 23g), particularly under high-current conditions.
Figure 23. (ac) Electrochemical cycling performance of Na/Na0.67Co1-xTixO2 cells with varying titanium substitution levels (x = 0, x = 0.05, and x = 0.1), illustrating specific discharge capacities as a function of cycle number, (d) specific capacity and coulombic efficiency over cycles for each composition, highlighting improved capacity retention and efficiency with titanium doping, (e) rate performance comparison of Na/Na0.67Co1-xTixO2 cells, emphasizing the enhanced rate capability with x = 0.05 and x = 0.1 substitutions. Reprinted from Ref. [134] with submission from the American Chemical Society, (f) discharge capacity retention of Na0.73CoO2 and calcium-doped variants Na0.68Ca0.02CoO2, Na0.63Ca0.05CoO2, Na0.60Ca0.07CoO2, and Na0.52Ca0.10CoO2 at 23.5 mA g −1, demonstrating capacity stabilization with optimized calcium doping levels, and (g) discharge (open markers) and charge (solid markers) capacities of Na0.73CoO2 (A) and Na0.60Ca0.07CoO2 (D) at varying current densities, where marked current values denote performance trends across rate capabilities. Reprinted from Ref. [136] with permission from Elsevier.
Figure 23. (ac) Electrochemical cycling performance of Na/Na0.67Co1-xTixO2 cells with varying titanium substitution levels (x = 0, x = 0.05, and x = 0.1), illustrating specific discharge capacities as a function of cycle number, (d) specific capacity and coulombic efficiency over cycles for each composition, highlighting improved capacity retention and efficiency with titanium doping, (e) rate performance comparison of Na/Na0.67Co1-xTixO2 cells, emphasizing the enhanced rate capability with x = 0.05 and x = 0.1 substitutions. Reprinted from Ref. [134] with submission from the American Chemical Society, (f) discharge capacity retention of Na0.73CoO2 and calcium-doped variants Na0.68Ca0.02CoO2, Na0.63Ca0.05CoO2, Na0.60Ca0.07CoO2, and Na0.52Ca0.10CoO2 at 23.5 mA g −1, demonstrating capacity stabilization with optimized calcium doping levels, and (g) discharge (open markers) and charge (solid markers) capacities of Na0.73CoO2 (A) and Na0.60Ca0.07CoO2 (D) at varying current densities, where marked current values denote performance trends across rate capabilities. Reprinted from Ref. [136] with permission from Elsevier.
Micromachines 16 00137 g023
Table 3 provides a summary of the capacities achieved with different phases of sodium cobalt oxide. These materials were all examined using non-aqueous electrolytes, such as organic or ionic solvents, with a sodium metal anode serving as the counter electrode.

2.6. NaVO2

Among the various reported NaxVO2 compounds, P2- and O3-type compositions have been thoroughly studied. The O3-type NaxVO2, obtained by reduction in NaVO3 under H2 atmosphere at 700 °C, demonstrates the highest initial discharge capacity (~140 mAh g−1) with full reversibility between Na0.5VO2 and Na1VO2 compositions, but suffers from poor cycling stability due to significant slab gliding during the charge–discharge process. The intermediate voltage window (1.6–2.6 V), although indicative of decent electrochemical performance, makes it less attractive for practical applications. Moreover, the O3 phase is sensitive to air, where oxygen slightly oxidizes the material, forming a small amount of an oxidized phase on the particle surface following the reaction:
NaVO2 + (1 − x)/4 O2 → NaxVO2 + (1 − x)/2 Na2O
Figure 24a,b illustrate the electrochemical performance of O3- and P2-NaxVO2 phases against Na+/Na. Figure 24a highlights a highly reversible Na insertion and de-insertion process between 1.2 and 2.4 V. The first charge curve shows a potential plateau indicative of a biphasic domain close to the Na2/3VO2 composition, followed by multiple potential plateaus associated with more complex behavior. Near the Na1/2VO2 composition, another single-phase potential step is observed. Figure 24b expands on this by showcasing the extended cycling behavior and capacity retention, underscoring the material’s electrochemical stability and phase transformation reversibility over successive cycles [141,142,143].
Despite these merits, the cell polarization remains significant, as shown in Figure 24c, even at low rates (C/100). Extending the voltage range beyond 2.5 V results in irreversible transitions, as only a small amount of Na can be reinserted (Figure 24d). The discharge curve shape changes significantly, similar to NaxTiO2 systems, attributed to Ti migration into Na layers. The structural changes are evidenced in Figure 24e, where XRPD patterns for the Na2/3VO2 and Na1/2VO2 phases are compared to the pristine NaVO2 phase.
The XRPD analysis reveals that for Na2/3VO2 (2/3 < x < 1), a mixture of starting O3-NaVO2 and Na2/3VO2 phases is formed. For smaller x values, NaxVO2 phases (1/2 < x < 2/3) crystallize in the C2/m space group, representing a distortion of the trigonal lattice in O3-NaVO2. The irreversible decrease in interslab distance suggests vanadium migration into vacancies caused by Na de-intercalation.
The P2-phase NaxVO2, while challenging to synthesize in a pure single-phase form, emerges as a more promising candidate. Its cycling behavior within the voltage range of 1.4–2.5 V (Figure 24b) shows minimal polarization and a reversible capacity of ~120 mAh g−1. However, at higher deintercalation voltages (V > 2.5 V), irreversible phase changes occur. The voltage–composition curve resembles the P2-NaxCoO2 system (Figure 24f), showing multiple potential plateaus (biphasic domains) and continuous voltage variations (solid-solution phases). Specific potential drops correlate with ordered Na/V ratios (1/2, 5/8, and 2/3), pointing to ordered Na ion arrangements and/or V3+/V4+ within the VO2 slabs [144,145,146,147].
Figure 24. (a) Voltage–composition curves for NaVO2 (O3), highlighting the reversible formation of Na1/2VO2 and Na2/3VO2 during Na de-intercalation and the onset of an electrochemically inactive phase at potentials above 2.4 V, (b) voltage–composition curves for Na0.7VO2 (P2), demonstrating its electrochemical behavior in the sodium-ion battery system. The inset figures show the delivered discharge capacity during limited cycling. Reprinted from Ref. [145] with permission from Elsevier, (c) electrochemical cycling of NaxVO2 (x = 0.92, as determined by lever rule) in the range 1.4–2.5 V at rate C/100. The differential curve dE/dx of the first charge is shown in the inset. NaxVO2 is a mixture of NaVO2 and Na2/3VO2, (d) electrochemical cycling of NaxVO2 (x = 0.95, as determined by lever rule) in the range 1.5–3.0 V at rate C/100, (e) X-ray powder diffractograms of the as-synthesized Na1VO2 phase and the electrochemically obtained Na2/3VO2 and Na1/2VO2 phases. Indexations in the corresponding space groups are shown. (*) is related to graphite peak in the XRD pattern. Reprinted from Ref. [146] with permission from IOP Publishing, (f) evolution of cell voltage as a function of sodium content in NaxVO2 over the 0.5 ≤ x ≤ 0.92 range. The limits of the biphasic domains and the solid solutions are shown by dashed lines. The three single phases for x = 1/2, 5/8, and 2/3 are emphasized by thick blue lines. The numbers 2, 3, and 4 indicate monophasic domains during sodium (de)intercalation. Reprinted from Ref. [144] with permission from Nature.
Figure 24. (a) Voltage–composition curves for NaVO2 (O3), highlighting the reversible formation of Na1/2VO2 and Na2/3VO2 during Na de-intercalation and the onset of an electrochemically inactive phase at potentials above 2.4 V, (b) voltage–composition curves for Na0.7VO2 (P2), demonstrating its electrochemical behavior in the sodium-ion battery system. The inset figures show the delivered discharge capacity during limited cycling. Reprinted from Ref. [145] with permission from Elsevier, (c) electrochemical cycling of NaxVO2 (x = 0.92, as determined by lever rule) in the range 1.4–2.5 V at rate C/100. The differential curve dE/dx of the first charge is shown in the inset. NaxVO2 is a mixture of NaVO2 and Na2/3VO2, (d) electrochemical cycling of NaxVO2 (x = 0.95, as determined by lever rule) in the range 1.5–3.0 V at rate C/100, (e) X-ray powder diffractograms of the as-synthesized Na1VO2 phase and the electrochemically obtained Na2/3VO2 and Na1/2VO2 phases. Indexations in the corresponding space groups are shown. (*) is related to graphite peak in the XRD pattern. Reprinted from Ref. [146] with permission from IOP Publishing, (f) evolution of cell voltage as a function of sodium content in NaxVO2 over the 0.5 ≤ x ≤ 0.92 range. The limits of the biphasic domains and the solid solutions are shown by dashed lines. The three single phases for x = 1/2, 5/8, and 2/3 are emphasized by thick blue lines. The numbers 2, 3, and 4 indicate monophasic domains during sodium (de)intercalation. Reprinted from Ref. [144] with permission from Nature.
Micromachines 16 00137 g024

2.7. NaMnO2

NaMnO2 is an intriguing sub-class of lithium transition metal oxides based on manganese, notable for its high theoretical capacity, environmental friendliness, and ease of synthesis. NaxMnO2 has been studied since the 1970s, offering a variety of stoichiometries and structures. Depending on sodium content, these materials can form double or triple chains creating 3D tunnel structures (Na/Mn < 0.5) or 2D layered structures (Na/Mn > 0.5). NaMnO2 exists in two main phases: at low temperatures, α-NaMnO2 adopts a distorted O3 layered structure with monoclinic distortion, attributed to the Jahn–Teller effect of Mn3+ ions and at high temperatures, it transforms into orthorhombic β-NaMnO2, a layered structure composed of MnO2 sheets made of edge-sharing MnO6 octahedra, where sodium ions occupy octahedral sites between adjacent MnO2 sheets. Density Functional Theory (DFT) calculations estimate similar formation energies for these phases, making it experimentally challenging to synthesize a pure β-NaMnO2 phase [148].
The charge–discharge characteristics of a non-aqueous Na/α-NaMnO2 cell at varying rates, obtained by solid-state reaction of Na2CO3 and Mn2O3 at 700 °C in air, are shown in Figure 25a,b. Almost 0.9 Na per formula unit is extracted during the first charge, delivering an initial charge capacity of 215 mAh g−1. Monoclinic NaMnO2 provides discharge capacities of 185 and 194 mAh g−1 within a voltage range of 2–3.8 V, achieving coulombic efficiencies of 86.4% and 92.8% at C/20 and C/30 rates, respectively (1 C = 244 mAh g−1). Coulombic efficiency decreases in subsequent cycles due to side reactions, likely involving the electrolyte, particularly at higher voltages. Lower efficiencies at C/30 suggest prolonged exposure to high voltage exacerbates parasitic reactions or degradation [149]. Overcharging beyond 3.8 V (up to 4.2 V at C/10) shows an additional charge capacity of ~90 mAh g−1 between 3.8 V and 4.2 V. However, this additional capacity is not recovered during discharge, indicating it arises from electrolyte decomposition rather than reversible Na-ion intercalation (Figure 25c). Monoclinic NaMnO2 exhibits stability issues, as shown in Figure 25d. Structural changes during sodium extraction from O′3-NaMnO2 generate internal stress, leading to capacity fading. This fading is linked to bulk structural changes and crystallite size rather than surface reactions. Gradual and irreversible changes in the charge–discharge profiles during cycling reflect the impact of bulk structural evolution [150].
Kubota et al. [150] investigated the phase transitions occurring in monoclinic-NaMnO2 during charging. They identified seven highly crystalline O′3-type phases (O′3(1) to O′3(7)) and less crystalline O′3-O1-like phases. The phase evolution proceeds as follows: In the first plateau around 2.6 V, O′3(2) gradually increases at the expense of O′3(1), undergoing a complete phase transition from O′3(1) to O′3(2). A similar two-phase transition process continues with progressive charging up to 3.8 V, involving seven O′3 phases and one O′3-O1 phase.
With progressive sodium extraction, electrostatic repulsion between adjacent MnO2 slabs increases, leading to an enlargement of the interslab distance. This trend persists throughout the desodiation process until the emergence of the O′3(7) phase. However, during the transition from O′3(7) to O′3-O1(8) at a voltage plateau around 3.6 V, the interslab distance decreases from 5.63 A° to 5.15 A° (Figure 25e). This transition, associated with decreased crystallinity and high Mn ion migration into the interslab space to form a modulated phase and/or the coexistence of differently oriented domains (Figure 25f), results in the formation of low-crystallinity O′3 phases, causing significant capacity degradation during cycling in a wider voltage range (above 3.55 V).
β-NaMnO2 exhibits a distinctive layered structure that sets it apart from the typical polymorphs of α-NaMnO2. Unlike the planar MnO6 octahedral layers in α-types, which adjust their stacking sequences to form polymorphs like O3, P2, or P3, β-NaMnO2 is characterized by zigzag layers of edge-sharing MnO6 octahedra. In this structure, Na+ ions occupy the octahedral sites between the zigzag MnO6 layers (Figure 26a). The electrochemical performance of β-NaMnO2, synthesized by firing a mixture of Mn2O3 and Na2CO3 first at 950 °C for 24 h, after a temperature ramp of 1 °C/min, and second at 950 °C for 24 h, after ramping the temperature at a faster heating rate of 5 °C/min in oxygen atmosphere, is demonstrated through the load curves in Figure 26b, showing Na extraction and reinsertion processes. During the first charge, a distinct plateau appears between the fully sodiated NaMnO2 phase and a desodiated phase close to Na0.57MnO2, with smaller steps observed at compositions around Na0.49MnO2 and Na0.39MnO2. This initial cycle experiences an irreversible capacity loss of approximately 25 mAh g−1 (equivalent to 0.1 Na per formula unit). Despite this initial loss, the load curves remain relatively stable over subsequent cycles, with only a slight reduction in capacity. A plateau at 2.7 V corresponds to a phase transition between Jahn–Teller distorted and undistorted structures, accompanied by minimal polarization (below 150 mV). The rate performance of β-NaMnO2 at varying charge–discharge rates (Figure 26c) highlights a gradual but irreversible capacity fade over time. Notably, the material does not recover its initial capacity, indicating some structural degradation during cycling, particularly under higher discharge rates (Figure 26d). These findings suggest that while β-NaMnO2 demonstrates interesting structural and electrochemical features, its cycling stability requires further investigation and optimization [151].
Figure 26. (a) Schematic representation of the β-NaMnO2 structure in the Pmnm space group. MnO6 octahedra are depicted in pink, NaO6 octahedra in yellow, and oxygen atoms in red. An intergrowth model on the right illustrates the structural relationship between α- and β-NaMnO2. Reprinted from Refs. [151,152] with permission from the American Chemical Society, (b) load curves for β-NaMnO2 during Na extraction/reinsertion cycles at a C/20 rate (10 mA g−1), showing the first (black), second (red), fifth (blue), and tenth (green) cycles, (c) cycling performance of β-NaMnO2 in the voltage range of 2–4.2 V vs. Na+/Na at various rates. Na extraction occurs at C/4, while reinsertion is performed at rates indicated above each dataset. For the C/20 dataset, both extraction and reinsertion occur at C/20, and (d) Specific discharge capacities of β-NaMnO2 over 100 cycles between 2 and 4.2 V vs. Na+/Na at room temperature. The plot illustrates capacity changes with cycling across different rates (C/20 to 10 C). Na extraction occurs at C/4, with reinsertion rates indicated by square markers: black (C/20), purple (C/4), green (C/4), and blue (10 C). Reprinted from Ref. [151] with permission from the American Chemical Society.
Figure 26. (a) Schematic representation of the β-NaMnO2 structure in the Pmnm space group. MnO6 octahedra are depicted in pink, NaO6 octahedra in yellow, and oxygen atoms in red. An intergrowth model on the right illustrates the structural relationship between α- and β-NaMnO2. Reprinted from Refs. [151,152] with permission from the American Chemical Society, (b) load curves for β-NaMnO2 during Na extraction/reinsertion cycles at a C/20 rate (10 mA g−1), showing the first (black), second (red), fifth (blue), and tenth (green) cycles, (c) cycling performance of β-NaMnO2 in the voltage range of 2–4.2 V vs. Na+/Na at various rates. Na extraction occurs at C/4, while reinsertion is performed at rates indicated above each dataset. For the C/20 dataset, both extraction and reinsertion occur at C/20, and (d) Specific discharge capacities of β-NaMnO2 over 100 cycles between 2 and 4.2 V vs. Na+/Na at room temperature. The plot illustrates capacity changes with cycling across different rates (C/20 to 10 C). Na extraction occurs at C/4, with reinsertion rates indicated by square markers: black (C/20), purple (C/4), green (C/4), and blue (10 C). Reprinted from Ref. [151] with permission from the American Chemical Society.
Micromachines 16 00137 g026
To mitigate phase transitions and enhance capacity retention in P2- and O3-type sodium manganese-based layered oxide cathode materials, extensive research has focused on elemental doping and transition metal combinations [107,153,154,155,156,157,158,159,160]. Li et al. [161] synthesized a sodium transition metal oxide, NaMn1/4Ni1/4Fe1/4Co1/4O2, featuring an equimolar ratio of Ni, Mn, Fe, and Co. This cathode material exhibited an initial discharge capacity of 180 mAh g−1, delivered at an average discharge voltage of 3.21 V (Figure 27a). Compared to NaFe0.5Co0.5O2 and NaMn0.5Ni0.5O2, the NaMn1/4Ni1/4Fe1/4Co1/4O2 cathode demonstrated superior cycling performance across both low and high upper cutoff voltages (Figure 27b,c). The improvements in capacity retention were attributed to the absence of multiple phase transitions (for example, from hexagonal to monoclinic structures) and the reduction in Jahn–Teller distortion of Mn ions.
Further advancements have been made by investigating Fe substitution within the NaMn1-xNi1-xFexO2 system. Yuan et al. [162] explored the effect of Fe substitution in NaMn0.5Ni0.5O2 cathode materials, finding that Fe doping successfully suppressed capacity decay during charge–discharge cycling. This enhancement was linked to improved structural stability, specifically the inhibition of MO2 layer gliding and phase conversion.
Additionally, Li et al. [163] developed a MoS2 coating for NaMn0.4Ni0.4Fe0.2O2 particles and studied its impact on the material’s electrochemical performance. The coating not only improved the material’s electrical conductivity but also enhanced its structural stability, even in air. After exposure to air for 10 days, the cathode retained a specific discharge capacity of 128.9 mAh g−1 (initially 148.4 mAh g−1 at 0.1 C within a 2–4.2 V voltage window). These improvements were attributed to the increased conductivity and reduced side reactions occurring in the electrolyte.
Figure 27. (a) Galvanostatic charge and discharge profiles at the 1st, 2nd, 5th, and 10th cycles for NaMn1/4Fe1/4Co1/4Ni1/4O2 (MFCN), (b) discharge capacity of MFCN cycled between 1.9 and 4.3 V at different cycles compared with NaMn0.5Ni0.5O2 (NM) cycled between 1.9 and 3.8 V and NaFe0.5Co0.5O2 (FC) cycled between 1.9 and 4.0 V, all at a C/10 rate. (c) Discharge capacity of MFCN cycled between 1.9 and 4.3 V at different cycles compared with NM and FC cycled between 1.9 and 4.2 V, all at a C/10 rate. Reprinted from Ref. [161] with permission from Elsevier, (d) charge–discharge curves for NaCu0.1Mn0.9O2 at a rate of C/20 (10 mA/g) in the voltage range of 2.0–4.0 V versus Na+/Na. The 1st, 2nd, 5th, and 10th Na extraction/reinsertion cycles are represented in black, red, blue, and green, respectively, (e) rate capability of the NaCu0.1Mn0.9O2 electrode material, (f) the capacity and coulombic efficiency versus cycle number at 10 mA g−1 for NaCu0.1Mn0.9O2, (g) the capacity versus cycle number at 100 mA g−1, 200 mA g−1, and 400 mA g−1 for NaCu0.1Mn0.9O2. Reprinted from Ref. [164] with permission from IOP Publishing.
Figure 27. (a) Galvanostatic charge and discharge profiles at the 1st, 2nd, 5th, and 10th cycles for NaMn1/4Fe1/4Co1/4Ni1/4O2 (MFCN), (b) discharge capacity of MFCN cycled between 1.9 and 4.3 V at different cycles compared with NaMn0.5Ni0.5O2 (NM) cycled between 1.9 and 3.8 V and NaFe0.5Co0.5O2 (FC) cycled between 1.9 and 4.0 V, all at a C/10 rate. (c) Discharge capacity of MFCN cycled between 1.9 and 4.3 V at different cycles compared with NM and FC cycled between 1.9 and 4.2 V, all at a C/10 rate. Reprinted from Ref. [161] with permission from Elsevier, (d) charge–discharge curves for NaCu0.1Mn0.9O2 at a rate of C/20 (10 mA/g) in the voltage range of 2.0–4.0 V versus Na+/Na. The 1st, 2nd, 5th, and 10th Na extraction/reinsertion cycles are represented in black, red, blue, and green, respectively, (e) rate capability of the NaCu0.1Mn0.9O2 electrode material, (f) the capacity and coulombic efficiency versus cycle number at 10 mA g−1 for NaCu0.1Mn0.9O2, (g) the capacity versus cycle number at 100 mA g−1, 200 mA g−1, and 400 mA g−1 for NaCu0.1Mn0.9O2. Reprinted from Ref. [164] with permission from IOP Publishing.
Micromachines 16 00137 g027
While significant research has been conducted on P2- and O3-type NaxMnO2 cathode materials, making them strong candidates for high capacity and stability, relatively few studies have addressed modifications of β-NaxMnO2. For instance, Jiang et al. [164] investigated the effects of Cu doping on the electrochemical performance of β-NaMnO2, specifically examining NaCu0.1Mn0.9O2. The charge–discharge curves, capacity retention, and rate performance are presented in Figure 27d–g. As shown in Figure 27d, NaCu0.1Mn0.9O2 exhibited an initial discharge capacity corresponding to 0.6 Na reversible intercalation/deintercalation with a coulombic efficiency of 94.1%, surpassing that of undoped β-NaMnO2. Remarkably, NaCu0.1Mn0.9O2 demonstrated excellent reversibility at the voltage plateau of ~3.5 V, unlike the irreversible behavior observed in β-NaMnO2.
Moreover, NaCu0.1Mn0.9O2 exhibited improved cycling stability, retaining 80.1% of its initial capacity after 100 cycles, compared to 68.4% retention in β-NaMnO2 (Figure 27f). It also showed superior rate performance (Figure 27e,g). This enhanced performance may stem from the absence of local structural rearrangements associated with planar defects in β-NaMnO2, which are mitigated by Cu doping.
Based on the discussion of transition metal oxide layered materials for SIBs, it is evident that Mn- and Ni-based compounds are excellent candidates for cathode materials. Wang et al. [116] investigated the substitution of both Li and Ti for Mn and Ni in the structure of NaMn0.5Ni0.5O2. The electrochemical performance of NaLi1/9Ni1/3Mn4/9Ti1/9O2 and NaMn0.5Ni0.5O2 is presented in Figure 28a,b.
Unlike the charge–discharge curves of NaMn0.5Ni0.5O2, which exhibit multiple voltage plateaus due to complex phase transitions—identified as the primary cause of capacity decay during cycling—the curves for NaLi1/9Ni1/3Mn4/9Ti1/9O2 are much smoother. This indicates reduced phase transitions, enabling more stable material operation.
Elemental substitution significantly improved the cycling stability of NaMn0.5Ni0.5O2. After 100 cycles at 1 C, the capacity retention increased from 46.97% for the pristine material to 80% for the substituted material, as shown in Figure 28c. The primary reason for this improvement is the expanded Na layer spacing, which enhances Na-ion diffusion and effectively suppresses complex phase transitions (Figure 28d,e).
Beyond conventional elemental doping strategies, structural modifications such as single crystalline and binary structures have been explored as an effective approach to improve transition metal layered oxide cathode materials [165,166]. Single-crystal structural synthesis has emerged as a promising solution to enhance capacity retention during cycling [167,168,169,170,171].
Darga et al. [113] reported that single-crystalline NaMn0.5Ni0.5O2 demonstrated significantly higher capacity retention compared to its polycrystalline counterpart, with 69% retention versus 35% after 200 cycles at 0.2C. The superior stability of the single-crystalline structure is attributed to its enhanced mechanical stability, as illustrated in Figure 28f,g.
As shown, the polycrystalline cathode particles experience pronounced microcracking due to volume expansion, exposing newly formed surfaces of primary particles to parasitic reactions with the electrolyte. This accelerates capacity fading. In contrast, the monocrystalline structured sample exhibited reduced cracking and minimal intragranular fractures, effectively mitigating these degradation mechanisms.
Compared to single-phase materials, heterostructured materials can leverage the multifunctional advantages of multiple phase structures. This design minimizes phase transformation in the material while simultaneously delivering exceptional electrochemical properties [172,173,174,175,176,177].
Figure 28. Charge–discharge curves of (a) NaMn0.5Ni0.5O2 and (b) NaLi1/9Ni1/3Mn4/9Ti1/9O2 during cycling, (c) cycling performance of NaMn0.5Ni0.5O2, NaLi1/9Ni1/3Mn5/9O2, and NaLi1/9Ni1/3Mn4/9Ti1/9O2 cathodes over 100 cycles at 1 C in the voltage range of 2.0–4.3 V. Lattice fringes of (d) NaMn0.5Ni0.5O2 and (e) NaLi1/9Ni1/3Mn4/9Ti1/9O2. Reprinted from Ref. [116] with permission from the American Chemical Society. Cross-sectional SEM image of (f) pristine polycrystalline particles dispersed in carbon/PVdF and (g) FIB-SEM image of polycrystalline particles after 200 cycles. (h) Cross-sectional SEM image of a pristine single-crystalline particle and (i) FIB-SEM image of single-crystalline particles after 200 cycles. Reprinted from Ref. [113] with permission from the American Chemical Society. (j) Rate capability, and (k) cycling performance of P2/O3-Na0.7Mn0.4Ni0.3Cu0.1Fe0.1Ti0.1O1.95F0.1 at different temperatures. Reprinted from Ref. [178] with permission from Elsevier.
Figure 28. Charge–discharge curves of (a) NaMn0.5Ni0.5O2 and (b) NaLi1/9Ni1/3Mn4/9Ti1/9O2 during cycling, (c) cycling performance of NaMn0.5Ni0.5O2, NaLi1/9Ni1/3Mn5/9O2, and NaLi1/9Ni1/3Mn4/9Ti1/9O2 cathodes over 100 cycles at 1 C in the voltage range of 2.0–4.3 V. Lattice fringes of (d) NaMn0.5Ni0.5O2 and (e) NaLi1/9Ni1/3Mn4/9Ti1/9O2. Reprinted from Ref. [116] with permission from the American Chemical Society. Cross-sectional SEM image of (f) pristine polycrystalline particles dispersed in carbon/PVdF and (g) FIB-SEM image of polycrystalline particles after 200 cycles. (h) Cross-sectional SEM image of a pristine single-crystalline particle and (i) FIB-SEM image of single-crystalline particles after 200 cycles. Reprinted from Ref. [113] with permission from the American Chemical Society. (j) Rate capability, and (k) cycling performance of P2/O3-Na0.7Mn0.4Ni0.3Cu0.1Fe0.1Ti0.1O1.95F0.1 at different temperatures. Reprinted from Ref. [178] with permission from Elsevier.
Micromachines 16 00137 g028
Zhou et al. [178] reported a high-entropy biphasic P2/O3 material with a phase ratio of 23:77, Na0.7Mn0.4Ni0.3Cu0.1Fe0.1Ti0.1O1.95F0.1, as an excellent cathode material exhibiting high rate capability and cycling performance (Figure 28j,k) across a wide temperature range of −40 to 50 °C and at high current densities.
The key factor in optimizing the performance of biphasic materials is achieving an ideal phase ratio. An imbalanced ratio can degrade the cycling stability of the multiphase structure, making it less stable than its single-phase counterparts, especially during prolonged cycling.

3. Conclusions and Perspectives

Sodium-ion batteries (SIBs) are gaining significant attention as a promising alternative to lithium-ion batteries (LIBs), particularly for large-scale applications. The primary focus has been on improving the cathode materials to enhance their capacity density, operating voltage, and overall energy density. Layered oxide transition metal compounds, widely used as cathode materials in LIBs, have emerged as strong candidates for SIBs.
This review has outlined the key challenges that hinder the widespread application of layered oxide cathode materials in SIBs. The most significant obstacle is the structural instability of these materials, which leads to multiple phase transitions that degrade capacity retention over cycling and under high current conditions. Understanding these phase transitions, especially at higher cutoff voltages, is crucial, as the transition behavior in SIBs differs from that of LIBs due to the distinct sizes of Na and Li ions. Consequently, different phase transition mechanisms are at play, affecting the long-term performance of the cathodes.
As layered oxide materials for SIBs are still in the early stages of development, it is crucial to investigate the side reactions occurring at the cathode–electrolyte interface. Furthermore, detailed research into the thermal stability of battery cells, particularly at elevated temperatures, is essential to ensure their reliability under real-world operating conditions.
Based on the materials discussed, layered oxide compounds containing Mn, Fe, and Ni show great promise as cathode materials for SIBs. These materials can be further optimized through strategies such as introducing electrochemically inactive elements like Nb, Al, and Ti, as well as structural modifications, including single-crystalline designs and biphasic structures combining P2 and O3 phases.
A key focus for new cathode materials is reducing production costs and eliminating the need for expensive elements like Co in their composition. Additionally, the development of coatings is necessary to enhance material conductivity during cycling and high-rate operation, as well as to mitigate parasitic reactions with the electrolyte.
Thus, continued research is imperative to advance the understanding of capacity degradation mechanisms and to refine modification techniques. Ultimately, further exploration into these aspects is critical to enabling the widespread adoption of layered oxide materials as cathodes in SIBs, especially for large-scale energy storage systems.

Author Contributions

Conceptualization, writing, methodology, M.A.; review and editing, M.Z.; supervision, H.L. All authors have read and agreed to the published version of the manuscript.

Funding

The work is funded by the National Science Foundations: NSF-OIA-2119688, NSF-OIA-2429592, NSF-CMMI-2215982, NSF-CHE-2216473; and Department of Energy: DE-FE-0032475.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Yabuuchi, N.; Ikeuchi, I.; Kubota, K.; Komaba, S. Thermal Stability of NaCrO2 for Rechargeable Sodium Batteries: Studies By Differential Scanning Calorimetry, and High-Temperature X-Ray Diffraction. In Proceedings of the 224th ECS Meeting, San Francisco, CA, USA, 27 October–1 November 2013. [Google Scholar]
  2. Wable, M.; Bal, B.; Capraz, Ö.Ö. Probing electrochemical strain generation in sodium chromium oxide (NaCrO2) cathode in Na-ion batteries during charge/discharge. Energy Adv. 2024, 3, 601–608. [Google Scholar] [CrossRef]
  3. Li, G.; Yue, X.; Luo, G.; Zhao, J. Electrode potential and activation energy of sodium transition-metal oxides as cathode materials for sodium batteries: A first-principles investigation. Comput. Mater. Sci. 2015, 106, 15–22. [Google Scholar] [CrossRef]
  4. Wang, S.; Chen, F.; He, X.-D.; Zhang, L.-M.; Chen, F.; Wang, J.-R.; Dong, J.-M.; Chen, C.-H. Self-template synthesis of NaCrO2 submicrospheres for stable sodium storage. ACS Appl. Mater. Interfaces 2021, 13, 12203–12210. [Google Scholar] [CrossRef]
  5. Wang, Y.; Li, W.; Hu, G.; Peng, Z.; Cao, Y.; Gao, H.; Du, K.; Goodenough, J.B. Electrochemical performance of large-grained NaCrO2 cathode materials for Na-ion batteries synthesized by decomposition of Na2Cr2O7·2H2O. Chem. Mater. 2019, 31, 5214–5223. [Google Scholar] [CrossRef]
  6. Fukunaga, A.; Nohira, T.; Hagiwara, R.; Numata, K.; Itani, E.; Sakai, S.; Nitta, K. Performance validation of sodium-ion batteries using an ionic liquid electrolyte. J. Appl. Electrochem. 2016, 46, 487–496. [Google Scholar] [CrossRef]
  7. Liang, J.; Liu, L.; Liu, X.; Meng, X.; Zeng, L.; Liu, J.; Li, J.; Shi, Z.; Yang, Y. O3-type NaCrO2 as a superior cathode material for sodium/potassium-ion batteries ensured by high structural reversibility. ACS Appl. Mater. Interfaces 2021, 13, 22635–22645. [Google Scholar] [CrossRef]
  8. Doi, T.; Zhao, L.; Zhou, M.; Okada, S.; Yamaki, J.-I. Quantitative studies on the thermal stability of the interface between graphite electrode and electrolyte. J. Power Sources 2008, 185, 1380–1385. [Google Scholar] [CrossRef]
  9. Klimpel, M.; Kovalenko, M.V.; Kravchyk, K.V. Advances and challenges of aluminum–sulfur batteries. Commun. Chem. 2022, 5, 77. [Google Scholar] [CrossRef] [PubMed]
  10. Mărgulescu, S.; Mărgulescu, E. Lithium Market Trends. Glob. Econ. Observ. 2016, 4, 117. [Google Scholar]
  11. Azevedo, M.; Baczyńska, M.; Hoffman, K.; Krauze, A. Lithium Mining: How New Production Technologies Could Fuel the Global EV Revolution; McKinsey & Company: New York, NY, USA, 2022; Volume 12. [Google Scholar]
  12. Gao, Y.; Wang, Z.; Lu, G. Atomistic understanding of structural evolution, ion transport and oxygen stability in layered NaFeO2. J. Mater. Chem. A. 2019, 7, 2619–2625. [Google Scholar] [CrossRef]
  13. Kim, J.-K.; Lim, Y.J.; Kim, H.; Cho, G.-B.; Kim, Y. A hybrid solid electrolyte for flexible solid-state sodium batteries. Energy Environ. Sci. 2015, 8, 3589–3596. [Google Scholar] [CrossRef]
  14. Lu, Y.; Alonso, J.A.; Yi, Q.; Lu, L.; Wang, Z.L.; Sun, C. A high-performance monolithic solid-state sodium battery with Ca2+-doped Na3Zr2Si2PO12 electrolyte. Adv. Energy Mater. 2019, 9, 1901205. [Google Scholar] [CrossRef]
  15. Xue, Z.; He, D.; Xie, X. Poly(ethylene oxide)-based electrolytes for lithium-ion batteries. J. Mater. Chem. A 2015, 3, 19218–19253. [Google Scholar] [CrossRef]
  16. Faglioni, F.; Merinov, B.V.; Goddard, W.A., III; Kozinsky, B. Factors affecting cyclic durability of all-solid-state lithium batteries using poly(ethylene oxide)-based polymer electrolytes and recommendations to achieve improved performance. Phys. Chem. Chem. Phys. 2018, 20, 26098–26104. [Google Scholar] [CrossRef] [PubMed]
  17. Long, L.; Wang, S.; Xiao, M.; Meng, Y. Polymer electrolytes for lithium polymer batteries. J. Mater. Chem. A 2016, 4, 10038–10069. [Google Scholar] [CrossRef]
  18. Manthiram, A.; Yu, X.; Wang, S. Lithium battery chemistries enabled by solid-state electrolytes. Nat. Rev. Mater. 2017, 2, 16103. [Google Scholar] [CrossRef]
  19. Yu, C.-Y.; Park, J.-S.; Jung, H.-G.; Chung, K.-Y.; Aurbach, D.; Sun, Y.-K.; Myung, S.-T. NaCrO2 cathode for high-rate sodium-ion batteries. Energy Environ. Sci. 2015, 8, 2019–2026. [Google Scholar] [CrossRef]
  20. Shibata, T.; Fukuzumi, Y.; Kobayashi, W.; Moritomo, Y. Fast discharge process of layered cobalt oxides due to high Na+ diffusion. Sci. Rep. 2015, 5, 9006. [Google Scholar] [CrossRef] [PubMed]
  21. L Li, Y.; Lu, Y.; Zhao, C.; Hu, Y.-S.; Titirici, M.-M.; Li, H.; Huang, X.; Chen, L. Recent advances of electrode materials for low-cost sodium-ion batteries towards practical application for grid energy storage. Energy Storage Mater. 2017, 7, 130–151. [Google Scholar] [CrossRef]
  22. Hirsh, H.S.; Li, Y.; Tan, D.H.S.; Zhang, M.; Zhao, E.; Meng, Y.S. Sodium-ion batteries paving the way for grid energy storage. Adv. Energy Mater. 2020, 10, 2001274. [Google Scholar] [CrossRef]
  23. Chen, C.-Y.; Matsumoto, K.; Nohira, T.; Hagiwara, R.; Fukunaga, A.; Sakai, S.; Nitta, K.; Inazawa, S. Electrochemical and structural investigation of NaCrO2 as a positive electrode for sodium secondary battery using inorganic ionic liquid NaFSA–KFSA. J. Power Sources 2013, 237, 52–57. [Google Scholar] [CrossRef]
  24. Shen, B.; Zhan, R.; Dai, C.; Li, Y.; Hu, L.; Niu, Y.; Jiang, J.; Wang, Q.; Xu, M. Manipulating irreversible phase transition of NaCrO2 towards an effective sodium compensation additive for superior sodium-ion full cells. J. Colloid Interface Sci. 2019, 553, 524–529. [Google Scholar] [CrossRef] [PubMed]
  25. Su, J.; Pei, Y.; Yang, Z.; Wang, X. First-principles investigation on the structural, electronic properties and diffusion barriers of Mg/Al doped NaCoO2 as the cathode material of rechargeable sodium batteries. RSC Adv. 2015, 5, 27229–27234. [Google Scholar] [CrossRef]
  26. Guo, S.; Yu, H.; Jian, Z.; Liu, P.; Zhu, Y.; Guo, X.; Chen, M.; Ishida, M.; Zhou, H. A High-Capacity, Low-Cost Layered Sodium Manganese Oxide Material as Cathode for Sodium-Ion Batteries. ChemSusChem 2014, 7, 2115–2119. [Google Scholar] [CrossRef]
  27. Yabuuchi, N.; Kajiyama, M.; Iwatate, J.; Nishikawa, H.; Hitomi, S.; Okuyama, R.; Usui, R.; Yamada, Y.; Komaba, S. P2-type Nax[Fe1/2Mn1/2]O2 made from earth-abundant elements for rechargeable Na batteries. Nat. Mater. 2012, 11, 512–517. [Google Scholar] [CrossRef] [PubMed]
  28. Ponrouch, A.; Marchante, E.; Courty, M.; Tarascon, J.-M.; Palacín, M.R. In search of an optimized electrolyte for Na-ion batteries. Energy Environ. Sci. 2012, 5, 8572–8583. [Google Scholar] [CrossRef]
  29. Dahbi, M.; Yabuuchi, N.; Kubota, K.; Tokiwa, K.; Komaba, S. Negative electrodes for Na-ion batteries. Phys. Chem. Chem. Phys. 2014, 16, 15007–15028. [Google Scholar] [CrossRef] [PubMed]
  30. Kim, Y.; Kim, Y.; Choi, A.; Woo, S.; Mok, D.; Choi, N.S.; Jung, Y.S.; Ryu, J.H.; Oh, S.M.; Lee, K.T. Tin phosphide as a promising anode material for Na-ion batteries. Adv. Mater. 2014, 26, 4139–4144. [Google Scholar] [CrossRef] [PubMed]
  31. Li, Y.; Gao, Y.; Wang, X.; Shen, X.; Kong, Q.; Yu, R.; Lu, G.; Wang, Z.; Chen, L. Iron migration and oxygen oxidation during sodium extraction from NaFeO2. Nano Energy 2018, 47, 519–526. [Google Scholar] [CrossRef]
  32. Wei, J.; Shaw, L.; Chen, W. First-principles comparative study of Cr migration in O3 and O/P hybrid-phased NaCrO2. Phys. Rev. Mater. 2022, 6, 095403. [Google Scholar] [CrossRef]
  33. Li, G.; Zhu, W. First-principles investigation on the crystal, electronic structures and diffusion barriers of F-doped NaMO2 (M= V, Cr, Co and Ni) for rechargeable Na-ion batteries. J. Solid State Chem. 2021, 302, 122440. [Google Scholar] [CrossRef]
  34. Sawicki, M.; Ortiz, A.L.; Luo, M.; Shaw, L.L. Structural-Defect-Controlled Electrochemical Performance of Sodium Ion Batteries with NaCrO2 Cathodes. ChemElectroChem 2017, 4, 3222–3230. [Google Scholar] [CrossRef]
  35. Wei, J.; Shaw, L.; Chen, W. First-principles prediction of Na diffusivity in doped NaCrO2 layered cathode materials with van der Waals interactions. J. Phys. Chem. C. 2020, 124, 12239–12248. [Google Scholar] [CrossRef]
  36. Zhao, C.; Yao, Z.; Wang, Q.; Li, H.; Wang, J.; Liu, M.; Ganapathy, S.; Lu, Y.; Cabana, J.; Li, B.; et al. Revealing high Na-content P2-type layered oxides as advanced sodium-ion cathodes. J. Am. Chem. Soc. 2020, 142, 5742–5750. [Google Scholar] [CrossRef]
  37. Bianchini, M.; Wang, J.; Clément, R.; Cede, G. A first-principles and experimental investigation of nickel solubility into the P2 NaxCoO2 sodium-ion cathode. Adv. Energy Mater. 2018, 8, 1801446. [Google Scholar] [CrossRef]
  38. Kubota, K.; Asari, T.; Yoshida, H.; Yabuuchi, N.; Shiiba, H.; Nakayama, M.; Komaba, S. Understanding the structural evolution and redox mechanism of a NaFeO2–NaCoO2 solid solution for sodium-ion batteries. Adv. Funct. Mater. 2016, 26, 6047–6059. [Google Scholar] [CrossRef]
  39. Yan, P.; Zheng, J.; Zhang, J.-G.; Wang, C. Atomic resolution structural and chemical imaging revealing the sequential migration of Ni, Co, and Mn upon the battery cycling of layered cathode. Nano Lett. 2017, 17, 3946–3951. [Google Scholar] [CrossRef]
  40. Tian, M.; Gao, Y.; Xiao, R.; Wang, Z.; Chen, L. Structural stability and stabilization of Li2MoO3. Phys. Chem. Chem. Phys. 2017, 19, 17538–17543. [Google Scholar] [CrossRef] [PubMed]
  41. Gao, Y.; Ma, J.; Wang, Z.; Lu, G.; Chen, L. Vacancy-induced MnO6 distortion and its impacts on structural transition of Li2MnO3. Phys. Chem. Chem. Phys. 2017, 19, 7025–7031. [Google Scholar] [CrossRef]
  42. Liu, Y.-F.; Han, K.; Peng, D.-N.; Kong, L.-Y.; Su, Y.; Li, H.-W.; Hu, H.-Y.; Li, J.-Y.; Wang, H.-R.; Fu, Z.-Q.; et al. Layered oxide cathodes for sodium-ion batteries: From air stability, interface chemistry to phase transition. InfoMat 2023, 5, 12422. [Google Scholar] [CrossRef]
  43. Jakobsen, C.L.; Andersen, B.P.; Johansen, M.; Christensen, C.K.; Drejer, A.Ø.; Karlsen, M.A.; Ravnsbæk. Structure and evolution of disordered deep-charge phases in NaxCrO2 Na-ion battery electrodes. J. Power Sources 2024, 591, 233875. [Google Scholar] [CrossRef]
  44. Yabuuchi, N.; Ikeuchi, I.; Kubota, K.; Komaba, S. Thermal stability of NaxCrO2 for rechargeable sodium batteries; studies by high-temperature synchrotron X-ray diffraction. ACS Appl. Mater. Interfaces 2016, 8, 32292–32299. [Google Scholar] [CrossRef]
  45. Luo, M.; Ortiz, A.L.; Shaw, L.L. Enhancing the electrochemical performance of NaCrO2 through structural defect control. ACS Appl. Energy Mater. 2020, 3, 7216–7227. [Google Scholar] [CrossRef]
  46. Zheng, L.; Bennett, J.C.; Obrovac, M.N. Stabilizing NaCrO2 by sodium site doping with calcium. J. Electrochem. Soc. 2019, 166, A2058. [Google Scholar] [CrossRef]
  47. Kubota, K.; Ikeuchi, I.; Nakayama, T.; Takei, C.; Yabuuchi, N.; Shiiba, H.; Nakayama, M.; Komaba, S. New insight into structural evolution in layered NaCrO2 during electrochemical sodium extraction. J. Phys. Chem. C 2015, 119, 166–175. [Google Scholar] [CrossRef]
  48. Komaba, S.; Nakayama, T.; Ogata, A.; Shimizu, T.; Takei, C.; Takada, S.; Hokura, A.; Nakai, I. Electrochemically reversible sodium intercalation of layered NaNi0.5Mn0.5O2 and NaCrO2. ECS Trans. 2009, 16, 43. [Google Scholar] [CrossRef]
  49. Komaba, S.; Takei, C.; Nakayama, T.; Ogata, A.; Yabuuchi, N. Electrochemical intercalation activity of layered NaCrO2 vs. LiCrO2. Electrochem. Commun. 2010, 12, 355–358. [Google Scholar] [CrossRef]
  50. Myung, S.-T.; Komaba, S.; Hirosaki, N.; Kumagai, N.; Arai, K.; Kodama, R.; Nakai, I. Structural Investigation of Layered Li1-δMnxCr1-xO2 by XANES and In Situ XRD Measurements. J. Electrochem. Soc. 2003, 150, A1560. [Google Scholar] [CrossRef]
  51. Zhou, Y.-N.; Ding, J.-J.; Nam, K.-W.; Yu, X.; Bak, S.-M.; Hu, E.; Liu, J.; Bai, J.; Li, H.; Fu, Z.-W.; et al. Phase transition behavior of NaCrO2 during sodium extraction studied by synchrotron-based X-ray diffraction and absorption spectroscopy. J. Mater. Chem. A. 2013, 1, 11130–11134. [Google Scholar] [CrossRef]
  52. Xia, X.; Dahn, J.R. NaCrO2 is a fundamentally safe positive electrode material for sodium-ion batteries with liquid electrolytes. Electrochem. Solid-State Lett. 2011, 15, A1. [Google Scholar] [CrossRef]
  53. Bo, S.-H.; Li, X.; Toumar, A.J.; Ceder, G. Layered-to-rock-salt transformation in desodiated NaxCrO2 (x 0.4). Chem. Mater. 2016, 28, 1419–1429. [Google Scholar] [CrossRef]
  54. Jakobsen, C.L.; Brighi, M.; Andersen, B.P.; Ducrest, G.; Černý, R.; Ravnsbæk, D.B. Expanded solid-solution behavior and charge-discharge asymmetry in NaxCrO2 Na-ion battery electrodes. J. Power Sources. 2022, 535, 231317. [Google Scholar] [CrossRef]
  55. Tsuchiya, Y.; Takanashi, K.; Nishinobo, T.; Hokura, A.; Yonemura, M.; Matsukawa, T.; Ishigaki, T.; Yamanaka, K.; Ohta, T.; Yabuuchi, N. Layered NaxCrxTi1-xO2 as Bifunctional Electrode Materials for Rechargeable Sodium Batteries. Chem. Mater. 2016, 28, 7006–7016. [Google Scholar] [CrossRef]
  56. Park, W.B.; Nathan, M.G.T.; Han, S.C.; Lee, J.-W.; Sohn, K.-S.; Pyo, M. Aliovalent-doped sodium chromium oxide (Na 0.9Cr0.9Sn0.1O2 and Na0.8Cr0.9Sb0.1O2) for sodium-ion battery cathodes with high-voltage characteristics. RSC Adv. 2020, 10, 43273–43281. [Google Scholar] [CrossRef]
  57. Xi, K.; Chu, S.; Zhang, X.; Zhang, X.; Zhang, H.; Xu, H.; Bian, J.; Fang, T.; Guo, S.; Liu, P.; et al. A high-performance layered Cr-based cathode for sodium-ion batteries. Nano Energy 2020, 67, 104215. [Google Scholar] [CrossRef]
  58. Wang, S.; Chen, F.; He, H.-Y.; Zhu, Y.-R.; Liu, H.-B.; Chen, C.-H. Enhancing structural stability of NaCrO2 by Nb-substituting for sodium-ion battery. J. Alloys Compd. 2022, 925, 166690. [Google Scholar] [CrossRef]
  59. Wang, Y.; Cui, P.; Zhu, W.; Feng, Z.; Vigeant, M.-J.; Demers, H.; Guerfi, A.; Zaghib, K. Enhancing the electrochemical performance of an O3-NaCrO2 cathode in sodium-ion batteries by cation substitution. J. Power Sources 2019, 435, 226760. [Google Scholar] [CrossRef]
  60. Ikhe, A.B.; Park, W.B.; Manasi, M.; Ahn, D.; Sohn, K.-S.; Pyo, M. Unprecedented cyclability and moisture durability of NaCrO2 sodium-ion battery cathode via simultaneous Al doping and Cr2O3 coating. ACS Appl. Mater. Interfaces 2023, 15, 14958–14969. [Google Scholar] [CrossRef]
  61. Lee, I.; Oh, G.; Lee, S.; Yu, T.-Y.; Alfaruqi, M.H.; Mathew, V.; Sambandam, B.; Sun, Y.-K.; Hwang, J.-Y.; Kim, J. Cationic and transition metal co-substitution strategy of O3-type NaCrO2 cathode for high-energy sodium-ion batteries. Energy Storage Mater. 2021, 41, 183–195. [Google Scholar] [CrossRef]
  62. Ma, C.; Li, X.-L.; Yue, X.-Y.; Bao, J.; Luo, R.-J.; Zhou, Y.-N. Suppressing O3- O’3 phase transition in NaCrO2 cathode enabling high rate capability for sodium-ion batteries by Sb substitution. Chem. Eng. J. 2022, 432, 134305. [Google Scholar] [CrossRef]
  63. Wang, Y.; Xiao, R.; Hu, Y.-S.; Avdeev, M.; Chen, L. P2-Na0.6[Cr0.6Ti0.4]O2 cation-disordered electrode for high-rate symmetric rechargeable sodium-ion batteries. Nat. Commun. 2015, 6, 6954. [Google Scholar] [CrossRef]
  64. Li, W.; Wang, Y.; Hu, G.; Peng, Z.; Cao, Y.; Zeng, Y.; Du, K. Ti-doped NaCrO2 as cathode materials for sodium-ion batteries with excellent long cycle life. J. Alloys Compd. 2019, 779, 147–155. [Google Scholar] [CrossRef]
  65. Wang, S.; Chen, F.; Zhu, T.-Y.; He, X.-D.; Liao, J.-Y.; Zhang, L.-M.; Ding, X.; Hu, Q.; Chen, C.-H. In situ-formed Cr2O3 coating on NaCrO2 with improved sodium storage performance. ACS Appl. Mater. Interfaces. 2020, 12, 44671–44678. [Google Scholar] [CrossRef]
  66. Cai, J.; Wei, Y.; Zhu, Y.; Tian, L.; Zhang, J.; Shen, Y.; Wang, M. Nitrogen-doped carbon-coating enables high-rate capability and long-cycle stability of NaCrO2 cathode for sodium-ion battery. Appl. Surf. Sci. 2025, 683, 161804. [Google Scholar] [CrossRef]
  67. Bhardwaj, A.; Panwar, A.K. Effect of carbon shell over NaCrO2 core by C2H2 decomposition to enhance electrochemical properties for rechargeable Sodium-ion batteries. Appl. Surf. Sci. 2022, 573, 151449. [Google Scholar] [CrossRef]
  68. Shi, Z.; Wang, Z.; Shaw, L.L.; Ashuri, M. On the Electrochemical Properties of Carbon-Coated NaCrO2 for Na-Ion Batteries. Batteries 2023, 9, 433. [Google Scholar] [CrossRef]
  69. Ding, J.-J.; Zhou, Y.-N.; Sun, Q.; Fu, Z.-W. Cycle performance improvement of NaCrO2 cathode by carbon coating for sodium ion batteries. Electrochem. Commun. 2012, 22, 85–88. [Google Scholar] [CrossRef]
  70. Wu, J.; Hu, G.; Du, K.; Peng, Z.; Huang, M.; Fan, J.; Gong, Y.; Guan, D.; Shi, Y.; Liu, R.; et al. Inhibiting electrochemical phase transition of NaCrO2 with long-cycle stability by surface fluorination treatment. Electrochim. Acta 2022, 403, 139641. [Google Scholar] [CrossRef]
  71. Yabuuchi, N.; Yano, M.; Yoshida, H.; Kuze, S.; Komaba, S. Synthesis and electrode performance of O3-type NaFeO2-NaNi1/2Mn1/2O2 solid solution for rechargeable sodium batteries. J. Electrochem. Soc. 2013, 160, A3131. [Google Scholar] [CrossRef]
  72. Okada, S.; Yamaki, J. Iron-Based Rare-Metal-Free Cathodes. In Lithium Ion Rechargeable Batteries; John Wiley & Sons: Hoboken, NJ, USA, 2010. [Google Scholar]
  73. Zhao, J.; Zhao, L.; Dimov, N.; Okada, S.; Nishida, T. Electrochemical and thermal properties of α-NaFeO2 cathode for Na-ion batteries. J. Electrochem. Soc. 2013, 160, A3077. [Google Scholar] [CrossRef]
  74. Watanabe, H.; Usui, H.; Domi, Y.; Nishida, T.; Uetake, K.; Tanaka, T.; Kurokawa, H.; Sakaguchi, H. NaFeO2: Possible Materials for Anode and Cathode of Na-Ion Batteries. ACS Electrochem. 2025, 1, 73–81. [Google Scholar] [CrossRef]
  75. McAuliffe, R.D.; Kamm, G.E.; McDermott, M.J.; Hermann, R.P.; Vasquez-Garcia, N.; Sacci, R.L.; Persson, K.A.; Chapman, K.W.; Veith, G.M. Direct Mechanochemical Synthesis, Phase Stability, and Electrochemical Performance of α-NaFeO2. Inorg. Chem. 2023, 62, 3358–3367. [Google Scholar] [CrossRef]
  76. Blesa, M.C.; Moran, E.; León, C.; Santamaria, J.; Tornero, J.D.; Menéndez, N. α-NaFeO2: Ionic conductivity and sodium extraction. Solid State Ion. 1999, 126, 81–87. [Google Scholar] [CrossRef]
  77. Okada, S.; Takahashi, Y.; Kiyabu, T.; Doi, T.; Yamaki, J.-I.; Nishida, T. Layered transition metal oxides as cathodes for sodium secondary battery. ECS Meet. Abstr. 2006, MA2006-02, 201. [Google Scholar] [CrossRef]
  78. Tabuchi, M.; Kataoka, R. Structure and electrochemical properties of α-NaFeO2 obtained under various hydrothermal conditions. J. Electrochem. Soc. 2019, 166, A2209. [Google Scholar] [CrossRef]
  79. Kuganathan, N.; Kelaidis, N.; Chroneos, A. Defect chemistry, sodium diffusion and doping behavior in NaFeO2 polymorphs as cathode materials for Na-ion batteries: A computational study. Materials 2019, 12, 3243. [Google Scholar] [CrossRef]
  80. Kikkawa, S.; Miyazaki, S.; Koizumi, M. Sodium deintercalation from α-NaFeO2. Mater. Res. Bull. 1985, 20, 373–377. [Google Scholar] [CrossRef]
  81. Takeda, Y.; Nakahara, K.; Nishijima, M.; Imanishi, N.; Yamamoto, O.; Takano, M.; Kanno, R. Sodium deintercalation from sodium iron oxide. Mater. Res. Bull. 1994, 29, 659–666. [Google Scholar] [CrossRef]
  82. Yabuuchi, N.; Yoshida, H.; Komaba, S. Crystal structures and electrode performance of alpha-NaFeO2 for rechargeable sodium batteries. Electrochem 2012, 80, 716–719. [Google Scholar] [CrossRef]
  83. Lee, E.; Brown, D.E.; Alp, E.E.; Ren, Y.; Lu, J.; Woo, J.-J.; Johnson, C.S. New insights into the performance degradation of Fe-based layered oxides in sodium-ion batteries: Instability of Fe3+/Fe4+ redox in α-NaFeO2. Chem. Mater. 2015, 27, 6755–6764. [Google Scholar] [CrossRef]
  84. Thorne, J.S.; Chowdhury, S.; Dunlap, R.A.; Obrovac, M.N. Structure and electrochemistry of NaxFexTi1-xO2 (1.0 ≥ x ≥ 0.75) for Na-Ion battery positive electrodes. J. Electrochem. Soc. 2014, 161, A1801. [Google Scholar] [CrossRef]
  85. Xu, J.; Han, Z.; Jiang, K.; Bai, P.; Liang, Y.; Zhang, X.; Wang, P.; Guo, S.; Zhou, H. suppressing cation migration and reducing particle cracks in a layered Fe-based cathode for advanced sodium-ion batteries. Small 2020, 16, 1904388. [Google Scholar] [CrossRef] [PubMed]
  86. Yoshida, H.; Yabuuchi, N.; Komaba, S. NaFe0.5Co0.5O2 as high energy and power positive electrode for Na-ion batteries. Electrochem. Commun. 2013, 34, 60–63. [Google Scholar] [CrossRef]
  87. Shevchenko, V.A.; Glazkova, I.S.; Novichkov, D.A.; Skvortsova, I.; Sobolev, A.V.; Abakumov, A.M.; Presniakov, I.A.; Drozhzhin, O.A.; Antipov, E.V. Competition between the Ni and Fe redox in the O3-NaNi1/3Fe1/3Mn1/3O2 cathode material for na-ion batteries. Chem. Mater. 2023, 35, 4015–4025. [Google Scholar] [CrossRef]
  88. Ono, Y.; Yui, Y.; Hayashi, M.; Asakura, K.; Kitabayashi, H.; Takahashi, K.I. Electrochemical properties of NaCuO2 for sodium-ion secondary batteries. ECS Trans. 2014, 58, 33. [Google Scholar] [CrossRef]
  89. Ono, Y.; Yui, Y.; Asakura, K.; Nakamura, J.; Hayashi, M.; Takahashi, K.I. Characterization of electrochemical cycling-induced new products of NaCuO2 cathode material for sodium secondary batteries. Am. J. Phys. Chem. 2014, 3, 61–66. [Google Scholar] [CrossRef]
  90. Dyer, L.D.; Borie, B.S., Jr.; Smith, G.P. Alkali metal-nickel oxides of the type MNiO2. J. Am. Chem. Soc. 1954, 76, 1499–1503. [Google Scholar] [CrossRef]
  91. Dick, S.; Müller, M.; Preissinger, F.; Zeiske, T. The structure of monoclinic NaNiO2 as determined by powder X-ray and neutron scattering. Powder Diffr. 1997, 12, 239–241. [Google Scholar] [CrossRef]
  92. Chappel, E.; Núñez-Regueiro, M.D.; Chouteau, G.; Isnard, O.; Darie, C. Study of the ferrodistorsive orbital ordering in NaNiO2 by neutron diffraction and submillimeter wave ESR. Eur. Phys. J. B 2000, 17, 615–622. [Google Scholar] [CrossRef]
  93. Sofin, M.; Jansen, M. New route of preparation and properties of NaNiO2. Z. Naturforsch. B 2005, 60, 701–704. [Google Scholar] [CrossRef]
  94. Vassilaras, P.; Ma, X.; Li, X.; Ceder, G. Electrochemical properties of monoclinic NaNiO2. J. Electrochem. Soc. 2012, 160, A207. [Google Scholar] [CrossRef]
  95. Miyazaki, S.; Kikkawa, S.; Koizumi, M. Chemical and electrochemical deintercalations of the layered compounds LiMO2 (M = Cr, Co) and NaM’O2 (M’ = Cr, Fe, Co, Ni). Synth. Met. 1983, 6, 211–217. [Google Scholar] [CrossRef]
  96. Han, M.H.; Gonzalo, E.; Casas-Cabanas, M.; Rojo, T. Structural evolution and electrochemistry of monoclinic NaNiO2 upon the first cycling process. J. Power Sources 2014, 258, 266–271. [Google Scholar] [CrossRef]
  97. Wang, L.; Wang, J.; Zhang, X.; Ren, Y.; Zuo, P.; Yin, G.; Wang, J. Unravelling the origin of irreversible capacity loss in NaNiO2 for high voltage sodium ion batteries. Nano Energy 2017, 34, 215–223. [Google Scholar] [CrossRef]
  98. Kim, D.H.; Kim, J.-Y.; Park, J.-H.; Kim, S.-O.; Kim, H.-S.; Kim, K.-B.; Chung, K.Y. RT-XAMF and TR-XRD studies of solid-state synthesis and thermal stability of NaNiO2 as cathode material for sodium-ion batteries. Ceram. Int. 2022, 48, 19675–19680. [Google Scholar] [CrossRef]
  99. Singh, G.; Aguesse, F.; Otaegui, L.; Goikolea, E.; Gonzalo, E.; Segalini, J.; Rojo, T. Electrochemical performance of NaFex (Ni0.5Ti0.5)1-xO2 (x = 0.2 and x = 0.4) cathode for sodium-ion battery. J. Power Sources 2015, 273, 333–339. [Google Scholar] [CrossRef]
  100. Chang, Y.; Zhou, Y.; Wang, Z.; Li, X.; Wang, D.; Duan, J.; Wang, J.; Yan, G. Synthesis of NaNi0.5Mn0.5O2 cathode materials for sodium-ion batteries via spray pyrolysis method. J. Alloys Compd. 2022, 922, 166283. [Google Scholar] [CrossRef]
  101. Wang, H.; Gu, M.; Jiang, J.; Lai, C.; Ai, X. An O3-type NaNi0.5Mn0.3Ti0.2O2 compound as new cathode material for room-temperature sodium-ion batteries. J. Power Sources 2016, 327, 653–657. [Google Scholar] [CrossRef]
  102. Zheng, L.; Fielden, R.; Bennett, J.C.; Obrovac, M.N. Hexagonal and monoclinic NaNi0.8Co0.15Al0.05O2 (Na-NCA) for sodium ion batteries. J. Power Sources 2019, 433, 226698. [Google Scholar] [CrossRef]
  103. Meghnani, D.; Singh, R.K. Influence of synthesis route on the structure and electrochemical performance of biphasic (O’3/O3) NaNi0.815Co0.15Al0.035O2 cathode for sodium-ion batteries. Electrochim. Acta 2022, 419, 140403. [Google Scholar] [CrossRef]
  104. Deng, C.; Gabriel, E.; Skinner, P.; Lee, S.; Barnes, P.; Ma, C.; Gim, J.; Lau, M.L.; Lee, E.; Xiong, H. Origins of Irreversibility in Layered NaNixFeyMnzO2 Cathode Materials for Sodium Ion Batteries. ACS Appl. Mater. Interfaces 2020, 12, 51397–51408. [Google Scholar] [CrossRef]
  105. Fielden, R.; Obrovac, M.N. Investigation of the NaNixMn1-xO2 (0≤ x≤ 1) System for Na-ion Battery Cathode Materials (Preprint). J. Electrochem. Soc. 2015, 162, A453. [Google Scholar] [CrossRef]
  106. Moeez, I.; Susanto, D.; Ali, G.; Jung, H.-G.; Lim, H.-D.; Chung, K.Y. Effect of the interfacial protective layer on the NaFe0.5Ni0.5O2 cathode for rechargeable sodium-ion batteries. J. Mater. Chem. A 2020, 8, 13964–13970. [Google Scholar] [CrossRef]
  107. Nanba, Y.; Iwao, T.; Mortemard de Boisse, B.; Zhao, W.; Hosono, E.; Asakura, D.; Niwa, H.; Kiuchi, H.; Miyawaki, J.; Harada, Y.; et al. Redox potential paradox in NaxMO2 for sodium-ion battery cathodes. Chem. Mater. 2016, 28, 1058–1065. [Google Scholar] [CrossRef]
  108. Guo, H.; Wang, Y.; Han, W.; Yu, Z.; Qi, X.; Sun, K.; Hu, Y.-S.; Liu, Y.; Chen, D.; Chen, L. Na-deficient O3-type cathode material Na0.8[Ni0.3Co0.2Ti0.5]O2 for room-temperature sodium-ion batteries. Electrochim. Acta 2015, 158, 258–263. [Google Scholar] [CrossRef]
  109. Maletti, S.; Sarapulova, A.; Schökel, A.; Mikhailova, D. Operando studies on the NaNi0.5Ti0.5O2 cathode for Na-ion batteries: Elucidating titanium as a structure stabilizer. ACS Appl. Mater. Interfaces 2019, 11, 33923–33930. [Google Scholar] [CrossRef]
  110. Ma, J.; Bo, S.-H.; Wu, L.; Zhu, Y.; Grey, C.P.; Khalifah, P.G. Ordered and disordered polymorphs of Na(Ni2/3Sb1/3)O2: Honeycomb-ordered cathodes for Na-ion batteries. Chem. Mater. 2015, 27, 2387–2399. [Google Scholar] [CrossRef]
  111. An, S.; Karger, L.; Dreyer, S.L.; Hu, Y.; Barbosa, E.; Zhang, R.; Lin, J.; Fichtner, M.; Kondrakov, A.; Janek, J. Improving cycling performance of the NaNiO2 cathode in sodium-ion batteries by titanium substitution. Mater. Futures 2024, 3, 035103. [Google Scholar] [CrossRef]
  112. Saadoune, I.; Maazaz, A.; Ménétrier, M.; Delmas, C. On the NaxNi0.6Co0.4O2 system: Physical and electrochemical studies. J. Solid State Chem. 1996, 122, 111–117. [Google Scholar] [CrossRef]
  113. Darga, J.; Manthiram, A. Facile synthesis of O3-type NaNi0.5Mn0.5O2 single crystals with improved performance in sodium-ion batteries. ACS Appl. Mater. Interfaces 2022, 14, 52729–52737. [Google Scholar] [CrossRef]
  114. Liang, X.; Yu, T.-Y.; Ryu, H.-H.; Sun, Y.-K. Hierarchical O3/P2 heterostructured cathode materials for advanced sodium-ion batteries. Energy Storage Mater. 2022, 47, 515–525. [Google Scholar] [CrossRef]
  115. Zhang, S.; Li, X.; Su, Y.; Yang, Y.; Yu, H.; Wang, H.; Zhang, Q.; Lu, Y.; Song, D.; Wang, S.; et al. Four-in-one strategy to boost the performance of Nax[Ni, Mn]O2. Adv. Funct. Mater. 2023, 33, 2301568. [Google Scholar] [CrossRef]
  116. Wang, Q.; Yu, G.; Luo, B.; Ji, W.; Liu, Z.; Li, M.; Nong, Y.; Tian, Y.; Wang, X.; Zhang, J.; et al. Suppression of Adverse Phase Transition of Layered Oxide Cathode via Local Electronic Structure Regulation for High-Capacity Sodium-Ion Batteries. ACS Nano 2024, 18, 18622–18634. [Google Scholar] [CrossRef]
  117. Yuan, T.; Li, P.; Sun, Y.; Che, H.; Zheng, Q.; Zhang, Y.; Huang, S.; Qiu, J.; Pang, Y.; Yang, J.; et al. Refining O3-Type Ni/Mn-Based Sodium-Ion Battery Cathodes via “Atomic Knife” Achieving High Capacity and Stability. Adv. Funct. Mater. 2024, 34, 2414627. [Google Scholar] [CrossRef]
  118. Wang, P.F.; Yao, H.R.; Liu, X.Y.; Zhang, J.N.; Gu, L.; Yu, X.Q.; Yin, Y.X.; Guo, Y.G. Ti-substituted NaNi0.5Mn0.5-xTixO2 cathodes with reversible O3−P3 phase transition for high-performance sodium-ion batteries. Adv. Mater. 2017, 29, 1700210. [Google Scholar] [CrossRef] [PubMed]
  119. Meghnani, D.; Srivastava, N.; Tiwari, R.K.; Mishra, R.; Patel, A.; Tiwari, A.; Samriddhi; Singh, S.P.; Yadav, V.; Singh, R.K. Surface-Coated NaNi0.815Co0.15Al0.035O2 Cathode-Based Sodium-Ion Batteries with Enhanced Performance. Energy Fuels 2023, 38, 694–706. [Google Scholar] [CrossRef]
  120. Li, M.; Qiu, X.; Yin, Y.; Wei, T.; Dai, Z. O3-Type Ni-Rich NaNi2/3Mn1/6Fe1/6O2: A high-performance cathode material for sodium-ion batteries. J. Alloys Compd. 2023, 969, 172406. [Google Scholar] [CrossRef]
  121. Bhide, A.; Hariharan, K. Physicochemical properties of NaxCoO2 as a cathode for solid state sodium battery. Solid State Ion. 2011, 192, 360–363. [Google Scholar] [CrossRef]
  122. Shibata, A.; Kobayashi, W.; Moritomo, Y. Sodium ion diffusion in layered NaxMnO2 (0.49 ≤ x≤ 0.75): Comparison with NaxCoO2. Appl. Phys. Express 2014, 7, 067101. [Google Scholar] [CrossRef]
  123. Zandbergen, H.W.; Foo, M.; Xu, Q.; Kumar, V.; Cava, R.J. Sodium ion ordering in NaxCoO2: Electron diffraction study. Phys. Rev. B 2004, 70, 024101. [Google Scholar] [CrossRef]
  124. Nwanya, A.C.; Ndipingwi, M.M.; Ezema, F.I.; Iwuoha, E.I.; Maaza, M. Bio-synthesized P2-Na0.57CoO2 nanoparticles as cathode for aqueous sodium ion battery. J. Electroanal. Chem. 2020, 878, 114600. [Google Scholar] [CrossRef]
  125. Su, J.-C.; Zhou, G.; Pei, Y.; Yang, Z.-H.; Wang, X.-Y. First-principles investigation on structural and electrochemical properties of NaCoO2 for rechargeable Na-ion batteries. J. Cent. South Univ. 2015, 22, 2036–2042. [Google Scholar] [CrossRef]
  126. Takahashi, Y.; Gotoh, Y.; Akimoto, J. Single-crystal growth, crystal and electronic structure of NaCoO2. J. Solid State Chem. 2003, 172, 22–26. [Google Scholar] [CrossRef]
  127. Berthelot, R.; Carlier, D.; Delmas, C. Electrochemical investigation of the P2-NaxCoO2 phase diagram. Nat. Mater. 2011, 10, 74–80. [Google Scholar] [CrossRef] [PubMed]
  128. Gao, L.; Chen, S.; Zhang, L.; Yang, X. High Areal Capacity Na0.67CoO2 Bundle Array Cathode Tailored for High-Performance Sodium-Ion Batteries. ChemElectroChem 2019, 6, 947–952. [Google Scholar] [CrossRef]
  129. Shiprath, K.; Manjunatha, H.; Nadh, R.V.; Babu, T.A.; Ramesh, S.; Naidu, K.C.B.; Ramesha, M. Synthesis and Electrochemical Characterization of NaCoO2 as Cathode Material in 2 M NaOH Aqueous Electrolyte. ChemistrySelect 2021, 6, 1874–1881. [Google Scholar] [CrossRef]
  130. Yabuuchi, N.; Kubota, K.; Dahbi, M.; Komaba, S. Research development on sodium-ion batteries. Chem. Rev. 2014, 114, 11636–11682. [Google Scholar] [CrossRef] [PubMed]
  131. Clément, R.J.; Bruce, P.G.; Grey, C.P. manganese-based P2-type transition metal oxides as sodium-ion battery cathode materials. J. Electrochem. Soc. 2015, 162, A2589. [Google Scholar] [CrossRef]
  132. Carlier, D.; Cheng, J.H.; Berthelot, R.; Guignard, M.; Yoncheva, M.; Stoyanova, R.; Hwang, B.J.; Delmas, C. The P2-Na2/3Co2/3Mn1/3O2 phase: Structure, physical properties and electrochemical behavior as positive electrode in sodium battery. Dalton Trans. 2011, 40, 9306–9312. [Google Scholar] [CrossRef]
  133. Beck, F.R.; Cheng, Y.Q.; Bi, Z.; Feygenson, M.; Bridges, C.A.; Moorhead-Rosenberg, Z.; Manthiram, A.; Goodenough, J.B.; Paranthaman, M.P.; Manivannan, A. Neutron diffraction and electrochemical studies of Na0.79CoO2 and Na0.79Co0.7Mn0.3O2 cathodes for sodium-ion batteries. J. Electrochem. Soc. 2014, 161, A961. [Google Scholar] [CrossRef]
  134. Kang, S.M.; Park, J.-H.; Jin, A.; Jung, Y.H.; Mun, J.; Sung, Y.-E. Na+/Vacancy Disordered P2-Na0.67Co1-xTixO2: High-Energy and High-Power Cathode Materials for Sodium Ion Batteries. ACS Appl. Mater. Interfaces 2018, 10, 3562–3570. [Google Scholar] [CrossRef] [PubMed]
  135. Matsui, M.; Mizukoshi, F.; Imanishi, N. Improved cycling performance of P2-type layered sodium cobalt oxide by calcium substitution. J. Power Sources 2015, 280, 205–209. [Google Scholar] [CrossRef]
  136. Han, S.C.; Lim, H.; Jeong, J.; Ahn, D.; Park, W.B.; Sohn, K.-S.; Pyo, M. Ca-doped NaxCoO2 for improved cyclability in sodium ion batteries. J. Power Sources 2015, 277, 9–16. [Google Scholar] [CrossRef]
  137. Ding, J.J.; Zhou, Y.N.; Sun, Q.; Yu, X.Q.; Yang, X.Q.; Fu, Z.W. Electrochemical properties of P2-phase Na0.74CoO2 compounds as cathode material for rechargeable sodium-ion batteries. Electrochim. Acta 2013, 87, 388–393. [Google Scholar] [CrossRef]
  138. Rai, A.K.; Anh, L.T.; Gim, J.; Mathew, V.; Kim, J. Electrochemical properties of NaxCoO2 (x~ 0.71) cathode for rechargeable sodium-ion batteries. Ceram. Int. 2014, 40, 2411–2417. [Google Scholar] [CrossRef]
  139. Hwang, S.; Lee, Y.; Jo, E.; Chung, K.Y.; Choi, W.; Kim, S.M.; Chang, W. Investigation of Thermal Stability of P2-NaxCoO2 Cathode Materials for Sodium Ion Batteries Using Real-Time Electron Microscopy. ACS Appl. Mater. Interfaces 2017, 9, 18883–18888. [Google Scholar] [CrossRef]
  140. Fang, Y.; Yu, X.-Y.; Lou, X.W. (David). A practical high-energy cathode for sodium-ion batteries based on uniform P2-Na0.7CoO2 microspheres. Angew. Chem. Int. Ed. 2017, 56, 5801–5805. [Google Scholar] [CrossRef]
  141. Onoda, M. Geometrically frustrated triangular lattice system NaxVO2: Super paramagnetism in x = 1 and trimerization in x ≈ 0.7. J. Phys. Condens. Matter. 2008, 20, 145205. [Google Scholar] [CrossRef]
  142. Fielden, R.; Cole, L.; Obrovac, M.N. Successive orbital ordering transitions in NaVO2. Phys. Rev. Lett. 2008, 101, 166402. [Google Scholar] [CrossRef]
  143. Fielden, R.; Cole, L.; Obrovac, M.N. Low voltage sodium intercalation in NaxVxTi1-xO2 (2/3 ≤ x ≤ 1). J. Electrochem. Soc. 2017, 164, A490. [Google Scholar] [CrossRef]
  144. Guignard, M.; Didier, C.; Darriet, J.; Bordet, P.; Elkaïm, E.; Delmas, C. P2-NaxVO2 system as electrodes for batteries and electron-correlated materials. Nat. Mater. 2013, 12, 74–80. [Google Scholar] [CrossRef]
  145. Hamani, D.; Ati, M.; Tarascon, J.-M.; Rozier, P. NaxVO2 as possible electrode for Na-ion batteries. Electrochem. Commun. 2011, 13, 938–941. [Google Scholar] [CrossRef]
  146. Didier, C.; Guignard, M.; Denage, C.; Szajwaj, O.; Ito, S.; Saadoune, I.; Darriet, J.; Delmas, C. Electrochemical Na-Deintercalation from NaVO2. Electrochem. Solid-State Lett. 2011, 14, A75. [Google Scholar] [CrossRef]
  147. Didier, C.; Guignard, M.; Darriet, J.; Delmas, C. O’3-NaxVO2 System: A Superstructure for Na1/2VO2. Inorg. Chem. 2012, 51, 11007–11016. [Google Scholar] [CrossRef] [PubMed]
  148. Shishkin, M.; Kumakura, S.; Sato, S.; Kubota, K.; Komaba, S.; Sato, H. Unraveling the role of doping in selective stabilization of NaMnO2 polymorphs: Combined theoretical and experimental study. Chem. Mater. 2018, 30, 1257–1264. [Google Scholar] [CrossRef]
  149. Ma, X.; Chen, H.; Ceder, G. Electrochemical properties of monoclinic NaMnO2. J. Electrochem. Soc. 2011, 158, A1307. [Google Scholar] [CrossRef]
  150. Kubota, K.; Miyazaki, M.; Kim, E.J.; Yoshida, H.; Barpanda, P.; Komaba, S. Structural change induced by electrochemical sodium extraction from layered O’3-NaMnO2. J. Mater. Chem. A 2021, 9, 26810–26819. [Google Scholar] [CrossRef]
  151. Billaud, J.; Clément, R.J.; Armstrong, A.R.; Canales-Vázquez, J.; Rozier, P.; Grey, C.P.; Bruce, P.G. β-NaMnO2: A high-performance cathode for sodium-ion batteries. J. Am. Chem. Soc. 2014, 136, 17243–17248. [Google Scholar] [CrossRef] [PubMed]
  152. Abakumov, A.M.; Tsirlin, A.A.; Bakaimi, I.; Van Tendeloo, G.; Lappas, A. Multiple twinning as a structure directing mechanism in layered rock-salt-type oxides: NaMnO2 polymorphism, redox potentials, and magnetism. Chem. Mater. 2014, 26, 3306–3315. [Google Scholar] [CrossRef]
  153. Yu, H.; Walsh, M.; Liang, X. Improving the comprehensive performance of Na0.7MnO2 for sodium ion batteries by ZrO2 atomic layer deposition. ACS Appl. Mater. Interfaces 2021, 13, 54884–54893. [Google Scholar] [CrossRef] [PubMed]
  154. Wen, Y.; Wang, B.; Zeng, G.; Nogita, K.; Ye, D.; Wang, L. Electrochemical and structural study of layered P2-type Na2/3Ni1/3Mn2/3O2 as cathode material for sodium-ion battery. Chem. Asian J. 2015, 10, 661–666. [Google Scholar] [CrossRef] [PubMed]
  155. Thorne, J.S.; Dunlap, R.A.; Obrovac, M.N. Structure and electrochemistry of NaxFexMn1-xO2 (1.0 ≤ x ≤ 0.5) for Na-ion battery positive electrodes. J. Electrochem. Soc. 2012, 160, A361. [Google Scholar] [CrossRef]
  156. Li, Z.-Y.; Zhang, J.; Gao, R.; Zhang, H.; Hu, Z.; Liu, X. Unveiling the Role of Co in Improving the High-Rate Capability and Cycling Performance of Layered Na0.7Mn0.7Ni0.3-xCoxO2 Cathode Materials for Sodium-Ion Batteries. ACS Appl. Mater. Interfaces 2016, 8, 15439–15448. [Google Scholar] [CrossRef] [PubMed]
  157. Han, M.H.; Gonzalo, E.; Sharma, N.; López del Amo, J.M.; Armand, M.; Avdeev, M.; Saiz Garitaonandia, J.J.; Rojo, T. High-performance P2-phase Na2/3Mn0.8Fe0.1Ti0.1O2 cathode material for ambient-temperature sodium-ion batteries. Chem. Mater. 2016, 28, 106–116. [Google Scholar] [CrossRef]
  158. Gao, X.; Chen, J.; Liu, H.; Yin, S.; Tian, Y.; Cao, X.; Zou, G.; Hou, H.; Wei, W.; Chen, L.; et al. Copper-substituted NaxMO2 (M= Fe, Mn) cathodes for sodium ion batteries: Enhanced cycling stability through suppression of Mn (III) formation. Chem. Eng. J. 2021, 406, 126830. [Google Scholar] [CrossRef]
  159. Bucher, N.; Hartung, S.; Franklin, J.B.; Wise, A.M.; Lim, L.Y.; Chen, H.-Y.; Nelson Weker, J.; Toney, M.F.; Srinivasan, M. P2-NaxCoyMn1-yO2 (y = 0, 0.1) as Cathode Materials in Sodium-Ion Batteries-Effects of Doping and Morphology To Enhance Cycling Stability. Chem. Mater. 2016, 28, 2041–2051. [Google Scholar] [CrossRef]
  160. Xu, S.; Chen, H.; Li, C.; Nie, R.; Yang, Y.; Zhou, M.; Zhang, X.; Zhou, H. Effect of Ti doping on the structural and electrochemical performance of O3-type Na (Ni0.3Fe0.2Mn0.5)1-xTixO2 cathode materials for sodium-ion batteries. J. Alloys Compd. 2023, 962, 171199. [Google Scholar] [CrossRef]
  161. Li, X.; Wu, D.; Zhou, Y.-N.; Liu, L.; Yang, X.-Q.; Ceder, G. O3-type Na (Mn0.25Fe0.25Co0.25Ni0.25)O2: A quaternary layered cathode compound for rechargeable Na ion batteries. Electrochem. Commun. 2014, 49, 51–54. [Google Scholar] [CrossRef]
  162. Yuan, D.D.; Wang, Y.X.; Cao, Y.L.; Ai, X.P.; Yang, H.X. Improved electrochemical performance of Fe-substituted NaNi0.5Mn0.5O2 cathode materials for sodium-ion batteries. ACS Appl. Mater. Interfaces 2015, 7, 8585–8591. [Google Scholar] [CrossRef] [PubMed]
  163. Li, X.; Ge, M.; Zhang, M.; Tang, X.; Liu, X.; Cui, Y.; Zhang, H.; Yang, Y.; Yin, Y.; Yang, S.-T. Prilling and Coating Strategy to Synthesize High-Performance Spherical NaNi0.4Fe0.2Mn0.4O2 Cathode Materials for Sodium Ion Batteries. Langmuir 2024, 40, 18610–18618. [Google Scholar] [CrossRef]
  164. Jiang, L.-W.; Lu, Y.-X.; Wang, Y.-S.; Liu, L.-L.; Qi, X.-G.; Zhao, C.-L.; Chen, L.-Q.; Hu, Y.-S. A high-temperature β-phase NaMnO2 stabilized by Cu doping and its Na storage properties. Chin. Phys. Lett. 2018, 35, 048801. [Google Scholar] [CrossRef]
  165. Wang, X.-Z.; Zuo, Y.; Qin, Y.; Zhu, X.; Xu, S.-W.; Guo, Y.-J.; Yan, T.; Zhang, L.; Gao, Z.; Yu, L.; et al. Fast Na+ Kinetics and Suppressed Voltage Hysteresis Enabled by a High-Entropy Strategy for Sodium Oxide Cathodes. Adv. Mater. 2024, 36, 2312300. [Google Scholar] [CrossRef] [PubMed]
  166. Ding, F.; Zhao, C.; Xiao, D.; Rong, X.; Wang, H.; Li, Y.; Yang, Y.; Lu, Y.-X.; Hu, Y.-S. Using high-entropy configuration strategy to design Na-ion layered oxide cathodes with superior electrochemical performance and thermal stability. J. Am. Chem. Soc. 2022, 18, 8286–8295. [Google Scholar] [CrossRef]
  167. Lamb, J.; Jarvis, K.; Manthiram, A. Molten-salt synthesis of O3-type layered oxide single crystal cathodes with controlled morphology towards long-life sodium-ion batteries. Small 2022, 43, 2106927. [Google Scholar] [CrossRef]
  168. Zhang, L.; Huang, J.; Song, M.; Lu, C.; Wu, W.; Wu, X. Single-crystal growth of P2-type layered oxides with increased exposure of {010} planes for high-performance sodium-ion batteries. ACS Appl. Mater. Interfaces 2023, 40, 47037–47048. [Google Scholar] [CrossRef]
  169. Yang, D.; Long, Y.; Gao, X.-W.; Zhao, Z.; Chen, H.; Lai, Q.; Li, C.; Niu, R.; Liu, Z.; Gu, Q.; et al. Single Crystal Sodium Layered Oxide Achieves Superior Cyclability at High Voltage. Adv. Energy Mater. 2024, 14, 2404999. [Google Scholar] [CrossRef]
  170. Xu, L.; Song, M.; Xie, J.; Chen, M.; Wu, W.; Tan, Z.; Qiu, S.; Wu, X. Suppressing the P2-OP4 phase transition of single-crystal P2-type Ni/Zn/Mn-based layered oxide for advanced sodium-ion batteries. Powder Technol. 2024, 448, 120314. [Google Scholar] [CrossRef]
  171. Zhang, F.; Lu, Y.; Guo, Y.; Li, C.; Liu, Y.; Yang, M.; Zhao, B.; Wu, W.; Wu, X. Highly stabilized single-crystal P2-type layered oxides obtained via rational crystal orientation modulation for sodium-ion batteries. Chem. Eng. J. 2023, 458, 141515. [Google Scholar] [CrossRef]
  172. Xiao, Y.; Wang, H.-R.; Hu, H.-Y.; Zhu, Y.-F.; Li, S.; Li, J.-Y.; Wu, X.-W.; Chou, S.-L. Formulating high-rate and long-cycle heterostructured layered oxide cathodes by local chemistry and orbital hybridization modulation for sodium-ion batteries. Adv. Mater. 2022, 33, 2202695. [Google Scholar] [CrossRef]
  173. Yang, L.; López del Amo, J.M.; Shadike, Z.; Bak, S.-M.; Bonilla, F.; Galceran, M.; Nayak, P.K.; Buchheim, J.R.; Yang, X.-Q.; Rojo, T.; et al. A Co- and Ni-free P2/O3 biphasic lithium stabilized layered oxide for sodium-ion batteries and its cycling behavior. Adv. Funct. Mater. 2020, 42, 2003364. [Google Scholar] [CrossRef]
  174. Zhou, Y.-N.; Wang, P.-F.; Niu, Y.-B.; Li, Q.; Yu, X.; Yin, Y.-X.; Xu, S.; Guo, Y.-G. A P2/P3 composite layered cathode for high-performance Na-ion full batteries. Nano Energy 2019, 55, 143–150. [Google Scholar] [CrossRef]
  175. Zhou, D.; Huang, W.; Lv, X.; Zhao, F. A novel P2/O3 biphase Na0.67Fe0.425Mn0.425Mg0.15O2 as cathode for high-performance sodium-ion batteries. J. Power Sources 2019, 421, 147–155. [Google Scholar] [CrossRef]
  176. Liang, X.; Sun, Y.-K. A novel pentanary metal oxide cathode with P2/O3 biphasic structure for high-performance sodium-ion batteries. Adv. Funct. Mater. 2022, 44, 2206154. [Google Scholar] [CrossRef]
  177. Liu, R.; Huang, W.; Liu, J.; Li, Y.; Wang, J.; Liu, Q.; Ma, L.; Kwon, G.; Ehrlich, S.N.; Wu, Y.; et al. Revealing the Nature of Binary-Phase on Structural Stability of Sodium Layered Oxide Cathodes. Adv. Mater. 2024, 36, 2401048. [Google Scholar] [CrossRef]
  178. Zhou, P.; Che, Z.; Liu, J.; Zhou, J.; Wu, X.; Weng, J.; Zhao, J.; Cao, H.; Zhou, J.; Cheng, F. High-entropy P2/O3 biphasic cathode materials for wide-temperature rechargeable sodium-ion batteries. Energy Storage Mater. 2023, 57, 618–627. [Google Scholar] [CrossRef]
Figure 1. (a) Comparison of redox potentials and gravimetric/volumetric capacities for Li, K, Na, Mg, Al, and S. Reprinted from Ref. [9] with permission from Nature, (b) price trends of lithium carbonate from 2005 onward. Reprinted from Ref. [10], and (c) top 10 countries with the largest lithium reserves, measured in million metric tons. Reprinted from Ref. [11].
Figure 1. (a) Comparison of redox potentials and gravimetric/volumetric capacities for Li, K, Na, Mg, Al, and S. Reprinted from Ref. [9] with permission from Nature, (b) price trends of lithium carbonate from 2005 onward. Reprinted from Ref. [10], and (c) top 10 countries with the largest lithium reserves, measured in million metric tons. Reprinted from Ref. [11].
Micromachines 16 00137 g001
Figure 2. A schematic view of SIB.
Figure 2. A schematic view of SIB.
Micromachines 16 00137 g002
Figure 3. (a) Abundance of elements in the Earth’s crust reprinted from Ref. [21] with permission from Elsevier, (b) global distribution of lithium and sodium reserves as of 2020, represented by red (lithium) and blue (soda ash) circles, with circle size proportional to reserve quantity in metric tons. Light blue shading indicates oceans as additional sodium sources from brine, and (c) price trends for sodium carbonate and lithium carbonate from 2005 to 2019, with an inset highlighting the percentage price changes of both materials over the past decade. Reprinted from Ref. [22] with permission from Wiley.
Figure 3. (a) Abundance of elements in the Earth’s crust reprinted from Ref. [21] with permission from Elsevier, (b) global distribution of lithium and sodium reserves as of 2020, represented by red (lithium) and blue (soda ash) circles, with circle size proportional to reserve quantity in metric tons. Light blue shading indicates oceans as additional sodium sources from brine, and (c) price trends for sodium carbonate and lithium carbonate from 2005 to 2019, with an inset highlighting the percentage price changes of both materials over the past decade. Reprinted from Ref. [22] with permission from Wiley.
Micromachines 16 00137 g003
Figure 5. Structural comparison of layered LiCrO2 (left) and NaCrO2 (right) crystal models. The illustration highlights the differences in interstitial tetrahedral dimensions and their impact on chromium ion migration during electrochemical cycling. Reprinted from Ref. [48] with permission from IOP Publishing.
Figure 5. Structural comparison of layered LiCrO2 (left) and NaCrO2 (right) crystal models. The illustration highlights the differences in interstitial tetrahedral dimensions and their impact on chromium ion migration during electrochemical cycling. Reprinted from Ref. [48] with permission from IOP Publishing.
Micromachines 16 00137 g005
Figure 6. (a) Galvanostatic charge and discharge profiles of a Li/LiCrO2 cell cycled at a current density of 20 mA g−1 within the voltage range of 3.0–4.5 V. Reprinted from Ref. [49] with permission from Elsevier, (b) voltage-specific capacity curve of a NaCrO2 electrode in a Na/NaCrO2 coin cell, cycled between 3.6 and 2.0 V at 25 mA g−1. The inset illustrates the specific capacity versus cycle number. Reprinted from Ref. [52] with permission from IOP Publishing, and (c) initial charge and discharge profiles of NaCrO2 in a Na/NaCrO2 cell cycled at 5.0 mA g−1. Reprinted from Ref. [48] with permission from IOP Publishing.
Figure 6. (a) Galvanostatic charge and discharge profiles of a Li/LiCrO2 cell cycled at a current density of 20 mA g−1 within the voltage range of 3.0–4.5 V. Reprinted from Ref. [49] with permission from Elsevier, (b) voltage-specific capacity curve of a NaCrO2 electrode in a Na/NaCrO2 coin cell, cycled between 3.6 and 2.0 V at 25 mA g−1. The inset illustrates the specific capacity versus cycle number. Reprinted from Ref. [52] with permission from IOP Publishing, and (c) initial charge and discharge profiles of NaCrO2 in a Na/NaCrO2 cell cycled at 5.0 mA g−1. Reprinted from Ref. [48] with permission from IOP Publishing.
Micromachines 16 00137 g006
Figure 7. (a) Galvanostatic charge and discharge curves of a Na/NaCrO2 cell recorded at a current density of 5.0 mA g−1 and (b) the corresponding variation in discharge capacity over multiple cycles, illustrating the cycling performance of the NaCrO2 electrode. Reprinted from Ref. [48] with permission from IOP Publishing.
Figure 7. (a) Galvanostatic charge and discharge curves of a Na/NaCrO2 cell recorded at a current density of 5.0 mA g−1 and (b) the corresponding variation in discharge capacity over multiple cycles, illustrating the cycling performance of the NaCrO2 electrode. Reprinted from Ref. [48] with permission from IOP Publishing.
Micromachines 16 00137 g007
Figure 8. (a) In situ X-ray diffraction (XRD) patterns collected during the first charge cycle of a Na/NaCrO2 cell up to 3.6 V at a C/12 rate. The displayed 2θ angles were recalculated to the corresponding values for Cu-Kα radiation (λ = 1.54 A°) from the synchrotron XRD wavelength (λ = 0.7747 A°). The voltage–composition profile is shown adjacent to the XRD patterns. On the far right, the schematic oxygen stacking arrangements for O3 and P3 structures are illustrated and (b) the evolution of lattice parameters (a, b, and c) as a function of composition during the first charge, derived from in situ XRD data. For clarity, the monoclinic lattice parameters were converted to their rhombohedral equivalents using the transformation equations: aM 3 aR, bM = bR, cM = cR/3 sinβ. Reprinted from Ref. [51] with permission from the Royal Society of Chemistry.
Figure 8. (a) In situ X-ray diffraction (XRD) patterns collected during the first charge cycle of a Na/NaCrO2 cell up to 3.6 V at a C/12 rate. The displayed 2θ angles were recalculated to the corresponding values for Cu-Kα radiation (λ = 1.54 A°) from the synchrotron XRD wavelength (λ = 0.7747 A°). The voltage–composition profile is shown adjacent to the XRD patterns. On the far right, the schematic oxygen stacking arrangements for O3 and P3 structures are illustrated and (b) the evolution of lattice parameters (a, b, and c) as a function of composition during the first charge, derived from in situ XRD data. For clarity, the monoclinic lattice parameters were converted to their rhombohedral equivalents using the transformation equations: aM 3 aR, bM = bR, cM = cR/3 sinβ. Reprinted from Ref. [51] with permission from the Royal Society of Chemistry.
Micromachines 16 00137 g008
Figure 9. (a) In situ XRD patterns collected during the first charge of NaCrO2 at a current density of C/50. A peak originating from the Al current collector overlaps with the 104 reflections of NaCrO2 at approximately 19° (2θ) and is marked with an asterisk and (b) the corresponding electrochemical curve, with scan numbers annotated above it. The graph is divided into four color-shaded regions to emphasize the dominant reactions associated with stacking variations during NaCrO2 desodiation. The critical compositions marking transitions between different reaction regimes are also indicated on the upper x-axis. Reprinted from Ref. [53] with permission from the American Chemical Society.
Figure 9. (a) In situ XRD patterns collected during the first charge of NaCrO2 at a current density of C/50. A peak originating from the Al current collector overlaps with the 104 reflections of NaCrO2 at approximately 19° (2θ) and is marked with an asterisk and (b) the corresponding electrochemical curve, with scan numbers annotated above it. The graph is divided into four color-shaded regions to emphasize the dominant reactions associated with stacking variations during NaCrO2 desodiation. The critical compositions marking transitions between different reaction regimes are also indicated on the upper x-axis. Reprinted from Ref. [53] with permission from the American Chemical Society.
Micromachines 16 00137 g009
Figure 10. (a) Cyclic voltammogram of the first five cycles for the O3-NaCrO2 electrode, ranging from 2.5 V to 3.7 V. The cycles are represented in black (cycle 1), red (cycle 2), green (cycle 3), blue (cycle 4), and magenta (cycle 5). Reprinted from Ref. [54] with permission from Elsevier. The cyclic voltammetry data of O3-NaCrO2 vs. Na/Na+ are shown for different upper potential limits. Two cycles are presented for each potential window, with the first cycle in black and the second in red. The recorded current is plotted against potential. The potential windows used are (b) 2.5–3.5 V, (c) 2.5–3.65 V, (d) 2.5–3.75 V, and (e) 2.5–4.0 V. In (b), the well-known crystalline-to-crystalline phase transformations (O3 → O′3, O′3 → O′3-E, and O′3-E → P′3) are noted. Reprinted from Ref. [43] with permission from Elsevier.
Figure 10. (a) Cyclic voltammogram of the first five cycles for the O3-NaCrO2 electrode, ranging from 2.5 V to 3.7 V. The cycles are represented in black (cycle 1), red (cycle 2), green (cycle 3), blue (cycle 4), and magenta (cycle 5). Reprinted from Ref. [54] with permission from Elsevier. The cyclic voltammetry data of O3-NaCrO2 vs. Na/Na+ are shown for different upper potential limits. Two cycles are presented for each potential window, with the first cycle in black and the second in red. The recorded current is plotted against potential. The potential windows used are (b) 2.5–3.5 V, (c) 2.5–3.65 V, (d) 2.5–3.75 V, and (e) 2.5–4.0 V. In (b), the well-known crystalline-to-crystalline phase transformations (O3 → O′3, O′3 → O′3-E, and O′3-E → P′3) are noted. Reprinted from Ref. [43] with permission from Elsevier.
Micromachines 16 00137 g010
Figure 12. (a) Galvanostatic charge–discharge curves of NaCrO2/NaClO4/Na cells at current densities of 5 mA g−1 (I), 12.5 mA g−1 (II), and 25 mA g−1 (III), (b) specific discharge capacity of NaCrO2/NaClO4/Na cells at a current density of 5 mA g−1 as a function of cycle number. Reprinted from Ref. [69] with permission from Elsevier, (c) electrical conductivity of NaCrO2 as a function of carbon content, (d) first charge and discharge curves comparing electrochemical performances of bare and C-NaCrO2, using Na metal as the counter electrode and 0.5 M NaPF6 in PC:FEC as the electrolyte. Cells were tested in the voltage range of 2.0–3.6 V at a charge-discharge rate of 20 mA g−1, (e) cycling plots showing capacity as a function of the carbon content, (f) rate capability as a function of the carbon content, and (g) rate performances of 3.4 wt% C-NaCrO2, which delivered the best results among various carbon-coated samples. Reprinted from Ref. [19] with permission from the Royal Society of Chemistry.
Figure 12. (a) Galvanostatic charge–discharge curves of NaCrO2/NaClO4/Na cells at current densities of 5 mA g−1 (I), 12.5 mA g−1 (II), and 25 mA g−1 (III), (b) specific discharge capacity of NaCrO2/NaClO4/Na cells at a current density of 5 mA g−1 as a function of cycle number. Reprinted from Ref. [69] with permission from Elsevier, (c) electrical conductivity of NaCrO2 as a function of carbon content, (d) first charge and discharge curves comparing electrochemical performances of bare and C-NaCrO2, using Na metal as the counter electrode and 0.5 M NaPF6 in PC:FEC as the electrolyte. Cells were tested in the voltage range of 2.0–3.6 V at a charge-discharge rate of 20 mA g−1, (e) cycling plots showing capacity as a function of the carbon content, (f) rate capability as a function of the carbon content, and (g) rate performances of 3.4 wt% C-NaCrO2, which delivered the best results among various carbon-coated samples. Reprinted from Ref. [19] with permission from the Royal Society of Chemistry.
Micromachines 16 00137 g012
Figure 13. (a) Mössbauer spectra of initial and fully charged α-NaFeO2 cathodes. The black line represents Fe3+, while the red line corresponds to Fe4+. Reprinted from Ref. [73] with permission from IOP Publishing, (b) schematic illustration of the O3-NaFeO2 structure, where Na and Fe ions are octahedrally coordinated with oxygen atoms, forming a layered structure with Na layers between O-Fe-O slabs, and (c) voltage profiles of NaFeO2 during cycling between 2.0 and 4.4 V versus Na+/Na at a current density of 20 mA g−1, showing potential plateaus and capacity characteristics. Reprinted from Ref. [31] with permission from Elsevier.
Figure 13. (a) Mössbauer spectra of initial and fully charged α-NaFeO2 cathodes. The black line represents Fe3+, while the red line corresponds to Fe4+. Reprinted from Ref. [73] with permission from IOP Publishing, (b) schematic illustration of the O3-NaFeO2 structure, where Na and Fe ions are octahedrally coordinated with oxygen atoms, forming a layered structure with Na layers between O-Fe-O slabs, and (c) voltage profiles of NaFeO2 during cycling between 2.0 and 4.4 V versus Na+/Na at a current density of 20 mA g−1, showing potential plateaus and capacity characteristics. Reprinted from Ref. [31] with permission from Elsevier.
Micromachines 16 00137 g013
Figure 14. (a) Initial charge–discharge profiles of Na/NaFeO2 cells at various cutoff voltages, measured at a rate of 12 mA g−1, (b) charge–discharge curves of Na/NaFeO2 cells with different cutoff voltages at the same rate, (c) discharge capacity retention over 30 cycles for cells charged to 3.4, 4.0, or 4.5 V and discharged to 2.5 V, and (d) XRD patterns of Na1-xFeO2 composite electrodes before and after the initial charge–discharge cycle at different cutoff voltages. The electrodes were charged (oxidized) to 3.4, 4.0, or 4.5 V and discharged to 2.5 V at a rate of 12.1 mA g−1. Reprinted from Ref. [82] with permission from the Electrochemical Society of Japan.
Figure 14. (a) Initial charge–discharge profiles of Na/NaFeO2 cells at various cutoff voltages, measured at a rate of 12 mA g−1, (b) charge–discharge curves of Na/NaFeO2 cells with different cutoff voltages at the same rate, (c) discharge capacity retention over 30 cycles for cells charged to 3.4, 4.0, or 4.5 V and discharged to 2.5 V, and (d) XRD patterns of Na1-xFeO2 composite electrodes before and after the initial charge–discharge cycle at different cutoff voltages. The electrodes were charged (oxidized) to 3.4, 4.0, or 4.5 V and discharged to 2.5 V at a rate of 12.1 mA g−1. Reprinted from Ref. [82] with permission from the Electrochemical Society of Japan.
Micromachines 16 00137 g014
Figure 15. (a) Initial charge and discharge curves of Na/NaFeO2 cells at a rate of 0.2 C (where 1 C corresponds to 100 mA g−1), (b) initial charge and discharge curves of Na/Na4FeRuO6 cells at a rate of 0.2 C, (c) rate performance of Na/Na4FeRuO6 cells at rates of 0.2 C, 0.5 C, 1 C, and 2 C, and (d) cycle performance of Na/Na4FeRuO6 and Na/NaFeO2 cells at a rate of 2 C. Reprinted from Ref. [85] with permission from Wiley, (e) charge–discharge curves of NaFeO2, NaFe0.5Co0.5O2, and NaCoO2 in Na cells at a rate of 12 mA g−1, (f) changes in the discharge capacity over 50 cycles, and (g) rate capability of NaFe0.5Co0.5O2. Cells were charged to 4.0 V at a rate of 12 mA g−1 and discharged at different rates (C/20 = 12 mA g−1 to 30 C = 7260 mA g−1). Reprinted from ref. [86] with permission from Elsevier.
Figure 15. (a) Initial charge and discharge curves of Na/NaFeO2 cells at a rate of 0.2 C (where 1 C corresponds to 100 mA g−1), (b) initial charge and discharge curves of Na/Na4FeRuO6 cells at a rate of 0.2 C, (c) rate performance of Na/Na4FeRuO6 cells at rates of 0.2 C, 0.5 C, 1 C, and 2 C, and (d) cycle performance of Na/Na4FeRuO6 and Na/NaFeO2 cells at a rate of 2 C. Reprinted from Ref. [85] with permission from Wiley, (e) charge–discharge curves of NaFeO2, NaFe0.5Co0.5O2, and NaCoO2 in Na cells at a rate of 12 mA g−1, (f) changes in the discharge capacity over 50 cycles, and (g) rate capability of NaFe0.5Co0.5O2. Cells were charged to 4.0 V at a rate of 12 mA g−1 and discharged at different rates (C/20 = 12 mA g−1 to 30 C = 7260 mA g−1). Reprinted from ref. [86] with permission from Elsevier.
Micromachines 16 00137 g015
Figure 16. (a) Crystal structure of NaCuO2 and (b) discharge–charge curves of a Na/NaCuO2 cell cycled between 0.75 and 3.0 V. The inset illustrates the cycling performance. Reprinted from Ref. [88] with permission from IOP Publishing.
Figure 16. (a) Crystal structure of NaCuO2 and (b) discharge–charge curves of a Na/NaCuO2 cell cycled between 0.75 and 3.0 V. The inset illustrates the cycling performance. Reprinted from Ref. [88] with permission from IOP Publishing.
Micromachines 16 00137 g016
Figure 17. (a) Voltage profiles of NaNiO2 during multiple cycles at C/10, with the cell galvanostatically cycled within the voltage range of 1.25–3.75 V and (b) voltage profiles of NaNiO2 cycled between 2.0–4.5 V. Reprinted from Ref. [94] with permission from IOP Publishing.
Figure 17. (a) Voltage profiles of NaNiO2 during multiple cycles at C/10, with the cell galvanostatically cycled within the voltage range of 1.25–3.75 V and (b) voltage profiles of NaNiO2 cycled between 2.0–4.5 V. Reprinted from Ref. [94] with permission from IOP Publishing.
Micromachines 16 00137 g017
Figure 18. (a) Cyclic voltammetry data of NaNiO2 (main plot) compared to the electrolyte (inset). The CV data illustrate the oxidation and reduction processes of NaNiO2 across multiple cycles, highlighting the reversibility of the material in the 2.0–3.75 V range, (b) XRD patterns of pristine NaNiO2 (black), NaNiO2 mixed with PTFE and carbon black (red), and the cathode material after the first cycle in the 2.0–3.75 V range (blue). The pristine phase corresponds to NaNiO2, while the mixed electrode includes an additional Na0.91NiO2 phase, attributed to minor sodium loss during electrode preparation. The cathode material post-cycling shows only the Na0.91NiO2 phase, with no evidence of reformation of the fully sodiated NaNiO2 phase. Broad peaks between 10° and 30° are due to the Kapton film used during the measurement, and (c) XRD patterns of the cathode after charging to 3.75 V (red) and 4.5 V (blue). Both patterns are largely similar, though the cathode charged to 4.5 V exhibits two additional peaks near 23.7° and 35.0°, indicative of the formation of a new phase. The structure of this new phase remains unidentified. The broad peaks between 10° and 30° are also attributed to the Kapton film. The lattice parameters for NaNiO2 in this study are reported as a = 2.84 A° and c = 20.8 A°. Reprinted from Ref. [94] with permission from IOP Publishing.
Figure 18. (a) Cyclic voltammetry data of NaNiO2 (main plot) compared to the electrolyte (inset). The CV data illustrate the oxidation and reduction processes of NaNiO2 across multiple cycles, highlighting the reversibility of the material in the 2.0–3.75 V range, (b) XRD patterns of pristine NaNiO2 (black), NaNiO2 mixed with PTFE and carbon black (red), and the cathode material after the first cycle in the 2.0–3.75 V range (blue). The pristine phase corresponds to NaNiO2, while the mixed electrode includes an additional Na0.91NiO2 phase, attributed to minor sodium loss during electrode preparation. The cathode material post-cycling shows only the Na0.91NiO2 phase, with no evidence of reformation of the fully sodiated NaNiO2 phase. Broad peaks between 10° and 30° are due to the Kapton film used during the measurement, and (c) XRD patterns of the cathode after charging to 3.75 V (red) and 4.5 V (blue). Both patterns are largely similar, though the cathode charged to 4.5 V exhibits two additional peaks near 23.7° and 35.0°, indicative of the formation of a new phase. The structure of this new phase remains unidentified. The broad peaks between 10° and 30° are also attributed to the Kapton film. The lattice parameters for NaNiO2 in this study are reported as a = 2.84 A° and c = 20.8 A°. Reprinted from Ref. [94] with permission from IOP Publishing.
Micromachines 16 00137 g018
Figure 19. (a) First charge–discharge cycle of NaNiO2 within the voltage range of 1.5–4.0 V at a C/10 rate. Numbers 1 to 8 indicate the specific voltage stops at which ex situ and relaxed in situ XRD measurements were performed. The charge–discharge process is divided into phases marked by plateaus a to d’ that correspond to key voltage ranges and phase transitions observed during the cycling and (b) in situ XRD patterns of the NaNiO2 electrode during the first charge–discharge process, showing corresponding phase transitions. The voltage–composition curve, combined with the XRD patterns, highlights the structural changes that occur as sodium is extracted and reinserted, corresponding to the different phases during charge and discharge, as detailed in the charge–discharge plateaus. The phase transitions are tracked in real-time, revealing the changes in the crystal structure at various stages of sodium deintercalation and reintercalation. Reprinted from Ref. [96] with permission from Elsevier.
Figure 19. (a) First charge–discharge cycle of NaNiO2 within the voltage range of 1.5–4.0 V at a C/10 rate. Numbers 1 to 8 indicate the specific voltage stops at which ex situ and relaxed in situ XRD measurements were performed. The charge–discharge process is divided into phases marked by plateaus a to d’ that correspond to key voltage ranges and phase transitions observed during the cycling and (b) in situ XRD patterns of the NaNiO2 electrode during the first charge–discharge process, showing corresponding phase transitions. The voltage–composition curve, combined with the XRD patterns, highlights the structural changes that occur as sodium is extracted and reinserted, corresponding to the different phases during charge and discharge, as detailed in the charge–discharge plateaus. The phase transitions are tracked in real-time, revealing the changes in the crystal structure at various stages of sodium deintercalation and reintercalation. Reprinted from Ref. [96] with permission from Elsevier.
Micromachines 16 00137 g019
Figure 20. (a,c) Time-resolved XRD patterns of Na0.91NiO2 and Na0.5NiO2 during heating from 30 to 600 °C, (b,d) selected regions showing the two phase transition ranges for each material. Reprinted from Ref. [98] with permission from Elsevier.
Figure 20. (a,c) Time-resolved XRD patterns of Na0.91NiO2 and Na0.5NiO2 during heating from 30 to 600 °C, (b,d) selected regions showing the two phase transition ranges for each material. Reprinted from Ref. [98] with permission from Elsevier.
Micromachines 16 00137 g020
Figure 21. (a) Cyclic performance of Na3PO4@NaNi0.815Co0.15Al0.035O2 at a current density of 100 mA g−1 at room temperature, (b) cyclic performance of Na3PO4@NaNi0.815Co0.15Al0.035O2 at a current density of 200 mA g−1 at room temperature. Reprinted from Ref. [119] with permission from the American Chemical Society, (c) first charge–discharge curves of NNMF-800, NNMF-830, NNMF-850, and NNMF-880 at a rate of 0.2C (1 C = 160 mA g−1), (d) capacity retention curves of NNMF-800, NNMF-830, NNMF-850, and NNMF-880 at a rate of 0.2 C, and (e) rate cycling plots for NNMF-800, NNMF-830, NNMF-850, and NNMF-880. Reprinted from Ref. [120] with permission from Elsevier.
Figure 21. (a) Cyclic performance of Na3PO4@NaNi0.815Co0.15Al0.035O2 at a current density of 100 mA g−1 at room temperature, (b) cyclic performance of Na3PO4@NaNi0.815Co0.15Al0.035O2 at a current density of 200 mA g−1 at room temperature. Reprinted from Ref. [119] with permission from the American Chemical Society, (c) first charge–discharge curves of NNMF-800, NNMF-830, NNMF-850, and NNMF-880 at a rate of 0.2C (1 C = 160 mA g−1), (d) capacity retention curves of NNMF-800, NNMF-830, NNMF-850, and NNMF-880 at a rate of 0.2 C, and (e) rate cycling plots for NNMF-800, NNMF-830, NNMF-850, and NNMF-880. Reprinted from Ref. [120] with permission from Elsevier.
Micromachines 16 00137 g021
Figure 25. (a) Voltage profile of NaMnO2 after multiple cycles at a C/20 rate, (b) voltage profile of NaMnO2 after multiple cycles at a C/30 rate. The cell is galvanostatically cycled between 2.0 V and 3.8 V, (c) comparison of the voltage profiles of NaMnO2 charged up to 3.8 V and 4.2 V, with the cell cycled at a C/20 rate. Reprinted from Ref. [149] with permission from IOP Publishing, and (d) cycling stability of O′3-NaMnO2 over 50 cycles at 12 mA g−1 in the voltage range of 2.0–3.8 V at 25 °C. Reprinted from Ref. [150] with permission from the Royal Society of Chemistry, (e) refinement results exhibiting the changes in interslab distance, and (f) in-plane Mn–Mn distance. Reprinted from Ref. [150] with permission from the Royal Society of Chemistry.
Figure 25. (a) Voltage profile of NaMnO2 after multiple cycles at a C/20 rate, (b) voltage profile of NaMnO2 after multiple cycles at a C/30 rate. The cell is galvanostatically cycled between 2.0 V and 3.8 V, (c) comparison of the voltage profiles of NaMnO2 charged up to 3.8 V and 4.2 V, with the cell cycled at a C/20 rate. Reprinted from Ref. [149] with permission from IOP Publishing, and (d) cycling stability of O′3-NaMnO2 over 50 cycles at 12 mA g−1 in the voltage range of 2.0–3.8 V at 25 °C. Reprinted from Ref. [150] with permission from the Royal Society of Chemistry, (e) refinement results exhibiting the changes in interslab distance, and (f) in-plane Mn–Mn distance. Reprinted from Ref. [150] with permission from the Royal Society of Chemistry.
Micromachines 16 00137 g025
Table 1. General comparison between sodium and lithium batteries.
Table 1. General comparison between sodium and lithium batteries.
CharacteristicSodium BatteriesLithium Batteries
Energy density150–200 Wh kg−1200–300 Wh kg−1
Nominal voltage 3.2–3.3 V3.6–3.7 V
Environmentally friendlyCan be transported at zero voltShould be always stored with a minimum charge
AvailabilityMostly worldwide available at about 500 times higher the lithiumAvailability is limited to some countries and not abundant
Raw material price~330 U.S. dollar per metric ton for Na2CO3~68,000 U.S. dollar per metric ton for Li2CO3
Lifespan5–10 years5–15 years
Table 2. Electrochemical performance of Cr-based and modified Cr-based cathode for SIBs.
Table 2. Electrochemical performance of Cr-based and modified Cr-based cathode for SIBs.
ResearchersPristine MaterialModified MaterialTesting ConditionPristine Material [mAh g−1]Modified Material [mAh g−1]
Ikhe et al. [60]NaCrO2NaCr0.98Al0.02O2@Cr2O32.0–3.6 V, RT *
600 mAh g−1
108.5 (initial) 56 (500 cycles)106.6 (initial) 107 (1000 cycles)
Li et al. [64]NaCrO2Na0.95Cr0.95Ti0.05O22.3–3.6 V, RT
120 mAh g−1
104 (initial) 52 (800 cycles)98 (initial) 78 (800 cycles)
Ma et al. [62]NaCrO2Na0.9Cr0.95Sn0.05O22.5–3.5 V, RT
600 mAh g−1
107 (initial) 50 (1000 cycles)114 (initial) 81 (1000 cycles)
Zheng et al. [46]NaCrO2Na0.9Ca0.05CrO22–3.6 V, RT
120 mAh g−1
110 (initial) 52 (500 cycles)118 (initial) 90 (500 cycles)
Cai et al. [66]NaCrO2NaCrO2@nitrogen doped carbon2.0–3.6 V, RT
1200 mAh g−1
90 (initial) 47.5 (1000 cycles)110 (initial) 100.2 (1000 cycles)
Lee et al. [61]NaCrO2Na0.9Ca0.035Cr0.97Ti0.03O21.5–3.8 V, RT
2400 mAh g−1
34 (initial) 22.1 (500 cycles)76 (initial) 61.7 (500 cycles)
Wu et al. [70]NaCrO2NaCrO2@PVDF-La2(CO3)3·8H2O2.3–3.5 V, RT
240 mAh g−1
119.6 (initial) 56.5 (400 cycles)111.8 (initial) 103.5 (900 cycles)
* RT stands for room temperature.
Table 3. Specific capacities reported for various phases of NaxCoO2.
Table 3. Specific capacities reported for various phases of NaxCoO2.
CompoundPhase TypeDischarge Specific Capacity (mAh g−1)Cycling Rate
Na0.67CoO2 [20]P2105.02 μA cm−2
Na0.74CoO2 [137]P2107.00.1 C
Na0.71CoO2 [138]P269.70.08 C
Na0.74CoO2 [135]P2134.011.1 mA g−1
Na0.67CoO2 [139]P2146.810 mAg−1
Na0.67CoO2 [134]P2146.0100 mA g−1
Na0.7CoO2 [140]P2125.05 mA g−1
Na0.99CoO2 [20]O3118.04.3 μA cm−2
NaCoO2 [86]O3140.012 mA g−1
Na0.57CoO2 [124]P268.00.7 C (0.1 A g−1)
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ahangari, M.; Zhou, M.; Luo, H. Review of Layered Transition Metal Oxide Materials for Cathodes in Sodium-Ion Batteries. Micromachines 2025, 16, 137. https://doi.org/10.3390/mi16020137

AMA Style

Ahangari M, Zhou M, Luo H. Review of Layered Transition Metal Oxide Materials for Cathodes in Sodium-Ion Batteries. Micromachines. 2025; 16(2):137. https://doi.org/10.3390/mi16020137

Chicago/Turabian Style

Ahangari, Mehdi, Meng Zhou, and Hongmei Luo. 2025. "Review of Layered Transition Metal Oxide Materials for Cathodes in Sodium-Ion Batteries" Micromachines 16, no. 2: 137. https://doi.org/10.3390/mi16020137

APA Style

Ahangari, M., Zhou, M., & Luo, H. (2025). Review of Layered Transition Metal Oxide Materials for Cathodes in Sodium-Ion Batteries. Micromachines, 16(2), 137. https://doi.org/10.3390/mi16020137

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

Article Metrics

Back to TopTop