Next Article in Journal
Towards Correlative Raman Spectroscopy–STEM Investigations Performed on a Magnesium–Silver Alloy FIB Lamella
Previous Article in Journal
Ultrathin Gold Nanowires
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mg2+ and Cr3+ Co-Doped LiNi0.5Mn1.5O4 Derived from Ni/Mn Bimetal Oxide as High-Performance Cathode for Lithium-Ion Batteries

1
College of Materials and Metallurgy, Guizhou University, Guiyang 550025, China
2
Guizhou Key Laboratory of Metallurgical Engineering and Process Energy Conservation, Guiyang 550025, China
3
Engineering Technology and Research Center of Manganese Material for Battery, Tongren 554300, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2025, 15(6), 429; https://doi.org/10.3390/nano15060429
Submission received: 16 February 2025 / Revised: 2 March 2025 / Accepted: 4 March 2025 / Published: 11 March 2025
(This article belongs to the Section Energy and Catalysis)

Abstract

:
In this study, pure and Mg2+/Cr3+ co-doped Ni/Mn bimetallic oxides were used as precursors to synthesize pristine and doped LNMO samples. The LNMO samples exhibited the same crystal structure as the precursors. XRD analysis confirmed the successful synthesis of LNMO cathode materials using Ni/Mn bimetallic oxides as precursors. FTIR and Raman spectroscopy reveal that Mg2+/Cr3+ co-doping promotes the formation of the Fd3m disordered phase, effectively reducing electrochemical polarization and charge transfer resistance. Furthermore, co-doping significantly lowers the Mn3+ content on the LNMO surface, thereby mitigating Mn3+ dissolution. Significantly, Mg2+/Cr3+ co-doping induces the emergence of high-surface-energy {100} crystal facets in LNMO grains, which promote lithium-ion transport and, finally, enhance rate capability and cycling performance. Electrochemical analysis indicates that the initial discharge capacities of LNMO-0, LNMO-0.005, LNMO-0.010, and LNMO-0.015 were 126.4, 125.3, 145.3, and 138.2 mAh·g−1, respectively, with capacity retention rates of 82.45%, 82.93%, 83.32%, and 82.08% after 100 cycles. Furthermore, the impedance of LNMO-0.010 prior to cycling was 97.38 Ω, representing a 14.35% reduction compared to the pristine sample. After 100 cycles, its impedance was only 58.61% of that of the pristine sample, highlighting its superior rate capability and cycling stability. As far as we know, studies on the synthesis of LNMO cathode materials via the design of Ni/Mn bimetallic oxides remain limited. Accordingly, this work provides an innovative approach for the preparation and modification of LNMO cathode materials. The investigation of Ni/Mn bimetallic oxides as precursors, combined with co-doping by Mg2+ and Cr3+, for the synthesis of high-performance LiNi0.5Mn1.5O4 (LNMO) aims to provide insights into improving rate capability, cycling stability, reducing impedance, and enhancing capacity retention.

1. Introduction

In recent years, lithium-ion batteries (LIBs) have found widespread application in portable smart devices, electric vehicles, plug-in hybrid vehicles, and energy storage systems due to their superior energy density and cycle life [1,2,3,4,5,6]. To meet the growing demands for enhanced energy density, extended cycle life, and high-rate performance in these applications, it is crucial to improve the performance of LIBs, which is primarily determined by the choice of cathode materials [7]. Among various lithium battery cathodes, the high-voltage spinel LNMO has emerged as one of the most promising candidates for next-generation high energy density LIBs. This is attributed to its high operating voltage (~4.7 V), substantial energy density (658 Wh∙kg−1), and three-dimensional lithium-ion diffusion pathways [8,9,10].
Despite these advantages, the electrochemical performance of synthesized LNMO encounters several significant challenges. The manganese (Mn) within LNMO can dissolve in the electrolyte at elevated temperatures through the disproportionation reaction 2Mn3+→Mn2+ + Mn4+, leading to Jahn–Teller distortion that alters the crystal structure. This alteration compromises the structural integrity of the cathode material, thereby diminishing its stability [11,12,13,14,15]. Moreover, at the high operating voltage of 4.7 V for LNMO, side reactions such as electrolyte decomposition are likely to occur at the electrode/electrolyte interface [16]. These side reactions lead to the formation of a cathode/electrolyte interphase (CEI) layer, which in turn increases the interfacial membrane resistance [17]. Strategies such as ion doping and surface coating have been demonstrated to effectively mitigate these challenges [18,19,20]. However, studies on Mg2+/Cr3+ co-doping in LNMO are relatively scarce, and the synergistic effects of Mg2+ and Cr3+ remain unclear.
The arrangement of Ni and Mn atoms within the spinel structure differs, resulting in LNMO materials that consist of both the disordered Fd3m phase and the ordered P4332 phase with distinct spatial symmetry [21]. In the P4332 ordered structure, Mn ions exhibit a valence state of +4, while in the Fd3m disordered structure, both Mn3+ and Mn4+ coexist. The presence of Mn3+ facilitates ionic transport, leading to ionic conductivity in the Fd3m disordered phase that is two orders of magnitude higher than that of the P4332 ordered phase, thereby enhancing the rate performance of LNMO [22]. However, excessive Mn3+ presence can negatively impact the cycle life of the battery [23]. Moreover, during the battery’s charge/discharge process, compared to the two-phase transition in the P4332 ordered structure, the Fd3m disordered structure undergoes a single-phase transition, which experiences less mechanical stress and benefits the cycling stability of the material [24,25]. Consequently, synthesizing LNMO materials with a higher proportion of the Fd3m disordered phase and appropriate Mn3+ content can simultaneously enhance both rate capability and cycling performance.
The general formula for bimetal oxides is AB2O4, where A represents divalent cations such as Mg2+, Fe2+, or Ni2+, and B represents trivalent cations such as Mn3+, Cr3+, or Al3+. These oxides share the Fd3m space’s group spinel structure with LNMO [26,27]. Bimetal oxides exhibit excellent performance in various fields [28]. Utilizing Ni/Mn bimetal oxides as a doping template facilitates ion doping and mitigates the stress effects associated with crystal structure changes during the calcination process for LNMO synthesis. Furthermore, in NiMn2O4 (NMO), both tetrahedral and octahedral sites can be occupied by Ni and Mn cations, leading to cation disorder within the oxide [29,30]. Importantly, NiMn2O4 exhibits a porous composite structure that enhances the intercalation and deintercalation of lithium-ions, promoting rapid ion transport. Consequently, utilizing Ni/Mn bimetal oxides as precursors for the synthesis of both LNMO and doped LNMO constitutes an effective and feasible strategy. As far as we know, the preparation of LNMO cathodes using Ni/Mn bimetal oxide precursors has rarely been investigated.
In this study, NH3·H2O was utilized as a complexing agent and precipitant to achieve a uniform mixing of Ni and Mn metal ions in solution. Air was subsequently introduced into the system to oxidize the metal hydroxides into metal oxide precipitates, which were then pre-calcined to obtain Ni/Mn bimetal oxides. The resultant Ni/Mn bimetal oxides were mixed with Li2CO3 in stoichiometric ratios and calcined, leading to the transformation of the Ni/Mn bimetal oxides into LNMO. To further enhance the electrochemical performance of LNMO, Mg2+/Cr3+ co-doped Ni/Mn bimetal oxide precursors were synthesized, resulting in a series of co-doped LiNi0.5-x/2Mgx/2Crx/2Mn1.5-2/xO4 (x = 0, 0.005, 0.01, 0.015). The results confirm the successful synthesis of the LNMO cathode material, with ionic co-doping effectively facilitating the formation of the Fd3m disordered structure and suppressing Mn3+ content. The co-doped LNMO demonstrates significantly improved cycling stability and rate performance.

2. Experimental Section

2.1. Material Synthesis

The specific preparation process is detailed in Scheme 1, where the pristine and co-doped LiNi0.5-x/2Mgx/2Crx/2Mn1.5-x/2O4 (x = 0, 0.005, 0.010, 0.015) was synthesized through liquid-phase oxidation followed by two-stage calcination. Initially, (0.5-x/2) mol of NiSO4·6H2O, (1.5-x/2) mol of MnSO4·H2O, (x/2) mol of MgSO4, and (x/2) mol of CrH3O12S3 were mechanically stirred and dissolved in 1000 mL of deionized water, designated solution A. Subsequently, NH3·H2O was mixed with deionized water in a 1:1 ratio to serve as a complexing and precipitating agent, referred to as solution B. A total of 500 mL of deionized water was added to the reactor, which was heated in a water bath while continuously supplying air until the temperature reached 70 °C. Solutions A and B were introduced into the reactor at rates of 1.33 mL∙min−1 and 2.71 mL∙min−1, respectively, and the reaction was maintained for 14 h. Upon completion of the reaction, the black-brown precipitate was filtered, washed several times with deionized water, and dried at 100 °C for 3 h. The precursor was subjected to calcination at 500 °C in a muffle furnace for 10 h to synthesize Ni/Mn bimetallic oxide precursors, designated NMO-0, NMO-0.005, NMO-0.01, and NMO-0.015, corresponding to varying doping content. This precursor was uniformly mixed with stoichiometric Li2CO3 (with an excess of 5 wt% to compensate for lithium loss) and calcined at 850 °C for 20 h. After natural cooling to room temperature, the powder was collected to obtain the product LiNi0.5-x/2Mgx/2Crx/2Mn1.5-x/2O4 (x = 0, 0.005, 0.01, 0.015). The resulting products were designated according to the co-doping content: LNMO-0, LNMO-0.005, LNMO-0.01, and LNMO-0.015.

2.2. Material Characterization and Electrochemical Analysis

All detailed information for the material characterization, computational methods, and electrochemical measurements are shown in the Supporting Information.

3. Results and Discussion

The morphology of the pristine and Mg2+/Cr3+ co-doped LNMO during the preparation process was characterized using scanning electron microscopy (SEM). As illustrated in Figure 1a,e,i,m, the morphology of the pristine and lightly co-doped precursors consists of nanosheets that adhere to and grow on the surfaces of nanospheres, whereas the precursor with 0.015 Mg2+/Cr3+ co-doping is entirely composed of nanosheets. With an increasing Mg2+/Cr3+ content, the size of the nanospheres diminishes, suggesting that Ni and Mn in the precursor exist as more micro-aggregated nanosheets. LNMO precursors prepared via coprecipitation generally exhibit a more disordered spherical morphology [31]. In contrast, the morphology obtained through this method, characterized by the coexistence of nanosheets and nanospheres, is more organized and provides an increased specific surface area for enhanced reactivity. Energy-dispersive spectroscopy (EDS) was employed to assess the chemical composition and elemental distribution of both the pristine and co-doped precursors. Figure S1a–m and Figure 1q–u reveal the presence of Ni, Mn, and O in the pristine precursor, while the co-doped Mg2+/Cr3+ precursor includes Ni, Mn, Mg, Cr, and O. EDS mapping demonstrates that Mg and Cr are uniformly distributed throughout the doped precursor, confirming the successful incorporation of Mg2+ and Cr3+ ions into the precursor.
Figure 1b,f,j,n illustrates the morphologies of the pristine and co-doped NMO precursors. All four samples consist of numerous small octahedral structures that aggregate into spherical shapes. As the Mg2+/Cr3+ content increases, the porosity of the spherical octahedral aggregates also increases, leading to a higher specific surface area. Figure S2 and Table S1 indicate that increasing the content of Mg2+/Cr3+ would enhance the BET surface area, pore volume, and average pore diameter of the NMO precursor, which is consistent with the SEM results. However, the data for NMO-0.010 show a downward trend, which is due to the fact that the voids in the morphology of NMO-0.005 are concentrated and relatively large, while the octahedral structure size of NMO-0.015 becomes smaller. The morphology of NMO-0.010 exhibits a uniform octahedral structure size and an even distribution of voids within the spherical aggregates, which can provide more active sites and diffusion pathways, potentially enhancing the capacity of LNMO. The structures of both pristine and co-doped NMO, as well as LNMO, were analyzed using X-ray diffraction (XRD), and the results for LNMO were subjected to Rietveld refinement. As shown in Figure 2a, the main diffraction peaks of all samples align perfectly with the spinel-type structures of NiMn2O4 (PDF#97-002-7813) and Mn3O4 (PDF#00-013-0162) in the Fd3m space group. It is observed that as the Mg2+/Cr3+ co-doping content increases, the XRD signals gradually shift toward higher angles, indicating the successful incorporation of Mg2+ and Cr3+ into the NMO precursor. Given that the ionic radius of Cr3+ (0.61 Å) is smaller than that of Mn3+ (0.65 Å), the crystal volume decreases [32]. The emergence of the Mn3O4 phase results from the stoichiometric ratio of Ni to Mn in solution A being 1:3, while in the bimetallic oxide NiMn2O4, the ratio is 1:2. This discrepancy leads to the calcination of excess Mn that cannot bond with Ni, resulting in the formation of Mn3O4.
Figure 1c,g,k,o illustrates the morphologies of the pristine and co-doped LNMO samples. All four samples consist of numerous small octahedral crystals that aggregate into spherical shapes, resembling the spherical octahedral aggregates of the NMO precursor; however, they exhibit fewer voids and a more compact arrangement of the octahedral crystals. The spherical aggregation in LNMO-0 and LNMO-0.005 is less pronounced, with the octahedral crystals appearing fragmented and disordered, accompanied by a significant amount of surface impurities. In contrast, LNMO-0.010 and LNMO-0.015 display more pronounced spherical aggregation and cleaner surfaces, free from other impurities. The smaller sizes of the spherical aggregates and octahedral crystals in LNMO-0.015 are attributed to the micro-aggregated nanosheet precursors and the smaller octahedral crystal size observed in the NMO-0.015 precursor. LNMO-0.010 exhibits a favorable morphology with clear spherical aggregation and larger, uniformly distributed octahedral crystals. The larger particles may reduce the contact area between the cathode material and the electrolyte, enhancing structural stability [33,34]. Figure 1d,h,l,p presents the crystal morphologies of LNMO-0, LNMO-0.005, LNMO-0.010, and LNMO-0.015. Notably, the morphology of the primary particles exhibits intriguing variations with increasing Mg2+/Cr3+ co-doping content. The crystal surface orientations correspond to the established {111} plane orientation of the octahedral face-centered cubic framework and their truncated surface orientations [35,36]. The primary particles of LNMO-0 are composed of {111} crystal facets, while the grains of LNMO-0.005, LNMO-0.010, and LNMO-0.015 form new {100} crystal facets at the truncated corners of the octahedral structure. Previous studies indicate that when doping with Cr3+ and calcining at temperatures above 800 °C, the {100} crystal planes, which possess higher surface energy, tend to grow preferentially over the {111} planes [37,38,39]. This suggests that during the calcination of NMO-0, the low-energy {111} planes preferentially grow, resulting in the sacrifice of the {100} and other higher-energy planes [40,41,42]. In contrast, after co-doping with Mg2+/Cr3+, the lattice structure becomes more stable, allowing for greater exposure of high-energy crystal planes during calcination. Among these, the LNMO-0.010 sample exhibits the highest exposure of high-energy crystal planes, which may facilitate the lithium-ion insertion/extraction process, thereby ensuring improved specific capacity and rate performance.
Figure 2b presents the XRD patterns for the synthesis of LiNi0.5-x/2Mgx/2Crx/2Mn1.5-x/2O4 (x = 0, 0.005, 0.01, 0.015). The main diffraction peaks of all samples align with the LiNi0.5Mn1.5O4 structure in the Fd3m space group (PDF#97-009-4762). Remarkably, the diffraction peaks for LNMO-0 and LNMO-0.010 exhibit enhanced intensity and sharpness, indicating a higher degree of crystallinity, which is advantageous for improving charge/discharge capacity and cycling performance [43,44]. As the co-doping content of Mg2+ and Cr3+ increases, a gradual shift of XRD signals toward higher angles is observed, confirming the successful incorporation of these ions into the NMO precursor, leading to a reduction in crystal volume. The LNMO-0.015 sample shows weaker diffraction peaks at 2θ = 35°, 43.5°, and 63.5°, which can be attributed to the presence of rock salt impurity phases (NixO, Ni6MnO8, or Li1-xNixO, marked with *) [45,46]. This phase indicates that, during high-temperature calcination, some Mn in LNMO-0.015 is reduced from Mn4+ to Mn3+ due to lithium volatilization, maintaining charge balance and leading to the formation of disordered phases with an increased Mn3+ content [47]. Figure 2c,d,e,f depicts the Rietveld refinement results for the pristine and Mg2+/Cr3+ co-doped LNMO samples based on the Fd3m space group, where the 8a site is occupied by Li, the 16d sites are occupied by Ni and Mn, and the 32e sites are occupied by O. The lower Rwp values obtained in all refinements indicate a high level of agreement between the measured and fitted values. Table 1 summarizes the refined structural parameters, showing a decreasing trend in crystal parameters, which is consistent with the SEM results. It can be inferred that Mg2+/Cr3+ co-doping contributes to the reduction of crystal structure parameters; specifically, the ionic radius of Cr3+ (0.61 Å) is smaller than that of Ni2+ (0.69 Å), leading to a decrease in crystal volume. Furthermore, this is mainly ascribed to fewer oxygen deficiencies caused by the higher bond strengths of Mg/O (389 kJ∙mol−1) and Cr/O (461 kJ∙mol−1) compared with Ni/O (382 kJ∙mol−1) and Mn/O (402 kJ∙mol−1), in view of the larger ionic radius of Mn3+ than that of Mn4+ [48]. Additionally, the stronger Mg/O and Cr/O bonds contribute to the decrease in crystal volume, thereby reducing the lattice parameters. The I311/I400 ratio serves as an indicator of the distortion degree within the cubic spinel structure and reflects the stability of the crystal structure [49,50]. A smaller I311/I400 ratio signifies a more stable structure and enhanced cycling performance for the LNMO materials. As indicated in Table 1, the I311/I400 ratio decreases with Mg2+/Cr3+ co-doping, suggesting that the crystal structures of the co-doped samples exhibit greater stability, which supports the selection of more high-energy crystal planes during the calcination process of the samples with Mg2+/Cr3+ co-doping.
LNMO materials are generally a mixture of the ordered P4332 and disordered Fd3m structures, with the proportion of the Fd3m space group playing a key role in achieving superior electrochemical performance. Studies indicate that the Fd3m disordered structure offers enhanced electrochemical properties due to its greater charge transport capacity [24]. Since XRD analysis alone cannot reliably distinguish between the Fd3m and P4332 space groups in synthesized LNMO samples [51], FTIR and Raman spectra were employed to assess the Ni/Mn ordering in both pristine and Mg2+/Cr3+ co-doped LNMO, as shown in Figure 3. According to Amatucci et al. [52], the P4332 ordered structure presents eight infrared absorption bands in the range of 400–700 cm−1, whereas the Fd3m disordered structure, due to peak broadening, exhibits only five absorption bands. The FTIR for LNMO-0, LNMO-0.005, LNMO-0.010, and LNMO-0.015 in Figure 3a reveals five absorption bands at 620, 582, 555, 493, and 465 cm−1, indicative of the Fd3m disordered structure. Peaks observed at 620, 582, 555, and 493 cm−1 correspond to the Fd3m disordered structure, while the 465 cm−1 peak is associated with the P4332 ordered structure [53]. Additionally, the signals at 620 cm−1 and 582 cm−1 correspond to symmetric stretching vibrations of Mn/O and Ni/O [10]. The FTIR intensity ratio at 620 cm−1 and 582 cm−1 (I620/I582) qualitatively reflects the disorder level in LNMO, with a higher I620/I582 ratio suggesting increased Ni/Mn disorder. From Figure 3a, the I620/I582 values for LNMO-0, LNMO-0.005, LNMO-0.010, and LNMO-0.015 are 1.43, 1.45, 1.56, and 2.80, respectively, indicating that Mg2+/Cr3+ co-doping partially substitutes Ni2+ and Mn4+, enhancing Ni/Mn disorder at the octahedral 16d sites and increasing the proportion of the Fd3m disordered structure. This trend aligns with the improved electrochemical performance observed upon Mg2+/Cr3+ co-doping. The Raman spectra of LNMO-0, LNMO-0.005, LNMO-0.010, and LNMO-0.015 are shown in Figure 3b. Mn/O stretching is observed at 634 cm−1 [54], while Ni2+/O stretching modes appear at 394 and 494 cm−1 [55]. Compared to the doped LNMO samples, LNMO-0 displays a pronounced peak at 162 cm−1, a feature of the P4332 space group attributed to the reduced symmetry from Ni/Mn ordering [56] and the P4332 polymorph [57,58]. This peak’s intensity diminishes significantly following Mg2+/Cr3+ co-doping, suggesting an increased disorder in LNMO. Additionally, the absence of distinct peak splitting in the 589–634 cm−1 range aligns with the Fd3m disordered structure. Thus, FTIR and Raman analyses confirm that all samples primarily exhibit the Fd3m disordered space group, with only a minor presence of the P4332 ordered structure.
In order to investigate the impact of Mg2+/Cr3+ co-doping on the surface elemental distribution, X-ray Photoelectron Spectroscopy (XPS) was employed to analyze the valence states of elements in both bare and Mg2+/Cr3+ co-doped LNMO. Figure 4a shows the full XPS spectrum of LNMO-0, revealing the presence of Mn, Ni, O, and C on the surface. Figure 4b,d,g,j depicts the Mn2p spectra for LNMO-0, LNMO-0.005, LNMO-0.010, and LNMO-0.015, respectively, all of which can be deconvoluted into two peaks. The characteristic peaks at approximately ~642.0 eV and ~653.7 eV correspond to Mn3+, while those at around ~643.2 eV and ~654.5 eV indicate Mn4+, confirming the simultaneous presence of both Mn3+ and Mn4+ ions on the surfaces of all four samples [59]. Figure 4c and Figure S3a–c display the Ni2p spectra for LNMO-0, LNMO-0.005, LNMO-0.010, and LNMO-0.015, where four characteristic peaks can be observed. The peaks at approximately ~854.7 eV and ~872.7 eV correspond to the main peaks of Ni2+ in the Ni 2p3/2 and Ni 2p1/2 states, respectively. The observed characteristic peaks at approximately ~861.2 eV and ~879.5 eV correspond to the satellite peaks of Ni2p3/2 and Ni2p1/2, respectively, attributed to the multiple splitting of energy levels in nickel oxide (NiO) [59]. These features confirm that Ni in all four samples exists in the +2 oxidation state. Figure 4e,h,k presents the Mg1s spectra for LNMO-0.005, LNMO-0.010, and LNMO-0.015. Remarkably, no Mg1s characteristic peak is observed in LNMO-0.005, likely due to the low doping content of Mg2+ being below the detection limit of XPS. However, the presence of Mg2+ is confirmed in LNMO-0.010 and LNMO-0.015, with Mg1s characteristic peaks appearing at approximately ~1303.5 eV [10,60]. Figure 4f,i,l illustrates the Cr2p spectra for LNMO-0.005, LNMO-0.010, and LNMO-0.015. The absence of Cr2p peaks in LNMO-0.005 is attributed to the Cr3+ doping content being below the detection threshold. In LNMO-0.010 and LNMO-0.015, the characteristic peaks for Cr2p3/2 and Cr2p1/2 appear at approximately ~575.1 eV and ~585.2 eV, respectively, confirming the presence of Cr3+ [2,61]. The characteristic peaks of Mg1s and Cr2p validate the successful doping of Mg2+/Cr3+. Furthermore, the peak areas of Mg1s and Cr2p in LNMO-0.015 are greater than those in LNMO-0.010, indirectly indicating a higher co-doping content of Mg2+/Cr3+ in LNMO-0.015. Since the dissolution of Mn2+ and Mn4+ is generally associated with the disproportionation of Mn3+, a higher Mn3+ content can negatively impact the cycling performance of LNMO materials. Therefore, reducing the Mn3+ content in the surface region may alleviate the dissolution of Mn2+ and Mn4+, thus enhancing the cycling performance of LNMO. The relative content of Mn3+ can be quantified based on the peak area ratio of Mn2p3/2 [62]. As observed in Figure 4b,d,g,j, the Mn3+ content on the particle surface decreases with 0.010 Mg2+/Cr3+ co-doping, whereas it exceeds that of the undoped sample at a doping content of 0.015.
To investigate the Li+ migration kinetics in pristine and Mg2+/Cr3+ co-doped LNMO, cyclic voltammetry (CV) measurements were performed on LNMO/Li half-cells over a voltage range of 3.5–5.0 V at scan rates of 0.05, 0.075, 0.10, 0.125, and 0.150 mV/s. As shown in Figure 5a–d, at a scan rate of 0.1 mV/s, all samples exhibit cathodic and anodic peaks near 4.6 V and 4.8 V, respectively. The redox peaks are split into two distinct peaks, corresponding to the Ni2+/Ni3+ and Ni3+/Ni4+ redox processes [63], with LNMO-0.010 displaying the most pronounced peak splitting. G.B. Zhong et al. reported that the potential difference between Ni2+/Ni3+ and Ni3+/Ni4+ redox peaks is greater in the Fd3m disordered phase than in the P4332 ordered phase [64]. These findings suggest that the relative content of the Fd3m disordered phase among the samples follows the order LNMO-0.015 > LNMO-0.010 > LNMO-0.005 > LNMO-0, consistent with FTIR and Raman spectroscopy results. With the increase in scan rate, the anodic peak (de-lithiation) potentials of the four samples shift to higher values, while the cathodic peak (lithiation) potentials shift to lower values in Figure 5a–d. The potential difference between the anodic and cathodic peaks serves as a reliable indicator of the electrode material’s polarization degree. Table 2 compares the potential differences between the anodic and cathodic peaks of the four samples at different scan rates. At scan rates of 0.05, 0.125, and 0.150 mV/s, LNMO-0.010 exhibits the smallest potential difference, indicating the lowest electrode polarization among all samples. Figure 5e illustrates the linear relationship between the peak current (ip) and the square root of the scan rate (ν1/2). The anodic peak current increases with ν1/2, while the cathodic peak current decreases correspondingly, demonstrating a well-defined linear trend. This relationship is commonly employed to calculate the Li+ diffusion coefficient (DLi) using the Randles–Sevcik equation [65]:
ip = 2.69 × 105An3/2DLi1/2
where A represents the surface area of the electrode (1.1304 cm2), n is the number of electrons transferred per reaction, and C is the volumetric concentration of LNMO (0.02378 mol∙cm⁻3). Table 3 presents the slopes of ip versus ν1/2 and the corresponding DLi values for Li+ extraction/insertion in the four samples. Mg2+/Cr3+ co-doping increases the Li+ extraction rate by an order of magnitude and also enhances the Li+ insertion rate, indicating that co-doping promotes faster Li+ diffusion. Among the co-doped samples, LNMO-0.010 exhibits the highest DLi value, suggesting that a Mg2+/Cr3+ co-doping content of 0.010 improves the electronic conductivity and structural stability of LNMO, leading to superior electrochemical performance. A minor redox peak near 4.0 V, attributed to the Mn3+/Mn4+ redox process, is observed in all samples. Studies have shown that a reduced Mn3+ content mitigates the Jahn–Teller effect and decreases Mn dissolution in the electrolyte [54]. Figure 5f provides a magnified view of the Mn3+/Mn4+ redox peaks at 0.1 mV/s. At doping contents of 0.005 and 0.010, the peak areas at 4.0 V decrease, indicating a reduction in Mn3+ content. However, at a doping content of 0.015, the Mn3+ content surpasses that of the pristine sample, consistent with the Mn3+ relative content obtained from the XPS peak area ratios. This increase can be attributed to the charge compensation mechanism, wherein Mg2+/Cr3+ substitution for Mn4+ leads to an increase in Mn3+ content.
To elucidate the beneficial effects of Mg2+/Cr3+ co-doping on LNMO, electrochemical impedance spectroscopy (EIS) was utilized to assess the impedance and interfacial behavior during the cycling process, activated by running five cycles before the test. Figure 6a,c displays the Nyquist plots obtained before cycling and after 100 cycles at 0.5C. All samples exhibit similar profiles, comprising a semicircle in the high-frequency region and an inclined line in the low-frequency region. The intercept in the high-frequency region corresponds to the ohmic resistance (Rs), which is attributed to the contributions from the electrolyte, electrode material, and separator. The semicircle represents the charge transfer resistance (Rct) at the electrode/electrolyte interface, while the inclined line reflects the Warburg diffusion resistance (Zw), indicative of lithium-ion solid-state diffusion within the electrode material [66]. The Nyquist plots were fitted using software, achieving a high degree of correlation between the experimental data and fitted results. The equivalent circuit model employed for fitting and the derived Rct values are shown in Figure 6a–d. Before cycling, the pristine LNMO (LNMO-0) exhibited a high impedance of 113.7 Ω, while the Rct values of LNMO-0.005 (118.7 Ω) and LNMO-0.015 (109.4 Ω) were comparable to that of LNMO-0. In contrast, LNMO-0.010 demonstrated a significantly lower impedance of 97.38 Ω, representing a reduction of 14.35% relative to the pristine sample. The reduced impedance of LNMO-0.010 can be attributed to the introduction of Mg2+/Cr3+, which enhances conductivity and promotes a disordered structure. After 100 cycles at 0.5C, all samples exhibited increased Rct values. The impedances of LNMO-0, LNMO-0.005, and LNMO-0.015 were approximately 500 Ω. However, LNMO-0.010 exhibited a markedly lower impedance of 279.5 Ω, which is only 58.61% of that of the pristine sample, indicating a substantial reduction. The inability of LNMO-0.005 and LNMO-0.015 to achieve similar impedance reductions is likely due to different factors. The low doping level in LNMO-0.005 failed to effectively substitute Ni and Mn sites, while the higher doping content in LNMO-0.015 led to an increased Mn3+ content, exacerbating Mn dissolution and interfacial side reactions. These findings highlight that Mg2+/Cr3+ co-doping at an optimal content of 0.010 effectively stabilizes the LNMO structure and the cathode/electrolyte interface, thereby enhancing the cycling performance.
The charge/discharge profile serves as a critical metric for assessing the electrochemical performance of LNMO materials. Figure 7a presents the initial charge/discharge curves of the four samples. All samples exhibit two distinct voltage plateaus at approximately 4.7 V, corresponding to the Ni2+/Ni3+ and Ni3+/Ni4+ redox couples. Additionally, a short voltage plateau appears at ~4.0 V in all four samples, attributed to the Mn3+/Mn4+ redox couple, indicating that the Fd3m disordered structure is predominant in all samples. This result is consistent with the findings from XRD, FTIR, Raman, and CV analyses. The Mn3+ content was calculated by dividing the discharge capacity within the 3.8~4.25 V range by the total discharge capacity [67]. Based on Figure 7a, the Mn3+ contents for LNMO-0, LNMO-0.005, LNMO-0.010, and LNMO-0.015 were determined to be 9.18%, 8.81%, 7.21%, and 9.33%, respectively. These values are in good agreement with the results obtained from XPS and CV. The reduced Mn3+ content in LNMO-0.010 effectively suppresses the Jahn–Teller effect and minimizes Mn dissolution, thereby achieving enhanced storage capacity.
The cycling performance of all samples was assessed through galvanostatic charge/discharge testing at 0.5C, as shown in Figure 7b. Prior to testing, each cell underwent 10 activation cycles at 0.2C to stabilize the electrodes. As presented in Figure 7b, the initial discharge capacities of LNMO-0, LNMO-0.005, LNMO-0.010, and LNMO-0.015 were 126.4, 125.3, 145.3, and 138.2 mAh∙g−1, respectively. After 100 cycles, the corresponding discharge capacities were 103.3, 104.84, 121.1, and 113.4 mAh∙g−1, with capacity retention rates of 82.45%, 82.93%, 83.32%, and 82.09%, respectively. The observed capacity degradation during cycling is primarily attributed to the progressive increase in charge transfer resistance (Rct) caused by interfacial reactions at the electrode/electrolyte interface. As shown in Figure 7b, Mg2+/Cr3+ co-doping notably enhances both the initial discharge capacity and cycling stability of the samples. This improvement is largely attributed to the higher binding energy of Mg/O and Cr/O bonds, which reinforces the structural stability of LNMO. Additionally, the three electrons in the third orbital of Cr3+ effectively mitigate the Jahn–Teller effect. Compared to LNMO-0.010, the performance of LNMO-0.005 is hindered by its insufficient doping level, which fails to fully capitalize on the advantages of Mg2+/Cr3+ co-doping. Conversely, LNMO-0.015 exhibits reduced cycling stability due to excessive doping, which increases Mn3+ content through charge compensation during Mg2+/Cr3+ substitution of Mn4+, thereby exacerbating side reactions and capacity fade. LNMO-0.010 exhibits better cycling performance, primarily due to the reduced Mn3+ content induced by doping and the increase in {100} crystal facets, which consequently reduce the proportion of {111} crystal facets. Previous studies have demonstrated that the cycling stability of LNMO is strongly influenced by surface orientation exposed to the electrolyte [35], and the exposed {111} crystal facets increase Mn dissolution and induce unstable interfacial behavior at higher voltages. This insight explains the rapid capacity decay observed in LNMO-0, whose particles predominantly expose {111} planes. Therefore, Mg2+/Cr3+ co-doping at an optimal concentration of 0.010 provides the most significant enhancement in both initial capacity and cycling stability, establishing LNMO-0.010 as the best-performing material. Figure 7b further depicts the Coulombic efficiency trends during cycling. The initial Coulombic efficiency is relatively low, which can be attributed to inevitable electrolyte decomposition and side reactions at the electrode/electrolyte interface under high voltage conditions [68]. However, a stable interfacial layer forms during cycling, leading to the subsequent stabilization of Coulombic efficiency in later cycles.
The differential capacity curve (dQ/dV) is a valuable tool for investigating the capacity fading mechanisms in LIBs, as the characteristic peaks on the curve correspond to the redox peaks observed in CV curves [69]. Figure 7c illustrates the dQ/dV curves of the four samples during the first and 100th cycles at 0.5C, revealing three pairs of redox peaks. The short plateau at approximately 4.0 V is associated with the Mn3+/Mn4+ redox couple, while the two redox peaks near 4.7 V are attributed to the Ni2+/Ni3+ and Ni3+/Ni4+ redox processes, respectively [70]. As shown in Figure 7c, after 100 cycles, the anodic peaks of all four samples shift to higher potentials while the cathodic peaks shift to lower potentials, and the potential difference (ΔV) increases, indicating an increased degree of electrode polarization after cycling. When considered alongside the observed increase in impedance after 100 cycles, it is evident that side reactions at the electrode/electrolyte interface and electrolyte decomposition play significant roles in this process. The deposition of decomposition products on the electrode surface obstructs Li+ transport and intensifies polarization, thereby contributing to capacity degradation in the LNMO samples.
To assess the rate performance of the prepared samples, the cells were first cycled at 0.2C for several cycles to activate the electrodes, followed by charge/discharge testing at sequential rates of 0.2C, 0.5C, 1C, 5C, and 10C, and finally returned to 0.2C. The corresponding rate performance curves are shown in Figure 7d. After the rate performance tests, the capacities of all samples nearly returned to their initial values, demonstrating the structural stability of the materials during rapid lithium insertion/extraction processes. The introduction of Cr3+ dopants extends the Li/O bond length, facilitating the separation of Li+ from O2− and enhancing lithium-ion migration within the lattice [71]. Furthermore, spinel-type materials with high surface energy provide advantageous conditions for Li+ diffusion during charge/discharge processes [35]. As a result, LNMO-0.010, characterized by a lower Mn3+ content, fewer impurity phases, a higher proportion of Fd3m disordered structure, and prominent {100} crystal planes, exhibited superior capacities across different rates. At low rates of 0.2C and 0.5C, all samples displayed excellent performance. However, at higher rates of 5C and 10C, the capacities of the materials declined significantly. The rate capability of the materials is primarily governed by their morphology, particle size, and the availability of three-dimensional Li+ diffusion pathways. As shown in Figure 1c,g,k,o, LNMO-0, LNMO-0.005, LNMO-0.010, and LNMO-0.015 are composed of spherical aggregates formed by densely packed micro-octahedral structures. This compact porous morphology slows Li+ migration, leading to reduced capacities at high charge/discharge rates.

4. Conclusions

Based on the principle of crystal structure consistency, we successfully synthesized pristine disordered-phase LNMO samples using Ni/Mn bimetallic oxides as precursors. Structural characterizations (XRD, FTIR, Raman) and electrochemical performance (initial discharge capacity of 126.4 mAh·g−1) indicate the significant application potential of this approach. Furthermore, Mg2+/Cr3+ co-doped Ni/Mn bimetallic oxides were employed to prepare co-doped LNMO samples. Compared with the pristine LNMO, Mg2+/Cr3+ co-doping enhanced the degree of Ni/Mn disorder in the LNMO cathode material, reduced the surface Mn3+ content, and decreased the charge transfer resistance, thereby improving both the rate capability and cycling stability. Additionally, SEM analysis further revealed that Mg2+/Cr3+ co-doping facilitated the formation of high-surface-energy {100} crystal facets in LNMO grains, which promote lithium-ion migration. The 0.010 Mg2+/Cr3+ co-doped sample (LNMO-0.010) exhibited a significantly improved initial discharge capacity of 145.3 mAh·g−1 at 0.5C. Additionally, due to reduced electrode polarization and suppressed Mn3+ dissolution, this sample exhibited less capacity degradation compared to the pristine LNMO, achieving capacity retention of 83.32% after 100 cycles, thus demonstrating superior rate capability and cycling performance. This study offers an innovative and effective strategy for the synthesis and modification of LNMO cathode materials.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano15060429/s1, Figure S1: EDS mapping of the pristine and co-doped precursors. (a–c) EDS mapping of the pristine precursor; (d–h) EDS mapping of the precursor for the 0.005 Mg-Cr co-doped sample; (i–m) EDS mapping of the precursor for the 0.015 Mg-Cr co-doped sample.; Figure S2: Isothermal adsorption-desorption curves and pore size distribution profiles for both pristine and doped NMO samples were obtained. Figure S3: Ni XPS spectra of LNMO-0.005, LNMO-0.010, LNMO-0.015. (a) Ni XPS spectra of LNMO-0.005; (b) Ni XPS spectra of LNMO-0.010; (c) Ni XPS spectra of LNMO-0.015. Table S1: The BET specific surface area, total pore volume, and average pore diameter of both pristine and co-doped NMO samples were evaluated.

Author Contributions

Conceptualization, Z.P., Q.W., G.Q., X.Z., K.Z. and J.L.; methodology, D.M., Q.W., K.Z. and J.L.; validation, D.M.; formal analysis, Z.P. and X.Z.; investigation, Z.P.; writing—original draft, D.M.; writing—review & editing, D.M.; supervision, H.W.; project administration, J.W. and H.W.; funding acquisition, J.W. and H.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Guizhou Provincial Basic Research Program ([2025] Key 095), Guizhou Provincial Central Government Guides Local Science and Technology Development Fund ([2025] 026), Guizhou Provincial Science and Technology Major Special Project ([2024] 93), Guizhou Provincial Key Science and Technology Support Program ([2022] Key 020), and Guizhou Provincial Major Special Project ([2022] 003).

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no competing financial interests.

References

  1. Bai, X.; Jiang, Z.Y.; Sun, Y.Z.; Liu, X.G.; Jin, X.; He, R.; Liu, Z.F.; Pan, J.Q. Clean Universal Solid-State Recovery Method of Waste Lithium-Ion Battery Ternary Positive Materials and Their Electrochemical Properties. ACS Sustain. Chem. Eng. 2023, 11, 3673–3686. [Google Scholar] [CrossRef]
  2. Gong, J.J.; Yan, S.P.; Lang, Y.Q.; Zhang, Y.; Fu, S.X.; Guo, J.L.; Wang, L.; Liang, G.C. Effect of Cr3+ doping on morphology evolution and electrochemical performance of LiNi0.5Mn1.5O4 material for Li-ion battery. J. Alloys Compd. 2021, 859, 157885. [Google Scholar] [CrossRef]
  3. Xiao, B.; Liu, H.; Liu, J.; Sun, Q.; Wang, B.; Kaliyappan, K.; Zhao, Y.; Banis, M.N.; Liu, Y.; Li, R. Nanoscale Manipulation of Spinel Lithium Nickel Manganese Oxide Surface by Multisite Ti Occupation as High-Performance Cathode. Adv. Mater. 2017, 29, 1703764. [Google Scholar] [CrossRef] [PubMed]
  4. Xiao, B.; Sun, X. Surface and subsurface reactions of lithium transition metal oxide cathode materials: An overview of the fundamental origins and remedying approaches. Adv. Energy Mater. 2018, 8, 1802057. [Google Scholar] [CrossRef]
  5. Zhou, S.; Mei, T.; Li, J.; Pi, W.; Wang, J.; Li, J.; Wang, X. Hierarchical LiNi0.5Mn1.5O4 micro-rods with enhanced rate performance for lithium-ion batteries. J. Mater. Sci. 2018, 53, 9710–9720. [Google Scholar] [CrossRef]
  6. Divakaran, A.M.; Minakshi, M.; Bahri, P.A.; Paul, S.; Kumari, P.; Divakaran, A.M.; Manjunatha, K.N. Rational design on materials for developing next generation lithium-ion secondary battery. Prog. Solid State Chem. 2021, 62, 100298. [Google Scholar] [CrossRef]
  7. Ma, J.; Hu, P.; Cui, G.; Chen, L. Surface and interface issues in spinel LiNi0.5Mn1.5O4: Insights into a potential cathode material for high energy density lithium ion batteries. Chem. Mater. 2016, 28, 3578–3606. [Google Scholar] [CrossRef]
  8. Hsiao, Y.-S.; Huang, J.-H.; Cheng, T.-H.; Hu, C.-W.; Wu, N.-J.; Yen, C.-Y.; Hsu, S.-C.; Weng, H.C.; Chen, C.-P. Cr-doped LiNi0.5Mn1.5O4 derived from bimetallic Ni/Mn metal-organic framework as high-performance cathode for lithium-ion batteries. J. Energy Storage 2023, 68, 107686. [Google Scholar] [CrossRef]
  9. Kim, J.H.; Pieczonka, N.P.; Yang, L. Challenges and approaches for high-voltage spinel lithium-ion batteries. ChemPhysChem 2014, 15, 1940–1954. [Google Scholar] [CrossRef]
  10. Wei, A.; Li, W.; Chang, Q.; Bai, X.; He, R.; Zhang, L.; Liu, Z.; Wang, Y. Effect of Mg2+/F co-doping on electrochemical performance of LiNi0.5Mn1.5O4 for 5 V lithium-ion batteries. Electrochim. Acta 2019, 323, 134692. [Google Scholar] [CrossRef]
  11. Li, B.Z.; Xing, L.D.; Xu, M.Q.; Lin, H.B.; Li, W.S. New solution to instability of spinel LiNi0.5Mn1.5O4 as cathode for lithium ion battery at elevated temperature. Electrochem. Commun. 2013, 34, 48–51. [Google Scholar] [CrossRef]
  12. Li, W.; Christiansen, T.L.; Li, C.; Zhou, Y.; Fei, H.; Mamakhel, A.; Iversen, B.B.; Watkins, J.J. High-power lithium-ion microbatteries from imprinted 3D electrodes of sub-10 nm LiMn2O4/Li4Ti5O12 nanocrystals and a copolymer gel electrolyte. Nano Energy 2018, 52, 431–440. [Google Scholar] [CrossRef]
  13. Lin, M.Y.; Ben, L.; Sun, Y.; Wang, H.; Yang, Z.; Gu, L.; Yu, X.; Yang, X.Q.; Zhao, H.; Yu, R.; et al. Insight into the Atomic Structure of High-Voltage Spinel LiNi0.5Mn1.5O4 Cathode Material in the First Cycle. Chem. Mater. 2015, 27, 292–303. [Google Scholar] [CrossRef]
  14. Xu, M.Q.; Liu, Y.L.; Li, B.; Li, W.S.; Li, X.P.; Hu, S.J. Tris (pentafluorophenyl) phosphine: An electrolyte additive for high voltage Li-ion batteries. Electrochem. Commun. 2012, 18, 123–126. [Google Scholar] [CrossRef]
  15. Yin, C.J.; Bao, Z.Q.; Tan, H.; Zhou, H.M.; Li, J. Metal-organic framework-mediated synthesis of LiNi0.5Mn1.5O4: Tuning the Mn3+ content and electrochemical performance by organic ligands. Chem. Eng. J. 2019, 372, 408–419. [Google Scholar] [CrossRef]
  16. Banerjee, A.; Shilina, Y.; Ziv, B.; Ziegelbauer, J.M.; Luski, S.; Aurbach, D.; Halalay, I.C. On the Oxidation State of Manganese Ions in Li-Ion Battery Electrolyte Solutions. J. Am. Chem. Soc. 2017, 139, 1738–1741. [Google Scholar] [CrossRef]
  17. Li, J.; Zhang, Q.; Xiao, X.; Cheng, Y.-T.; Liang, C.; Dudney, N.J. Unravelling the Impact of Reaction Paths on Mechanical Degradation of Intercalation Cathodes for Lithium-Ion Batteries. J. Am. Chem. Soc. 2015, 137, 13732–13735. [Google Scholar] [CrossRef]
  18. Benavides, L.A.; Sergio Moreno, M.; Cuscueta, D.J. Improving the electrochemical performance of LiNi0.5Mn1.5O4 by ZnO nanocrystals coating. Mater. Sci. Eng. B 2024, 303, 117307. [Google Scholar] [CrossRef]
  19. Qureshi, Z.A.; Ali, M.E.S.; Shakoor, R.A.; AlQaradawi, S.; Kahraman, R. Impact of synergistic interfacial modification on the electrochemical performance of LiNi0.5Mn1.5O4 cathode materials. Ceram. Int. 2024, 50, 17818–17835. [Google Scholar] [CrossRef]
  20. Zhang, Z.; Chang, C.; Zheng, J. Stabilizing the Mn4+/Mn3+ redox at 2.7 V plateau in LiNi0.5Mn1.5O4 cathode material for high energy density and long cycle lifetime via Mo6+ doping. J. Energy Storage 2024, 99, 113226. [Google Scholar] [CrossRef]
  21. Liang, G.; Didier, C.; Guo, Z.; Pang, W.K.; Peterson, V.K. Understanding Rechargeable Battery Function Using in Operando Neutron Powder Diffraction. Adv. Mater. 2020, 32, 1904528. [Google Scholar] [CrossRef] [PubMed]
  22. Liu, H.; Zhu, G.; Zhang, L.; Qu, Q.; Shen, M.; Zheng, H. Controllable synthesis of spinel lithium nickel manganese oxide cathode material with enhanced electrochemical performances through a modified oxalate co-precipitation method. J. Power Sources 2015, 274, 1180–1187. [Google Scholar] [CrossRef]
  23. Lee, J.; Kim, C.; Kang, B. High electrochemical performance of high-voltage LiNi0.5Mn1.5O4 by decoupling the Ni/Mn disordering from the presence of Mn3+ ions. NPG Asia Mater. 2015, 7, e211. [Google Scholar] [CrossRef]
  24. Kim, J.-H.; Myung, S.-T.; Yoon, C.; Kang, S.; Sun, Y.-K. Comparative study of LiNi0.5Mn1.5O4−δ and LiNi0.5Mn1.5O4 cathodes having two crystallographic structures: Fd 3m and P4332. Chem. Mater. 2004, 16, 906–914. [Google Scholar] [CrossRef]
  25. Wu, J.; Tsai, C.-J. Qualitative modeling of the electrolyte oxidation in long-term cycling of LiCoPO4 for high-voltage lithium-ion batteries. Electrochim. Acta 2021, 368, 137585. [Google Scholar] [CrossRef]
  26. Cabral, A.J.F.; Remédios, C.M.R.; Gratens, X.; Chitta, V.A. Effects of microstructure on the magnetic properties of polycrystalline NiMn2O4 spinel oxides. J. Magn. Magn. Mater. 2019, 469, 108–112. [Google Scholar] [CrossRef]
  27. O’Neill, H.S.C.; Navrotsky, A. Simple spinels; crystallographic parameters, cation radii, lattice energies, and cation distribution. Am. Mineral. 1983, 68, 181–194. [Google Scholar]
  28. Watcharatharapong, T.; Minakshi Sundaram, M.; Chakraborty, S.; Li, D.; Shafiullah, G.M.; Aughterson, R.D.; Ahuja, R. Effect of Transition Metal Cations on Stability Enhancement for Molybdate-Based Hybrid Supercapacitor. ACS Appl. Mater. Interfaces 2017, 9, 17977–17991. [Google Scholar] [CrossRef]
  29. Brabers, V.A.M.; van Setten, F.M.; Knapen, P.S.A. X-ray photoelectron spectroscopy study of the cation valencies in nickel manganite. J. Solid State Chem. 1983, 49, 93–98. [Google Scholar] [CrossRef]
  30. Sinha, A.; Sanjana, N.; Biswas, A. On the structure of some anganites. Acta Crystallogr. 1957, 10, 439–440. [Google Scholar] [CrossRef]
  31. Liu, J.; Han, Y.; Ke, W.; Tang, J.; Xu, G.; Deng, L.; Zhao, L.; Wang, Z. Suppressed phase separation in high-voltage LiNi0.5Mn1.5O4 cathode via Co-doping. J. Alloys Compd. 2024, 993, 174587. [Google Scholar] [CrossRef]
  32. Li, J.; Li, S.; Xu, S.; Huang, S.; Zhu, J. Synthesis and Electrochemical Properties of LiNi0.5Mn1.5O4 Cathode Materials with Cr3+ and F Composite Doping for Lithium-Ion Batteries. Nanoscale Res. Lett. 2017, 12, 414. [Google Scholar] [CrossRef]
  33. Kim, G.-H.; Myung, S.T.; Bang, H.J.; Prakash, J.; Sun, Y.-K. Synthesis and Electrochemical Properties of Li[Ni1/3Co1/3Mn(1/3-x)Mgx] O2-y Fy via Coprecipitation. Electrochem. Solid State Lett. 2004, 7, A477. [Google Scholar] [CrossRef]
  34. Luo, Y.; Li, H.; Lu, T.; Zhang, Y.; Mao, S.S.; Liu, Z.; Wen, W.; Xie, J.; Yan, L. Fluorine gradient-doped LiNi0.5Mn1.5O4 spinel with improved high voltage stability for Li-ion batteries. Electrochim. Acta 2017, 238, 237–245. [Google Scholar] [CrossRef]
  35. Chen, Z.; Zhao, R.; Du, P.; Hu, H.; Wang, T.; Zhu, L.; Chen, H. Polyhedral LiNi0.5Mn1.5O4 with excellent electrochemical properties for lithium-ion batteries. J. Mater. Chem. A 2014, 2, 12835–12848. [Google Scholar] [CrossRef]
  36. Kim, J.-S.; Kim, K.; Cho, W.; Shin, W.H.; Kanno, R.; Choi, J.W. A Truncated Manganese Spinel Cathode for Excellent Power and Lifetime in Lithium-Ion Batteries. Nano Lett. 2012, 12, 6358–6365. [Google Scholar] [CrossRef]
  37. Li, Y.; Tan, H.; Yang, X.-Y.; Goris, B.; Verbeeck, J.; Bals, S.; Colson, P.; Cloots, R.; Van Tendeloo, G.; Su, B.-L. Well Shaped Mn3O4 Nano-octahedra with Anomalous Magnetic Behavior and Enhanced Photodecomposition Properties. Small 2011, 7, 475–483. [Google Scholar] [CrossRef] [PubMed]
  38. Liu, H.; Wang, J.; Zhang, X.; Zhou, D.; Qi, X.; Qiu, B.; Fang, J.; Kloepsch, R.; Schumacher, G.; Liu, Z.; et al. Morphological Evolution of High-Voltage Spinel LiNi0.5Mn1.5O4 Cathode Materials for Lithium-Ion Batteries: The Critical Effects of Surface Orientations and Particle Size. ACS Appl. Mater. Interfaces 2016, 8, 4661–4675. [Google Scholar] [CrossRef]
  39. Liu, Y.; Li, J.; Zeng, M.; Huang, Y.; Xu, X.; Yan, M.; Guo, J.; Deng, J.; Yang, J. Octahedral nano-particles constructed LiNi0.5Mn1.5O4 microspheres as high-voltage cathode materials for long-life lithium-ion batteries. Ceram. Int. 2018, 44, 20043–20048. [Google Scholar] [CrossRef]
  40. Chen, Z.; Qiu, S.; Cao, Y.; Ai, X.; Xie, K.; Hong, X.; Yang, H. Surface-oriented and nanoflake-stacked LiNi0.5Mn1.5O4 spinel for high-rate and long-cycle-life lithium ion batteries. J. Mater. Chem. 2012, 22, 17768–17772. [Google Scholar] [CrossRef]
  41. Chung, K.Y.; Yoon, W.-S.; Lee, H.S.; Yang, X.-Q.; McBreen, J.; Deng, B.H.; Wang, X.Q.; Yoshio, M.; Wang, R.; Gui, J.; et al. Comparative studies between oxygen-deficient LiMn2O4 and Al-doped LiMn2O4. J. Power Sources 2005, 146, 226–231. [Google Scholar] [CrossRef]
  42. Huang, M.-R.; Lin, C.-W.; Lu, H.-Y. Crystallographic facetting in solid-state reacted LiMn2O4 spinel powder. Appl. Surf. Sci. 2001, 177, 103–113. [Google Scholar] [CrossRef]
  43. Sun, Y.-K.; Lee, K.-H.; Moon, S.-I.; Oh, I.-H. Effect of crystallinity on the electrochemical behaviour of spinel Li1.03Mn2O4 cathode materials. Solid State Ion. 1998, 112, 237–243. [Google Scholar] [CrossRef]
  44. Thackeray, M.M. Structural Considerations of Layered and Spinel Lithiated Oxides for Lithium Ion Batteries. J. Electrochem. Soc. 1995, 142, 2558. [Google Scholar] [CrossRef]
  45. Jafta, C.; Mathe, M.; Manyala, N.; Roos, W.; Ozoemena, K. Microwave-Assisted Synthesis of High-Voltage Nanostructured LiMn1.5Ni0.5O4 Spinel: Tuning the Mn3+ Content and Electrochemical Performance. ACS Appl. Mater. Interfaces 2013, 5, 7592–7598. [Google Scholar] [CrossRef]
  46. Kim, J.-H.; Huq, A.; Chi, M.; Pieczonka, N.P.W.; Lee, E.; Bridges, C.A.; Tessema, M.M.; Manthiram, A.; Persson, K.A.; Powell, B.R. Integrated Nano-Domains of Disordered and Ordered Spinel Phases in LiNi0.5Mn1.5O4 for Li-Ion Batteries. Chem. Mater. 2014, 26, 4377–4386. [Google Scholar] [CrossRef]
  47. Song, J.; Shin, D.W.; Lu, Y.; Amos, C.D.; Manthiram, A.; Goodenough, J.B. Role of Oxygen Vacancies on the Performance of Li[Ni0.5–xMn1.5+x]O4 (x = 0, 0.05, and 0.08) Spinel Cathodes for Lithium-Ion Batteries. Chem. Mater. 2012, 24, 3101–3109. [Google Scholar] [CrossRef]
  48. Nie, X.; Zhong, B.; Chen, M.; Yin, K.; Li, L.; Liu, H.; Guo, X. Synthesis of LiCr0.2Ni0.4Mn1.4O4 with superior electrochemical performance via a two-step thermo polymerization technique. Electrochim. Acta 2013, 97, 184–191. [Google Scholar] [CrossRef]
  49. Bao, S.-J.; Li, C.-M.; Li, H.-L.; Luong, J.H.T. Morphology and electrochemistry of LiMn2O4 optimized by using different Mn-sources. J. Power Sources 2007, 164, 885–889. [Google Scholar] [CrossRef]
  50. Feng, J.; Huang, Z.; Guo, C.; Chernova, N.A.; Upreti, S.; Whittingham, M.S. An Organic Coprecipitation Route to Synthesize High Voltage LiNi0.5Mn1.5O4. ACS Appl. Mater. Interfaces 2013, 5, 10227–10232. [Google Scholar] [CrossRef]
  51. Wang, J.; Lin, W.; Wu, B.; Zhao, J. Syntheses and electrochemical properties of the Na-doped LiNi0.5Mn1.5O4 cathode materials for lithium-ion batteries. Electrochim. Acta 2014, 145, 245–253. [Google Scholar] [CrossRef]
  52. Kunduraci, M.; Al-Sharab, J.F.; Amatucci, G.G. High-Power Nanostructured LiMn2-xNixO4 High-Voltage Lithium-Ion Battery Electrode Materials:  Electrochemical Impact of Electronic Conductivity and Morphology. Chem. Mater. 2006, 18, 3585–3592. [Google Scholar] [CrossRef]
  53. Wei, A.; Mu, J.; He, R.; Bai, X.; Liu, Z.; Zhang, L.; Wang, Y.; Liu, Z. Enhancing electrochemical performance and structural stability of LiNi0.5Mn1.5O4 cathode material for rechargeable lithium-ion batteries by boron doping. Ceram. Int. 2021, 47, 226–237. [Google Scholar] [CrossRef]
  54. Xiao, B.; Liu, J.; Sun, Q.; Wang, B.; Banis, M.N.; Zhao, D.; Wang, Z.; Li, R.; Cui, X.; Sham, T.-K.; et al. Unravelling the Role of Electrochemically Active FePO4 Coating by Atomic Layer Deposition for Increased High-Voltage Stability of LiNi0.5Mn1.5O4 Cathode Material. Adv. Sci. 2015, 2, 1500022. [Google Scholar] [CrossRef]
  55. Zheng, J.; Xiao, J.; Yu, X.; Kovarik, L.; Gu, M.; Omenya, F.; Chen, X.; Yang, X.-Q.; Liu, J.; Graff, G.L.; et al. Enhanced Li+ ion transport in LiNi0.5Mn1.5O4 through control of site disorder. Phys. Chem. Chem. Phys. 2012, 14, 13515–13521. [Google Scholar] [CrossRef]
  56. Wang, L.; Li, H.; Huang, X.; Baudrin, E. A comparative study of Fd-3m and P4332 “LiNi0.5Mn1.5O4”. Solid State Ion. 2011, 193, 32–38. [Google Scholar] [CrossRef]
  57. Keppeler, M.; Nageswaran, S.; Kim, S.-J.; Srinivasan, M. Silicon Doping of High Voltage Spinel LiNi0.5Mn1.5O4 towards Superior Electrochemical Performance of Lithium Ion Batteries. Electrochim. Acta 2016, 213, 904–910. [Google Scholar] [CrossRef]
  58. Wang, J.; Lin, W.; Wu, B.; Zhao, J. Porous LiNi0.5Mn1.5O4 sphere as 5 V cathode material for lithium ion batteries. J. Mater. Chem. A 2014, 2, 16434–16442. [Google Scholar] [CrossRef]
  59. Hao, X.; Austin, M.H.; Bartlett, B.M. Two-step hydrothermal synthesis of submicron Li1+xNi0.5Mn1.5O4−δ for lithium-ion battery cathodes (x = 0.02, δ = 0.12). Dalton Trans. 2012, 41, 8067–8076. [Google Scholar] [CrossRef]
  60. Minakshi, M.; Mitchell, D.R.G.; Munnangi, A.R.; Barlow, A.J.; Fichtner, M. New insights into the electrochemistry of magnesium molybdate hierarchical architectures for high performance sodium devices. Nanoscale 2018, 10, 13277–13288. [Google Scholar] [CrossRef]
  61. Xu, Y.-D.; Zhang, J.; Wu, Z.-G.; Xu, C.-L.; Li, Y.-C.; Xiang, W.; Wang, Y.; Zhong, Y.-J.; Guo, X.-D.; Chen, H. Stabilizing the Structure of Nickel-Rich Lithiated Oxides via Cr Doping as Cathode with Boosted High-Voltage/Temperature Cycling Performance for Li-Ion Battery. Energy Technol. 2020, 8, 1900498. [Google Scholar] [CrossRef]
  62. Deng, Y.-F.; Zhao, S.-X.; Xu, Y.-H.; Gao, K.; Nan, C.-W. Impact of P-Doped in Spinel LiNi0.5Mn1.5O4 on Degree of Disorder, Grain Morphology, and Electrochemical Performance. Chem. Mater. 2015, 27, 7734–7742. [Google Scholar] [CrossRef]
  63. Sha, O.; Tang, Z.; Wang, S.; Yuan, W.; Qiao, Z.; Xu, Q.; Ma, L. The multi-substituted LiNi0.475Al0.01Cr0.04Mn1.475O3.95F0.05 cathode material with excellent rate capability and cycle life. Electrochim. Acta 2012, 77, 250–255. [Google Scholar] [CrossRef]
  64. Zhong, G.B.; Wang, Y.Y.; Yu, Y.Q.; Chen, C.H. Electrochemical investigations of the LiNi0.45M0.10Mn1.45O4 (M=Fe, Co, Cr) 5V cathode materials for lithium ion batteries. J. Power Sources 2012, 205, 385–393. [Google Scholar] [CrossRef]
  65. Sun, H.Y.; Kong, X.; Wang, B.S.; Luo, T.B.; Liu, G.Y. Cu doped LiNi0.5Mn1.5-xCuxO4 (x = 0, 0.03, 0.05, 0.10, 0.15) with significant improved electrochemical performance prepared by a modified low temperature solution combustion synthesis method. Ceram. Int. 2018, 44, 4603–4610. [Google Scholar] [CrossRef]
  66. Chu, C.-T.; Mondal, A.; Kosova, N.V.; Lin, J.-Y. Improved high-temperature cyclability of AlF3 modified spinel LiNi0.5Mn1.5O4 cathode for lithium-ion batteries. Appl. Surf. Sci. 2020, 530, 147169. [Google Scholar] [CrossRef]
  67. Xiao, J.; Chen, X.; Sushko, P.V.; Sushko, M.L.; Kovarik, L.; Feng, J.; Deng, Z.; Zheng, J.; Graff, G.L.; Nie, Z.; et al. High-Performance LiNi0.5Mn1.5O4 Spinel Controlled by Mn3+ Concentration and Site Disorder. Adv. Mater. 2012, 24, 2109–2116. [Google Scholar] [CrossRef]
  68. Manthiram, A.; Chemelewski, K.; Lee, E.-S. A perspective on the high-voltage LiMn1.5Ni0.5O4 spinel cathode for lithium-ion batteries. Energy Environ. Sci. 2014, 7, 1339–1350. [Google Scholar] [CrossRef]
  69. Kim, J.-H.; Pieczonka, N.P.W.; Li, Z.; Wu, Y.; Harris, S.; Powell, B.R. Understanding the capacity fading mechanism in LiNi0.5Mn1.5O4/graphite Li-ion batteries. Electrochim. Acta 2013, 90, 556–562. [Google Scholar] [CrossRef]
  70. Ando, K.; Matsuda, T.; Imamura, D. Degradation diagnosis of lithium-ion batteries with a LiNi0.5Co0.2Mn0.3O2 and LiMn2O4 blended cathode using dV/dQ curve analysis. J. Power Sources 2018, 390, 278–285. [Google Scholar] [CrossRef]
  71. Li, F.; Ma, J.; Lin, J.; Zhang, X.; Yu, H.; Yang, G. Exploring the origin of electrochemical performance of Cr-doped LiNi0.5Mn1.5O4. Phys. Chem. Chem. Phys. 2020, 22, 3831–3838. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Synthesis process of LiNi0.5-x/2Mgx/2Crx/2Mn1.5-x/2O4 (x = 0, 0.005, 0.01, 0.015) samples.
Scheme 1. Synthesis process of LiNi0.5-x/2Mgx/2Crx/2Mn1.5-x/2O4 (x = 0, 0.005, 0.01, 0.015) samples.
Nanomaterials 15 00429 sch001
Figure 1. Morphological evolution during the preparation of pristine and Mg2+/Cr3+ co-doped LNMO. (ac) Morphologies of the pristine precursor, NMO-0, and LNMO-0; (d) particle morphology of LNMO-0; (eg) morphologies of the precursor, NMO-0.005, and LNMO-0.005 in 0.005 Mg2+/Cr3+ co-doped samples; (h) particle morphology of LNMO-0.005; (ik) morphologies of the precursor, NMO-0.010, and LNMO-0.010 in 0.010 Mg2+/Cr3+ co-doped samples; (l) particle morphology of LNMO-0.010; (mo) morphologies of the precursor, NMO-0.015, and LNMO-0.015 in 0.015 Mg2+/Cr3+ co-doped samples; (p) particle morphology of LNMO-0.015; (qu) EDS mapping of 0.010 Mg2+/Cr3+ co-doped precursor.
Figure 1. Morphological evolution during the preparation of pristine and Mg2+/Cr3+ co-doped LNMO. (ac) Morphologies of the pristine precursor, NMO-0, and LNMO-0; (d) particle morphology of LNMO-0; (eg) morphologies of the precursor, NMO-0.005, and LNMO-0.005 in 0.005 Mg2+/Cr3+ co-doped samples; (h) particle morphology of LNMO-0.005; (ik) morphologies of the precursor, NMO-0.010, and LNMO-0.010 in 0.010 Mg2+/Cr3+ co-doped samples; (l) particle morphology of LNMO-0.010; (mo) morphologies of the precursor, NMO-0.015, and LNMO-0.015 in 0.015 Mg2+/Cr3+ co-doped samples; (p) particle morphology of LNMO-0.015; (qu) EDS mapping of 0.010 Mg2+/Cr3+ co-doped precursor.
Nanomaterials 15 00429 g001
Figure 2. Structural analysis of pristine and Mg2+/Cr3+ co-doped NMO and LNMO, including the Rietveld refinement results for both pristine and Mg2+/Cr3+ co-doped LNMO. (a) XRD patterns of pristine and Mg2+/Cr3+ co-doped NMO; (b) XRD patterns of pristine and Mg2+/Cr3+ co-doped LNMO; (cf) Rietveld refinement results for pristine and Mg2+/Cr3+ co-doped LNMO.
Figure 2. Structural analysis of pristine and Mg2+/Cr3+ co-doped NMO and LNMO, including the Rietveld refinement results for both pristine and Mg2+/Cr3+ co-doped LNMO. (a) XRD patterns of pristine and Mg2+/Cr3+ co-doped NMO; (b) XRD patterns of pristine and Mg2+/Cr3+ co-doped LNMO; (cf) Rietveld refinement results for pristine and Mg2+/Cr3+ co-doped LNMO.
Nanomaterials 15 00429 g002
Figure 3. FTIR and Raman spectra of the pristine and Mg2+/Cr3+ co-doped LNMO. (a) FTIR; (b) Raman spectra.
Figure 3. FTIR and Raman spectra of the pristine and Mg2+/Cr3+ co-doped LNMO. (a) FTIR; (b) Raman spectra.
Nanomaterials 15 00429 g003
Figure 4. XPS spectra of pristine and Mg2+/Cr3+ co-doped LNMO. (ac) Survey spectra, Mn 2p, and Ni 2p XPS spectra of LNMO-0; (df) Mn, Mg, and Cr XPS spectra of LNMO-0.005; (gi) Mn, Mg, and Cr XPS spectra of LNMO-0.010; (jl) Mn, Mg, and Cr XPS spectra of LNMO-0.015.
Figure 4. XPS spectra of pristine and Mg2+/Cr3+ co-doped LNMO. (ac) Survey spectra, Mn 2p, and Ni 2p XPS spectra of LNMO-0; (df) Mn, Mg, and Cr XPS spectra of LNMO-0.005; (gi) Mn, Mg, and Cr XPS spectra of LNMO-0.010; (jl) Mn, Mg, and Cr XPS spectra of LNMO-0.015.
Nanomaterials 15 00429 g004
Figure 5. The electrochemical characterization of pristine and Mg2+/Cr3+ co-doped LNMO. (a) CV curves of LNMO-0 at different scan rates; (b) CV curves of LNMO-0.005 at different scan rates; (c) CV curves of LNMO-0.010 at different scan rates; (d) CV curves of LNMO-0.015 at different scan rates; (e) the linear relationship between the anodic and cathodic peak currents and the square root of the scan rate; (f) magnified CV curves of the four samples at the 4.0V peak under a scan rate of 0.100 mV/s.
Figure 5. The electrochemical characterization of pristine and Mg2+/Cr3+ co-doped LNMO. (a) CV curves of LNMO-0 at different scan rates; (b) CV curves of LNMO-0.005 at different scan rates; (c) CV curves of LNMO-0.010 at different scan rates; (d) CV curves of LNMO-0.015 at different scan rates; (e) the linear relationship between the anodic and cathodic peak currents and the square root of the scan rate; (f) magnified CV curves of the four samples at the 4.0V peak under a scan rate of 0.100 mV/s.
Nanomaterials 15 00429 g005
Figure 6. EIS curves (Nyquist plots) and Rct values of pristine and Mg2+/Cr3+ co-doped LNMO. (a) Nyquist plots before cycling; (b) Rct values before cycling; (c) Nyquist plots after 100 cycles at 0.5C; (d) Rct values after 100 cycles at 0.5C.
Figure 6. EIS curves (Nyquist plots) and Rct values of pristine and Mg2+/Cr3+ co-doped LNMO. (a) Nyquist plots before cycling; (b) Rct values before cycling; (c) Nyquist plots after 100 cycles at 0.5C; (d) Rct values after 100 cycles at 0.5C.
Nanomaterials 15 00429 g006
Figure 7. Electrochemical performance of pristine and Mg2+/Cr3+ co-doped LNMO. (a) Initial charge/discharge curves at 0.5C. (b) Cycling performance and Coulombic efficiency at 0.5C. (c) dQ/dV curves for the first and 100th cycles at 0.5C. (d) Rate performance.
Figure 7. Electrochemical performance of pristine and Mg2+/Cr3+ co-doped LNMO. (a) Initial charge/discharge curves at 0.5C. (b) Cycling performance and Coulombic efficiency at 0.5C. (c) dQ/dV curves for the first and 100th cycles at 0.5C. (d) Rate performance.
Nanomaterials 15 00429 g007
Table 1. Rietveld refinement results for the pristine and Mg2+/Cr3+ co-doped LNMO.
Table 1. Rietveld refinement results for the pristine and Mg2+/Cr3+ co-doped LNMO.
Samplesa (Å)V (Å3)I311/I400 Rwp
LNMO-08.174959546.330.978.10
LNMO-0.0058.173774546.090.959.68
LNMO-0.0108.168736545.090.8710.73
LNMO-0.0158.168580545.050.8613.75
Table 2. Potential differences (ΔV, V) between anodic peaks (φpa, V) and cathodic peaks (φpc, V).
Table 2. Potential differences (ΔV, V) between anodic peaks (φpa, V) and cathodic peaks (φpc, V).
ν (mV∙s−1)LNMO-0LNMO-0.005LNMO-0.010LNMO-0.015
φpaφpcΔVφpaφpcΔVφpaφpcΔVφpaφpcΔV
0.054.8064.6220.1844.8164.6540.1624.7914.6450.1464.8064.6270.179
0.0754.8164.6080.2084.7904.6830.1074.7984.6400.1584.8284.6070.221
0.14.8394.5860.2534.8254.6420.1834.8194.6280.1914.8534.5840.269
0.1254.8524.5950.2574.8334.6300.2034.8204.6210.1994.8674.580.287
0.1504.8684.5770.2914.8384.6210.2174.8344.6090.2254.8864.5690.317
Table 3. The diffusion coefficients of Li+ in all samples.
Table 3. The diffusion coefficients of Li+ in all samples.
SamplesSlopepaSlopepcLi-Deintercalation
DLi (cm2∙s−1)
Li-Intercalation
DLi (cm2∙s−1)
LNMO-01.976 × 10−2−2.308 × 10−27.468 × 10−121.019 × 10−11
LNMO-0.0052.636 × 10−2−2.641 × 10−21.329 × 10−111.333 × 10−11
LNMO-0.0103.180 × 10−2−3.122 × 10−21.934 × 10−111.864 × 10−11
LNMO-0.0152.458 × 10−2−2.956 × 10−21.156 × 10−111.671 × 10−11
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

Ma, D.; Wang, J.; Wang, H.; Qian, G.; Zhou, X.; Pei, Z.; Zheng, K.; Wang, Q.; Lu, J. Mg2+ and Cr3+ Co-Doped LiNi0.5Mn1.5O4 Derived from Ni/Mn Bimetal Oxide as High-Performance Cathode for Lithium-Ion Batteries. Nanomaterials 2025, 15, 429. https://doi.org/10.3390/nano15060429

AMA Style

Ma D, Wang J, Wang H, Qian G, Zhou X, Pei Z, Zheng K, Wang Q, Lu J. Mg2+ and Cr3+ Co-Doped LiNi0.5Mn1.5O4 Derived from Ni/Mn Bimetal Oxide as High-Performance Cathode for Lithium-Ion Batteries. Nanomaterials. 2025; 15(6):429. https://doi.org/10.3390/nano15060429

Chicago/Turabian Style

Ma, Dehua, Jiawei Wang, Haifeng Wang, Guibao Qian, Xingjie Zhou, Zhengqing Pei, Kexin Zheng, Qian Wang, and Ju Lu. 2025. "Mg2+ and Cr3+ Co-Doped LiNi0.5Mn1.5O4 Derived from Ni/Mn Bimetal Oxide as High-Performance Cathode for Lithium-Ion Batteries" Nanomaterials 15, no. 6: 429. https://doi.org/10.3390/nano15060429

APA Style

Ma, D., Wang, J., Wang, H., Qian, G., Zhou, X., Pei, Z., Zheng, K., Wang, Q., & Lu, J. (2025). Mg2+ and Cr3+ Co-Doped LiNi0.5Mn1.5O4 Derived from Ni/Mn Bimetal Oxide as High-Performance Cathode for Lithium-Ion Batteries. Nanomaterials, 15(6), 429. https://doi.org/10.3390/nano15060429

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