Next Article in Journal
Recent Advances in Supercritical CO2 Extraction of Pigments, Lipids and Bioactive Compounds from Microalgae
Next Article in Special Issue
Coronavirus-like Core–Shell-Structured Co@C for Hydrogen Evolution via Hydrolysis of Sodium Borohydride
Previous Article in Journal
Investigating the Conformations of a Family of [M2L3]4+ Helicates Using Single Crystal X-ray Diffraction
Previous Article in Special Issue
Selectivity of Oxygen Evolution Reaction on Carbon Cloth-Supported δ-MnO2 Nanosheets in Electrolysis of Real Seawater
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mnx+ Substitution to Improve Na3V2(PO4)2F3-Based Electrodes for Sodium-Ion Battery Cathode

School of Materials Science and Engineering, Jiangsu Province Engineering Research Center of Intelligent Manufacturing Technology for the New Energy Vehicle Power Battery, Changzhou Key Laboratory of Intelligent Manufacturing and Advanced Technology for Power Battery, Changzhou University, Changzhou 213164, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(3), 1409; https://doi.org/10.3390/molecules28031409
Submission received: 31 December 2022 / Revised: 27 January 2023 / Accepted: 29 January 2023 / Published: 1 February 2023
(This article belongs to the Special Issue Energy-Relevant Advanced Materials)

Abstract

:
Na3V2(PO4)2F3 (NVPF) is an extremely promising sodium storage cathode material for sodium-ion batteries because of its stable structure, wide electrochemical window, and excellent electrochemical properties. Nevertheless, the low ionic and electronic conductivity resulting from the insulated PO43− structure limits its further development. In this work, the different valence states of Mnx+ ions (x = 2, 3, 4) doped NVPF were synthesized by the hydrothermal method. A series of tests and characterizations reveals that the doping of Mn ions (Mn2+, Mn3+, Mn4+) changes the crystal structure and also affects the residual carbon content, which further influences the electrochemical properties of NVPF-based materials. The sodiation/desodiation mechanism was also investigated. Among them, the as-prepared NVPF doped with Mn2+ delivers a high reversible discharge capacity (116.2 mAh g−1 at 0.2 C), and the capacity retention of 67.7% after 400 cycles at 1 C was obtained. Such excellent performance and facile modified methods will provide new design ideas for the development of secondary batteries.

1. Introduction

It is a big challenge to fulfill the rising demand for renewable energy with the current market mainstream lithium-ion batteries (LIBs) due to their high cost and low-capacity retention at high rates. Meanwhile, sodium-ion batteries (SIBs) occupy several advantages over LIBs, for instance, rich resources, wide distribution, and low prices [1,2,3,4,5]. Therefore, the SIBs were considered an extremely promising new energy-storage technology. However, the radius of Na+ (1.02 Å) is larger than that of Li+ (0.76 Å), which requires large sodium-ion transport channels, and inevitably causes volume change after long-term cycling [6]. More importantly, the breakthrough point for the electrochemical performance of the battery depends on the cathode materials. Therefore, considerable efforts have been put into finding suitable cathode materials for SIBs to solve these problems [7].
In recent times, transition metal oxides [8,9,10], polyanionic compounds [11,12], Prussian blue analogues [13,14], and organic compounds [15,16] make up the mainstream of cathode materials for SIBs [17]. Among them, Na3V2(PO4)2F3 (NVPF) and Na3V2(PO4)3 (NVP) with the NASICON type phosphates are considered to be the most favorable cathode materials owing to their three-dimensional open frameworks, which can boost sodium ion transport by delivering large interstitial spaces [18]. NVPF has also been demonstrated to possess a higher average working potential of ~3.9 V, and an ideal theoretical capacity of 128 mAh g−1, which is higher than that of NVP (~3.4 V, 117.6 mAh g−1) [17,19,20].
However, low electronic conductivity and sluggish Na+ diffusion limit the high rate performance and long-term cycling life of the NVPF material [21,22]. To improve the electrochemical performance, several methods are frequently used, heteroatom doping, particle nanosizing, and carbon coating [23,24]. Carbon coating is the most straightforward strategy to construct an external conductive framework to increase the surface electronic conductivity, while too much carbon coating prevents the electrolyte from being exposed to the electrode material entirely. Particle nanosizing can limit the material scale to nanometer, which can shorten the Na+ diffusion path and increase the exposure of electrodes to electrolytes at the same time. However, this will also lead to cracking at the electrode surface or particle agglomeration during cycling.
Heteroatom doping is a practical way to increase Na+ diffusion and structural stability. The common practice is using ions with similar radius size to replace sodium ions, which can increase the interplanar distance and facilitate the Na+ extraction/insertion [25,26]. However, limited by element species, fewer elements can be doped at sodium sites, and vanadium sites taking place is also widely studied (e.g., Mn2+, Zr3+, Fe3+, Tix+). Some works found that transition metal elements such as Mn, Fe, and Ti are often used to modulate the width of the forbidden band. Moreover, transition metal ion doping has almost no effect on the material structure, but restricts the growth of microstructures in specific directions to obtain specific morphologies such as flakes and blocks, which generates more active sites at the edges of the electrode material to boost the intrinsic electronic conductivity [27,28,29].
Zhang et al. investigate the effect of Tix+ (x = 2, 3, 4) doping on the NVPF compound properties. The appropriate amount of Tix+ doping will inhibit the growth of NVPF crystals. At the same time, the particle size of the research object increases with the increase of the valence state, and the carbon content decreases with the increase of the valence state, resulting in the optimal cycle performance and rate performance of NVPF-Ti2+ [30]. Starting from the morphology design, Zhang et al. design a hollow microsphere structure of NVPF materials consisting of thin nanoplates and nanoparticles through a tetraglycol-assisted hydrothermal method, and then chemical vapor deposit is used to deposit a carbon layer on the surface of the designed NVPF to further optimize the electrochemical performance of the cathode materials with an initial reversible capacity of 122.9 mAh g−1 and a high coulombic efficiency of 99.1% even after 500 cycles at 5 C [31]. However, the effect of manganese ions with different valence states (Mnx+ x = 2, 3, 4) doping on the NVPF materials to modify the morphology by hydrothermal method and, accordingly, affect the electrochemical properties, has not been studied.
In this work, we report the synthesis of different valence states of manganese ions (Mn2+, Mn3+, Mn4+) doping in NVPF materials using a one-step hydrothermal method (hereafter referred to as NVM2PF@C, NVM3PF@C, NVM4PF@C). Moreover, the crystal structure, morphology, and carbon connect of the doped samples are also studied. Among them, NVM2PF@C exhibits an optimal initial discharge capacity of 116.2 mAh g−1 at 0.2 C and is maintained with a retention ratio of 67.7% even after 400 cycles at 1 C due to microcuboid morphologies and uniform carbon coating.

2. Results and Discussion

Figure 1a shows the XRD patterns of various NVPF samples. All diffraction peaks well correspond to tetragonal NVPF phases with a group of P42/mnm (JCPDS NO: 89-8485) [24,32]. For these samples, no impurity peaks are detected, indicating that a low dose of Mnx+(x = 2, 3, 4) doping has no effect on the crystal structure of NVPF [33]. Moreover, no carbon-related diffraction peaks are observed, indicating that the presence of carbon may exist in the form of disordered carbon coating. In addition, embedding Mn ions can result in a larger cell volume since the radius of Mn ions (0.97 Å) is larger than that of V ions (0.78 Å). To verify this point, Jade software was used to study the influence of the doping of different valence states of manganese ions on the lattice parameters and cell volume (Table S1). It can be observed that all the samples doped with Mn ions show expansions both in cell parameters and volume, confirming that Mn ions were doped into the crystal lattice of NVPF successfully. Accordingly, the increased cell parameters and volume contribute to an increased Na+ migration rate and Na+ storage performance [34].
As shown in Figure 1b, the crystal structure of Na3V2(PO4)2F3 consists of bioctahedral units [V2O8F3] and tetrahedral units [PO4]. [V2O8F3] bioctahedrons are connected by two [VO4F2] octahedrons via F atoms, and the O atoms are all located on the [PO4] tetrahedrons. The Na(1) is located on a triangular prismatic site surrounded by two F atoms and four O atoms, and the other two extended triangular prismatic sites connected to the square prism of the F atoms are denoted as Na(2) sites [35,36]. Turning to the three sodium ions in NVPF, one pair of ions occupy the Na(1) site completely, while the remaining ion conquers half of the two Na(2) sites to make it a whole ion [37]. The vanadium ions with a smaller ionic radius are partially replaced by the Manganese ions with a larger ionic radius, broadening the transport of sodium ions and improving the transportation efficiency of sodium ions. At the same time, a low dose of Mn ions occupying the V sites shortens the bond length of V-O and P-O and lengthens the Na-O bond, which improves the intrinsic electronic conductivity [18].
The effects of Mnx+(x = 2, 3, 4) doping on morphologies of the NVPF materials can be observed by SEM. Figure 2a depicts the morphology of the pristine NVPF@C. The overall structures are in blocky forms, and many broken particles are attached to the intact NVPF@C. From further high-magnification images in Figure 2b,c, it can be observed that both NVM2PF@C and NVM3PF@C materials show microcuboid morphologies with a length of about 2 μm. Compared with the pristine one, the modified materials are more uniformly dispersed and smaller in size, indicating that the introduction of Mnx+ (x = 2, 3) suppresses the crystal growth during a hydrothermal reaction. While in Figure 2d, the morphology of NVM4PF@C is close to that of the unmodified NVPF@C and even becomes more agglomerated and irregular.
More information about lattice fringes and carbon coating could be obtained by TEM. As shown in Figure 3a, the pristine NVPF@C is in the form of block particles with a smooth surface, and its lattice planes with a spacing of 0.543 nm correspond to the (002) plane in tetragonal NVPF crystal. Simultaneously, the 3.9 nm thickness carbon coating can be observed. Turning to the NVM2PF@C, the Mn2+-doped NVMPF@C presents the same lattice planes (002) with the pristine NVPF@C material, and a thicker carbon layer (4.2 nm) is shown in Figure 3b. Similarly, Figure 3c presents the Mn3+-doped NVMPF@C sample. In particular, it can be seen that two lattice planes exist in the Mn4+-doped NVMPF@C with the spacing of 0.545 nm and 0.279 nm corresponding to the lattice plane (002) and (222), respectively, from Figure 3d. This phenomenon can be attributed to the strong oxide MnO2, which not only deteriorates the growth of NVPF crystals to induce growth in specific directions but also consumes a portion of citric acid leading to a 2.9 nm thickness carbon coating [38].
As shown in Figure 4, Raman spectra were characterized to investigate the carbon coating state. Two distinct peaks can be seen at 1350 cm−1 and 1600 cm−1 [39,40], corresponding to the disordered band (D-band) and graphite band (G-band), respectively. Furthermore, the D-band integral area to G-band integral area ratio (AD/AG) is used to describe the degree of graphitization. Figure 4 also illustrates the AD/AG of the as-prepared NVMx+PF@C (x = 0, 2, 3, 4) samples, and the NVM2+PF@C possesses the highest AD/AG value of 2.81, indicating that Mn2+-doped sample contains more electronic conductive disordered carbon [41].
The valence states of the series samples were examined using X-ray photoelectron spectroscopy (XPS) (Figure S1). From Figure S1a, the full XPS spectra show the characteristic peaks corresponding to the Na, V, P, O, and F elements, which verifies the NVPF material has been successfully synthesized. The high-resolution XPS spectra were used to analyze the Mn element of pristine and Mnx+ (x = 2, 3, 4) doped ones. Figure S1b,c show the F1s spectra of NVPF@C and NVM2PF@C. Compared with the pristine one, the divalent manganese ions doped sample possesses an extra Mn-F peak, which confirms the manganese ions were successfully doped into the pristine one [42]. Peaks at 653.8 eV and 645.9 eV in Figure S1d are associated with Mn2+ 2p1/2 and Mn2+ 2p3/2, respectively, both of which are characteristic peaks of Mn2+ 2p and only the peaks of Mn2+ can be seen in the NVM2PF@C sample [43]. Figure S1e shows the XPS spectrum of NVM3PF@C, excluding two characteristic peaks of Mn2+ 2p, the peak of Mn3+ locates at 641.8 eV [44]. In addition to the distinctive peaks of Mn2+ and Mn3+, Figure S1f also shows a typical peak caused by Mn4+ at 649.6 eV. The reason for this phenomenon is that citric acid has a reducing effect, and excess added citric acid can reduce Mn ions from a high valance state to a low valance state. Consequently, all of the valence states of manganese ions can be found in the NVM4PF@C sample. Similarly, both trivalent and divalent manganese ions are present in the NVM3PF@C sample. For NVM2PF@C, only the divalent valence state can be discovered, because the divalent is the lowest valence state of the manganese ion. Therefore, the XPS energy spectra of the manganese ion present multivalent states.
The carbon content of NVPF@C and NVMPF@C was evaluated by thermogravimetric analysis (TGA). As shown in Figure S2, the samples doped with different manganese ions possess different carbon content, and the value of NVPF@C, NVM2PF@C, NVM3PF@C and NVM4PF@C is 4.77%, 5.09%, 4.62%, and 4.23%, respectively (Table S2). At the same time, Figure S2 corroborates the presence of amorphous carbon coating and illustrates the slight difference in carbon content coincides with the thickness of carbon coating observed by TEM characterization. The XPS spectra also show that Mn4+ consumes the greatest amount of citric acid of all the samples, making the NVM4PF@C sample have the lowest carbon content (4.23%). The citric acid in the pristine and NVM2PF@C samples was used to reduce V5+ to V3+ and acted as a carbon source, so these two samples have a high carbon content of 4.77% and 5.09%, respectively. Turing to NVM3PF@C, whose carbon content (4.62%) is between that of NVM2PF@C and NVM4PF@C samples.
Figure 5a shows the cyclic voltammetry (CV) curves of Mn-doped NVPF@C and pristine NVPF@C materials measured in the voltage range of 2.5 to 4.3 V (vs. Na+/Na) at 0.1 mV s−1. The redox peak positions for the NVM2PF@C electrode are at 3.596/3.758 V and 3.867/4.102 V, which is assigned to the V3+/V4+ redox couple [40], as can be observed from the CV curves. NVM2PF@C has a cathodic potential difference of 162 mV and an anodic potential difference of 235 mV, which are 59 mV and 10 mV, respectively, less than the pristine NVPF@C electrode. Additional details regarding the cathodic and anodic potential differences between various samples were calculated, which is shown in Table S3. By comparison, the NVM2PF@C electrode has the fewest cathodic and anodic potential differences, indicating a decreased polarization and enhanced Na+ diffusion kinetics [45,46,47].
Figure 5b shows the initial charge/discharge curves of NVPF@C, NVM2PF@C, NVM3PF@C, and NVM4PF@C at 0.2 C. The pristine electrodes possess two obvious charge/discharge plateaus, 3.57/3.79 V and 3.83/4.07 V. While, the charge/discharge plateaus of NVM2PF@C corresponding to 3.60/3.76 V and 3.87/4.10V, which has a smaller polarization compared with the NVPF@C. The result is consistent with the obtained data from the CV plots and table. In addition, the divalent manganese ion doping electrode has the best initial discharge capacity of all the electrodes, as shown by the longer charge/discharge plateau of NVM2PF@C than other samples, which is 116.2 mAh g−1 at 0.2 C, close to the theoretical value (128 mAh g−1). While, the values of the other three ones are 106.8 mAh g−1, 108.1 mAh g−1, and 94.5 mAh g−1 corresponding to NVPF@C, NVM3PF@C, and NVM4PF@C, respectively. Furthermore, the increased initial specific capacity is benefiting from smaller particles, which can broaden the region in contact between the electrode and electrolyte, thereby enhancing sodium ion transport [48,49].
The cycling performance of the as-prepared objects is depicted in Figure 5c (NVPF@C, NVM2PF@C, NVM3PF@C, NVM4PF@C). The NVM2PF@C electrode shows the highest initial discharge capacity of 116.2 mAh g−1 and remains 93.2 mAh g−1 at 0.2 C after 100 cycles. The capacity retention is 80.2%. While, the NVPF@C, NVM3PF@C, and NVM4PF@C electrodes display the initial discharge capacity of 106.8 mAh g−1, 108.1 mAh g−1, and 94.5 mAh g−1 after cycling, corresponding to 77.8%, 79.3%, and 73.9% capacity retention, respectively (Table S4). Benefiting from the small particles with micro-sized distribution, the NVM2PF@C and NVM3PF@C samples possess a favorable cycling performance. The manganese ion doped samples induce zero-dimensional defects that can lead to more active sites, larger sodium-ion transport channels, and a more stable crystal structure [50]. However, the tetravalent manganese ions doped sample shows an inferior result because of low carbon content and agglomerated and broken particles, deteriorating the electrochemical performance in contrast with the NVPF@C material.
Figure 5d displays the rate performance of the four electrodes. At the current rate of 0.2, 0.5, 1, 2, and 5 C, the capacity ranking presented by rate curves among the four samples is in line with the cycling performance at 0.2 C (from first to fourth is NVM2PF@C, NVM3PF@C, NVPF@C, and NVM4PF@C). The data reveal that both the NVPF@C materials doped with divalent and trivalent manganese ions present high rate performance at different current rates. The reason behind this phenomenon could be related to the enhanced structural stability carried on by the partial replacement of manganese ions for vanadium sites. In particular, the NVM2PF@C electrode possesses the optimal capacity compared with the other three samples. Classical chemical structure theory states that structure determines performance, during a high current rate, the structural stability is expressed as cycling stability. When the rate was shifted from 5 C to 0.2 C, the capacities of NVM2PF@C recovered to 99.3 mAh g−1, revealing an outstanding electrochemical reversibility. The long-life cycling performance of the NVM2PF@C electrode is provided in Figure 5e. The discharge capacity reaches 83.9 mAh g−1 and a retention ratio of 67.7% after 400 cycles at 1 C, indicating that the divalent manganese ions replacing part of the vanadium ions can deliver better long-term cycling life.
The XRD pattern and the SEM images were used to study the phase analysis and morphology changes of the NVM2PF@C cathode before and after 400 cycles at 1 C. The diffraction peak positions of the NVM2PF@C all match well with the Aman space group (JCPDS NO: 89-8485), and no impurity peaks can be observed in Figure S3. It can be concluded that the NVM2PF@C electrode is still maintained in the NVPF phase, which proves that divalent manganese ions doping is beneficial to structural stability and cycling stability [51]. However, the diffraction peak positions of the NVM2PF@C sample after cycling shift to lower angles, indicating that the cell parameters and the lattice spacing get increased. This probably results from the sodium-ion transport channels that have been further expanded after long cycling, which makes the crystal structure larger. Simultaneously, the relative intensity of the diffraction peaks after cycling is weaker than that before cycling, demonstrating a slightly decreased crystallinity. This phenomenon is mainly associated with the deteriorated crystal structure, and the other possible reason is the coating of mixed slurry (conductive agent and binder). The NVM2PF@C sample exhibits a uniform cubic morphology when comparing the morphology before and after cycling, despite the fact that a lot of broken particles have adhered to the surface of the long-term cycling electrode. This outcome is consistent with the favorable reversible cycling performance and the electrochemical performance analysis discussed above.
Electrochemical impedance spectroscopy (EIS) was used to future investigate the four samples’ electrochemical behavior and kinetic reaction process. The Nyquist plots, as illustrated in Figure 6a, consist of a semicircle at high frequency and a sloping line at low frequency. The semicircle at high frequency is associated with the diffusive migration of sodium ions via the solid electrolyte interface (SEI), and the sloping line at low frequency is connected with the solid diffusion process of sodium ions inside the active material particles. In the inset figure, the equivalent circuit is shown with Re, Rct, Rf, CPE1, CPE2, and Zw, which stand for, electrolyte resistance, interfacial charge transfer resistance, film resistance of Na+ diffusion, capacitance of the passivation film, double layer and Warburg impedance, respectively [52]. The Rct of the NVM2PF@C electrode is 82.5 Ω (Table S5), which is much lower than that of the other three (NVPF@C 110.0 Ω NVM3PF@C 91.7 Ω NVM4PF@C 136.9 Ω), implying the Mn2+-doped NVPF@C possesses the enhanced Na+ transfer kinetic and higher electronic conductivity. The smaller the radius of the semicircle, the less the sloping line’s angle of inclination is less than 45°, which reflects the roughness of the electrode surface. It seems that Mn-doped samples’ particle surface roughness and specific surface area are increased.
The coefficient diffusion of sodium ion (DNa+) could be obtained in view of Equation (1) [30]:
D = R 2 T 2 2 A 2 n 4 F 4 C 2 σ 2  
The following Equation (2) can be used to compute the Warburg factor (σ) linked to Z’. Where R is the gas constant, T is the absolute temperature, A is the surface area, n is the number of electrons transferred per molecule, F is the Faraday constant, and C is the concentration of Na+
Z re = R e + R ct + σ ω - 1 / 2
where stands for the angle frequency. The Warburg factor can be observed in Figure 6b, which reveals the relationship between the linearly fitting Zre and ω−1/2 in the low frequency, and the calculated results for the series of NVMx+PF@C are displayed in Table S5. The NVM2PF@C electrode possesses the highest sodium-ion diffusion coefficient reaching 1.03 × 10−13 cm2 s−1. While the values of the others are much lower than that of the NVM2PF@C. The uniform and porous grain morphology after a proper Mn2+ concentration doping effectively reduces interfacial resistance and improves the Na+ diffusion kinetics. At the same time, a low dose of the divalent manganese ion doping helps to stabilize the Na+ diffusion channel and optimize the NASICON structure [37]. Furthermore, the uniform carbon coating plays a very important part in the charge/discharge process, which significantly increases the intrinsic electronic conductivity. Therefore, the NVM2PF@C and NVM3PF@C electrodes have improvements in both reversible capacity and cycling performance compared with the pristine NVPF@C electrode.

3. Experimental

The NVPF@C was prepared by a hydrothermal method. The components are NaF (Macklin 99.5%), NH4H2PO4 (Macklin 99.5%), NH4VO3 (Macklin, 99.5%), C6H8O7 (Macklin, Ar), and manganese source including (CH3COO)2Mn, (CH3COO)3Mn•2H2O and MnO2(Macklin, Ar) in the molar ratio of 3:2:1.95:1.5:0.05. Firstly, NH4VO3 and C6H8O7 were dissolved in an appropriate amount of deionized water, stirring at 70 °C to obtain a transparent solution. Then, NH4H2PO4, NaF, and manganese sources were put into the above solution. The color of the solution changed from faint yellow to brick red after 30 min of stirring. Next, a 100 mL Teflon-lined autoclave was filled with the previously prepared solution and heated at 200 °C for 12 h. After that, until the as-prepared samples returned to 25 °C, the dark green product was obtained then centrifuged and washed to acquire the precursor. The above precursor was put into a tube furnace for calcination in Ar/H2. The first step was preheated at 300 °C for 4 h, and the second step was sintered to 600 °C for 8 h. Then, the black powder product Na3V1.95Mn0.05(PO4)2F3@C (referred to as NVMPF@C) was obtained. According to the different valence states of manganese sources (Mn2+, Mn3+, Mn4+), a series of NVMPF@C materials (NVM2PF@C, NVM3PF@C, NVM4PF@C) can be obtained.
Phase composite and crystal structure of all NVMx+PF@C (x = 0, 2, 3, 4) samples were performed via X-ray diffraction (XRD). Jade software was used to study the lattice parameters. X-ray photoelectron spectroscopy (XPS) analysis was identified by a Thermo VG Scientific ESCALab 250. The Thermogravimetric (TG) data of the series of NVPF@C samples were tested by Labsys Evo/seteram thermal analyzer. Raman spectra were obtained by using a Raman spectrometer (LabRAM-HR) with an excitation wavelength of 532 nm. High Resolution field transmission electron microscopy (FETEM, a JEOL 2100) and field emission scanning electron microscopy (FESEM, SUPRA55) were employed to explore the surface morphology of the NVMx+PF@C and electrodes.
The active substance, conductive agent (Super p), and binder (PVDF) were weighed at a mass ratio of 7:2:1 and thoroughly ground in a mortar. An appropriate amount of NMP solvent was added drop by drop and mixed evenly. The above slurry was coated on aluminum foil and put into a vacuum oven to evaporate solvent and water. After 12 h, the as-prepared electrode was cut into disks. In our experiment, the average mass loading of the active materials was 1.2–1.5 mg cm−2 on a 12 mm disk. The batteries were assembled in the argon-filled glove box (BRAUN). A proper amount of electrolyte (1 M NaClO4 in EC:DMC = 1:1 with 5%FEC) was added to obtain sodium ion half-cell coin cells (CR2032 type). Electrochemical performance measurements were evaluated by a battery test system (LAND), and the potential range was set from 2.5 V to 4.3 V. Cyclic voltammogram (CV, 2.5–4.3 V) was tested by an electrochemical workstation (CHI 600A) with a sweep rate of 0.1 mV s−1. Electrochemical impedance spectroscopy (EIS) was investigated by an electrochemical workstation (CHI 600A), and the frequency ranged from 0.1 MHz to 105 MHz.

4. Conclusions

In summary, by using a single hydrothermal method, manganese ions with different valence states doped NVPF@C were successfully synthesized. The manganese with different valence states was used to investigate the effects of NVPF, such as morphologies, the thickness of carbon coating layer, and electrochemical performance. It was found that the Mn2+ can effectively guide the formation of microcuboid NVPF and restrict the size growth of NVPF crystal. In addition, the Mn2+ doping strategy also improved the electrochemical performance by lower Rct and larger Na+ diffusion kinetics. Thus, the NVM2PF@C sample performed the best cycling and rate performance with the initial discharge capacity of 116.2 mAh g−1 and the capacity retention (67.7% after 400 cycles at 1 C). The improved electrochemical performance is attributed to the uniform carbon coating and stable crystal structure, indicating that divalent manganese ions doping NVPF@C can be a promising cathode for SIBs.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28031409/s1, Figure S1: Full XPS spectrum of NVPF@C; high-resolution XPS spectra of NVPF@C and NVMPF@C; Figure S2: The TG curves of NVPF@C and NVMPF@C; Figure S3: XRD and SEM images of NVM2PF@C before and after the cycles; Table S1: Cell parameters of NVPF@C and NVMPF@C; Table S2: Carbon content of NVPF@C and NVMPF@C; Table S3: Potential difference of NVPF@C and NVMPF@C; Table S4: Long-term cycling performance of NVPF@C and NVMPF@C; Table S5: The impedance parameters from the simulated equivalent circuit and sodium diffusion coefficients of NVPF@C and NVMPF@C.

Author Contributions

Conceptualization, R.S., P.W. and Y.R.; Methodology, R.S.; Software, R.S. and W.Z.; Validation, R.S.; Formal analysis, R.S., K.L. and W.Z.; Investigation, R.S., K.L., W.Z. and J.L.; Data curation, R.S. and W.Z.; Writing—original draft, R.S.; Writing—review & editing, Y.R., K.L. and P.W.; Supervision, Y.R., K.L., P.W., J.L. and W.L.; Project administration, J.L. and Y.R.; Funding acquisition, W.L. and Y.R. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (No. 91961126, U22A20420), the Science and Technology Plan Project of Chang-zhou (No. CJ20220148). In addition, the authors thank Jiangsu Development & Reform Commission for their support.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Armand, M.; Tarascon, J.M. Building better batteries. Nature 2008, 451, 652–657. [Google Scholar] [CrossRef] [PubMed]
  2. Larcher, D.; Tarascon, J.M. Towards greener and more sustainable batteries for electrical energy storage. Nat. Chem. 2015, 7, 19–29. [Google Scholar] [CrossRef] [PubMed]
  3. Yang, Z.G.; Zhang, J.L.; Kintner-Meyer, M.C.W.; Lu, X.C.; Choi, D.W.; Lemmon, J.P.; Liu, J. Electrochemical Energy Storage for Green Grid. Chem. Rev. 2011, 111, 3577–3613. [Google Scholar] [CrossRef] [PubMed]
  4. Pan, H.L.; Hu, Y.S.; Chen, L.Q. Room-temperature stationary sodium-ion batteries for large-scale electric energy storage. Energy Environ. Sci. 2013, 6, 2338–2360. [Google Scholar] [CrossRef]
  5. Fang, C.; Huang, Y.H.; Zhang, W.X.; Han, J.T.; Deng, Z.; Cao, Y.L.; Yang, H.X. Routes to High Energy Cathodes of Sodium-Ion Batteries. Adv. Energy Mater. 2016, 6, 1501727. [Google Scholar] [CrossRef]
  6. Zhao, H.-S.; Qi, Y.-L.; Liang, K.; Zhu, W.-K.; Wu, H.-B.; Li, J.-B.; Ren, Y.-R. Phosphorus-doping and oxygen vacancy endowing anatase TiO2 with excellent sodium storage performance. Rare Met. 2021, 41, 1284–1293. [Google Scholar] [CrossRef]
  7. Zhao, H.; Qi, Y.; Liang, K.; Li, J.; Zhou, L.; Chen, J.; Huang, X.; Ren, Y. Interface-Driven Pseudocapacitance Endowing Sandwiched CoSe2/N-Doped Carbon/TiO2 Microcubes with Ultra-Stable Sodium Storage and Long-Term Cycling Stability. ACS Appl. Mater. Interfaces 2021, 13, 61555–61564. [Google Scholar] [CrossRef]
  8. Xiao, Y.; Zhu, Y.F.; Yao, H.R.; Wang, P.F.; Zhang, X.D.; Li, H.L.; Yang, X.N.; Gu, L.; Li, Y.C.; Wang, T.; et al. A Stable Layered Oxide Cathode Material for High-Performance Sodium-Ion Battery. Adv. Energy Mater. 2019, 9, 1803978. [Google Scholar] [CrossRef]
  9. Guo, S.H.; Yu, H.J.; Jian, Z.L.; Liu, P.; Zhu, Y.B.; Guo, X.W.; Chen, M.W.; Ishida, M.; Zhou, H.S. A High-Capacity, Low-Cost Layered Sodium Manganese Oxide Material as Cathode for Sodium-Ion Batteries. Chemsuschem 2014, 7, 2115–2119. [Google Scholar] [CrossRef]
  10. Han, M.H.; Gonzalo, E.; Singh, G.; Rojo, T. A comprehensive review of sodium layered oxides: Powerful cathodes for Na-ion batteries. Energ Environ. Sci. 2015, 8, 81–102. [Google Scholar] [CrossRef]
  11. Ni, Q.; Bai, Y.; Wu, F.; Wu, C. Polyanion-Type Electrode Materials for Sodium-Ion Batteries. Adv. Sci. 2017, 4, 1600275. [Google Scholar] [CrossRef]
  12. Chen, M.; Chen, L.; Hu, Z.; Liu, Q.; Zhang, B.; Hu, Y.; Gu, Q.; Wang, J.L.; Wang, L.Z.; Guo, X.; et al. Carbon-Coated Na3.32 Fe2.34 (P2O7)2 Cathode Material for High-Rate and Long-Life Sodium-Ion Batteries. Adv. Mater. 2017, 29, 1605535. [Google Scholar] [CrossRef]
  13. Tang, X.; Liu, H.; Su, D.; Notten, P.H.L.; Wang, G. Hierarchical sodium-rich Prussian blue hollow nanospheres as high-performance cathode for sodium-ion batteries. Nano Res. 2018, 11, 3979–3990. [Google Scholar] [CrossRef]
  14. Nie, P.; Yuan, J.; Wang, J.; Le, Z.; Xu, G.; Hao, L.; Pang, G.; Wu, Y.; Dou, H.; Yan, X.; et al. Prussian Blue Analogue with Fast Kinetics Through Electronic Coupling for Sodium Ion Batteries. ACS Appl. Mater. Interfaces 2017, 9, 20306–20312. [Google Scholar] [CrossRef]
  15. Knight, J.C.; Therese, S.; Manthiram, A. Chemical extraction of Zn from ZnMn2O4-based spinels. J. Mater. Chem. A 2015, 3, 21077–21082. [Google Scholar] [CrossRef]
  16. Luo, W.; Allen, M.; Raju, V.; Ji, X. An Organic Pigment as a High-Performance Cathode for Sodium-Ion Batteries. Adv. Energy Mater. 2014, 4, 1400554. [Google Scholar] [CrossRef]
  17. Jiang, M.; Xu, D.; Yang, B.; Zhang, C.; Cao, M. Oxygen Vacancy Engineering in Na3V2(PO4)3 for Boosting Sodium Storage Kinetics. Adv. Mater. Interfaces 2021, 8, 2100188. [Google Scholar] [CrossRef]
  18. Puspitasari, D.A.; Patra, J.; Hung, I.M.; Bresser, D.; Lee, T.-C.; Chang, J.-K. Optimizing the Mg Doping Concentration of Na3V2–xMgx(PO4)2F3/C for Enhanced Sodiation/Desodiation Properties. ACS Sustain. Chem. Eng. 2021, 9, 6962–6971. [Google Scholar] [CrossRef]
  19. Broux, T.; Fauth, F.; Hall, N.; Chatillon, Y.; Bianchini, M.; Bamine, T.; Leriche, J.B.; Suard, E.; Carlier, D.; Reynier, Y.; et al. High Rate Performance for Carbon-Coated Na3V2(PO4)2F3 in Na-Ion Batteries. Small Methods 2019, 3, 1800215. [Google Scholar] [CrossRef]
  20. Deng, L.; Yu, F.-D.; Xia, Y.; Jiang, Y.-S.; Sui, X.-L.; Zhao, L.; Meng, X.-H.; Que, L.-F.; Wang, Z.-B. Stabilizing fluorine to achieve high-voltage and ultra-stable Na3V2(PO4)2F3 cathode for sodium ion batteries. Nano Energy 2021, 82, 105659. [Google Scholar] [CrossRef]
  21. Park, Y.U.; Seo, D.H.; Kwon, H.S.; Kim, B.; Kim, J.; Kim, H.; Kim, I.; Yoo, H.I.; Kang, K. A New High-Energy Cathode for a Na-Ion Battery with Ultrahigh Stability. J. Am. Chem. Soc. 2013, 135, 13870–13878. [Google Scholar] [CrossRef] [PubMed]
  22. Duan, W.; Zhu, Z.; Li, H.; Hu, Z.; Zhang, K.; Cheng, F.; Chen, J. Na3V2(PO4)3@C core-shell nanocomposites for rechargeable sodium-ion batteries. J. Mater. Chem. A 2014, 2, 8668–8675. [Google Scholar] [CrossRef]
  23. Qi, Y.; Mu, L.; Zhao, J.; Hu, Y.-S.; Liu, H.; Dai, S. pH-regulative synthesis of Na3(VPO4)2F3 nanoflowers and their improved Na cycling stability. J. Mater. Chem. A 2016, 4, 7178–7184. [Google Scholar] [CrossRef]
  24. Cai, Y.; Cao, X.; Luo, Z.; Fang, G.; Liu, F.; Zhou, J.; Pan, A.; Liang, S. Caging Na3V2(PO4)2F3 Microcubes in Cross-Linked Graphene Enabling Ultrafast Sodium Storage and Long-Term Cycling. Adv. Sci. 2018, 5, 1800680. [Google Scholar] [CrossRef]
  25. Li, Y.; Chen, M.; Liu, B.; Zhang, Y.; Liang, X.; Xia, X. Heteroatom Doping: An Effective Way to Boost Sodium Ion Storage. Adv. Energy Mater. 2020, 10, 2000927. [Google Scholar] [CrossRef]
  26. Zhu, L.; Wang, H.; Sun, D.; Tang, Y.; Wang, H. A comprehensive review on the fabrication, modification and applications of Na3V2(PO4)2F3 cathodes. J. Mater. Chem. A 2020, 8, 21387–21407. [Google Scholar] [CrossRef]
  27. Xin, M. Effect of Cu and Mn doping on optical properties of ZnS:Cu, Mn nanoparticles synthesized by the hydrothermal technique. J. Sulfur Chem. 2021, 43, 206–214. [Google Scholar] [CrossRef]
  28. Bhat, A.A.; Tomar, R. Mn and Ce doping in hydrothermally derived CaSnO3 perovskite nanostructure. A facile way to enhance optical, magnetic and electrochemical properties. J. Alloys Compd. 2021, 876, 160043. [Google Scholar] [CrossRef]
  29. Vanetsev, A.; Põdder, P.; Oja, M.; Khaidukov, N.M.; Makhov, V.N.; Nagirnyi, V.; Romet, I.; Vielhauer, S.; Mändar, H.; Kirm, M. Microwave-hydrothermal synthesis and investigation of Mn-doped K2SiF6 microsize powder as a red phosphor for warm white LEDs. J. Lumin. 2021, 239, 118389. [Google Scholar] [CrossRef]
  30. Yi, H.M.; Ling, M.X.; Xu, W.B.; Li, X.F.; Zheng, Q.; Zhang, H.M. VSC-doping and VSU-doping of Na3V2-xTix(PO4)2F3 compounds for sodium ion battery cathodes: Analysis of electrochemical performance and kinetic properties. Nano Energy 2018, 47, 340–352. [Google Scholar] [CrossRef]
  31. Zhang, Y.; Guo, S.R.; Xu, H.Y. Synthesis of uniform hierarchical Na3V1.95Mn0.05(PO4)2F3@C hollow microspheres as a cathode material for sodium-ion batteries. J. Mater. Chem. A 2018, 6, 4525–4534. [Google Scholar] [CrossRef]
  32. Liang, K.; Wang, S.; Zhao, H.; Huang, X.; Ren, Y.; He, Z.; Mao, J.; Zheng, J. A facile strategy for developing uniform hierarchical Na3V2(PO4)2F3@carbonized polyacrylonitrile multi-clustered hollow microspheres for high-energy-density sodium-ion batteries. Chem. Eng. J. 2022, 428, 131780. [Google Scholar] [CrossRef]
  33. Fuentes, R.O.; Figueiredo, F.M.; Marques, F.M.B.; Franco, J.I. Influence of microstructure on the electrical properties of NASICON materials. Solid State Ion. 2001, 140, 173–179. [Google Scholar] [CrossRef]
  34. Barker, J.; Saidi, M.Y.; Swoyer, J.L. A comparative investigation of the Li insertion properties of the novel fluorophosphate phases, NaVPO4F and LiVPO4F. J. Electrochem. Soc. 2004, 151, A1670–A1677. [Google Scholar] [CrossRef]
  35. Song, W.X.; Ji, X.B.; Chen, J.; Wu, Z.P.; Zhu, Y.R.; Ye, K.F.; Hou, H.S.; Jing, M.J.; Banks, C.E. Mechanistic investigation of ion migration in Na3V2(PO4)2F3 hybrid-ion batteries. Phys. Chem. Chem. Phys. 2015, 17, 159–165. [Google Scholar] [CrossRef]
  36. Song, W.X.; Cao, X.Y.; Wu, Z.P.; Chen, J.; Zhu, Y.R.; Hou, H.S.; Lan, Q.; Ji, X.B. Investigation of the Sodium Ion Pathway and Cathode Behavior in Na3V2(PO4)2F3 Combined via a First Principles Calculation. Langmuir 2014, 30, 12438–12446. [Google Scholar] [CrossRef]
  37. Song, W.X.; Ji, X.B.; Wu, Z.P.; Yang, Y.C.; Zhou, Z.; Li, F.Q.; Chen, Q.Y.; Banks, C.E. Exploration of ion migration mechanism and diffusion capability for Na3V2(PO4)2F3 cathode utilized in rechargeable sodium-ion batteries. J. Power Sources 2014, 256, 258–263. [Google Scholar] [CrossRef]
  38. Toupin, M.; Brousse, T.; Belanger, D. Charge storage mechanism of MnO2 electrode used in aqueous electrochemical capacitor. Chem. Mater. 2004, 16, 3184–3190. [Google Scholar] [CrossRef]
  39. Zhang, L.L.; Ma, D.; Li, T.; Liu, J.; Ding, X.K.; Huang, Y.H.; Yang, X.L. Polydopamine-Derived Nitrogen-Doped Carbon-Covered Na3V2(PO4)2F3 Cathode Material for High-Performance Na-Ion Batteries. ACS Appl. Mater. Interfaces 2018, 10, 36851–36859. [Google Scholar] [CrossRef]
  40. Shen, C.; Long, H.; Wang, G.C.; Lu, W.; Shao, L.; Xie, K.Y. Na3V2(PO4)2F3@C dispersed within carbon nanotube frameworks as a high tap density cathode for high-performance sodium-ion batteries. J. Mater. Chem. A 2018, 6, 6007–6014. [Google Scholar] [CrossRef]
  41. Li, Y.; Liang, X.; Chen, G.; Zhong, W.; Deng, Q.; Zheng, F.; Yang, C.; Liu, M.; Hu, J. In-situ constructing Na3V2(PO4)2F3/carbon nanocubes for fast ion diffusion with high-performance Na+-storage. Chem. Eng. J. 2020, 387, 123952. [Google Scholar] [CrossRef]
  42. Gu, Z.Y.; Guo, J.Z.; Sun, Z.H.; Zhao, X.X.; Wang, X.T.; Liang, H.J.; Zhao, B.; Li, W.H.; Pan, X.M.; Wu, X.L. Aliovalent-Ion-Induced Lattice Regulation Based on Charge Balance Theory: Advanced Fluorophosphate Cathode for Sodium-Ion Full Batteries. Small 2021, 17, 2102010. [Google Scholar] [CrossRef] [PubMed]
  43. Kaliyappan, K.; Liu, J.; Xiao, B.W.; Lushington, A.; Li, R.Y.; Sham, T.K.; Sun, X.L. Enhanced Performance of P2-Na0.66(Mn0.54Co0.13Ni0.13)O2 Cathode for Sodium-Ion Batteries by Ultrathin Metal Oxide Coatings via Atomic Layer Deposition. Adv. Funct. Mater. 2017, 27, 1701870. [Google Scholar] [CrossRef]
  44. Shen, W.; Li, H.; Guo, Z.Y.; Li, Z.H.; Xu, Q.J.; Liu, H.M.; Wang, Y.G. Improvement on the high-rate performance of Mn-doped Na3V2(PO4)3/C as a cathode material for sodium ion batteries. Rsc. Adv. 2016, 6, 71581–71588. [Google Scholar] [CrossRef]
  45. Zhu, C.B.; Song, K.P.; van Aken, P.A.; Maier, J.; Yu, Y. Carbon-Coated Na3V2(PO4)3 Embedded in Porous Carbon Matrix: An Ultrafast Na-Storage Cathode with the Potential of Outperforming Li Cathodes. Nano Lett. 2014, 14, 2175–2180. [Google Scholar] [CrossRef]
  46. Aragon, M.J.; Lavela, P.; Ortiz, G.F.; Tirado, J.L. Benefits of Chromium Substitution in Na3V2(PO4)3 as a Potential Candidate for Sodium-Ion Batteries. Chemelectrochem 2015, 2, 995–1002. [Google Scholar] [CrossRef]
  47. Shin, J.; Yang, J.; Sergey, C.; Song, M.S.; Kang, Y.M. Carbon Nanofibers Heavy Laden with Li3V2(PO4)3 Particles Featuring Superb Kinetics for High-Power Lithium Ion Battery. Adv. Sci. 2017, 4, 1700128. [Google Scholar] [CrossRef]
  48. Zheng, S.Y.; Zhong, G.M.; McDonald, M.J.; Gong, Z.L.; Liu, R.; Wen, W.; Yang, C.; Yang, Y. Exploring the working mechanism of Li+ in O3-type NaLi0.1Ni0.35Mn0.55O2 cathode materials for rechargeable Na-ion batteries. J. Mater. Chem. A 2016, 4, 9054–9062. [Google Scholar] [CrossRef]
  49. Yi, T.f.; Shi, L.; Han, X.; Wang, F.; Zhu, Y.; Xie, Y. Approaching High-Performance Lithium Storage Materials by Constructing Hierarchical CoNiO2@CeO2 Nanosheets. Energy Enviro. Mater. 2020, 4, 586–595. [Google Scholar] [CrossRef]
  50. Yi, T.f.; Qiu, L.Y.; Mei, J.; Qi, S.Y.; Cui, P.; Luo, S.; Zhu, Y.R.; Xie, Y.; He, Y.B. Porous spherical NiO@NiMoO4@PPy nanoarchitectures as advanced electrochemical pseudocapacitor materials. Sci. Bull. 2020, 65, 546–556. [Google Scholar] [CrossRef] [Green Version]
  51. Gao, H.C.; Li, Y.T.; Park, K.; Goodenough, J.B. Sodium Extraction from NASICON-Structured Na3MnTi(PO4)3 through Mn(III)/Mn(II) and Mn(IV)/Mn(III) Redox Couples. Chem. Mater. 2016, 28, 6553–6559. [Google Scholar] [CrossRef]
  52. Wu, L.Y.; Dong, S.Y.; Pang, G.; Li, H.S.; Xu, C.Y.; Zhang, Y.D.; Dou, H.; Zhang, X.G. Rocking-chair Na-ion hybrid capacitor: A high energy/power system based on Na3V2O2(PO4)2F@PEDOT core-shell nanorods. J. Mater. Chem. A 2019, 7, 1030–1037. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns of NVMPF@C; (b) crystal structure of NVMPF.
Figure 1. (a) XRD patterns of NVMPF@C; (b) crystal structure of NVMPF.
Molecules 28 01409 g001
Figure 2. SEM patterns of (a) NVPF@C; (b) NVM2PF@C; (c) NVM3PF@C; (d) NVM4PF@C.
Figure 2. SEM patterns of (a) NVPF@C; (b) NVM2PF@C; (c) NVM3PF@C; (d) NVM4PF@C.
Molecules 28 01409 g002
Figure 3. TEM images of (a) NVPF@C; (b) NVM2PF@C; (c) NVM3PF@C; (d) NVM4PF@C.
Figure 3. TEM images of (a) NVPF@C; (b) NVM2PF@C; (c) NVM3PF@C; (d) NVM4PF@C.
Molecules 28 01409 g003
Figure 4. Raman images of (a) NVPF@C; (b) NVM2PF@C; (c) NVM3PF@C; (d) NVM4PF@C.
Figure 4. Raman images of (a) NVPF@C; (b) NVM2PF@C; (c) NVM3PF@C; (d) NVM4PF@C.
Molecules 28 01409 g004
Figure 5. Electrochemical performances of NVPF@C and NVMPF@C electrode materials; (a) CV curves of different samples at 0.1 mV s−1; (b) charge-discharge curves of electrode materials; (c) cycling performance at 0.2 C; (d) rate capability of electrode materials; (e) long-cycling performance of NVM2PF@C at 1 C.
Figure 5. Electrochemical performances of NVPF@C and NVMPF@C electrode materials; (a) CV curves of different samples at 0.1 mV s−1; (b) charge-discharge curves of electrode materials; (c) cycling performance at 0.2 C; (d) rate capability of electrode materials; (e) long-cycling performance of NVM2PF@C at 1 C.
Molecules 28 01409 g005
Figure 6. (a) Nyquist plots; (b) the relationship between Zre and ꞷ−1/2 in the low-frequency region of different samples charged to 4.1 V.
Figure 6. (a) Nyquist plots; (b) the relationship between Zre and ꞷ−1/2 in the low-frequency region of different samples charged to 4.1 V.
Molecules 28 01409 g006
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

Su, R.; Zhu, W.; Liang, K.; Wei, P.; Li, J.; Liu, W.; Ren, Y. Mnx+ Substitution to Improve Na3V2(PO4)2F3-Based Electrodes for Sodium-Ion Battery Cathode. Molecules 2023, 28, 1409. https://doi.org/10.3390/molecules28031409

AMA Style

Su R, Zhu W, Liang K, Wei P, Li J, Liu W, Ren Y. Mnx+ Substitution to Improve Na3V2(PO4)2F3-Based Electrodes for Sodium-Ion Battery Cathode. Molecules. 2023; 28(3):1409. https://doi.org/10.3390/molecules28031409

Chicago/Turabian Style

Su, Renyuan, Weikai Zhu, Kang Liang, Peng Wei, Jianbin Li, Wenjun Liu, and Yurong Ren. 2023. "Mnx+ Substitution to Improve Na3V2(PO4)2F3-Based Electrodes for Sodium-Ion Battery Cathode" Molecules 28, no. 3: 1409. https://doi.org/10.3390/molecules28031409

APA Style

Su, R., Zhu, W., Liang, K., Wei, P., Li, J., Liu, W., & Ren, Y. (2023). Mnx+ Substitution to Improve Na3V2(PO4)2F3-Based Electrodes for Sodium-Ion Battery Cathode. Molecules, 28(3), 1409. https://doi.org/10.3390/molecules28031409

Article Metrics

Back to TopTop