Materials Characterizations. The designed NASICON Na4 − 3xMnCr(PO4 − xFx)3 (x = 0, 0.01, 0.02, 0.05, 0.1, 0.15 and 0.2) materials were synthesized by a simple sol-gel method combined with a high-temperature sintering process under an argon atmosphere. The X-ray diffraction (XRD) results (x = 0 ~ 0.05) in Supplementary Fig. 1 present analogous patterns, which determined a pure Na4MnCr(PO4)3 phase within these samples and proved the partial introduction of F had no impact on the structure within this x value range. Of note, NaPO3 (JCPDS NO. 00-003-0688) and Na3PO4 (JCPDS NO. 00-030-1232) impurity phases appear with x value higher than 0.05, in line with the corresponding relatively larger values of Rwp in the XRD Rietveld refinement results (Supplementary Fig. 2). In Supplementary Fig. 3, it can be clearly found that the rate properties improved initially with an increased F ratio (x = 0, 0.01, 0.02, 0.05) and reached a peak value at x = 0.05, accompanied by a descending trend when further increasing x. With the purpose of disclosing the underling mechanism, the bond lengths of Na(2)-O in different samples were calculated (Supplementary Fig. 4). It can be clearly seen that at the initial stage, there was an increasing trend with a maximum Na(2)-O distance of 2.637 Å (x = 0.05) and then the distance decreased with higher fluorine content. According to previous works, longer the bond length of Na(2)-O (weaker bond strength) is associated with faster Na+ diffusion, which explains the performance discrepancy of different samples in Supplementary Fig. 3 41–43. Therefore, the material with x = 0.05 was selected as the optimal one (NMCPF; the baseline sample (x = 0) is denoted as NMCP).
XRD Rietveld refinements determined that both NMCPF (Fig. 1a) and NMCP (Supplementary Fig. 2, x = 0) were indexed to the R_3c space group with low Rwp values of 1.52% and 1.48%, respectively. The calculated geometric parameters, atomic coordinates and crystallographic information are listed in Supplementary Table 1 and Supplementary Table 4. Figure 1b provides an illustration of their crystalline structure based on the XRD Rietveld refinement results. Both samples crystalized in a similar rhombohedral NASICON structure: Cr and Mn atoms share the same 12c site with 50% fraction for each position; [PO4] or [PO3.95F0.05] tetrahedra are corner-shared with [CrO6] or [MnO6] octahedra to construct the lantern-like [MnCr(PO4)3] or [MnCr(PO3.95F0.05)3] constituents, and thus two Na+ host sites are formed. Na(1) (6b site) is sixfold coordinated while the other Na(2) (18e site) is eightfold coordinated which tends to be more easily extracted/inserted from/to the NASICON structure due to the weaker bonding force of Na(2)-O28. As anticipated, F atoms partially occupy in O1 and O2 sites with the fractions of 1.2% for each position from Supplementary Table 4. According to the valence-induced mechanism, the less negative charge stemming from the substitution of O2− by F− should be compensated by cation deficiency (i.e., Na+ deficiency) to form Na3.85MnCr(PO3.95F0.05)3, which coincides with previous reports42, 43.
Inductively Coupled Plasma- Optical Emission Spectrometry (ICP-OES) was carried out to further determine the chemical formula. As shown in Supplementary Fig. 5, the atomic ratios were measured to be Na(3.84), Mn(1), Cr(1.01) and P(2.95) for NMCPF and Na(3.96), Mn(1), Cr(1.01) and P(2.97) for NMCP, which agree well with previous discussions. These results indicate that some Na vacancies exist in NMCPF as illustrated in Fig. 1b. The Rietveld refinements results of all samples could be found in Supplementary Fig. 2; their detailed structural information was listed in Supplementary Table 1–6. Transmission electron microscopy (TEM) energy dispersive spectrometry (EDS) mapping images of NMCPF (Fig. 1f) and NMCP (Supplementary Fig. 6) present a uniform distribution of all elements throughout these particles, which confirms that F has been successfully introduced into NMCPF. Solid-state nuclear magnetic resonance (NMR) spectroscopy of 19F was performed to further study the fluorine substitution. The broad resonance of -80 ~ -180 ppm in Supplementary Fig. 7 determined that F atoms partially substituted O atoms in NMCPF42. The combination of XRD Rietveld refinements, EDS mapping, and NMR results validates that O atoms were partially occupied by F atoms within NMCPF structure, which provides compelling evidence that this light-weight F doping strategy could be applied to other phosphates or even wider polyanionic materials.
Furthermore, TEM was employed to probe the microstructure and morphology of NMCPF and NMCP. As displayed in Fig. 1c-d and Supplementary Fig. 6a-b, both NMCPF and NMCP particles are homogenously coated with an amorphous carbon layer of 2 ~ 5 nm. Similar irregular morphologies as they displayed, the size of NMCPF (0.5 ~ 1 µm) is smaller than that of NMCP (1 ~ 2 µm), indicating that the F dopant is favorable for suppressing the particle growth to some extent and thereby achieved a shortened diffusion length for Na+ in NMCPF44. Besides, high-resolution (HR) TEM images (Fig. 1d and Supplementary Fig. 6b) show a lattice fringe of 0.32 nm, which corresponds to the (20_4) lattice plane in both NMCPF and NMCP. The selected area electron diffraction (SAED) in Fig. 1e and Supplementary Fig. 6c determined the well-crystallized NMCPF and NMCP with the R_3c space group.
Thermogravimetric analysis (TGA) was adopted to determine the contents of carbon within the samples. From Supplementary Fig. 8, the amounts of carbon were shown to be 10.85% (NMCPF) and 8.83% (NMCP), respectively. As shown in Supplementary Fig. 9, NMCPF and NMCP possessed identical Fourier transform infrared (FT-IR) spectra: the stretching or bending vibration signals of [CrO6] or [MnO6] octahedra and [PO4] tetrahedra are located at ~ 1000 cm− 1 31, 37. Raman spectra in Supplementary Fig. 10 reveal two representative peaks of 1340 cm− 1 (D-band) and 1590 cm− 1 (G-band); the intensity ratios were calculated to be 0.74 (NMCPF) and 0.70 (NMCP), demonstrative of partial graphitization of carbon for these two samples and enhanced electronic conductivity for NMCPF31, 41. X-ray photoelectron spectroscopy (XPS) results (Supplementary Fig. 11) display the existence of Na, Cr, Mn, P, O, C for both NMCPF and NMCP. But it is of note that F only exists in NMCPF and the F1s has a characteristic peak at 683.93 eV, which is in good agreement with previous F-doped works42, 44. On the contrary, there are no distinguishable signals of F in NMCP. In addition, Mn 2p and Cr 2p spectra identified the signals of Mn2+ and Cr3+ in both NMCPF and NMCP45, 46. Thus, the above results confirmed that both NMCPF and NMCP were successfully prepared.
Electrochemical Performances. The CR2032 type coin cells with metallic Na anode were fabricated to evaluate the effectiveness of our strategy. In Fig. 2a, two redox couples of 3.75/3.44 V and 4.24/4.10 V could easily be found in the CV curves, which correspond to the oxidation/ reduction of Mn2+/Mn3+ and Mn3+/Mn4+, respectively37. The highly overlapped profiles indicated an excellent reversibility of de-/sodiation processes for both NMCPF and NMCP; the slightly deviated CV curves can be ascribed to the formation of the solid electrolyte interphase (SEI) layer18. And the narrowed potential gap of NMCPF (0.01 V; 0.03 V for NMCP) showed an improved kinetics after the introduction of fluorine within the NMCP structure. In Fig. 2b, two distinguishable plateaus located at 3.65 V and 4.2 V could be observed as well, thus delivering specific capacities of 110.7 mA h g− 1 (NMCPF) and 107.1 mA h g− 1 (NMCP) at 0.1 C accompanied by successive phase transformations of Na3.85MnCr(PO3.95F0.05)3 ↔ Na2.85MnCr(PO3.95F0.05)3 ↔ Na1.85MnCr(PO3.95F0.05)3 and Na4MnCr(PO4)3 ↔ Na3MnCr(PO4)3 ↔ Na2MnCr(PO4)3, respectively, which will be analyzed in detail later.
Rate properties are also of great significance for practical applications. As displayed in Fig. 2c, in general, NMCPF showed much enhanced rate performances as compared to NMCP while they delivered analogous initial capacities at 0.1 C: 110.6 mA h g− 1 for NMCPF and 107.5 mA h g− 1 for NMCP. Increasing current densities from 0.1 C to 0.2 C, 0.5 C, 1 C, 3 C, 5 C, 10 C, 20 C and 30 C, accordingly enhanced the capacity gap between NMCPF and NMCP. Notably, even under an ultrahigh current rate of 40 C, NMCPF could still deliver 60.4 mA h g− 1, which outweighs all of previous reports on Na4MnCr(PO4)336–38, 47. In comparison, NMCP only revealed a “NEAR ZERO” capacity of 15 mA h g− 1 under the same test condition. When the current rate went back to 0.5 C, a capacity of 89.1 mA h g− 1 for NMCPF was still recovered, demonstrative of the remarkable reversibility. Moreover, as depicted in Fig. 2d, the corresponding charge-discharge curves proved a restrained electrochemical polarization of NMCPF18. The above rate performances again determined a boosted kinetics after F doping in NMCP. Figure 2e also compared the performances of NMCPF with other reported cathode materials hinged in Na half-cells. It turns out that NMCPF possessing higher energy densities surpassed the majority of the proposed materials.
To meet the demands of large-scale ESSs, cycling stabilities under various current densities play pivotal roles as well. In Supplementary Fig. 12, NMCPF showed a capacity of 110.8 mA h g− 1 for the initial cycle at 0.1 C; it could stably operate for 65 cycles. Nevertheless, NMCP possessed a much poorer cyclability and only a low capacity of 81.5 mA h g− 1 could be maintained after 65 cycles under the identical condition. Figure 2f (at 1 C), Supplementary Fig. 13 (at 0.5 C) and Supplementary Fig. 14 (at 5 C) showed apparently improved capacity retentions and cycling stabilities for NMCPF in comparison to NMCP. As displayed in Supplementary Fig. 15, we also subjected the batteries to a higher rate of 10 C to investigate their long-term cyclabilities. Markedly, NMCPF could steadily cycle 500 times at 10 C (Supplementary Fig. 15) and work for 1000 cycles at 20 C (Fig. 2g) with boosted Coulombic efficiency. In contrast, capacity retentions of merely 56.7% at 10 C and 51.3% at 20 C were achieved for NMCP.
In addition, the batteries were tested within a wider voltage window of 1.5–4.5 V so as to further determine the merits of our work. In Supplementary Fig. 16a-b, both the CV profiles and charge-discharge curves exhibit three pairs of redox couples and three apparent voltage platforms. Apart from Mn2+/Mn3+ and Mn3+/Mn4+, a high-potential Cr3+/Cr4+ (4.4 V) was also activated in these two samples thereby 143 mA h g− 1 for NMCPF and 133.6 mA h g− 1 for NMCP were obtained, which coincides with previous works36, 38, 47. Furthermore, rate performances (Supplementary Fig. 16c) and the cycling stability (Supplementary Fig. 16d) of NMCPF were better than those of NMCP. Further improvements reply on the development of novel high-voltage electrolyte to endure such a harsh cut-off voltage.
Therefore, the above results confirmed the effectiveness of the F doping strategy on NMCPF. Two typical approaches were carried out to further evaluate the Na+ diffusion coefficient (DNa+) of NMCPF and NMCP. The first one is hinged on galvanostatic intermittent titration technique (GITT) method (Supplementary Fig. 17a) after five activation cycles. As seen in Supplementary Fig. 17b, overall, NMCPF displayed higher DNa+ (10− 13~10− 8 cm2 s− 1) than NMCP (10− 14~10− 8 cm2 s− 1), which indicates the enhanced Na+ kinetics after F doping in agreement with Fig. 1b and Supplementary Fig. 4. Also, Supplementary Fig. 17c showed suppressed overpotentials for NMCPF. Likewise, the batteries were subjected to various scan rates from 0.1-1.0 mV s− 1 under CV tests (Supplementary Fig. 19). The linear fitting profiles of Supplementary Fig. 19b and Supplementary Fig. 19d reveal diffusion-controlled processes during Na+ extraction/intercalation for both NMCPF and NMCP48. Accordingly, NMCPF showed higher DNa+ values than NMCP in Supplementary Fig. 17d, which were computed based on the Randles-Sevcik equation. This enhancement could be on account of the construction of Na vacancies (Fig. 1b) due to F substitution, which truly facilitated Na+ diffusion42.
Structural Evolution and Charge Compensation Analysis. Aiming to unravel the structural evolution of NMCPF and NMCP upon de-/sodiation processes, in-situ XRD tests based on a specially designed cell were performed within 1.5–4.5 V (vs. Na+/Na) for one cycle (left sides of Fig. 3a, c). Supplementary Fig. 20 depicts the full-scope patterns. Figure 3a and 3c clearly showed the reflections of (104), (2_13), (20_4), (3_11), (2_16) and (3_14) for both NMCPF and NMCP and all of them were highly reversible in general, which can be ascribed to their robust NASICON structures and explains the good reversibility of electrochemical properties as discussed above. All peaks could be indexed to the R_3c group with a typical NASICON rhombohedral phase for both fresh NMCPF and NMCP. At first upon charging to ~ 3.65 V, all reflections underwent a right-shift, which corresponded to a solid-solution reaction. It should be noted that (20_4) reflection was split into two peaks from the end of ~ 3.65 V; both (3_11) and (3_14) vanished during the first voltage platform stage, which confirmed that a second phase appeared with coexistence of the initial phase. At the end of the first voltage plateau, all vanished reflections appeared again. When charging to high voltage (3.72 V ~ 4.5 V), all the peaks gradually shifted to higher 2θ values, corresponding to solid-solution reactions. During discharging, all reflections underwent completely obsequent processes and returned to the initial positions. Therefore, the above results revealed a combination of solid-solution reactions and two-phase reactions during extraction/intercalation of Na+ for both NMCPF and NMCP; while solid-solution reactions were in the majority, which ensures the good reversibility and fast kinetics for Na+ storge. The residual Na+ in the structure at deep charged states are deemed to stabilize the overall framework as the binding pillars. The above results confirmed that F doping has no impact on Na+ storage mechanisms for NMCP.
In addition, with the purpose of obtaining a clearer understanding regarding their structural transitions, we collected the lattice parameters at all stages, as presented in Fig. 3a and c. As displayed in Fig. 3b and 3d, two samples underwent analogous transitions for all lattices parameters: to illustrate, at the beginning upon charging, a (= b)-axis slightly decreased but the c-axis kept invariant thus the decreased v (volume) could be obtained, corresponding to the structural shrinkage; when further charging, the decreasing rate was facilitated for a (= b)-axis as well as v although the c-axis was only slightly increased, which is in good agreement with solid-solution reactions; upon discharging, all parameters experienced the reverse processes and finally returned to their initial states, supportive of good reversibility for both NMCPF and NMCP. As a consequence, the volume change of NMCPF (8.9%) was much lower than that of NMCP (10.6%), which strongly confirmed that the F doping strategy is conducive to improve structural stability. This may be due to the high electronegativity of F and the increased F-metal interaction, which will be discussed in detail afterwards. The extent of peak deviation for NMCPF (Fig. 3a) was also smaller than that of NMCP (Fig. 3c), which coincides well with the results in Fig. 3b and 3d. The ex-situ TEM and SAED results of fully charged and discharged NMCPF cathodes are shown in Fig. 3e-i. Some obvious lattice fringes and diffraction spots could be clearly identified, demonstrating that the well-crystallized structure was maintained when even subjected to a deep charge state. Such an enhanced structural stability guaranteed the boosted cyclability of NCMPF after fluorine doping as displayed in Fig. 2.
Additionally, ex-situ XPS was carried out to further determine the charge compensation mechanisms of NMCPF and NMCP (Supplementary Fig. 21). The transformations of Na3.85MnCr(PO3.95F0.05)3 → Na2.85MnCr(PO3.95F0.05)3 → Na1.85MnCr(PO3.95F0.05)3 → Na0.85MnCr(PO3.95F0.05)3 or Na4MnCr(PO4)3 → Na3MnCr(PO4)3 → Na2MnCr(PO4)3 → Na1MnCr(PO4)3 are anticipated to be accompanied with the changes of valence states of transition metals (i.e., Mn and Cr) during charging38. The initial binding energies of Mn 2p3/2 (641.68 eV) and Cr 2p3/2 (577.73 eV) of NMCPF and NMCP are the same as that of the corresponding powders, indicating Mn2+ and Cr3+ in the fresh electrodes45, 46. When charged to 3.67 V and 4.26 V, the peaks of Mn 2p3/2 shifted to higher binding energies of 642 eV and 642.2 eV, respectively, indicative of oxidations of manganese from pristine Mn2+ to Mn3+ and further to Mn4+ 45. While the binding energies of Cr 2p3/2 were kept constant (577.73 eV) in this process, which demonstrated that this cut-off voltage is not high enough to activate the redox behavior of Cr3+ 37, 46. Whereafter, when the battery was subjected to 4.5 V, there was a left-shift in the Cr 2p3/2 peak (578.3 eV), a signal of Cr4+ 49. But Mn 2p3/2 remained at the same position as that for Mn4+. Therefore, the first two platforms were mainly attributed to the successive oxidations of Mn2+ to Mn3+ and Mn4+ while the last high-voltage plateaus coincided with the oxidation of Cr3+ to Cr4+. The calculated magnetic moments (µB) on the Mn and Cr sites for both NMCPF and NMCP with different Na contents were collected in Supplementary Table 7 and Supplementary Table 8, respectively. The changes of µB were in good agreement with the ex-situ XPS results. Besides, it should be noted that the whole redox reaction processes were identical for both NMCPF and NMCP (Supplementary Fig. 13). Thus, one can conclude that appropriate F modification doesn’t alter the charge compensation mechanism of NMCP.
DFT calculations of the Na + storage mechanism. To obtain in-depth insights on the physicochemical properties of NMCPF and NMCP from a theoretical perspective, we further employed density functional theory (DFT) calculations. The bond valence (BV) method has been widely adopted to identify the ionic diffusion pathways within structures as a typical empirical approach. From the BV maps in Fig. 4a, it can be clearly found that both NMCPF and NMCP possess 3D well-interconnected pathways for Na+ with the lowest energy regions except for some subtle differences, which is the feature of typical NASICONs23, 26. Convex-hull phase diagrams of NMCPF (Fig. 4b) and NMCP (Fig. 4c) were constructed hinged on the formation energy at different de-/sodiation states, which accompanied the structural transformations when cycling. All the metastable phases from Na4MnCr(PO4)3 (or Na3.85MnCr(PO3.95F0.05)3) to Na0MnCr(PO4)3 (or Na0MnCr(PO3.95F0.05)3) could be electrochemically achieved on account of the negative values of formation energies, which may be the intrinsic origin for multi-electron redox reactions as displayed in Fig. 2 and Supplementary Fig. 16. Of note, Na2MnCr(PO4)3 or Na1.85MnCr(PO3.95F0.05)3 emerged as the most stable phase due to the lowest formation energy, which well explained the relatively good performance between 1.5–4.3 V.
The voltage profiles of NMCPF (Fig. 4d) and NMCP (Fig. 4e) were also calculated with regard to Mn2+/Mn3+, Mn3+/Mn4+ and Cr3+/Cr4+ redox couples, both of which fitted well with the experimental (GITT) results (with slight deviations). In general, the calculated voltage platforms of NMCPF were relatively higher than that of NMCP, which may be ascribed to the high electronegativity of F (similar to Na3V2(PO4)3 and Na3V2(PO4)2F3 cases).50 An ultrahigh voltage plateaus at ~ 5 V is anticipated to be activated from the calculation if an optimized electrolyte that endures such a high voltage could be developed, which would greatly increase the overall energy density. In addition, the density of states (DOS) diagrams (Fig. 4f and Fig. 4g) were obtained, from which we could find the hybridized Na 3s, Mn 3d, Cr 3d, P 2p, and O 2p orbitals in both NMCPF and NMCP; one more F 1s orbital existed in NMCPF. It turned out that the forbidden band gap was greatly decreased from 1.52 eV in NMCP to 0.22 eV in NMCPF, which contributed to the enhanced electronic conductivity after fluorine modification. The rest DOS patterns with different Na contents are displayed in Supplementary Fig. 14. The electron density difference in NMCPF (Fig. 4h) and NMCP (Fig. 4i) further supported the modification of electron density through F doping. These results accounted for better electrochemical properties of NMCPF from a kinetics perspective.
To further reveal the origin of the capacity decay of the pristine NMCP, we dissembled the cycled batteries in the glove box to collect the electrolyte and the separator and measure the separate manganese concentration through ICP-OES method. Like other Mn-based materials, NMCP also suffers from severe Mn dissolution with the concentration of 0.24 mg L− 1 in the electrolyte and 8 ppm in the separator after cycling, which could account for the inferior electrochemical performances in Fig. 251. In sharp contrast, the manganese dissolution issue in NMCPF was effectively suppressed with a low concentration of 0.05 mg L− 1 in the electrolyte and 1 ppm in the separator after cycling. In addition, we employed a theoretical approach to explain this phenomenon. The Crystal Orbital Hamilton Population (COHP) approach has been acknowledged as a reliable tool to visualize the chemical bonding in the battery field in recent years52–54. The Kohn-Sham states were initially projected into atomic orbitals and subsequently the mutual orbital overlap population was inspected. The COHP and ICOHP (integrated COHP) results are shown in Fig. 5b-f, supplementary Fig. 23, supplementary Fig. 24 and supplementary Table 9–10 based on the adjacent six O/F atoms around a specific Mn atom. As reported, the negative value of COHP of a Mn-O/F pair represents constructive (namely, bonding) interference of atomic orbitals; the modulus of COHP corresponds to the degree of covalency between Mn and O/F and the zero point of COHP suggests non-interacting bonds or a pure ionic bond. In each pattern of Fig. 5c-f, supplementary Fig. 23, and supplementary Fig. 24, the intersection between ICOHP curve and Fermi level is the actual -ICOHP value. From Fig. 5b, NMCPF possesses a higher accumulated -ICOHP value (all six Mn-O/F pairs including spin-up and spin-down direction) of 4.91 eV than NMCP (4.38 eV). This result indicate that F doping strengthens the overall chemical bonding of Mn-O/F bonds in the local scope, which may be associated with the high electronegativity of F. The strengthened chemical bonding between adjacent O/F and Mn helps explain the suppressed manganese dissolution.