Next Article in Journal
An Ad-Hoc Initial Solution Heuristic for Metaheuristic Optimization of Energy Market Participation Portfolios
Previous Article in Journal
The Comparison Study of Short-Term Prediction Methods to Enhance the Model Predictive Controller Applied to Microgrid Energy Management
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Synthesis, Structure, and Sodium Mobility of Sodium Vanadium Nitridophosphate: A Zero-Strain and Safe High Voltage Cathode Material for Sodium-Ion Batteries

1
Helmholtz Institute Ulm, Helmholtzstraße 11, 89081 Ulm, Germany
2
Karlsruhe Institute of Technology (KIT), P.O. Box 3640, 76021 Karlsruhe, Germany
*
Authors to whom correspondence should be addressed.
Energies 2017, 10(7), 889; https://doi.org/10.3390/en10070889
Submission received: 6 June 2017 / Revised: 22 June 2017 / Accepted: 28 June 2017 / Published: 30 June 2017
(This article belongs to the Section D: Energy Storage and Application)

Abstract

:
Herein, the nitridophosphate Na3V(PO3)3N is synthesized by solid state method. X-ray diffraction (XRD) and Rietveld refinement confirm the cubic symmetry with P213 space group. The material exhibits very good thermal stability and high operating voltage of 4.0 V vs. Na/Na+ due to V3+/V4+ redox couple. In situ X-ray diffraction studies confirm the two-phase (de-)sodiation process to occur with very low volume changes. The refinement of the sodium occupancies reveal the low accessibility of sodium cations in the Na2 and Na3 sites as the main origin for the lower experimental capacity (0.38 eq. Na+, 28 mAh g−1) versus the theoretical one (1.0 eq. Na+, 74 mAh g−1). These observations provide valuable information for the further optimization of this materials class in order to access their theoretical electrochemical performance as a potentially interesting zero-strain and safe high-voltage cathode material for sodium-ion batteries.

Graphical Abstract

1. Introduction

The growing demand for sustainable energy leads to an increasing interest in efficient energy storage systems (EES). Among these, lithium-ion batteries (LIBs) have received strong scientific attention and, indeed, have undergone rapid development since their commercialization in 1991 [1]. However, the fact of continued lithium resource consumption and its limited abundance in earth’s crust has driven people to renew their interest in beyond-lithium chemistries [2]. Sodium ion batteries (NIBs) have, therefore, been reconsidered as promising alternatives due to the low cost and high abundance of Na resource, and good electrochemical performance [3,4]. The main challenge associated with these systems is to optimize energy density, rate capability, and sustainability of the active materials [5,6]. Since the energy of a battery is largely dependent on the gravimetric and volumetric capacities and operating voltages of the electrode materials, exploration on suitable active materials is highly-targeted.
In this pursuit, there have been intensive efforts on expanding the range of available structural families. One strategy is to develop Na-containing host materials from lithium analogues. Such efforts have resulted in tremendous successes in realizing some known polyanionic frameworks, such as phosphate NaMPO4/Na3M2(PO4)3 [7,8], pyrophosphate Na2MP2O7 [9], fluorophosphate NaMPO4F/Na3M2(PO4)2F3 [10,11], sulfate NaxM(SO4)2 [12], fluorosulfate NaMSO4F [13], silicate Na2M(SiO4) [14], and mixed polyanionic Na4M3(PO4)2(P2O7) [15,16] compounds (where M represents the transition metal ion). Basically, the electrochemical properties of anticipated polyanion electrode materials, including the operating potential and specific capacity, rely on the Mn+/M(n+x)+ redox process and the natures of polyanionic frameworks, i.e., electronegativity and molecular weight [17]. Among these reported polyanion cathodes, the NASICON-type Na3V2(PO4)3 in a rhombohedral structure with the space group R-3c attracts particular interest due to the vanadium redox couple, enabling a relatively high working voltage of ~3.4 V (vs. Na+/Na), and the framework of PO4, enabling superior thermal stability [8,18]. During the course of the investigation of the vanadium-based phosphate family, it was found that doping with the more electronegative fluoride can enhance the operating potential of electroactive materials. For example, the Na3V2(PO4)2F3 compound shows promising potential improvement due to the high electronegativity of F ions [19,20].
Of particular interest are the members of a new compounds class with general formula AI2MII2(PO3)3N, where A = Na or Li, M = Mg, Fe, Mn, and Co, which are addressed as potential electrode materials for lithium ion batteries. In fact, the polyanionic structure with N3−, arising from three PO3N tetrahedra sharing one N3−, can provide intriguing voltage as cathodes. The divalent nitridophosphate analogue was firstly investigated as cathode for lithium ion batteries, with the Fe2+/Fe3+ redox couple [21]. Due to the strong inductive effect of PO3N groups, Li2Fe2(PO3)3N delivers a high potential, as high as 3.5 V (vs. Li+/Li), which is 0.1 V higher than the olivine LiFePO4, and a high theoretical capacity of 142 mAh g−1, demonstrating the possibility to design high potential cathode materials for Li-ion batteries. On the Na-ion battery side, cubic Na3Ti(PO3)3N exhibits good reversibility for sodium ion (de-)insertion with extremely small volume changes during the sodiation and desodiation process [22]. However, this material only exhibits an average working potential of 2.7 V vs. Na+/Na originating from the Ti3+/Ti4+ redox couple, and a theoretical capacity of 74 mAh g−1. To maximize the energy density of this kind of materials, the most significant improvement lies in the flexibility of increasing the open-circuit voltage by manipulation of the 3d-metal Mn+/M(n+1)+ redox couple in the polyanionic framework.
Herein, we report for the first time, the electrochemical properties of cubic Na3V(PO3)3N as cathodes for NIBs, with V3+/4+ redox activities. The nitridophosphate was synthesized by conventional solid state method. As a cathode in non-aqueous NIBs, this material exhibits an average operating potential around 4.0 V vs. Na+/Na, while 0.74 Na+ can be extracted from the host structure.

2. Materials and Methods

2.1. Mateirals Synthesis

The Na3V(PO3)3N powder was synthesized via a conventional one-step solid state method. Typically, sodium polyphosphate ((NaPO3)n, pure, Acros Organics, Morris Plains, NJ, USA) and vanadium oxide (V2O5, ≥99.6%, ~10 mesh, Alfa Aesar, Karlsruhe, Germany) in a molar ratio of 6:1 were manually ground with mortar and pestle at ambient temperature for 30 min. Then the yellow powders were placed in an alumina crucible with excessive melamine (5 molar equivalents, C3H6N6, Sigma Aldrich, ≥99%, St. Louis, MO, USA) as nitrogen sources, followed by a pre-heating treatment at 200 °C for 3 h and final calcination at 700 °C for 10 h under extremely slow Ar flow (heating rate: 5 °C min−1).

2.2. Structural Characterization

X-ray diffraction analysis was performed in Bragg–Brentano geometry using a Bruker D8 Advance diffractometer (Bruker, Karlsruhe, Germany). All diffractograms were recorded with CuKα radiation (λ = 0.154 nm) in the 2θ range of 10–90° with a step size of 0.01° and a time per step of 0.5 s. The structures were refined using the Rietveld method with the DIFFRAC PLUS TOPAS 4.2 software (Bruker, Germany). Morphological characterization was conducted on ZEISS 1550VP Field Emission SEM (Carl Zeiss, Jena, Germany). Thermal stability was determined by a thermogravimetric analyzer (TGA-209F, Netzsch, Selb, Germany) from RT to 800 °C (heating rate = 5 °C min−1) under steady N2 and O2 flows (20 mL min−1), respectively. In situ XRD analyses, coupled with galvanostatic sodiation and desodiation, were performed.

2.3. Electrochemical Characterization

The electrochemical tests were performed using three electrode Swagelok-type cells assembled in an argon-filled glove box. Sodium half-cells were assembled by using glass fiber (Whatman) as separator, sodium metal (99.8%, Acros Organics) as counter and reference electrodes, and 1 M NaPF6 (99%, Alfa Aesar) in propylene carbonate (PC) with 2 wt.% fluoroethylene carbonate (FEC) (Battery Grade, BASF, Ludwigshafen, Germany) as electrolyte. The working electrodes were prepared by doctor-blade casting a slurry of 80 wt.% of active material, 10 wt.% of carbon black (Super C65, IMERYS, Bodio, Switzerland) and 10 wt.% of polyvinylidenefluoride (6020 Solef, Arkema Group, Paris, France) in N-methyl-2-pyrrolidone (anhydrous, Sigma-Aldrich, St Louis, MO, USA) on Al foil as current collector. After drying, electrodes with a diameter of 12 mm were punched, pressed, and dried at 120 °C in vacuum. The mass loading of the electrodes is ~3.5 mg cm−2. A Maccor 4000 Battery system (Maccor, Tulsa, OK, USA) was used for the galvanostatic charge/discharge test at 10 mA g−1 and 20 mA g−1. Cyclic voltammetry was performed on a multichannel potentiostat/galvanostat (VMP3, Biologic Science Instruments, Claix, France). All measurements were performed in climatic chambers at a temperature of 20 ± 1 °C.

3. Results and Discussion

Pristine Na3V(PO3)3N powders were synthesized by conventional sold state method, but different from other reports on the nitridophosphate analogues synthesis, without flowing ammonia gas [21,22,23]. The crystal structure of Na3V(PO3)3N was determined by X-ray diffraction (XRD, Figure 1), including Rietveld refinement of the obtained XRD pattern. The results of the refinement confirm, that Na3V(PO3)3N adopts a cubic structure (a = 9.440 (1) Å) with a space group of P213 and, thus, is isostructural to Na3Ti(PO3)3N (ICSD: 172056) [22]. The refinement shows satisfactory agreement factors (Rwp (weighted-profile R-factor) = 8.7%, GOF (goodness of fit) = 1.41). The diffractogram in Figure 1 does not depict crystalline impurity phases, suggesting that the simplified synthesis method herein used is successful to obtain pure cubic Na3M(PO3)3N phase. The refined cell parameters and atomic occupancy values are summarized in Supplementary Materials Tables S1 and S2, respectively. The SEM image (inset in Figure 1) shows that the synthesized powder is composed of irregular-shaped microscale particles. The refined crystal structure and the connectivity of trimers of PO3N tetrahedra with the local environments of Na and V atoms are shown in Figure 2. The structure of this compound is composed of VO6 octahedra and PO3N tetrahedra. Three PO3N tetrahedra build up a nitridotriphosphate group, (PO3)3N sharing a common N3−-anion. As shown in Figure 2b, Na+ ions occupy three distinct sites. One Na+ ion is coordinated by six O atoms and one N atom (Na1), while the other two Na+ cations are surrounded by six (Na2) and nine O anions (Na3: six more long-distanced O atoms are not shown in the structure), respectively [24]. Based on the calculation results on the Na3Ti(PO3)3N system, Na1 and Na3 sites show high mobility for Na migration, while Na2 site is inaccessible due to the high energy barrier. This cubic structure, bridged by trimers of PO3N tetrahedra, offers a three-dimensional framework for Na-ion migration and contributes to the structural stability [24].
Thermal gravimetric analysis (TGA) was used to assess the thermal stability of the material under both inert (N2) and oxygen (O2) atmospheres. The TGA curves in the temperature range of RT-800 °C are displayed in Figure S1 (see Supplementary Materials). It should be noted that there is no weight decrease under N2 flow up to 800 °C, indicating the superior thermal stability of the material. However, when exposed to O2, a weight increase starting from 400 °C is observed, resulting from the oxidation of V3+ to V5+ and V4+, which was also observed in Na3V2(PO4)3 system [25]. Apart from this oxidization process, it should be highlighted that no thermal decomposition is evident during heating, illustrating the excellent thermal stability, which is why Na3V(PO3)3N can be regarded as safe electrode materials in batteries. It should be noted that the melamine completely decomposes between 250–350 °C and does not introduce any additional carbon or impurities into the material, as indicated by the thermogravimetric analysis of pristine melamine in N2 atmosphere (Figure S1, Supplementary Materials).
The electrochemical performance of pristine Na3V(PO3)3N electrode was investigated in T-type cells with Na-metal as counter and reference electrode. Cyclic voltammograms at various scan rates between 3.2–4.5 V vs. Na/Na+ are presented in Figure 3a. The CV sweeps reveal that the Na3V(PO3)3N electrode shows electrochemical activity centered at 4.0 V with only one pair of redox peaks associated to the V3+/4+ redox couple. Despite the slight cell polarization, the reversibility of the sodium insertion is maintained at a higher scan rate. Figure 3b shows the first galvanostatic charge and discharge (GCD) curve of the cell between 3.2 and 4.5 V at a current of 10 mA g−1. The electrode exhibits a single plateau around 4.0 V in both charge and discharge. During the first charge to 4.5 V, 0.74 eq. of Na (55 mAh g−1) are extracted from the host structure. However, only half (0.38 eq.) of the Na ions were re-inserted, leading to a high irreversible capacity for the first cycle. The theoretical capacity for the reversible shuttling of 1.0 eq. Na+ per each formula unit of Na3V(PO3)3N is 72 mAh g−1. The lower charge capacity in the first cycle, hence, indicates that not all Na+ can be extracted from the host structure. Increasing the current density to 20 mA g−1, this material delivers a discharge capacity of 24 mAh g−1 (see Supplementary Materials Figure S2). It is worth mentioning that, although the capacity is low, the measured discharge capacities are stable over 10 cycles, suggesting that after optimization such electrodes could work with high cycling stability after the initial capacity loss. For instance, nano-sizing [26], carbon coating [27], and doping [28] are promising strategies to improve the rate capability due to shortened ion diffusion pathways and enhanced electronic and ionic conductivities, respectively.
The apparent Na+ diffusion coefficients may be derived from the CV profiles at various scan rates. Indeed, if the charge transfer at the interface is fast enough, the relationship of the peak current and the CV sweep rate can be described via the simplified Randles–Sevcik Equation (1) [25,29]
i p = k   n 3 / 2   A   D N a + 1 / 2   C N a +   v 1 / 2
where ip is the peak current (A), k is a constant value of 2.69 × 105 C mol1 V1/2 under standard conditions (25 °C), n is the number of electrons (1), A is the contact area between electrode and electrolyte (1.13 cm2), and CNa+ is the concentration of sodium ions in the material (2.37 × 10−4 mol cm−3 in Na3V(PO3)3N), ν is the scan rate (V s−1). When plotting the peak current ip versus the square root of the scan rate ν, a linear behavior is observed (see Figure S3). The apparent Na+ diffusion coefficients are 1.15 × 10−12 cm2 s−1 (cathodic) and 7.13 × 10−13 cm2 s−1 (anodic), respectively. These values are much lower than those for Na3Ti(PO3)3N but similar to those for Na3V2(PO4)3 [22,25].
In situ XRD measurement was conducted to understand the structural changes during the initial sodiation and desodiation processes. The XRD patterns (Figure 4b) were collected during the first cycle, each after extraction or insertion of 0.025 eq. Na+ (in total: 44 diffraction patterns). The voltage profile resulting from the galvanostatic steps during the in situ characterization is presented in Figure 4a. From Figure 4b, only imperceptible changes of the intensity and position of the reflections are observed. This suggests that only small structural and volume changes occur upon (de-)sodiation and supports a stable two-phase reaction (Na-poor and Na-rich), as also indicated by the flat charge-discharge potential profiles [30]. The small intensity changes in the XRD powder patterns for the desodiated and sodiated samples (Figure S3) are consistent with the observed electrochemical reversibility of the system. To further investigate the desodiated and sodiated phases, Rietveld refinements of the XRD patterns of electrodes in fully charged (4.5 V) and discharged (3.2 V) state were performed. The according Na occupancies and lattice parameters are listed in Table 1. All diffractograms were successfully refined with the space group P213, i.e., cubic lattice parameters. The results demonstrate that Na+ ions at Na1 and Na3 sites are firstly extracted when charged to 4.5 V, resulting in a total amount of about 0.7 eq. Na+. Na cations at these two sites are, therefore, more mobile. During the discharge process, 0.3 eq. of Na+ cannot be electrochemically re-inserted into the host structure, which is in agreement with the irreversible capacity in the first cycle as shown in Figure 3b. This preliminary result indicates that less capacity contribution is from Na3 site resulting from the higher diffusion barrier for Na3 than Na1, and there is almost no Na migration occurring at the Na2 site, in accordance with the previous study on Na3Ti(PO3)3N [22]. However, the Na ion mobility at Na1 sites is highly reversible. This insight may provide a strategy to improve the Na-migration by tuning the structural properties via selective doping, for example. Moreover, the volume change (ΔV/V), obtained from the calculated cell volume difference before (OCV) and after (4.5 V) Na removal by Rietveld refinement, that occurs on Na-ion removal is approximately 0.2%, a remarkably small volume shrinkage given the large ionic radius of Na ions. This is contributed by the framework constructed of PO3N tetrahedra trimers.

4. Conclusions

Summarizing, the nitridophosphate Na3V(PO3)3N with a cubic structure can be successfully synthesized via a conventional solid-state method without flowing ammonia gas as nitrogen source. This compound shows a high operating potential of 4.0 V vs. Na/Na+ due to V3+/4+ redox reactions. In the first charge, 0.74 eq. Na+ can be extracted but only 0.38 eq. Na+ (i.e., 28 mAh g−1) can be re-inserted into the framework and reversibly shuttled. It should be noted that the potential of V3+/4+ in most polyanionic cathodes is lower than 4.0 V vs. Na/Na+. The nitridophosphate Na3V(PO3)3N exhibits a V3+/4+ redox reaction at 4.0 V, higher than that of Na3V2(PO4)3 (3.4 V vs. Na/Na+). Although this kind of compound offers a limited reversible capacity, room to improve the performance and narrow the gap between the practical and theoretical capacities is existing. The most obvious is to improve the conductivity and optimize the particle size with morphology control via, e.g., nano-sizing and carbon-coating. However, the excellent thermal stability and very small volume changes during (de)sodiation, i.e., “zero strain”, with high operating voltage, make this kind of compounds a potentially interesting safe high-voltage cathode material for sodium-ion batteries.

Supplementary Materials

The following are available online at www.mdpi.com/1996-1073/10/7/889/s1, Table S1: Crystallographic data of Na3V(PO3)3N, Table S2: Atomic parameters for Na3V(PO3)3N, Figure S1: TG curves of Na3V(PO3)3N and melamine under N2 and O2 flows of 20 mL min−1 within the temperature range of 30–800 °C at 5 °C min−1 heating rate, Figure S2: Galvanostatic charge-discharge profiles of Na3V(PO3)3N at current density of 10 mA g−1 and 20 mA g−1 for the 1st and 10th cycles, Figure S3: Peak current (Ip, A cm2) vs. square root of scan rate (v1/2, V1/2 s1/2) and related linear fit corresponding to the V3+/4+ redox processes, Figure S4: In situ X-ray diffraction patterns of Na3V(PO3)3N at OCV, after the 1st charge (4.5 V) and 1st discharge state (3.2 V).

Acknowledgments

The author, Huang Zhang, gratefully acknowledges the financial support from Chinese State Scholarship Fund. The financial support of Helmholtz Association is also acknowledged.

Author Contributions

Huang Zhang designed and performed the experiments; Huang Zhang wrote the original manuscript; Daniel Buchholz and Stefano Passerini revised the paper and supervised the work. All authors read and approved the final manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Tarascon, J.M.; Armand, M. Issues and challenges facing rechargeable lithium batteries. Nature 2001, 414, 359–367. [Google Scholar] [CrossRef] [PubMed]
  2. Choi, J.W.; Aurbach, D. Promise and reality of post-lithium-ion batteries with high energy densities. Nat. Rev. Mater. 2016, 1, 16013. [Google Scholar] [CrossRef]
  3. Slater, M.D.; Kim, D.; Lee, E.; Johnson, C.S. Sodium-ion batteries. Adv. Funct. Mater. 2013, 23, 947–958. [Google Scholar] [CrossRef]
  4. Yabuuchi, N.; Kubota, K.; Dahbi, M.; Komaba, S. Research development on sodium-ion batteries. Chem. Rev. 2014, 114, 11636–11682. [Google Scholar] [CrossRef] [PubMed]
  5. Ong, S.P.; Chevrier, V.L.; Hautier, G.; Jain, A.; Moore, C.; Kim, S.; Ma, X.; Ceder, G. Voltage, stability and diffusion barrier differences between sodium-ion and lithium-ion intercalation materials. Energy Environ. Sci. 2011, 4, 3680–3688. [Google Scholar] [CrossRef]
  6. Palomares, V.; Serras, P.; Villaluenga, I.; Hueso, K.B.; Carretero-Gonzalez, J.; Rojo, T. Na-ion batteries, recent advances and present challenges to become low cost energy storage systems. Energy Environ. Sci. 2012, 5, 5884–5901. [Google Scholar] [CrossRef]
  7. Lee, K.T.; Ramesh, T.N.; Nan, F.; Botton, G.; Nazar, L.F. Topochemical synthesis of sodium metal phosphate olivines for sodium-ion batteries. Chem. Mater. 2011, 23, 3593–3600. [Google Scholar] [CrossRef]
  8. Saravanan, K.; Mason, C.W.; Rudola, A.; Wong, K.H.; Balaya, P. The first report on excellent cycling stability and superior rate capability of Na3V2(PO4)3 for sodium ion batteries. Adv. Energy Mater. 2013, 3, 444–450. [Google Scholar] [CrossRef]
  9. Kim, H.; Shakoor, R.A.; Park, C.; Lim, S.Y.; Kim, J.-S.; Jo, Y.N.; Cho, W.; Miyasaka, K.; Kahraman, R.; Jung, Y.; et al. Na2FeP2O7 as a promising iron-based pyrophosphate cathode for sodium rechargeable batteries: A combined experimental and theoretical study. Adv. Funct. Mater. 2013, 23, 1147–1155. [Google Scholar] [CrossRef]
  10. Ellis, B.L.; Makahnouk, W.R.M.; Rowan-Weetaluktuk, W.N.; Ryan, D.H.; Nazar, L.F. Crystal structure and electrochemical properties of A2MPO4F fluorophosphates (A = Na, Li; M = Fe, Mn, Co, Ni). Chem. Mater. 2010, 22, 1059–1070. [Google Scholar] [CrossRef]
  11. Chihara, K.; Kitajou, A.; Gocheva, I.D.; Okada, S.; Yamaki, J.-I. Cathode properties of Na3M2(PO4)2F3 [M = Ti, Fe, V] for sodium-ion batteries. J. Power Sources 2013, 227, 80–85. [Google Scholar] [CrossRef]
  12. Singh, P.; Shiva, K.; Celio, H.; Goodenough, J.B. Eldfellite NaFe(SO4)2: An intercalation cathode host for low-cost Na-ion batteries. Energy Environ. Sci. 2015, 8, 3000–3005. [Google Scholar] [CrossRef]
  13. Barpanda, P.; Chotard, J.-N.; Recham, N.; Delacourt, C.; Ati, M.; Dupont, L.; Armand, M.; Tarascon, J.-M. Structural, transport, and electrochemical investigation of Novel AMSO4F (A = Na, Li; M = Fe, Co, Ni, Mn) metal fluorosulphates prepared using low temperature synthesis routes. Inorg. Chem. 2010, 49, 7401–7413. [Google Scholar] [CrossRef] [PubMed]
  14. Li, S.; Guo, J.; Ye, Z.; Zhao, X.; Wu, S.; Mi, J.-X.; Wang, C.-Z.; Gong, Z.; McDonald, M.J.; Zhu, Z.; et al. A zero-strain Na2FeSiO4 as novel cathode material for sodium ion batteries. ACS Appl. Mater. Interfaces 2016, 8, 17233–17238. [Google Scholar] [CrossRef] [PubMed]
  15. Wood, S.M.; Eames, C.; Kendrick, E.; Islam, M.S. Sodium ion diffusion and voltage trends in phosphates Na4M3(PO4)2P2O7 (M = Fe, Mn, Co, Ni) for possible high-rate cathodes. J. Phys. Chem. C 2015, 119, 15935–15941. [Google Scholar] [CrossRef]
  16. Zhang, H.; Hasa, I.; Buchholz, D.; Qin, B.; Geiger, D.; Jeong, S.; Kaiser, U.; Passerini, S. Exploring the Ni redox activity in polyanionic compounds as conceivable high potential cathodes for Na rechargeable batteries. NPG Asia Mater. 2017, 9, e370. [Google Scholar] [CrossRef]
  17. Barpanda, P.; Ati, M.; Melot, B.C.; Rousse, G.; Chotard, J.N.; Doublet, M.L.; Sougrati, M.T.; Corr, S.A.; Jumas, J.C.; Tarascon, J.M. A 3.90 V iron-based fluorosulphate material for lithium-ion batteries crystallizing in the triplite structure. Nat. Mater. 2011, 10, 772–779. [Google Scholar] [CrossRef] [PubMed]
  18. Zatovsky, I. NASICON-type Na3V2(PO4)3. Acta Crystallogr. Sect. E 2010, 66, i12. [Google Scholar] [CrossRef] [PubMed]
  19. Gover, R.K.B.; Bryan, A.; Burns, P.; Barker, J. The electrochemical insertion properties of sodium vanadium fluorophosphate, Na3V2(PO4)2F3. Solid State Ion. 2006, 177, 1495–1500. [Google Scholar] [CrossRef]
  20. Liu, Z.; Hu, Y.-Y.; Dunstan, M.T.; Huo, H.; Hao, X.; Zou, H.; Zhong, G.; Yang, Y.; Grey, C.P. Local structure and dynamics in the Na ion battery positive electrode material Na3V2(PO4)2F3. Chem. Mater. 2014, 26, 2513–2521. [Google Scholar] [CrossRef]
  21. Liu, J.; Yu, X.; Hu, E.; Nam, K.-W.; Yang, X.-Q.; Khalifah, P.G. Divalent iron nitridophosphates: A new class of cathode materials for Li-ion batteries. Chem. Mater. 2013, 25, 3929–3931. [Google Scholar] [CrossRef]
  22. Liu, J.; Chang, D.; Whitfield, P.; Janssen, Y.; Yu, X.; Zhou, Y.; Bai, J.; Ko, J.; Nam, K.-W.; Wu, L.; et al. Ionic conduction in cubic Na3TiP3O9N, a secondary Na-ion battery cathode with extremely low volume change. Chem. Mater. 2014, 26, 3295–3305. [Google Scholar] [CrossRef]
  23. Massiot, D.; Conanec, R.; Feldmann, W.; Marchand, R.; Laurent, Y. NMR characterization of the Na3AlP3O9N and Na2Mg2P3O9N nitridophosphates:  Location of the (NaAl)/Mg2 substitution. Inorg. Chem. 1996, 35, 4957–4960. [Google Scholar] [CrossRef] [PubMed]
  24. Kim, M.; Kim, S.-J. Synchrotron powder study of Na3V(PO3)3N. Acta Crystallogr. Sect. E 2013, 69, i34. [Google Scholar] [CrossRef] [PubMed]
  25. Zhang, H.; Hasa, I.; Qin, B.; Diemant, T.; Buchholz, D.; Behm, J.; Passerini, S. Excellent cycling stability and superior rate capability of Na3V2(PO4)3 cathodes enabled by nitrogen-doped carbon interpenetration for sodium-ion batteries. ChemElectroChem 2017, 4, 1256–1263. [Google Scholar] [CrossRef]
  26. Wang, Y.; Cao, G. Developments in nanostructured cathode materials for high-performance lithium-ion batteries. Adv. Mater. 2008, 20, 2251–2269. [Google Scholar] [CrossRef]
  27. Gaberscek, M.; Dominko, R.; Jamnik, J. Is small particle size more important than carbon coating? An example study on LiFePO4 cathodes. Electrochem. Commun. 2007, 9, 2778–2783. [Google Scholar] [CrossRef]
  28. Wagemaker, M.; Ellis, B.L.; Lützenkirchen-Hecht, D.; Mulder, F.M.; Nazar, L.F. Proof of supervalent doping in olivine LiFePO4. Chem. Mater. 2008, 20, 6313–6315. [Google Scholar] [CrossRef]
  29. Hung, T.-F.; Cheng, W.-J.; Chang, W.-S.; Yang, C.-C.; Shen, C.-C.; Kuo, Y.-L. Ascorbic acid-assisted synthesis of mesoporous sodium vanadium phosphate nanoparticles with highly sp2-coordinated carbon coatings as efficient cathode materials for rechargeable sodium-ion batteries. Chem. Eur. J. 2016, 22, 10620–10626. [Google Scholar] [CrossRef] [PubMed]
  30. Recham, N.; Chotard, J.N.; Dupont, L.; Delacourt, C.; Walker, W.; Armand, M.; Tarascon, J.M. A 3.6 V lithium-based fluorosulphate insertion positive electrode for lithium-ion batteries. Nat. Mater. 2010, 9, 68–74. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Rietveld refinement and powder X-ray diffraction (XRD) pattern for Na3V(PO3)3N powders. The inset figure displays the SEM images.
Figure 1. Rietveld refinement and powder X-ray diffraction (XRD) pattern for Na3V(PO3)3N powders. The inset figure displays the SEM images.
Energies 10 00889 g001
Figure 2. (a) Refined crystal structure of Na3V(PO3)3N and (b) the connectivity of trimers of PO3N tetrahedra with the local environments of Na, V atoms.
Figure 2. (a) Refined crystal structure of Na3V(PO3)3N and (b) the connectivity of trimers of PO3N tetrahedra with the local environments of Na, V atoms.
Energies 10 00889 g002
Figure 3. (a) Cyclic voltammograms in the potential range of 3.2–4.5 V at various scan rates and (b) first cycle galvanostatic discharge/charge profile between 3.2 and 4.5 V at a current of 10 mA· g−1 of Na3V(PO3)3N.
Figure 3. (a) Cyclic voltammograms in the potential range of 3.2–4.5 V at various scan rates and (b) first cycle galvanostatic discharge/charge profile between 3.2 and 4.5 V at a current of 10 mA· g−1 of Na3V(PO3)3N.
Energies 10 00889 g003
Figure 4. (a) Typical charge/discharge profile and (b) in situ X-ray diffraction patterns of Na3V(PO3)3N electrode during the first cycle.
Figure 4. (a) Typical charge/discharge profile and (b) in situ X-ray diffraction patterns of Na3V(PO3)3N electrode during the first cycle.
Energies 10 00889 g004
Table 1. Refined Na occupancies, lattice parameter, and cell volume extracted from the Rietveld refinement of diffraction patterns from 10–45° 2θ of Na3V(PO3)3N electrodes at OCV, fully charged and fully discharged state in the first cycle.
Table 1. Refined Na occupancies, lattice parameter, and cell volume extracted from the Rietveld refinement of diffraction patterns from 10–45° 2θ of Na3V(PO3)3N electrodes at OCV, fully charged and fully discharged state in the first cycle.
StateNa1 occNa2 occNa3 occa ( Å)V ( Å3)Rwp/%GOF
OCV1.05 (3)1.00 (2)0.91 (3)9.445 (1)842.7 (2)3.381.14
1st charge: 4.5 V0.58 (4)0.97 (2)0.76 (3)9.438 (1)840.8 (3)3.431.16
1st discharge: 3.2 V0.85 (3)1.01 (2)0.80 (3)9.440 (1)841.4 (2)3.271.10

Share and Cite

MDPI and ACS Style

Zhang, H.; Buchholz, D.; Passerini, S. Synthesis, Structure, and Sodium Mobility of Sodium Vanadium Nitridophosphate: A Zero-Strain and Safe High Voltage Cathode Material for Sodium-Ion Batteries. Energies 2017, 10, 889. https://doi.org/10.3390/en10070889

AMA Style

Zhang H, Buchholz D, Passerini S. Synthesis, Structure, and Sodium Mobility of Sodium Vanadium Nitridophosphate: A Zero-Strain and Safe High Voltage Cathode Material for Sodium-Ion Batteries. Energies. 2017; 10(7):889. https://doi.org/10.3390/en10070889

Chicago/Turabian Style

Zhang, Huang, Daniel Buchholz, and Stefano Passerini. 2017. "Synthesis, Structure, and Sodium Mobility of Sodium Vanadium Nitridophosphate: A Zero-Strain and Safe High Voltage Cathode Material for Sodium-Ion Batteries" Energies 10, no. 7: 889. https://doi.org/10.3390/en10070889

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