Open access peer-reviewed chapter

High-Voltage Cathodes for Na-Ion Batteries: Sodium– Vanadium Fluorophosphates

Written By

Paula Serras, Verónica Palomares and Teófilo Rojo

Submitted: 30 September 2015 Reviewed: 28 January 2016 Published: 01 June 2016

DOI: 10.5772/62317

From the Edited Volume

Alkali-ion Batteries

Edited by Dongfang Yang

Chapter metrics overview

3,035 Chapter Downloads

View Full Metrics

Abstract

This chapter analyses the main advances made in the field of sodium–vanadium fluorophosphates as cathodes for Na-ion batteries and tries to clarify some discrepancies and common errors published about these compounds. The sodium–vanadium fluorophosphate family can be divided in two main members: Na3V2(PO4)2F3 (V+3 extreme phase) and Na3V2O2(PO4)2F (V4+ extreme phase). Na3V2O2x(PO4)2F3-2x, where 0 < x < 1 would correspond to intermediate V3+/4+ mixed valence phases. Among them, the V3+ extreme has demonstrated to be difficult to isolate, whereas the V4+ and mixed valence phases can be more easily prepared by different synthesis methods and from different vanadium sources. In terms of electrochemical performance, mixed valent compound provides good performance, with high specific capacity at moderate/high cycling rates, and long cycle life. The future perspectives for this family of compounds are discussed in terms of raw materials availability, price, and performance relative to other cathode systems for Na-ion batteries.

Keywords

  • Na-ion batteries
  • cathodes
  • fluorophosphates
  • vanadium
  • High voltage

1. Introduction

Energy production is a key issue for the support of our society and way-of-life. Driven by the need to reduce the emissions of CO2 and increase energy security, policy makers have implemented different measures in order to shift to low-carbon energy resources. One of these policies is to promote the use of renewable energy sources at the expense of the carbon-based fuels [1,2]. However, the increasing use of these clean energy sources entails other problems, such as modulating time-variable energy production from renewable resources due to their dependence on the weather to integrate them into the grid. Thus, in order to solve the problem concerning the intermittency of these resources, additional energy storage devices are needed. This way, electrochemical energy storage is a strategic research area that will help to increase the weight of renewable energy sources in the energy mix.

In this field, batteries are one of the most promising systems to act as buffer to regulate the variable energy income that can be obtained from the wind, sun and from the ocean waves. For this purpose, one of the main technologies under development is Na-ion batteries. In the last 20 years, Li-ion batteries have attracted all the attention in energy storage, but recently, Na-based compounds have made a comeback because of controversial debates regarding the size of reserves and higher cost to obtain Li [3]. The use of Na instead of Li in rocking chair batteries could mitigate the feasible shortage of lithium in an economic way for many reasons. The first one, the unlimited sodium sources, being the sodium content in the earth’s crust and water of 28,400 mg·kg−1 and 11,000 mg·L−1 compared to 20 mg·kg−1 and 0.18 mg·L−1 for lithium [4]. Moreover, sodium presents lower price than lithium and it is easy to recover [5].

Although sodium (Na)-ion batteries are considered a promising alternative to lithium-ion systems, it is usually said that they possess limited electrochemical activity compared to Li due to two intrinsic reasons. First, Na has a lower ionization potential than Li (5.139eV vs. 5.392 eV) [6], leading to lower operating voltages and thus lower energy densities. Second, Na+ ions are heavier and larger than Li+ ions, leading to slow diffusion within a solid electrode during cycling and larger volume expansion of the electrode [7]. These two factors would induce lower gravimetric energy densities, so renewed research on these systems has been focused on their utility as buffering systems in the electrical grid, where battery volume is not so crucial due to the stationary nature of the practical use of these batteries.

However, and in spite of the great efforts devoted to the development of higher performance Li-ion batteries, if practical energy densities achieved up till now for both Li- and Na-based systems are compared, it can be seen that Na-ion batteries are catching up with the practical energy densities of Li-ion in a quite short time (Figure 1).

This great advance in gravimetric energy densities for Na-ion batteries is due to the extensive research that has been done in the last years to identify and optimise the best materials for these systems, so they can be soon commercially viable. As it has been said, the incorporation of this kind of systems to the electrical grid would lead to a significant increase of the proportion of renewable energies on the energy mix, with the positive environmental consequences that this fact would result in. Thus, it is a priority to continue searching for and optimising the best-performing materials to be used in the different components of a battery: cathode, electrolyte and anode.

In the case of the cathode, a good performing material should provide high energy density to the battery, that is, it should be able to charge and discharge the highest possible specific capacities with also high operating voltage and for many cycles (long cycle life). In this sense, sodium–vanadium fluorophosphates have demonstrated to fulfil all of these characteristics.

Figure 1.

Theoretical and (estimated) practical energy densities of different rechargeable batteries: Pb–acid (lead acid), NiMH (nickel metal hydride), Na-ion (estimate derived from data for Li-ion assuming a slightly lower cell voltage), Li-ion (average over different types), HT-Na/S8 (high temperature sodium–sulphur battery), Li/S8 and Na/S8 (lithium–sulphur and sodium–sulphur battery assuming Li2S and Na2S as discharge products), Li/O2 and Na/O2 (lithium–oxygen battery, where theoretical values include the weight of oxygen and depend on the stoichiometry of the assumed discharge product, i.e., oxide, peroxide or superoxide). Reproduced from reference [8].

Advertisement

2. High-voltage cathode materials for Na-ion batteries

In the last decade, three dimensional frameworks built on transition metals (M) and polyanions (XO4)n- have become a hot topic in the research field of electroactive materials for lithium and Na-ion batteries. The smaller theoretical gravimetric capacity achieved due to the presence of polyanion groups is compensated by the positive features presented by these compounds, such as very stable frameworks and high inductive effect [9]. The “inductive effect” is a concept proposed by Goodenough et al. consisting on the tune of the redox potential of the 3d metal centre (M) depending on the electronegativity of the X-O bonding [10].

On one hand, sulphate-based cathodes offer a good combination of sustainable syntheses and high energy density owing to their high-voltage operation due to electronegative SO42− units [11]. Na2Fe2(SO4)3 has demonstrated to be the most promising compound being operable at 3.8 V versus sodium and offering a reversible capacity of over 100 mAhg−1 with a 50% capacity retention at high rates (20C) [12]. However, the main issues concerning sulphur-containing cathodes are the low utilisation of active material, volume expansion during extraction/insertion of Na ions and the dissolution of polysulphide into the electrolyte [13].

On the other hand, framework materials based on the phosphate polyanion have also been identified as promising electro-active materials for sodium metal and Na-ion battery applications. It is the strong inductive effect of the PO43− polyanion that moderates the energetic of the transition metal redox couple to generate relatively high operating potentials for these compounds [14]. NaMnPO4 is one of the most studied compounds in this field due to the high operating voltage provided by its lithium analogue: LiMnPO4 which works at 4.1 V versus lithium [15]. However, contrary to LiMnPO4, NaMnPO4 possesses two structure modifications: marycite-type and olivine-type structures being this latter one the electroactive one. Boyadzhieva et al. have recently reported, for the first time, the reversible lithium and Na ion intercalation into the olivine NaMnPO4 phase when testing it versus lithium. However, the electrochemical performance of this material is not yet competitive compared to the well-known LiMnPO4 analogue so that further optimisation is needed [16].

Na4Co3(PO4)2P2O7 pyrophosphate is another promising candidate due to its high working potential region (between 4.1 and 4.7 V versus sodium). In the work presented by Nose et al. [17], this compound offers 95 mAhg−1 reversible capacity with negligible fading after 100 cycles and 84% of capacity retention at rates of 25C. Moreover, the battery presented may be capable of running at even higher potentials (~5.1 V) in order to utilise the Co3+/4+ redox couple leading to the extraction of 4 Na ions from the Na4Co3(PO4)2P2O7 phase. Something similar is proposed by Xu et al. in the case of Na3V2O2(PO4)2F sodium–vanadium fluorophosphate. Due to the difficulty to access the third sodium in the formula unit some theoretical calculations suggested the substitution of oxygen with chlorine to form Na3V2Cl2(PO4)2F which would increase the energy density from 520 to 758 Whkg−1[18]. However, a suitable electrolyte would be needed in both cases in order to work in such high-voltage conditions.

Among the framework materials fluorophosphates possess even higher operating voltages than phosphates, because the inductive effect of fluorine is added to the effect of phosphate. This latter feature makes them a key to solve the energy density issue of sodium-based batteries. Fe, Mn, and V have been the most investigated transition metals but the three of them present different structures: Whereas the sodium–iron fluorophosphate possesses a two-dimensional layered structure [19], the sodium–manganese fluorophosphate presents a three-dimensional tunnel structure [20]. However, manganese compound has demonstrated to be poorly electrochemically active so that more studies have been performed for the iron-based compound. For the moment, one of the best results achieved has been for carbon coated porous hollow spheres of Na2FePO4F phase. This nanostructured material contained about 6–8 wt.% of carbon and presented a specific surface area of 8.7 m2g−1 which led to a specific capacity of 90 mAhg-1 at C/10 and a good cycling stability for 100 cycles [21].

Regarding sodium–vanadium fluorophosphates, three phases have been described in the literature: NaVPO4F, Na3V2O2(PO4)2F and Na3V2(PO4)2F3 with theoretical specific capacity values of 143, 130 and 128 mAh·g−1, respectively. Barker et al. first proposed NaVPO4F as a tetragonal structure with I4/mmm space group [22]. On the other hand, several groups reported also the existence of a NaVPO4F polymorph which they indexed with C2/c space group [2326]. The Na3V2O2(PO4)2F phase has been described as a tetragonal symmetry compound with I4/mmm space group by Sauvage et al. [27] and Massa et al. [28] but, more recently, Tsirlin et al. [29] proposed the P42/mnm space group for the room temperature phase among the diverse polymorphs of this material at different temperatures. The third of the sodium–vanadium fluorophosphates mentioned in the literature, Na3V2(PO4)2F3, is also described as a tetragonal symmetry with P42/mnm space group [3032] but more recently some authors have proposed a different space group for this compound: Amam [33,34].

A deep study of the bibliographic data related to these three sodium fluorophosphates leads to doubt about the real existence of these three different compounds. It is worth noting that the existence of the NaVPO4F phase has already been questioned by authors such as Sauvage et al. [27]. Furthermore, there are no structural data about the suggested monoclinic polymorph of NaVPO4F [2326]. Besides, the diffractogram shown in their works as well as the electrochemical curves obtained match with the one of Na3V2(PO4)3 NASICON compound. In this case, the formation of a NASICON type phase instead of NaVPO4F could be possible because the samples were prepared by the ceramic method and sublimation of VF3 could occur, as it has been described for other lithium–vanadium fluorophosphates [35]. Thus, structural data of only two compounds are found in the literature: Na3V2O2(PO4)2F [29] and Na3V2(PO4)2F3 [30]. The diffractograms of both phases are quite similar that it is possible to establish a relationship between them.

Electrochemical studies on the three mentioned tetragonal structure compounds have been performed under different conditions versus lithium or sodium [2224,27,3639] and in all cases the charge–discharge curves display two voltage plateaux of similar length at the same voltages with little differences depending on the anode material used (ca. 3.6 and 4.1 V versus Na/Na+).

Advertisement

3. Sodium–vanadium fluorophosphates as cathode materials for Na-ion batteries

The study of sodium–vanadium fluorophosphates is especially relevant because of the high operating voltages offered by these compounds that could lead the way to high energy Na-ion batteries.

The similarity of the X-ray diffraction patterns as well as the almost identical electrochemical data for Na3V2O2(PO4)2F and Na3V2(PO4)2F3 suggested that both materials could belong to the same family of compounds, where the fluorine content is modulated by the presence of V3+ and VO2+ (V4+) leading to the following general formula Na3V2O2x(PO4)2F3-2x which was proposed by our group for the first time [40]. The extreme members (x = 0 and x = 1) would correspond to the mentioned phases, Na3V2(PO4)2F3 for x = 0 (Figure 2a) and Na3V2O2(PO4)2F for x = 1 (Figure 2b), whereas intermediate compounds (0 < x < 1) would be V3+/V4+ mixed valence phases (Figure 3). The existence of a mixed valence family of compounds between Na3V2(PO4)2F3 and Na3V2O2(PO4)2F phases was also reported by Park et al. [41] and more recently, by Qi et al. [42].

Figure 2.

Crystallographic structure of a) Na3V2(PO4)2F3 and b) Na3V2O2(PO4)2F. Oxygen in red and fluorine in green. Vanadium octahedron is represented in blue and phosphate tetrahedron in grey.

Figure 3.

Vanadium octahedron for mixed valent Na3V2O2x(PO4)2F3-2x (0 < x < 1) compounds. Vanadium is represented in blue, oxygen in red and fluorine in green.

In order to understand the structure and properties of this family of compounds, a series of sodium–vanadium fluorophosphate samples were hydrothermally prepared, varying the type and amount of carbon used as reductive agent during the synthesis [40]. This way, the possible influence of that carbon on both the final properties of the sample and the attainment of different compositions in the family of general formula Na3V2O2x(PO4)2F3-2x (0 < x < 1) was analysed. Intermediate amounts of carbon present in the Na3V2O2x(PO4)2F3-2x final product (1 – 50 wt.% C) led to V3+/V4+ mixed valence phases (0 < x < 1) while a sample containing almost only V4+ was obtained in a composite with less than 1 wt.% of carbon. The electrochemical study of these samples showed the presence of two reaction voltages at 3.6 and 4.1 V versus Na/Na+ and good specific capacity values up to 100 mAhg−1 at 1C when an electrochemical grade carbon was employed. The material containing a moderate amount of this type of carbon (~6 wt.%) showed the best electrochemical results. A deeper analysis of this sample with an average oxidation state of +3.8 for the vanadium, showed excellent specific capacity values from 100 to 75 mAhg−1 for charging rates of C/20 and 5C, respectively. It is worth mentioning the high capacity retention up to 95% after 200 cycles at 1C [43]. A combined experimental and theoretical study on this family of compounds was also carried out by Park et al. [44,45]. Different samples with compositions x = 0, 0.2, 0.5, 0.8 and 1 were synthesised by solid state reactions. The obtained compounds were characterised by the use of different techniques such as X-ray Diffraction (XRD), Transmission Electron Microscopy (TEM), Nuclear Magnetic Resonance (NMR), etc. The electrochemical study showed specific capacity values of approximately 100 mAhg−1 only in the case of the V4+ composition. The specific capacity value was maintained for all the compounds during 150 cycles at C/2. Recently and in order to achieve different compositions of the family of sodium–vanadium fluorophosphates a new synthesis based on the solvothermal method at low-temperature (60-120°C) has been employed [42]. The size of the particles obtained with this method was in the range of 20–50 nanometers. The electrochemical analysis of these compounds showed a rate capability of 73 mAhg−1 at 10C rate and long cycle stability over 1200 cycles at 2C.

Regarding the V4+ phase, Na3V2O2(PO4)2F, the first electrochemical results were presented by Sauvage et al. getting a capacity of 87 mAhg−1 at C/100 through a two voltage-step behaviour at 3.6 and 4.0 V vs. Na/Na+. They found that the obtained electrochemical curve was almost identical to the ones reported for Na3V2(PO4)2F3 and NaVPO4F phases giving rise to doubt about the real existence of this latter one in light of lack reliable information about its structure [27]. Some years after, a novel single-step hydrothermal treatment was discovered to obtain the Na3V2O2(PO4)2F phase [46]. The absence of in-situ carbon in the synthesised product made it necessary to employ an ex-situ carbon coating in order to improve the conductivity of the material and hence its electrochemical performance. However, control of the thermal treatment process resulted determinant to avoid the degradation of the material and the appearance of secondary phases such as NASICON Na3V2(PO4)3. The carbon coated material prepared by impregnation and a flash thermal treatment, exhibited good rate capabilities between 100 and 65 mAhg−1 for rates of C/20 and 1C respectively and good cycling stability (Figure 4).

Figure 4.

Comparison of discharge specific capacity values between raw Na3V2O2(PO4)2F and carbon coated flash thermally treated Na3V2O2(PO4)2F/C samples.

Additional works showed that the reversible capacity of Na3V2O2(PO4)2F/graphene electrodes exceed 100 mAhg−1 after 200 cycles at a C/20 current rate [37]. Apart from that, a theoretical and experimental study on a solvothermally obtained Na3V2O2(PO4)2F phase was presented were structural and electrochemical characterisation of this material was carried out [18]. Moreover, computational studies revealed the possibility of extracting the third sodium of the mentioned phase at 5.3 V versus sodium. However, this would be experimentally complicated due to the overcoming of the stability window of the electrolyte (4.8 V). However, a hypothetical new vanadium fluorophosphate phase, Na3V2Cl2(PO4)2F, where the oxygen is substituted by chlorine , was proposed from the theoretical study. This latter compound would provide 188 mAh·g−1 theoretical specific capacity and 758 Wh·kg−1 energy density. Finally, electrochemical results of RuO2-coated Na3V2O2(PO4)2F nanowires were presented were a reversible capacity of 120 mAh·g−1 was achieved at 1C rate during more than 1000 cycles [47].

Concerning the V+3 phase, Na3V2(PO4)2F3 , the electrochemical mechanism of the reversible Na extraction was described by Chihara et al. [48] and Shakoor et al. [38]. Both groups reported that the electrochemical reaction occurs through a single-phase reaction with negligible volume variation (~2%). More recently, the ion migration mechanism and diffusion capability of the Na3V2(PO4)2F3 phase have been explored [49,50]. However, the presented results could correspond to any of the sodium–vanadium fluorophosphates included in the general formula Na3V2O2x(PO4)2F3-2x. Moreover, one of the three discharge plateaux that they achieve in the phase (the one at 3.4 V) could correspond to the non fluorinated NASICON Na3V2(PO4)3 phase as recently reported[46]. On the other hand, an ex situ high resolution 23Na and 31P solid state NMR study on this Na3V2(PO4)2F3 phase was reported [51]. It is mentioned that 23Na NMR spectrum of pristine Na3V2(PO4)2F3 phase exhibits two major resonances at 92 and 146 ppm, which are assigned to partially filled Na(2) site and fully occupied Na(1) site, respectively. However, in recent works concerning ssNMR of the mixed valent sodium–vanadium fluorophosphate phase the peaks found in the spectra were assigned to the presence of different vanadium environments instead of the occupancy of the Na cations [44,52]. Similar results were found in the case of the Na3V2O2(PO4)2F phase containing only V4+ . The results of the 23Na ssNMR spectrum obtained in this phase show the presence of a unique peak with an occupancy factor of the two Na cations of 1 for Na(1) site and ½ for Na(2) site (see Figure 5) [53].

Figure 5.

23Na ssNMR spectrum of Na3V2O2(PO4)2F sample (black line). Fittings for Na(1) and Na(2) are shown in green and blue, respectively.

The crystal structure of the Na3V2(PO4)2F3 phase at high temperature (400 K) obtained from high resolution synchrotron radiation measurements in operando revealed the existence of a combined solid solution region and four intermediate phases during the charge process [33,34]. In addition, a new space group (Amam) was achieved for the Na3V2(PO4)2F3 phase . On the other hand, the electrochemical results obtained in a full cell using Na3V2(PO4)2F3 as cathode, hard carbon as anode and an optimised electrolyte showed a capacity of 97 mAh·g−1 after more than 120 cycles at C/5 and 70 mAh·g−1 at 5C rate. [39].

Concerning ex-situ structural study of the aforementioned phases, a detailed study was recently presented [44,45] to determine the sodium extraction/insertion mechanism of the mixed valent fluorophosphate materials. Examination of the ex-situ XRD patterns of V3.8+ Na3V2O1.6(PO4)2F1.4 compound revealed the coexistence of a two-phase region in the first half of the charge indicating the presence of a stable intermediate phase between the two plateaux at 3.6 and 4.1 V versus Na/Na+. It was also reported that the V3+ present in the V3.8+ Na3V2O1.6(PO4)2F1.4 phase creates a local environment similar to the one presented by Na3V2(PO4)2F3 phase [38] leading to a high-voltage V3+/V4+ redox reaction. It is assumed that 2.4 electrons are supplied from the V3.8+ / V5+ redox couple leading to a theoretical specific capacity of 155.6 mAhg−1. On the other hand, X-Ray Absorption Near Edge Spectroscopy (XANES) studies revealed that an oxidation process happens while charging and that the reverse process occurs when discharging but no evolution of the value of the vanadium oxidation state was determined.

Concerning Na3V2O2(PO4)2F phase, different space group symmetries have been described in the literature to define it [28,29]. In terms of sodium insertion/extraction, the I4/mmm is shown to reversibly cycle 1.12 Na ions per formula unit (Na3V2O2(PO4)2F) [27] while the P42/nmn polymorph with a carbon-coated synthetic procedure, reversibly cycles 1.56 Na ions per formula unit [46].

Finally, different time-resolved in-situ synchrotron X-ray powder diffraction studies have been carried out in order to analyse reaction mechanism evolution, lattice parameters ,and sodium extraction of V3.8+( Na3V2O1.6(PO4)2F1.4), V4+ Na3V2O2(PO4)2F and V+3 Na3V2(PO4)2F3 materials [34,54,55].

The electrochemical behaviour of the two electrode materials studied versus Na/Na+ showed similar electrochemical reaction mechanism evolution where a simultaneous solid solution and two-phase region were observed. However, the maximum sodium amount that can be extracted when a potentiostatic step is applied (2 hours at 4.3 V) is higher for V3.8+(Na3V2O1.6(PO4)2F1.4) sample than for the V4+ Na3V2O2(PO4)2F one, 2.23 versus 1.75, respectively. This fact can be due to several reasons, such as better intrinsic conductivity, the presence of in-situ carbon in the mixed valent sample or to the more rigid structure of the only V4+ containing sample. Nevertheless, both materials present high reversibility and flexibility during charge/discharge process. The effects of overcharging up to 4.8 V and cycling-related degradation of the same two compounds, V3.8+( Na3V2O1.6(PO4)2F1.4) and V4+ Na3V2O2(PO4)2F [56] have been recently reported. The V3.8+(Na3V2O1.6(PO4)2F1.4) phase presents less stress than V4+ phase due to the more disordered structure and the presence of in situ carbon. Thus, this phase is able to reach the maximum theoretical Na+ extraction limit when charging up to 4.8 V while in the case of V4+ Na3V2O2(PO4)2F phase this limit is established at the usual 4.3 V cut-off voltage. Moreover, mid-term cycling shows the formation of secondary products due to the degradation of the materials in both cases.

Figure 6.

Potential−composition electrochemical curves obtained upon Na+ extraction from Na3VPF. The first cell was cycled (GITT) between the compositions Na3VPF and Na2VPF while recording 58 5 min diffraction scans; the second one (galvanostatic cycling) between the compositions Na2VPF and Na1VPF while recording 95 5 min scans. Significant single-phase compositions are highlighted by coloured circles. For the first cell they are obtained by the electrochemistry. Cycling of the second cell resulted in Δx > 1 because of electrolyte oxidation, hence in this case reported compositions are those obtained from Rietveld refinement. Top line describes the nature of the observed deintercalation mechanism. Reproduced from reference [34].

To finish with, the structural evolution of V3+ Na3V2(PO4)2F3 phase has recently been analysed by using high angular and intensity resolution synchrotron radiation [34]. The two voltage domains (at 3.7 and 4.2 V versus Na/Na+) have been studied separately. In the case of the low voltage plateau (3.7 V) three biphasic reactions have been observed while in the higher voltage region (4.2 V) a small solid solution zone together with a biphasic zone is shown (Figure 6). Unexpectedly, the structure of the NaV2(PO4)2F3 end member is different from the previous ones revealing a different space group (Cmc21) and two different vanadium environments (V3+ and V5+).

Advertisement

4. Conclusions

Polyanions (XO4)n- based on transition metals (M) are one of the most promising candidates for high-voltage Na-ion batteries. Among them, sulphate and phosphate based framework materials stand out. Regarding phosphate-based materials, and as it has been previously shown in this chapter, there has been some controversy about the structural and electrochemical features of sodium–vanadium fluorophosphates. This family of materials show very high energy density due to the high voltage provided by the adding inductive effect of fluorine. Moreover, these materials have demonstrated to be very stable and present outstanding electrochemical properties. V+3 Na3V2(PO4)2F3, V+3.8 Na3V2O1.6(PO4)2F1.4 and V+4 Na3V2O2(PO4)2F phases included in the family of compounds Na3V2O2x(PO4)2F3-2x have been structurally and electrochemically deeply analysed. Despite the similarity of all these materials, up to date studies show that the electrochemical behaviour of these electrode materials is clearly dependant not only on the sodium extraction/insertion mechanism, the occupancy and distribution of sodium and the electrochemical cycling history but also on the synthesis process employed for the obtaining of the starting material that determines its properties.

Advertisement

5. Future perspectives for sodium–vanadium fluorophosphates as cathodes for Na-ion batteries

As it has been presented, sodium–vanadium fluorophosphates are more than promising cathodic materials for near future commercial Na-ion batteries. However, the appearance of real commercial Na-ion batteries involves the development of the whole battery, that includes advancing in finding the appropriate components for a Na-ion cell based on a sodium–vanadium fluorophosphate cathode. In the first place, an anode matching this cathode must be chosen. This anode material should present relatively low operating voltage in order to provide the battery with the maximum available voltage. In the second place, the specific capacity of this anode and its rate capability should be in balance with the cathode performance in order to make an easier mass balance to prepare the cells. These two conditions will be necessary to assure that the battery, and thus its components, present the best possible performance in terms of battery voltage and power.

On the other hand, a special effort must be done to search for a long-lasting and high-voltage working electrolyte for these systems. Since the recovery of Na-ion battery research, great efforts have been directed to the search of new electrode materials whereas studies dealing with the electrolyte are much scarcer. One of the reasons for this trend can be that it has been shown that the SEI formed on carbonaceous electrodes is markedly different for sodium- and lithium-based electrolytes even using the same solvent [57,58]. Non-aqueous Na electrolytes presently used are mainly based on NaClO4 or NaPF6 salts in propylene carbonate or other solvents and mixtures such as ethylene carbonate: dimethyl carbonate. In general, a good electrolyte should exhibit: i) good ionic conductivity, ii) a large electrochemical window (i.e., high and low onset potential for electrolyte decomposition through oxidation and reduction at high and low voltages, respectively), iii) no reactivity towards the cell components, iv) thermal stability (i.e. melting point and boiling point lower and higher than the standard temperatures for the cell utilization, respectively). Finally, it should be intrinsically safe, have as low toxicity as possible and meet cost requirements for the targeted applications. All these features are intrinsically dependent on the nature of the salt and the solvent(s) and the possible use of additives [59].

Another argument pointed out as a drawback for Na-ion batteries based on sodium–vanadium fluorophosphates is the use of a vanadium compound in them. It is said that vanadium is not a green element that needs careful handling after battery life cycle, but the same can be said from the electrolytes used in both sodium- and lithium-based systems. In any case, a battery will always need to be collected to be treated as a residue after its cycle life, as it is nowadays done. Apart from that, it is also commented that vanadium cost is another drawback for the development of these batteries. But it must be recalled that there exist commercial vanadium-rich energy storage systems, such as redox flow batteries, that comprise the use of high amounts of vanadium for their two electrodes [60]. Thus, a careful cost analysis should be done to confirm or discard this statement once the appropriate anode and electrolyte materials are chosen and a viable cell prototype is proposed.

Finally, and now focusing our attention on the commercial development of only sodium–vanadium fluorophosphates compounds, a last challenge that must be overcome is to develop a synthetic method that allows an easy and economical mass production of these materials. For this purpose, different parameters will have to be taken into account: the different possible synthetic methods (hydrothermal, microwave or ceramic processing); the different possible vanadium starting reactants; and the yields obtained by crossing these latter two parameters (Figure 7).

Figure 7.

Main parameters that must be adjusted to get a commercially viable production method for sodium–vanadium fluorophosphates.

To sum up, sodium–vanadium fluorophosphates family is one of the main possible future cathodic materials for high-voltage Na-ion batteries. Their excellent electrochemical properties and performance have been perfectly well demonstrated by different research groups, as it has been told in this chapter. Now, the great defy for these materials lays on the industrial pace, that is, in the assembly and building of a real Na-ion full cell based on these materials and the development of a commercially viable process to produce these compounds in great amounts.

Advertisement

Acknowledgments

This work was financially supported by the Ministerio de Economía y Competitividad (MAT2013-41128-R) and the Gobierno Vasco/Eusko Jaurlaritza (IT570-13). University of the Basque Country (UPV/EHU) is acknowledged for funding under project UFI11/15.

References

  1. 1. Pavley F, Nunez F. California Assembly Bill No. 32-Global Warming Solutions Act of 2006. 2006.
  2. 2. CARPS, California Codes: Public Utilities Code. Section 399.11-399-31. California Renewables Portfolio Standard Program. 2009.
  3. 3. Ellis B L, Makahnouk W R M, Makimura Y, Toghill K, Nazar L F. A multifunctional 3.5 V iron-based phosphate cathode for rechargeable batteries. Nature Materials. 2007: 6: 749–753. DOI: 10.1038/nmat2007
  4. 4. Holleman A F, Wiberg N. Lehrbuch der Anorganischen Chemie. 102nd ed. Berlin: De Gruyter; 2007. ISBN: 978-3-11-017770-1
  5. 5. Palomares V, Serras P, Villaluenga I, Hueso K B, Carretero-González J, Rojo T. Na-ion batteries, recent advances and present challenges to become low cost energy storage systems. Energy & Environmental Science. 2012: 5:5884–5901. DOI: 10.1039/c2ee02781j
  6. 6. Ong S.P, Chevrier V.L, Hautier G, JainA, Moore C, Kim S, Ma X, Ceder G. Voltage, stability and diffusion barrier differences between sodium-ion and lithium-ion intercalation materials. Energy & Environmental Science. 2011: 4: 3680–3688. DOI: 0.1039/c1ee01782a.
  7. 7. Sawicki M, Shaw L L. Advances and challenges of sodium ion batteries as post lithium ion batteries. RSC Advances. 2015: 5: 53129–53154. DOI: 10.1039/c5ra08321d.
  8. 8. Adelhelm P, Hartmann P, Bender C L, Busche M, Eufinger C, Janek J. From lithium to sodium: cell chemistry of room temperature sodium–air and sodium–sulfur batteries. Beilstein Journal of Nanotechnology. 2015: 6: 1016–1055. DOI: 10.3762/bjnano.6.105
  9. 9. Masquelier C, Croguennec L. Polyanionic (phosphates, silicates, sulfates) frameworks as electrode materials for rechargeable Li (or Na) batteries. Chemical Reviews, 2013: 113: 6552–6591. DOI: 10.1021/cr3001862.
  10. 10. Manthiram A, Goodenough J.B. Lithium insertion into Fe2(SO4)3 frameworks. Journal of Power Sources, 1989: 26: 403–408. DOI: 10.1016/0378-7753(89)80153-3.
  11. 11. Barpanda P. Sulfate Chemistry for high-voltage insertion materials: Synthetic, structural and electrochemical insights. Israel Journal of Chemistry. 2015: 55: 537–557. DOI: 10.1002/ijch.201400157
  12. 12. Barpanda P, Oyama G, Nishimura S.I, Chung S.C., Yamada A. A 3.8-V earth-abundant sodium battery electrode. Nature Communications. 2014:5: 1–8. DOI: 10.1038/ncomms5358.
  13. 13. Sawicki M, Shaw L.L. Advances and challenges of sodium ion batteries as post lithium ion batteries. RSC Advances. 2015:5: 53129–53154. DOI: 10.1039/c5ra08321d.
  14. 14. Yamada A, Chung S.C, Hinokuma K. J. Optimised LiFePO4 for lithium battery cathodes. Electrochemical Society. 2001: 148: A224–A229. DOI: 10.1149/1.1348257.
  15. 15. Lee K.T, Ramesh T.N, Nan F, Botton G, Nazar L.F. Topochemical synthesis of sodium metal phosphate olivines for sodium–-ion batteries. Chemistry of Materials. 2011: 23: 3493–3600. DOI: 10.1021/cm200450y.
  16. 16. Boyadzhieva T, Koleva V, Zhecheva E, Nihtianova D, Mihaylovc L, Stoyanova R. Competitive lithium and sodium intercalation into sodium manganese phospho-olivine NaMnPO4 covered with carbon black. RSC Advances. 2015: 5: 87694–87705. DOI: 10.1039/c5ra17299c.
  17. 17. Nose M, Nakayama H, Nobuhara K, Yamaguchi H, Nakanishi S, Iba H. Na4Co3(PO4)2P2O7: A novel storage material for sodium-ion batteries. Journal of Power Sources. 2013: 234: 175–179. DOI: 10.1016/j.jpowsour.2013.01.162.
  18. 18. Xu M, Xiao P, Stauffer S, Song J, Henkelman G, Goodenough J.B. Theoretical and experimental study of vanadium-based fluorophosphate cathodes for rechargeable batteries. Chemistry of Materials. 2014: 26: 3089–3097. DOI: 10.1021/cm500106w.
  19. 19. 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). Chemistry of Materials. 2010: 22: 1059–1070. DOI: 10.1021/cm902023h.
  20. 20. Yakubovich O.V, Karimova O.V, Mel’nikov O.K. The mixed anionic framework in the structure of Na2MnF(PO4). Acta Crystallographica Section C. 1997: 53: 395–397. DOI: 10.1107/S0108270196014102.
  21. 21. Langrock A, Xu Y, Liu Y, Ehrman S, Manivannan A, Wang C. Carbon coated hollow Na2FePO4F spheres for Na–ion battery cathodes. Journal of Power Sources. 2013: 223: 62–67. DOI: 10.1016/j.jpowsour.2012.09.059.
  22. 22. Barker J, Saidi M.Y, Swoyer J.L. A sodium-ion cell based on the fluorophosphate compound NaVPO4 F. Electrochemical and Solid-State Letters. 2003: 6: A1–A4. DOI: 10.1149/1.1523691.
  23. 23. Zhuo H, Wang X, Tang A, Liu Z, Gamboa S, Sebastian P.J. The preparation of NaV1−xCrxPO4F cathode materials for sodium-ion battery. Journal of Power Sources. 2006: 160: 698–703. DOI: 10.1016/j.jpowsour.2005.12.079.
  24. 24. Liu Z, Wang X, Wang Y, Tang A, Yang S, He L. Preparation of NaV1−xAlxPO4F cathode materials for application of sodium-ion battery. Transactions of Nonferrous Metals Society of China. 2008: 18: 346–350. DOI: 10.1016/S1003-6326(08)60060-6.
  25. 25. Lu Y, Zhang S, Li Y, Xue L, Xu G, Zhang X. Preparation and characterisation of carbon-coated NaVPO4F as cathode material for rechargeable sodium-ion batteries. Journal of Power Sources. 2013: 247: 770–777. DOI: 10.1016/j.jpowsour.2013.09.018.
  26. 26. Xu M, Cheng C.-J, Sun Q.-Q, Bao S.-J, Niu Y.-B, He H, Li Y, Song J. A 3D porous interconnected NaVPO4F/C network: preparation and performance for Na–ion batteries. RSC Advances. 2015;5: 40065–40069. DOI: 10.1039/c5ra05161d.
  27. 27. Sauvage F, Quarez E, Tarascon J, Baudrin E. Crystal structure and electrochemical properties vs. Na+ of the sodium fluorophosphate Na1.5VOPO4F0.5. Solid State Sciences. 2006; 8: 1215–1221. DOI: 10.1016/j.solidstatesciences.2006.05.009.
  28. 28. Massa W, Yakubovich O, Dimitrova O. Crystal structure of a new sodium vanadyl(IV) fluoride phosphate Na3V2O2F[PO4]2. Solid State Sciences. 2002;4: 495–501. DOI: 10.1016/S1293-2558(02)01283-9.
  29. 29. Tsirlin A, Nath R, Abakumov A, Furukawa Y, Johnston D, Hemmida M, Krug von Nidda H.-A, Loidl A, Geibel C, Rosner H. Phase separation and frustrated square lattice magnetism of Na1.5VOPO4F0.5. Physical Review B. 2011;84: 014429/1-014429/16. DOI: 10.1103/PhysRevB.84.014429.
  30. 30. Le Meins J.-M, Crosnier-Lopez M.-P, Hemon-Ribaud A, Courbion G. Phase Transitions in the Na3M2(PO4)2F3 Family (M=Al3+, V3+, Cr3+, Fe3+, Ga3+): Synthesis, Thermal, Structural, and Magnetic Studies. Journal of Solid State Chemistry. 1999;148: 260–277. DOI: 10.1006/jssc.1999.8447.
  31. 31. Barker J, Gover R.K.B, Burns P, Bryan A.J. Hybrid-Ion: A lithium–ion cell based on a sodium insertion material. Electrochemical Solid-State Letters. 2006;9: A190-A192. DOI: 10.1149/1.2168288.
  32. 32. Jiang T, Chen G, Li A, Wang C, Wei Y. Sol–gel preparation and electrochemical properties of Na3V2(PO4)2F3/C composite cathode material for lithium ion batteries. Journal of Alloys and Compounds. 2009;478: 604–607. DOI: 10.1016/j.jallcom.2008.11.147.
  33. 33. Bianchini M, Brisset N, Fauth F, Weill F, Elkaim E, Suard E, Masquelier C, Croguennec L. Na3V2(PO4)2F3 revisited: A high resolution diffraction study. Chemistry of Materials. 2014;26:4238–4247. DOI: 10.1021/cm501644g.
  34. 34. Bianchini M, Fauth F, Brisset N, Weill F, Suard E, Masquelier C, Croguennec L. Comprehensive investigation of the Na3V2(PO4)2F3 - NaV2(PO4)2F3 system by operand high resolution synchrotron X-ray diffraction. Chemistry of Materials. 2015;27:3009–3020. DOI: 10.1021/acs.chemmater.5b00361.
  35. 35. Ateba J.-M, Masquelier C, Suard E, Croguennec L. Synthesis and crystallographic study of homeotypic LiVPO4F and LiVPO4O. Chemistry of Materials. 2012;24: 1223–1234. DOI: 10.1021/cm3003996.
  36. 36. Ruan Y.-L, Wang K, Song S.-D, Han X, Cheng B.W. Graphene modified sodium vanadium fluorophosphate as high voltage cathode material for sodium ion batteries. Electrochimica Acta. 2015;160:330-336. DOI: 10.1016/j.electacta.2015.01.186
  37. 37. Xu M, Wang L, Zhao X, Song J, Xie H, Lu Y, Goodenough J.B. Na3V2O2(PO4)2F/graphene sandwich structure for high-performance cathode of a sodium-ion battery. Physical Chemistry Chemical Physics. 2013;15: 13032–13037. DOI: 10.1039/C3CP52408F.
  38. 38. Shakoor R.A, Seo D.-H, Kim H, Park Y.U, Kim J, Kim S.-W, Gwon H, Lee S, Kang K. A combined first principles and experimental study on Na3V2(PO4)2F3 for rechargeable Na batteries. Journal of Materials Chemistry. 2012;22: 20535–20541. DOI: 10.1039/C2JM33862A.
  39. 39. Ponrouch A, Dedryvère R, Monti D, Demet A.E, Ateba J.M, Croguennec L, Masquelier C, Johansson P, Palacín M.R. Towards high energy density sodium ion batteries through electrolyte optimisation. Energy and Environmental Science. 2013;6: 2361–2369. DOI: 10.1039/C3EE41379A.
  40. 40. Serras P, Palomares V, Goñi A, Gil de Muro I, Kubiak P, Lezama L, Rojo T. High voltage cathode materials for Na–ion batteries of general formula Na3V2O2x(PO4)2F3-2x. Journal of Materials Chemistry. 2012;22: 22301–22308. DOI: 10.1039/C2JM35293A.
  41. 41. Park Y.U, Seo D.H, Kim B, Hon K.P, Kim H, Lee S, Shakoor R.A, Miyasaka K, Tarascon J.M, Kang K. Tailoring a fluorophosphate as a novel 4 V cathode for lithium-ion batteries. Scientific Reports. 2012;2: 704. DOI: 10.1038/srep00704.
  42. 42. Qi Y, Mu L, Zhao J, Hu Y.-S, Liu H, Dai S. Superior Na-storage performance of low-temperature-synthesized Na3(VO1-xPO4)2F1+2x (0 ≤ x ≤ 1) nanoparticles for Na–ion batteries. Angewandte Chemie International Edition. 2015;54: 9911–9916. DOI: 10.1002/anie.201503188.
  43. 43. Serras P, Palomares V, Goñi A, Kubiak P, Rojo T. Electrochemical performance of mixed valence Na3V2O2x(PO4)2F3-2x/C as cathode for sodium-ion batteries. Journal of Power Sources. 2013;241: 56–60. DOI: 10.1016/j.jpowsour.2013.04.094.
  44. 44. 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. Journal of the American Chemical Society. 2013;135: 13870–13878. DOI: 10.1021/ja406016j.
  45. 45. Park Y. U, Seo D. H, Kim B, Kim J, Lee S, Kim B, Kang K. A family of high-performance cathode materials for Na–ion batteries, Na3(VO1−xPO4)2 F1+2x (0 ≤ x ≤ 1): combined first-principles and experimental study. Advanced Functional Materials. 2014;24: 4603–4614. DOI: 10.1002/adfm.201400561.
  46. 46. Serras P, Palomares V, Kubiak P, Lezama L, Rojo T. Enhanced electrochemical performance of vanadyl (IV) Na3(VO)2(PO4)2F by ex-situ carbon coating. Electrochemistry Communications. 2013;34: 344–347. DOI: 10.1016/j.elecom.2013.07.010.
  47. 47. Peng M, Li B, Yan H, Zhang D, Wang X, Xia D, Guo G. Ruthenium-oxide-coated sodium vanadium fluorophosphate nanowires as high power cathode materials for sodium ion batteries. Angewandte Chemie International Edition. 2015;54: 6452–6456. DOI: 10.1002/anie.201411917.
  48. 48. 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. DOI: 10.1016/j.jpowsour.2012.10.034.
  49. 49. Song W, Ji X, Wu Z, Yang Y, Zhou Z, Li F, Chen Q, Banks C.E. Exploration of ion migration mechanism and diffusion capability for Na3V2(PO4)2F3 cathode utilized in rechargeable sodium ion batteries. Journal of Power Sources. 2014;256:258–263. DOI: 10.1016/j.jpowsour.2014.01.025.
  50. 50. Song W, Ji X, Wu Z, Zhu Y, Li F, Yao Y, Banks C.E. Multifunctional dual Na3V2(PO4)2F3 cathode for both lithium-ion and sodium-ion batteries. RSC Advances. 2014;4:11375–11383. DOI: 10.1039/c3ra47878e.
  51. 51. 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. Chemistry of Materials. 2014;26:2513–2521. DOI: 10.1021/cm403728w.
  52. 52. Serras P, Palomares V, Alonso J, Sharma N, Lopez del Amo J.M, Kubiak P, Fdez-Gubieda M.L, Rojo T. Electrochemical Na extraction/insertion of Na3V2O2x(PO4)2F3−2x. Chemistry of Materials. 2013;25:4917–4925. DOI: 10.1021/cm403679b.
  53. 53. Serras Malillos P. High voltage cathodes for Na–ion batteries: Na3V2O2x(PO4)2F3-2x system [thesis]. Leioa: University of the Basque Country; 2014.
  54. 54. Serras P, Palomares V, Rojo T, Brand H.E.A, Sharma N. Structural evolution of high energy density V3+/V4+ mixed valent Na3V2O2x(PO4)2F3−2x (x = 0.8) sodium vanadium fluorophosphate using in situ synchrotron X-ray powder diffraction. Journal of Materials Chemistry A. 2014;2:7766–7779. DOI: 10.1039/c4ta00773e.
  55. 55. Sharma N, Serras P, Palomares V, Brand H.E.A, Alonso J, Kubiak P, Fdez-Gubieda M.L, Rojo T. Sodium distribution and reaction mechanisms of a Na3V2O2(PO4)2F electrode during use in a sodium-ion battery. Chemistry of Materials. 2014;26:3391–3402. DOI: 10.1021/cm5005104.
  56. 56. Palomares V, Serras P, Brand H.E.A, Rojo T, Sharma N. Structural evolution of mixed valent (V3+/V4+) and V4+ sodium vanadium fluorophosphates as cathodes in sodium-ion batteries: comparisons, overcharging and mid-term cycling. Journal of Materials Chemistry A. 2015;3:23017–23027. DOI: 10.1039/c5ta03780h.
  57. 57. Komaba S, Murata W, Ishikawa T, Yabuuchi N, Ozeki T, Nakayama T, Ogata A, Gotoh K, Fujiwara K. Electrochemical Na insertion and solid electrolyte interphase for hard-carbon electrodes and application to Na–Ion batteries. Advanced Functional Materials. 2011;21: 3859–3867. DOI: 10.1002/adfm.201100854.
  58. 58. Xia X, Obrovac M N, Dahn J R. Comparison of the reactivity of NaxC6 and LixC6 with non-aqueous solvents and electrolytes. Electrochemical and Solid State Letters. 2011;14: A130-A133. DOI.: 10.1149/1.3606364.
  59. 59. Ponrouch A, Marchante E, Courty M, Tarascon J-M, Palacín M R. In search of an optimised electrolyte for Na–ion batteries. Energy & Environmental Science. 2012;5: 8572–8583. DOI: 10.1039/C2EE22258B.
  60. 60. Alotto P G, Guarnieri M, Moro F. Redox flow batteries for the storage of renewable energy: A review. Renewable and Sustainable Energy Reviews. 2014;29: 325–335. DOI: 10.1016/j.rser.2013.08.001.

Written By

Paula Serras, Verónica Palomares and Teófilo Rojo

Submitted: 30 September 2015 Reviewed: 28 January 2016 Published: 01 June 2016