Brought to you by:

Electrochemical Redox Processes Involved in Carbon-Coated KVPO4F for High Voltage K-Ion Batteries Revealed by XPS Analysis

, , , , , and

Published 1 October 2020 © 2020 The Electrochemical Society ("ECS"). Published on behalf of ECS by IOP Publishing Limited
, , Citation Laure Caracciolo et al 2020 J. Electrochem. Soc. 167 130527 DOI 10.1149/1945-7111/abbb0c

1945-7111/167/13/130527

Abstract

K-ion batteries need efficient positive electrode materials with open structural frameworks to accommodate the large ionic radius of K+. In that direction, polyanionic compounds are of great interest. Among them, KVPO4F is the most studied one. Its electrochemical redox processes still remain unclear, especially at high voltage. To tackle this issue, X-ray photoelectron spectroscopy was used to reveal the electrochemical redox processes of KxVPO4F. First, a carbon coating was performed and allowed increasing the overall electrochemical performance while mitigating electrolyte degradation at high voltage. Then, XPS analysis showed a high reversibility of the redox processes although the K+ extraction (and insertion) from x = 0.5 to 0 was hindered, possibly by structural constraints while electrolyte degradation occurred mostly above 4.5 V.

Export citation and abstract BibTeX RIS

The development of new electrochemical energy storage systems as alternatives to the Li-ion technology remains a key challenge due to the growth of energy storage demand. Alternative batteries based on abundant and thus low cost raw materials are required. Na-ion and K-ion batteries have attracted increasing interest during the last decade. Thanks to the low potential of the K+/K redox couple vs the other alkali metals in non-aqueous solvents, high voltage K-ion batteries (KIBs) could be developed.1 Moreover, K+ has the lowest Lewis acidity and desolvation energy compared to Na+ and Li+,2 and thus higher ionic conductivity and faster electrode/electrolyte interface diffusion kinetics that would allow developing high power KIBs.

The high reactivity of K metal requires considering results obtained in half-cells with caution, especially the interaction between the electrolyte and the electrode surface. Indeed, migration of electrolyte degradation products can occur from the K metal to the working electrode.3 Among negative electrode materials,4 graphite is of great interest as it reversibly intercalates K+ to form KC8.57 Promising electrochemical performance is obtained, with a reversible capacity of ∼250 mAh.g−1 up to hundreds of cycles and even for high practical loading (>20 mg cm−2).713 Regarding the safety issue, the average discharge potential of ∼0.3 V vs K+/K makes graphite even more attractive for the application.

In fact, the major drawback of K-ion batteries is the need of positive electrode materials with open structural frameworks to accommodate the large ionic radius of K+ (1.38 Å) compared to Na+ (1.02 Å) and Li+ (0.76 Å),1,14 which is detrimental to the volumetric energy density.15 Polyanionic compounds are of great interest as they are suitable host structures for the reversible K+ intercalation as well as to deliver high voltage.16 For instance, KVPO4F (KVPF) showed reversible capacity up to ∼105 mAh.g−1 with an average discharge potential of ∼4.3 V vs K+/K and excellent rate performance.1722 In a full-cell configuration using graphite, the energy density has even been evaluated at 300 Wh.kg−1, suggesting a promising material for further full-cell studies.23 Regarding its electrochemical behaviour as a positive electrode material, KVPF shows 4 distinct potential domains in both charge and discharge, corresponding to successive phase transitions with an interesting limited lattice volume change of ∼6.5%,19 and the formation of intermediate phases KxVPO4F at x = 0.75, x = 0.625 and x = 0.5. The reaction occurring at high voltage, between the composition K0.5VPO4F and VPO4F still remains unclear, however, with simultaneous active material redox reaction and large electrolyte degradation. It may explain why the theoretical capacity of KVPF (131 mAh.g−1) was never reached so far. Therefore, it is of great interest to better understand the electrochemical processes involved in KVPF positive electrode materials in KIBs in order to improve its performance.

The present work proposes to investigate the electrochemical redox processes involved in KVPF upon cycling in K metal batteries using X-ray photoelectron spectroscopy (XPS). The impact of a carbon-coating on the electrochemical performance of KVPF was first evaluated, regarding its ability to improve the overall performance and mitigate the electrolyte degradation.

Experimental

KVPO4F and carbon coated KVPO4F (KVPO4F-C) were prepared by a two-step reaction. First, VPO4 was obtained by mixing V2O5 (from Sigma-Aldrich, 99.6%) and NH4H2PO4 (from Sigma-Aldrich, 99.99%) by high energy ball-milling (SPEX 8000 M mixer/mill at 1425 rpm) for 1 h followed by a thermal treatment at 800 °C for 7 h under Ar/H2 (5%). Second, stoichiometric amounts of KF.2H2O (from Alfa Aesar, 98.5%, dried overnight under vacuum at 250 °C and stored in a glovebox under Ar) and VPO4 were mixed without (KVPF) or with (KVPF-C) the addition of 10 wt.% sucrose (from Sigma-Aldrich, ACS reagent). The mixture was obtained by SPEX ball-milling for 1 h, pelletized and thermal treated at 650 °C for 8 h under Ar. Rietveld refinement of the X-ray diffraction patterns was performed using the FullProf software and considering an orthorhombic unit cell described in the Pna21 space group.24 It confirmed the nature of the grey (KVPF) and black (KVPF-C) powders obtained (Fig. S1 is available online at stacks.iop.org/JES/167/130527/mmedia). The smaller cell parameters and especially volume cell obtained for KVPO4F-C vs the KVPO4F phase reported and characterized in-depth by Fedotov et al.17 suggest a partial O2− substitution for F and thus the formation of KVPO4F1−yOy. y was estimated to 0.1 from the study of the solid solution KVPO4F - KVPO4O (Table SI, and from 31P MAS NMR data not reported here). Note that the stoichiometry in K/V/P was determined by chemical analyses (Fig. S1). In addition, thermogravimetric analyses and scanning electron microscopy images confirmed the formation of a carbon coating (2.5 wt.% of C) for KVPF-C compared to KVPF (Fig. S2). Electrodes were prepared by mixing KVPF or KVPF-C, carbon black (Super P, BET = 62 m2 g−1, Alfa Aesar) and poly(vinylidene fluoride) according to a 70:25:5 weight ratio, in N-methyl-2-pyrrolidone (32:68 solvent:material weight ratio), by ball-milling for 1 h. The obtained slurry was cast on an aluminium current collector, dried for 24 h under Ar, and finally electrodes of 1.5 mgKVPO4F/cm2 were punch out and dried under vacuum at 80 °C for 12 h. 2032 coin-type cells (316L stainless steel) were assembled under Ar using a Whatman (grade GF/D) and a polypropylene membrane (Celgard) as separators, K metal (99.95%, Alfa Aesar) as negative electrode and 100 μl of 0.8 M KPF6 in a mixture of EC:DEC by volume as electrolyte (KPF6, Sigma-Aldrich, ≥99%; EC, Sigma-Aldrich, anhydrous, 99%; DEC, Sigma-Aldrich, anhydrous, ≥99%). All potentials reported in the text are referred to K+/K. Note that K metal polarization in symmetrical K//K cell was found negligible during open circuit voltage (<6 mV, Fig. S3) compared to the literature.25 During plating/stripping at 0.04 mA cm−2 (corresponding to the current used to cycle K//KVPF cells), however, a polarization of 0.1 V was measured. Cycling tests were performed at C/5 (26 mA.g−1, i.e. the exchange of 1 K+ in 5 h) and 20 °C using a VMP3 (Biologic, SAS France). Two formation cycles were first performed between 3.5–5 V vs K+/K to minimize electrolyte degradation impact (Figs. S4 and 3), as discussed later. KVPF-C cells were then stopped at different potentials during the 3rd cycle and electrodes were recovered from the cells and washed twice with 1 ml of DEC during 30 s. XPS analysis was performed using an Escalab 250 Xi spectrometer, with a monochromatized Al Kα radiation (hν = 1486.6 eV). Electrodes were placed on a sample holder using uPVC insulation tape (3 M part number 655), and transferred to Ar filled glove box connected to the spectrometer. Using the standard charge compensation mode, core spectra were recorded with a 0.15 eV step and a constant 20 eV pass energy. Using CasaXPS software, the binding energy scale was calibrated from the KVPF oxygen peak at 530.0 eV. It is reminded that XPS probes about 5 to 10 nm depth of the sample surface, and here between 6–7 nm for the vanadium, i.e. not only the extreme surface atomic layers (0.5–1 nm) but also the deeper ones, which means a representative part of the bulk is also probed (Fig. S5). Moreover, at the extreme surface, the presence of defects may lead to vanadium with an oxidation state higher or lower than in the bulk. V 2p3/2 core spectra of KVPF electrodes were thus fitted using peak positions and line shapes constraints corresponding to different oxidation states as determined from reference samples with Vn+ in a PO4 environment (Fig. S5 and Table SII) in agreement with the literature.2630 This thorough procedure allows reliably following the evolution of the mean oxidation state upon cycling, especially considering that the mean oxidation state obtained by XPS (3.1) for the pristine KVPF-C material is in good agreement with the XRD analysis (KVPO4F1−yOy with y = 0.1).

Results and Discussion

The impact of the carbon coating was evaluated regarding its ability to optimize the electrochemical performance of the material in KIBs and to mitigate the KVPF active material surface reactivity against the electrolyte. Overall, both KVPF and KVPF-C showed a similar charge/discharge profile at C/5 with 4 potential domains (Fig. S4), in very good agreement with the literature.19 However KVPF-C showed a lower average polarization of ∼230 mV compared to ∼300 mV for KVPF (Fig. S4) and a better capacity retention after 70 cycles compared to KVPF (76% vs 62%, (Fig. S6). For both systems, a significant high-potential hysteresis is observed, suggesting that the main capacity loss would be related to the high-potential domain. Since KVPF and KVPF-C have close particles size (Fig. S2), the carbon coating was shown—as expected—to promote better transport properties within the positive electrode, lower polarization of the cell, smaller electrolyte degradation at high voltage and thus optimized performances. XPS analysis confirmed a significantly lower solid electrolyte interphase coverage at the KVPF-C surface compared to KVFP after the first cycle (Table SIII). Note that the lower electrolyte degradation is explained by a passivation effect of the carbon coating at the KVPF surface. Based on these results, KVPF-C was therefore selected to study the electrochemical redox processes involved in KVPO4F upon cycling.

Figure 1a shows the typical 3rd charge/discharge cycle obtained for K//KVPF-C cells and indicates at which potentials the prepared cells were stopped to perform the analysis of the redox processes involved in KVPF-C. Note that charge/discharge profiles obtained for all the cells showed very close polarization and capacities (Fig. S7), highlighting the great reproducibility of the electrochemical results so that reliable XPS analysis is expected on materials recovered ex situ from this series of cells. Figure 1b shows the V 2p3/2 XPS core spectra of KxVn+PO4F-C electrodes as recovered from K//KVPF-C cells during the third cycle at C/5, as function of the K content. For more information, the full V 2p core spectra can be found in Fig. S8. XPS analysis revealed that the pristine KVPF-C showed the presence of V2+, V3+ and V4+ (Fig. 1b) although only V3+ and V4+ were expected based on the KVPO4F0.9O0.1-C formula. Note that V2+ was also observed for the VPO4 precursor (Fig. S5). However, considering the very low amount (<1wt.%) of VPO4 impurity in the KVPF-C material (Fig. S1), the presence of V2+ is likely to originate from the KVPF-C extreme surface. This was supported by the decrease of the V2+ contribution during oxidation as discussed thereafter.

Figure 1.

Figure 1. (a) Typical 3rd charge/discharge cycle (at C/5, 26 mA.g−1) obtained for K//KVPF-C cells indicating at which potentials cells were stopped for the analysis of the redox processes involved in KVPF-C. (b) Vanadium 2p3/2 XPS core spectra of KxVn+PO4F0.9O0.1-C electrodes as recovered from K//KVPF-C cells during the 3rd cycle, as function of the potential.

Standard image High-resolution image

After charge of the cell to 4.10 V, i.e. during the 1st oxidative potential domain, the peak associated with V4+ increased while the V3+ and V2+ ones decreased (respectively +7%, −4% and −3%). At almost the end of the 2nd potential domain at 4.30 V (i.e. at the composition of about K0.625VPO4F), the peak associated with V4+ increased while the peaks associated with V3+ and V2+ decreased (respectively +15%, −12% and −3%). At the end of the 3rd potential domain at 4.46 V (i.e. for approx. K0.5VPO4F), a new peak, attributed to V5+ appeared (+13%) while the V4+, V3+ and V2+ peaks decreased (−11%, −1% and −1%, respectively). This phenomenon further continued after the 4th potential domain at 5.00 V, so that the V5+, V4+, V3+ and V2+ peaks represent 20%, 50%, 23% and 7%, respectively. Note that despite the remaining V3+ and V2+ even at the end of charge, the corresponding mean oxidation states were in good agreement with the theoretical/electrochemical ones, as discussed thereafter. Note that remaining of V3+ and V2+ at the end of charge should arise from specific vanadium environment at the KVPF surface. During discharge, opposite phenomena were observed (Fig. 1b), indicating a high reversibility of the redox processes involved in KVPF-C upon cycling in the potential window 3.5–5 V vs K+/K.

The evolution of the vanadium mean oxidation state as the K content changed during the 3rd cycle for KVPO4F0.9-O0.1C and determined from both the charge/discharge capacity (DOelec) and XPS analysis (DOxps) is reported in Fig. 2.

Figure 2.

Figure 2. Vanadium mean oxidation state (DO) of KVPO4F0.9-O0.1-C, as the K content changed during the 3rd cycle, determined from the charge/discharge capacity on one side and from the analysis of the XPS spectra on the other side. Grey lines represent the theoretical DO based on the KVPO4F0.9-O0.1-C formula.

Standard image High-resolution image

Until the end of the 3rd potential domain, i.e. at 4.46 V, a very good agreement was observed between the vanadium mean oxidation states determined from the electrochemistry and from XPS analysis. On the contrary, at the end of the 4th potential domain, i.e. at 5.00 V, DOelec was +3.9 while DOxps was only +3.8. This discrepancy indicates that about 25% of the electrons exchanged in the last potential domain are not related to K+ deintercalation from KVPF-C but to electrolyte degradation. This is in agreement with the high irreversible capacity observed for the 4th potential domain between the charge and the next discharge (Fig. 1a). Moreover, the evolution of the capacities associated with the different potential domains for the first 20 cycles showed that electrolyte degradation is mainly associated with the 4th potential domain as shown in Fig. 3 by the large difference in height between the full and hatched black bars. Nevertheless, during the first two cycles, irreversibility was observed throughout the reaction, the latter being much more pronounced at high voltage (about 60%, Fig. 3). During the next cycles, a stabilization was observed with an excellent efficiency below 4.6 V (same height for the full and hatched blue and green bars), and thus a capacity loss almost exclusively associated with the high voltage domain (about 90%) and the composition range K0.5VPO4F-C—VPO4F-C. This highlights the relevance of using the 3rd cycle to perform the redox processes study instead of the 1st cycle. During discharge, DOxps decreased back to +3.1 (Fig. 2), indicating a complete reversibility of the redox processes. This result fully supports the limitation of the participation of the electrolyte degradation to high voltage, only after the 2nd cycle as shown in Fig. 3.

Figure 3.

Figure 3. Capacity (mAh.g−1) related to the 1st, 2nd, 3rd and 4th redox processes as obtained from the derivative of the galvanostatic charge/discharge curves. Note that the 2nd and 3rd potential domainss were combined as they cannot be perfectly dissociated after few cycles (not shown). Corresponding capacity loss (%) between oxidation and reduction is also reported by the curves with the red points. Theoretical capacity of the different redox processes based on the KVPO4F0.9-O0.1C formula (118 mAh.g−1) are also reported for comparison.

Standard image High-resolution image

Importantly, only 0.7 K+ were extracted from KVPO4F0.9-O0.1-C (Fig. 2) instead of 0.9 theoretically. Especially, the capacity associated with the 1st, 2nd and 3rd potential domains nearly reached the theoretical ones and remained almost constant from cycles 3 to 20 (Fig. 3). On the contrary, the 4th potential domain capacity continuously decreased to about half the theoretical one from cycle 3 to 20 (Fig. 3). Interestingly, no polarization increase (due to the accumulation of electrolyte degradation products) was observed (Fig. S9). Thus, the 4th potential domain capacity decrease is not explained by the cut-off potential being reached before the end of the domain. Instead, these results likely originate from structural constraints hindering the extraction (and insertion) of K+ from x = 0.5 to x = 0 as proposed previously.19,31 To evaluate the possibility to further extract the remaining K+, a cell was hold at 5 V for 18 h during the 3rd cycle and the corresponding DOxps was +3.9 (Fig. S10), in agreement with the evolution of the cell parameters determined by XRD (not shown). This result highlights that despite the electrolyte degradation process, it is still possible to extract the potassium but the process is much slower, more likely due to a structural constraint such as cell volume reduction. This structural constraint could increase the K+ migration barrier, inducing a decrease in the K+ diffusion coefficient,31 slowing the global extraction process.

Conclusions

In conclusion, the carbon coating of the KVPF material was found beneficial to the overall electrochemical performance while mitigating electrolyte degradation at high voltage. Then, by comparing the average oxidation state derived from electrochemical and XPS analysis, the electrochemical redox processes of KxVPO4.1F0.9-C were revealed. Interestingly, a high reversibility was found although the K+ extraction (and insertion) from x = 0.5 to x = 0 was hindered, possibly by structural constraints. Moreover, the severe electrolyte degradation issue, observed mostly above 4.5 V, is currently under further investigation, as well as possible KVPF dissolution in order to successfully mitigate these irreversible degradation reactions using electrolyte additives and enable highly efficient KVPF based full K-ion cells.

Acknowledgments

The authors thank the financial support of E2S-UPPA, Région Nouvelle Aquitaine and the French National Research Agency (STORE-EX Labex Project ANR-10-LABX-76-01 and TROPIC project ANR-19-CE05-0026-01). The authors thank Cathy DENAGE, Etienne DURAND and Eric LEBRAUD (ICMCB) for their help with SEM, TGA-FTIR and XRD analyses, respectively. Authors also thank Joël GAUBICHER (IMN-Nantes) for fruitful discussions about redox processes in KVPF.

Please wait… references are loading.
10.1149/1945-7111/abbb0c