Kohei
Shibuya
ab,
Kazuaki
Kisu
*c,
Takara
Shinohara
ab,
Kosuke
Ishibashi
d,
Hiroshi
Yabu
d and
Shin-ichi
Orimo
*ad
aInstitute for Materials Research (IMR), Tohoku University, Katahira 2-1-1, Aoba-ku, Sendai 980-8577, Japan
bIchikawa Research Centre, Sumitomo Metal Mining Co. Ltd., Nakakokubun 3-18-5, Ichikawa, Chiba 272-8588, Japan
cCollege of Engineering, Shibaura Institute of Technology, Toyosu 3-7-5 Koto-ku, Tokyo 135-8548, Japan. E-mail: kkisu@shibaura-it.ac.jp
dAdvanced Institute for Materials Research (AIMR), Tohoku University, Katahira 2-1-1, Aoba-ku, Sendai 980-8577, Japan. E-mail: shin-ichi.orimo.a6@tohoku.ac.jp
First published on 12th May 2025
Ca-metal batteries have emerged as promising candidates for post-Li-ion batteries owing to their high energy density, cost-effectiveness, and natural abundance. However, few cathode materials have been evaluated for full-cell configurations with Ca-metal anodes, and most of these materials exhibit large voltage hysteresis and sloping voltage profiles. In this study, we propose polyanionic Na superionic conduction (NASICON)-type NaTi2(PO4)3 (NTP) as a cathode material for Ca-metal batteries, leveraging its relatively large lattice size among NASICON compounds to accelerate Ca2+ diffusion. Nanoparticulate NTP/carbon nanotube (CNT) composites were synthesised via solvothermal synthesis. The composite structure consisted of NTP nanoparticles uniformly dispersed within a conductive CNT network, which provided high electronic conductivity and short Ca2+ diffusion pathways, enabling fast electrochemical reactions. By combining NTP cathodes with a Ca-metal anode and a hydride-based electrolyte, i.e. Ca(CB11H12)2 in 1,2-dimethoxyethane/tetrahydrofuran, capable of supporting reversible Ca plating/stripping at room temperature, we fabricated Ca-metal batteries that achieved a reversible Ca storage capacity of 40.6 mA h g−1 at 5 mA g−1. Notably, these cells exhibited small voltage hysteresis and flat voltage profiles, indicating rapid Ca2+ diffusion. Ex situ analysis, including X-ray diffraction and X-ray photoelectron spectroscopy, confirmed the reversible insertion and extraction of Ca2+ within the NTP framework. This study provides valuable insights into the design of practical Ca cathodes with enhanced ion-transport properties, paving the way for the development of high-performance Ca-metal batteries.
A model cathode plays a critical role in the advancement of battery technology, as exemplified by Mo6S8 in Mg-ion batteries, which has served as a benchmark for understanding Mg2+ insertion/extraction.8–10 To serve as an effective model cathode for Ca-metal batteries, a material must meet several key criteria, including a simple reaction mechanism, a flat voltage plateau, low hysteresis, high structural stability, and an appropriate operating voltage. However, no reported Ca-metal battery cathode meets these requirements, presenting a major obstacle to further research and development.
Research on cathodes using full-cell configurations remains in its early stages due to the limited availability of electrolytes compatible with Ca plating/stripping, restricting cathode evaluations primarily to half-cell studies with activated-carbon electrodes as substitutes for Ca metal. While half-cell measurements provide insight into intrinsic properties of cathode materials,11–17 they do not fully capture the electrochemical complexities of practical full-cell operation. Consequently, many materials that exhibit reversible Ca2+ insertion/extraction in half-cells demonstrate poor or non-reversible performance in full-cell tests,18–20 underscoring the necessity of full-cell evaluations.
Recent advancements in electrolytes such as Ca(BH4)2,21,22 Ca[B(hfip)4]2,23,24 and Ca(CB11H12)2 (ref. 25 and 26) have enabled full-cell testing, shifting research beyond half-cell screening. Various candidates, including sulfides (Sulfur (S),27,28 CuS,29,30 and VS4 (ref. 31)), Se,32 C-based materials (graphite33 and freestanding lattice-expanded graphitic carbon fiber (FLEGCF)34), organic compounds (1,4-polyanthraquinone (14PAQ),35 poly(anthraquinonyl sulfide) (PAQS),36 polytriphenylamine/graphene nanoplates (PTGNP),37 and 3,4,9,10-perylenetetracarboxylic dianhydride (PTCDA)38), metal oxides (CaV6O16·2.8H2O (CVO)18), and fluorophosphates (NaV2PO4.8F0.7,19 NaV2(PO4)2F3,20, (NVPF) and KxVPO4F (x ∼ 0) (K0VPF)39) have been explored. However, each of these materials faces fundamental limitations. Sulfides suffer from low cycling efficiency due to conversion-type reactions. Se undergoes dissolution and pulverisation, leading to rapid capacity fade. Carbon-based materials either operate at low voltages (e.g., graphite) or suffer from high polarisation shifts (e.g., FLEGCF). Organic compounds, while offering good cycling stability, dissolve in electrolytes, causing performance degradation. Fluorophosphates, despite their high voltage operation, exhibit large voltage hysteresis and sloping voltage profiles, which reduce energy efficiency. Given these limitations, the lack of a model cathode remains a critical bottleneck in the field.
Among polyanionic compounds, NASICON-type materials have gained attention due to their three-dimensional interconnected channels, which enhance Ca2+ mobility. While fluorophosphates offer the advantage of high voltage,40 NASICON-type polyanionic compounds have higher ionic conductivity.41 NaV2(PO4)3 (NVP) has been extensively investigated as a potential cathode for Ca-ion batteries due to its structural stability and ability to accommodate multivalent ions. NVP has demonstrated reversible Ca2+ insertion/extraction in half-cells, confirming the viability of NASICON-type frameworks for Ca2+ storage. However, in full-cell configurations, NVP exhibits substantial voltage hysteresis and a sloping voltage profile, indicating significant kinetic limitations in Ca2+ mobility. This suggests that while the NASICON framework itself is promising, improving ion diffusion within the structure is critical for practical application.
Building upon this insight, we propose NaTi2(PO4)3 (NTP) as an improved NASICON-type cathode for Ca-metal batteries. Compared to NVP, NTP features a larger lattice size, which provides superior rate performance in Na-ion batteries42 and is expected to enhance Ca2+ diffusivity, thereby mitigating the kinetic limitations observed in NVP. This fast-ion diffusion property can enable flatter voltage profiles and reduce voltage hysteresis. Furthermore, NTP offers a moderate average working voltage compared to NVP (NTP: 2.1 V vs. Na+/Na, NVP: 3.4 V vs. Na+/Na),42 which may contribute to reduce electrolyte decomposition. These advantages position NTP as a strong candidate for a model cathode. However, despite these advantages, NTP has only been reported as a Na+ host,43–45 and its feasibility for Ca2+ storage remains unexplored.
In this study, the electrochemical properties of NTP as a Ca2+ host material were investigated using a Ca full-cell configuration consisting of a Ca-metal anode and a hydride-based electrolyte (Ca(CB11H12)2 in DME/THF), which exhibits high compatibility with Ca metal and minimises the impact of electrolyte–Ca metal interactions in this system. This study provides the first demonstration of reversible Ca2+ insertion/extraction in NTP, exhibiting minimal voltage hysteresis and a flat voltage plateau, both of which are key characteristics for a practical model cathode. The galvanostatic intermittent titration technique (GITT) test revealed that the fast Ca2+ diffusion rate in NTP is comparable to the Li+ diffusion rate in conventional cathodes for Li-ion batteries. Ex situ X-ray diffraction (XRD) and X-ray photoelectron spectroscopy (XPS) analyses, conducted before and after discharge and charge, confirmed the reversible structural changes in NTP and the corresponding valence state of Ti. These findings highlight the potential of NTP as a model cathode material for Ca-metal batteries, offering small voltage hysteresis and a single voltage plateau enabled by fast Ca2+ diffusion.
For the battery tests, the electrodes were assembled with a separator, electrolyte, and a Ca-metal anode in a stainless-steel electrochemical-cell holder. The tests were performed at a current density of 5 mA g−1 at 25 °C within the voltage range of 1.6–2.5 V using a battery test system (580 Battery Test System, Scribner Associates, Inc). To verify NTP's capability as a Na+ host and to elucidate the specific advantages and challenges of Ca insertion/extraction, Na half-cells were fabricated using activated carbon instead of the Ca-metal anode and NaTFSI in DME as the electrolyte. The electrochemical evaluation of the Na half-cell was conducted at 10 times the rate of the Ca cell.
Cyclic voltammetry (CV) and electrochemical impedance spectroscopy (EIS) were performed using a multichannel potentiostat (1470E Cell Test System, Solartron Analytical) and a multichannel parallel impedance analyser (1455A, Solartron Analytical). CV was conducted at a scan rate of 0.1 mV s−1 over a voltage range of 1.5–2.6 V vs. Ca2+/Ca. The impedance spectra were recorded by applying an alternating-current (AC) voltage of 5 mV over a frequency range of 1 MHz to 10 mHz. The GITT test was performed using a galvanostat (SP-200, Biologic). During the GITT, the battery cell was charged at a current density of 5 mA g−1 for 30 min, followed by an open-circuit rest period of 10 min.
SEM observations of ncNTP revealed NTP nanoparticles embedded within the CNT matrices, which were used to analyse the morphology of the composite (Fig. 1c). The EDS spectrum confirmed the presence of key elements in NTP, exhibiting signals corresponding to Na at 1.04 keV (Na Kα), P at 2.01 keV (P Kα), and Ti at 4.51 keV (Ti Kα) and 4.93 keV (Ti Kβ) (Fig. S2†). In addition, an Al signal at 1.48 keV originating from the sample stage was detected. EDS mapping revealed a uniform distribution of all elements constituting NTP throughout the CNT at the micron scale. In contrast, an SEM image of lpNTP revealed a larger, chunk-like NTP particle exceeding 20 μm in size. EDS mapping confirmed the uniform distribution of all NTP elements in the particles (Fig. S3†).
To investigate the nanoscale structure and particle size of ncNTP, TEM analysis was conducted (Fig. 1d). The observed black particles were identified as NTP nanoparticles, with sizes ranging from approximately 10 to 50 nm. The fibrous structures visible in the TEM images were attributed to CNTs. TEM observations confirmed that NTP nanoparticles are embedded within the CNT matrix. This entanglement and dispersion of NTP nanoparticles within the CNT matrix suggest the establishment of efficient electron conduction pathways and a short diffusion length of Ca2+ in NTP crystals (Fig. 1e). While a well-dispersed NTP/CNT composite was achieved, aggregation of NTP particles was also observed, which may hinder electrochemical performance owing to insufficient contact with the CNT matrix (Fig. S4†).
The galvanostatic discharge and charge performance of ncNTP (Fig. 2c) was evaluated within the voltage range of 1.5–2.6 V at a current density of 5 mA g−1 and room temperature. The theoretical capacity of a two-electron reaction involving one Ca2+ insertion per formula unit of NTP is 132.8 mA h g−1. The initial discharge and charge capacities are 53.4 and 40.6 mA h g−1, corresponding to 0.40 and 0.30 mol of Ca2+ per 1 mol of NTP, respectively. These values fall short of the theoretical capacity, suggesting that the relatively large NTP particles impede the full progression of the reaction. Furthermore, lpNTP failed to function in a Ca full-cell, likely because of the large particles (Fig. 2c inset). In contrast, both ncNTP and lpNTP exhibited good performance in the Na half-cell (Fig. S8 and S9†). This discrepancy is attributed to three factors. (1) The diffusion coefficient of Ca2+ (DCa2+) is lower than that of Na+ (DNa+). (2) Higher reaction energy for Ca2+ insertion/extraction than that for Na+, indicated by a lower average reaction voltage (Ca: 1.7 V vs. Na+/Na, Na: 2.1 V vs. Na+/Na, Fig. S10†). (3) The significant changes in the crystal structure induced by Ca2+ insertion may lead to material degradation (discussed in detail in the following section). These Na half-cell evaluations are validated through comparison with full-cell evaluations (Fig. S11†).
To further assess the electrochemical performance of ncNTP, galvanostatic charge and discharge measurements were conducted at various current densities (Fig. S12†). At moderate rates (10 and 20 mA g−1), a distinct plateau was observed, suggesting that ncNTP exhibits relatively high Ca2+ diffusivity, enabling Ca2+ insertion and extraction even at elevated rates. However, as the rate increased, a progressive capacity fade was observed, which can be attributed to the expansion of unreacted regions within the electrode. Several factors may contribute to this phenomenon: (i) structural distortions and phase transitions – the insertion of Ca2+ induces structural strain and phase transitions, leading to localised increases in diffusion barriers that hinder further ion transport. (ii) Delayed Ca2+ supply from the electrolyte – since Ca2+ is divalent and exhibits a larger solvation structure than monovalent cations, its desolvation and transport kinetics are inherently sluggish. This limitation becomes more pronounced at high rates, exacerbating the capacity fade. Additionally, a notable increase in overpotential was observed during discharge compared to charge, suggesting that Ca2+ insertion is kinetically more hindered than its extraction. This asymmetry implies that additional energy is required to overcome the barriers associated with Ca2+ intercalation, likely due to stronger electrostatic interactions and steric hindrances within the host framework. These findings highlight the favorable Ca2+ diffusion properties of ncNTP at moderate rates while also revealing the kinetic limitations that arise at higher current densities, providing insight into the rate-limiting factors in Ca-ion storage using NTP.
The voltage profile of the full-cell using ncNTP and Ca metal, as shown in Fig. 2c, exhibited small voltage hysteresis and a non-sloping curve. To evaluate this characteristic quantitatively, calculations were performed as described below, and the results were compared with previous reports on cathodes for Ca-metal batteries (Fig. 2d). Voltage hysteresis (Vhys) is defined as the voltage difference between the peak tops of the charge and discharge voltages in the dQ/dV curves. The dQ/dV profile of ncNTP exhibited redox peaks with maxima at 1.87 and 1.91 V, corresponding to a hysteresis of only 0.04 V (Fig. 2e). These values were obtained using a three-electrode configuration. Considering the potential of the counter electrode during discharge (0.01 V) and charge (−0.18 V), the discharge and charge voltages in a two-electrode configuration were estimated to be 1.86 and 2.09 V, respectively, corresponding to a hysteresis of 0.23 V. If the dQ/dV curve is not presented in the paper, Vhys is defined as the difference between the charge and discharge voltages at a state of charge (SOC) of 50%. The plateau flatness was defined as the average of the discharge and charge flatness, which was calculated as the voltage change (Vdis and Vcha) divided by the capacity change (Qdis and Qcha) within the plateau region from 20% to 80% SOC (Fig. 2f). As shown in Fig. 2d, Ca-metal batteries using ncNTP exhibited smaller voltage hysteresis and flatter plateaus than previously reported cathodes for Ca-metal batteries.
To investigate the Ca2+ diffusion kinetics in ncNTP, the GITT test was performed, and the Ca2+ diffusion coefficient (DCa2+) was calculated (Fig. 2g). During discharge, two distinct regions were observed: a plateau in the first half of the curve and a broad region in the second half. These regions are also reflected in the DCa2+ profile, suggesting that the Ca2+ diffusion pathway changes with the discharge reaction mechanism. In contrast, during charging, a single region was observed in both the GITT and DCa2+ profiles. The average DCa2+ values during discharge and charge in the plateau region (discharge: SOC 20–60%, charge: SOC 20–80%) were calculated to be 5.7 × 10−13 and 9.3 × 10−13 cm2 s−1, respectively. These values are comparable to the Li-ion diffusion coefficient (DLi+) in conventional LIB cathode materials, such as LiCoO2 (10−13–10−11 cm2 s−1)48 and LiFePO4 (10−16–10−14 cm2 s−1).49 In addition, GITT measurements of ncNTP were conducted to calculate the Na-ion diffusion coefficient (DNa+). The average DNa+ values during discharge and charge in the plateau region (SOC 20–80%) were 5.4 × 10−13 and 2.8 × 10−12 cm2 s−1, respectively (Fig. S13†). The obtained DCa2+ is lower than DNa+, suggesting that Ca2+ diffusion in NTP is slower than Na+ diffusion, which may explain the performance discrepancy between the Ca full-cells and Na half-cells.
The FE-SEM images confirmed that the embedded NTP nanoparticles did not undergo significant morphological changes (Fig. S14†). The morphology of the NTP nanoparticles after both discharging and charging remained similar to that of their as-deposited state, indicating that no significant damage or deformation of NTP particles occurred during the discharging and charging processes. To quantify the amount of Ca2+ inserted and extracted within the NTP, EDS analysis was performed (Fig. S15†). The EDS spectra were normalised to the Ti signal. After discharge, Ca signals were observed at 3.7 keV (Ca Kα) and 4.0 keV (Ca Kβ). After charging, the intensity of the Ca signal was reduced compared to that in the discharge state, indicating the extraction of Ca2+ from the NTP framework. The amount of Ca quantified by EDS agreed well with calculated values from electrochemical studies. Additionally, the Na content remained relatively constant after both discharging and charging, suggesting that the reaction predominantly involved Ca2+ transport.
Ex situ XRD analysis was conducted after discharging and charging to investigate the structural changes during the insertion and extraction of Ca2+ within the NTP framework (Fig. 3a). In the pristine state, peaks corresponding to NTP were observed, along with a broad peak at approximately 18°, attributed to the CNT and dome of the sample holder, and a peak at approximately 43°, originating from the sample holder. After discharge, the peak intensity of NTP was reduced, and peaks corresponding to new phases appeared at 23.7°, 28.6°, 32.1°, and 35.2°. For Na+ insertion into NTP, the main peak corresponding to the (113) plane at 24.2° disappeared, and a new peak associated with the (113) plane of Na3Ti2(PO4)3 appeared at 23.5°.44 In this study, a similar phenomenon was observed for Ca2+ insertion into NTP: the peak corresponding to the (113) plane at 24.3° disappeared, and a new peak emerged at 23.7°. This suggests that CaxNaTi2(PO4)3 (CaNTP), which is formed by the insertion of Ca2+ into NTP, has a crystal structure similar to that of Na3Ti2(PO4)3. In addition, the (024) and (116) planes of Na3Ti2(PO4)3 at 28.5° and 31.8°, respectively, were similar to the new peaks of CaNTP at 28.6° and 32.1°, which is consistent with the above discussion. Furthermore, for Ca2+ insertion into NaV2(PO4)3, which has the same crystal structure as NTP, it has been reported that the peak shifts during Na+ and Ca2+ insertion are similar, supporting the above interpretation.50–52 The peak at 35.2° is attributed to the formation of CaCO3 as part of the cathode–electrolyte interphase (CEI) on the electrode surface. Residual peaks of NTP were also observed, indicating the presence of unreacted NTP, possibly owing to the large number of NTP particles. After charging, the peaks corresponding to the new phase disappeared, and the intensities of the NTP peaks increased, suggesting the successful extraction of Ca2+ from NTP.
To investigate the insertion and extraction of Ca2+ within NTP, as well as the changes in the oxidation state of Ti during discharging and charging, XPS analyses were conducted (Fig. 3b and c). For the pristine electrode, no Ca 2p peak was observed, and the Ti 2p spectrum showed the presence of only Ti4+. Upon discharging to 1.5 V, Ca 2p peaks appeared at 347.8 and 351.4 eV, indicating the presence of Ca on the electrode after discharge. Additionally, the Ti 2p spectrum shifted to a lower binding energy, exhibiting two peaks corresponding to Ti4+ and Ti3+, which suggested that the Ti in NTP was partially reduced by Ca insertion. After charging, the intensity of the Ca 2p peaks was reduced, indicating the extraction of Ca2+ from NTP. However, residual Ca 2p peaks remained, indicating the presence of partially unreacted NTP or the formation of a CEI layer containing Ca(CB11H12)2 salts. According to the changes in the valence state of Ti in NTP calculated through the deconvolution of the XPS profiles, the reaction can be formulated as follows.
Discharge: NaTi2(PO4)3 + 0.62Ca2+ + 1.24e− → 0.62CaNaTi2(PO4)3 + 0.38NaTi2(PO4)3 |
Charge: 0.62CaNaTi2(PO4)3 + 0.38NaTi2(PO4)3 → 0.26CaNaTi2(PO4)3 + 0.74NaTi2(PO4)3 + 0.36Ca2+ + 0.72e− |
To further investigate the CEI, additional XPS analyses were performed on C and O before and after charging (Fig. S16†). The C 1s peak remained unchanged, indicating that the CNT and binder in the electrode were chemically stable and did not degrade. The deconvoluted O 1s peak revealed contributions from PO, O from PO4 units in the NTP lattice, C
O, and COO groups, as well as O from the CNT and binder, including adsorbed O and surface oxides. The P
O component did not change during the discharge and charge cycles, demonstrating the high stability of the NTP structure. In contrast, the C
O component decreased during discharging and charging, suggesting the reduction of surface oxides on the CNT and the binder and/or adsorbed oxygen. The C–O–C component increased during discharging and charging, implying the formation and accumulation of a polymeric CEI derived from the ether-based solvents, i.e. DME and THF.
The voltage profiles from the 1st to 5th cycles are shown in Fig. 3d. Reversible charging and discharging were observed in each cycle. An increase in the overvoltage was noted during discharge. It has been proposed that the highly Ca-rich NVP becomes electrochemically inert. It is likely that the increase in overvoltage is related to the accumulation of electrochemically inert, highly Ca-rich NTP over successive cycles. The cycling performance of NTP is shown in Fig. 3e. A gradual increase in the discharge capacity was observed during the 1st to 3rd cycles, which was not observed for the Na half-cell (Fig. S8†). These results suggest that electrolyte decomposition and CEI-layer growth contribute to the observed behaviour. Beyond the 3rd cycle, capacity degradation was observed during the discharge and charge cycles.
To investigate the degradation mechanisms during charge–discharge cycling, EIS measurements were performed after each cycle, and the results were analysed using an equivalent circuit model. The Nyquist plots obtained after discharge for different cycles (Fig. 3f and S17a†) exhibited characteristic features, including a high-frequency semicircle and mid-frequency semicircle, which correspond to electrolyte resistance (Rs), solid electrolyte interphase (SEI) resistance (RSEI), and charge transfer resistance (Rct), respectively. In the initial cycle, the EIS spectrum showed a relatively small Rct, indicating efficient transfer at the electrode–electrolyte interface. However, with increasing cycle number, Rct exhibited a progressive increase, suggesting a deterioration in interfacial charge transfer kinetics. This degradation is likely caused by structural distortions or the loss of active sites for Ca2+ intercalation/extraction, which may result from the accumulation of residual Ca2+ in NTP. In contrast, during charging, no significant increase in the Rct was observed after the 2nd cycle (Fig. S17b†), indicating that the degradation mechanism is mainly associated with the discharge process (Ca2+ insertion) rather than the charge process. These findings highlight the importance of optimizing the discharge process to improve long-term cycling performance.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ta00829h |
This journal is © The Royal Society of Chemistry 2025 |