Shichen Sun,
Boyu Wang and
Kevin Huang*
Department of Mechanical Engineering, University of South Carolina, SC29201, USA. E-mail: huang46@cec.sc.edu
First published on 21st July 2025
Aqueous zinc-ion batteries (AZIBs) have been actively studied in recent years as a promising solution for next-generation stationary energy storage due to their inherent safety, low cost, and high energy density. However, their practical deployment remains hindered by the limited cycling stability of cathode materials. Overcoming this challenge requires a detailed understanding of cathodic electrokinetics and degradation mechanisms. In this study, we investigate the electrokinetic behavior of a NaCa0.6V6O16·3H2O (NaCaVO) cathode in ZnSO4 electrolyte through a combined application of the galvanostatic intermittent titration technique (GITT) and electrochemical impedance spectroscopy (EIS). For the first time, we quantify the exchange current density (i0) and interfacial charge-transfer resistance (RCT) of NaCaVO as a function of states of charge (SOCs). The results reveal that the V4+ ⇌ V3+ redox reaction exhibits significantly slower kinetics than the V5+ ⇌ V4+ counterpart. Further GITT-EIS studies using D2O–ZnSO4 electrolyte, complemented by X-ray diffraction (XRD) and X-ray photoelectron spectroscopy (XPS), indicate that the sluggish V4+ ⇌ V3+ process is predominantly associated with proton insertion. Distribution of Relaxation Time (DRT) analysis correlates the increased interfacial resistance with the intermediate phase formation induced by this proton insertion. The electrokinetic insights obtained in this work fill critical knowledge gaps in AZIB research and provide a foundation for designing more durable and efficient cathode materials in the future.
Despite their promise, the commercial development of AZIBs faces serious challenges, particularly due to degradations associated with the cathode and its interfacial interactions with the electrolyte.15–17 Therefore, the advancement of AZIB technology critically hinges on the development of high-capacity, durable cathode materials, which in turn requires a thorough understanding of cathodic degradation mechanisms.1,18–21 Additionally, there remains a substantial knowledge gap in the understanding of cathodic kinetics – especially those that govern the ion storage processes at various states of charge (SOCs).
To date, a major effort in aqueous zinc-ion battery (AZIB) research has been focused on engineering a stable zinc-anode/electrolyte interface to suppress the hydrogen evolution reaction (HER) at the zinc anode and mitigate zinc corrosion. Strategies such as surface coatings and electrolyte additives have been widely explored for this purpose. However, many of these additives inadvertently hinder cathodic kinetics, ultimately compromising overall battery performance despite effectively suppressing HER and reducing corrosion.22,23
In comparison, cathode-focused research has primarily centered on compositional modifications of MnO2− and V2O5− based materials to enhance capacity and cycling stability. A notable example is the pre-insertion of alkaline and alkaline-earth metal cations into V2O5-based structures.8,24–29 In these studies, a variety of in situ and ex situ surface and bulk characterization techniques – such as X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), Electron Microscopy, Fourier Transform Infrared Spectroscopy (FTIR), and X-ray absorption spectroscopy (XAS) – are routinely employed to probe material properties and elucidate underlying reaction mechanisms. In parallel, conventional electrochemical techniques30 such as Galvanostatic/Potentiostatic Intermittent Titration (GITT/PITT),31,32 electrochemical impedance spectroscopy (EIS),33 and cyclic voltammetry (CV)34,35 are used to study electrochemical behaviors. Despite these efforts, rigorous investigation of cathodic electrokinetics during actual charge and discharge processes remains rare. Furthermore, many early studies only evaluated charge-transfer resistance (RCT) under open-circuit conditions, which do not accurately represent real operating states of AZIBs, thereby limiting their value for mechanistic understanding and battery performance modeling.
In this study, we present a combined GITT and EIS approach to probe cathodic electrokinetics under actual operating conditions. The cathode material of choice is NaCa0.6V6O16·3H2O (NaCaVO), previously identified in our work as a stable and promising candidate for AZIB applications with outstanding cyclability (e.g., 94% capacity retention after 2000 cycles at 2 A g−1 and 83% after 10000 cycles at 5 A g−1).36 Electrokinetic measurements were conducted using a three-electrode configuration, with NaCaVO as the working electrode (WE), zinc metal as the counter electrode (CE), Ag/AgCl as the reference electrode (RE), and 2 M ZnSO4 as the electrolyte – an industry-relevant formulation. To deconvolute overlapping electrochemical processes and better interpret cathodic behavior, we also apply Distribution of Relaxation Times (DRT) analysis to EIS spectra, correlating observed features with the formation of interfacial secondary phases. Additionally, to investigate the co-intercalation mechanisms of Zn2+ and H+ into the NaCaVO structure, we employ water isotope solvent D2O in the ZnSO4 electrolyte system to distinguish the contributions of zinc ions from those of protons.
![]() | ||
Fig. 1 H2O–ZnSO4 system: (a) CV and (b) Ip vs. v1/2 of NaCaVO|2 M ZnSO4|Zn cell scanned at different rates. RE: Ag/AgCl; CE: Zn; WE: NaCaVO. |
Moreover, to investigate the ion diffusion characteristics in the NaCaVO cathode in H2O–ZnSO4 system, cyclic voltammetry (CV) conducted at varying scan rates (0.1–1.0 mV s−1) was further analyzed by plotting the peak currents corresponding to distinct redox reactions vs. square root of scanning rate to determine diffusion coefficient of the electroactive species of a diffusion-controlled redox process using Randles–Ševčík equation:42,43
Ip = (2.69 × 105) × n3/2 × A × D1/2 × C × v1/2 | (1) |
Fig. 1(b) shows that a plot of Ip versus v1/2 of the four redox peaks follow a linear relationship, indicating semi-infinite diffusion as the rate-limiting process. The calculated diffusion coefficients related to the four redox peaks are 4.74 × 10−10 and 1.37 × 10−10 cm2 s−1 for C1 and D1, respectively, located at the high-potential region and 1.14 × 10−11 and 3.11 × 10−11 cm2 s−1 for C2 and D2, respectively, located at the low-potential peaks. The higher D values for the high-potential V4+/V5+ redox couple than the V3+/V4+ counterpart suggests a facile kinetics for the former, which will be further discussed in the following sections.
The potential (E) profiles collected by GITT during charge and discharge cycle are shown in Fig. 2(a). As potential E increases and decreases with time during charge and discharge cycle, respectively, pseudo-plateaus regions (quasi-flat regions) show good alignment with the two redox peaks identified in the CV (see Fig. 1(a)). During the charge cycle, the profile exhibits a mid-E pseudo-plateau around 0.85–0.95 V (corresponding to V4+ → V5+ oxidation peak) and a lower pseudo-plateau near 0.60–0.70 V (corresponding to V3+ → V4+ oxidation peak). Similar pseudo-plateaus are also observed during the discharge cycle, corresponding to V5+ → V4+ and V4+ → V3+ reduction peaks, respectively. The greater E changes in lower E region than in higher E region imply higher polarization and sluggish kinetics in the region.
![]() | ||
Fig. 2 ZnSO4–H2O system: (a) E vs. time profile collected under ±50 mA g−1; EIS spectra measured with ±50 mA g−1 bias during charge (b) and discharge (c) cycles. |
The EIS spectra collected at each GITT step during charge and discharge cycled at ±50 mA g−1 (equivalent to ±0.024 mA cm−2) are shown in Fig. 2(b) and (c), respectively, as an example. Each spectrum is measured under different bias currents to obtain RCT= RCT(i) at a specific E (or SOC) created by the preceding galvanic polarization. From Fig. 2(b) collected during the charging cycle, it is evident that EIS spectra feature a charge transfer impedance at high frequencies, followed by low-frequency Warburg impedance (related to semi-infinitive surface diffusion of the active species) over low E range (step 1–10, 0.53–0.70 V). As E increases, the spectra gradually transit to a mixed Warburg (step 11–14, 0.72–0.78 V) and charge-transfer only feature (step 15–28, 0.80–0.93 V) over intermediate E range, and finally to charge-transfer only feature over high E range (step 29–45, 0.95–1.50 V). During the discharging cycle, the above trend remains, i.e., Warburg impedance appears at low E, whereas charge-transfer-only feature appears at high E. Note that the appearance of Warburg impedance at low E range implies diffusion limitation to the active species; we will further correlate it with the formation of an intermediate phase in the following section. The original E vs. time profiles and corresponding EIS spectra collected under OCV and other current densities such as 75–125 mA g−1 can be found in Fig. S1.† By comparing Fig. 2 and .S1,† it is concluded that the shape of all profiles (two distinct pseudo-plateau regions) remains the same, suggesting that the electrochemical reaction mechanisms remain unchanged by the applied current.
With the equivalent circuit model shown in Fig. 2(b-1), RCT under low E is extracted from the interception length on Z′-axis of the fitted curve (dotted), and for the high E cases, RCT (sum of RCT1 and RCT2) is similarly taken using the equivalent circuit model in Fig. 2(c-1). The obtained RCT is then plotted against E and shown in Fig. 3 for both charge and discharge cycle in a current range of 0–125 mA g−1. A clear dependence of RCT on E is observed. In both charge and discharge cycles, the RCT–E profiles can be divided into three stages. Stage-1, the lowest E range (0.40–0.60 V for both charge and discharge), RCT is generally the highest but decreases with E. Stage-2, the intermediate E range (0.68–0.91 V for charge and 0.60–0.83 V for discharge), a maximum is observed. Stage-3, the high E range (>0.91 V for charge and >0.83 V for discharge), a plateau is observed. The minimal RCT appears to occur at a high E close to the V5+ ⇌ V4+ as indicated in CV, see Fig. 1(a). In contrast, high RCT appears at lower E range close to the V4+ ⇌ V3+. Therefore, we can conclude from Fig. 3 that the V4+ ⇌ V3+ redox reaction is more sluggish than the V5+ ⇌ V4+ counterpart. This finding could be explained by a greater difficulty in inserting/extracting ions into/from NaCaVO at a deeper SOC. Another reason could be related to the formation of intermediate phase on the surface of NaCaVO cathode due to proton insertion; we will show evidence to support this assertion in a later section. On the other hand, in the higher E and shallow SOC regime, the charge transfer kinetics is facile with low RCT, suggesting a lower presence of the intermediate phase. In addition, the ohmic resistance RO of all cells studied was roughly ∼0.45 Ω cm2 (see Fig. 2 and S1†), independent of DC bias applied.
![]() | ||
Fig. 3 H2O–ZnSO4 system: RCT vs. E during (a) charge and (b) discharge cycle under different current densities. |
The trend of enhanced charge transfer kinetics by the applied current agrees with Butler–Volmer law that predicts higher overpotential drives a lower RCT. The agreement also suggests that the charge transfer kinetics is a rate limiting step of the ion insertion process into NaCaVO. Although the shapes of two RCT–E profiles between charge and discharge cycle are close to a mirror image, the discharge cycle does exhibit a slightly higher RCT than the charge cycle at the same E, indirectly reflecting influence of the intermediate phase formation on charge transfer process.
η = (RT/nF) × i/i0) | (2) |
![]() | ||
Fig. 4 H2O–ZnSO4 system. (a) RCT vs. i, (b) η vs. i and (c) i0 comparison at 0.68 V and 0.91 V during charge and 0.83 V and 0.6 V during discharge. |
A comparison of i0 at different E of the charge and discharge cycle is given in Fig. 4(c); it informs that i0 is higher at higher E than at lower E. Between the charge and discharge cycle for the same pair of redox reaction, i0 is higher for the charge cycle than the discharge cycle. Again, these comparisons suggest that the V5+ ⇌ V4+ is a more facile redox reaction than the V4+ ⇌ V3+ counterpart. We hypothesize that the sluggish V4+ ⇌ V3+ redox reaction could be attributed to the intermediate phase formation induced by the H+ insertion at a deeper SOC. To verify this, we performed a further analysis on EIS spectra with DRT method and replacing of H2O with D2O as a solvent for the ZnSO4 electrolyte.
![]() | (3) |
![]() | ||
Fig. 6 H2O–ZnSO4 system. Ex situ XRD patterns of NaCaVO cathode after (a) charge and discharge at different SOCs and (b) cycled at low and high E compared with the pristine. |
The XPS spectra shown in Fig. S4† indicate a weaker V-2p signal in NaCaVO after cycling at low E than those of the pristine and cycled at high E, suggesting possible blocking effect by the Zn-LDH formed. Furthermore, the ratio of V3+/V4+ in NaCaVO becomes significantly higher after cycling at low E compared to the pristine (with V5+ and trace V4+) and the high-E-cycled one (with much lower V3+/V4+ ratio), which is consistent with our previous results.36 These XPS results further suggest that the low-E redox reaction is associated with V3+ ⇌ V4+ and high-E redox reaction is related to V5+ ⇌ V4+.
The E-profiles collected from the D2O–ZnSO4 system under 50 mA g−1 by GITT during charge and discharge cycle are shown in Fig. 7 as an example; similar plots at other current densities can be found in Fig. S3.† The two-step pseudo-plateau feature remains, suggesting two active charge transfer processes: V3+ ⇌ V4+ and V4+ ⇌ V5+. However, a quick glance of EIS spectra indicate that RCT at low E is appreciably higher than the H2O–ZnSO4 counterpart, inferring that the V3+ ⇌ V4+ kinetics has become more sluggish by the slower D+ (de)insertion process.
A noticeable difference of EIS spectra of the D2O–ZnSO4 system from H2O–ZnSO4 system, see Fig. 7(b), is the appearance of a high-frequency semicircle (highlighted by red dotted circles) during the charge cycle. The magnitude of this small semicircle does not seem to change appreciably with SOC. Since D+ extraction from NaCaVO occurs during the charge cycle, the appeared semicircle (an extra charge transfer process) implies a reduced D+ extraction kinetics, which leads to a more sluggish V3+ ⇌ V4+ redox reaction and increased relaxation time constant. Another notable observation is that the Warburg impedance becomes less pronounced as E increases, the same trend as the H2O–ZnSO4 system.
For the subsequent discharge cycle, see Fig. 7(c), the EIS collected under high E (Fig. 7(c-1), step 1–19, 1.5–0.83 V) show a much less pronounced, small high-frequency semicircle than that observed during the charge cycle. As E decreases, see Fig. 7(c-2) (steps 20–39, 0.82–0.66 V) and (c-3) (steps 40–56, 0.65–0.4 V), this small high-frequency semicircle appears to increase slightly, while the Warburg impedance becomes more pronounced, like the H2O–ZnSO4 system. A comparison of EIS spectra vs. SOC between the H2O–ZnSO4 and D2O–ZnSO4 systems suggests similar charge transfer mechanisms, but the heavier D+ is more difficult to extract from the cathode, thus increasing the relaxation time constant of the process and depicting itself on the EIS spectrum within the frequency range studied. Since D2O has the most pronounced impact on RCT at low E, it is reasonable to speculate that H+/D+ (de)insertion into the cathode leads to the V3+ ⇌ V4+ redox reaction.
To better illustrate how RCT vary with SOC, we plot RCT vs. E under different current densities in Fig. 8. In general, the overall trend of each curve resembles Fig. 3 of the H2O–ZnSO4 system, i.e., they can be divided into three stages, featuring monotonically decreasing, peaking and flattening of RCT with SOC. However, the magnitude of RCT for the V3+ ⇌ V4+ is significantly higher (by nearly 2×) in the D2O–ZnSO4 system than the H2O–ZnSO4 counterpart, whereas RCT of the V4+ ⇌ V5+ is close for the two electrolyte systems. This observation further supports the assertion that it is H+/D+ that are mainly responsible for the V3+ ⇌ V4+ redox reaction.
![]() | ||
Fig. 8 D2O–ZnSO4 system: RCT vs. E during (a) charge and (b) discharge cycle under different current densities. |
The dependence of RCT and η on the applied current density for the D2O–ZnSO4 system is plotted in Fig. 9(a) and (b), where RCT for the charge and discharge cycle at high E is seen to remain close but notably different at lower E, i.e. RCT during the charge is lower than the discharge at a given SOC, which is like the H2O–ZnSO4 system.
Using the “low-field” approximation, the obtained i0 of the D2O–ZnSO4 system is shown in Fig. 9(c). The results indicate a greater decrease in i0 than the H2O–ZnSO4 system, i.e., it is −38% and −42% for the charge and discharge cycle, respectively, for the V3+ ⇌ V4+ occurring at low E and it is −25% and −23% for the charge and discharge cycle, respectively, for the V4+ ⇌ V5+ occurring at high E. Such an increased difference could be attributed to the slower D+ (de)insertion kinetics than H+.
NaCaVO cathode ink is prepared by thoroughly mixing 66 wt% NaCaVO with 20 wt% Super-P and 14 wt% polyvinylidene fluoride (PVDF) in N-methyl pyrrolidone (NMP) solvent. The resultant slurry is then coated uniformly onto ϕ10 mm stainless steel meshes with ∼0.6 mg cm−2 active mass loading, followed by vacuum drying at 120 °C for about 12 h and compression under 10 MPa. The aqueous H2O–ZnSO4 and D2O–ZnSO4 electrolytes are prepared by dissolving 2 M ZnSO4–7H2O (Sigma-Aldrich, ACS reagent, 99%) in DI water and D2O (Sigma-Aldrich, 99.9 atom% D), respectively.
To illustrate the phase compositions and the variations of V-oxidation state with SOCs, we employe XRD and XPS, respectively, to examine NaCaVO cycled at high voltage range (1.5–0.75 V at 0.2 A g−1 for 50 cycles) and low voltage range (0.75–0.4 V at 0.2 A g−1 for 50 cycles). The cycled samples are thoroughly rinsed with DI water before examinations to avoid interference from the electrolyte.
The surface oxidation states of the sample are investigated with AXIS Ultra DLD XPS (Kratos Analytical) instrument. The XPS system is equipped with a non-monochromatic Al Kα source (1486.8 eV) operated at 150 W X-ray gun power, a hemispherical analyzer and a load lock chamber for rapid introduction of samples without breaking vacuum. The X-rays were incident at an angle of 45°, with respect to the surface normal. Analysis was performed at a pressure of ∼1 × 10−9 Torr and high-resolution core level spectra were acquired in the constant analyzer energy mode using a pass energy of 10 eV and a 0.07 eV step size (for survey scans 80 eV pass energy is used with 0.08 eV step size). The XPS experiments were performed by using a low energy electron beam, directed at the sample, for charge neutralization. The Binding energies (BE) of all peaks are corrected in reference to the C ls peak at 285 eV and given with an accuracy of ±0.2 eV. The curve fitting procedure was carried out using the Avantage software.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ta04992j |
This journal is © The Royal Society of Chemistry 2025 |