Amol S.
Vedpathak
ab,
Shrishreshtha A.
Sahu
c,
Tanuja N.
Shinde
b,
Shubham S.
Kalyane
b,
Sambhaji S.
Warule
d,
Ramchandra S.
Kalubarme
c,
Aditya Narayan
Singh
e,
Ravindra N.
Bulakhe
*f,
Ji Man
Kim
*f and
Shrikrishna D.
Sartale
*b
aSymbiosis Centre for Nanoscience and Nanotechnology, Symbiosis International (Deemed University), Pune 412115, India
bThin Films and Nanomaterials Laboratory, Department of Physics, Savitribai Phule Pune University, Pune 411 007, India. E-mail: shrikrishna.sartale@unipune.ac.in
cCentre for Materials for Electronics Technology (C-MET), Panchavati, Off. Dr. Homi Bhabha Road, Pashan, Pune 411008, India
dDepartment of Physics, Nowrosjee Wadia College, Pune 411 001, India
eDepartment of Energy and Materials Engineering, Dongguk University-Seoul, Seoul, 04620, Republic of Korea
fDepartment of Chemistry, Sungkyunkwan University, Suwon, 440-746, Republic of Korea. E-mail: jimankim@skku.edu; bulakhe@skku.edu
First published on 18th March 2025
Multifunctional layered nanostructures have attracted great attention for next-generation electrochemical supercapacitors and metal-ion batteries. Herein, we use a hydrothermal method to demonstrate the synthesis of 1D and layered sodium vanadate (NaV8O20) nanobelts architecture. These NaV8O20 nanobelts demonstrate outstanding electrochemical performance in supercapacitors (SCs) and sodium-ion batteries (SIBs). The possible formation mechanism of NaV8O20 nanobelts is briefly discussed. Benefiting from the 1D and layered nanostructure, pre-inserted cations, significantly enhanced electrochemical conductivity, and high electroactive surface area, the prepared NaV8O20 electrode material exhibited excellent charge storage capacity, favorable rate, and cyclic stability performance. The NaV8O20 nanobelts displayed outstanding electrochemical characteristics, including 676 F g−1 of specific capacitance, 45 W h kg−1 of energy density and 5224 W kg−1 of power density. Additionally, on testing in Na-ion batteries, the NaV8O20 nanobelts exhibit a discharge capacity of 110 mA h g−1 at 10 mA g−1 and retain ∼52% capacity after 100 cycles. Along with this, the galvanostatic intermittent titration technique (GITT) measurements reveal a high diffusion coefficient for Na+ ions, highlighting the efficient Na+ ions transportation within the NaV8O20 structure. To our knowledge, this is the first report on the use of NaV8O20 nanobelts for both SIBs and SCs, marking a significant contribution to the development of multifunctional materials for energy storage applications.
Transition metal vanadates, a class of bimetallic TMOs, have gathered great scientific attention as electrode materials for SCs and MIBs. Their appeal stems from their high surface area, low-dimensional and layered structures, excellent pseudocapacitive properties, enhanced electrical conductivity, and superior cyclic stability. These characteristics enable rapid and reversible surface redox reactions of electroactive species, making them highly suitable for advanced energy storage applications.14,15 Currently, different binary transition metal vanadates such as Co3V2O8, FeVO4, MnV2O8, Ni3V2O8, and Cu3V2O8 have proven their ability in SCs and MIBs applications owing to their wide working potential window, high specific capacitance, and energy storage capacity compared to their pristine metal oxide counterparts.16–20 For example, quasi-cuboidal CoV2O6 has been shown to exhibit a specific capacitance of 223 F g−1 at 1 A g−1, along with remarkable cyclic stability of 123% after 15000 cycles, making it suitable for SC applications.21 Similarly, 3D flower-like Ni3V2O8@CNTs have proven effective in enhancing energy storage capacity due to their high specific capacitance of 1054 F g−1 at 1 A g−1 and retain high cyclic stability of 89% after 10
000 cycles.22 Pseudocapacitive layered Fe5V15O39(OH)9·9H2O nanosheets exhibit significant potential as a cathode material for lithium-ion batteries, achieving a high reversible capacity of 350 mA h g−1 at 100 mA g−1, indicating its capability for long-term cycling stability.23 Similarly, Zr–NH4V4O10 has been identified as a promising cathode material for sodium-ion batteries, achieving a high specific capacity of 342 mA h g−1 at 100 mA g−1, positioning it as an appealing option for large-scale energy storage systems.24 However, despite these advancements, several challenges persist, including the complex synthesis procedures, use of toxic and hazardous materials and high cost of precursor materials, which hinder the feasibility of these materials for widespread industrial applications.15 Due to its numerous significant benefits, the hydrothermal method is the most frequently utilized for synthesizing metal vanadates. This approach enables accurate regulation of the morphology and particle size of resultant nanostructures, which is essential for enhancing the electrochemical performance of energy storage devices. This process is environmentally friendly and scalable, capable of being conducted at relatively low temperatures and pressures, contributing to its energy efficiency and cost-effectiveness. Furthermore, the hydrothermal approach facilitates the creation of well-crystallised nanostructures that exhibit high purity, thereby improving their applicability in energy storage applications. Considering these aspects, low-cost and environment-friendly alkali metals (Li, Na, K) can be appropriate alternatives to toxic and expensive metals for developing metal vanadate materials. However, alkali metal-inserted vanadate materials have been less explored for electrochemical SC and MIB applications.
In this study, we address these above-mentioned challenges by developing a simple, cost-effective hydrothermal route to prepare the 1D and layered structural sodium vanadate (NaV8O20) nanobelts without any post-annealing treatment. Sodium vanadate is particularly gaining attention due to its pre-inserted sodium ions, which enhances electrochemical performance and its suitability as a low-cost charge storage material for SCs and sodium-ion battery (SIB) applications. To the best of our knowledge, no prior publication has been reported that explored the electrochemical SCs and sodium-ion battery (SIB) applications of the 1D and layered NaV8O20 structures. Owing to the unique 1D layered structure and nanobelts morphology features, the NaV8O20 electrode displayed outstanding electrochemical features with high specific capacitance (Csp) and cyclic stability. The NaV8O20 material was investigated using various electrochemical and physicochemical characterization techniques to evaluate its structural, morphological and electrochemical characteristics. We have also successfully demonstrated the practical applicability of NaV8O20 nanobelts in electrochemical applications, particularly in solid-state asymmetric SCs, highlighting their potential for advanced EES systems.
![]() | ||
Fig. 1 Schematic representation of (a) hydrothermal synthesis of NaV8O20 nanobelts, (b) preparation of working electrode, (c) fabrication of ASC device and (d) fabrication of Na-ion battery. |
Fig. 1c and d depicts the schematic for the device manufacture processes for SC and SIB coin cells of the NaV8O20 nanobelts material. As shown in Fig. 1c, the AC//NaV8O20 asymmetric SC (ASC) device was fabricated using NaV8O20 nanobelts as cathode (∼1.2 mg of mass loading) and AC as an anode (∼2 mg of mass loading) materials prepared onto a circular 16 mm carbon paper current collector. These electrodes were separated using a cellulose membrane. The PVA–NaClO4 gel electrolyte was prepared by mixing 2 g NaClO4 and 2 g PVA in 25 ml DDW and then heated at 80 °C for 3 h. Finally, the AC//NaV8O20 asymmetric SC device was fabricated with a CR2032 coin cell using the PVA–NaClO4 gel electrolyte.
The sodium-ion storage performance of NaV8O20 nanobelts as the active cathode material was investigated. The electrode was fabricated by combining NaV8O20, PVDF, and AC with a weight ratio of 8:
1
:
1 and thoroughly mixed in NMP solvent. The prepared homogeneous slurry was uniformly coated onto a current collector (aluminium foil) using the doctor blade method and subsequently dried under a vacuum oven at 100 °C for 12 hours. The dried film was then cut into 16 mm diameter discs with a mass loading of approximately 0.745 mg cm−2. Sodium metal was utilized as the counter and reference electrode, while a 19 mm diameter glass-fiber Whatman separator was employed. The 2032-type coin cells were assembled in an argon-filled glove box, where oxygen and moisture levels were controlled at 0.1 ppm. A 1 M solution of NaClO4 in ethylene carbonate (EC) and diethyl carbonate (DEC) (1
:
1 v/v), with the addition of 5% fluoroethylene carbonate (FEC), was used as the electrolyte. A schematic representation of the Na+ ion battery assembly process is provided in Fig. 1d.
![]() | ||
Fig. 2 (a) XRD patterns of V2O5 and NaV8O20 powder samples and (b) XRD refinement analysis of NaV8O20 powder sample. |
Fig. 3b represents the high-resolution FESEM image of the NaV8O20 nanobelts. The width of NaV8O20 nanobelts was measured using ImageJ software and the Gaussian function was used to calculate the average width of nanobelts. The calculated average width (Fig. 3c) of the NaV8O20 nanobelt was ∼49 nm. Additionally, to identify the surface area and pore size of the NaV8O20 nanobelts sample, N2 adsorption/desorption isotherm analysis was used. Fig. 3d shows the mesopores nature of NaV8O20 nanobelts with a specific surface area of 12.41 m2 g−1 and pore size of ∼43.3 nm (inset of Fig. 3d). Furthermore, crystallinity, morphological and elemental analysis of the NaV8O20 nanobelts sample was investigated through TEM analysis. The low and high-magnification TEM images (Fig. 3e and f) display the nanobelts-like morphology of the NaV8O20 material. The corresponding HRTEM images (Fig. 3g–i) depict the well-observable lattice fringes. The measured d-spacings of 0.22, 0.45, and 0.36 nm correspond to the (204), (201), and (003) lattice planes of the NaV8O20 nanobelts. The selected area electron diffraction (SAED) pattern (Fig. 3k) further confirms the crystallinity of the NaV8O20 nanobelts, revealing their polycrystalline nature. Furthermore, the elemental mapping images (Fig. 3l–o) show a uniform spreading of Na, V, and O elements all over the NaV8O20 nanobelts surface.
The X-ray photoelectron spectroscopy analysis of the NaV8O20 nanobelts sample is provided in the ESI, S2.†
A hydrothermal synthesis route was used to prepare 1D and layered NaV8O20 nanobelts. In the typical synthesis procedure, bulk V2O5 powder (vanadium precursor) was pre-treated with ultrasonication in double distilled water to break bulk vanadium particles into smaller particles. This ultrasonication process promotes the dissociation and formation of free vanadium ionic species such as decavanadate ([V10O28]6−), polyoxovanadate (V8O204−), vanadyl ions (VO2+), hydrated vanadium species etc., because of partial reduction and condensation of vanadium.29,30
Here, the presence of water plays a vital role in the hydrolysis of V2O5. When Na2SO4 is added to the ionic solution of vanadium, the Na+ ions interact with V8O204− ions, forming hydrated sodium vanadate complexes. Here, the free sulfate ions (SO42−) do not participate directly in the formation of the final product and water molecules (crystal water) help to stabilize these complexes by maintaining charge balance, leading to the formation of the NaV8O20·2H2O (complex), where water integrates sodium ions into the polyvanadate structure.
During the hydrothermal treatment, the system is subjected to high temperature and pressure, promoting dehydration and crystallization. The removal of crystal water during this step drives the organization of sodium vanadate species into well-defined and layered nanobelt structures. The dehydration process facilitates the reorganization of the vanadate layers, promoting anisotropic and preferentially growth NaV8O20 nanobelts along the c direction, confirmed by XRD with an intense peak at 8.8° with Miller indices (001).
Afterward, the FESEM image reveals that the NaV8O20 nanobelts adopt a self-assembled morphology. This may be due to the removal of crystal water during crystallization, facilitating the formation of smooth, well-defined nanobelts. The self-assembly process is further driven by the depreciation of structural energy during crystallization, with sodium ions playing a role in charge stabilization.31
The CV curves of the NaV8O20 electrode reveal a hybrid charge storage behaviour, combining pseudocapacitive and electric double-layer (EDLC) characteristics. The Trasatti method is employed to understand the charge storage mechanism of the NaV8O20 nanobelts. This method helps distinguish between the contributions of surface-capacitive and diffusion-controlled processes. According to Trasatti, the overall capacitance arises from a blend of these two mechanisms: surface capacitive reactions and diffusion-driven charge storage.33,34Fig. 5b depicts the Trasatti plot, plotted using Csp against the square root of the scan rate. The y-axis intercept of the linear fit represents the surface-controlled contribution. Based on this intercept, the proportion of surface capacitive and diffusion-controlled charge storage components was calculated across various scan rates. These results are displayed in Fig. 5c as a histogram. This histogram shows that the surface capacitive (EDLC) contribution has the dominant charge storage mechanism over the diffusion-controlled contribution. By increasing the scan rate from 2.5 to 50 mV s−1, the surface capacitive contributions improved from 56 to 85% with no noticeable CV curve distortion detected, revealing a better rate performance of the NaV8O20 electrode. Furthermore, at lower scan rates (2.5 to 20 mV s−1), broad redox peaks and quasi-rectangular CV curves were observed (ESI, Fig. S3(d)†). This is due to surface faradaic reactions, clarifying the extrinsic pseudocapacitive nature. During the cathodic CV scans (ESI, Fig. S3(d)†), two reduction peaks at ∼0.2 and ∼0.35 V vs. SCE arose because of the reduction of vanadium ionic species from V5+ → V4+ and V4+ → V3+. Whereas, in the anodic CV scans, two oxidation peaks at ∼0.1 and ∼0.24 V vs. SCE were detected due to the oxidation of vanadium ionic species from V3+ → V4+ and V4+ → V5+. Interestingly, by increasing the scan rates, the cathodic peaks gradually shifted towards the higher potentials, indicating that rapid and scan rate-dependent redox reactions occurred near the NaV8O20 electrode surface.35Fig. 5d shows the GCD curves of the NaV8O20 electrode, performed 0.5 to 32 A g−1 of current densities. All the GCD curves show an ideal triangular and symmetric nature with a slight hump at ∼0.4 V vs. SCE, which indicates fast faradaic reactions and high reversibility of Na+ ions on the NaV8O20 electrode surface. Additionally, the GCD curves show extended charging–discharging characteristics with a negligible iR drop, signifying the outstanding rate capability. The Csp values were calculated using eqn (E2) (see ESI†) and summarized in the ESI, Table S3.† The NaV8O20 electrode displays an optimal Csp of 676 F g−1 at a current density of 0.5 A g−1; afterwards, it decreases with an increase in current density (as shown in Fig. 5e). At a 32 A g−1 current density, the NaV8O20 electrode still provides the admirable Csp of 142 F g−1. We believe these excellent Csp values mainly arise from the pre-insertion of Na+ into the V8O20 structure, which offers a high electroactive surface area, enhances the overall electronic conductivity and simplifies the Na+ ions insertion/de-insertion process. The electrode material's low electric and electrochemical resistances are typically preferred for the supercapacitor's longer cyclic lifespan and versatile applicability. The electrochemical resistance characteristics of the NaV8O20 electrode material were evaluated through EIS measurements. The Nyquist plot for the NaV8O20 electrode is presented in Fig. 5f. In the high-frequency region of the Nyquist plot, two key components are observed: the equivalent series resistance (Rs) and the charge transfer resistance (Rct). Rs represents the overall internal (ohmic) resistance, which encompasses the active electrode material's intrinsic resistance, the electrolyte's ionic resistance, and the contact resistance at the electrode/electrolyte interface.36–38 Meanwhile, the diameter of the semicircle represents the charge transfer resistance, which is associated with faradaic reactions. The simulated circuit provided that the NaV8O20 electrode possesses Rs = 4.9 Ω and Rct = 0.55 Ω. These low resistance values are attributed to the high electronic conductivity, large electroactive surface area, enhanced electron transport rate, and reduced charge transfer resistance of the NaV8O20 electrode, which collectively enable rapid faradaic reactions at the electrode/electrolyte interface. In the Nyquist plot, the Warburg impedance (W) is associated with the vertical line in the low-frequency region, corresponding to the intercalation or diffusion of electrolyte ions inside the active electrode material. The slope of the Warburg impedance approaches a 90° angle, signifying the capacitive behaviour of the NaV8O20 electrode.
Long cyclic life is one of the significant parameters for electrochemical supercapacitor applications. The electrochemical cycling stability over the NaV8O20 electrode (Fig. 5g) was performed at a scan rate of 50 mV s−1 over 10000 cycles. In the initial ∼1000 cycles, the Csp gradually increases and then tends to stabilize for further CV cycling. This is because of the activation of the NaV8O20 electrode through the insertion/de-insertion of Na+ ions and complete infiltration of Na+ electrolyte ions inside the active electrode material surface.21 The NaV8O20 electrode retains 90% of its initial Csp after 10
000 CV cycles. This outstanding cycling stability is mainly ascribed to the layered and 1D nanobelts structured assembly of the NaV8O20 electrode, which offers significant redox active sites and efficiently accommodates the possible volume deviations through Na+ ions insertion/de-insertion process, which ultimately enhances the charge storage capacity of the NaV8O20 electrode material.
Fig. 6a shows the CV profile for the AC and NaV8O20 electrodes carried out in a three-electrode system at a scan rate of 50 mV s−1 in 1 M Na2SO4 electrolyte. The CV curves of NaV8O20 and AC electrodes are nearly rectangular, measured between the voltage windows of 0 to 0.6 V vs. SCE and −1 to 0 V vs. SCE, respectively. These results imply that integrating the cathode and anode can provide a 1.6 V working voltage. After the fabrication of the ASC device, first, CV measurements with different voltages from 1 to 1.6 V (Fig. 6b) were performed, which indicates the actual working voltage can be extended up to 1.6 V without showing any alteration in CV curves due to aqueous electrolyte redox processes. Fig. 6c displays the CV curves of the AC//NaV8O20 ASC device, measured at 10 to 100 mV s−1 scan rates. These CV curves retain the quasi-rectangular shape observed for NaV8O20 nanobelts electrodes in a three-electrode setup, confirming the pseudocapacitive behaviour of NaV8O20 nanobelt material. Fig. 6d displays the GCD curves of the AC//NaV8O20 ASC device measured at 0.5 to 16 A g−1 of current densities, demonstrating good symmetric and triangular GCD curves because of the high reversible charge/discharge processes. The Csp of the AC//NaV8O20 ASC device (ESI, Table S3†) was calculated using eqn (E2) (ESI†). Fig. 6e shows the change in Csp with current densities. The highest Csp of 126 F g−1 was achieved at 0.5 A g−1 of current density. Energy (Ed) and power densities (Pd) of the AC//NaV8O20 ASC device were calculated using the following equations:39
![]() | (1) |
![]() | (2) |
Vanadate material | Method | Morphology | Three-electrode system | Supercapacitor device | Ref. | |||||
---|---|---|---|---|---|---|---|---|---|---|
Electrolyte | C sp (F g−1 @ A g−1) | Electrode stability (% @ cycles) | Voltage window (V) | E d (W h kg−1) | P d (W kg−1) | Device stability (% @ cycles) | ||||
LiVO3 | Chemical route | Nanorods | 1 M Li2SO4 | 426 @ 0.5 | 80 @ 2000 | 1.6 | 25 | 3600 | 80 @ 5000 | 40 |
Li3VO4/carbon | Solid-state | Nanofibers | 0.1 mol per L (NH4)2SO4 | — | — | 3.8 | 110 | 3870 | 86 @ 2400 | 41 |
β-Na0.33V2O5 | Hydrothermal | Nanobelts | 1 M LiCO4/PC | 320 @ 5 mV s−1 | — | 1 | 47 | 5000 | 66 @ 4000 | 42 |
β-Na0.33V2O5 | Spin coating | Nanowires | 1 M LiCO4/PC | 498 @ 0.4 | 90 @ 1500 | — | — | — | — | 43 |
NaV3O8 | Electrochemical conversion | Nanosheets | 1 M Na2SO4 | 640.10 @ 0.5 | 87.7 @ 2000 | 1.6 | 62.6 | 7200 | 81 @ 5000 | 44 |
Na2V6O16 | Chemical route | Nanobelts | 1 M Na2SO4 | 455 @ 0.5 | 90 @ 5000 | 1.6 | 42 | 4300 | 80 @ 5000 | 27 |
Na6V10O28 | Chemical route | Microrods | 1 M LiClO4 + PC | 354 @ 0.1 | — | 2.8 | 73 | 6238 | 70 @ 5000 | 45 |
KV3O8 | Chemical route | Nanobelts | 0.5 M K2SO4 | 677 @ 0.5 | 89 @ 5000 | 1.8 | 51 | 4127 | 90 @ 5000 | 46 |
Cu3V2O8 | SILAR | Nanopebbles | 1.5 M NaClO4 | 443 @ 5 mV s−1 | 85 @ 2000 | 1.15 | 8.4 | 345 | 76 @ 4000 | 47 |
S–Co3V2O8 | Hydrothermal | Nanosheets | 6 M KOH | 410 mA h g−1 at 2A g−1 | 94 @ 4000 | 1.5 | 36 | 11![]() |
98 @ 4000 | 48 |
Mn2V2O7 | Hydrothermal | Nanoparticles | 3 M KOH | 327 C g−1 at 1 A g−1 | 93 @ 10![]() |
1.5 | 44.3 | 807.4 | 158 @ 10![]() |
49 |
Ni3−xCoxV2O8 | Hydrothermal | Oval-shaped | 3 M KOH | 391 @ 0.5 | 98 @ 8000 | 1.3 | 25.5 | 7500 | 85 @ 10![]() |
50 |
Ni3V2O8/Ni | Hydrothermal | Nanosheets | 3 M KOH | 1300 @ 1 | 80 @ 7000 | 1.2 | 33.2 | 2400 | 73 @ 10![]() |
51 |
NiCo2V2O8 | Microwave-assisted | Microspheres | 2 M KOH | 113.2 mA h g−1 at 0.5 A g−1 | 82 @ 10![]() |
1.6 | 34.8 | 4000 | 80 @ 10![]() |
52 |
NaV 8 O 20 | Hydrothermal | Nanobelts | 1 M Na 2 SO 4 | 676 @ 0.5 |
90 @ 10![]() |
1.6 | 45 | 5224 |
89 @ 10![]() |
This work |
A detailed comparative analysis of the NaV8O20 nanobelts developed in this study with the various metal vanadate electrodes for SC devices previously reported in the literature is presented in Table 1.
The comparative results suggest that, compared with reported metal vanadate electrode materials, the NaV8O20 nanobelts consistently demonstrate better electrochemical performance. The NaV8O20 nanobelts exhibit significantly higher capacitance with an operating window of 1.6 V, surpassing many vanadium-based alternatives. They offer excellent cyclic stability, retaining their electrochemical performance over 10000 cycles, making them ideal for high-performance applications. Here, the key contributor to this exceptional electrochemical performance is the use of the Na2SO4 electrolyte, which provides numerous advantages, such as high ionic conductivity, high ionic mobility, and less corrosive compared to KOH. It also ensures compatibility with the Na+ pre-insertion in the electrode material, allowing for efficient ion transfer and enhanced electrochemical reactions. Moreover, the excellent electrochemical performance of the NaV8O20 nanobelts can be primarily ascribed to the pre-insertion of Na+ ions during synthesis. This pre-insertion enhances the redox active sites, provides an efficient diffusion path for Na+ ions, and ensures optimal interlayer spacing, facilitating effective Na+ ion intercalation. Additionally, the high electrical conductivity of the NaV8O20 nanobelts further improves their overall electrochemical behaviour, leading to enhanced charge storage and cyclic stability. These features collectively position NaV8O20 nanobelts as an auspicious material for next-generation energy storage technologies.
The CV measurements for 0.1 to 1 mV s−1 scan rates were also performed (ESI, Fig. S3(e)†) to examine the rate performance and reaction kinetics. Fig. S3(e)† indicates that the peak positions slightly shift towards higher potentials, specifying that the Na+ ions' insertion and de-insertion kinetics are closely related at different scan rates. To evaluate the Na+ ions diffusion characteristics for the NaV8O20 electrode material, the diffusion coefficient (D) is determined by the Randles–Sevcik equation:55
![]() | (3) |
To calculate D from the CV measurements, the above formula can be modified as follows:
![]() | (4) |
![]() | (5) |
Electrode material | Current density | GITT-pulse duration/relaxation time | Diffusion coefficient (DNa+) | Ref. |
---|---|---|---|---|
NaVPO4F | 14.3 mA g−1 | 600 s/2 h | 10−11 to 10−10 cm2 s−1 | 57 |
Na3MnTi(PO4)3 | 11.7 mA g−1 | 10 min/30 min | 10−10 to 10−15 m2 s−1 | 58 |
Fe-HCF (iron-based hexacyanoferrate) | — | 600 s/2 h | ∼10−11 cm2 s−1 | 59 |
NaFePO4 | 3.85 mA g−1 | 1 h/15 h | ∼10−17 cm2 s−1 | 60 |
Na0.66Ca0.05[Ni0.25Li0.11Mn0.64]O1.95F0.05 | 13 mA g−1 | 30 min/5 h | 10−11 to 10−12 cm2 s−1 | 61 |
Na8Fe5(SO4)9@rGO | — | — | 10−11 to 10−12 cm2 s−1 | 62 |
NaV 8 O 20 | 100 mA g− 1 | 300 s/2 h | ∼10− 9 cm 2 s− 1 | This work |
Fig. 8e displays the GCD behavior of second cycles for different rates from 10 mA g−1 to 1 A g−1. The NaV8O20 cathode provides a maximum discharge capacity of 105, 87, 70, 58, 48, 36, and 28 mA h g−1 for the current densities of 10, 20, 50, 100, 200, 500, and 1000 mA g−1. Additionally, the rate capability of the NaV8O20 cathode is shown in Fig. 8f; three cycles with each rate were tested, and the nature shows the electrochemical stability at a particular rate, but the capacity decreases with an increase in the rate. As the current density reverts to 10 mA g−1, the NaV8O20 cathode recovers ∼83.2% (84 mA h g−1) of its initial capacity. In addition, the cyclic stability measurement was performed over the fabricated Na+ ion battery. Fig. 8g displays the cyclic stability of the Na+ ion battery performed for 100 cycles at a current density of 100 mA g−1. The Na+ ion battery device displays ∼100% coulombic efficiency while retaining ∼52.4% of the initial capacity. The observed capacity fading can be attributed to the forming of a passivation layer on the electrode surface and a reduction in the number of electroactive sites. Additionally, EIS was employed to analyze the reaction kinetics of the NaV8O20 cathode, both before and after cycling, as shown in Fig. 8h. By comparing both the EIS spectra, it can be found that the Rct (quasi-semicircle) for the diffusion of sodium ions around ∼650 Ω, which after cycling gradually decreased to ∼562 Ω and shows an increase in ion and electron transport mobility as cycling is undertaken. This may be due to the decrease in the charge transfer characteristic in terms of lack of the electrolyte ions at the surface of the electrodes, activation, or reduction in active surface area during a prolonged cycling test.
Footnote |
† Electronic supplementary information (ESI) available: S1: (a) chemicals and reagents, (b) formulae for Csp calculation of working electrode and ASC device, and (c) characterization techniques, S2: XPS studies of NaV8O20 nanobelts sample, S3: electrochemical measurements of carbon paper, bulk V2O5 and NaV8O20 electrode and Na ion battery, Table S1: crystallographic data of the NaV8O20 powder obtained from Rietveld refinement, Table S2: Csp values of NaV8O20 electrode from CV curves, Table S3: Csp values of NaV8O20 electrode from GCD curves, Table S4: Csp, energy and power density values of AC//NaV8O20 ASC coin cell device. See DOI: https://doi.org/10.1039/d4ta08624d |
This journal is © The Royal Society of Chemistry 2025 |