Open Access Article
Pooja Kumaria,
Mustafizur R. Hazarikaa,
Chandan Sahaa,
Harishchandra Singhb and
Kaushik Mallick
*a
aDepartment of Chemical Sciences, University of Johannesburg, P. O. Box: 524, Auckland Park, 2006, South Africa. E-mail: kaushikm@uj.ac.za
bNano and Molecular Systems Research Unit, University of Oulu, FIN-90014, Finlan
First published on 1st December 2025
Electrochemical capacitors or supercapacitors have become a prominent area of research in energy storage because of their extended cycle life and rapid electrochemical performance along with significant energy and power density. This report provides a comprehensive synthesis and analysis of silver chromate (Ag2CrO4) nanoparticles and investigates their electrochemical performance as an electrochemical capacitor. Ag2CrO4 is a chemically stable material that exhibits high electrical conductivity and low solubility, making it an ideal candidate for supercapacitor applications. Nanosized Ag2CrO4 was synthesized using an organic molecule mediated complexation route where particles ranging from 10 to 30 nm in size are encapsulated within the organic matrix. For the Ag2CrO4-based three-electrode system, at a scan rate and current density of 5 mV s−1 and 3 A g−1, the specific capacity values obtained were 131 and 203C g−1, respectively. An asymmetric supercapacitor device was fabricated utilizing Ag2CrO4 as the cathode and activated carbon as the anode material, which displayed a specific capacity of 36C g−1 at 0.4 A g−1 and 31C g−1 at 10 mV s−1. The assembled device displayed a maximum power density of 361 W kg−1 at an energy density of 1.5 Wh kg−1. The performance of the device was successfully demonstrated by powering a red LED, confirming its capability as an energy storage component. Furthermore, the ability of the device to attenuate high-frequency signals was demonstrated through its integration into a resistor–capacitor (RC) circuit, where it operated as a low-pass filter, emphasizing its potential for use in frequency-selective electronic applications.
Most commercially used supercapacitors are made of carbon-based electrodes, which offer numerous advantages, including low cost, excellent corrosion resistance, good cycling stability, high electrical conductivity and strong mechanical properties.5,6 Metal oxides are known for their superior electrochemical properties, offering higher specific capacitance, enhanced chemical stability and greater energy density.7 Among these, transition metal oxides, such as nickel oxide,8 manganese oxide,9 ruthenium dioxide,10 iridium oxide,11 cobalt oxide,12 copper oxide,13 zinc oxide,14 chromium oxide,15 etc. stand out as promising candidates for use as electrode materials in electrochemical supercapacitors. These oxides are highly regarded for their ability to store and deliver energy efficiently, making them ideal for advanced supercapacitor applications. However, the primary obstacles associated with transition metal-based electrode materials include their limited specific energy and high internal resistance.16,17 To provide high specific energy, which is crucial for the performance of supercapacitor devices, the electrodes must also facilitate rapid electron transport to ensure the overall effectiveness of the supercapacitor.18 To achieve this goal, silver-containing materials have been developed because the reduced product (metallic silver) can form a network that supports fast electron transport.18
The energy storage behavior of silver vanadate as a cathode material involves a reversible intercalation–deintercalation process coupled with an irreversible displacement reaction, forming a conductive silver matrix that enhances high rate performance.19 Silver molybdate (Ag2MoO4) nanoparticle based novel electrodes were used to fabricate a high-performance asymmetric supercapacitor device that provided an energy density of 72.1 Wh kg−1.20 It was reported that mesoporous β-Ag2MoO4 as a supercapacitor electrode displayed a specific capacity of 2610C g−1 at 1 A g−1.21
Among the various silver-based materials, Ag2CrO4 stands out as an attractive photocatalyst for the degradation of organic pollutants.22–24 Despite its promising photocatalytic properties, there are no reports available in the literature regarding the performance of Ag2CrO4 as an energy storage material. The exploration of Ag2CrO4 for supercapacitor applications is justified by its favorable electronic structure, redox-active nature, and excellent chemical stability, which are crucial characteristics for efficient charge storage. Moreover, materials that exhibit strong photocatalytic activity often possess high surface area and efficient charge separation properties that are also advantageous for supercapacitor electrodes.
In this work, Ag2CrO4 nanoparticles were synthesized using silver nitrate and potassium chromate as precursors. The synthesis was carried out through a hexamethylenetetramine mediated complexation route, which enabled the controlled formation of the nanoparticles. The Ag2CrO4 nanoparticles, stabilized by an organic molecule, were used as the cathode material in the construction of an asymmetric electrochemical capacitor for energy storage applications. Silver chromate was chosen as the electrode material because of its unique redox activity, excellent electrochemical properties and good electrical conductivity. The ability of the capacitor to attenuate high-frequency signals was evaluated through integration into a resistor–capacitor circuit, demonstrating potential for low-pass filter applications.
:
1 ratio), dried under vacuum at 60 °C and characterized using different analytical techniques. The synthesized product was used as an active ingredient (cathode material) for supercapacitor application.
:
1
:
1 (wt%) in the presence of 10 µL of N-methyl pyrrolidine (NMP) and coated onto Ni-foam with a mass loading of 1.2 mg, which was used as the active working electrode. The ASC was assembled in a coin cell configuration. To fabricate the ASC, the as-prepared Ag2CrO4 and activated carbon (AC) were used as the cathode and anode, respectively. The cathode electrode material for the device was prepared as described above. The anode electrode was prepared by mixing AC and PVDF (95
:
5 wt%) in NMP. The prepared electrode materials were deposited on two nickel foams and vacuum-dried at 60 °C. The electrodes were then separated using KOH-soaked filter paper and assembled into a CR2032-type coin cell. The mass balance of both the electrodes was calculated using the equation m+/m− = CS− × V−/CS+ × V+ (CS+, m+, V+ and CS−, m−, V− are the specific capacity, mass and potential window of the cathode and anode electrodes, respectively). The total mass loading of the device was 3.6 mg.
The electrochemical parameters were calculated utilizing the following equations. Specific capacity (CS in C g−1) = ∫IdV/m × v (i) and Δt × I/m (ii), where ∫IdV is the area under the cyclic voltammetry (CV) curves and Δt, I, m and v represent the discharge time, current, mass of the electroactive material and scan rate, respectively. Coulombic efficiency (CE) = (td/tc) × 100 (iii), where td and tc are discharging and charging times, respectively. Energy density (ED in Wh kg−1) = V × ∫Idt/2 × 3.6 × m (iv) and power density (PD in W kg−1) = (ED/Δt) × 3600 (v).
The XRD analysis of the synthesized material, recorded in the 2θ range of 30° to 90°, Fig. 1, revealed diffraction peaks that matched the JCPDS file 00-026-0952, confirming the formation of Ag2CrO4, which corresponds to an orthorhombic crystal structure associated with the space group Pmnb (62).25 The figure shows the unit cell illustration of the Ag2CrO4 structure, projected along the b-axis, revealing a three-dimensional framework consisting of chromate (CrO4−2) groups connected with two distinct silver atoms, Ag(X) and Ag(y). The Ag(X) atom adopts a tetragonal bipyramidal geometry, whereas Ag(y) is arranged in a distorted tetrahedral configuration with an O–Ag–O bond angle of 107°. The CrO4−2 group also exhibits a distorted tetrahedral structure, characterized by Cr–O bond lengths of 1.66 Å and bond angles ranging from 106.1° to 111.6°. The calculated unit cell parameters are a = 7.022 Å, b = 10.065 Å, and c = 5.538 Å, with α = β = γ = 90°, resulting in a cell volume of ∼391.4 Å3.26
XPS analysis was utilized to examine the elemental composition and oxidation states of the synthesized Ag2CrO4. The survey spectra, Fig. 2A, display the peaks corresponding to Cr, Ag, O, N and C. The characteristic peaks for nitrogen and carbon originated from the organic molecule (hexamine). The high-resolution Cr 2p spectrum (Fig. 2A, inset) displays two peaks positioned at 588.4 eV and 578.3 eV, corresponding to 2p1/2 and 2p3/2, respectively, representing the Cr(VI) oxidation state, which supports the formation of Ag2CrO4.27–29 Fig. 2B shows the high-resolution Ag 3d spectrum with two distinct peaks present at 367.5 eV and 373.3 eV corresponding to the Ag 3d5/2 and Ag 3d3/2 states, respectively, which are the characteristic peaks for Ag(I).23,29 The deconvoluted high resolution O 1s spectrum, Fig. 2C, exhibits peaks at 530.2 and 531.6 eV corresponding to lattice oxygen and the hydroxyl groups adsorbed on the surface of Ag2CrO4, respectively.30,31
![]() | ||
| Fig. 2 (A) XPS survey spectrum of Ag2CrO4 (main panel). Inset: high-resolution Cr 2p spectrum. High-resolution XPS spectrum of (B) Ag 3d, (C) O 1s, (D) N 1s and (E) C 1s. | ||
Transmission electron microscopy images, Fig. 3(A and B), reveal that the silver chromate nanoparticles are predominantly spherical in shape with a broad size distribution. The particles are moderately well-dispersed, although some degree of aggregation is noticeable, which is common in nanoscale systems due to high surface energy. The histogram, Fig. 3A, inset, shows the particle size distribution, indicating that most of the nanoparticles fall within the range of 10–30 nm. A small fraction of the particles also exhibits a hexagonal shape, as observed in Fig. 3B, highlighted within a circle. Fig. 4A, shows the scanning electron microscopy image of the Ag2CrO4 based hybrid system. The image highlights the surface morphology of the system at a microscopic level. Elemental analysis using the energy dispersive X-ray spectroscopy technique shows the presence of Ag, Cr and O in the sample, Fig. 4B. The elemental mapping shows uniform distributions of Ag, Cr and O in Ag2CrO4, Fig. 4(C–E). The elemental mapping also ensures that Ag, Cr and O are well-integrated at a microscopic level.
![]() | ||
| Fig. 3 TEM images (A and B) of Ag2CrO4 particles at different magnifications. The inset in figure (A) shows the histogram (particle size as a function of particle frequency). | ||
![]() | ||
| Fig. 4 (A) SEM image of hexamine stabilized Ag2CrO4. Energy dispersive X-ray spectra (B) and corresponding element mapping of silver (C), chromium (D) and oxygen (E). | ||
The nitrogen adsorption–desorption isotherm of Ag2CrO4 is shown in Fig. 5. Based on IUPAC classification, the isotherm is classified as type IV, featuring a distinct H1 hysteresis loop between relative pressures of 0.4 and 1.0. The surface area, measured by the Brunauer–Emmett–Teller (BET) technique, was calculated to be 72.7 m2 g−1. Further analysis using the Barrett–Joyner–Halenda (BJH) method indicated a maximum pore volume of 0.035 cm3 g−1 and pore diameters distributed within the range of 19–64 Å.
![]() | ||
| Fig. 5 Nitrogen adsorption–desorption isotherm and the pore size analysis (inset) of the synthesized material (Ag2CrO4). | ||
| CrO2−4 + 2H2O + 2e− ↔ CrO2 + 4OH− | (I) |
| Ag+ + 2OH− ↔ AgO + 2H2O + e− | (II) |
| 2AgO + 2H2O + 2e− ↔ Ag2O + 2OH− | (III) |
Nitrogen adsorption–desorption analysis was performed to interpret the electrochemical behavior of the Ag2CrO4-modified electrode and its correlation with ion transport kinetics. The electrode material exhibited a moderately high surface area of 72.7 m2 g−1 with a pore size distribution in the range of 19 to 64 Å. These features are crucial for charge storage, where the high surface area provides electroactive sites for faradaic reactions, while the mesoporous structure creates efficient ion-diffusion pathways that reduce resistance and allow deeper access to internal active sites, particularly at low scan rates. The combined electrochemical and BET results confirm that hierarchical porosity facilitates smooth intercalation and deintercalation of electrolyte ions, improving ion accessibility and resulting in higher capacitance.
The electrochemical kinetics of the free-standing electrode were examined using Dunn's method, which evaluated the charge storage performance of Ag2CrO4 through both surface-controlled (capacitive contribution) and diffusion-controlled processes. We have applied the power-law equation, Ip = avb, (a is a variable parameter, b is an indicator to evaluate the charge storage kinetics, v and Ip are the scan rate and peak current, respectively) to examine the diffusive and capacitive contributions from the CV data at scan rates from 5.0 to 50 mV s−1.36 The calculated b-value of 0.6, Fig. 6A, inset, demonstrates that the diffusion-controlled process dominates the charge-storage mechanism.
The GCD profile, Fig. 6B, of the Ag2CrO4 modified electrode was studied within the potential window of −0.1 to 0.6 V at various current densities (CDs) from 12–3 A g−1. The nonlinear nature of the charge–discharge curves indicates that the storage (energy) mechanism is predominantly associated with the electrochemical oxidation–reduction process, similar to battery-type electrodes. The appearance of a potential drop in the discharge curve along with a sudden potential increase is due to a quasi–conversion reaction, an inherent electrochemical behaviour of the silver-based materials. The CS values obtained from charge–discharge curves were 59 and 203C g−1, at CD values of 12 and 3 A g−1, respectively, calculated using equation (ii).
The dependence of CS on CD and scan rate is demonstrated in Fig. 6(C and D). The electrode was stable and showed high reproducibility across three batches with a standard deviation below 3%. This trend can be attributed to the limited time available for charge carriers to participate in redox reactions at higher rates, leading to incomplete utilization of the active material and hindering the efficiency of faradaic charge storage processes. The aforementioned factors contribute to the observed reduction in specific capacity under high-rate operating conditions. The cycling stability of the electrode is an important characteristic for real time applications. In this case, the GCD study was applied to evaluate the long-term stability of the electrode. At a CD of 9 A g−1, the Ag2CrO4 electrode showed a capacity retention and CE of 91% and 99%, respectively, after 2000 cycles, Fig. 6E. The inset figure displays the first five (1–5) and last five (1996–2000) charge–discharge cycles. The electrode exhibited high reproducibility with a standard deviation below 4% for three consecutive batches. We further performed X-ray diffraction (XRD) analysis, Fig. S1, SI, after completing the galvanostatic charge–discharge cycles. The obtained XRD pattern did not show any noticeable shift or emergence of new diffraction peaks, indicating that the crystalline phase of the electrode material remained unchanged. This observation confirms the structural stability and robustness of the electrode material during repeated electrochemical cycling.
The impedance spectroscopy study was carried out across a frequency range of 100 mHz–200 kHz. The expanded view of the high frequency region of the Nyquist plot is displayed in Fig. 7A, inset (I). The Nyquist data of the device were fitted with the equivalent circuit model (Randle circuit), inset (II), consisting of solution resistance (RS), double layer capacitance (Cdl), charge transfer resistance (Rct) and Warburg impedance (W). The low-frequency and high-frequency regions indicate the ion diffusion mechanism and charge transfer process, respectively. The RS and Rct values were 0.96 Ω and 2.83 Ω, respectively, obtained from the fitting curve. The Bode plot, Fig. 7B, of the Ag2CrO4 based electrode shows the intercept at low frequency with a phase angle of −60°, which indicates efficient charge storage performance and quick ion diffusion.37,38 The plot also shows a prominent peak (time constant) at 37 Hz, indicating the behaviour of double layer capacitance.39
| Electrode material (*) | Electrolyte (M) | Three- electrode | Two-electrode (SC device) | Energy density | Power density | Retention (%) and cycle number | Ref. |
|---|---|---|---|---|---|---|---|
| a (*) CrOXNY: chromium oxynitride; MIL-101(Cr, Mg)-rGO: Mg-doped chromium-based metal organic framework-reduced graphene oxide composite; Co–Cr-LDH: 2-D cobalt-chromium layered double hydroxide and poly-oxo-vanadate anions; Ni2Cr1-LDNs: nickel-chromium layered double hydroxide nanoflakes. CoCr2O4: cobalt chromite; Cr2O3: chromium oxide; Cr2O3–Co3O4 NC: chromium oxide-cobalt oxide-based nanocomposite; Cr2O3–MoO2: chromium oxide–molybdenum dioxide. | |||||||
| CrOXNY | KOH (3) | 105 F g−1 @ 1 A g−1 | 146 F g−1 @ 5 mV s−1 (symmetric) | 8 Wh kg−1 | 28.8 kW kg−1 | 98% @ 10 000 |
40 |
| MIL-101(Cr, Mg)-rGO | KOH (6) | 261.4 F g−1 @ 1 A g−1 | — | 4.6 Wh kg−1 | 7.6 kW kg−1 | 86% @ 2000 | 41 |
| Co–Cr-LDH | KOH (2) | 732C g−1 @ 1 A g−1 | 148.5 F g−1 @ 0.8 A g−1 (asymmetric) | 35.1 Wh kg−1 | 1.2 kW kg−1 | 90% @ 5000 | 42 |
| (Fe, Cr)2O3 | KOH (1) | 45.9 mF cm−2 @ 5 mV s−1 | 16.8 mF cm−2 @ 5 mV s−1 (symmetric) | 0.57 mWh cm−2 | 200 mW cm−2 | 115% @ 5000 | 43 |
| Ni2Cr1-LDNs | KOH (6) | 1525 F g−1 @ 2 A g−1 | 155 F g−1 @ 0.5 A g−1 (asymmetric) | 55.33 Wh kg−1 | 400 W kg−1 | 81% @ 5000 | 44 |
| NiCrO3 | KOH (6) | 2862 F g−1 @ 1 A g−1 | 102.7 F g−1 @ 2 A g−1 (asymmetric) | 32.9 Wh kg−1 | 1.5 kW kg−1 | 48% @ 50 000 |
45 |
| CoCr2O4 | Na2SO4 (1) | 883 F g−1 @ 5 mV s−1 | 74 F g−1 @ 0.5 A g−1 (asymmetric) | 26.3 Wh kg−1 | 400 W kg−1 | 91% @ 5000 | 46 |
| Cr2O3 | KOH (2) | 300 F g−1 @ 1 mV s−1 | 50 F g−1 @ 1.0 A g−1 (asymmetric) | 14.2 Wh kg−1 | 568.2 W kg−1 | 98% @ 3000 | 47 |
| Cr2O3 | KOH (6) | 453 F g−1 @ 1 A g−1 | 79 F g−1 @ 1 A g−1 (symmetric) | 15.8 Wh kg−1 | 600.5 W kg−1 | 90% @ 5000 | 48 |
| Cr2O3–Co3O4 NC | KOH (1) | 619.4 F g−1 @ 10 mV s−1 | 33.07 F g−1 @ 10 mV s−1 (asymmetric) | 4.3 Wh kg−1 | 200 W kg−1 | 74.8% @ 1000 | 49 |
| Cr2O3–MoO2 | Na2SO4 (1) | 340 F g−1 @ 2 mA cm−2 | 74.5 F g−1 @2 mA cm−2 (asymmetric) | 37.35 Wh kg−1 | 9708 W kg−1 | 91% @ 20 000 |
50 |
| Ag2CrO4 | KOH (1) | 203C g−1 @ 3 A g−1 | 36C g−1 @ 0.4 A g−1 (asymmetric) | 13 Wh kg−1 | 361 W kg−1 | 85% @ 5000 | This work |
Furthermore, to enhance the performance of such capacitors, several effective strategies can be implemented. These include optimizing the electrode surface area to maximize the number of active sites, incorporating conductive additives to enhance electron transport pathways, and refining the synthesis methodology by altering the complexation agent to control the morphology and increase the pore size of the active material, thereby facilitating efficient ion diffusion. In addition, employing redox-active or multi-electrolyte systems can further improve the charge storage capability, resulting in a significant boost in specific capacitance, energy density and long-term cycling stability of the device.
We further calculated the total charge storage mechanism of the device using the power law equation, Ip = avb,51 which can be derived as log(Ip) = log(avb) = log(a) + b
log(v). The slope of the log(Ip) vs. log(v) plot defines the b-value, which is 0.5, Fig. 9C, validating the diffusion controlled charge storage mechanism. The capacitive and diffusion-controlled contributions to the total capacity at a given potential at each scan rate were further evaluated using the equation Ip = k1v + k2v1/2, where k1v and k2v1/2 represent capacitive (surface-controlled) and diffusion controlled processes, respectively. The above equation can be rearranged as, Ip/v1/2 = k1v1/2 + k2. The values of k1 and k2 can be obtained by plotting Ip/v1/2 vs. v1/2. The Ag2CrO4-based device displayed a surface-controlled capacitive contribution of 34% at 10 mV s−1, which increased to 70% at 200 mV s−1, Fig. 10A. At lower scan rates, the majority of ions in the device remain on the outer surface of the electrode, with ion diffusion behaviour dominating over the capacitive contribution. As the scan rate increased, the device exhibited a steady rise in capacitive contribution, which is advantageous for achieving high power density.52 Fig. 10(B–D) illustrates the cyclic voltammograms of the Ag2CrO4 based device at 10, 100 and 200 mV s−1. The red shaded area represents the diffusion controlled contribution and the blue shaded area of the curve represents capacitive controlled contribution.
The pattern of the Nyquist plot for the free-standing electrode and device are almost identical across the frequency range between 100 mHz and 200 kHz, Fig. 11A. The magnified image of the high-frequency region (semicircle pattern) is shown in Fig. 11A, inset (I). The RS and Rct values are 2.25 Ω and 9.47 Ω, respectively, obtained from the intercept and diameter of the semicircle. All the fitting parameters are mentioned within the figure, obtained from the Randles circuit, Fig. 11A, inset (II). Fig. 11B, main panel, shows the relationship between imaginary, C″(ω), and real, C′(ω), capacitance as a function of frequency. The imaginary and real capacitances are determined from the equations Z′(ω) = 2πfC″(ω)|Z(ω)| and Z″(ω) = 2πfC′(ω)|Z(ω)|, respectively, where ω is the angular frequency, f is the frequency and Z represents the impedance.53 The real capacitance (C′)
denotes the available capacitance that can be supplied to the device, which is 0.3 mF. The graphical representation of C″ vs. frequency reveals a peak at 1.03 Hz with a relaxation time (τ = 1/f) of 0.15 s, which identifies the time to charge the device. For practical application, two fabricated asymmetric supercapacitor devices (C1 and C2) were connected in series to power a red LED (light emitting diode). It was observed that after being charged for 30 seconds at an applied voltage of 2.6 V and a current density of 0.6 A g−1, the assembled device was capable of powering a red LED for approximately 120 seconds (Fig. 11C). This practical demonstration highlights the ability of the device to store and deliver energy for real-world applications, reflecting its potential for use in low-power electronic systems and backup energy sources.
0.07
and 0.2 Hz. At 0.02 Hz (below cut-off frequency), the capacitor was almost fully charged and discharged in each cycle, and the output waveform (Vout) closely follows the input signal. At 0.07 Hz (at the cut-off frequency), incomplete charging–discharging produced amplitude reduction and a noticeable phase lag. The output waveform amplitude decreased by 69.78%, closely matching the simulated output signal drop to 70.7% of the input amplitude, corresponding to −3 dB attenuation.55 At 0.2 Hz (above the cut-off frequency), the output waveform exhibited strong attenuation, preventing the higher frequency components of the input signal to pass through. The circuit connected with the silver chromate-based capacitor, effectively preserved low-frequency signals and attenuated higher-frequency components, demonstrating its performance as an ideal low-pass filter. A comparative table based on the simulated and experimental output peak voltage values has been included in the manuscript to serve as a ready reference for the readership, Table 2.
![]() | ||
| Fig. 12 Experimental setup of the circuit (main panel) and the circuit diagram of the RC low pass filter (inset). | ||
| Frequency (Hz) | Simulated Vout-peak (V) | Experimental Vout-peak (V) | Deviation (%) |
|---|---|---|---|
| 0.02 | 0.59 | 0.58 | 1.72 |
| 0.07 | 0.44 | 0.43 | 2.33 |
| 0.2 | 0.24 | 0.23 | 4.35 |
| This journal is © The Royal Society of Chemistry 2026 |