Abstract
Despite their critical importance, developing sustainable high-performance supercapacitor (SC) electrodes with long-term stability poses significant challenges. Herein, we report a novel ternary composite electrode in which Ag/Ni-doped manganese oxide (Ag/NiO x @Mn y O z ) is supported on human hair-derived activated carbon (HHC). This composite is synthesized via a one-pot hydrothermal process followed by thermal annealing at 800 °C, a strategy that creates a conductive Ag/Ni bimetallic network and abundant oxygen vacancies in the NiO x and Mn y O z phases. During operation, operando X-ray absorption spectroscopy (XAS) confirms reversible dual-ion redox transitions (Mn2+/Mn3+ and Ni0/Ni2+) in the cathode, highlighting the material’s enhanced redox activity. As a result, HHC-supported Ag/NiO x @Mn y O z exhibits an exceptional specific capacitance (Cs) of 1770 F g–1 at 5 mV s–1 in three-electrode tests. When assembled into an asymmetric hybrid supercapacitor (AHSC), the device delivers a high energy density of 37.53 Wh kg–1 and a power density of 2251.8 W kg–1 at 3 A g–1 while retaining ∼82% of its initial capacitance after 5000 charge–discharge cycles. These results confirm the effectiveness of our sustainable HHC-supported Ag/NiO x @Mn y O z framework in addressing the enduring trade-off between energy density, power density, and cycling stability in next-generation SCs.
Keywords: sustainability, biobased carbon, Ag impregnation, manganese oxide, nickel oxide, operando XAS, energy storage


1. Introduction
Due to environmental concerns associated with fossil fuel consumption, the growing demand for advanced energy storage devices has intensified. , Across various industries, rechargeable batteries, supercapacitors (SCs), and other energy storage technologies, including consumer electronics and automotive applications, are widely utilized. − In particular, SCs are emerging as promising candidates owing to their high power density, rapid charge/discharge capabilities, and exceptional cycling stability. , SCs are classified into two primary types based on their mechanisms of charge storage: electric double-layer capacitors (EDLCs) and pseudocapacitors (PCs). EDLCs store charge through ion accumulation at the electrode–electrolyte interface. In contrast, PCs rely on fast and reversible redox reactions within active materials. ,
As pseudocapacitive materials, transition metal organic frameworks, double-layered hydroxides, and transition metal oxides (TMOs) are widely explored. − Multiple oxidation states of transition metals enable rapid redox reactions for charge storage. Of the various TMOs (RuO2, MnO2, Fe3O4, WO3, V2O5, MoO3), manganese(II) oxide (MNO) and manganese(IV) oxide (MnO2) are the most extensively studied metal oxides for SC electrodes due to their favorable properties. Manganese exists in several stable oxidation states, such as MnO, MnO2, MnO3, and MnO4, each exhibiting distinct electrochemical behavior. Among these oxides, MnO theoretically offers the highest specific capacitance due to redox reactions between Mn2+ and Mn3+. Enhanced charge storage performance has also been ascribed to the intercalation and deintercalation of electrolyte ions, driven by the reversible redox transition between Mn2+ and Mn3+ within the electrode material. However, the practical application of unmodified MnO as a supercapacitor electrode is limited by its low electrical conductivity (10–5–10–6 S/cm) and poor structural stability, both of which severely hinder its overall performance. Likewise, MnO2 offers fast and reversible redox kinetics, robust chemical and thermal stability, excellent charge–discharge behavior, a wide operating potential (∼1 V), and a high theoretical specific capacitance of 1370 F g1. , However, pure MnO2 has limited electronic conductivity, which can hinder its rate capability and the utilization of its full theoretical capacitance in practical devices. Strategies to improve the electrochemical performance of MnO2 have therefore focused on electronic structure optimization, such as noble metal incorporation, defect engineering, and composite formation.
Noble metal nanoparticles are of great interest as conductive additives in nanocomposites with carbon materials, conducting polymers, and metal oxides/hydroxides due to their exceptional electrical conductivity. , Such nanoparticles facilitate efficient electron transport from redox-active sites to the current collector, thereby enhancing charge transfer and utilization of the active material. Among noble metals, silver (Ag) stands out for its superior conductivity and comparatively lower cost. , Nanoscale Ag particles can form conductive bridges and electrical contact points within an electrode, improving overall electrochemical performance. , The integration of Ag with carbon-based substrates or metal oxides has been extensively investigated, demonstrating significant enhancements in both SCs and lithium-ion batteries. In general, doping, or modifying MnO2 with highly conductive components like noble metals, and creating structural defects such as oxygen vacancies, is an effective way to increase its electrical conductivity and expose more active sites for redox reactions. ,
Forming composite electrodes is another promising approach to overcome the limitations of a single metal oxide. By hybridizing two or more electroactive components, one can leverage synergistic effects that boost capacitive performance beyond what each component could achieve alone. , In this context, nickel oxide (NiO) is widely combined with other metal oxides. For example, NiO@MnO2, NiO/ZnO, NiO@Co3O4, and NiO/NiCo2O4 composites have been reported to take advantage of multiple redox processes within an electrode. NiO is attractive because of its low toxicity, environmental benignity, and excellent chemical and thermal stability. However, NiO suffers from low intrinsic electrical conductivity that leads to high internal resistance and subpar rate capability; as a result, its practical capacitance often falls well below the theoretical maximum. , Integrating NiO, therefore, with additional redox-active oxides and combining both with conductive matrices, viz, carbon supports, is necessary to provide extra electron pathways and active sites, thus increasing overall pseudocapacitive behavior.
As sustainable alternatives to conventional carbons for energy storage, biomass-derived carbon materials have gained much attention, offering high electronic conductivity, tunable porosity, good stability, and low-cost production. As such, human hair, a widely available waste material that decomposes slowly, presents a promising precursor for porous carbon. Improper disposal of human hair waste can cause environmental problems, namely, harmful emissions from burning and airborne particulate hazards, but repurposing this waste into useful carbon materials addresses both waste management and energy storage needs. Chemically, human hair consists predominantly of keratin (>90 wt %), along with lipids (∼2 wt %), nucleic acids, and carbohydrates. Its elemental composition is roughly 50 wt % carbon along with substantial heteroatoms: ∼22 wt % oxygen, 16 wt % nitrogen, and 3–4 wt % sulfur. These heteroatoms, when retained in the carbon structure, can contribute faradaic pseudocapacitance by providing additional redox-active functional groups, improving charge storage capacity. , Utilizing human hair-derived carbon (HHC) for electrode applications not only mitigates environmental pollution but also yields a heteroatom-rich, high-surface-area conductive scaffold that can enhance electrochemical performance in SCs.
Despite progress in developing nanocomposites and using biobased carbons, achieving a holistic solution that combines high capacitance, excellent rate performance, and long-term cycling stability in a sustainable electrode is challenging. Many MnO2- or NiO-based composites still face trade-offs. For instance, improved energy density often comes at the expense of cycle life; incorporating noble metals can increase cost or complexity if not executed judiciously. Moreover, the full potential of waste-derived carbon supports in such multicomponent systems is yet to be realized. Notably, an electrode design that combines the good pseudocapacitance properties of MnO2, the complementary redox activity of NiO, and the superior conductivity of Ag is crucial. Incorporating these materials within a sustainable framework aims to maximize surface area and electrochemical response.
This work investigates a novel ternary composite electrode that synergistically integrates Mn and Ni oxides with Ag nanoparticles on HHC support. The novelty of our approach lies in two key aspects: (1) Waste-derived HHC is used as a high-surface area, heteroatom-doped support for the metal oxides. (2) In-situ formation of Ag/Ni bimetallic framework and oxygen-vacancy-rich oxide phases, achieved through thermal tuning, promotes dual redox reactions at the electrode. This sustainable design simultaneously improves the electrode’s surface area, electronic conductivity, and electrochemical stability. The synergistic effect of the HHC support and high-temperature annealing yields a robust composite architecture with optimized charge transport pathways, emphasizing its capability for future applications. Our findings demonstrate significant advancement in electrode design for SCs, highlighting the effectiveness of combining bioderived carbon supports with bimetallic oxide networks to achieve high energy and power densities concurrently with remarkable cycling stability.
2. Experimental Section
2.1. Chemicals/Materials
All chemicals utilized were used as purchased without further refining. The chemicals/reagents include manganese chloride (MnCl2, ≥98%, grade AR, Kemaus), silver nitrate (AgNO3, 99.9%, POCH), nickel nitrate hexahydrate (Ni(NO3)3·6H2O, 98%, Loba Chemie PVT LTD), sodium hydroxide (NaOH, 98%, Loba Chemie PVT Ltd.), urea (Co(NH2)2, grade AR, QRec), ethanol (95% C2H5OH), potassium hydroxide (KOH, ≥99.0%, Kemaus), N-methyl-2-pyrrolidone (NMP, QReC), carbon black (Super P, TIMCAL), human hair (middle-aged African male), and nickel foam (NF) (thickness = 0.6 mm).
2.2. Synthesis of Human Hair-Based Activated Carbon (HHC)
In Figure , the step-by-step procedure for the synthesis of HHC is depicted. Human hair was collected from a middle-aged African and meticulously cleaned using deionized (DI) water and acetone to eliminate dust and oil. The hair was then dried in an oven at 80 °C for 24 h. After drying, the hair was placed in a quartz boat and subjected to pyrolysis within a tubular furnace under an inert (N2) atmosphere, following a two-stage process. Initially, the furnace was purged with nitrogen for 15 min to eliminate residual oxygen. In the first stage of pyrolysis, the temperature was elevated to 220 °C at a rate of 5 °C per min and maintained at that temperature for 30 min to enable degasification and remove moisture and volatile substances.
1.
Schematic illustration for the synthesis of HHC and C–Ag/NiO x @Mn y O z composites.
The second stage involved heating the material further to 900 °C at the same heating rate and maintaining it for 1 h to ensure complete carbonization. The resulting carbonized hair was then subjected to chemical activation by mixing it with KOH pellets in a weight ratio of 1:4. This KOH-infused carbon was reheated at 900 °C under identical conditions to bolster its porosity. Following activation, the sample was neutralized with 1 M HCl, thoroughly washed with DI water, and filtered. Finally, the material was dried at 80 °C for 24 h.
2.3. Synthesis of HHC-Supported Ag/NiO x @Mn y O z Composites
First, equimolar solutions of 0.2 M MnCl2, Ni(NO3)2·6H2O, and AgNO3 were combined with a sonicated suspension of 200 mg of HHC in 30 mL of DI water. Then, the mixture was stirred continuously at 230 rpm using a magnetic stirrer for 30 min. After that, solutions of 0.4 M CO(NH2)2 and 5 M NaOH were added to the mixture and stirred for an additional 30 min. The resultant colloidal suspension was transferred into a PTFE-lined autoclave and heated in an oil bath at 180 °C for 6 h. The precipitate was collected, washed thoroughly with water and ethanol, and dried in an oven at 80 °C for 24 h. The dried material was then divided into three portions, two of which were subjected to different annealing temperatures of 400 and 800 °C for 1 h under identical annealing conditions described in Section .
The samples were labeled as C–Ag/NiO x @Mn y O z , C–Ag/NiO x @Mn y O z @400 °C, and C–Ag/NiO x @Mn y O z @800 °C. For comparison, a control sample (Ag/NiO x @Mn y O z ) was prepared following the same procedure but without the inclusion of HHC and annealing.
2.4. Sample Characterization and Electrochemical Investigation
Detailed descriptions of the material characterization techniques and the methodologies employed to assess the electrochemical performance of the synthesized materials are provided in the Supporting Information.
2.5. Assemblage of C–Ag/NiO x @Mn y O z @800 °C/HHC AHSC
AHSC was fabricated using a C–Ag/NiO x @Mn y O z @800 °C-coated nickel foam (NF) as the cathode, with a mass loading of ∼1.5–2.0 mg, and an HHC-coated NF as the anode, with a mass loading of ∼2.0–2.5 mg. A Whatman glass microfiber, presoaked in 2 M aqueous KOH electrolyte solution for 30 min, was employed as the separator. The assembled components, including the electrodes and separator, were then integrated into a stainless steel CR2025 lithium coin cell casing.
3. Results and Discussion
3.1. X-ray Diffraction (XRD) Analysis
The synthesized samples were systematically analyzed via XRD, as illustrated in Figure a,b. In Figure a, two broad diffraction peaks at ∼2θ = 25.7 and 43.1° are displayed, corresponding to the (002) and (100) planes of the turbostratic carbon structure. These peaks confirm the presence of crystallites, as indicated by the graphite-like (002) and (100) reflections observed in the HHC spectrum. The peak at 2θ = 43.1° specifically denotes a higher degree of interlayer condensation within HHC, a structural attribute expected to significantly augment its electrical conductivity. Additionally, the peak observed at 2θ = 10.5° suggests a transition of the amorphous phase in HHC toward graphitic structures. However, the relatively low intensity of the 2θ = 43.1° peak suggests that only a minimal amount of graphitic material is present within the overall structure. ,
2.
XRD patterns and XPS spectra of synthesized materials: (a) XRD patterns of human HHC, (b) XRD patterns of C–Ag/NiO x @Mn y O z composite materials at different conditions, (c) XPS wide scan, (d) Mn 2p XPS spectra of C–Ag/NiO x @Mn y O z @800 °C, (e) Ni 2p XPS spectra of C–Ag/NiO x @Mn y O z @800 °C, (f) Ag 3d XPS spectra of C–Ag/NiO x @Mn y O z @800 °C, (g) C 1s XPS spectra of C–Ag/NiO x @Mn y O z @800 °C, and (h) O 1s XPS spectra of C–Ag/NiO x @Mn y O z @800 °C.
In Figure b, the presence of metallic silver (Ag0) is confirmed by the diffraction peaks observed at 2θ angles of 38.1, 64.1, 77.1, and 81.4°, corresponding to the (111), (220), (311), and (222) planes, respectively. These peaks are in close agreement with the reference data provided by JCPDS card No. 04–0783. Further, as the annealing temperature increases up to 800 °C, distinct differences in the Mn oxide phases are revealed. The diffraction peaks of Mn y O z in the composites annealed at 400 °C, occurring at 2θ = 20.5, 29.4, and 44.0°, are indicative of the tetragonal phase of Mn3O4 as referenced by JCPDS Card No. 00-024-0734. This observation suggests the coexistence of Mn2O3 and MnO within these samples. However, at 800 °C, a pronounced molecular transformation from Mn2O3 to MnO occurs, which is attributed to thermal reduction. The transition from Mn2O3 to MnO is expected to improve electrochemical performance by amplifying electrical conductivity and oxygen vacancy density, resulting in quicker charge transfer and ion diffusion. The lower oxidation state of MnO compared to Mn2O3 is capable of promoting more reversible redox processes, increasing specific capacitance and cycle stability. This phase shift also improves structural flexibility, allowing the electrode to tolerate several charge–discharge cycles with minimum degradation. The thermal treatment process also promotes the formation of metallic Ni, as indicated by the appearance of a diffraction peak, corresponding to the (200) plane of Ni. This peak is exclusively observed in the C–Ag/NiO x @Mn y O z @800 °C sample, confirming the influence of high-temperature annealing on structural modification and phase evolution.
3.2. X-ray Photoelectron Spectroscopy (XPS) Analysis
To gain a comprehensive understanding of the electronic structure and elemental composition of the composites, XPS analyses were conducted. As depicted in Figure c, the XPS survey spectrum reveals distinct signals corresponding to Mn, Ag, Ni, C, and O, affirming their successful incorporation into the composite matrix. In Figure d, the high-resolution XPS spectrum of Mn 2p for the C–Ag/NiO x @Mn y O z @800 °C sample was deconvoluted into three peaks located at 641.4, 643.1, and 646.7 eV, which are ascribed to Mn2+, Mn3+, and Mn4+ species, respectively. The high peak intensity of the Mn2+ species demonstrates the prevalence of MnO within the sample, aligning well with the XRD findings.
In Figure e, the high-resolution XPS spectrum of Ni 2p is illustrated, which comprises two spin–orbit doublets at binding energies of 855.7 and 874.2 eV, with a spin energy separation of 18.5 eV, corresponding to the Ni 2p3/2 and Ni 2p1/2 components. Satellite peaks detected at 861.3 and 879.7 eV are credited to the shakeup phenomena of Ni 2p3/2 and Ni 2p1/2. A meticulous peak fitting analysis of the Ni 2p3/2 component reveals two peaks at 855.7 and 857.9 eV, consistent with the binding energy values associated with Ni2+ and Ni3+ species. The presence of atomic Ni at 853.3 eV on the surface of the C–Ag/NiO x @Mn y O z @800 °C sample results from the thermal reduction of NiO at elevated annealing temperature. In Figure S1, the comparative XPS spectra of Mn 2p of C–Ag/NiO x @Mn y O z @400 °C and C–Ag/NiO x @Mn y O z @800 °C samples are seen to exhibit different patterns, signifying the impact of thermal tuning derived from high-temperature annealing.
In Figure f, the XPS spectrum of Ag 3d is presented, showing two prominent peaks at 368.2 and 374.2 eV, attributed to Ag 3d5/2 and Ag 3d3/2 of metallic silver (Ag0). This observation highlights the dominance of metallic silver within the sample. In Figure g, the C 1s spectrum is shown, providing additional chemical information, with peaks observed at 289.0, 287.6, 286.4, and 284.9 eV, which are ascribed to O–C–O, C = O, C–O, and C–C bonds, respectively. In Figure h, the O 1s spectrum reveals peaks at 532.9, 531.4, and 529.9 eV, which conform to adsorbed oxygen, oxygen associated with defect sites (O2-containing species), and metal–oxygen bonds (O–Mn or O–Ni). , The combined results from XRD and XPS analyses verify the successful integration of metallic silver (Ag) into the HHC-supported NiO x and Mn y O z nanostructure.
3.3. Morphology Characterization
In Figure , a comprehensive analysis of the surface morphology and compositional attributes of the synthesized materials, utilizing FE-SEM, EDS, and HR-TEM techniques, is presented. In Figure a,b, the FE-SEM micrographs of HHC are depicted, exhibiting a porous nanostructure essential for enhancing surface area and promoting electrochemical performance. In Figure c, EDS elemental mapping is illustrated, confirming the consistent distribution of Ag, Mn, Ni, O, and C within the composite, affirming their successful incorporation. In Figure S2, the EDS spectra and percentage weight composition of all the constituent elements are given. Figure S3 presents the relative concentrations of constituent metals (Ag, Mn, and Ni) in the samples, as determined by inductively coupled plasma (ICP) analysis. The results indicate a high degree of compositional homogeneity across the samples.
3.
FE-SEM Micrographs and EDS mapping of synthesized materials: (a, b) FE-SEM micrographs of HHC, (c) EDS mapping of Ag/NiO x @Mn y O z @800 °C, (d) FE-SEM micrograph of Ag/NiO x @Mn y O z , (e) FE-SEM micrograph of C–Ag/NiO x @Mn y O z , (f, g) FE-SEM micrographs of C–Ag/NiO x @Mn y O z @400 °C, (h, i) FE-SEM micrographs of C–Ag/NiO x @Mn y O z @800 °C, (j, k) HR-TEM micrograph of C–Ag/NiO x @Mn y O z @400 °C, and (l) HR-TEM micrographs of C–Ag/NiO x @Mn y O z @800 °C.
In addition, In Figure d shows the compact structure of the Ag/NiO x @Mn y O z composite synthesized without HHC. The absence of HHC results in a more aggregated and less crystalline morphology.
Under identical conditions, the incorporation of HHC with Ag/NiO x @Mn y O z produces a more loosely packed and crystalline architecture (Figure e), as evident in the more pronounced XRD peaks associated with the C–Ag/NiO x @Mn y O z sample. This enhancement in structural characteristics emphasizes the critical role of HHC as an effective support matrix for the composite. In addition, In Figure f,g, the formation of hexagonal NiO crystals attached to the Mn3O4 nanosheets, when annealed at 400 °C, is observed. Notably, Figure h,i reveal substantial impregnation of Ag and Ni on MnO as the partial reduction of Mn3O4 to MnO occurs at 800 °C. Furthermore, HR-TEM analysis offers a detailed visualization of the crystalline nanostructures. In Figure j,k, the overlapping hexagonal NiO molecules formed at 400 °C are highlighted. In Figure l, the uniform distribution of Ag nanospheres and Ni nanoparticles across the MnO surface is demonstrated.
3.4. Surface Area and Particle Size Measurements
In Figure a, a characteristic type I Brunauer–Emmett–Teller (BET) isotherm, corresponding to the HHC sample, is illustrated. This isotherm portrays a remarkable specific surface area (SSA) of 2605.93 m2 g–1 with an average pore size of 2.18 nm. The negligible hysteresis observed in the isotherm indicates a predominantly closed-ended pore structure with similar adsorption and desorption mechanisms. In Figure b, a comparative BET analysis conducted on the composites, with and without HHC support, is presented. Both samples exhibit type IV isotherms, which are indicative of mesoporous materials. Significantly, the composite incorporating HHC exhibits a substantially higher SSA (91.4 m2 g–1): approximately double that of the sample synthesized without HHC (40.6 m2 g–1). This outcome highlights the critical function of HHC as a conductive supporting material, capable of boosting the specific surface area and electrochemical properties of the composites. The corresponding pore diameters measure 7.08 and 11.09 nm for the samples, with and without HHC, respectively. These results further demonstrate the beneficial impact of HHC inclusion.
4.
BET and particle size analyses (PSA) of the composites: (a) BET adsorption–desorption isotherm for HHC, (b) Combined BET adsorption–desorption isotherms for Ag/NiO x @Mn y O z and C–Ag/NiO x @Mn y O z composites, (c) PSA of Ag/NiO x @Mn y O z , (d) PSA of C–Ag/NiO x @Mn y O z , (e) PSA of C–Ag/NiO x @Mn y O z @400 °C, and (f) PSA of C–Ag/NiO x @Mn y O z @800 °C.
In Figure c–g, an analysis of particle size is shown. Average particle sizes of 1.26, 2.08, 2.35, and 12.59 μm for Ag/NiO x @Mn y O z , C–Ag/NiO x @Mn y O z @, C–Ag/NiO x @Mn y O z @400 °C, and C–Ag/NiO x @Mn y O z @800 °C are displayed. These findings are consistent with the FE-SEM analysis results, affirming that the agglomeration of Ag and Ni nanoparticles becomes increasingly prominent at elevated annealing temperatures. The trend of increasing particle size, as observed in Figure c–f, reveals a clear correlation between annealing temperature and particle growth. This outcome indicates that thermal treatment facilitates the agglomeration and coalescence of metallic nanoparticles, particularly Ag and Ni. Such particle growth at elevated temperatures can significantly enhance electrochemical performance by reducing interfacial resistance and improving electron transport pathways. Larger particles typically possess a reduced number of grain boundaries, which act as barriers to charge mobility. Consequently, the observed increase in particle size at elevated temperature thermal treatment can contribute to a more continuous and conductive framework within the electrode material with improved crystallinity and structural stability, crucial for long-term cycling and capacitive retention.
In Figure h, a broad particle size distribution is observed. This trend can be attributed to a combination of rapid particle growth dynamics and concurrent phase transformations within the composite’s constituents. At such a high annealing temperature, a diverse growth rate of particle sizes is expected, especially in a ternary composite, as some particles grow more extensively than others due to local energy and compositional differences. Moreover, XRD and subsequent ex-situ XAS analyses confirm that at 800 °C, the formation of a new molecular phase, MnO, becomes evident alongside the reduction of NiO x to atomic Ni. These transformations involve substantial structural rearrangements that contribute to nonuniform growth rates and morphological heterogeneity. The emergence of MnO and metallic Ni not only introduces new nucleation sites but also alters the thermodynamic and kinetic growth behavior of the existing phases, resulting in varied crystallite sizes.
3.5. Electrochemical Measurements in Half Cell
In Figure a–e, for all samples, the comparative cyclic voltammetry (CV) curves are delineated. The CV curves were measured within a specified voltage window of 1.0 V for HHC and 0.6 V for the composite materials, respectively. In Figure S4, a three-electrode configuration used for electrochemical analysis is depicted. As shown in Figure a, the CV curves for HHC display nearly rectangular profiles, characteristic of EDLC typically observed in carbon-based materials. In contrast, in Figure b–e, the CV curves of the composite samples reveal the presence of two prominent oxidation and reduction peaks, particularly at lower scan rates. This observation indicates the occurrence of at least two simultaneous redox reactions within the system, improving overall capacitance.
5.
Three-electrode’s CV and GCD features of the electrode: (a) CV of the HHC, (b) CV of Ag/NiO x @Mn y O z , (c) CV of C–Ag/NiO x @Mn y O z , (d) CV of C–Ag/NiO x @Mn y O z @400 °C, (e) CV of C–Ag/NiO x @Mn y O z @800 °C, (f) Specific capacitance of the electrodes at various scan rate, (g) GCD plot of HHC, (h) GCD curve of Ag/NiO x @Mn y O z , (i) GCD curve of C–Ag/NiO x @Mn y O z , (j) GCD curve of C–Ag/NiO x @Mn y O z @400 °C, (k) GCD curve of C–Ag/NiO x @Mn y O z @800 °C, and (l) Specific capacitance of the electrodes at various specific currents.
In Figure f, HHC exhibits a Cs of 418 F g–1 at a scan rate of 5 mV s–1. While all composites demonstrated impressive specific capacitances, the highest Cs of 1770 F g–1 at 5 mV s–1 was achieved by the Ag/NiO x @Mn y O z @800 °C sample. This exceptional performance can be attributed to the presence of impregnated, free-standing redox-active Ni atoms on the surface of HHC-supported porous MnO molecules, which facilitate enhanced charge storage. In Figures S5 and S6, an average b-value of 0.54 is denoted, signifying a combination of capacitive and diffusion-controlled processes. Moreover, Figure S7 illustrates the proportion of current signal contributed by capacitive and diffusion processes at varying sweep rates. In Table S1, a comparison of HHC with other bioderived electrode materials is provided.
Galvanostatic charge–discharge (GCD) analysis of HHC, conducted at a current density range of 1–10 A g–1, reveals symmetrical and triangular profiles with a nearly steady specific capacitance of 271 F g–1 at 1 A g–1. During charge–discharge processes, the GCD curves of the composite materials display distinct two-stage profiles, indicating dual redox-active transitions associated with Mn2+/Mn3+ and Ni0/Ni2+ within the electrode (Figure g–k). In Figure l, the specific capacitance values for each composite sample are summarized. In Table , a comparative analysis of the electrochemical performance of this study against recent literature is provided.
1. Comparative Electrochemical Outputs of This Study with Recently Reported Literature.
| material | synthesis method | electrolyte | specific capacitance Cs(F g –1 ) | energy density(Wh kg–1)/Cs retention | refs |
|---|---|---|---|---|---|
| MnO2/CNTs | electrodeposition | 0.5 M Na2SO4 | 3310 | –/65.4% | |
| Na+-MnO2 | solvothermal | 1 M KOH | 655 | 27/95% | |
| Cu-MnO x | electrodeposition | 1 M Na2SO4 | 92.3 | 25.13/44.9% | |
| MnO2@Ti3C2T x | solution immersion | 1 M KOH | 324.1 | 30.8/95.98% | |
| MnO2/NiCo | hydrothermal | 1 M KOH | 2572.7 | –/78.9 | |
| MnO2/MnS | hydrothermal | 977.6 | 122.23/91.5 | ||
| Al-doped δ-MnO2 | hydrothermal | 6 M KOH | 174 | 18.5/90% | |
| MnO2/CNTs | chemical reduction | 617.7 | 111/71.4% | ||
| PCDWT/MnO2/PANI | pyrolysis | trifluoroacetic acid (TFA) | 369.6 | 36.7/87.7% | |
| C–Ag/NiO x @Mn y O z @800 °C | hydrothermal | 2 M KOH | 1770 | 37.53/82% | this study |
In Figure a, a schematic representation of the redox reactions occurring at the C–Ag/NiO x @Mn y O z @800 °C/KOH interface is presented. The illustration highlights the dual role of Mn and Ni species in promoting OH– reduction when a potential difference is applied across the cell terminals. In Figure d, the electrochemical impedance spectroscopy (EIS) results are presented through Nyquist plots. The C–Ag/NiO x @Mn y O z @800 °C sample exhibits the lowest equivalent series resistance (R1) of ∼0.48 Ω and a charge transfer resistance (R2) of 1.41 Ω, which is significantly lower than that of the sample annealed at 400 °C (Table S2). This observation is consistent with the fact that the infused metallic Ni derived at 800 °C acts as a superior electrical conductor compared to the NiO present in the composite material annealed at 400 °C.
6.
Reaction mechanisms, diffusion coefficient, charge transfer coefficient, and reaction rate constant evaluation plots: (a) Simultaneous redox reactions occurring at the cathode, (b) Relationship between peak current (A) and υ0.5, (c) Relationship between peak potentials (E p) and ln (υ), (d) EIS analysis of the composites, and (e) EIS-fitted equivalent circuit.
Further, both C–Ag/NiO x @Mn y O z @400 °C and C–Ag/NiO x @Mn y O z @800 °C samples exhibit nearly identical spectral profiles, characterized by low electrochemical impedance across R1, R2, and double-layer resistance (R3), as detailed in Table S2. These findings indicate that increasing the annealing temperature from 400 to 800 °C does not negatively affect the material’s conductivity. Instead, the higher annealing temperature amplifies both the material’s electrical conductivity and charge transport capabilities by generating oxygen vacancies through thermal reduction. In Figure e, the fitted EIS equivalent circuit is illustrated, providing a visual representation of the circuit elements that impede current flow. The excellent impedance feature of the C–Ag/NiO x @Mn y O z @800 °C sample highlights its potential for efficient charge transfer, thereby improving the overall electrochemical performance of the composite material.
The mass loading ratio of both electrodes, which forms the basis of electrochemical performance indices, was estimated at 0.39, using eq :
| 1 |
where C m+ and C m– represent the capacitance of the positive and negative electrodes, respectively, corresponding to their voltage windows ΔV + and ΔV –. By applying Laviron equations, good linear regressions were observed in the plots of both anodic and cathodic peak potentials against ln (υ), as shown in Figure b, yielding the following linear eqs and :
| 2 |
| 3 |
By utilizing the gradient of either E pc (V) or E pa (V) in eqs and and assuming α = 0.5, the number of electrons transferred (n) is estimated at ∼2, indicating a two-electron reduction process at the cathode. The charge transfer coefficient (α) was subsequently evaluated at 0.56. The heterogeneous electron transfer rate constant (k s) was determined using eq , as given by Reddy et al. In Table S3, for all scan rates, the values of the heterogeneous reaction rate constant (k s) are summarized. Notably, k s increased with increasing scan rate, confirming the rapid redox kinetics at higher sweep rates:
| 4 |
| 5 |
| 6 |
where E° is the formal potential, R is the universal gas constant, T is the absolute temperature, α is the electron transfer coefficient, n is the electron transfer number, and F is the Faraday constant. In addition, the diffusion coefficient (D), which quantifies the ion transport kinetics within the electrode interface, was determined using the Randles-Sevcik equation, as expressed in eq . By analyzing the linear relationship between the peak current (i p) and the square root of the scan rate (υ), the diffusion coefficient was estimated to be 1.41 × 10–4 cm2 s–1 (Figure c). This value highlights the efficient ion mobility within the electrode, reinforcing the material’s suitability for high-performance electrochemical applications:
| 7 |
where n = number of transferred electrons, A = area of electrode (cm2), C = bulk concentration (mol cm–3) D = diffusion coefficient (cm2 s–1), and v = scan rate (mV s–1).
The suspected redox reactions occurring within the system can be represented as
| 8 |
where x, y, and z are positive integers.
3.6. Electrochemical Measurements of AHSC
In Figure a, a comparative cyclic voltammogram of HHC and C–Ag/NiO x @Mn y O z @800 °C, at a scan rate of 5 mV s–1, is presented. The substantially enhanced peak current achieved by the composite material, in comparison to the HHC anode, is highlighted. Notably, the current signals of both materials effectively complement one another, enabling a broader voltage window of 1.5 V. In Figure b, CV performed across a 1.5 V potential window at sweep rates ranging from 5 to 100 mV s–1, demonstrates a consistent and stable pattern, indicating the robustness of AHSC. This stability is attributed to the synergistic interaction between the EDLC mechanism at the HHC anode and the faradaic redox reactions occurring at the C–Ag/NiO x @Mn y O z @800 °C cathode. AHSC exhibits a maximum Cs of 150.7 F g–1 at a scan rate of 5 mV s–1, as shown in Figure S8.
7.
Electrochemical analyses of AHSC: (a) Combined CV curves of the HHC and C–Ag/NiO x @Mn y O z @800 °C at 30 mV s–1, (b) CV curves of AHSC, (c) GCD curve of AHSC, (d) AHSC assembling and illumination of 3.2 V LED bulb, (e) Capacitance retention after 5000 charge–discharge cycles, and (f) Specific capacitance, energy density, and power density values of AHSC.
In Figure c, GCD tests were conducted over a range of current densities from 3 to 15 A g–1, within the same potential window. The GCD curves, particularly at lower current densities, exhibit nontriangular profiles, confirming the supercapacitive behavior of the AHSC device. Figure c exhibits an asymmetric feature due to different mechanisms of charge and discharge. In supercapacitors, discharging typically occurs faster than charging, reflecting an inherent asymmetry in their operation. In AHSC, during charging, OH– and K+ migrate and adsorb onto the electrode surfaces to undergo redox reactions, a process slowed by ion diffusion and the porous structure of the electrodes. In contrast, discharging happens more rapidly as the electrostatic forces holding the ions weaken, allowing for fast ion desorption and electron flow. Factors like internal resistance and electrode polarization also accelerate voltage drop during discharge, further reducing its duration compared to charging. However, the electrochemical outputs of C–Ag/NiO x @Mn y O z @800 °C/HHC AHSC are irrespective of the asymmetric feature of the charge–discharge curves.
Furthermore, in Figure d, the practical assembly of AHSC and its capability to successfully illuminate a 3.2 V LED bulb, is shown. In Figure S9, the practical applicability of the device is highlighted. The exceptional performance of this device is attributable to the availability of high volume of oxygen vacancy observed in C–Ag/NiO x @Mn y O z @800 °C. Figure S10 shows the spectra an electron paramagnetic resonance (EPR) conducted to assess the presence of oxygen vacancies. Initial evidence of vacancy formation was detected in C–Ag/NiO x @Mn y O z @400 °C. Using a comparable mass loading of approximately 94–98 mg across samples, the spectrum of the 800 °C-annealed sample was significantly distorted and out of range, suggesting the presence of a high concentration of unpaired electrons due to oxygen vacancies. Consequently, the mass loading for this sample was reduced to 1.5 mg to obtain a symmetrical and interpretable spectrum. The EPR signals for C–Ag/NiO x @Mn y O z @800 °C and @400 °C appeared at g-values of 2.0 and 2.002, respectively–both characteristic of unpaired electrons resulting from oxygen vacancies.
In Figure S11, the EIS spectra for both precycling and postcycling analyses through Nyquist plots are presented. Notably, before cycling, EIS analysis of the AHSC device shows an R1 value of 0.48 Ω, with minor increases in R2 and R3, recorded at 3.66 and 11.48 Ω, respectively (Table S3). This result underscores the stability of the C–Ag/NiO x @Mn y O z @800 °C sample and the excellent electrochemical interaction between the electrode and electrolyte as well as between the electrode material and the NF current collector. The minimal semicircle observed in the high-frequency region signifies low charge transfer resistance, corroborated by the recorded value of 3.66 Ω obtained via equivalent circuit fitting.
In the low-frequency region, the nearly vertical profile of the spectrum suggests enhanced ion diffusion capability within the electrode material. As expected, the postcycling EIS spectrum displayed increased impedance values for all three components, with R1, R2, and R3 measured at 0.75, 3.9, and 25.6 Ω, respectively. Such an increase is likely due to irreversible redox reactions, the potential formation of a passivation layer, and electrolyte depletion, which can impede the redox activity of the active species at the cathode. This trend emphasizes the impact of long-term cycling on the electrode’s electrochemical performance. As shown in Figure e, cyclic stability tests over 5000 charge–discharge cycles demonstrated remarkable durability, with AHSC retaining 82% of its initial capacitance after 5000 cycles. This outstanding performance confirms AHSC’s robust long-term operational stability and reliability.
In Figure f, Cs, energy density, and power density values achieved at various current densities are summarized. It is significant that AHSC exhibited an impressive energy density of 37.53 Wh kg–1 and a power density of 2251.8 W kg–1 at 3 A g–1. The electrochemical performance of the C–Ag/NiO x @Mn y O z @800 °C/HHC composite surpassed several previously reported values for manganese oxide-based materials, indicating its superior performance and potential for practical applications.
3.7. Ex-situ and Operando XAS Analysis of C–Ag/NiO x @Mn y O z @800 °C Electrode
In Figure S12a, the Ag L-edge XANES spectra provide critical insights into the electronic state and local structural environment of Ag within the C–Ag/NiO x @Mn y O z @400 °C and C–Ag/NiO x @Mn y O z @800 °C samples, compared to reference materials: Ag foil, Ag2O, and AgO. The spectra of both samples display pre-edge and postedge features analogous to those of Ag foil, indicating the predominance of metallic silver (Ag0) within the composites. In both samples, however, a subtle shift of the absorption edge toward lower energy, relative to the Ag foil, was detected. In the local electronic environment of Ag atoms, this shift indicates minor perturbations, possibly induced by structural modifications in the Ag crystal lattice.
To further elucidate the oxidation state and phase composition of Ag, linear combination fitting (LCF) of the XANES spectra for the C–Ag/NiO x @Mn y O z @400 °C and C–Ag/NiO x @Mn y O z @800 °C samples was carried out over an energy range of 3331–3381 eV, using ATHENA software. The fitting procedure, conducted without floating the E0 parameter and constrained to a sum of one, yielded high-quality fits with R-factors below 0.05. LCF analysis indicated that the C–Ag/NiO x @Mn y O z @400 °C sample comprised ∼81% Ag0 (from Ag foil) and 19% Ag2O. In contrast, the C–Ag/NiO x @Mn y O z @800 °C sample consisted of 74% Ag0 and 26% AgO. This outcome demonstrates the subtle variation of Ag between the two samples attributable to different annealing conditions.
Initially, the hydrothermal synthesis yielded a composite predominantly featuring the Mn3O4 phase, as confirmed by XRD analysis. Subsequent thermal treatments induced controlled molecular transitions that were systematically examined. At 400 °C, the material transformed from Mn3O4 into Mn2O3, as revealed by the ex-situ XAS spectra at the Mn K-edge region. This intermediate phase is crucial for facilitating subsequent reduction and enhancing redox-active sites.
In Figure b, ex-situ Mn K-edge XANES analysis of the C–Ag/NiO x @Mn y O z @800 °C sample reveals spectral features closely resembling those of MnO, with pre-edge and postedge characteristics as well as excitation energy (∼6554.3 eV), mirroring the standard MnO reference. This outcome signifies that at 800 °C, the desired MnO phase became predominant, with a quantified phase composition of 95.6% determined via linear combination fitting (LCF) of the XANES spectra for C–Ag/NiO x @Mn y O z @800 °C. At this temperature, the MnO phase was also clearly detected in the XRD pattern, showing a distinct diffraction peak of MnO at a 2θ angle of 40.5°, which corresponds to the (200) plane (Figure b). This progressive thermal reduction process not only enables precise control over the redox-active species but also facilitates the formation of metallic Ni0, thus establishing the synergistic redox interplay between Mn2+/Mn3+ and Ni0/Ni2+. This synergy is essential in amplifying electrochemical performance, as it promotes improved charge storage capacity and cycling stability of the electrode material.
8.
Ex-situ XAS analysis and operando charge-storage mechanism assessment via synchrotron light: (a) Assemblage and operando XAS study of AHSC, (b) Normalized ex-situ Mn K-edge XANES spectra, (c) Mn K-edge XANES spectra recorded via operando XAS technique during 7 complete charging cycles, (d) Normalized ex-situ Ni K-edge XANES spectra, (e) Ni K-edge XANES spectra recorded via operando XAS technique during 7 complete charging cycles, (f) FT Mn K-edge EXAFS spectra of C–Ag/NiO x @Mn y O z @800 °C, and (g) FT Ni K-edge EXAFS spectra of C–Ag/NiO x @Mn y O z @800 °C.
These results deviate slightly from that of the C–Ag/NiO x @Mn y O z @400 °C sample; the Mn K-edge spectrum mimics the patterns of the standard Mn2O3 reference (Figure S12b). Moreover, the Ni K-edge XANES spectra of the C–Ag/NiO x @Mn y O z @800 °C sample exhibit remarkable similarity to the spectral patterns of the reference Ni foil, indicating the predominance of metallic Ni within the sample. This observation aligns well with the XRD findings, which display a prominent peak of Ni at a 2θ angle of 57.1°, corresponding to the (200) plane of metallic Ni. It is noted that the C–Ag/NiO x @Mn y O z @400 °C sample exhibits an XANES spectral pattern identical to that of the NiO reference (Figure S12c).
Fourier transform (FT) EXAFS analysis of Mn K-edge spectra further supports the presence of MnO and metallic Ni within the C–Ag/NiO x @Mn y O z @800 °C sample, with a recorded Mn–O bond length of 1.68 Å, consistent with the MnO reference (Figure f). Similarly, the FT EXAFS of Ni K-edge spectra highlight congruent patterns between the C–Ag/NiO x @Mn y O z @800 °C sample and the Ni foil reference, both displaying a Ni–TM bond length of 2.17 Å.
To elucidate the charge-storage mechanism and redox activity of both Mn and Ni components within the cathode material of the AHSC, synchrotron operando XAS was conducted at Beamline 2.2 of the Synchrotron Light Research Institute (SLRI), Thailand (Figure S13). Measurements were conducted using a custom-designed cell optimized for transmission XAS mode (Figure a). To ensure the acquisition of high-quality XAS data, the cathode material loading was meticulously controlled at ∼ 3.0 mg cm–2. The collected XAS data included XANES spectra, providing critical insights into the oxidation states and local electronic structure of the active materials. Data processing and interpretation were carried out using Athena–Demeter (version 0.9.26): a sophisticated software suite for XAS analysis.
The operando XAS technique was employed to investigate the structural evolution of Mn and Ni within the C–Ag/NiO x @Mn y O z @800 °C samples under real-cell conditions. Mn K-edge XANES spectra reveal a noticeable increase in absorption edge energy upon charging, corresponding to the oxidation state evolution of Mn (Figure c). When compared to standard references of MnO, Mn2O3, and MnO2, the absorption K-edge energy of C–Ag/NiO x @Mn y O z @800 °C, during charging, overlaps with that of Mn2O3. This observation suggests a valence transition of Mn from ∼+2 to +3, indicating that the charge-storage redox mechanism is primarily governed by the valence evolution of Mn within the solid phase.
While the white-line region in the Mn K-edge spectrum does not directly provide information about Mn 3d orbitals, its intensity serves as a useful indicator of 3d orbital activity, particularly when Mn is coordinated with O atoms. During charging, the increased white-line intensity points to a higher number of unoccupied Mn 3d orbitals due to electron transfer to electrolyte species. This observation supports the gradual formation of an intercalated state where mobile ions from the electrolyte are incorporated into the C–Ag/NiO x @Mn y O z @800 °C layers. In general, the pre-edge peak arises from both the quadrupole transition and the dipole transition associated with d–p hybridization. In the case of Ni foil, the pre-edge peak is primarily attributed to the direct dipole transition of Ni 1s → O 2p, resulting from strong hybridization, which is notably absent in the K-edge spectra of the reference NiO and C–Ag/NiO x @Mn y O z @800 °C (Figure e). The distinct alignment of the white line between C–Ag/NiO x @Mn y O z @800 °C and the standard NiO reference indicates the extensive accommodation of Ni 3d unoccupied orbitals by O atoms. This finding implies a gradual transition in Ni’s oxidation state from 0 to +2, resulting from a partial reduction reaction during the charging process.
3.8. Aging Effects on Electrochemical Properties of AHSC
To evaluate the aging effects on AHSC, postcycling electrochemical characterization was conducted after two months, including CV and GCD analyses. Results are presented in Figure S14a–S14d. Here, critical insights into the long-term electrochemical stability and performance retention of the device are provided, including Cs at different scan rates and current densities. The CV profile of AHSC is seen to retain its characteristic oxidation and reduction peaks, confirming the sustained redox activity of Mn2+/Mn3+ and Ni0/Ni2+ species within the electrode. As shown in Figure S14a, the CV curves maintain stable redox behavior across all scan rates, ranging from 5 to 100 mV s–1, demonstrating the preserved electrochemical activity of the redox-active materials over time. The recorded Cs value of 51.2 F g–1, at a scan rate of 5 mV s–1, indicates the robustness of the electrode material (Figure S14b).
Likewise, in Figure S14c, the GCD curves retain their characteristic nontriangular profiles, further validating the supercapacitive behavior of AHSC. After two months, the device achieved a specific capacitance of 47 F g–1 at a current density of 3 A g–1 (Figure S14d). However, compared to the freshly prepared cell, the discharge time decreased across all current densities. This decline suggests a gradual decrease in energy storage capacity, which can be attributed to natural aging effects such as minor structural modifications within the electrode or electrolyte depletion over time. Overall, the electrochemical stability of AHSC remained well-preserved, reinforcing its potential for long-term applications.
4. Conclusions
In summary, we successfully developed a high-performance AHSC electrode by integrating a bimetallic Ag/Ni–MnO nanostructure with HHC support. The incorporation of HHC as a sustainable conductive scaffold significantly enhanced the composite’s surface area, electronic conductivity, and electrochemical stability. Comprehensive structural and spectroscopic analyses, applying FE-SEM, HR-TEM, XRD, XPS, and operando XAS revealed that the 800 °C thermal treatment induces a phase transformation of Mn2O3 to MnO and the partial reduction of NiO to metallic Ni as well as the generation of oxygen vacancies. These thermally driven changes give rise to a highly conductive framework and facilitate dual redox activity (Mn2+/Mn3+ and Ni0/Ni2+) within the electrode. As a result, the optimized Ag/NiO x @Mn y O z @800 °C–HHC composite achieved exceptional Cs of 1770 F g–1 at 5 mV s–1 in three-electrode measurements. When assembled into an AHSC, using an HHC anode and the C–Ag/NiO x @Mn y O z @800 °C cathode, the device delivered a Cs of ∼ 150.7 F g–1 at 5 mV s–1, an energy density of 37.53 Wh kg–1 at a discharge current of 3 A g–1, and a corresponding power density of 2251.8 W kg–1. AHSC also demonstrated outstanding cycling stability, retaining about 82% of its initial capacitance after 5000 cycles. These findings provide a valuable pathway for designing next-generation energy storage devices with sustainable materials. In essence, they reinforce the potential of HHC-supported multicomponent composites for durable SCs.
Supplementary Material
Acknowledgments
This research is funded by the Thailand Science Research and Innovation Fund Chulalongkorn University (IND_FF_68_169_2100_028). The Program Management Unit for Human Resources and Institutional Development, Research, and Innovation (B49G680109) is acknowledged. A.S. expresses gratitude to Chulalongkorn University’s Graduate Scholarship Program for ASEAN or non-ASEAN countries as well as the Chulalongkorn University’s Graduate School for the Overseas Research Experience Scholarship. K.T. expresses gratitude to the Council for Science, Technology and Innovation (CSTI), Cross-ministerial Strategic Innovation Promotion Program (SIP), the third period of SIP, Creating a Materials Innovation Ecosystem for Industrialization Grant Number JPJ012307. M.Y.C would like to acknowledge the funding support from the Korea Basic Science Institute (National research Facilities and Equipment Center) Grant funded by the Ministry of Education (No. RS-2024-00434932) and National Research Foundation of Korea (NRF) (2022R1A2C2010686).
The authors declare that all data supporting the findings of this study are available within the paper and its Supporting Information file.
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsami.5c07064.
Additional experimental details, comprising the descriptions of morphological characterization equipment, photos of the experimental setup, and additional experimental results (PDF)
The authors declare no competing financial interest.
References
- Poolnapol L., Kao-ian W., Somwangthanaroj A., Mahlendorf F., Nguyen M. T., Yonezawa T., Kheawhom S.. Silver Decorated Reduced Graphene Oxide as Electrocatalyst for Zinc–Air Batteries. Energies. 2020;13(2):462. doi: 10.3390/en13020462. [DOI] [Google Scholar]
- Radenahmad N., Khezri R., Mohamad A. A., Nguyen M. T., Yonezawa T., Somwangthanaroj A., Kheawhom S.. A Durable Rechargeable Zinc-Air Battery via Self-Supported MnOx-S Air Electrode. J. Alloys Compd. 2021;883:160935. doi: 10.1016/j.jallcom.2021.160935. [DOI] [Google Scholar]
- Lao-atiman W., Julaphatachote T., Boonmongkolras P., Kheawhom S.. Printed Transparent Thin Film Zn-MnO2 Battery. J. Electrochem. Soc. 2017;164(4):A859–A863. doi: 10.1149/2.1511704jes. [DOI] [Google Scholar]
- Bachankar S., Malavekar D., Lokhande V., Ji T.. Empowering Supercapacitors: Strategic B-Site Modulations of Transition Metal Atoms in Perovskite Oxide Based Electrodes. J. Alloys Compd. 2024;1003:175708. doi: 10.1016/j.jallcom.2024.175708. [DOI] [Google Scholar]
- Łańcucki Ł., Pasierb P.. Effect of Microstructure on Performance and Working Mechanism of MnO2-PEDOT Supercapacitors Based on Nonaqueous Electrolytes. J. Energy Storage. 2025;106:114824. doi: 10.1016/j.est.2024.114824. [DOI] [Google Scholar]
- Lv Y., He J., Yang Y., Huang M., He D., Wang Y.. Facile Synthesis of MnO2@Ti3C2Tx Composite Electrodes for Superior Performance Supercapacitor. J. Solid State Chem. 2025;344:125198. doi: 10.1016/j.jssc.2025.125198. [DOI] [Google Scholar]
- Guo Y., Chen J., Liang Z., Su X., Sun M., Tang J., Li Z., Dang D., Yu L.. Hierarchical Porous MnO2/3D Graphene Oxide/Carbon Cloth Used as Practical Mass-Loading Supercapacitor Electrodes. J. Alloys Compd. 2025;1010:177125. doi: 10.1016/j.jallcom.2024.177125. [DOI] [Google Scholar]
- Wang M., Liu X., Wu X.. Surface Selenization of NiCo-Layered Double Hydroxide Nanosheets for High-Performance Supercapacitors. Batteries. 2023;9(1):49. doi: 10.3390/batteries9010049. [DOI] [Google Scholar]
- Sanni A., Govindarajan D., Nijpanich S., Limphirat W., Mohamad A. A., Theerthagiri J., Choi M. Y., Kheawhom S.. Unveiling a Paradigm Shift in Supercapacitor Dynamics: γ-Al2O3-Infused ZnO Nanorods with Redox-Active K4Fe(CN)6 Alkaline Electrolytes. J. Alloys Compd. 2025;1010:177892. doi: 10.1016/j.jallcom.2024.177892. [DOI] [Google Scholar]
- Sanni A., Govindarajan D., Nijpanich S., Limphirat W., Theerthagiri J., Choi M. Y., Kheawhom S.. Al-Doped ZnO@CuO Nanoflower/Nanorod Heterostructures on CNTs as High-Performance Supercapacitor Electrodes in Redox-Supporting Electrolytes. J. Energy Storage. 2025;109:115184. doi: 10.1016/j.est.2024.115184. [DOI] [Google Scholar]
- Zhang Q., Liu S., Huang J., Fu H., Fan Q., Zong H., Guo H., Zhang A.. In Situ Selective Selenization of ZIF-Derived CoSe2 Nanoparticles on NiMn-Layered Double hydroxide@CuBr2 Heterostructures for High Performance Supercapacitors. J. Colloid Interface Sci. 2024;655:273–285. doi: 10.1016/j.jcis.2023.11.008. [DOI] [PubMed] [Google Scholar]
- Zhang A., Zhang Q., Fu H., Zong H., Guo H.. Metal-Organic Frameworks and Their Derivatives-Based Nanostructure with Different Dimensionalities for Supercapacitors. Small. 2023;19(48):2303911. doi: 10.1002/smll.202303911. [DOI] [PubMed] [Google Scholar]
- Zhang A., Zhang Q., Huang J., Fu H., Zong H., Guo H.. NiMnCo-LDH in-Situ Derived from ZIF-67@ZnO as Self-Supporting Electrode for Asymmetric Supercapacitor Device. Chem. Eng. J. 2024;487:150587. doi: 10.1016/j.cej.2024.150587. [DOI] [Google Scholar]
- Sanni A., Govindarajan D., Nijpanich S., Limphirat W., Tipplook M., Teshima K., Sangaraju S., Kheawhom S.. Designing Dual-Phase ZnO-Al2O3-CuO Nanostructures for Enhanced Supercapacitor Performance. Mater. Res. Bull. 2025;189:113443. doi: 10.1016/j.materresbull.2025.113443. [DOI] [Google Scholar]
- Nechikott A. A., Nayak P. K.. Superior Pseudo-Capacitive Performance of Co-Free P2-Na2/3Mn2/3Ni1/3O2 for Aqueous Dual-Ion Hybrid Supercapacitors. J. Power Sources. 2024;623:235402. doi: 10.1016/j.jpowsour.2024.235402. [DOI] [Google Scholar]
- Li P., Zhu C., Yang L., Shi J., Hu W., Li Z., Wu J., Wang H.. One Single-Atom Mn Doping Strategy Enabling Two Functions of Oxygen Reduction Reaction and Pseudocapacitive Performance. Energy Storage Mater. 2024;71:103639. doi: 10.1016/j.ensm.2024.103639. [DOI] [Google Scholar]
- Zhang B., Jiang T., Zhou X., Fan X., Jia B., Li L.. Unlocking Supercapacitive Energy Storage Potential: Catalyzing Electrochemically Inactive Manganese Oxides to Active MnO2 via Heterostructure Reconstruction. Nano Res. 2024;17(7):5897–5906. doi: 10.1007/s12274-024-6577-6. [DOI] [Google Scholar]
- Liu J., Bao J., Zhang X., Gao Y., Zhang Y., Liu L., Cao Z.. MnO2 -Based Materials for Supercapacitor Electrodes: Challenges, Strategies and Prospects. RSC Adv. 2022;12(55):35556–35578. doi: 10.1039/D2RA06664E. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li X., Lin Z., Wang C., Wang H., Feng S., Li T., Ma Y.. Hierarchical Graphene Oxide/Manganese Dioxide/Cobalt-Nickel Layered Double Hydroxide Electrodes for High Energy Density Asymmetric Flexible Supercapacitor. Chem. Eng. J. 2024;484:149430. doi: 10.1016/j.cej.2024.149430. [DOI] [Google Scholar]
- Ivandini T. A., Ulfah F. F., Nahda D. P. N., Luhur S. P., Sanjaya A. R., Krisnandi Y. K., Sumboja A., Zulfia A., Ryu K.-S., Akbar Z. A.. Hierarchical Rod-like Structure MnO2/NiCo-Layered Double Hydroxide on Nickel Foam for a High-Performance Supercapacitor Electrode. Mater. Res. Bull. 2025;184:113289. doi: 10.1016/j.materresbull.2024.113289. [DOI] [Google Scholar]
- Zia A. W., Anestopoulos I., Bowen L., Panayiotidis M. I., Birkett M.. Precision Depth-Controlled Isolated Silver Nanoparticle-Doped Diamond-like Carbon Coatings with Enhanced Ion Release, Biocompatibility, and Mechanical Performance. Surf. Coat. Technol. 2024;493:131281. doi: 10.1016/j.surfcoat.2024.131281. [DOI] [Google Scholar]
- Aslan N., Kurt M. Ş., Koç M. M.. Silver-Doped Diamond-Like Carbon (DLC:Ag) Nanocomposite Films for Solar Tracking Applications. J. Electron. Mater. 2023;52(4):2807–2818. doi: 10.1007/s11664-023-10243-x. [DOI] [Google Scholar]
- Jing P. P., Gong Y. L., Deng Q. Y., Zhang Y. Z., Huang N., Leng Y. X.. The Formation of the “Rod-like Wear Debris” and Tribological Properties of Ag-Doped Diamond-like Carbon Films Fabricated by a High-Power Pulsed Plasma Vapor Deposition Technique. Vacuum. 2020;173:109125. doi: 10.1016/j.vacuum.2019.109125. [DOI] [Google Scholar]
- Kerli S., Bhardwaj S., Lın W., Gupta R. K.. Silver-Doped Reduced Graphene Oxide/Pani Composite Synthesis and Their Supercapacitor Applications. J. Organomet. Chem. 2023;995:122725. doi: 10.1016/j.jorganchem.2023.122725. [DOI] [Google Scholar]
- Alam A., Saeed G., Lim S.. Screen-Printed Activated Carbon/Silver Nanocomposite Electrode Material for a High Performance Supercapacitor. Mater. Lett. 2020;273:127933. doi: 10.1016/j.matlet.2020.127933. [DOI] [Google Scholar]
- Ascencio-Flores Y. F., Carhuayal-Alvarez S. M., Quiroz-Aguinaga U., Calderon-Zavaleta S. L., López E. O., Ponce-Vargas M., Cardenas-Riojas A. A., Baena-Moncada A. M.. Simultaneous Square Wave Voltammetry Detection of Azo Dyes Using Silver Nanoparticles Assembled on Carbon Nanofibers. Electrochim. Acta. 2023;441:141782. doi: 10.1016/j.electacta.2022.141782. [DOI] [Google Scholar]
- Mariappan A., Dharman R. K., Oh T. H., Prabu S., Chiang K.-Y.. Silver Nanoparticles Embedded in Phosphorus and Nitrogen-Doped Hierarchical Hollow Porous Carbon for Efficient Supercapacitor and Electrocatalytic Water Oxidation. Colloids Surf., A. 2023;671:131579. doi: 10.1016/j.colsurfa.2023.131579. [DOI] [Google Scholar]
- Wang Z., Wang H., Pei D., Wan S., Wang Y., Yu M., Lu H.. Hydrothermal-Deposition Synthesis of MnO2/MnS Nanorods as High-Performance Asymmetric Supercapacitor Positive Electrode Materials. Prog. Nat. Sci.:Mater. Int. 2024;34(5):1057–1065. doi: 10.1016/j.pnsc.2024.07.023. [DOI] [Google Scholar]
- Wu Q., Li F., Sheng H., Qi Y., Yuan J., Bi H., Li W., Xie E., Lan W.. In Situ Fabrication of Hierarchical CuO@CoNi-LDH Composite Structures for High-Performance Supercapacitors. ACS Appl. Mater. Interfaces. 2024;16(18):23241–23252. doi: 10.1021/acsami.4c01533. [DOI] [PubMed] [Google Scholar]
- Mulla N. R., Patel N. N., Bhosale S. B., Patil U. M., Patil R. S.. Morphologically Tuned MnO2 Thin Film Electrodes Prepared by Growth Kinetic Dependent SILAR Approach for High-Performance Extrinsic Pseudocapacitors. J. Alloys Compd. 2024;1006:176261. doi: 10.1016/j.jallcom.2024.176261. [DOI] [Google Scholar]
- Bulla M., Kumar V., Devi R., Kumar S., Sisodiya A. K., Dahiya R., Mishra A. K.. Natural Resource-Derived NiO Nanoparticles via Aloe Vera for High-Performance Symmetric Supercapacitor. Sci. Rep. 2024;14(1):7389. doi: 10.1038/s41598-024-57606-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dai J., Yang C., Xu Y., Wang X., Yang S., Li D., Luo L., Xia L., Li J., Qi X., Cabot A., Dai L.. MoS2@Polyaniline for Aqueous Ammonium-Ion Supercapacitors. Adv. Mater. 2023;35(39):2303732. doi: 10.1002/adma.202303732. [DOI] [PubMed] [Google Scholar]
- Ning W., Xu S., Wang P., Ma H., Yang X., Sun X., Yang C., Shi X.-R.. MOF Derived Tri-Metallic CoNiMn Hydroxide Assembled on Carbon Cloth for Hybrid Supercapacitor and Hydrogen Evolution. J. Energy Storage. 2024;96:112758. doi: 10.1016/j.est.2024.112758. [DOI] [Google Scholar]
- Liu R., Shi X.-R., Wen Y., Shao X., Su C., Hu J., Xu S.. Trimetallic Synergistic Optimization of 0D NiCoFe-P QDs Anchoring on 2D Porous Carbon for Efficient Electrocatalysis and High-Energy Supercapacitor. J. Energy Chem. 2022;74:149–158. doi: 10.1016/j.jechem.2022.07.015. [DOI] [Google Scholar]
- Paul A., Samanta P., Kolya H., Kang C.-W., Murmu N. C., Kuila T.. Rational Design of Asymmetric Aqueous Supercapacitor Based on the Composites of Reduced Graphene Oxide with Manganese-Doped Nickel Sulfide-Tin Sulfide as Positive Electrode and Manganese Sulfide Negative Electrode. J. Power Sources. 2023;580:233352. doi: 10.1016/j.jpowsour.2023.233352. [DOI] [Google Scholar]
- Zhang Y., Jing X., Yan X., Gao H., Gao K., Cao Y., Hu S., Zhang Y.. Rational Design of NiMn-Based Electrode Materials for High-Performance Supercapacitors. Coord. Chem. Rev. 2024;499:215494. doi: 10.1016/j.ccr.2023.215494. [DOI] [Google Scholar]
- Zhu M., Yang Y., Ma Y.. Salt-Assisted Synthesis of Advanced Carbon-Based Materials for Energy-Related Applications. Green Chem. 2023;25(24):10263–10303. doi: 10.1039/D3GC03080F. [DOI] [Google Scholar]
- Sinha P., Yadav A., Tyagi A., Paik P., Yokoi H., Naskar A. K., Kuila T., Kar K. K.. Keratin-Derived Functional Carbon with Superior Charge Storage and Transport for High-Performance Supercapacitors. Carbon. 2020;168:419–438. doi: 10.1016/j.carbon.2020.07.007. [DOI] [Google Scholar]
- Verma K. D., Sinha P., Ghorai M. K., Kar K. K.. Mesoporous Electrode from Human Hair and Bio-Based Gel Polymer Electrolyte for High-Performance Supercapacitor. Diamond Relat. Mater. 2022;123:108879. doi: 10.1016/j.diamond.2022.108879. [DOI] [Google Scholar]
- Qian W., Sun F., Xu Y., Qiu L., Liu C., Wang S., Yan F.. Human Hair-Derived Carbon Flakes for Electrochemical Supercapacitors. Energy Environ. Sci. 2014;7(1):379–386. doi: 10.1039/C3EE43111H. [DOI] [Google Scholar]
- Brebu M., Spiridon I.. Thermal Degradation of Keratin Waste. J. Anal. Appl. Pyrolysis. 2011;91(2):288–295. doi: 10.1016/j.jaap.2011.03.003. [DOI] [Google Scholar]
- Lesbayev B., Auyelkhankyzy M., Ustayeva G., Yeleuov M., Rakhymzhan N., Maral Y., Tolynbekov A.. Modification of Biomass-Derived Nanoporous Carbon with Nickel Oxide Nanoparticles for Supercapacitor Application. J. Compos. Sci. 2023;7(1):20. doi: 10.3390/jcs7010020. [DOI] [Google Scholar]
- Tian W., Ren P., Hou X., Xue R., Chen Z., Guo Z., Jin Y., Ren F.. MnO2 Porous Carbon Composite from Cellulose Enabling High Gravimetric/Volumetric Performance for Supercapacitor. Int. J. Biol. Macromol. 2024;261:129977. doi: 10.1016/j.ijbiomac.2024.129977. [DOI] [PubMed] [Google Scholar]
- Ghaly H. A., El-Deen A. G., Souaya E. R., Allam N. K.. Asymmetric Supercapacitors Based on 3D Graphene-Wrapped V2O5 Nanospheres and Fe3O4@3D Graphene Electrodes with High Power and Energy Densities. Electrochim. Acta. 2019;310:58–69. doi: 10.1016/j.electacta.2019.04.071. [DOI] [Google Scholar]
- Yang B., Zhang D., Li Y., He Y., Li Y., Li Z., Wang B., Han Z., Wang K.. Locally Graphitized Biomass-Derived Porous Carbon Nanosheets with Encapsulated Fe3O4 Nanoparticles for Supercapacitor Applications. Chem. Eng. J. 2024;479:147662. doi: 10.1016/j.cej.2023.147662. [DOI] [Google Scholar]
- Han Z., Zhang X., Tang G., Liang J., Wu W.. Hollow Co3O4 Nanoparticles on CNF with Polydopamine Assistance for Enhanced Supercapacitor Electrodes. Electrochim. Acta. 2024;500:144778. doi: 10.1016/j.electacta.2024.144778. [DOI] [Google Scholar]
- Cuéllar-Herrera L., Arce-Estrada E. M., Pacheco-Catalán D. E., Vivas L., Ortiz-Landeros J., Jiménez-Lugos C., Romero-Serrano A.. Chemical Synthesis and Electrochemical Performance of Hausmannite Mn3O4/rGO Composites for Supercapacitor Applications. Int. J. Electrochem. Sci. 2024;19(9):100737. doi: 10.1016/j.ijoes.2024.100737. [DOI] [Google Scholar]
- Sunil Kumar K., Pundareekam Goud J., Roy N., Jong Su K., Joo S. W.. In-Situ Synthesis of SnO2/CoFe2O4/Fe3O4 Nanograss Array Composite: A Redox-Active Electrode Material for Battery-Type Supercapacitors. Ceram. Int. 2024;50(11):20535–20546. doi: 10.1016/j.ceramint.2024.03.176. [DOI] [Google Scholar]
- Kumar R., Joanni E., Tan W. K., Matsuda A.. Microwave-Aided Ultra-Fast Synthesis of Fe3O4 Nanoparticles Attached Reduced Graphene Oxide Edges as Electrode Materials for Supercapacitors. Mater. Today Commun. 2024;38:108438. doi: 10.1016/j.mtcomm.2024.108438. [DOI] [Google Scholar]
- Liu X., Lu Y., Xiao G., Zhao H., Hu Z., Zhu J., Liu Z.. MnO2 Loaded Flower-like Carbon Microspheres Derived from Polyimides for High-Performance Supercapacitors. Electrochim. Acta. 2023;454:142417. doi: 10.1016/j.electacta.2023.142417. [DOI] [Google Scholar]
- Liu J.-H., Yang J.-L., Yang F., Feng J., Yuan X.-Y., Zhou J., Ma X.-J., Zhang W.-B.. A Renewable Bifunctional Structural Supercapacitor Using the Recycled Geopolymer Membrane and MnO2/CNTs Electrode. J. Energy Storage. 2024;102:114203. doi: 10.1016/j.est.2024.114203. [DOI] [Google Scholar]
- Wang K., Wang J., Qian J., Yu Q., Bai J., Wei Y., Chen J., Wu M., Sun S., Mao C.-J.. Boosted Na+-MnO2 Supercapacitor Performance via Strong Metal Support Interaction. J. Colloid Interface Sci. 2025;682:865–874. doi: 10.1016/j.jcis.2024.11.252. [DOI] [PubMed] [Google Scholar]
- Barclay M., Firestein K., Jadhav S. D., Motta N., Ostrikov K. Ken.. Development of Novel Plasma Process for Modification of Solid and Liquid Interfaces for High-Energy-Density Cu-MnOx Supercapacitor Electrodes. Electrochim. Acta. 2024;506:144985. doi: 10.1016/j.electacta.2024.144985. [DOI] [Google Scholar]
- Cheng X., Zhang L., Li L., Wu H., Zheng J., Yao J., Li G.. One-Step Hydrothermal Synthesis of Al-Doped δ-MnO2with Three-Dimensional Open Network Structure for High Performance Asymmetric Supercapacitors. Colloids Surf., A. 2025;709:136163. doi: 10.1016/j.colsurfa.2025.136163. [DOI] [Google Scholar]
- Xie M., Zhong J., Wang H., Liu X., Li M., Kang L., Du L.. Oxygen Vacancies-Enriched Manganese Dioxide/Carbon Nanotubes Enabling High-Energy Aqueous Ammonium-Ion Hybrid Supercapacitors. J. Power Sources. 2025;632:236315. doi: 10.1016/j.jpowsour.2025.236315. [DOI] [Google Scholar]
- Zhang D., Wang Y., Yang M., Zhang J., Wu J., Liu H., Wu J., Li P.. Preparation of Sandwich-Structured Ternary Nanocomposites Porous Carbon-Derived from Waste Tires/Manganese Dioxide/Polyaniline as Electrode for Supercapacitor. J. Electroanal. Chem. 2024;974:118692. doi: 10.1016/j.jelechem.2024.118692. [DOI] [Google Scholar]
- Sarkar A., Gopal Khan G.. Synthesis of BiFeO 3 Nanoparticle Anchored TiO2-BiFeO3 Nano-Heterostructure and Exploring Its Different Electrochemical Aspects as Electrode. Mater. Today: Proc. 2018;5(3):10177–10184. doi: 10.1016/j.matpr.2017.11.016. [DOI] [Google Scholar]
- Reddy S., Kumara Swamy B. E., Jayadevappa H.. CuO Nanoparticle Sensor for the Electrochemical Determination of Dopamine. Electrochim. Acta. 2012;61:78–86. doi: 10.1016/j.electacta.2011.11.091. [DOI] [Google Scholar]
- Marecka-Migacz A., Mitkowski P. T., Nędzarek A., Różański J., Szaferski W.. Effect of pH on Total Volume Membrane Charge Density in the Nanofiltration of Aqueous Solutions of Nitrate Salts of Heavy Metals. Membranes. 2020;10(9):235. doi: 10.3390/membranes10090235. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ravel B., Newville M.. ATHENA, ARTEMIS, HEPHAESTUS: Data Analysis for X-Ray Absorption Spectroscopy Using IFEFFIT . J. Synchrotron Radiat. 2005;12(4):537–541. doi: 10.1107/S0909049505012719. [DOI] [PubMed] [Google Scholar]
- Gougoussis C., Calandra M., Seitsonen A., Brouder Ch., Shukla A., Mauri F.. Intrinsic Charge Transfer Gap in NiO from Ni K-Edge X-ray Absorption Spectroscopy. Phys. Rev. B. 2009;79(4):045118. doi: 10.1103/PhysRevB.79.045118. [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
The authors declare that all data supporting the findings of this study are available within the paper and its Supporting Information file.








