Startseite Naturwissenschaften Synthesis and electrochemical characterization of iron oxide/poly(2-methylaniline) nanohybrids for supercapacitor application
Artikel Open Access

Synthesis and electrochemical characterization of iron oxide/poly(2-methylaniline) nanohybrids for supercapacitor application

  • Ashour M. Ahmed , Doaa Essam , Mohamed A. Basyooni-M. Kabatas EMAIL logo , Ahmed A. Abdel-Khaliek , Mohamed Shaban , Mamduh J. Aljaafreh und Mohamed Rabia
Veröffentlicht/Copyright: 21. Juli 2025
Veröffentlichen auch Sie bei De Gruyter Brill

Abstract

This study reports the synthesis of a nanohybrid material composed of poly(2-methylaniline) (P(2MA)) and iron oxide (Fe2O3) as electrodes for supercapacitors using a simple and cost-effective method. Various characterization techniques were employed to analyze the samples. The results revealed that the Fe2O3/P(2MA) nanohybrid exhibits nanofiber structures, while pure P(2MA) displays a porous hollow sphere morphology. Furthermore, the analysis confirmed the effective dispersion of Fe2O3 nanoparticles within the polymer matrix. The electrochemical properties of the Fe2O3/P(2MA) nanohybrid were found to surpass those of pure P(2MA) in both NaCl and HCl electrolytes. Notably, the nanohybrid demonstrated longer discharge times and higher oxidation/reduction currents in HCl than NaCl. The gravimetric and areal capacitances were measured at 998.4 F g−1 and 1497.6 mF cm−2 in 0.5 M HCl at a current density of 0.6 A g−1. Furthermore, the nanohybrid retained 99.9% of its initial specific capacitance after 2,000 cycles. These findings underscore the significant potential of the Fe2O3/P(2MA) nanohybrid as a high-performance supercapacitor electrode for energy storage applications.

1 Introduction

Currently, electrochemical supercapacitors (ECs) have a wide range of modern applications [13]. Supercapacitors can be categorized into electric double-layer capacitors (EDLCs) and pseudocapacitors (PCs) based on the different electrochemical energy storage mechanisms they employ [4,5]. EDLCs utilize carbon-based materials such as graphene and activated carbon as electrodes, enabling reversible adsorption and desorption of ions at the electrode–electrolyte interface [6,7]. The charge storage primarily occurs within the electric double layer formed at this interface. The high specific surface area, good electrical conductivity, and mechanical stability of carbon-based EDLC materials have attracted significant research interest [8,9]. However, these materials often suffer from high internal resistance and relatively low capacitance, which restricts their broader application [10]. PCs store energy through quick and reversible Faradaic interactions at the surface of active electrodes. Conducting polymers (CPs) have been extensively studied as potential materials for PCs, including polypyrrole (PPy), polyindole, poly(3,4-ethylenedioxythiophene), polyaniline (PANI), and poly(2-methyl aniline) (P(2MA)) [11,1218]. However, CPs often suffer from cycling instability due to the swelling and shrinking of the polymeric chains during the charging–discharging process [19]. Metal oxides (MOs) and metal sulfides investigated as supercapacitor electrodes include CoS, PbS, RuO2, NiO, Al2O3, TiO2, MnO2, Mn3O4, and V2O5 [19,20,21,22,23]. Although MOs offer specific energy and capacitance advantages, they suffer from low conductivity and are prone to easy agglomeration during interactions [24]. Many researchers have investigated blending conductive polymers with MOs to address these limitations. The combination of CPs/MOs is promising for improving electrical conductivity and enabling fast, reversible redox reactions with enhanced specific capacitance [25,26]. In this context, Fathy et al. used cobalt nickel oxide combined with poly(m-toluidine) as an electrode material to enhance supercapacitor performance [27]. Bathula et al. reported that poly(3-dodecylthiophene)-wrapped cobalt oxide (P3DDT-Co3O4) exhibited a specific capacitance of 294 F g−1, substantially higher than that of pristine Co3O4 (174 F g−1) at a current density of 1 A g−1 [28]. Chenyang et al. described 3D-structured NiCo2O4-PANI nanosheets as a standing electrode for supercapacitors, delivering a higher specific capacitance than NiCo2O4 alone [29]. Additionally, Merlin et al. presented Ag2WO4 and an Ag2WO4/PANI nanocomposite as supercapacitor electrodes, with the Ag2WO4/PANI nanocomposite showing about five times the capacitance of Ag2WO4 alone [30]. In these works, the capacitance remains relatively low and requires further improvements for better supercapacitor performance.

On the other hand, P(2MA) is a promising CP derived from the monomer m-toluidine through oxidative polymerization. It belongs to the polyaniline family and shares many similar properties. Notably, P(2MA) exhibits electrical conductivity due to the delocalization of electrons along its polymer chain, facilitating the flow of electric charge. The electron-donating methyl group on its phenyl ring lowers resistance and enhances capacitance [31]. Moreover, these methyl groups contribute additional electrons, increasing the surface area of the electroactive sites, enhancing polymer stability, and accelerating the charging and discharging processes [32,33]. P(2MA) also demonstrates advantageous characteristics such as high stability, resistance to oxidation, reversible redox reactions, intense absorption in the visible region, and mechanical flexibility. These attributes make P(2MA) suitable for diverse applications, including optoelectronic systems, electrochemical devices, hydrogen generation, sensors, and conductive coatings [27,31,32,34,35]. However, despite its promising qualities, its potential as an energy storage electrode for supercapacitors remains underexplored.

Iron oxide (Fe2O3) nanoparticles have gained considerable attention for their potential in supercapacitors, fuel cells, electronic devices, and photocatalysis [36]. Iron oxide offers notable advantages such as cost-effectiveness, abundant availability, and environmental friendliness [37,38]. Its suitability for supercapacitor electrodes is attributed to its high theoretical capacity and the ability to exist in multiple oxidation states. Vittaya et al. developed flower-shaped Fe2O3 nanomaterials as supercapacitor electrodes, achieving a capacitance of 218.49 F g−1 at 1 A g−1 [39]. Wang et al. fabricated trigonal α-Fe2O3 structures via high-temperature thermal decomposition, resulting in a specific capacitance of 149.3 F g−1 at 1 A g−1 [40]. Saxena et al. synthesized Fe2O3 nanocubes with a specific capacitance of 288 F g−1 at 1 A g−1 [41]. Similarly, Surender et al. examined Fe2O3 nanomaterials as supercapacitor electrodes, achieving a capacitance of approximately 123 F g−1 at 1 A g−1 [42]. Nevertheless, Fe2O3 exhibits drawbacks, including poor Coulombic efficiency and rapid capacity decay during charge–discharge cycles [43,44]. Its limited electrical conductivity and moderate specific capacitance have also restricted its broader adoption in supercapacitor technologies.

It is expected that incorporating Fe2O3 into the P(2MA) matrix to form a Fe2O3/P(2MA) nanohybrid will effectively address the limitations of both materials for supercapacitor applications. This integration enhances the contact between the nanohybrid and the current collector, improving overall electrical conductivity [45,46]. Additionally, it expands the surface area available for the reversible redox faradaic reactions. The synergistic effects of the Fe2O3/P(2MA) nanohybrid result in several benefits, such as facilitating more effective charge transfer and increased stability over longer-term operations.

Most researchers initially synthesized MO nanoparticles and then incorporated them into polymer chains during polymerization to fabricate nanocomposites [27,47,48]. This study uses a one-pot photopolymerization approach to develop Fe2O3/P(2MA) nanohybrids for supercapacitor applications. A key innovation lies in combining photopolymerization with chemical oxidative polymerization, enabling efficient, controllable synthesis. In this method, FeCl3 serves a dual role as the oxidizing agent and the precursor for in situ Fe2O3 nanoparticle formation, facilitating the direct construction of a nanofibrous Fe2O3/P(2MA) nanohybrid. P(2MA), as an alternative to the widely studied polyaniline, takes advantage of its high electrical conductivity, thermal stability, and ease of processing. The resulting nanohybrid features uniform Fe2O3 dispersion and strong polymer–oxide interfacial bonding, forming a porous, high-surface-area architecture that improves ion diffusion and charge transport. This synergistic structure significantly boosts specific capacitance and cycling stability. Additionally, the single-step, cost-effective, and environmentally friendly method eliminates the need for separate oxidants and reduces fabrication complexity compared to traditional multi-step procedures.

The characteristics of both pure P(2MA) and Fe2O3/P(2MA) nanohybrids were analyzed using several characterization techniques, including X-ray diffraction (XRD), Fourier transform infrared spectroscopy (FTIR), X-ray photoelectron spectroscopy (XPS), scanning electron microscopy (SEM), and UV spectroscopy. The electrochemical performance of the nanohybrid electrode was evaluated using different aqueous electrolytes, specifically NaCl and HCl. After 2,000 charge/discharge cycles, the Fe2O3/P(2MA) nanohybrid exhibited high stability in HCl electrolytes.

2 Experimental section

2.1 Synthesis of the P(2MA) nanopowder

The pure P(2MA) nanopowder was synthesized using the photopolymerization method. Initially, a solution of 0.12 M m-toluidine (C6H4CH3NH2) and 0.5 M acetic acid (CH3COOH) was stirred, while a separate solution of 0.12 M potassium persulfate (KPS) dissolved in 0.5 M CH3COOH acted as the oxidizing agent. The two solutions were then combined and stirred for 2 h. Next, the resulting mixture was exposed to UV light for 15 h to ensure complete polymerization. The mixture was filtered to obtain the pure P(2MA) nanopowder and was finally dried at 70°C.

2.2 Synthesis of the Fe2O3/P(2MA) nanohybrid

The synthesis of the Fe2O3/P(2MA) nanohybrid followed a slightly different procedure. Initially, 0.12 M m-toluidine was dissolved in 0.5 M acetic acid to create a solution. Simultaneously, another solution was prepared, containing 0.12 M iron chloride hexahydrate (FeCl3.6H2O) and 0.5 M acetic acid. Equimolar concentrations of 0.12 M FeCl₃·6 H₂O and 0.12 M m‑toluidine were intentionally selected to balance the redox capacity of Fe2O3 with the electrical conductivity of P(2MA). An excessive amount of Fe2O3 can hinder electron transport and promote nanoparticle aggregation, while an overabundance of the polymer may dilute redox-active sites and reduce overall capacitance [11,22]. This equimolar ratio thus represents a near-optimal compromise, resulting in high specific capacitance and excellent cycling stability. These two solutions were then mixed for 2 h. Subsequently, the combined solutions were exposed to UV light for 15 h. Finally, the mixture was filtered, purified, and dried overnight at 70°C to obtain the Fe2O3/P(2MA) nanohybrid. Based on the initial molar concentrations and accounting for Fe2O3 formation (0.06 mol × 159.69 g/mol ≈ 9.58 g) and polymer yield (≈12.86 g), the final mass ratio of Fe2O3 was approximately 43 wt% of the nanohybrid.

2.3 Characterization of the synthesized nanostructures

Characterization of the synthesized nanostructures involved several techniques to assess their properties. The morphological features were observed using an SEM (Axioskop 40 POL, Zeiss) and a transmission electron microscope (TEM, JOEL JEM-2010). The compositional structure properties of both the pure P(2MA) nanopowder and the Fe2O3/P(2MA) nanohybrid were analyzed using X-ray diffraction (XRD, Philips X’Pert Pro MRD) and Fourier-transform infrared spectroscopy (FTIR, Bruker Vertex 70). A UV-Vis spectrophotometer (Lambda 950, Perkin Elmer) was employed to investigate the optical characteristics. Furthermore, XPS (Axis Ultra DLD) was performed to evaluate the elemental state of the nanostructures.

2.4 Supercapacitor design

To create a well-mixed catalyst ink, 250 µg of active nanopowder material (P(2MA) or Fe2O3/P(2MA)) and 25 µg of graphite nanopowder were combined with 0.350 mL of ethanol and 0.050 mL of Nafion. The resulting mixture was then vigorously shaken in a sealed bottle for 1 day to ensure thorough homogenization. Collectors currently employ two gold (Au) sheets. Au was selected for its high electrical conductivity and chemical stability, ensuring no reaction with the electrolytes. The current collector played a vital role in a supercapacitor by facilitating efficient electron transfer between the external circuit and electrode materials during the charge and discharge processes [49]. It ensures uniform current distribution across the electrode surface.

Additionally, it provides mechanical support to the electrode materials and aids in dissipating heat generated during operation. Approximately 56.0 µL of the homogeneous slurry was carefully applied to etch the Au sheets (1 cm2). Two pieces of filter paper were immersed in separate solutions overnight. One piece was soaked in a 0.5 M sodium chloride (NaCl) electrolyte, while the other was submerged in a 0.5 M hydrochloric acid (HCl) electrolyte. A strip of electrolyte-soaked filter paper was inserted between the two newly constructed electrodes to design the supercapacitor device.

2.5 Electrochemical description

Various electrochemical tests were conducted using an electrochemical workstation (CHI 660E; CH Instruments, China) to evaluate the performance of the pure P(2MA) and Fe2O3/P(2MA) nanohybrid. Galvanostatic charge/discharge (GCD) measurements were performed at different current densities in different aqueous electrolyte solutions. Cyclic voltammetry (CV) experiments were conducted at various scan rates within a voltage window ranging from 0 to 1 V.

3 Results and discussion

3.1 Morphological and structural section

SEM was employed to analyze the morphological features of both pure P(2MA) and the Fe2O3/P(2MA) nanohybrid, as shown in Figure 1. The SEM image (Figure 1a) revealed that pure P(2MA) exhibited spherical shapes with hollow interior structures. The surfaces of these hollow spheres were open and featured small holes. Similar structures have been observed in previous studies for polymers such as poly(o-toluidine) and polyaniline when KPS or ammonium persulfate were used as oxidants [50,51]. Figure S1 clearly illustrates these hollow spheres with surface holes. The hole sizes ranged from 0.2 to 0.35 µm, while the hollow sphere diameters ranged from 0.5 to 0.9 µm. The hollow interior in these spheres allows solutions to penetrate inside. The morphology of porous hollow sphere nanoparticles is helpful for various applications where solution infiltration or encapsulation is desired.

Figure 1 
                  SEM images of (a) pure P(2MA), (b) Fe2O3/P(2MA) nanohybrid, (c) TEM micrograph, and (d) SAED of Fe2O3/P(2MA).
Figure 1

SEM images of (a) pure P(2MA), (b) Fe2O3/P(2MA) nanohybrid, (c) TEM micrograph, and (d) SAED of Fe2O3/P(2MA).

SEM micrographs of the Fe2O3/P(2MA) nanohybrid (Figure 1b) illustrate the formation of a spongy and irregularly accumulated structure of nanofibers. This structure indicates an improved attachment of Fe2O3 to the P(2MA) chain. The corresponding microstructure exhibits highly desirable characteristics for energy storage systems [52]. The presence of aggregated porosity areas enhances electrical conductivity because when a material has aggregated porosity, a network of interconnected pores is formed. This arrangement can significantly improve electrical conductivity by providing several channels for electron transport [5357]. Nanofiber-based materials reduce the ion diffusion pathway, enabling faster ion intercalation kinetics [58]. The high surface area of the spongy structure helps decrease charge-transfer resistance at the electrode/electrolyte interface [59,60].

The transition from KPS to FeCl3 plays a critical role in driving the morphological shift from hollow spheres in pure P(2MA) to spongy nanofibers in Fe2O3/P(2MA) during photopolymerization [61]. This transformation is multifactorial and complex. In pure P(2MA) synthesis, KPS functions solely as a source of sulfate radicals that initiate the polymerization process, without exerting significant templating or structural influence. However, since 2MA dissolves in acid, some monomer droplets are dispersed across the surface, and their oxidation in the presence of KPS leads to the formation of hollow spheres via the water-in-oil method [50,51,62]. In contrast, FeCl3 in the Fe2O3/P(2MA) nanohybrid serves a dual function: it acts as an oxidizing agent and a precursor for the in situ generation of Fe2O3 nanoparticles [63]. These photochemically generated nanoparticles serve as coordination-templating sites that guide polymer growth into one-dimensional fibers through strong polymer–nanoparticle interactions [64]. The simultaneous formation of the polymer and MO phases creates a dynamic reaction environment fundamentally distinct from the pure polymer system. Changes in solvent polarity, viscosity, ionic strength, and pH induced by FeCl3 influence polymer chain behavior, modifying the propagation mechanism during growth and altering the polymerization kinetics. This favors linear propagation and suppresses the cross-linking processes that typically lead to the formation of closed hollow spheres. Additionally, introducing Fe³⁺ ions facilitates the formation of coordination complexes with the amine groups in the monomer [65]. Therefore, the observed morphological transition is likely the result of a synergistic interplay among the oxidizing agent, nanoparticle formation, and polymer–MO interactions, rather than the effect of the oxidizing agent alone [66].

The TEM morphology of the nanohybrid is presented in Figure 1c. The TEM image reveals nanofibers of the nanohybrid at the edges, as indicated by the red circle. The darkish area corresponds to Fe2O3 nanoparticles, while the gray-colored regions represent the P(2MA) particles. Figure 1(d) presents the selected area electron diffraction (SAED) distribution of the nanohybrid, confirming the presence of polycrystalline structures. The diffraction rings observed in the SAED pattern align closely with the crystalline planes of the Fe2O3/P(2MA) nanohybrid, specifically corresponding to the (012) plane for Fe2O3 and the [67] plane for P(2MA). This correlation is consistent with the XRD results presented in Figure 2 and previous results [5].

Figure 2 
                  (a) XRD patterns of P(2MA) and (b) the Fe2O3/P(2MA) nanohybrid.
Figure 2

(a) XRD patterns of P(2MA) and (b) the Fe2O3/P(2MA) nanohybrid.

3.2 XRD properties

Figure 2 presents the results of the XRD test conducted to examine the crystal structure of the pure P(2MA) and the Fe2O3/P(2MA) nanohybrid. For the pure P(2MA), the diffraction peaks at 2θ = 14° and 25° correspond to the parallel and perpendicular periodicities of the polymer backbone, respectively, as depicted in Figure 2(a) [68,69]. The broad peaks in the polymer structure indicate its semi-crystalline nature [68]. The XRD pattern of the Fe2O3/P(2MA) nanohybrid, shown in Figure 2(b), confirms the presence of crystalline Fe2O3. A series of well-defined diffraction peaks appear at 2θ = 23.8°, 31.0°, 42.0°, 61.5°, 65.7°, 70.6°, and 77.1°. These peaks are indexed to the (012), (1040, (113), (214), (300), (028), and (220) planes, respectively, and are consistent with the standard patterns listed in JCPDS card numbers 86-0550 and 33-0664 [70,71]. These planes are characteristic of Fe2O3 and confirm its successful incorporation into the nanohybrid. Additionally, the P(2MA) component, being largely semi-crystalline, does not contribute sharp peaks but may contribute a broad hump around 20°–25°, which is typical of CPs. The retention of sharp Fe2O3 peaks indicates that the polymer matrix does not disrupt the crystalline structure of the MO, allowing for effective nanohybrid formation with preserved crystallinity. Figure S2 provides the standard XRD patterns for Card No. JCPDS 86-0550 of Fe2O3.

To calculate the crystallite size (CS), the Scherrer equation was utilized as follows:

(1) CS = ( 0.9 λ ) / ( β cos θ ) ,

where λ represents the X-ray wavelength (CuKα = 0.15405 nm), θ is the Bragg’s angle, and β is the full width at half-maximum [72]. The calculated CS was 3.55 nm for Fe2O3/P(2MA).

3.3 FTIR analysis

Figure 3 illustrates the distinctive bands obtained from the FTIR analysis of pure P(2MA) and the Fe2O3/P(2MA) nanohybrid. In pure P(2MA), a broad band at 3,423 cm−1 corresponds to the stretching vibration of the (–NH–) group. The quinoid ring of pure P(2MA) displays a band at 2,937 cm−1, which indicates C–H stretching vibrations and reflects the polymer’s conjugation nature. Additionally, the band at 1,107 cm−1 is associated with in-plane C–H bending vibrations [73]. The band at 1,633 cm−1 is attributed to C═C stretching vibrations within the conjugated system. C–N stretching vibrations of the benzenoid ring are observed at 1,368 cm−1, while the band at 833 cm−1 is contributed by the methyl group attached to the phenyl ring.

Figure 3 
                  FTIR spectra of (a) P(2MA) and (b) Fe2O3/P(2MA) nanohybrid.
Figure 3

FTIR spectra of (a) P(2MA) and (b) Fe2O3/P(2MA) nanohybrid.

The Fe2O3/P(2MA) nanohybrid FTIR spectrum reveals characteristic vibrational bands that confirm the successful incorporation of Fe2O3 into the polymer matrix. A broad band at 3,222 cm−1 is attributed to the N–H stretching vibration of the amine group in P(2MA), which is slightly shifted from that of the pure polymer due to interactions with the Fe2O3 nanoparticles [74,75]. The band at 2,972 cm−1 corresponds to C–H stretching in the aromatic rings [73]. The peak at 1,590 cm−1 indicates C═C stretching vibrations of the quinoid and benzenoid rings, suggesting the preservation of the polymer’s conjugated backbone. In contrast, the band at 1,475 cm−1 is associated with C–N stretching within the benzenoid unit. The bands at 1,305 cm−1, 1,222 cm−1, and 1,161 cm−1 are attributed to in-plane bending of the C–H and C–N bonds, which are sensitive to interactions with MOs [74,75]. The band at 812 cm−1 reflects the C–H out-of-plane bending modes, consistent with substituted aromatic rings. Furthermore, the strong band at 594 cm−1 is assigned to Fe–O stretching vibrations, confirming the presence of Fe2O3 nanoparticles within the polymer matrix [76]. These shifts and the emergence of new bands in the nanohybrid spectrum, compared to pure P(2MA), indicate strong interactions between the polymer and Fe2O3, suggesting effective nanohybrid formation that may enhance electron transport and redox behavior.

3.4 Optical properties

The optical spectra of pure P(2MA) and the Fe2O3/P(2MA) nanohybrid were analyzed, as shown in Figure S3. In the pure P(2MA) (blue) spectrum, three distinct absorption peaks are observed at approximately 335, 433, and 582 nm. These peaks correspond to the electronic transitions of the benzenoid ring, the P(2MA) emeraldine salt polaron–π* transition, and the excitation at the quinonoid portions, respectively [77,78]. In the case of the Fe2O3/P(2MA) nanohybrid, the absorption peak is broader and exhibits higher absorbance across the entire wavelength range. Fe2O3 has a broad absorption band in the visible region [79,80]. Consequently, the Fe2O3/P(2MA) nanohybrid exhibits a broad absorption band from the UV to the visible and near-infrared (NIR) regions. This suggests that the Fe2O3 nanoparticles are well-blended into the polymeric chain, increasing absorption.

The Tauc equation was employed to determine the direct optical bandgap of pure P(2MA) and the Fe2O3/P(2MA) nanohybrid as follows [81]:

(2) α E ph = B ( E ph E g ) 0.5 .

Here, α represents the absorption coefficient and E ph represents the energy of the incident light. The constant B represents the absorption threshold width, and E g denotes the optical bandgap energy [82]. The Tauc plot is presented in Figure 3S. After incorporating Fe2O3 nanoparticles, the direct optical bandgap energy of P(2MA) decreases from 2.9 to 2.5 eV, a value consistent with that reported in the literature [76,79]. The electronic interaction between Fe2O3 and P(2MA) leads to a reduction in the energy gap of the nanohybrid.

3.5 XPS

Figure 4 illustrates the full XPS spectrum of the Fe2O3/P(2MA) nanohybrid to visualize the elemental configuration and bonding state. There are five main peaks. The Fe 2p3 peak is observed at a binding energy of 715.8 eV. A significant C 1s peak is observed at a binding energy of 285.36 eV, with an atomic ratio of 82.85%. Additionally, an O 1s peak is detected at 532.93 eV, and the N 1s peak appears at 399.86 eV, with atomic ratios of 4.69 and 10.57%, respectively. The Cl 2p peak, originating from the iron chloride precursor, is observed at a binding energy of 198.38 eV, with an atomic ratio of 1.43%.

Figure 4 
                  Full survey XPS characterization of the Fe2O3/P(2MA) nanohybrid, which showed peaks of Fe 2p, C 1s, O 1s, and N 1s spectra.
Figure 4

Full survey XPS characterization of the Fe2O3/P(2MA) nanohybrid, which showed peaks of Fe 2p, C 1s, O 1s, and N 1s spectra.

Figure S4 presents the high-resolution XPS spectra of the Fe2O3/P(2MA) nanohybrid. Figure S4(a) presents a high-resolution characteristic spectrum of the Fe 2p for the Fe2O3/P(2MA) nanohybrid. The Fe 2p peak can be deconvoluted into six peaks. The two peaks with higher atomic ratios, located at 711.97 and 724.64 eV, respectively, correspond to Fe 2p3/2 and Fe 2p1/2. These peaks are characteristic of the Fe+3 state, indicating the presence of Fe2O3 nanoparticles within the polymer chain [43,83]. The high-resolution XPS of the C 1s spectrum is shown in Figure S4(b), and it reveals four different carbon species: C–C/C═C (284.24 eV), C–N (285.17 eV), and C–O (289.19 eV) [84]. The Fe–O bond exhibits a prominent peak in the O 1s spectra at 532 eV, while the O–H group displays a shoulder peak at 531 eV, as depicted in Figure S4(c) [83]. The high-resolution N1s spectra (Figure S4(d)) displayed one peak at 398.6 eV caused by the N–H and C–N bonds associated with the amine (–NH–) groups and cationic species (N+) of P(2MA) [27].

3.6 Electrochemical performance

3.6.1 Galvanostatic charge/discharge profiles

Galvanostatic charge/discharge (CD) testing is performed to determine the charging and discharging time of the electrode material. Additionally, it is used to determine the capacitance of the electrode material. Figure 5 demonstrates the charge/discharge curves of the Fe2O3/P(2MA) nanohybrid compared to pure P(2MA) in a 0.5 M HCl solution at 0.4 Ag−1. The total charge/discharge time of P(2MA) is approximately 42 s, while the total charge/discharge time of the Fe2O3/P(2MA) nanohybrid is around 670 s.

Figure 5 
                     Charge/discharge process of pure P(2MA) and Fe2O3/P(2MA) in 0.5 M HCl at 0.4 A g−1.
Figure 5

Charge/discharge process of pure P(2MA) and Fe2O3/P(2MA) in 0.5 M HCl at 0.4 A g−1.

The electrochemical mechanisms of the Fe2O3/P(2MA) nanohybrid involve a combination of faradaic redox reactions from both components, contributing to its high specific capacitance. Iron oxide (Fe2O3) plays a critical role by undergoing reversible redox transitions between Fe3+ and Fe2+ states, contributing significant pseudocapacitance through surface-controlled reactions [85]. Concurrently, P(2MA) participates in rapid and reversible doping/dedoping transitions between its leucoemeraldine, emeraldine, and pernigraniline oxidation states. These transitions contribute additional pseudocapacitive charge storage via fast faradaic reactions at the polymer backbone [86]. The associated redox reactions can be summarized as follows:

(3) P ( 2 MA ) P ( 2 MA ) + + e ,

(4) Fe 2 O 3 + H + + e Fe 3 O 4 ,

(5) Fe 3 O 4 + H + + e Fe ( OH ) 2 .

The Fe2O3/P(2MA) nanohybrid morphology, as revealed by TEM and SEM imaging, plays a pivotal role in facilitating these electrochemical mechanisms. The nanohybrid forms a 3D spongy nanofiber network with interconnected meso- and macropores, significantly increasing the electrochemically active surface area. This structure shortens ion-diffusion pathways and enhances electrolyte penetration, both essential for high-rate performance. The in situ formation of Fe2O3 within the P(2MA) matrix promotes uniform nanoparticle dispersion, preventing agglomeration and ensuring full utilization of the redox-active oxide. Additionally, the intimate polymer–oxide interface, formed during one-pot photopolymerization, minimizes interfacial charge-transfer resistance [87]. The flexible polymer backbone further accommodates volumetric expansion and contraction during cycling, preserving structural integrity and enhancing long-term stability [88]. Hence, the Fe2O3/P(2MA) nanohybrid represents a promising candidate for next-generation supercapacitor electrodes.

The electrolyte is crucial for the performance of supercapacitors, as it significantly influences charge transport [89]. It also affects conductivity and capacitance [90,91]. Aqueous electrolytes are widely employed in electrochemical energy storage systems due to their high ionic conductivity, low viscosity, inherent safety, and environmental friendliness compared to organic electrolytes [89,92]. Furthermore, they offer predictable ionic behavior, faster ion diffusion, and reduced fabrication complexity, making them particularly suitable for early-stage materials screening and fundamental electrochemical investigations [91].

Figure 6(a) and (b) illustrates the charge/discharge (CD) behavior profiles of the Fe2O3/P(2MA) nanohybrid electrode in two different electrolytic solutions (0.5 M NaCl and 0.5 M HCl). The CD measurements were conducted under various applied current densities ranging from 0.2 to 1.0 A g−1, within a potential window of 0–1 V. The CD curves demonstrate nearly triangular shapes with slight variations in behavior observed in the two electrolytes. This characteristic pattern signifies the presence of pseudocapacitive behavior [93]. This indicates that the charge storage mechanism involves rapid and reversible redox reactions at the interface between the electrode and the electrolyte [94]. The charge/discharge time decreases in the same electrolyte as the current densities increase. Specifically, in the 0.5 M NaCl electrolyte, the charge/discharge time decreases from 37 to 18 s as the current density increases from 0.2 to 0.4 A g−1. This can be attributed to the limited ability of electrolyte ions to access the nanohybrid interior structure, resulting in only the outer active surface being utilized for ion diffusion at higher currents [93,95].

Figure 6 
                     Galvanostatic charge/discharge profiles of Fe2O3/P(2MA) were measured at different current densities in (a) 0.5 M NaCl and (b) 0.5 M HCl. The relation between the gravimetric and areal capacitances of Fe2O3/P(2MA) in response to various current densities in (c) 0.5 M NaCl and (d) 0.5 M HCl.
Figure 6

Galvanostatic charge/discharge profiles of Fe2O3/P(2MA) were measured at different current densities in (a) 0.5 M NaCl and (b) 0.5 M HCl. The relation between the gravimetric and areal capacitances of Fe2O3/P(2MA) in response to various current densities in (c) 0.5 M NaCl and (d) 0.5 M HCl.

Comparing the discharge time of the nanohybrids in HCl and NaCl at the same current density, it is observed that the discharge time in HCl is higher than that in NaCl, as shown in Figure 6(b). Several factors can explain this. In the NaCl electrolyte, the high surface charge density of Na+ ions leads to strong interactions with water molecules, resulting in larger hydration spheres and less mobile ions. On the other hand, the H+ ions in the HCl electrolyte have smaller hydration spheres, which increases their diffusion rates due to their higher kinetic movements. Additionally, the smaller hydration spheres of H+ ions enable them to penetrate the electrode material, accessing more active regions [90].

Furthermore, H+ ions have higher ionic conductivity compared to Na+ ions [16,96,97]. The higher discharge time observed in HCl indicates a higher capacitance and the ability to store more energy in the supercapacitor device. The nature of the GCD shape is a crucial indicator of the charge storage mechanism, directly corresponding to the type of supercapacitor. The Fe2O3/P(2MA) nanohybrid demonstrates the characteristics of a nanohybrid supercapacitor, combining both pseudocapacitance and electrical double-layer capacitance (EDLC).

Based on the CD curves, the gravimetric ( C wt ) and areal ( C A ) capacitances of the nanohybrid in NaCl and HCl electrolytes were determined using the following equations [4]:

(6) C wt = 4 I m ( E / t ) ,

(7) C A = 4 I A ( E / t ) .

In the equations, I represent the applied constant current (A), ( E / t ) represents the slope of the discharge curve, A represents the electrode footprint area (cm−2), and m (g) represents the mass of the electroactive material on each electrode [98].

The gravimetric and areal capacitances gradually reduce as the current densities increase. This phenomenon can be attributed to the slow electrochemical activity kinetics observed at higher current densities and the limited ability of most nanohybrids to effectively participate in the electrochemical reactions under these higher current densities. Additionally, increasing current density hinders the charges from having sufficient time to traverse through the pores, decreasing the specific capacitance [99,100].

Figure 6(c,d) illustrates the gravimetric and areal capacitance variation of the Fe2O3/P(2MA) nanohybrid at different current densities. The nanohybrid exhibits higher capacitance values at a lower current density (0.2 A g−1). In the NaCl electrolyte, the gravimetric capacitance is approximately 32 F g−1, while the areal capacitance is around 48 mF cm−2 at 0.2 A g−1. Conversely, the nanohybrid demonstrates significantly higher capacitance values in the HCl electrolyte (Figure 6d). At a current density of 0.2 A g−1, the gravimetric capacitance reaches approximately 1385.6 F g−1, and the areal capacitance is approximately 2078.5 mF cm−2. Notably, the gravimetric and areal capacitances of the Fe2O3/P(2MA) nanohybrid in the HCl electrolyte are much greater than those in the NaCl electrolyte. This substantial increase can be attributed to the considerably longer discharge time observed in the HCl electrolyte. The hydrogen ions in HCl can more effectively penetrate the electrode material and participate in electrochemical reactions compared to sodium ions.

To broaden the scope of the study and enhance its comparative depth across different aqueous electrolytes, the electrochemical performance of the Fe2O3/P(2MA) nanohybrid was also evaluated in potassium hydroxide (KOH) as a basic electrolyte, as shown in Figure S5. In 0.5 M KOH, the Fe2O3/P(2MA) nanohybrid exhibited a charge/discharge time of approximately 140 s at 0.2 A g−1, corresponding to a gravimetric capacitance of around 30 F g−1. These values are significantly lower than those obtained in 0.5 M HCl electrolyte.

Several studies have investigated polymer-based MO nanohybrids for supercapacitor applications, highlighting their potential in energy storage. Table 1 summarizes the electrochemical performance of some previous works in this field. The Fe2O3/P(2MA) nanohybrid demonstrates impressive performance, outperforming or closely matching other reported materials. It exhibits a specific capacitance of 998.4 F g−1 at 0.6 A g−1 in 0.5 M HCl, with 99.9% retention after 2,000 cycles. For instance, a P(2MA)/Ag–Ag2O nanocomposite achieved 443 F g−1 at 0.4 A g−1 with 89.9% retention after 1,000 cycles [5], while a poly(m-toluidine)/Co–Ni oxide system reached 308 F g−1 at 0.6 A g−1 with 98% retention after 1,000 cycles [27]. A PANI/α-Fe2O3 nanohybrid recorded 473.6 F g−1 at 1 A g−1 and 98.2% retention over 5,000 cycles in 1 M Na2SO4 [101], while the PPy/Fe2O3 nanocomposite reported a capacitance of 395 F g−1 [67]. These results demonstrate that the proposed Fe2O3/P(2MA) nanohybrid delivers a remarkable combination of high capacitance and excellent cycling durability. This positions it as a competitive and viable candidate for next-generation supercapacitor technologies.

Table 1

Electrochemical performance of supercapacitor electrode materials based on the polymer and MO

Nanomaterial Current density (A g−1) Electrolyte Specific capacitance (F g−1) Ref.
PMT/Ag–Ag2O 0.4 0.5 M HCl 443 [5]
PMT/(Co–Ni) 0.6 0.5 M HCl 308 [27]
PANI/α-Fe2O3 1 1.0 M Na2SO4  473.6 [101]
PANI/α-Fe2O3 0.2 1.0 M H2SO4 857 [102]
Fe2O3@PPy 3 M KOH 395.45 [67]
Fe2O3/poly-2 aminothiophenol 0.3 1.0 M NaOH 46.4 [103]
Fe2O3/PPy/carbon cloth 1.0 M Na2SO4 640 [104]
Carbon foam/Fe2O3 0.2 6.0 M KOH 225 [105]
N-doped carbon nanofiber/Fe3C/Fe2O3 2 2.0 M KOH 590.1 [106]
Fe2O3/P(2MA) 0.6 0.5 M HCl 998.4 This work

3.6.2 CV profiles

Figure S6 presents the CV curves of P(2MA) and Fe2O3/P(2MA) electrodes. The CV curve was recorded at a scan rate of 30 mV/s in 0.5 M HCl electrolyte over a potential window of 0.0–1.0 V. The Fe2O3/P(2MA) electrode exhibits a significantly higher current response and a larger enclosed area under the CV curve than the pure P(2MA) electrode. The reaction rate of an electrochemical process is proportional to the electrochemically active surface area (ECSA) of the electrode, as a larger surface area provides more active sites for the reaction. The ECSA of an electrode can be evaluated using CV, where a greater area under the CV curve typically indicates a larger surface area [27,49]. This result suggests that the spongy structure of the Fe2O3/P(2MA) electrode contributes to its high surface area, leading to superior charge storage capacity and specific capacitance.

Figure 7(a) and (b) depicts the CV of the Fe2O3/P(2MA) nanohybrid in NaCl and HCl electrolytes within a voltage range of 0–1 V, using different scan rates ranging from 30 to 500 mV/s. The surface area under the CV curve increases with an increase in the scan rate, indicating the desired capacitive behavior of the electrode. At higher scan rates, the diffusion rate of electrolytic ions becomes more significant than the reaction rate. As a result, many electrolytic ions reach the electrode/electrolyte interface, while only a tiny fraction of these ions participate in the charge transfer reaction. According to the Randles–Sevcik equation, this increases the peak current in the CV curve and the area under the CV curve at higher scan rates. The slow scan rate provides sufficient time for the electrolyte to penetrate all of the pores and fully interact with the electrode. In the HCl electrolyte, the CV curves of the nanohybrid exhibit stronger oxidation and reduction currents compared to NaCl at the same scan rates, as shown in Figure 7(a,b).

Figure 7 
                     CV in (a) 0.5 M NaCl and (b) 0.5 M HCl at different scan rates for the Fe2O3/P(2MA) nanohybrid.
Figure 7

CV in (a) 0.5 M NaCl and (b) 0.5 M HCl at different scan rates for the Fe2O3/P(2MA) nanohybrid.

Moreover, the HCl electrolyte shows a larger surface area under the CV curve, indicating enhanced electrochemical activity in HCl. This observation suggests that the specific cationic species in the electrolyte solution play a crucial role in these electrochemical processes. As seen in Figure 7(b), the integral area of the CV curve for the Fe2O3/P(2MA) nanohybrid in 0.5 M HCl at a scan rate of 500 mV/s is significantly larger. This result indicates a greater capacity for electrochemical energy storage in the HCl electrolyte.

The following equation is used to determine specific capacitance based on CV data [107,108]:

(8) C wt = 4 ms V v 1 v n i d v ,

where v 1 v n i d v represents the integrated area under the CV curve, s represents the scan rate (V/s) applied, and the active material mass on each electrode is denoted m(g).  V is double the potential range covered in the CV measurement. It starts from V1, reaches V2, and then returns to V1. At a scan rate of 500 mV/s, the specific capacitance of the nanohybrid in 0.5 M NaCl electrolyte was determined to be 51.4 F g−1 using the CV integral area. In contrast, the specific capacitance of the nanohybrid was measured to be significantly higher at 225 F g−1 in a 0.5 M HCl electrolyte.

3.6.3 Ragone plot

The Ragone plot is a graphical representation used to compare the energy density of various energy storage devices. It shows the amount of energy that can be stored relative to the power output of each device. In this plot, the specific energy (energy per unit mass) is plotted on the vertical axis, while the specific power (power per unit mass) is plotted on the horizontal axis. The Ragone plot is utilized to examine the electrochemical performance of the Fe2O3/P(2MA) nanohybrid electrode in NaCl and HCl electrolytic solutions, as presented in Figure 8(a) and (b). The specific energy and power of the nanohybrid electrode can be calculated using the following equations [109,110]:

(9) E wt ( W h kg 1 ) = 0.0347 C wt ( E ) 2 ,

(10) P wt ( W kg 1 ) = 3,600 E wt t .

Here, E represents the discharge potential after considering the IR drop, and t denotes the discharge time. In the NaCl electrolyte, the Fe2O3/P(2MA) nanohybrid electrode exhibits average specific energy and power values of 0.13  and 180 W kg−1, respectively. However, higher values of 3.5and 140 W kg−1 are obtained using the HCl electrolyte. Utilizing the Fe2O3/P(2MA) nanohybrid in an HCl electrolyte for energy storage is more efficient. This is primarily due to the material’s significantly higher specific energy than its NaCl counterpart.

Figure 8 
                     Ragon plot profile of the Fe2O3/P(2MA) nanohybrid in (a) NaCl and (b) HCl electrolytes.
Figure 8

Ragon plot profile of the Fe2O3/P(2MA) nanohybrid in (a) NaCl and (b) HCl electrolytes.

3.6.4 Impedance

Electrochemical impedance spectroscopy (EIS) is a versatile technique for analyzing electrochemical systems, offering insights into charge transfer rates, redox reactions, diffusion processes, and capacitive behavior. EIS expresses impedance as a combination of real and imaginary components, commonly visualized using a Nyquist plot, where the real impedance is plotted on the X-axis and the negative imaginary impedance on the Y-axis [111]. Figure 9(a) presents the Nyquist plot of the Fe2O3/P(2MA) nanohybrid electrode, recorded over a frequency range from 10 MHz to 100 kHz in a 0.5 M HCl electrolyte. The plot employs orthonormal scaling for clear visualization. In the low-frequency region, the Nyquist plot exhibits nearly linear behavior (right side of the plot). At the same time, a distinct semicircular arc appears in the high-frequency region (left side of the plot). The diameter of the semicircle corresponds to the charge-transfer resistance, while the slope of the low-frequency tail reflects ion transport through the electrode/electrolyte interface [112].

Figure 9 
                     (a) Nyquist plot with equivalent circuit and (b) cyclic stability of the Fe2O3/P(2MA) nanohybrid in HCl.
Figure 9

(a) Nyquist plot with equivalent circuit and (b) cyclic stability of the Fe2O3/P(2MA) nanohybrid in HCl.

The EIS data were fitted using the EC-Lab software to a modified Randle’s circuit comprising four resistive elements, one double-layer capacitance, and two constant phase elements. The extracted values, listed in Table S1, help interpret the physical and electrochemical processes occurring in the electrode. R 1 (2.473 Ω) represents the solution resistance, and its low value confirms the good ionic conductivity of the electrolyte. R 2 (0.10 Ω) corresponds to the charge-transfer resistance at the polymer–electrolyte interface. This very low resistance highlights the strong interfacial coupling between Fe2O3 and the conductive P(2MA) matrix, contributing to the nanohybrid’s improved electrical conductivity and electrochemical stability [113]. R 3 (49.13 Ω) and R 4 (1,574 Ω) are associated with ion diffusion and faradaic resistance, respectively. The capacitance C 1 (15.56 μF) reflects the double-layer capacitance, indicating good electrochemical interface formation. The constant phase elements Q 1 (2.22 F s) and Q 2 (0.02 mF s) with exponents a 1 (0.55) and a 2 (0.59) represent non-ideal capacitive behavior due to surface roughness and porosity in the electrode structure. Hence, the EIS results strongly support the electrochemical behavior observed in CV and GCD tests and are consistent with the porous nanofiber morphology. These characteristics make the Fe2O3/P(2MA) nanohybrid a strong candidate for high-performance electrochemical energy storage systems.

3.6.5 Stability

The cycling stability of the Fe2O3/P(2MA) nanohybrid electrode was evaluated in a 0.5 M HCl electrolyte at a specific current density of 0.6 A g−1. The stability was assessed by monitoring the preservation of gravimetric capacitance and the capacitance retention (y-axis) in response to charge–discharge cycles for up to 2,000 cycles, as shown in Figure 9(b). The inset in the figure compares the first cycle to the final cycle. The nanohybrid electrode exhibited 100% cycling stability during the first 300 cycles. After 2,000 cycles, the nanohybrid electrode achieved a remarkable cycling stability of approximately 99.9% (998 F g−1) compared to its initial specific capacitance of 998.4 F g−1 (100%). This implies that the Fe2O3/P(2MA) nanohybrid electrode exhibits good cycling electrochemical stability.

The observed long-term cycling stability of the Fe2O3/P(2MA) nanohybrid arises from the synergistic interplay between its structural architecture and chemical resilience. The P(2MA) matrix is a redox-active yet chemically robust host. Its intrinsically conductive backbone facilitates rapid charge transfer while preserving the integrity of the polymer chain. As demonstrated by Rahayu et al., P(2MA) maintains high electrical conductivity in 0.5 M HCl without undergoing acid-induced degradation, confirming its electrochemical stability under harsh conditions [114]. Simultaneously, Fe2O3 offers corrosion resistance, substantially reducing the risk of metal-ion leaching during prolonged operation. As confirmed by TEM and SEM imaging, the formation of a spongy nanofiber architecture in the Fe2O3/P(2MA) matrix reveals that Fe2O3 nanoparticles are tightly embedded within the polymer framework. This structure reflects strong interfacial adhesion between Fe2O3 and P(2MA), which effectively suppresses nanoparticle agglomeration, mitigates structural disintegration, and reduces the risk of mechanical detachment during repetitive redox cycling.

Furthermore, the flexible polymeric scaffold accommodates volumetric changes during Fe3+/Fe2+ redox transitions, preventing microcracking during long-term cycling. Notably, the polymer matrix also serves as a protective sheath, shielding Fe2O3 from direct contact with the acidic electrolyte and enhancing the overall system stability [115]. This architectural strategy aligns with previous studies that demonstrate improved durability of MOs encapsulated in polymer matrices [27,115,116,117]. Using 0.5 M HCl as the electrolyte maximizes capacitance and helps mitigate localized electrolyte depletion at the electrode-electrolyte interface. The low charge-transfer resistance observed from EIS measurements, even after 2,000 cycles, indicates that the electronic and ionic transport pathways remain unobstructed, thereby validating the structural and electrochemical integrity of the nanohybrid over prolonged cycling. Additionally, the device was meticulously sealed to prevent any electrolyte loss due to evaporation (Figure S7), ensuring stable ionic conductivity throughout the testing period. This suggests that the Fe2O3/P(2MA) nanohybrid electrode is highly suitable as an effective electrode material in supercapacitor applications.

The post-PEC SEM analysis confirmed the structural integrity of the Fe2O3/P(2MA) nanohybrid after the electrochemical measurements, as illustrated in Figure S8. The electrode maintained its morphology with no microcracking or particle agglomeration. Additionally, it remained well-adhered to the substrate without degradation. These observations indicate strong corrosion resistance and affirm the long-term mechanical stability of the electrode.

The superior performance of the Fe2O3/P(2MA) nanohybrid arises from multiple synergistic factors, as supported by structural, morphological, and electrochemical evidence. P(2MA) provides a highly conductive polymeric backbone enriched with redox-active sites. A methyl substituent enhances charge delocalization and offers steric protection against overoxidation. The flexible polymer chains also buffer volumetric fluctuations during charge–discharge cycling, contributing to mechanical stability and improved capacitance retention [98]. Fe2O3 offers high theoretical capacitance due to its reversible Fe3+/Fe2+ redox transitions, but is limited by low electrical conductivity and a tendency to agglomerate when used alone. However, when Fe2O3 nanoparticles are generated in situ during photopolymerization and uniformly embedded within the P(2MA) matrix, they contribute significant pseudocapacitance while avoiding aggregation. This integration results in an interpenetrating nanofiber network with high surface area, shortened ion-diffusion pathways, and improved electrolyte accessibility to active sites [22]. SEM and TEM analyses confirm the spongy nanostructure, which is crucial for rapid ion transport and enhanced rate capability. Furthermore, the one-pot photopolymerization process ensures intimate polymer–oxide interfacial contact, reducing interfacial resistance and enhancing electron mobility. The EIS supports this finding, showing low charge-transfer resistance, indicative of efficient charge transport within the nanohybrid structure.

4 Conclusions

This work successfully presents a one-pot photopolymerization strategy for synthesizing Fe2O3/P(2MA) nanohybrid aimed at high-performance supercapacitor applications. The iron precursor acts dually as both the oxidant and the source of Fe2O3, simplifying fabrication while promoting strong polymer–oxide integration. The combination of photochemical and oxidative polymerization successfully yields robust, interpenetrating Fe2O3/P(2MA) spongy nanofibers, as confirmed by structural and morphological analyses. The resulting architecture exhibits excellent electrochemical performance. Tests were conducted in HCl, KOH, and NaCl electrolytes. The Fe2O3/P(2MA) nanohybrid delivers a specific capacitance of 998.4 F g−1 at 0.6 A g−1 in 0.5 M HCl, with energy and power densities of 3.5 W hkg−1 and 140 W kg−1, respectively. Remarkably, it retains 99.9% of its initial capacitance after 2,000 cycles, demonstrating exceptional long-term stability. These findings underscore the novelty, effectiveness, and scalability of the synthesis method, establishing Fe2O3/P(2MA) as a promising electrode material for next-generation supercapacitor technologies.



  1. Funding information: This work was supported and funded by the Deanship of Scientific Research at Imam Mohammad Ibn Saud Islamic University (IMSIU) (grant number IMSIU-DDRSP2503).

  2. Author contributions: A.M. Ahmed: writing – review and editing, software, visualization, and funding acquisition. D. Essam: methodology, writing – original draft, visualization, and software. M.A. Basyooni-M. Kabatas: review and editing, formal analysis, data curation, and software. A.A. Abdel-Khaliek: investigation, supervision, and formal analysis. M. Shaban: resources, visualization, and validation. M.J. Aljaafreh: data curation, visualization, and formal analysis. M. Rabia: conceptualization, supervision, and writing – review and editing. All authors have accepted responsibility for the entire content of this manuscript and approved its submission.

  3. Conflict of interest: The authors state no conflict of interest.

  4. Data availability statement: The datasets generated and/or analyzed during the current study are available from the corresponding author on reasonable request.

References

[1] Abdah MAAM, Azman NHN, Kulandaivalu S, Sulaiman Y. Review of the use of transition-metal-oxide and conducting polymer-based fibres for high-performance supercapacitors. Mater Des. 2020;186:108199.Suche in Google Scholar

[2] Mao X, Liu H, Xu J, Mengwei L, Yang W. Surface reconstruction of Co3O4/rGO heterointerface enabling high-performance asymmetric supercapacitors. J Energy Storage. 2024;102:114128.10.1016/j.est.2024.114128Suche in Google Scholar

[3] Thalji MR, Ali GAM, Shim J-J, Chong KF. Cobalt-doped tungsten suboxides for supercapacitor applications. Chem Eng J. 2023;473:145341.10.1016/j.cej.2023.145341Suche in Google Scholar

[4] Atta A, Abdelhamied MM, Essam D, Shaban M, Alshammari AH, Rabia M. Structural and physical properties of polyaniline/silver oxide/silver nanocomposite electrode for supercapacitor applications. Int J Energy Res. 2022;46(5):6702–10.10.1002/er.7608Suche in Google Scholar

[5] Essam D, Ahmed AM, Abdel-khaliek AA, Shaban M. One pot synthesis of poly m-toluidine incorporated silver and silver oxide nanocomposite as a promising electrode for supercapacitor devices. 2025;15:2698.10.1038/s41598-024-84848-5Suche in Google Scholar PubMed PubMed Central

[6] Huang Y, Li H, Wang Z, Zhu M, Pei Z, Xue Q, et al. Nanostructured Polypyrrole as a flexible electrode material of supercapacitor. Nano Energy. 2016;22:422–38. 10.1016/j.nanoen.2016.02.047.Suche in Google Scholar

[7] Afif A, Rahman SM, Tasfiah Azad A, Zaini J, Islam MA, Azad AK. Advanced materials and technologies for hybrid supercapacitors for energy storage – A review. J Energy Storage. 2019;25(July):100852. 10.1016/j.est.2019.100852.Suche in Google Scholar

[8] Karri SN, Ega SP, Perupogu V, Srinivasan P. Enhancing the electrochemical performance of polyaniline using fly ash of coal waste for supercapacitor application. ChemistrySelect. 2021;6(10):2576–89.10.1002/slct.202100513Suche in Google Scholar

[9] Sun J, Luo B, Li H. A review on the conventional capacitors, supercapacitors, and emerging hybrid ion capacitors: past, present, and future. Adv Energy Sustainable Res. 2022;3(6):2100191.10.1002/aesr.202100191Suche in Google Scholar

[10] Mao X, Hatton T, Rutledge G. A review of electrospun carbon fibers as electrode materials for energy storage. Curr Org Chem. 2013;17(13):1390–401.10.2174/1385272811317130006Suche in Google Scholar

[11] Shoeb M, Fouzia M, Khan MN, Kim B-J, Jeong C. Waste to energy strategy: Graphene-supported Au-Ag2O polyIndole nanocomposites for antimony adsorption and their sequential utilization in supercapacitors device. Sep Purif Technol. 2024;329:123080.10.1016/j.seppur.2024.128656Suche in Google Scholar

[12] Zhou Q, Wei T, Yue J, Sheng L, Fan Z. Polyaniline nanofibers confined into graphene oxide architecture for high-performance supercapacitors. Electrochim Acta. 2018;291:234–41. 10.1016/j.electacta.2018.08.104.Suche in Google Scholar

[13] Kim BH, Kim CH, Yang KS, Rahy A, Yang DJ. Electrospun vanadium pentoxide/carbon nanofiber composites for supercapacitor electrodes. Electrochim Acta. 2012;83:335–40. 10.1016/j.electacta.2012.07.093.Suche in Google Scholar

[14] Guo FM, Xu RQ, Cui X, Zhang L, Wang KL, Yao YW, et al. High performance of stretchable carbon nanotube-polypyrrole fiber supercapacitors under dynamic deformation and temperature variation. J Mater Chem A. 2016;4(23):9311–8.10.1039/C6TA02437HSuche in Google Scholar

[15] Balan BK, Chaudhari HD, Kharul UK, Kurungot S. Carbon nanofiber-RuO2-poly(benzimidazole) ternary hybrids for improved supercapacitor performance. RSC Adv. 2013;3(7):2428–36.10.1039/c2ra22776bSuche in Google Scholar

[16] Radhamani AV, Shareef KM, Rao MSR. ZnO@MnO2 core-shell nanofiber cathodes for high performance asymmetric supercapacitors. ACS Appl Mater Interfaces. 2016;8(44):30531–42.10.1021/acsami.6b08082Suche in Google Scholar PubMed

[17] Yang Y, Yang F, Hu H, Lee S, Wang Y, Zhao H, et al. Dilute NiO/carbon nanofiber composites derived from metal organic framework fibers as electrode materials for supercapacitors. Chem Eng J. 2017;307:583–92.10.1016/j.cej.2016.08.132Suche in Google Scholar

[18] Zhou H, Liu Y, Ren M, Zhai HJ. Mechanically exfoliated graphite paper with layered microstructures for enhancing flexible electrochemical energy storage. Inorg Chem Front. 2022;9:1346–57.10.1039/D1QI01601FSuche in Google Scholar

[19] Bashir T, Shakoor A, Ahmed E, Niaz NA, Akhtar MS, Raza MR, et al. Polypyrrole-Fe2O3 Nanocomposites with high dielectric constant: In situ chemical polymerisation. Polym Polym Compos. 2018;26(3):233–41.10.1177/096739111802600303Suche in Google Scholar

[20] Liao M, Zhang K, Yan W, Yue H, Luo C, Wu G, et al. Constructing amorphous/crystalline heterointerface in cobalt sulfide for high-performance supercapacitors. J Power Sources. 2025;625:235663.10.1016/j.jpowsour.2024.235663Suche in Google Scholar

[21] Gamal A, Shaban M, BinSabt M, Moussa M, Ahmed AM, Rabia M, et al. Facile fabrication of polyaniline/pbs nanocomposite for high-performance supercapacitor application ahmed. Micromachines. 2022;1:17.10.3390/nano12050817Suche in Google Scholar PubMed PubMed Central

[22] Ahmad SA, Fouzia M, Ansari MA, Alothman AA, Shoeb M, Jeong C. Morus alba-mediated defect-rich Fe-doped V2O5 nanoparticles on woven carbon fiber for high-performance supercapacitor with enhanced charge storage mechanism. J Alloy Compd. 2025;1020:179515.10.1016/j.jallcom.2025.179515Suche in Google Scholar

[23] Fouzia M, Shoeb M, Khan JA, Gondal MA, Jeong C. Chemical reduction-induced defect-rich and synergistic effects of reduced graphene oxide based Cu-doped NiO nanocomposite (RGO@Cu-NiO NCs) decorated on woven carbon fiber for supercapacitor device and their charge storage mechanism. J Energy Storage. 2024;104:114578.10.1016/j.est.2024.114578Suche in Google Scholar

[24] Mohd Abdah MAA, Azman NHN, Kulandaivalu S, Sulaiman Y. Review of the use of transition-metal-oxide and conducting polymer-based fibres for high-performance supercapacitors. Mater Des. 2020;186:108199. 10.1016/j.matdes.2019.108199 Suche in Google Scholar

[25] Lu XF, Chen XY, Zhou W, Tong YX, Li GR. α-Fe2O3@PANI core–shell nanowire arrays as negative electrodes for asymmetric supercapacitors. ACS Appl Mater Interfaces. 2015;7(27):14843–50.10.1021/acsami.5b03126Suche in Google Scholar PubMed

[26] Shinde SK, Karade SS, Truong NTN, Veer SS, Shaikh SF, Al-Enizi AM, et al. Highly stable and thin layer of conducting PANI decorative on porous 3D nanoflakes like-NiCo2O4/CF nanomaterials for efficient Hybrid Supercapacitors application. J Energy Storage. 2023;78:109960.10.1016/j.est.2023.109960Suche in Google Scholar

[27] Fathy A, Ahmed AM, Ghanem MA, Hamdy H, Shaban M, Mohammed KMH, et al. Supercapacitance performance of novel core/shell poly (m-toluidine)/( Co-Ni) nanoplatelets/nanoneedles nanocomposite. J Mol Liq. 2023;386:122466.10.1016/j.molliq.2023.122466Suche in Google Scholar

[28] Bathula C, Reddy R, Naik S, Naushad M, Kim H. Deep eutectic solvent-assisted synthesis of transition metal oxide and its integration into polymer matrix for supercapacitor. J Mol Liq. 2024;403(March):124899. 10.1016/j.molliq.2024.124899.Suche in Google Scholar

[29] Li Y, Zhang Z, Chen Y, Chen H, Fan Y, Li Y, et al. Facile synthesis of a Ni-based NiCo2O4-PANI composite for ultrahigh specific capacitance. Appl Surf Sci. 2020;506:144646.10.1016/j.apsusc.2019.144646Suche in Google Scholar

[30] Dhineshkumar S, Rajkumar S, Sathiyan A, Merlin JP. Fabrication of ­ Ag 2 WO 4/PANI composite with enhanced supercapacitor performance. J Mater Sci Mater Electron. 2024;35(20):1–15. 10.1007/s10854-024-13202-2.Suche in Google Scholar

[31] Sivakumar C, Nian J, Teng H. Poly (o -toluidine) for carbon fabric electrode modification to enhance the electrochemical capacitance and conductivity. J Power Sources. 2005;144:295–301.10.1016/j.jpowsour.2004.12.010Suche in Google Scholar

[32] Andriianova AN, Biglova YN. Effect of structural factors on the physicochemical properties of functionalized polyanilines. 2020;10:7468–91.10.1039/C9RA08644GSuche in Google Scholar PubMed PubMed Central

[33] Bober P, Trchová M, Prokeš J, Varga M, Stejskal J. Polyaniline–silver composites prepared by the oxidation of aniline with silver nitrate in solutions of sulfonic acids. Electrochim Acta. 2011;56(10):3580–5.10.1016/j.electacta.2010.08.041Suche in Google Scholar

[34] Rabia M, Elsayed AM, Alnuwaiser MA. Decoration of poly-3-methyl aniline with As ( III) oxide and hydroxide as an effective photoelectrode for electroanalytical photon sensing with photodiode-like behavior. Micromachines. 2023;14(iii).10.3390/mi14081573Suche in Google Scholar PubMed PubMed Central

[35] Abdelazeez AAA, Hadia NMA, Mourad AI, El-fatah GA, Shaban M, Ahmed AM, et al. Effect of Au plasmonic material on Poly M-toluidine for photoelectrochemical hydrogen generation from sewage water. Polymers. 2022;14:17.10.3390/polym14040768Suche in Google Scholar PubMed PubMed Central

[36] Pujari SS, Bobade RG, Shaikh SF, Al-Enizi AM, Ambare RC, Lokhande BJ. A binderless Ru:α-Fe2O3 binary nanocomposite electrode for supercapacitor applications. J Mater Sci Mater Electron. 2024;35:2162–75.10.1007/s10854-024-13857-xSuche in Google Scholar

[37] Sephra PJ, Baraneedharan P, Sivakumar M, Thangadurai TD, Nehru K. In situ growth of hexagonal-shaped α-Fe2O3 nanostructures over few layered graphene by hydrothermal method and their electrochemical performance. J Mater Sci Mater Electron. 2018;29(8):6898–908.10.1007/s10854-018-8676-1Suche in Google Scholar

[38] Shaban M, Binsabt M, Ahmed AM. Recycling rusty iron with natural zeolite heulandite to create a unique nanocatalyst for green hydrogen production. Nanomaterials. 2021;11(12):3445.10.3390/nano11123445Suche in Google Scholar PubMed PubMed Central

[39] Phakkhawan A, Suksangrat P, Srepusharawoot P, Ruangchai S, Pawinee K, Pimanpang S, et al. Reagent-and solvent-mediated Fe2O3 morphologies and electrochemical mechanism of Fe2O3 supercapacitors. J Alloy Compd. 2022;919:165702.10.1016/j.jallcom.2022.165702Suche in Google Scholar

[40] Lu PA, Manikandan MR, Yang PF, He YL, Yang F, Dang ST, et al. Synthesis, analysis and characterization of alpha-Fe2O3 nanoparticles and their applications in supercapacitors. 2023;826:1–13.10.1007/s10854-023-10246-8Suche in Google Scholar

[41] Singh U, Ramadesigan V, Chakraborty AK, Shukla S, Saxena S. High-performance α-Fe2O3 nano cubes as anode for supercapacitors. J Energy Storage. 2024;102:114124.10.1016/j.est.2024.114124Suche in Google Scholar

[42] Ponnarasi K, Elangovan P, Surender S. Results in chemistry development of CNT/Fe2O3 nanocomposites as electrode materials for advanced supercapacitors. Results Chem. 2024;7(December 2023):101355. 10.1016/j.rechem.2024.101355.Suche in Google Scholar

[43] Palem RR, Ramesh S, Yadav HM, Kim JH, Sivasamy A, Kim HS, et al. Nanostructured Fe2O3@nitrogen-doped multiwalled nanotube/cellulose nanocrystal composite material electrodes for high-performance supercapacitor applications. J Mater Res Technol. 2020;9(4):7615–27. 10.1016/j.jmrt.2020.05.058.Suche in Google Scholar

[44] Guo G, Su Q, Zhou W, Wei M, Wang Y. Cycling stability of Fe2O3 nanosheets as supercapacitor sheet electrodes enhanced by MgFe2O4 nanoparticles. RSC Adv. 2023;13(6):3643–51.10.1039/D2RA07383HSuche in Google Scholar PubMed PubMed Central

[45] Lorkit P, Panapoy M, Ksapabutr B. Iron oxide-based supercapacitor from ferratrane precursor via sol-gel-hydrothermal process. Energy Proc. 2014;56(C):466–73.10.1016/j.egypro.2014.07.180Suche in Google Scholar

[46] Azimov F, Kim J, Choi SM, Jung HM. Synergistic effects of Fe2O3 nanotube/polyaniline composites for an electrochemical supercapacitor with enhanced capacitance. Nanomaterials. 2021;11(6):1557.10.3390/nano11061557Suche in Google Scholar PubMed PubMed Central

[47] Nandagudi A, Nagarajarao SH, Santosh MS, Basavaraja BM, Malode SJ, Mascarenhas RJ. Hydrothermal synthesis of transition metal oxides, transition metal oxide/carbonaceous material nanocomposites for supercapacitor applications. Mater Today Sustainable. 2022;19:100214.10.1016/j.mtsust.2022.100214Suche in Google Scholar

[48] ur Rehman MN, Munawar T, Nadeem MS, Mukhtar F, Maqbool A, Riaz M, et al. Facile synthesis and characterization of conducting polymer-metal oxide based core-shell PANI-Pr2O–NiO–Co3O4 nanocomposite: As electrode material for supercapacitor. Ceram Int. 2021;47(13):18497–509.10.1016/j.ceramint.2021.03.173Suche in Google Scholar

[49] Zhu P, Zhao Y. Cyclic voltammetry measurements of electroactive surface area of porous nickel: Peak current and peak charge methods and diffusion layer effect. Mater Chem Phys. 2019;233:60–7.10.1016/j.matchemphys.2019.05.034Suche in Google Scholar

[50] Hwang JJ, Soto C, Lafaurie D, Stephen M, Sarno DM. Porous microspheres of polyaniline, poly(o-toluidine), and poly(m-toluidine) prepared from double emulsions stabilized by toluidine isomers as the single surfactant. J Colloid Interface Sci. 2018;513:331–41. 10.1016/j.jcis.2017.11.029.Suche in Google Scholar PubMed

[51] Ramli RA. Hollow polymer particles: A review. RSC Adv. 2017;7(83):52632–50.10.1039/C7RA10358ASuche in Google Scholar

[52] Lu Y, Chen C, Zhang X. Functional Nanofibers for Energy Storage. In: Handbook of smart textiles. Singapore: Springer; 2015. p. 513–47.10.1007/978-981-4451-45-1_18Suche in Google Scholar

[53] Dhahebi A, Convergence N, Mohammed A, Dhahebi A, Chandra S, Gopinath B. Graphene impregnated electrospun nanofiber sensing materials: a comprehensive overview on bridging laboratory set ‑ up to industry. Nano Convergence. 2020;7:27. 10.1186/s40580-020-00237-4.Suche in Google Scholar PubMed PubMed Central

[54] Li Y, Fu Z, Su B. Hierarchically structured porous materials for energy conversion and storage. Adv Funct Mater. 2012;22(22):4634–67.10.1002/adfm.201200591Suche in Google Scholar

[55] Manuscript A. Environmental Science.Suche in Google Scholar

[56] Article R. Chem Soc Rev materials from energy storage and conversion. 2016.Suche in Google Scholar

[57] Han S, Wu D, Li S, Zhang F, Feng X. Porous graphene materials for advanced electrochemical energy storage and conversion devices. Adv Mater. 2014;26(6):849–64.10.1002/adma.201303115Suche in Google Scholar PubMed

[58] Feng N, Gao M, Zhong J, Gu C, Zhang Y, Liu B. Stable structure and fast ion diffusion: a flexible MoO2@carbon hollow nanofiber film as a binder-free anode for sodium-ion batteries with superior kinetics and excellent rate capability. Polymers. 2024;16(11):1452.10.3390/polym16111452Suche in Google Scholar PubMed PubMed Central

[59] Liu M, Zhang P, Qu Z, Yan Y, Lai C, Liu T, et al. Ion battery. Nat Commun. 2019;10:3917. 10.1038/s41467-019-11925-z.Suche in Google Scholar PubMed PubMed Central

[60] Yang S, Zhao S, Chen S. Chemical science recent advances in electrospinning nano fiber materials for aqueous zinc ion batteries. Chem Sci. 2023;14:13346–66.10.1039/D3SC05283DSuche in Google Scholar

[61] Zhang X, Zhang J, Song W, Liu Z. Controllable synthesis of conducting polypyrrole nanostructures. J Phys Chem. 2006;110(3):1158–65.10.1021/jp054335kSuche in Google Scholar PubMed

[62] Pan L, Qiu H, Dou C, Li Y, Pu L, Xu J, et al. Conducting polymer nanostructures: template synthesis and applications in energy storage. 2010;11(7):2636–57.10.3390/ijms11072636Suche in Google Scholar PubMed PubMed Central

[63] Ali MA, Kim HH, Lee CY, Soh HS, Lee JG, et al. Effects of the FeCl3 concentration on the polymerization of conductive poly(3,4-ethylenedioxythiophene) thin films on (3-aminopropyl) trimethoxysilane monolayer-coated SiO2 surfaces. Met Mater Int. 2009;15(6):977–81.10.1007/s12540-009-0977-8Suche in Google Scholar

[64] Suhasini A, Kumar KPV, Maiyalagan T. Synthesis, thermal and magnetic behavior of iron oxide-polymer nanocomposites. Sci Eng Compos Mater. 2018;25(1):189–95.10.1515/secm-2015-0531Suche in Google Scholar

[65] Georgia Institute of Technology. Researchers find new mechanism governing particle growth in nanocomposites. Rockville, MD, USA: ScienceDaily, LLC; 2005. p. 1–4.Suche in Google Scholar

[66] Kumar A. Nanoarchitectured transition metal oxides and their composites for supercapacitors. Electrochem Sci Adv. 2022;2(6):e2100187.10.1002/elsa.202100187Suche in Google Scholar

[67] Balasubramanian N, Prabhu S, Sakthivel N, Ramesh R. Electrochemical performance of Fe2O3@PPy nanocomposite as an effective electrode material for supercapacitor electrochemical performance of Fe2O3@PPy nanocomposite as an effective electrode material for supercapacitor. ECS J Solid State Sci Technol. 2022;11(9):091001.10.1149/2162-8777/ac8837Suche in Google Scholar

[68] Katowah DF, Hussein MA, Alam MM, Sobahi TR, Gabal MA, Asiri AM, et al. Poly(pyrrole-co-o-toluidine) wrapped CoFe2O4/R(GO-OXSWCNTs) ternary composite material for Ga3+ sensing ability. RSC Adv. 2019;9(57):33052–70.10.1039/C9RA03593ASuche in Google Scholar

[69] Mahudeswaran A, Vivekanandan J, Jeeva. A. Synthesis, characterization, optical and electrical properties of nanostructured poly(aniline-co-o-bromoaniline) prepared by in-situ polymerization method. Optik. 2016;127:3984–8.10.1016/j.ijleo.2016.01.095Suche in Google Scholar

[70] Haydar MS, Ali S, Mandal P, Roy D, Roy MN, Kundu S, et al. Fe–Mn nanocomposites doped graphene quantum dots alleviate salt stress of Triticum aestivum through osmolyte accumulation and antioxidant defense. Sci Rep. 2023;13(1):1–26. 10.1038/s41598-023-38268-6.Suche in Google Scholar PubMed PubMed Central

[71] Liu L, Fu S, Lv X, Yue L, Fan L, Yu H, et al. A gas sensor With Fe2O3 Nanospheres based on trimethylamine detection for the rapid assessment of spoilage degree in fish. Front Bioeng Biotechnol. 2020;8:567584.10.3389/fbioe.2020.567584Suche in Google Scholar PubMed PubMed Central

[72] Altowyan AS, Shaban M, Abdelkarem K, El Sayed AM. The impact of co doping and annealing temperature on the electrochemical performance and structural characteristics of SnO2 nanoparticulate photoanodes. Material. 2022;15(19):6534.10.3390/ma15196534Suche in Google Scholar PubMed PubMed Central

[73] Correa CM, Faez R, Bizeto MA, Camilo FF. One-pot synthesis of a polyaniline-silver nanocomposite prepared in ionic liquid. RSC Adv. 2012;2(7):3088–93.10.1039/c2ra00992gSuche in Google Scholar

[74] Annamalai R. Synthesis, characterization and electrical properties of nano metal and metal-oxide doped with conducting polymer composites by in-situ chemical polymerization. MOJ Polym Sci. 2017;1(6):4–7.10.15406/mojps.2017.01.00031Suche in Google Scholar

[75] Rasheed MH, Hashim FS, Abass KH. Impact of Fe2O3 nanoparticles on the spectral and optical properties of electrospun nanofibrous poly vinyl alcohol/poly acrylamide. SSRN Electron J. 2023. 10.2139/ssrn.4403005.Suche in Google Scholar

[76] Soflaee F, Farahmandjou M, Firoozabadi TP. Polymer-mediated synthesis of iron oxide (Fe2O3) nanorod. Chin J Phys. 2015;53(4):178–86.Suche in Google Scholar

[77] Wang H, Wen H, Hu B, Fei G, Shen Y, Sun L, et al. Facile approach to fabricate waterborne polyaniline nanocomposites with environmental benignity and high physical properties. Sci Rep. 2017;7(January):1–12.10.1038/srep43694Suche in Google Scholar PubMed PubMed Central

[78] Wadatkar NS, Waghuley SA. Characterizing the electro-optical properties of polyaniline/poly(vinyl acetate) composite films as-synthesized through chemical route. Results Surf Interfaces. 2021;4(June):100016. 10.1016/j.rsurfi.2021.100016.Suche in Google Scholar

[79] Mirzaei A, Janghorban K, Hashemi B, Hosseini SR, Bonyani M, Leonardi SG, et al. Synthesis and characterization of mesoporous α-Fe2O3 nanoparticles and investigation of electrical properties of fabricated thick films. Process Appl Ceram. 2016;10(4):209–18.10.2298/PAC1604209MSuche in Google Scholar

[80] Mohamed F, Hassaballa S, Shaban M, Ahmed AM. Highly efficient photocatalyst fabricated from the chemical recycling of iron waste and natural zeolite for super dye degradation. Nanomaterials. 2022;12(2):235.10.3390/nano12020235Suche in Google Scholar PubMed PubMed Central

[81] Alenad AM, Taha TA, Zayed M, Gamal A, Shaaban M, Ahmed AM. Impact of carbon nanotubes concentrations on the performance of carbon nanotubes/zinc oxide nanocomposite for photoelectrochemical water splitting. J Electroanal Chem. 2023;943:117579.10.1016/j.jelechem.2023.117579Suche in Google Scholar

[82] Atta A, Abdelhamied MM, Abdelreheem AM, Berber MR. Flexible methyl cellulose/polyaniline/silver composite films with enhanced linear and nonlinear optical properties. Polymers. 2021;13(8):1–16.10.3390/polym13081225Suche in Google Scholar PubMed PubMed Central

[83] Wang H, Mao J, Zhang Z, Zhang Q, Zhang L, Zhang W, et al. Photocatalytic degradation of deoxynivalenol over dendritic-like α-Fe2O3 under visible light irradiation. Toxins. 2019;11(2):1–9.10.3390/toxins11020105Suche in Google Scholar PubMed PubMed Central

[84] Zhan T, Ding G, Cao W, Li J, She X, Teng H. Amperometric sensing of catechol by using a nanocomposite prepared from Ag/Ag2O nanoparticles and N,S-doped carbon quantum dots. Microchim Acta. 2019;186(11):743.10.1007/s00604-019-3848-0Suche in Google Scholar PubMed

[85] Martins AFN, Diniz FB, Rodrigues AR. Reaction between Fe3+and aniline in the synthesis of PANI-γFe₂O₃ and PANI-Fe₃O₄ nanocomposites: Mechanistic studies and evaluation of parameters. Nano-Struct Nano-Objects. 2025;42:101477.10.1016/j.nanoso.2025.101477Suche in Google Scholar

[86] EMI shielding performance of graphene oxide reinforced polyaniline/polystyrene solution cast thin films. Synth Met. 2023;296:117369.10.1016/j.synthmet.2023.117369Suche in Google Scholar

[87] Okafor OB, Popoola API, Popoola OM, Adeosun SO. Review on the recent development on polyaniline and transition metal oxides composite electrode for supercapacitor application. Mater. 2025;6(July 2024):100389. 10.1016/j.nxmate.2024.100389.Suche in Google Scholar

[88] Li J, Xu GX. Research progress on applications of hDPSCs in cornea reconstruction. Int Eye Sci. 2017;17:1655–8.Suche in Google Scholar

[89] Pal B, Yang S, Ramesh S, Thangadurai V, Jose R. Electrolyte selection for supercapacitive devices: A critical review. Nanoscale Adv. 2019;1(10):3807–35.10.1039/C9NA00374FSuche in Google Scholar PubMed PubMed Central

[90] Xia L, Yu L, Hu D, Chen GZ. Electrolytes for electrochemical energy storage. Mater Chem Front. 2017;1(4):584–618.10.1039/C6QM00169FSuche in Google Scholar

[91] Xiong T, Tan TL, Lu L, Siang W, Lee V, Xue J. Harmonizing energy and power density toward 2.7 v asymmetric aqueous supercapacitor. Adv Energy Mater. 2018;1702630:1–10.10.1002/aenm.201702630Suche in Google Scholar

[92] Smiatek J, Heuer A, Winter M. Properties of ion complexes and their impact on charge transport in organic solvent-based electrolyte solutions for lithium batteries: insights from a theoretical perspective. Batteries. 2018;4(4):62.10.3390/batteries4040062Suche in Google Scholar

[93] Kulandaivalu S, Suhaimi N, Sulaiman Y. Unveiling high specific energy supercapacitor from layer-by-layer assembled polypyrrole/graphene oxide|polypyrrole/manganese oxide electrode material. Sci Rep. 2019;9(1):1–10.10.1038/s41598-019-41203-3Suche in Google Scholar PubMed PubMed Central

[94] Lu L, Xu S, An J, Yan S. Electrochemical performance of CNTs/RGO/MnO2 composite material for supercapacitor. Nanomater Nanotechnol. 2016;6:1–7.10.1177/1847980416663687Suche in Google Scholar

[95] Jiang DB, Zhang BY, Zheng TX, Zhang YX, Xu X. One-pot synthesis of η-Fe2O3 nanospheres/diatomite composites for electrochemical capacitor electrodes. Mater Lett. 2018;215:23–6. 10.1016/j.matlet.2017.12.059.Suche in Google Scholar

[96] Goriparti S, Miele E, De Angelis F, Di Fabrizio E, Zaccaria RP, Capiglia. C. Review on recent progress of nanostructured anode materials for Li-ion batteries. J Power Sources. 2014;257:421–43. 10.1016/j.jpowsour.2013.11.103.Suche in Google Scholar

[97] Zhu J, Xu Y, Wang J, Lin J, Sun X, Mao S. Effect of various electrolyte cations on electrochemical performance of polypyrrole/RGO based supercapacitors. R Soc Chem. 2015;17(43):28666–73.10.1039/C5CP04080ASuche in Google Scholar

[98] Shoeb M, Mashkoor F, Jeong H, Khan MN, Jeong C. Investigating the effect of carbon nanotubes decorated SmVO4-MoS2 nanocomposite for energy storage enhancement via VARTM-fabricated Solid-state structural supercapacitors using woven carbon fiber. Small. 2025;21(3):e2408283.10.1002/smll.202408283Suche in Google Scholar PubMed

[99] Bohinc K, Kralj-Iglič V, Iglič A. Thickness of electrical double layer. Effect of ion size. Electrochim Acta. 2001;46(19):3033–40.10.1016/S0013-4686(01)00525-4Suche in Google Scholar

[100] Ratajczak P, Suss ME, Kaasik F, Béguin F. Carbon electrodes for capacitive technologies. Energy Storage Mater. 2019;16:126–45. 10.1016/j.ensm.2018.04.031.Suche in Google Scholar

[101] Yang Z, Tang L, Ye J, Shi D, Liu S, Mingqing C. Hierarchical nanostructured α-Fe2O3/polyaniline anodes for high performance supercapacitors. Electrochim Acta. 2018;269:21–29.10.1016/j.electacta.2018.02.144Suche in Google Scholar

[102] Prasanna BP, Avadhani DN, Raghu MS, Yogesh KK. Synthesis of polyaniline/α-Fe2O3 nanocomposite electrode material for supercapacitor applications; Mater Today Commun. 2017;12:72–8.10.1016/j.mtcomm.2017.07.002Suche in Google Scholar

[103] Hameed SA, Ewais HA, Rabia M. Dumbbell-like shape Fe2O3/poly-2-aminothiophenol nanocomposite for two-symmetric electrode supercapacitor application. J Mater Sci Mater Electron. 2023;34(14):1–8. 10.1007/s10854-023-10586-5.Suche in Google Scholar

[104] Wu C, Pei Z, Lv M, Huang D, Wang Y, Yuan S. Polypyrrole-coated low-crystallinity iron oxide grown on carbon cloth enabling enhanced electrochemical supercapacitor performance. Molecules. 2023;28(1):434.10.3390/molecules28010434Suche in Google Scholar PubMed PubMed Central

[105] Wu C, Xiong W, Li H. A novel synthesis of carbon foam@Fe2O3 via hydrolysis-driven emulsion polymerization for supercapacitor electrodes. Carbon Lett. 2023;33(7):2327–34. 10.1007/s42823-023-00566-4.Suche in Google Scholar

[106] Li L, Xie F, Wu H, Zhu Y, Zhang P, Li Y, et al. N-doped porous carbon-nanofiber-supported Fe3C/Fe2O3 nanoparticles as anode for high-performance supercapacitors. Molecules. 2023;28(15):1–14.10.3390/molecules28155751Suche in Google Scholar PubMed PubMed Central

[107] Tao Y, Xie X, Lv W, Tang DM, Kong D, Huang Z, et al. Towards ultrahigh volumetric capacitance: Graphene derived highly dense but porous carbons for supercapacitors. Sci Rep. 2013;3:1–8.10.1038/srep02975Suche in Google Scholar PubMed PubMed Central

[108] Apelgren P, Amoroso M, Säljö K, Montelius M, Lindahl A, Stridh Orrhult L, et al. 3D Nanoporous graphene films converted from liquid-crystalline holey graphene oxide for thin and high-performance supercapacitors. Mater Today Proc. 2019;27:0–31.Suche in Google Scholar

[109] Yu L, Chen GZ. Redox electrode materials for supercapatteries. J Power Sources. 2016;326:604–12. 10.1016/j.jpowsour.2016.04.095.Suche in Google Scholar

[110] Benchikh I, Ezzat AO, Sabantina L, Benmimoun Y. Investigation of hybrid electrodes of polyaniline and reduced graphene oxide with bio-waste-derived activated carbon for supercapacitor applications. 2024;16(3):421.10.3390/polym16030421Suche in Google Scholar PubMed PubMed Central

[111] Charmaine L, Hussain I, Ma X, Zhang K. Properties, functions, and challenges: current collectors. Mater today Chem. 2022;26:101152.10.1016/j.mtchem.2022.101152Suche in Google Scholar

[112] Torop J, Summer F, Zadin V, Koiranen T, Jänes A, Lust E, et al. Low concentrated carbonaceous suspensions assisted with carboxymethyl cellulose as electrode for electrochemical flow capacitor. Eur Phys J E. 2019;42(1):17–20.10.1140/epje/i2019-11766-2Suche in Google Scholar PubMed

[113] Zhou J, Li J, Liu K, Lan L, Song H, Chen. X. Free-standing cobalt hydroxide nanoplatelet array formed by growth of preferential-orientation on graphene nanosheets as anode material for lithium-ion batteries. J Mater Chem A. 2014;2:9416–23.10.1039/C4TA05073HSuche in Google Scholar

[114] Rahayu I, Eddy DR, Novianty AR, Rukiah, Anggreni A, Bahti H, et al. The effect of hydrochloric acid-doped polyaniline to enhance the conductivity. IOP Conf Ser Mater Sci Eng. 2019;509(1):012051.10.1088/1757-899X/509/1/012051Suche in Google Scholar

[115] Vorobiov VK, Bugrov AN, Kasatkin IA. Effect of α‑Fe2O3 nanoparticles on the mechanism of charge storage in polypyrrole ‑ based hydrogel. Polym Bull. 2021;78(5):2389–404. 10.1007/s00289-020-03216-1.Suche in Google Scholar

[116] Zenasni M, Belhadj H, Kiari M, Alelyani M, Alhailiy AB, Benyoucef A, et al. Synthesis, characterization, and enhanced electrochemical behavior of polypyrrole doped ZrO2–ZnO electrode materials for supercapacitor applications. Front Energy Res. August: 2023;11.10.3389/fenrg.2023.1244699Suche in Google Scholar

[117] Das SK, Pradhan L, Jena BK, Basu S. Polymer derived honeycomb-like carbon nanostructures for high capacitive supercapacitor application. Carbon. 2023;213:118214.10.1016/j.carbon.2022.09.004Suche in Google Scholar

Received: 2025-02-10
Revised: 2025-05-14
Accepted: 2025-06-10
Published Online: 2025-07-21

© 2025 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Artikel in diesem Heft

  1. Research Articles
  2. MHD radiative mixed convective flow of a sodium alginate-based hybrid nanofluid over a convectively heated extending sheet with Joule heating
  3. Experimental study of mortar incorporating nano-magnetite on engineering performance and radiation shielding
  4. Multicriteria-based optimization and multi-variable non-linear regression analysis of concrete containing blends of nano date palm ash and eggshell powder as cementitious materials
  5. A promising Ag2S/poly-2-amino-1-mercaptobenzene open-top spherical core–shell nanocomposite for optoelectronic devices: A one-pot technique
  6. Biogenic synthesized selenium nanoparticles combined chitosan nanoparticles controlled lung cancer growth via ROS generation and mitochondrial damage pathway
  7. Fabrication of PDMS nano-mold by deposition casting method
  8. Stimulus-responsive gradient hydrogel micro-actuators fabricated by two-photon polymerization-based 4D printing
  9. Physical aspects of radiative Carreau nanofluid flow with motile microorganisms movement under yield stress via oblique penetrable wedge
  10. Effect of polar functional groups on the hydrophobicity of carbon nanotubes-bacterial cellulose nanocomposite
  11. Review in green synthesis mechanisms, application, and future prospects for Garcinia mangostana L. (mangosteen)-derived nanoparticles
  12. Entropy generation and heat transfer in nonlinear Buoyancy–driven Darcy–Forchheimer hybrid nanofluids with activation energy
  13. Green synthesis of silver nanoparticles using Ginkgo biloba seed extract: Evaluation of antioxidant, anticancer, antifungal, and antibacterial activities
  14. A numerical analysis of heat and mass transfer in water-based hybrid nanofluid flow containing copper and alumina nanoparticles over an extending sheet
  15. Investigating the behaviour of electro-magneto-hydrodynamic Carreau nanofluid flow with slip effects over a stretching cylinder
  16. Electrospun thermoplastic polyurethane/nano-Ag-coated clear aligners for the inhibition of Streptococcus mutans and oral biofilm
  17. Investigation of the optoelectronic properties of a novel polypyrrole-multi-well carbon nanotubes/titanium oxide/aluminum oxide/p-silicon heterojunction
  18. Novel photothermal magnetic Janus membranes suitable for solar water desalination
  19. Green synthesis of silver nanoparticles using Ageratum conyzoides for activated carbon compositing to prepare antimicrobial cotton fabric
  20. Activation energy and Coriolis force impact on three-dimensional dusty nanofluid flow containing gyrotactic microorganisms: Machine learning and numerical approach
  21. Machine learning analysis of thermo-bioconvection in a micropolar hybrid nanofluid-filled square cavity with oxytactic microorganisms
  22. Research and improvement of mechanical properties of cement nanocomposites for well cementing
  23. Thermal and stability analysis of silver–water nanofluid flow over unsteady stretching sheet under the influence of heat generation/absorption at the boundary
  24. Cobalt iron oxide-infused silicone nanocomposites: Magnetoactive materials for remote actuation and sensing
  25. Magnesium-reinforced PMMA composite scaffolds: Synthesis, characterization, and 3D printing via stereolithography
  26. Bayesian inference-based physics-informed neural network for performance study of hybrid nanofluids
  27. Numerical simulation of non-Newtonian hybrid nanofluid flow subject to a heterogeneous/homogeneous chemical reaction over a Riga surface
  28. Enhancing the superhydrophobicity, UV-resistance, and antifungal properties of natural wood surfaces via in situ formation of ZnO, TiO2, and SiO2 particles
  29. Synthesis and electrochemical characterization of iron oxide/poly(2-methylaniline) nanohybrids for supercapacitor application
  30. Impacts of double stratification on thermally radiative third-grade nanofluid flow on elongating cylinder with homogeneous/heterogeneous reactions by implementing machine learning approach
  31. Synthesis of Cu4O3 nanoparticles using pumpkin seed extract: Optimization, antimicrobial, and cytotoxicity studies
  32. Cationic charge influence on the magnetic response of the Fe3O4–[Me2+ 1−y Me3+ y (OH2)] y+(Co3 2−) y/2·mH2O hydrotalcite system
  33. Pressure sensing intelligent martial arts short soldier combat protection system based on conjugated polymer nanocomposite materials
  34. Magnetohydrodynamics heat transfer rate under inclined buoyancy force for nano and dusty fluids: Response surface optimization for the thermal transport
  35. Fly ash and nano-graphene enhanced stabilization of engine oil-contaminated soils
  36. Enhancing natural fiber-reinforced biopolymer composites with graphene nanoplatelets: Mechanical, morphological, and thermal properties
  37. Performance evaluation of dual-scale strengthened co-bonded single-lap joints using carbon nanotubes and Z-pins with ANN
  38. Computational works of blood flow with dust particles and partially ionized containing tiny particles on a moving wedge: Applications of nanotechnology
  39. Hybridization of biocomposites with oil palm cellulose nanofibrils/graphene nanoplatelets reinforcement in green epoxy: A study of physical, thermal, mechanical, and morphological properties
  40. Design and preparation of micro-nano dual-scale particle-reinforced Cu–Al–V alloy: Research on the aluminothermic reduction process
  41. Spectral quasi-linearization and response optimization on magnetohydrodynamic flow via stenosed artery with hybrid and ternary solid nanoparticles: Support vector machine learning
  42. Ferrite/curcumin hybrid nanocomposite formulation: Physicochemical characterization, anticancer activity, and apoptotic and cell cycle analyses in skin cancer cells
  43. Enhanced therapeutic efficacy of Tamoxifen against breast cancer using extra virgin olive oil-based nanoemulsion delivery system
  44. A titanium oxide- and silver-based hybrid nanofluid flow between two Riga walls that converge and diverge through a machine-learning approach
  45. Enhancing convective heat transfer mechanisms through the rheological analysis of Casson nanofluid flow towards a stagnation point over an electro-magnetized surface
  46. Intrinsic self-sensing cementitious composites with hybrid nanofillers exhibiting excellent piezoresistivity
  47. Research on mechanical properties and sulfate erosion resistance of nano-reinforced coal gangue based geopolymer concrete
  48. Impact of surface and configurational features of chemically synthesized chains of Ni nanostars on the magnetization reversal process
  49. Porous sponge-like AsOI/poly(2-aminobenzene-1-thiol) nanocomposite photocathode for hydrogen production from artificial and natural seawater
  50. Multifaceted insights into WO3 nanoparticle-coupled antibiotics to modulate resistance in enteric pathogens of Houbara bustard birds
  51. Synthesis of sericin-coated silver nanoparticles and their applications for the anti-bacterial finishing of cotton fabric
  52. Enhancing chloride resistance of freeze–thaw affected concrete through innovative nanomaterial–polymer hybrid cementitious coating
  53. Development and performance evaluation of green aluminium metal matrix composites reinforced with graphene nanopowder and marble dust
  54. Morphological, physical, thermal, and mechanical properties of carbon nanotubes reinforced arrowroot starch composites
  55. Influence of the graphene oxide nanosheet on tensile behavior and failure characteristics of the cement composites after high-temperature treatment
  56. Central composite design modeling in optimizing heat transfer rate in the dissipative and reactive dynamics of viscoplastic nanomaterials deploying Joule and heat generation aspects
  57. Double diffusion of nano-enhanced phase change materials in connected porous channels: A hybrid ISPH-XGBoost approach
  58. Synergistic impacts of Thompson–Troian slip, Stefan blowing, and nonuniform heat generation on Casson nanofluid dynamics through a porous medium
  59. Optimization of abrasive water jet machining parameters for basalt fiber/SiO2 nanofiller reinforced composites
  60. Enhancing aesthetic durability of Zisha teapots via TiO2 nanoparticle surface modification: A study on self-cleaning, antimicrobial, and mechanical properties
  61. Nanocellulose solution based on iron(iii) sodium tartrate complexes
  62. Combating multidrug-resistant infections: Gold nanoparticles–chitosan–papain-integrated dual-action nanoplatform for enhanced antibacterial activity
  63. Novel royal jelly-mediated green synthesis of selenium nanoparticles and their multifunctional biological activities
  64. Direct bandgap transition for emission in GeSn nanowires
  65. Synthesis of ZnO nanoparticles with different morphologies using a microwave-based method and their antimicrobial activity
  66. Numerical investigation of convective heat and mass transfer in a trapezoidal cavity filled with ternary hybrid nanofluid and a central obstacle
  67. Halloysite nanotube enhanced polyurethane nanocomposites for advanced electroinsulating applications
  68. Low molar mass ionic liquid’s modified carbon nanotubes and its role in PVDF crystalline stress generation
  69. Green synthesis of polydopamine-functionalized silver nanoparticles conjugated with Ceftazidime: in silico and experimental approach for combating antibiotic-resistant bacteria and reducing toxicity
  70. Evaluating the influence of graphene nano powder inclusion on mechanical, vibrational and water absorption behaviour of ramie/abaca hybrid composites
  71. Dynamic-behavior of Casson-type hybrid nanofluids due to a stretching sheet under the coupled impacts of boundary slip and reaction-diffusion processes
  72. Influence of polyvinyl alcohol on the physicochemical and self-sensing properties of nano carbon black reinforced cement mortar
  73. Advanced machine learning approaches for predicting compressive and flexural strength of carbon nanotube–reinforced cement composites: a comparative study and model interpretability analysis
  74. Review Articles
  75. A comprehensive review on hybrid plasmonic waveguides: Structures, applications, challenges, and future perspectives
  76. Nanoparticles in low-temperature preservation of biological systems of animal origin
  77. Fluorescent sulfur quantum dots for environmental monitoring
  78. Nanoscience systematic review methodology standardization
  79. Nanotechnology revolutionizing osteosarcoma treatment: Advances in targeted kinase inhibitors
  80. AFM: An important enabling technology for 2D materials and devices
  81. Carbon and 2D nanomaterial smart hydrogels for therapeutic applications
  82. Principles, applications and future prospects in photodegradation systems
  83. Do gold nanoparticles consistently benefit crop plants under both non-stressed and abiotic stress conditions?
  84. An updated overview of nanoparticle-induced cardiovascular toxicity
  85. Arginine as a promising amino acid for functionalized nanosystems: Innovations, challenges, and future directions
  86. Advancements in the use of cancer nanovaccines: Comprehensive insights with focus on lung and colon cancer
  87. Membrane-based biomimetic delivery systems for glioblastoma multiforme therapy
  88. The drug delivery systems based on nanoparticles for spinal cord injury repair
  89. Green synthesis, biomedical effects, and future trends of Ag/ZnO bimetallic nanoparticles: An update
  90. Application of magnesium and its compounds in biomaterials for nerve injury repair
  91. Micro/nanomotors in biomedicine: Construction and applications
  92. Hydrothermal synthesis of biomass-derived CQDs: Advances and applications
  93. Research progress in 3D bioprinting of skin: Challenges and opportunities
  94. Review on bio-selenium nanoparticles: Synthesis, protocols, and applications in biomedical processes
  95. Gold nanocrystals and nanorods functionalized with protein and polymeric ligands for environmental, energy storage, and diagnostic applications: A review
  96. An in-depth analysis of rotational and non-rotational piezoelectric energy harvesting beams: A comprehensive review
  97. Advancements in perovskite/CIGS tandem solar cells: Material synergies, device configurations, and economic viability for sustainable energy
  98. Deep learning in-depth analysis of crystal graph convolutional neural networks: A new era in materials discovery and its applications
  99. Review of recent nano TiO2 film coating methods, assessment techniques, and key problems for scaleup
  100. Antioxidant quantum dots for spinal cord injuries: A review on advancing neuroprotection and regeneration in neurological disorders
  101. Rise of polycatecholamine ultrathin films: From synthesis to smart applications
  102. Advancing microencapsulation strategies for bioactive compounds: Enhancing stability, bioavailability, and controlled release in food applications
  103. Advances in the design and manipulation of self-assembling peptide and protein nanostructures for biomedical applications
  104. Photocatalytic pervious concrete systems: from classic photocatalysis to luminescent photocatalysis
  105. Corrigendum
  106. Corrigendum to “Synthesis and characterization of smart stimuli-responsive herbal drug-encapsulated nanoniosome particles for efficient treatment of breast cancer”
  107. Special Issue on Advanced Nanomaterials for Carbon Capture, Environment and Utilization for Energy Sustainability - Part III
  108. Efficiency optimization of quantum dot photovoltaic cell by solar thermophotovoltaic system
  109. Exploring the diverse nanomaterials employed in dental prosthesis and implant techniques: An overview
  110. Electrochemical investigation of bismuth-doped anode materials for low‑temperature solid oxide fuel cells with boosted voltage using a DC-DC voltage converter
  111. Synthesis of HfSe2 and CuHfSe2 crystalline materials using the chemical vapor transport method and their applications in supercapacitor energy storage devices
  112. Special Issue on Green Nanotechnology and Nano-materials for Environment Sustainability
  113. Influence of nano-silica and nano-ferrite particles on mechanical and durability of sustainable concrete: A review
  114. Surfaces and interfaces analysis on different carboxymethylation reaction time of anionic cellulose nanoparticles derived from oil palm biomass
  115. Processing and effective utilization of lignocellulosic biomass: Nanocellulose, nanolignin, and nanoxylan for wastewater treatment
  116. Retraction
  117. Retraction of “Aging assessment of silicone rubber materials under corona discharge accompanied by humidity and UV radiation”
Heruntergeladen am 25.12.2025 von https://www.degruyterbrill.com/document/doi/10.1515/ntrev-2025-0189/html
Button zum nach oben scrollen