Open Access Article
Amit
Sahoo
a,
Achyuta N.
Acharya
*a,
Priyambada
Jena
a,
M.
Moonis
a,
J. P.
Biswal
a,
S.
Swain
b and
M. C.
Tripathy
b
aSchool of Basic Sciences & Humanities (Chemistry), Odisha University of Technology and Research, Bhubaneswar-751029, Odisha, India. E-mail: aacharya@outr.ac.in; amitsahoochemistry@gmail.com
bSchool of Electronic Science, Odisha University of Technology and Research, Bhubaneswar, India
First published on 14th August 2025
A novel electrocatalyst, zirconium ferrite nanoparticles (NPs) (ZrFe2O5 NPs), was synthesized through coprecipitation and calcination processes at 300 °C and 500 °C using iron rust. The ZrFe2O5 NPs were used as catalysts for the hydrogen evolution reaction. Furthermore, these NPs in an alkaline medium exhibited superior properties of a fractional order supercapacitor, based on which a prototype device was fabricated to demonstrate its energy storage applications. The capacitance and phase graphs of the fractional-order supercapacitor exhibit a peak value of 1.5 F s−1−α in the mid-frequency range, followed by a decrease at increased frequencies. Furthermore, these NPs were found to be the most effective agents for removing fluoride ions. In a modified one-pot synthetic approach, the hydrogel (Hy) was synthesized using PVA and glycerine in an aqueous medium at 100 °C. The composite Hy membranes were prepared by mixing Hy with these NPs, which were applied to remove fluoride ions efficiently within the pH range of 1–10 from the aqueous medium through adsorption, yielding defluorinated water. The adsorption process of fluoride ions was correlated with the Freundlich and Langmuir isotherms, with pseudo-second-order kinetics as the preferred approach. The zirconium ferrite NPs and the composite Hy membranes were characterized through various characterization techniques, i.e., XRD, XPS, FTIR, solid UV-visible spectroscopy, STA, FESEM, HRTEM, and BET surface area analysis.
Human survival relies on water safety and purity, with concerns about fluoride concentration in drinking water.6 Excessive fluoride in industrialized and developing nations causes dental fluorosis. Artificial fluoridation sources, sourced from phosphate industry waste, are not of pharmaceutical quality.7,8 Over time, various approaches have been investigated to remove excess fluoride from water, including adsorbents, filtration methods such as reverse osmosis and nanofiltration, and electrochemical methods. The adsorption method is a practical and economical technique for removing heavy metals and fluoride, especially with polymers like chitosan, PVA, and their composites.6,9–14 Nanocellulose-based materials with surface alterations exhibit higher adsorption capabilities for heavy metals, fluoride, and dyes,15,16 while advanced modifications such as crosslinking and grafting have improved adsorption capacity and durability.6,9–14 Bi- and multi-metal ferrites and nanocomposites have demonstrated favorable fluoride adsorption from aqueous solutions.10,17
Hydrogen, a clean and sustainable energy source, can be produced through water electrolysis.18 Water is oxygenated to O2 at the anode, releasing protons and electrons that are used at the cathode to produce H2.19 However, the efficiency of the anodic oxygen evolution reaction (OER) is limited by its slow kinetics, which produces a significant overpotential. Non-noble metal OER electrocatalysts, such as Ni-based (oxy)hydroxides, make alkaline water electrolysis (AWE) an inexpensive option; however, it still has an overpotential of over 200 mV.20 Hydrogen gas is a sustainable energy source that can substitute for fossil fuels in applications such as fuel cells, energy storage, chemical manufacturing, and petroleum refining. The primary method of hydrogen production is natural gas cracking, which is not environmentally sustainable.21–24 Renewable energy sources such as wind, tidal, or solar can provide electrochemical water splitting (EWS) for hydrogen gas production. Developing modern agents as catalysts for increased EWS reactions could mitigate the energy challenges.25
Bi-metal oxides, particularly ferrites,26 are used as electrocatalysts for water splitting, as they are affordable, environmentally friendly, and highly effective.8–11,17,18,27,28 Debnath et al.29 investigated cobalt ferrite (CoFe2O4) nanoparticles with controlled oxygen vacancies, synthesized via coprecipitation and inert calcination, which display bifunctional electrocatalytic performance for the hydrogen evolution reaction (HER) and oxygen evolution reaction (OER), necessitating minimal overpotentials and showcasing exceptional durability in water electrolysis. Fatima et al.30 reported the development of a composite electrode comprising cobalt ferrite nanoparticles attached to carbon nanotube fibers, which enabled enhanced electrocatalytic efficacy for water splitting in alkaline environments, with notable improvements in conductivity and stability. Li et al.31 studied the synthesis of MnFe2O4 nanofibers using electrospinning, resulting in a three-dimensional porous structure with enhanced electrocatalytic performance for the oxygen evolution reaction and hydrogen peroxide reduction.
Hybrid supercapacitors (HSCs)32,33 prevail in the energy storage sector, although generation-I electrochemical energy conversion and storage systems (EECSs) represent the most environmentally sustainable technology. Strategic methodologies and electrode compositions drive the advancement of generation-II electrodes and hybrid energy devices.34 Freeborn et al.35 conducted a comprehensive review of existing fractional-order models employed in hybrid and renewable energy systems, encompassing supercapacitors, batteries, and fuel cells. These models optimally fit experimentally observed impedance values and transient responses, utilizing one or more constant phase elements (CPEs), also known as fractional-order capacitors.
A ZrCoFe2O4/NrGO nanocomposite, designed as an electrode material for supercapacitors, was investigated by Kenari et al.36 and exhibited a specific capacitance of 307 F g−1 at 1 A g−1. After 5000 cycles, its stability remained at approximately 97%, indicating strong electrochemical stability. Allagui et al.37 studied the charging voltage profile that affects the discharge voltage pattern of a supercapacitor, which displays fractional-order behavior. Due to its history-dependent dynamic behavior, the charging waveform pattern can yield unique information. The information storage mechanism is explained using an analytical model based on fractional calculus, which may result in new information retrieval and storage techniques.
It is the first of its kind that a waste material, i.e., iron rust, in which iron exists mainly Fe2O3,38 is used as the source of Fe3+ for the synthesis of zirconium ferrite nanoparticles (NPs). The hydrogel (Hy) is synthesized using a modified method, which is a novel, cost-effective, and one-pot synthesis method that utilizes polyvinyl alcohol (PVA) and glycerine. The composite of zirconium ferrite NPs with Hy is synthesized, and its membrane is used for the efficient adsorption of fluoride ions. These NPs exhibit inimitable properties, such as energy storage applications, producing defluorinating water with efficient removal of fluoride ions from aqueous solutions, and promoting sustainable energy production by producing green hydrogen gas through EWS. These NPs can also be utilized as fractional-order supercapacitors.
000 rpm for 10 minutes, and dried in an oven at 90 °C for 1 hour (Scheme 1).
However, the composite of zirconium ferrite NPs with Hy was synthesized in one step wherein 1 gram of M2/M1 NPs was added to the PVA–glycerine mixture (as described above) with stirring for 3–5 min to form a uniform mixture. Then, this was heated for one hour at 100 °C to synthesize Hy(M2)/Hy(M1) [Hy(M1): the composite of M1 zirconium ferrite NPs with Hy; and Hy(M2): the composite of M2 zirconium ferrite NPs with Hy].44 In both the composites, Hy(M2) or Hy(M1), the Hy: zirconium ferrite NP ratio was 1
:
1 (w/w in grams). The membranes of Hy and both composites were prepared using a Petri dish (Scheme 2).
:
2 g to fabricate the electrode-active material, where NMP (N-methyl-2-pyrrolidone) was used as a binder solvent. The uniform mixture of PVDF, NMP, and M2/M1 NP was coated on the surface of the graphite rod, in which each graphite electrode's tip was covered with a specific area of 0.48 cm2 (1.2 cm in length and 0.4 cm in width). The coated electrodes were dried in a hot oven at 90 °C to promote adhesion and remove any remaining solvent.
The electrodes were gently inserted through two holes drilled into the glass sample tube's top cover to include them in the configuration. Then, 15 mL of 3 M NaOH solution was added to the completed cell to submerge the active material and guarantee complete electrolyte immersion. The electrochemical cell was then used to test the charge storage and energy retention properties of a fractional-order supercapacitor45 (Scheme 3).
The elemental detection and oxidation states of Zr, and Fe in the M2 and M1 NP samples were determined using X-ray Photoelectron Spectroscopy (XPS) with an X-ray photoelectron spectrometer (Thermo Fisher Scientific Instruments, UK, Model-K ALPHA). A source–analyzer setup at a 45° angle was used for these investigations.
Fourier Transform Infrared Spectra (FTIR) were recorded at approximately 25 °C between 400 and 4000 cm−1 using a PerkinElmer Spectrum Two instrument with potassium bromide as the reference material. The bandgap energy of the M2 and M1 samples was determined by recording the UV-visible (solid) spectra using a UV-Visible spectrometer (Jasco, V-770). The band gap energy of the M2 and M1 samples was determined using the Tauc plot method.
The thermal stability of the M2 and M1 samples, as assessed using a Simultaneous Thermal Analyzer (STA) (TGA: thermogravimetric analysis and DSC: Differential Scanning Calorimetry), was evaluated using a PerkinElmer STA 6000 instrument in the temperature range of 30–900 °C at a heating rate of 10 °C min−1. The thermal behavior of the M2 and M1 samples was characterized by analyzing the time versus weight loss percentage and time versus heat flow graphs.
A JEOL FESEM (JSM-7610F) instrument was used to examine the surface morphology of the M2, M1 NPs, Hy, Hy(M2), and Hy(M1) samples utilizing the Field Emission Scanning Electronic Microscope (FESEM) technique. This technique involved placing the samples on carbon tape and then coating them with gold for around 60 s.
The internal morphology of the M2 and M1 NPs was examined using High-Resolution Transmission Electron Microscopy (HRTEM) using a JEOL HRTEM (JEM-2100 Plus) instrument. In this case, a 300-mesh carbon-coated copper grid (Merck, India) served as the standard/reference.
An Autosorb iQ Station 1 was used in the Brunauer–Emmett–Teller (BET) surface area analysis, which uses gas adsorption to determine an analyte's specific surface area. Using this technique, a solid sample is either suspended in a known volume of gas or passed over an inert gas, such as nitrogen.
An Epsilon (E21170 model) three-electrode system workstation was used for the electrochemical analysis of the M2 and M1 NP samples in the presence of a NaOH solution as the electrolyte. For electrochemical analysis, the electrodes are made using a graphite rod, PVDF, NMP, and NP material M2 and M1. The “Fabrication of two electrode-based devices for electronic application” section discusses the electrode fabrication process. The Ag/AgCl electrode (Basi, MF-2052) served as the reference electrode, and the platinum electrode (Basi, MW-1032) acted as the counter electrode for the analysis.46
The concentration of F− ions (ppm) in the solution at room temperature was measured using a Thermo Scientific Orion Versastar Pro advanced electrochemistry meter with an Orion 9609BNWP ionplus sure-flow fluoride electrode.47
| Sample name | No. | 2 Theta [°] | FWHM [°] | d-Spacing [Å] | Miller indices (h k l) (cubic-Zr4O8) | Miller indices (h k l) (cubic-O32Fe21.34) | Intensity (%) | Crystalline grain size (nm) |
|---|---|---|---|---|---|---|---|---|
| M2 | 1 | 31.69738 | 0.072 | 2.8206 | (1 1 1) | (0 2 2) | 100 | 2.00 |
| 2 | 35.62557 | 0.072 | 2.51808 | (0 2 0) | (1 3 1) | 2.72 | 2.02 | |
| 3 | 45.44542 | 0.096 | 1.99419 | (− − −) | (0 4 0) | 73.79 | 1.56 | |
| 4 | 56.45956 | 0.096 | 1.62851 | (− − −) | (1 5 1) | 18.14 | 1.63 | |
| 5 | 62.42435 | 0.096 | 1.48646 | (2 2 2) | (0 4 4) | 0.63 | 1.68 | |
| M1 | 1 | 31.70494 | 0.096 | 2.81994 | (1 1 1) | (0 2 2) | 100 | 1.50 |
| 2 | 35.43682 | 0.096 | 2.53105 | (0 2 0) | (1 3 1) | 2.65 | 1.51 | |
| 3 | 45.44399 | 0.096 | 1.99425 | (− − −) | (0 4 0) | 54.12 | 1.56 | |
| 4 | 56.46284 | 0.096 | 1.62842 | (− − −) | (1 5 1) | 11.87 | 1.63 | |
| 5 | 62.41375 | 0.144 | 1.48669 | (2 2 2) | (0 4 4) | 0.31 | 1.12 |
| Sample name | Crystal system | Lattice constant | Standard values (Å) | Calculated value (Å) |
|---|---|---|---|---|
| M2 | Cubic-Zr4O8 | a = b = c | 5.0900 | 5.0361 |
| Cubic-O32Fe21.34 | a = b = c | 8.4050 | 8.4619 | |
| M1 | Cubic-Zr4O8 | a = b = c | 5.0900 | 5.0621 |
| Cubic-O32Fe21.34 | a = b = c | 8.4050 | 8.4615 |
The XRD patterns of the Hy, Hy(M1), and Hy(M2) membranes in the 2θ range of 5–90° are presented in Fig. 2. Due to the copolymer Hy networks, the pure Hy XRD pattern exhibits a prominent broad peak at ∼20°, indicating the formation of three humps in the amorphous region. The M1 and M2 NP XRD peaks were observed with the Hy peak in the case of their composites, Hy(M1) and Hy(M2).
The XRD analysis confirmed the zirconium ferrite compositions for M2 and M1 NPs. In the M2 NP XRD pattern, there was another small peak (very low intensity) obtained at around 30° for the ZrO2 detaching from the zirconium ferrite NP crystal lattice due to the high-temperature calcination. Kaur et al.49 studied zirconium ferrite nanoparticles synthesized by the hydrothermal synthesis method and calcined at temperatures between 300 °C and 700 °C. The crystal phases typically range from 1.1 to 2.0 nm, and this range increases with an increase in temperature (Table 1). Strong peaks, corresponding to spinel ferrite phases, hematite α-Fe2O3, and tetragonal ZrO2 phases, were observed at higher temperatures after the initially amorphous materials were calcined. The presence of M2 and M1 NP XRD peaks with the Hy peak in the case of Hy(M2) and Hy(M1) confirmed the binding of the NPs in the Hy sample.50
There are two broad peaks in the 3000–3600 cm−1 range due to the O–H stretching vibration of the hydroxyl group of glycerine and PVA. There is a diminished peak at around 2942 cm−1 for the aliphatic C–H stretching vibration of the Hy membrane due to the crosslinking of PVA and glycerine.59–62 A carbon–hydrogen bond located on a hydrocarbon chain in the polymer network of Hy is known as an “aliphatic CH bond”.44 C
C bond stretching vibrations of the Hy membrane are responsible for the initial vibration peak at ∼1635 cm−1.44 The ether group (–C–O–C–) stretching vibration peaks in the 1000–1200 cm−1 range result from the formation of ether group and are characteristic of crosslinking between PVA and glycerine in the Hy membrane.59–62 The shift in positions and changes in intensities of the peaks of Hy in the case of Hy(M2) and Hy(M1) may correspond to the binding of the NPs in the Hy membrane.
The strong thermal stability of M2 and M1 NP samples in the 30–750 °C temperature range was confirmed by the STA study. The peak at around 750–900 °C was due to an exothermic reaction corresponding to the decomposition of M1 and M2 NPs into a new crystal material.28 The exothermic reaction of M1 NPs at 750–900 °C has also been confirmed from the DSC curve (Fig. 7c).
![]() | ||
| Fig. 8 FESEM images: M2 NPs (a–d), M1 NPs (e–j), Hy (k, l), Hy(M2) (m–p), and Hy(M1) (q, r) samples. | ||
The 100% purity of the M2 and M1 NPs was found (SI in Table 2) from the EDS (energy-dispersive X-ray spectroscopy) analysis, as shown in SI Figures S1a and S1b, respectively. The carbon (C) and gold (Au) signals in the EDS patterns are obtained due to the carbon tape and gold coating, which provide better resolution for the M2 and M1 NP samples.
Due to its hydrophilicity, the FESEM image of Hy exhibits a net-like structure with pores (Fig. 8k and l). The surface shape of Hy changes dramatically when M1 NPs (Fig. 8m–p) are loaded, and the presence of nanoparticles helps to restore the internal network structure. However, no surface changes are obtained in the case of Hy (M2) (Fig. 8q and r).
The flower-like surface morphology of both NPs is promising for excellent electrochemical, electronic, and adsorption applications.68 The FESEM and HRTEM analysis data indicated that the sizes of star- and needle-shapes of M1 NPs are larger in comparison to those of M2 NPs. This may be due to the high calcination temperatures (500 °C) in the case of M2 NPs, where the larger, star- and needle-shaped NPs are broken down into smaller sizes.
![]() | ||
| Fig. 11 The nitrogen adsorption and desorption isotherms with pore diameter distribution plots of the (a) M2 NPs and (b) M1 NPs samples. | ||
The peak current density significantly decreased when the concentration of NaOH was reduced from 3 M to 1 M. This supports a lower rate of charge transfer, which is similar to that of other metal oxide-based electrodes in alkaline media.69,70 The variation of anodic and cathodic peak potentials suggests changes in reaction kinetics with higher internal resistance.71–77 This was comparable to how larger overpotentials in the electrochemical systems resulted from decreased OH− availability. The broader peaks of the CV curve and smaller enclosed areas demonstrated decreased capacitance.69,70
In the case of M1 and M2 NPs in both 1 M and 3 M NaOH solutions, the Csp value (Fig. 12e) decreases with an increase in scan rate due to faster reaction times, but the electrolyte does not have enough time to reach all active electrode sites. Lower scan rates are generally preferred for measuring maximum Csp, as they allow more time for ion diffusion into the electrode.71–77 Compared to a 3 M NaOH solution, the 1 M solution exhibits a decreased Csp value for M2 NPs at all scan rates (Fig. 12e). A contrasting behavior is observed for the Csp values of M1 NPs in 3 M and 1 M NaOH solutions, i.e., with an increase in electrolyte concentration, the Csp values decrease for 3 M solutions within a 50–500 mV s−1 range scan rate. The Csp values are nearly equal for both electrolyte concentrations for M1 NPs at a 600 mV s−1 scan rate (Fig. 12e).
The charge storage mechanism in M1 nanoparticles may be influenced by ion diffusion restrictions, surface accessibility, and electrochemical kinetics rather than exhibiting a straightforward double-layer or pseudocapacitive behavior due to concentration effects. While the capacitive response may become more surface-controlled at higher scan rates, producing similar Csp values for both concentrations, the decreased capacitance in 3 M NaOH at lower scan rates may suggest restricted ion penetration due to increased ionic strength.
The CD curve analysis reveals that for samples M1 and M2 NPs in 3 M and 1 M NaOH solutions (Fig. 12f–m), the material retains some of its initial charge, known as the memory effect, even after a discharge time. The source exhibits a nonlinear triangular shape, similar to that of a supercapacitor with pseudo-capacitive behavior. The curve does not exhibit battery-like behavior, with a flat voltage region and a plateau-like shape.71–77 A small IR drop (/ohmic drop) (where I represents the current and R denotes the resistance) indicates good conductivity (Fig. 12f, g, j and k), while a gradual voltage drop suggests pseudo-capacitive behavior (Fig. 12h, i, l and m).78
Linear Sweep Voltammetry (LSV) provides key insights into the electrochemical behavior of the M1 and M2 samples, particularly for the oxygen evolution reaction (OER) and hydrogen evolution reaction (HER).18,79 The M1 sample exhibited exceptional charge storage and electrocatalytic performance, as evidenced by its reduced overpotential in the OER and increased specific capacitance, indicating its potential applications in integrated energy storage and conversion. These results indicate that M1 in 3 M NaOH (Fig. 13a and c) is a more effective electrocatalyst for OER and HER applications, showcasing reduced energy demands and enhanced charge transfer rates. In addition, it revealed that the onset potential for M1 in 3 M NaOH (440 mV) was lower than that in 1 M NaOH (480 mV) (Fig. 13b and d), suggesting an increase in catalytic efficiency attributed to the greater availability of OH− ions. The increased electrolyte concentrations improved OER performance by promoting quicker charge transfer kinetics.18,79 The faradaic zone significantly expanded with the increasing scan rate (50–600 mV s−1), suggesting increased redox contributions. These results are similar to those observed in hybrid electrochemical materials, indicating that zirconium ferrite NPs exhibit capacitive and battery-like charge storage behavior.18,79
In contrast to 1 M NaOH (−1255 mV), the HER onset potential for M1 in 3 M NaOH (−1200 mV) was less negative, suggesting enhanced catalytic activity at higher electrolyte concentrations. Similarly, M2 NPs (Fig. 13e and f) showed more negative onset potentials in both electrolytes compared to M1 NPs, possibly due to less HER activity. These are consistent with increased overpotentials resulting from decreased charge transfer efficiency at lower electrolyte concentrations.18,79
Chronopotentiometry (CPE) measures the current response when a constant voltage is applied to investigate the long-term stability and endurance of electrocatalysts. Through the identification of appropriate potential ranges for stability testing, LSV (Linear Sweep Voltammetry) aids in CPE. LSV determines the overpotential needed for a given current density (10 mA cm−2 for OER or −10 mA cm−2 for the HER) to ensure that CPE is carried out under realistic operating circumstances. By tracking voltage drift, CPE data shed light on catalyst stability; if the potential remains constant over time, the material is robust; if it rises, degradation occurs. CPE for M1 and M2 in 1 M versus 3 M NaOH demonstrates how electrolyte concentration affects stability; the more negligible drift in 3 M NaOH indicates improved ionic conductivity and long-term performance. The CPE data show steady potential over a long period and strong catalyst durability (Fig. 13i–l).
The CV and CD analysis confirmed that the supercapacitor exhibited pseudo-capacitive behavior in the M2 NP samples in NaOH solutions (both 1 M and 3 M). The M1 NP samples show pseudo-capacitive behavior only in 1 M NaOH solution and a combination of capacitive and pseudo-capacitive behavior in 3 M NaOH solution, functioning as fractional order supercapacitors in electronic application. Long-term electrochemical performance is enhanced in concentrated alkaline media due to higher ionic conductivity. The electrode's ability to sustain a constant reaction rate confirms the continuous production of H2 and O2 during the HER and the OER.18,79
The proposed capacitor model provides the magnitude and phase responses, as shown in Fig. 14a and b. The phase characteristics of M1 NPs show a constant phase (CP) behavior of −15° over a functional frequency range of 103 Hz to 105 Hz. On the other hand, M2 shows a more resistive type behavior, with phase values ranging from 0° to −4°. Unlike M2, the magnitude of M1 shows some unreal characteristics; therefore, capacitance was determined across varying frequencies. The model fabricated using M2 does not qualify as a capacitor due to its low phase shift and nearly zero magnitude. It was observed, as shown in Fig. 14c, that the capacitance value decreases as the frequency increases and remains within the range of 1–1.5 F within the CP functional frequency range. From the phase characteristics, it is confirmed to behave as a fractional capacitor; therefore, the variations of fractional order (α)80 concerning frequency have been reported in Fig. 14d.
The fractional-order supercapacitor reveals its unique frequency-dependent characteristics, highlighting its behavior as a non-ideal capacitor with fractional-order dynamics. The capacitance plot indicates a peak value of approximately 1.5 F s−1−α in mid-frequency ranges (105 to 106 Hz), followed by a decline at higher frequencies, demonstrating a frequency-dependent capacitance consistent with fractional-order behavior. The phase plot, which decreases from approximately −10° to −40° across the frequency range, confirms its fractional-order nature, with a constant phase angle deviation indicative of a fractional-order parameter (α = 0.16).80 This characteristic phase lag and frequency-dependent capacitance enable the supercapacitor to exhibit memory effects and non-local behaviors (Fig. 14e), making it suitable for applications that require precise control of reactive power and harmonic suppression. These properties enhance its potential for advanced energy storage and power electronics, particularly in dynamic systems and harmonic-rich environments.
The circuit diagram of the fractional-capacitor-based harmonic notch filter is shown in SI Fig. S3a, whereas the frequency domain analysis is reported in SI Fig. S3b, as obtained from the MATLAB simulations. A harmonic notch filter tuned to a center frequency of 50 Hz can effectively serve as a power factor correction (PFC) circuit by suppressing harmonic currents and compensating reactive power. It allows the fundamental frequency to pass with minimal attenuation while blocking higher-order harmonics (e.g., 150 Hz, 250 Hz) and frequencies within the −3 dB bandwidth around 50 Hz. This improves current waveform quality, reduces total harmonic distortion (THD), and aligns the current phase with the voltage, enhancing the power factor. The experimental setup of the harmonic notch filter is depicted in SI Fig. S4a, and its magnitude frequency response is shown in SI Fig. S4b. Here, the response indicates that the experimentally obtained results closely align with the simulated ones. By bypassing harmonic currents and preventing them from flowing back into the source, the filter improves system efficiency, reduces losses, and ensures compliance with power quality standards, making it vital for modern power systems.
The amount of adsorbent used is a key factor in quantifying its adsorption capacity from a fluoride ion solution. When the amount of adsorbent increases, the percentage of fluoride ions removed through adsorption and the number of adsorption sites also increase. The ability of the adsorbent to adsorb the least amount and its cost-effectiveness are studied. The effect of the amount of adsorbent (Fig. 15b) is studied by varying the amounts of M2 (Fig. 15a) and M1 (Fig. 15b) in a 20 mL solution of 10 ppm fluoride ions for up to 10 min and 20 min, respectively. The decrease in the concentration of fluoride ion solution (10 ppm in 20 mL) with varying amounts of adsorbent M2 and M1 is shown in Fig. 15b.
In the present study, the M2 sample exhibited 33% initial adsorption at 0.05 g, which increased to 50% at 0.25 g, and reached saturation above this dosage. Hence, the optimized dose for the M2 adsorbent is 0.25 g per 20 mL of a 10 ppm fluoride ion solution for 10 min at room temperature. Similarly, the M1 sample achieved 99% initial adsorption at 0.05 g which enhanced to 100% at 0.2 g. After 20 min at room temperature, the optimal dose of M1 was 0.2 g per 20 mL of a 10 ppm fluoride ions solution. From an economic perspective, this increase is significant; therefore, using M1 adsorbents at 0.2 g per 20 mL of a 10 ppm fluoride ion solution would be the best course of action, i.e., achieving 100% removal of the fluoride ions.
The number of accessible sites on an adsorbent surface directly affects the initial adsorption of fluoride ions. Saturation of adsorption sites causes the percentage of adsorption or removal to decrease as the concentration of initial fluoride ions increases. However, the high driving force for mass transfer extends the capacity of the adsorbents. Fig. 15c illustrates the influence of fluoride ion concentrations in the 5–100 ppm range on its absorption; however, at 100 ppm, it reaches saturation at 48% for M1 and 4.2% for M2 adsorbents. The impact of pH (2 ≤ pH ≤ 11) on the adsorption of 10 ppm of fluoride ions by the optimal quantity of M2 and M1 samples was investigated. The results (Fig. 15d) showed that in a pH range of 1–10, fluoride ions adsorption is almost constant at ∼100% for M1 NPs, ∼50% for M2 NPs, and slightly decreases or saturates with raising the pH to 11. This is most likely because the fluoride ions are drawn towards the M2 and M1 samples due to their surface charges within the pH range of 1–10.
The Hy membranes do not adsorb fluoride ions. Fluoride ion adsorption was conducted for 10 cycles to ascertain the heterogeneity of the composite adsorbents, Hy(M2) and Hy(M1) membranes (Fig. 15e), in a 20 mL solution of 10 ppm fluoride ions for 20 min. Furthermore, the percentage of adsorption of the fluoride ions in the case of Hy(M1) is higher than that of Hy(M2) across all cycles, indicating that Hy(M1) is a better reservoir of fluoride ions than Hy(M2). The Hy(M1) adsorbed fluoride ions at ∼100% up to 6 cycles; thereafter, the efficiency decreased to ∼70% at 10 cycles.
An empirical model that considers adsorption onto a heterogeneous surface with several active sites of different energies is the Freundlich isotherm (SI as E2) (shown in Fig. 16b). The equilibrium concentration of the adsorbate in solution is Ce (mg L−1). The adsorbed quantity per unit mass of the adsorbent is denoted by Qe (mg g−1), while the adsorption capacity and intensity are indicated by the Freundlich constants kF and n, respectively. The graph of ln(Qe) versus ln(Ce) is a straight line whose slope and intercept yield kF and 1/n, supporting that higher adsorption occurs at lower concentrations.
![]() | ||
| Fig. 16 (a) Langmuir and (b) Freundlich adsorption isotherm study of adsorption of F− ions using M2 and M1 NPs as adsorbents. | ||
The Langmuir isotherm, on the other hand, assumes that monolayer adsorption occurs on a uniform surface with adsorption sites (SI as E3) (shown in Fig. 16a). The maximum adsorption capacity when a monolayer covers the surface is indicated by Qmax (mg g−1). The Langmuir constant linked to adsorption energy is kL (L g−1). Qmax and kL values are obtained from a plot of Ce/Qeversus Ce. Furthermore, the feasibility of adsorption is determined by the equilibrium parameter RL (SI as E4), where C0 is the starting adsorbate concentration and aL (= kL/Qmax) is a Langmuir constant. The adsorption process is divided into three categories: irreversible adsorption (RL = 0), unfavorable or linear adsorption (RL > 1), and favorable adsorption (RL < 1).
The Freundlich and Langmuir models (Table 3) adequately describe the adsorption behavior; however, the Langmuir model offers a more potent connection based on statistical factors. The M1 NPs have a considerably higher Qmax value than M2 NPs in the Langmuir adsorption isotherm model, indicating that the former is a more effective adsorbent for adsorbing F− ions. The Hy(M1) is an efficient reservoir of F− ions.
| Absorbent | Langmuir | Freundlich | ||||||
|---|---|---|---|---|---|---|---|---|
| a L (L mg−1) | K L (L g−1) | Q max (mg g−1) | r L 2 | R L | K F | 1/n | r F 2 | |
| M1 | 20.03 | 383 | 19.11 | 0.98895 | 0.00989 | 651 | 0.950 | 0.90735 |
| M2 | 12.11 | 219 | 18.12 | 0.95721 | 0.01625 | 651 | 0.786 | 0.99901 |
![]() | ||
| Fig. 17 Linear pseudo-second-order kinetics study of adsorption of F− ions using M2 and M1 NPs as adsorbent. | ||
| Absorbent | k 2 (g mg−1 min−1) | Q e (mg g−1) | R 2 |
|---|---|---|---|
| M1 | 0.05891 | 4.27223 | 0.98612 |
| M2 | 0.22404 | 1.21181 | 0.98562 |
The engineered fractional-order supercapacitor may be utilized in future real-time power factor correction circuit research.
Equations E1 to E5, Fig. S1 to S4, and Tables 1 to 8 are submitted as the SI. See DOI: https://doi.org/10.1039/d5na00578g.
| This journal is © The Royal Society of Chemistry 2025 |