Translate this page into:
Synthesis, structural modification, and dielectric property enhancement of manganese and tin-doped bentonite composites for high-performance energy storage applications
*Corresponding author: E-mail address: Abishil@taibahu.edu.sa (A. Bashal)
-
Received: ,
Accepted: ,
Abstract
This study explores the structural, morphological, optical, and dielectric properties of pure Bentonite (BT), Mn-doped Bentonite (Mn-BT), and Sn-doped Bentonite (Sn-BT) to evaluate their potential in advanced applications. X-ray diffraction (XRD) confirms the presence of Montmorillonite (M), Si, and Quartz (Q) phases, with slight peak shifts in Mn-BT and Sn-BT indicating successful incorporation of metal ions. Crystallite sizes increased from 16.07 nm for BT to 17.58 nm and 19.88 nm for Mn-BT and Sn-BT, respectively, suggesting structural modifications. Scanning electron microscopy (SEM) and Energy Dispersive X-ray (EDX) spectroscopy reveal a porous structure with homogeneously distributed elements, dominated by Si and Al, with minor quantities of Fe, Mg, Ca, and Na. UV-Vis analysis shows a reduction in bandgap energy from 4.14 eV for BT to 4.01 eV and 4.07 eV for Mn-BT and Sn-BT, respectively, indicating enhanced optical properties. Dielectric studies demonstrate significantly improved permittivity, with ε’ values higher in Mn-BT and Sn-BT compared to BT and reduced dielectric losses across a broad frequency range. Electrical conductivity increases with frequency, with hopping conduction dominating beyond a critical frequency. These results highlight the potential of Mn-BT and Sn-BT for applications in high-performance capacitors, energy storage devices, and environmental remediation.
Keywords
Bentonite
Dielectric properties
Energy storage applications
Manganese
Recycle
Tin

1. Introduction
Over the past two decades, there have been continuous advancements in environmental treatment methodologies at both academic and industrial levels. Clean energy alternatives, particularly for electricity storage, must be efficiently stored or converted into other forms of energy to meet the growing demands of consumers. Among electronic components, electrochemical supercapacitors (ECs), also known as ultracapacitors, have gained prominence due to their ability to charge and discharge rapidly while reliably storing energy over extended periods [1]. Research efforts have primarily focused on enhancing the energy storage capacity of these capacitors to meet higher performance requirements. These advancements represent a significant industrial and commercial challenge, necessitating the widespread adoption of these devices [2]. Within the diverse microelectronic devices utilized for energy storage, supercapacitors and batteries represent significant advancements in contemporary scientific research [3]. Current studies aim to enhance energy storage efficiency by developing, optimizing, and manufacturing different suitable materials. Within this context, bentonite (BT) emerges as a noteworthy substance, categorized as a clay mineral belonging to the smectite group. It is composed generally of aluminum hydro silicate and the mineral montmorillonite (M) [4]. The swelling and permeability properties of BTs are closely dependent on the nature of the compensating cation. The term BT was proposed in 1898 to designate a soap-like clay belonging to the “Benton shale” formation, outcropping in the Rock River region (Wyoming, USA), where the first BT deposit was reportedly discovered in 1890, also known as fuller’s earth. This formation is named after Fort Benton, located about 650 km north of Rock River. This material is characterized by an octahedral layer of Al atoms sandwiched between two tetrahedral layers of Si atoms. The isomorphic substitution of Al3+ by Si4+ and Mg2+ by Al3+ in the octahedral layer leads to a net negative charge on the surface of the BT. This charge imbalance is compensated for exchangeable cations (Na+, Ca2+, etc.) on the BT surface. BT is mainly composed of SiO2 (∼70 %), with a smaller fraction of Al2O3 (∼15 %) [5]. Its general formula is Si4(Al(2-x)Rx)(O10, H2O)(Cex, nH2O) or Si4(Al(2-x)Rx)(H2O)n, with, R = Mg, Fe, Mn, Zn, and Ni, and Ce (exchangeable cations) = Ca, Na, and Mg. These materials possess abundant, easily accessible, medium-sized pores, making them cost-effective and environmentally friendly due to their widespread availability [6].
As mentioned earlier, they hold significant potential for energy storage applications, such as in supercapacitors. A crucial factor to consider is the level of electrical conductivity, which indicates the potential material of conducting electric current. This factor is particularly important in the study of BT, as the presence of inorganic elements can influence its conductivity. The introduction of inorganic elements, such as alkaline-earth or transition metals, can lead to substantial changes in BT’s electrical conductivity. Transition metal oxides are widely used in various industrial processes, serving as electrode materials in ECs. In our previous works, we examined the doping effect on BT with various elements. A significant improvement in its electrical properties was observed. Specifically, doping with 5% Sr and 5% Ru enhances the permittivity of BT, indicating an increased ability to store electrical energy [7]. Similarly, a 5% Zn doping also improves permittivity, further highlighting BT’s potential as an efficient dielectric material [8]. Additionally, doping with Fe notably enhances the conductivity of BT, making it more effective in conducting electric currents [9]. However, further studies could explore the effects of doping with other elements on the physical features of BT clay, helping to identify the optimal doping level that enhances the desired physical properties and boosts performance. In this study, BT was chosen as the host material with the goal of developing a new BT-based compound that exhibits enhanced dielectric properties, a higher dielectric constant, and a reduced energy band gap, making it a promising candidate for optoelectronic applications at high temperatures. In this work, wet impregnation methodology was applied to synthesize 5% Manganese/bentonite (Mn-BT) and 5% Tin/bentonite (Sn-BT). The structural, optical, dielectric, and conductivity analyses were done.
2. Materials and Methods
2.1. Materials
BN [CAS Number: 1302-78-9, Molecular weight: 1801 g/mol], SnCl2 [CAS Number: 7772-99-8, Molecular weight: 189.62 g/mol], and MnCl2 [CAS Number: 7773-01-5, Molecular weight: 125.84 g/mol] were obtained from Sigma-Aldrich (USA). Deionized water was used to prepare all aqueous solutions.
2.2. Composites preparation
Mn-BT and Sn-BT composites were prepared through the incipient wet impregnation method, following the procedure described in our previous work [7]. First, BT was dispersed in distilled water under constant stirring at room temperature. A precise amount of MnCl₂ and SnCl₂ salts (5 wt%) were then dissolved in the solution, followed by vigorous stirring for 3 h to ensure homogeneity. The mixture was subsequently dried at 393 K for 12 h. After drying, the obtained solid material was ground into a fine powder using a mortar and pestle. Finally, 0.30 g of the powdered composite was pressed into pellets using a hydraulic die-set for dielectric measurements.
2.3. Characterization techniques
The X-ray diffraction (XRD) patterns were obtained utilizing a Philips X’Pert-Pro diffractometer. The analysis used Cu-Kα radiation with a wavelength of 1.5418 Å, operating at 40 kV and 40 mA. Data were collected over a 2θ range of 10° to 65° with a scan step size of 0.02° and scan speed of 12°/min. The optical investigation was performed through an Ocean Optics (2000 USB) spectrometer covering the wavelength width of 200–1100 nm. On the other hand, a Shimadzu FTIR-8400 spectrophotometer was employed for the Fourier transform infrared (FT-IR) measurements in the wavenumber range 500 – 4000 cm-1. Additionally, we investigated the dielectric and electrical properties in the temperature range 25-120°C and in the frequency range from 1 kHz to 1 MHz. The samples were circular films, roughly 0.5 cm radius and approximately 1 mm thick. In our investigation, copper electrodes were attached to the films. For impedance measurements, copper electrodes were attached to both sides of the films using a thin adhesive layer to ensure good electrical contact. This method was preferred over using conductive paste (such as silver paste), which could potentially seep into the sample and produce unwanted signals in the impedance spectrum.
3. Results and Discussion
3.1. XRD analysis
Figure 1 depicts the XRD patterns for the pure BT, Mn-BT, and Sn-BT at room temperature. One can observe the presence of the M, Si, and Quartz (Q) phases within the three samples. For pure BT, the prominent peaks at 2θ values of approximately 19.78°, 27.99°, 35.35°, 54.16°, and 61.98° correspond to M. Diffraction peaks at 21.93° were attributed to the presence of Si (JCPDS no. 27-0605) [10]. The peak at approximately 26.60° indicates the presence of Q [7]. The additional diffraction peaks in the undoped BT are most likely caused by impurities like cristobalite, feldspar, and illite, consistent with the findings of previous BT investigations [11]. The same diffraction peaks were observed in Mn-BT and Sn-BT samples with a slight shift in 2θ values. This peak shift in 2θ values suggests structural modifications due to the interaction of metal ions with the BT. For Mn-BT, the peaks shift towards higher 2θ values, which suggests a decrease in the interlayer spacing (d-spacing). This contraction may result from the incorporation of Mn2⁺ ions, leading to stronger electrostatic interactions and a reduction in the distance between clay layers. Conversely, in Sn-BT, the peaks shift towards lower 2θ values, indicating an expansion of the interlayer spacing. This expansion could be attributed to the larger ionic radius of Sn2⁺ compared to Mn2⁺, which may disrupt the M structure and increase the basal spacing. Additionally, the shift may be due to the aggregation of clay layers caused by a shortage of water in the interlayer region, which might be a consequence of adding Mn and Sn metals [12,13]. All the diffraction peaks have been summarized in Tables 1–3 for BT, Mn-BT, and Sn-BT, respectively. To quantify these structural modifications, the shifts in 2θ values (Δ2θ) for key diffraction peaks indicate noticeable changes in the crystal structure of BT upon doping with Mn and Sn. For the M peak initially observed at approximately 19.78°, a shift of +0.062° is recorded for Mn-BT, whereas a slight decrease of -0.026° is observed for Sn-BT. Similarly, the Si peak at around 21.94° shifts by +0.123° for Mn-BT but exhibits a more pronounced decrease of -0.465° for Sn-BT. The Q peak near 26.61° experiences a minor shift of +0.056° for Mn-BT and +0.015° for Sn-BT, indicating minimal structural distortion. Conversely, the M peak at approximately 27.99° shifts by -0.282° for Mn-BT and -0.341° for Sn-BT, further suggesting variations in interlayer spacing and potential structural rearrangements induced by the incorporation of metal ions. Further, no distinct peaks corresponding to Mn or Sn were detected in the diffraction patterns of Mn-BT and Sn-BT, suggesting the successful inclusion of Mn and Sn ions into the BT structure. According to Scherrer’s equation, the average crystallite size of the BT, Mn-BT, and Sn-BT materials increased from 16.07 nm, 17.58 nm, and 19.88 nm, respectively. Similar behavior was noted in reference [9] for BT doped with Fe, Zn, Cd, and Cu metals.

- XRD spectra of BT, Mn-BT, and Sn-BT samples at room temperature.
| 2θ (°) | d (Å) | Phase |
|---|---|---|
| 19.788(5) | 4.482(9) | M |
| 21.937(0) | 4.045(8) | Si |
| 26.608(7) | 3.347(2) | Q |
| 27.994(5) | 3.184(6) | M |
| 35.354(5) | 2.536(8) | M |
| 54.163(0) | 1.692(1) | M |
| 61.981(3) | 1.496(0) | M |
| 2θ (°) | d (Å) | Phase |
|---|---|---|
| 19.850(5) | 4.469(0) | M |
| 22.060(4) | 4.026(7) | Si |
| 26.664(1) | 3.340(0) | Q |
| 27.712(1) | 3.216(4) | M |
| 35.336(6) | 2.538(0) | M |
| 54.172(0) | 1.691(9) | M |
| 62.010(3) | 1.495(4) | M |
| 2θ (°) | d (Å) | Phase |
|---|---|---|
| 19.762(4) | 4.488(8) | M |
| 21.471(9) | 4.134(0) | Si |
| 26.623(6) | 3.345(5) | Q |
| 27.653(7) | 3.223(2) | M |
| 35.318(6) | 2.539(3) | M |
| 54.105(0) | 1.693(7) | M |
| 61.862(0) | 1.498(7) | M |
3.2. SEM analysis
Figures 2 and 3 illustrate a detailed characterization of Mn-BT and Sn-BT, respectively, using scanning electron microscopy (SEM) and energy-dispersive X-ray spectroscopy (EDX). For the Mn-BT sample, the surface morphology shows a rough and irregular texture with multiple layers and flakes, indicative of its natural, non-uniform formation, which highlights the varying particle sizes and shapes, with some particles appearing larger and more irregular. The porous texture is evident, which is characteristic of BT and contributes to its adsorption properties. The layering and aggregation observed suggest a stratified structure, typical of clay minerals. However, for the Sn-BT compound, the surface morphology reveals a rough, granular surface with clusters of particles dispersed throughout the field of view. These clusters appear to be irregular in shape and size, indicating a heterogeneous distribution typical of BT materials. The image also shows cracks and voids within the structure, suggesting a porous nature, which is a key characteristic of BT and contributes to its high surface area and adsorption capacity. The detection of all constituents within the samples, such as Si, Al, Mn, N, Mg, Na, Fe, Ca, F, Zn, Cl, and O, indicates that no elements were lost throughout the experiment. Consequently, the composition of doped BT is mainly dominated by large amounts of Si and Al, with tiny quantities of Fe, Mg, Ca, and Na. In addition, all the elements are distributed homogeneously across the surface of the compounds. It is noteworthy that the SEM analysis for pure BT was investigated in our previous work [8].

- SEM analysis of Mn-doped BT. The image displays the morphological and elemental characterization. (a) shows the SEM micrograph of (Mn/BT) revealing the surface morphology and mapping of all presents elements, such as (b) aluminum, (c) calcium, (d) iron, (e) carbon, (f) manganese, (g) magnesium (h) silica, (i) sodium, (j) nitrogen, (k) sulphur, and (l) Oxygen. (m) are elemental full mapping images illustrating the distribution of each elements and their combinations within the nanocomposite structure.

- SEM analysis of Sn-doped BT. The image displays the morphological and elemental characterization. (a) shows the SEM micrograph of (Sn/BT) revealing the surface morphology and mapping of all presents elements, such as (b) aluminum, (c) carbon, (d) calcium, (e) chlorine, (f) fluorine, (g) iron (h) zinc, (i) silica, (j) magnesium, (k) nitrogen, (l) sodium, and (m) oxygen. (n) are elemental full mapping images illustrating the distribution of each elements and their combinations within the nanocomposite structure.
3.3. FT-IR analysis
Figure 4 shows the FT-IR spectra of the prepared samples in the wavenumber range 500 - 4000 cm-1. Distinct absorption bands characteristic of crystalline impurities and those related to the clay phase can be identified. The spectra for all three samples display prominent bands around 3625 cm-1 and 3415 cm-1, indicative of the stretching vibrations of OH groups, particularly those associated with water molecules in the clay structure [14]. A notable band around 1632 cm-1 corresponds to the H-O-H deformation vibrations of water molecules adsorbed between the clay layers [14]. Additionally, the strong absorption band around 1121 cm-1 is attributed to the Si-O stretching vibrations in the tetrahedral layers, while the band around 915 cm-1 is linked to the Al-O-H bending vibrations, characteristic of M [15].

- -FT-IR spectra of BT, Mn-BT and Sn-BT samples at room temperature.
Furthermore, the deformation vibrations of Si-O-Al manifest in M with a peak at 550 cm-1 [16]. It is noteworthy that specific bands appearing at around 1437 cm-1 are assigned to Na-M [17]. However, there are distinct differences in the absorption bands between the samples. In Mn-BT, there are slight shifts and variations in the intensity of the OH stretching and Si-O stretching bands compared to pure BT, indicating interactions or substitutions involving Mn ions. These shifts suggest that Mn incorporation alters the local electronic structure by modifying the bonding environment and electron density distribution around oxygen atoms. Similarly, the Sn-BT shows shifts and intensity changes in the same regions, reflecting the influence of Sn ions on the clay’s structure. These modifications arise due to differences in electronegativity and ionic radii between Mn2⁺/Sn2⁺ and the original ions they replace, leading to changes in dipole moments and local polarization effects. The differences in the clay’s structure highlight the effect of replacing metal ions on the structural and vibrational properties of the BT samples.
3.4. UV-Vis analysis
The absorbance spectra of BT, Mn-BT, and Sn-BT were recorded over the wavelength range of 200 to 1100 nm (Figure 5a). Notably, all three samples exhibit a strong absorbance peak in the UV region, specifically around 200 to 300 nm, which is characteristic of Si–O and the Al–O charge-transfer electronic within the clay minerals [18,19]. The BT sample shows a distinct absorbance peak at approximately 210 nm, indicating the presence of structural features and possibly impurities that absorb strongly in this region. Comparatively, the Mn-BT and Sn-BT samples demonstrate similar absorbance behavior but with slightly higher intensity and slight shifts in the peak positions. This behaviour suggests an increased absorbance due to the incorporation of Mn and Sn ions.

- (a) Absorbance spectra; (b) band gap analysis at room temperature of BT, Mn-BT and Sn-BT samples.
The band gap energies of our materials were determined by applying the Tauc plot method based on the Tauc Eq. (1) [20]:
α represents the optical absorption coefficient, h is Planck’s constant, ʋ denotes the frequency of the light, Eg stands for the band-gap energy, and A is a material-dependent constant. Figure 5(b) displayed the Tauc plot for BT, Mn-BT, and Sn-BT.
The extrapolation of the linear portion of the curve to the energy axis (x-axis) gives the bandgap energy (Eg) for each sample. From the graph, the intercepts of the linear fits for BT, Mn-BT, and Sn-BT are around 4.14 eV, 4.01 eV, and 4.07 eV, respectively. This indicates that the doping of BT with Mn and Sn slightly reduces the bandgap energy compared to pure BT. The decrease in bandgap energy suggests that doping introduces new energy levels within the bandgap, which can enhance the material’s electronic and optical properties [21]. These impurity levels facilitate electronic transitions at lower energies, contributing to the observed bandgap narrowing. Additionally, lattice distortions induced by the dopants may further influence the electronic structure. While charge transfer interactions cannot be entirely ruled out, the gradual nature of the bandgap reduction suggests that impurity levels are the dominant factor. A slight reduction in the bandgap allows the material to absorb light over a broader wavelength range, facilitating easier excitation of electrons from the valence band to the conduction band under lower photon energy. This improvement in optical response is particularly beneficial for applications in optoelectronic devices, photocatalysis, and sensors, where enhanced light interaction is desirable [22]. Our findings on the influence of metal substitution on BT’s band gap energy are consistent with previous research reported by M. Kaur et al. [23], where the incorporation of MgFe2O4 nanoparticles into BT reduced the band gap.
3.5. Dielectric analysis
3.5.1. Dielectric permittivity
The dielectric analysis represents a significant source of valuable information on conduction processes, as it can be used to understand the origin of dielectric losses and the electrical and dipolar relaxation times [24]. Figure 6 illustrates the real part of the permittivity of BT, Mn-BT, and Sn-BT versus frequency at various temperatures. All the samples have the same behaviour. As frequency increases, the decreases consistently across all temperatures. This trend is typical in dielectric materials, where higher frequencies result in reduced permittivity due to the inability of dipolar entities to align with the rapidly oscillating electric field [25]. At lower frequencies, the permittivity values are higher, indicating better alignment of dipoles with the electric field, contributing to higher dielectric constant values [26].

- Frequency evolution of at various temperatures of (a) BT, (b) Mn-BT, (c) Sn-BT, (d) frequency evolution of at room temperature.
Debye relaxation describes the dielectric response of material that pertains to an ideal dielectric subjected to an external alternating electric field, without the mutual interaction of the dipole population. It is generally described by the expression (Eq. 2):
Where is the static permittivity, is the high-frequency permittivity, is the characteristic relaxation time of the material. However, most dielectric materials do not follow the Debye expression and are often described by a more general expression of permittivity known as the Cole-Cole law (Eq. 3) [27]:
Where is the direct current electrical conductivity and is the vacuum permittivity. α represents the distribution of relaxation times (0 < α < 1) and can be interpreted as a dispersion factor around the average relaxation time. From the Eq. (4), the real part of the permittivity may be expressed as:
We observe that Equation II.16 consists of two terms. The first term corresponds to thermal polarization, while the second term is related to electrical conduction [28]. The solid lines on the curves in Figure 6 represent the simulated results for each temperature based on a Cole-Cole model. This indicates a strong correlation between experimental and calculated results. Tables 4–6 show the fit parameters, such as εₛ, ε∞, α, τ, and σdc for BT, Mn-BT, and Sn-BT, respectively. Specifically, the increase in relaxation time (τ) for Mn-BT and Sn-BT suggests a shift in polarization mechanisms, while the variations in σdc indicate changes in charge transport properties. These findings are consistent with previous studies on doped dielectric materials [29-33]. As shown in Figure 6(d), the dielectric constants for Mn-BT and Sn-BT are significantly higher than that of pure BT, indicating that doping with Mn and Sn enhances the dielectric properties of BT.
| T (°C) | (F.m-1) | (F.m-1) | α | τ (s) | (S.m-1) |
|---|---|---|---|---|---|
| 25 | 92.99 | 3.00 | 0.71 | 2.09×10-6 | 3.61×10-6 |
| 30 | 115.26 | 17.09 | 0.75 | 2.27×10-6 | 2.52×10-6 |
| 40 | 115.81 | 17.00 | 0.74 | 1.47×10-6 | 2.84×10-6 |
| 50 | 96.96 | 0.65 | 0.76 | 1.26×10-6 | 2.59×10-6 |
| 60 | 93.19 | 1.81 | 0.74 | 1.04×10-6 | 2.84×10-6 |
| 70 | 86.22 | 8.78 | 0.72 | 1.11×10-6 | 2.59×10-6 |
| 80 | 67.20 | 27.80 | 0.72 | 1.05×10-6 | 1.19×10-6 |
| 90 | 65.27 | 29.73 | 0.70 | 1.31×10-6 | 1.12×10-6 |
| 100 | 63.18 | 31.82 | 0.68 | 1.73×10-6 | 1.17×10-6 |
| 110 | 61.38 | 33.62 | 0.65 | 2.30×10-6 | 1.24×10-6 |
| 120 | 58.03 | 36.97 | 0.69 | 4.63×10-6 | 9.66×10-7 |
| 130 | 57.16 | 37.84 | 0.77 | 6.16×10-5 | 4.07×10-7 |
| T (°C) | (F.m-1) | (F.m-1) | α | τ (s) | (S.m-1) |
|---|---|---|---|---|---|
| 25 | 85.61 | 9.39 | 0.74 | 7.20×10-4 | 3.98×10-7 |
| 30 | 78.42 | 16.58 | 0.65 | 6.35×10-6 | 1.38×10-6 |
| 40 | 78.64 | 16.36 | 0.66 | 1.89×10-6 | 2.60×10-6 |
| 50 | 78.13 | 16.87 | 0.69 | 1.62×10-6 | 2.30×10-6 |
| 60 | 77.03 | 17.97 | 0.73 | 1.56×10-6 | 1.90×10-6 |
| 70 | 75.81 | 19.19 | 0.73 | 1.28×10-6 | 2.36×10-6 |
| 80 | 73.97 | 21.03 | 0.74 | 1.08×10-6 | 3.58×10-6 |
| 90 | 73.95 | 22.05 | 0.77 | 1.43×10-6 | 4.82×10-6 |
| 100 | 88.39 | 7.61 | 0.91 | 3.70×10-6 | 6.18×10-6 |
| 110 | 85.11 | 9.89 | 0.89 | 4.12×10-6 | 5.47×10-6 |
| 120 | 76.26 | 18.74 | 0.79 | 1.49×10-5 | 5.42×10-6 |
| 130 | 76.06 | 18.94 | 0.70 | 6.89×10-6 | 6.46×10-6 |
| T (°C) | (F.m-1) | (F.m-1) | α | τ (s) | (S.m-1) |
|---|---|---|---|---|---|
| 25 | 74.44 | 20.56 | 0.75 | 2.98×10-4 | 1.48×10-6 |
| 30 | 80.87 | 14.13 | 0.75 | 8.77×10-5 | 1.00×10-6 |
| 40 | 79.55 | 15.45 | 0.75 | 5.36×10-5 | 1.16×10-6 |
| 50 | 78.59 | 16.41 | 0.75 | 2.95×10-5 | 1.22×10-6 |
| 60 | 75.65 | 19.35 | 0.71 | 5.22×10-6 | 2.78×10-6 |
| 70 | 73.59 | 21.41 | 0.69 | 2.81×10-6 | 4.79×10-6 |
| 80 | 72.08 | 22.92 | 0.69 | 2.74×10-6 | 5.22×10-6 |
| 90 | 70.55 | 24.45 | 0.71 | 3.70×10-6 | 5.42×10-6 |
| 100 | 68.45 | 26.55 | 0.72 | 3.55×10-6 | 6.93×10-6 |
| 110 | 66.83 | 28.17 | 0.73 | 4.16×10-6 | 7.51×10-6 |
| 120 | 65.26 | 29.74 | 0.74 | 5.39×10-6 | 7.83×10-6 |
| 130 | 65.87 | 29.13 | 0.80 | 7.38×10-5 | 9.30×10-6 |
At 25°C and 4×10⁴ Hz, the dielectric constant of pure BT is 4.46, while Sn-BT exhibits a value of 12.56. Similarly, Mn-BT shows a dielectric constant of 16.30. These results highlight the effectiveness of Mn and Sn doping in enhancing the dielectric response of BT. This remarkable improvement suggests that Mn-BT and Sn-BT composites have great potential for advanced applications in high-performance capacitors and energy storage devices [34,35]. In comparison with other energy storage technologies, supercapacitors offer high power density, rapid charge-discharge cycles, and extended lifespan, making them superior for applications requiring fast energy storage and release. Unlike conventional capacitors, which rely solely on electrostatic charge storage, or batteries, which involve slow faradaic reactions, supercapacitors benefit from both mechanisms, allowing enhanced energy efficiency. The enhanced dielectric permittivity observed in Mn-BT and Sn-BT composites directly correlates with improved charge storage capacity. A higher dielectric constant increases the ability of the material to store an electric charge, thereby enhancing the energy density of the capacitor. This suggests that Mn-BT and Sn-BT composites could contribute to improving the performance of supercapacitors by enabling greater charge accumulation and reducing energy losses. A detailed comparison with other energy storage technologies can be found in recent studies [36-40]. Figure 7 depicts the temperature evolution of at distinctive frequencies for BT, Mn-BT, and Sn-BT. The observed trend suggests a pronounced decrease in with increasing temperature, signifying a transition toward a more conductive state at higher thermal energy levels. This can be attributed to the enhanced mobility of charge carriers within the samples lattice at elevated temperatures, leading to a diminished capacity for storing electrical energy. Furthermore, the higher values observed at lower frequencies indicate a more prominent contribution from relaxation phenomena, where the materials have sufficient time to reorient their internal dipoles in response to the alternating electric field. This response becomes progressively weaker at higher frequencies as the dipoles encounter difficulty in synchronizing with the rapidly fluctuating external field, leading to a decrease in .

- Temperature evolution of at various frequencies of (a) BT, (b) Mn-BT, (c) Sn-BT.
3.5.2. Dielectric loss
Figure 8 depicts the trend of dielectric losses as a function of frequency at various temperatures for BT, Mn-BT, and Sn-BT samples.

- Frequency evolution of at various temperatures of (a) BT, (b) Mn-BT, (c) Sn-BT.
Tan δ are notably high at low frequencies and elevated temperatures but decrease significantly at higher frequencies regardless of the temperature. This unique pattern in dielectric loss is consistent with Koop’s phenomenological theory [41].
3.6. Conductivity analysis
Electrical conductivity measurement in BT samples is a key property that affects their use in various applications, including environmental and material sciences. By examining the electrical behavior of undoped and metal-doped BT, researchers can gain insights into charge transport mechanisms and the impact of dopants on conductivity, leading to improved material performance and new applications. Figure 9 illustrates the electrical conductivity versus frequency at temperature range 25-130°C for BT, Mn-BT, and Sn-BT samples. Across all samples, the conductivity shows a gradual increase with rising frequency, followed by a sharp rise beyond the hopping frequency (F > Fh). This rise in conductivity at elevated frequencies could be assigned to electron hopping between grain sites, as described by Almond-West law (Eq. 5) [42,43]:

- Frequency evolution of at various temperatures of (a) BT, (b) Mn-BT, (c) Sn-BT.
Here, is a pre-exponential factor that determines the strength of polarizability, Fh denotes the hopping frequency, which distinguishes the direct current conduction from the dispersive regime, and p serves as the universal parameter, affecting the strength of the connection between mobile charges and the lattice (0 < p < 1). If p 1, the electron transition occurs through a sudden jump via translation. In contrast, for p >1, the electron transition is confined to neighboring sites [44]. By applying Eq. (5) to fit the conductivity data (solid line in Figure 9), the exponent p was obtained. Figure 10 depicts the thermal evolution of the exponent p for BT, Mn-BT, and Sn-BT. It is evident that p rises with increasing temperature before declining, indicating that the non-overlapping small polaron tunneling (NSPT) and correlated barrier hopping (CBH) models apply to all samples [45-47]. The same behavior was observed in our previous work Ce and Zn-doped BT [8]. However, doping significantly reduces the p values, suggesting that Mn and Sn incorporation alter the charge transport pathways. The lower p values in Mn-BT and Sn-BT imply a decrease in the effective energy barrier for charge hopping, which may be attributed to the introduction of additional localized states that facilitate charge mobility.

- Temperature evolution of exponent p of BT, Mn-BT and Sn-BT samples.
4. Conclusions
The investigation into Mn and Sn-doped BT composites has revealed substantial improvements in dielectric properties, which are critical for energy storage applications. The successful incorporation of Mn and Sn ions into the BT matrix was confirmed through various characterization techniques, highlighting changes in structural, optical, and electrical properties. Enhanced dielectric permittivity and reduced loss factors suggest that these composites can efficiently store and manage electrical energy, making them suitable for advanced electronic devices. Importantly, the scalability of these composites appears promising. The synthesis process can be adapted for industrial-scale production using cost-effective and abundant raw materials, such as BT. Moreover, the low toxicity and ready availability of Mn and Sn support the potential for eco-friendly manufacturing practices. These factors make Mn-BT and Sn-BT composites strong candidates for integration into capacitor production, offering a pathway toward high-performance and sustainable energy storage solutions. Furthermore, given their improved dielectric response, these materials hold promise for high-frequency electronic applications, such as capacitors for power electronics, resonators, and microwave devices. Their ability to operate efficiently at elevated frequencies makes them attractive for modern communication technologies and advanced energy storage systems. Future research could explore the doping effects of other transition metals, such as cobalt, nickel, or vanadium, based on their ionic radius and electronegativity, to further optimize BT’s dielectric properties. Additionally, impedance spectroscopy could be employed to gain deeper insight into the charge transport mechanisms and relaxation phenomena in these materials. These investigations would help refine the understanding of conduction processes and enhance the performance of BT-based energy storage materials. The promising results from this study pave the way for the development of high-performance, environmentally friendly energy storage materials.
CRediT authorship contribution statement
Ali H. Bashal: Conceptualization, methodology, validation, formal analysis, investigation, data curation, writing—original draft, writing—review & editing, supervision. Talat H. Habeeb: Software, writing—original draft, writing—review & editing, visualization. Abdulaziz Almalki: Conceptualization, methodology, validation, formal analysis, investigation, data curation, writing—original draft, writing—review & editing, supervision.
Declaration of competing interest
The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.
Declaration of Generative AI and AI-assisted technologies in the writing process
The authors confirm that there was no use of artificial intelligence (AI)-assisted technology for assisting in the writing or editing of the manuscript and no images were manipulated using AI.
References
- Double-layer and pseudocapacitance types of electrochemical capacitors and their applications to the development of hybrid devices. Journal of Solid State Electrochemistry. 2003;7:637-644. https://doi.org/10.1007/s10008-003-0395-7
- [Google Scholar]
- Electrochemical capacitors utilising transition metal oxides: An update of recent developments. RSC Advances. 2011;1:1171. https://doi.org/10.1039/c1ra00664a
- [Google Scholar]
- Electrical conductivity of metal (hydr)oxide–activated carbon composites under compression A comparison study. Materials Chemistry and Physics. 2015;152:113-122. https://doi.org/10.1016/j.matchemphys.2014.12.022
- [Google Scholar]
- Dielectric properties of non-swelling bentonite: The effect of temperature and water saturation. Journal of Non-Crystalline Solids. 2008;354:5533-5541. https://doi.org/10.1016/j.jnoncrysol.2008.03.050
- [Google Scholar]
- Adsorption of basic violet 14 from aqueous solution on bentonite. Comptes Rendus de l’Académie des Sciences - Series IIC - Chemistry. 2008;11:125-129.
- [Google Scholar]
- Influences of bentonite on conductivity of composite solid alkaline polymer electrolyte PVA-bentonite-KOH-H2O. Electrochimica Acta. 2007;52:7315-7321. https://doi.org/10.1016/j.electacta.2007.06.004
- [Google Scholar]
- Tailoring the dielectric properties of bentonite for advanced capacitor applications: The role of ruthenium and strontium incorporation. Journal of Saudi Chemical Society. 2024;28:101874. https://doi.org/10.1016/j.jscs.2024.101874
- [Google Scholar]
- Enhancing dielectric properties of bentonite with Ce and Zn: Structural insights and industrial applications. Journal of Sol-Gel Science and Technology. 2024;110:473-488. https://doi.org/10.1007/s10971-024-06372-2
- [Google Scholar]
- The impact of different metal dopants on the structural, dielectric, and electrical characteristics of bentonite: Electrical measurements supported by tight-binding calculations. Journal of Electronic Materials. 2024;53:4696-4705. https://doi.org/10.1007/s11664-024-11196-5
- [Google Scholar]
- A green, porous and eco-friendly magnetic geopolymer adsorbent for heavy metals removal from aqueous solutions. Journal of Cleaner Production. 2019;215:1233-1245. https://doi.org/10.1016/j.jclepro.2019.01.084
- [Google Scholar]
- Characterization of the cation-exchanged bentonites by XRPD, ATR, DTA/TG analyses and BET measurement. Chemical Engineering Journal. 2009;149:242-248. https://doi.org/10.1016/j.cej.2008.10.028
- [Google Scholar]
- Fixation of Zn2+ and Pb2+ by a ca-montmorillonite in brines and dilute solutions: Preliminary results. Applied Clay Science. 1996;11:117-126. https://doi.org/10.1016/s0169-1317(96)00014-2
- [Google Scholar]
- An investigation of Cu (II) adsorption by raw and acid-activated bentonite: a combined potentiometric, thermodynamic, XRD, IR, DTA study. Journal of Hazardous Materials. 2008;151:682. https://doi.org/10.1016/j.jhazmat.2007.06.040
- [Google Scholar]
- Structural characterization of aniline-bentonite composite by FTIR, DTA/TG, and PXRD analyses and BET measurement. Spectroscopy Letters. 2011;44:77-82. https://doi.org/10.1080/00387010903555953
- [Google Scholar]
- FTIR techniques in clay mineral studies, Vib. Spectros.. 2003;31:1-10. https://doi.org/10.1016/S0924-2031(02)00065-6
- [Google Scholar]
- In-situ FTIR analyses of bentonite under high-pressure. Applied Clay Science. 2011;51:202-208. https://doi.org/10.1016/j.clay.2010.11.017
- [Google Scholar]
- General introduction: clays, clay minerals, and clay science. Developments in Clay Science. 2006;1:1-18. https://doi.org/10.1016/S1572-4352(05)01001-9
- [Google Scholar]
- Preparation of al-magadiite material, copper ions exchange and effect of counter-ions: Antibacterial and antifungal applications. Research on Chemical Intermediates. 2019;45:633-644. https://doi.org/10.1007/s11164-018-3634-3
- [Google Scholar]
- TiO2 supported on HZSM-11 zeolite as efficient catalyst for the photodegradation of chlorobenzoic acids. Journal of the Brazilian Chemical Society. 2015;26:1191-1200.
- [Google Scholar]
- Structural, morphological, optical and dielectric properties of sodium bismuth titanate ceramics. Inorganic Chemistry Communications. 2022;146:110119. https://doi.org/10.1016/j.inoche.2022.110119
- [Google Scholar]
- Deep insights into the Cu- and w-doped Na0.5Bi0.5TiO3 solid solution: A study focusing on optical, dielectric and electrical properties. Journal of Molecular Structure. 2023;1294:136319. https://doi.org/10.1016/j.molstruc.2023.136319
- [Google Scholar]
- Dielectric, optical, piezoelectric and ferroelectric studies of NBT–BT ceramics near MPB. Ceramics International. 2015;41:10710-10717. https://doi.org/10.1016/j.ceramint.2015.05.005
- [Google Scholar]
- Insight into the structural, optical, adsorptive, and photocatalytic properties of MgFe2O4-bentonite nanocomposites. Journal of Physics and Chemistry of Solids. 2021;154:110060. https://doi.org/10.1016/j.jpcs.2021.110060
- [Google Scholar]
- Preparation and characterization of transparent ZnO thin films obtained by spray pyrolysis. Thin Solid Films. 2003;426:68. https://doi.org/10.1016/S0040-6090(02)01331-7
- [Google Scholar]
- Singh, investigation of crystal structure, dielectric properties, impedance spectroscopy and magnetic properties of (1–x) BaTiO3–(x)Ba0.9Ca0.1Fe12O19 multiferroic composites. Ceramics International. 2021;47:23088-23100. https://doi.org/10.1016/j.ceramint.2021.04.311
- [Google Scholar]
- On the dispersion of resistivity and dielectric constant of some semiconductors at audiofrequencies. Physical Review. 1951;83:121-124. https://doi.org/10.1103/physrev.83.121
- [Google Scholar]
- Interconnection between frequency-domain havriliak-negami and time-domain kohlrausch-williams-watts relaxation functions. Physical review. B, Condensed matter. 1993;47:125-130. https://doi.org/10.1103/physrevb.47.125
- [Google Scholar]
- Dielectric phenomena in solids. USA: Elsevier Academic Press; 2004. https://books.google.com.sa/books?hl=en&lr=&id=MxvRST5PT1cC&oi=fnd&pg=PP1&dq=K.C.+Kao,+Dielectric+phenomena+in+solids,+Elsevier+Academic+Press,+USA,+(2004).&ots=W8XV3tgmVx&sig=EqdLSqWkirCvz_I7idpouJI0I70&redir_esc=y#v=onepage&q&f=false
- Synthesis of M-type hexaferrite reinforced graphene oxide composites for electromagnetic interference shielding. Journal of Physics and Chemistry of Solids. 2024;185:111783. https://doi.org/10.1016/j.jpcs.2023.111783
- [Google Scholar]
- Synergistic effects of li-based ferrite and graphene oxide in microwave absorption applications. Synthetic Metals. 2024;307:117674. https://doi.org/10.1016/j.synthmet.2024.117674
- [Google Scholar]
- Innovative nanostructuring of Li–Zn ferrite/graphene composites with tunable properties. Ceramics International. 2024;50:39564-39573. https://doi.org/10.1016/j.ceramint.2024.07.335
- [Google Scholar]
- Optimizing ZnSe nanorods with La doping for structural, electrical, and dielectric properties in device applications. Surfaces and Interfaces. 2024;54:105275. https://doi.org/10.1016/j.surfin.2024.105275
- [Google Scholar]
- Investigating Al-doped Y-type hexaferrite/polyaniline composites: Synthesis, structural, and dielectric properties. Ionics. 2025;31:2319-2335. https://doi.org/10.1007/s11581-024-06025-y
- [Google Scholar]
- A critical review on multifunctional composites as structural capacitors for energy storage. Composite Structures. 2018;188:126-142. https://doi.org/10.1016/j.compstruct.2017.12.072
- [Google Scholar]
- Reliability evaluation of large scale battery energy storage systems. IEEE Transactions on Smart Grid. 2017;8:2733-2743. https://doi.org/10.1109/tsg.2016.2536688
- [Google Scholar]
- Performance optimization of Nd-doped LaNiO3 as an electrode material in supercapacitors. Solid State Ionics. 2023;395:116227.
- [Google Scholar]
- Advances in graphene-based electrode materials for high-performance supercapacitors: A review. Journal of Energy Storage. 2023;72:108731. https://doi.org/10.1016/j.est.2023.108731
- [Google Scholar]
- Recent developments in transition metal oxide-based electrode composites for supercapacitor applications. Journal of Energy Storage. 2024;81:110430. https://doi.org/10.1016/j.est.2024.110430
- [Google Scholar]
- Exploring new frontiers in supercapacitor electrodes through MOF advancements. Journal of Energy Storage. 2024;76:109822. https://doi.org/10.1016/j.est.2023.109822
- [Google Scholar]
- Elucidating an efficient super-capacitive response of a Sr2 Ni2 O5/rGO composite as an electrode material in supercapacitors. RSC advances. 2023;13:25316-25326. https://doi.org/10.1038/306456a0
- [Google Scholar]
- Development of cuboidal KNbO3@α-Fe2O3 hybrid nanostructures for improved photocatalytic and photoelectrocatalytic applications. ACS omega. 2020;5:20491-20505. https://doi.org/10.1021/acsomega.0c02646
- [Google Scholar]
- Anomalous conductivity prefactors in fast ion conductors. Nature. 1983;306:456-457. https://doi.org/10.1038/306456a0
- [Google Scholar]
- The determination of hopping rates and carrier concentrations in ionic conductors by a new analysis of ac conductivity. Solid State Ionics. 1983;8:159-164. https://doi.org/10.1016/0167-2738(83)90079-6
- [Google Scholar]
- Electrical conduction in (Na0.125Bi0.125 Ba0.65Ca0.1)(Nd0.065Ti0.87Nb0.065)O3 ceramic. Bulletin of Materials Science. 2006;29:35-41. https://doi.org/10.1007/BF02709353
- [Google Scholar]
- AC conductivity and dielectric measurements of bulk magnesium phthalocyanine (MgPc) Journal of Alloys and Compounds. 2009;480:564-567. https://doi.org/10.1016/j.jallcom.2009.01.124
- [Google Scholar]
- Jump relaxation in solid electrolytes. Progress in Solid State Chemistry. 1993;22:111-195. https://doi.org/10.1016/0079-6786(93)90002-9
- [Google Scholar]
- A.c. conduction in amorphous chalcogenide and pnictide semiconductors. Advances in Physics. 1987;36:135-217. https://doi.org/10.1080/00018738700101971
- [Google Scholar]
