Rima Altaliba,
Imen Ibrahmi*a,
Arafet Ghoudi
a,
Sami Znaidiab,
Walid Rekikc,
Jerome Lhoste
d and
Abderrazek Oueslati
a
aLaboratory of Spectroscopic Characterization and Optical Materials, Faculty of Sciences, University of Sfax, BP 1171, 3000 Sfax, Tunisia. E-mail: arafetghoudi199@gmail.com
bLaboratoire de recherche (LR 18ES19), Synthese Asymetrique et Ingenierie Moleculaire de Materiaux Organiques our L'electroniques Organiques, Facult des sciences de Monastir, 5000, Tunisia
cLaboratory Physical-Chemistry of Solid State, Chemistry Department, Faculty of Sciences of Sfax, University of Sfax, BP 1171, 3000, Sfax, Tunisia
dInstitut des Molécules et Matériaux du Mans (IMMM) – UMR-6283 CNRS, Le Mans Université, Avenue Olivier Messiaen, F-72085 Le Mans Cedex 9, France
First published on 11th September 2025
The hybrid compound (C6H9N2)3[BiBr6]H2O was synthesized via slow evaporation and structurally characterized using single-crystal X-ray diffraction. It crystallizes in the monoclinic C2/c space group and adopts a zero-dimensional architecture composed of isolated [BiBr6]3− octahedra, protonated organic cations (C6H9N2)+, and water molecules. These components are interconnected through hydrogen bonding and π–π interactions. Optical absorption measurements reveal a direct band gap of 2.81 eV, confirming the semiconducting nature of the material. Impedance spectroscopy, performed over a frequency range of 0.4 Hz to 3 MHz and a temperature range of 318 K to 363 K, reveals separate contributions from grains and grain boundaries. These were modeled using an equivalent circuit, indicating non-Debye relaxation behavior. The DC conductivity follows an Arrhenius-type behavior with activation energies of 0.96 and 0.51 eV. AC conductivity obeys Jonscher's power law, and the temperature-dependent decrease in the frequency exponent (s) supports the correlated barrier hopping (CBH) mechanism. The material exhibits enhanced dielectric permittivity, suggesting promising potential for optoelectronic and energy storage applications.
Bismuth(III) halide complexes coordinated with organic cations exhibit remarkable structural diversity. Halobismuthate(III) compounds typically consist of anionic sublattices formed by [BiX6]3− octahedra (X = halogen), which can adopt various connectivity modes such as corner-sharing, edge-sharing, or face-sharing. These arrangements give rise to a variety of dimensionalities, including discrete zero-dimensional (0D) units, one-dimensional (1D) chains, and two-dimensional (2D) layered frameworks.12 The resulting crystal packing is stabilized by a combination of hydrogen bonding, van der Waals forces, electrostatic interactions, and halide–halide contacts.13 Importantly, the choice of organic cation not only influences the structural organization but also plays a pivotal role in modulating the material's electronic and dielectric properties.
Aromatic amine-derived cations, such as substituted pyridinium and imidazolium species, offer distinct advantages over their aliphatic counterparts. Their relatively high dielectric constants reduce dielectric confinement, leading to lower exciton binding energies. Moreover, their π-conjugated and rigid structures facilitate stronger π–π stacking and hydrogen bonding interactions, enhancing structural stability, charge transport, and dielectric screening—critical factors for optoelectronic performance in layered halide systems.14–19
Several hybrid halobismuthate(III) materials incorporating aromatic amines have been reported, showing encouraging characteristics for optoelectronic applications. Examples include (C9H12N4)2[BiBr6]Cl4H2O,10[C13H16N2]5(BiCl6)3Cl,20 (C6H20N3)BiI6H2O,21 and [C4H10N]3[BiCl6],22 all of which demonstrate good environmental stability and functional potential. Notably, M. Hamdi et al.23 reported a novel organic–inorganic hybrid material [C6H10N2]7[Bi2Cl11]2·4[Cl], exhibiting a dielectric constant exceeding 104 at 423 K. This breakthrough has further heightened interest in such materials for advanced applications.
Within this promising field of research, we report the synthesis and characterization of a new hybrid compound (C6H9N2)3[BiBr6]·H2O, composed of aromatic organic molecules as the organic component, with bismuth(III) bromide as the inorganic counterpart. To elucidate its structural and physicochemical properties, a range of experimental techniques has been employed, including single-crystal X-ray diffraction, UV-visible spectroscopy, and impedance spectroscopy. The results provide valuable insight into the compound's charge transport behavior, luminescent behavior, and charge transport mechanisms. These findings contribute to the rational design of environmentally benign, lead-free hybrid materials with strong potential for optoelectronic and energy-related applications.
3(C6H8N2) + BiBr3 + 3HBr + H2O → (C6H9N2)3[BiBr6]·H2O | (1) |
Formula | (C6H9N2)3[BiBr6]·H2O |
---|---|
Color/shape | Yellow/prism |
Formula weight (g mol−1) | 1033.91 |
Crystal system | Monoclinic |
Space group | C2/c |
Density | 2.285 |
Crystal size (mm) | 0.22 × 0.18 × 0.14 |
Temperature (K) | 296(2) |
Diffractometer | Bruker APEXII |
a (Å) | 25.208(4) |
b (Å) | 12.9804(13) |
c (Å) | 19.878(2) |
β (°) | 112.470(4) |
V (Å3) | 6010.3(13) |
Z | 8 |
Radiation type | Mo Kα (0.71073 Å) |
Absorption correction | Multi-scan |
θ range for data collection (°) | 1.748 ≤ θ ≤ 27.514 |
Measured reflections | 53![]() |
Independent reflections | 6904 |
Observed data [I > 2σ(I)] | 4527 |
Index ranges | h = −32 → 32 |
k = −16 → 16 | |
l = −25 → 25 | |
F(000) | 3840 |
Number of parameters | 297 |
R1 | 0.0400 |
wR2 | 0.0769 |
Goof | 1.018 |
BiBr6 octahedra | Organic moieties | ||
---|---|---|---|
a Symmetry codes: I 1 − x, y, 1/2 − z: II 3/2 − x, 1/2 − y, 1 − z. | |||
Bi1–Br3 | 2.7537(10) | C5–C1 | 1.489(11) |
Bi1–Br1 | 2.8341(7) | N2–C6 | 1.327(9) |
Bi1–Br1I | 2.8342(7) | C1–C4 | 1.330(10) |
Bi1–Br2I | 2.8472(7) | C1–C2 | 1.393(10) |
Bi1–Br2 | 2.8473(7) | C2–C3 | 1.353(10) |
Bi1–Br4 | 2.9841(11) | C3–C6 | 1.383(10) |
Bi2–Br5II | 2.8474(7) | C4–N11 | 1.347(9) |
Bi2–Br5 | 2.8475(7) | N11–C6 | 1.328(9) |
Bi2–Br7II | 2.8534(7) | C7–C10 | 1.342(8) |
Bi2–Br7 | 2.8534(7) | C7–C9 | 1.395(8) |
Bi2–Br6 | 2.8559(7) | C7–C8 | 1.506(8) |
Bi2–Br6II | 2.8559(7) | N10–C12 | 1.346(8) |
Br3–Bi1–Br1 | 90.681(16) | N10–C10 | 1.357(8) |
Br3–Bi1–Br1I | 90.681(16) | C9–C11 | 1.366(8) |
Br1–Bi1–Br1I | 178.64(3) | C11–C12 | 1.404(8) |
Br3–Bi1–Br2I | 88.043(17) | C12–N4 | 1.324(8) |
Br1–Bi1–Br2I | 90.65(3) | C14–C13 | 1.308(11) |
Br1I–Bi1–Br2I | 89.40(2) | C14–N9 | 1.328(11) |
Br3–Bi1–Br2 | 88.043(17) | N9–C18 | 1.289(18) |
Br1–Bi1–Br2 | 89.40(3) | C13–C17 | 1.360(9) |
Br1–Bi1–Br2I | 90.65(3) | C13–C15 | 1.528(13) |
Br2–Bi1–Br2I | 176.09(3) | N–C18 | 1.374(19) |
Br3–Bi1–Br4 | 180.0 | C18–C16 | 1.340(10) |
Br1–Bi1–Br4 | 89.319(16) | C16–C17 | 1.33(3) |
Br1–Bi1–Br4I | 89.319(16) | C4–C1–C2 | 115.2(8) |
Br2–Bi1–Br4I | 91.957(17) | C4–C1–C5 | 122.2(8) |
Br2–Bi1–Br4 | 91.957(17) | C2–C1–C5 | 122.5(7) |
Br5–Bi2–Br5II | 180.0 | C3–C2–C1 | 122.2(7) |
Br5II–Bi2–Br7II | 93.31(2) | C2–C3–C6 | 120.3(8) |
Br5–Bi2–Br7II | 86.69(2) | C1–C4–N11 | 122.6(7) |
Br5II-Bi2-Br7 | 86.69(2) | C6–N11–C4 | 123.3(7) |
Br5–Bi2–Br7 | 93.31(2) | N2–C6–N11 | 120.3(8) |
Br7–Bi2–Br7II | 180.0 | N2–C6–C3 | 123.4(8) |
Br5–Bi2–Br6II | 90.96(2) | N11–C6–C3 | 116.3(8) |
Br5–Bi2–Br6 | 89.04(2) | C10–C7–C9 | 117.2(6) |
Br7–Bi2–Br6II | 92.54(2) | C10–C7–C8 | 122.3(6) |
Br7–Bi2–Br6 | 87.456(19) | C9–C7–C8 | 120.6(6) |
Br5II–Bi2–Br6II | 89.04(2) | C12–N10–C10 | 123.7(5) |
Br5II–Bi2–Br6 | 90.96(2) | C11–C9–C7 | 122.0(6) |
Br7II–Bi2–Br6II | 87.458(19) | C7–C10–N10 | 121.0(6) |
Br7II–Bi2–Br6 | 92.54(2) | C9–C11–C12 | 119.3(6) |
Br6–Bi2–Br6II | 180.0 | N4–C12–N10 | 119.5(6) |
N4–C12–C11 | 123.8(7) | ||
N10–C12–C11 | 116.7(6) | ||
C13–C14–N9 | 121.8(9) | ||
C18–N9–C14 | 120.9(12) | ||
C14–C13–C17 | 116.1(10) | ||
C14–C13–C15 | 120.2(12) | ||
C17–C13–C15 | 123.7(13) | ||
N9–C18–C16 | 121.0(17) | ||
N9–C18–N | 117(2) | ||
C16–C18–N | 122(3) | ||
C17–C16–C18 | 116.9(19) | ||
C16–C17–C13 | 123.1(16) |
CCDC 2371583 contains supplementary crystallographic data for (C6H9N2)3[BiBr6]·H2O. These data can be obtained free of charge from the Cambridge Crystallographic Data Center via https://www.ccdc.cam.ac.uk/data_request/cif.
For electrical characterization, the powder obtained from ground crystals was compressed into a pellet (8 mm in diameter and 1.1 mm thick) using a uniaxial hydraulic press under a pressure of 3500 tons per cm2. Thin silver layers, a few nanometers thick, were manually applied to both flat surfaces of the pellet. The prepared pellet was then positioned between two platinum electrodes to perform electrical measurements. Complex impedance spectra were collected across a wide frequency range [0.4–3 × 106 Hz] and at different temperatures [318–363 K] using a Solartron 1260 frequency response analyzer.
In the present work, a complete substitution of Sb3+ by Bi3+ leads to a new compound with a distinct chemical formula and crystal structure. Specifically, the combination of 2-amino-5-picoline with bismuth tribromide yielded a novel organic–inorganic hybrid material, (C6H9N2)3[BiBr6]·H2O. This compound crystallizes in the monoclinic system, centrosymmetric space group C2/c, with unit cell parameters: a = 25.208(4) Å, b = 12.9804(13) Å, c = 19.878(2) Å, β = 112.470(4)°, and V = 6010.3(13) Å3. The crystal structure consists of [BiBr6]3− octahedra, protonated 2-amino-5-picolinium cations (C6H9N2)+, and free water molecules. These building units are interconnected through a network of hydrogen bonds and π π⋯π π interactions, resulting in a stable zero-dimensional (0D) architecture (Fig. 1).
The asymmetric unit of (C6H9N2)3[BiBr6]·H2O, depicted in Fig. 2, comprises two crystallographically independent bismuth(III) ions. The first ion, Bi1, occupies a special position on a twofold axis (Wyckoff site 4e). It is coordinated to four bromide ions: two located at general positions (Br1 and Br2) and two situated on special positions on the same twofold axis (Br3 and Br4). Its octahedral coordination sphere is completed by two additional bromide ions generated by the symmetry code (1 − x, y, 1/2 − z).
![]() | ||
Fig. 2 The asymmetric unit of (C6H9N2)3[BiBr6]·H2O. Hydrogen bonds are represented by dashed lines (symmetry codes: I 1 − x, y, 1/2 − z: II 3/2 − x, 1/2 − y, 1 − z). |
The second bismuth(III) ion, Bi2, occupies a special position on an inversion center (Wyckoff site 4c) and is directly bonded to three bromide ions (Br5, Br6, Br7) located at general positions. Three more bromide ions (Br5′, Br6′, Br7′) generated by an inversion center complete the coordination sphere around Bi2.
As shown in Fig. 1, the hexabromidobismuthate(III) ions, [BiBr6]3−, are arranged in such a way that they form inorganic layers parallel to the (1 0 −1) plane and are isolated from one another with a minimum intermetallic Bi–Bi distance of 8.7694(8) Å. Within the [BiBr6]3− anions, the Bi–Br bond distances range from 2.7537(10) to 2.9841(11) Å for Bi1Br6, and from 2.8474(7) to 2.8559(7) Å for Bi2Br6 while the cis-Br–Bi–Br angles vary between 88.043(17) and 91.957(17)° and between 86.69(2) and 93.31(2)° for Bi1Br6 and Bi2Br6, respectively (Table 2). These geometric features are consistent with those observed in other compounds containing BiBr6 octahedra.30–32 The Bi–Br bond distance in this study, 2.8512(7) Å, is longer than the Sb–Br bond distance found in (C6H9N2)2[SbBr4]Br, where Sb–Br distances range from 2.5698(5) to 2.709(4) Å, with a mean value of 2.6817(4) Å.29
This difference is expected, as the Bi3+ ion is more voluminous than the Sb3+ ions (rBi3+ = 1.2 Å and rSb3+ = 0.76 Å). Considering the geometrical characteristics of the [BiBr6]3− anions in our compound, the calculated average Baur indices (DI) were obtained using the following equation:33
![]() | (2) |
![]() | (3) |
The calculated distortion indices, DI (Bi1–Br) = 0.0156 and DI (Bi2–Br) = 0.0011, along with DI (Br–Bi–Br) values of 0.012 and 0.025 for Bi1Br6 and Bi2Br6, respectively, indicate that the coordination geometry of bismuth is a slightly distorted octahedron. This distortion is due to the environment surrounding the [BiBr6]3− octahedra and particularly the intermolecular hydrogen bonds formed with the organic cations and water molecules.
The negative charges of the anionic [BiBr6]3− octahedra are compensated by the protonated amines, (C6H9N2)+, which are arranged to form organic cationic layers parallel to the (1 0 −1) plane (Fig. 1). The crystal structure of the title compound can thus be described as an alternating pattern of organic and inorganic layers, which are directed towards the [1 0 1] direction (Fig. 1). The bond distances and angles characteristic of the protonated amines, listed in Table 2, are consistent with those found in the other compounds containing the same organic cation.29,34–37 Unlike the anionic octahedra, which are isolated from one another, the organic cations are connected through π⋯π interactions in a parallel-displaced configuration of the amine aromatic rings.38,39 Indeed, the shortest distance between the planes of two adjacent protonated amine aromatic rings is 3.6005(3) Å (Fig. S1). The π⋯π interactions in the title compound are slightly weaker than those found in the previously reported antimony-based compound, (C6H9N2)2[SbBr4]Br, where the shortest centroid-to-centroid distance between aromatic rings is 3.4422(2) Å.29
The asymmetric unit of (C6H9N2)3[BiBr6]·H2O contains only one free water molecule, which plays a crucial role in the cohesion of the crystal structure. Specifically, it facilitates connections between the anionic octahedra and the cationic organic entities through hydrogen bonds. This water molecule acts as a donor in Ow–H⋯Br hydrogen bonds and as an acceptor in N–H⋯Ow interactions (Fig. 3(a)). Notably, the anionic and cationic entities are directly linked via N–H⋯Br hydrogen bonds and weak non-covalent C–H⋯Br interactions. Within these intermolecular hydrogen bonds, distances are as follows: N⋯Br distances range from 3.324(11) to 3.926(6) Å, C⋯Br distances are comprised between 3.695(6) and 3.881(6) Å, O⋯Br distances vary between 3.476(7) and 3.827(11) Å, and the N⋯O distance measures 2.834(11) Å. The D–H⋯A angles fall within the 121.1–171.3° (Table S2).
![]() | ||
Fig. 3 Hydrogen bonds established by (a) the water molecule and (b) the protonated amine in (C6H9N2)3[BiBr6]·H2O. |
In semiconductors and insulating materials, the energy band gap (Eg) is the smallest amount of energy required to excite an electron from the valence band to the conduction band through photon absorption. This gap significantly influences the optical properties of the material, including its absorption characteristics, color, and transparency. As a result, Eg is an important parameter across various fields, such as solar energy conversion, light emission technologies, and laser systems.
According to the method proposed by Marotti, Henríque, and their collaborators,44,45 the optical bandgap energy (Eg) can be estimated from the reflectance spectrum (R(λ)) by identifying the maximum of the function(1/R(λ))(dR(λ)/dλ). As shown in Fig. S2, this function exhibits a prominent peak at 439 nm for the studied material. Using the standard relation E(eV) = 1240/λ(nm),46 the corresponding band gap energy is calculated to be 2.82 eV. This value falls within the typical range for semiconductors (0.5–5 eV) and is comparable to that of the related compound (C9H12N4)2[BiBr6]·Cl·4H2O (3.483 eV),10 although it is lower than that of [C8H12N]3BiCl6 (4.42 eV).47
To determine whether the optical band transition mode of the studied materials is direct or indirect, Tauc's law was applied,48 as expressed by:
(F(R)hν)1/n = C(hν − Eg) | (4) |
The optical band gap is determined using the Tauc plot method, where [F(R(λ))hν]2 (direct transitions) and [F(R(λ))hν]2 (indirect transitions) are plotted against photon energy. The band gap was estimated by extrapolating the linear portion of the Tauc plot to the energy axis, where [F(R)hν]n = 0 (Fig. 5). For (C6H9N2)3[BiBr6]·H2O, this analysis yields an indirect band gap of 2.87 eV and a direct band gap of 2.81 eV, consistent with the reflectance-derived value of 2.82 eV (Fig. S2).
Although both types of transitions were identified, the direct transition dominates the optical behavior of the compound. This conclusion is substantiated by its exact correspondence with the reflectance-derived Eg (2.82 eV, Fig. S2) and the characteristic steep absorption onset at 440 nm, hallmarks of direct-gap semiconductors. The marginally higher indirect gap may originate from phonon-assisted processes but remains spectroscopically secondary. Consequently, all subsequent optoelectronic interpretations reference the direct transition at 2.81 eV.
![]() | ||
Fig. 6 Variation of the Nyquist plots measured at different temperatures, 318–363 K, with the proposed equivalent circuits for the investigated compound. |
Fig. 7 presents the frequency dependence of the real part of the complex impedance (Z′) at various temperatures. The data reveal three distinct regions, each characterized by different conduction behaviors influenced by both frequency and temperature:
• Low-frequency region: at low frequencies, Z′ exhibits a strong temperature dependence, while remaining largely independent of frequency. A plateau is often observed, suggesting that the response is dominated by grain boundary effects. This behavior is typically attributed to charge carrier accumulation at grain boundary interfaces, leading to interfacial polarization and pronounced low-frequency impedance dispersion.
• Intermediate-frequency region (∼102 to 105 Hz): in this region, both frequency and temperature significantly influence the Z′ values. Z′ decreases with increasing frequency and also diminishes with rising temperature, indicating enhanced charge carrier mobility. This behavior corresponds to an increase in AC conductivity, which becomes more prominent with thermal activation.
• High-frequency region: at higher frequencies, the Z′ values exhibit convergence across all temperatures, reflecting reduced impedance contributions. This convergence may be associated with the release of space charges and a corresponding reduction in interfacial polarization effects. Additionally, the temperature-induced lowering of potential barriers facilitates charge transport, resulting in a further decrease in AC resistance.51
The imaginary component of the complex impedance (−Z′′) as a function of frequency at various temperatures is presented in Fig. 8. The spectra exhibit well-defined peaks corresponding to the maximum of −Z′′ (denoted as ) at specific frequencies (ωmax), which are associated with the material's electrical relaxation processes. These peaks occur within the dispersion region of the real part of impedance (Z′). As the temperature increases, the relaxation peaks shift toward higher frequencies, indicating that the relaxation mechanism is thermally activated.52 The observed peaks are asymmetric and broadened over the entire temperature range, which is characteristic of a non-Debye type relaxation behavior, implying the presence of a distribution of relaxation times. Furthermore, the convergence of −Z′′ at high frequencies suggests the possible contribution of space charge relaxation effects.53 This behavior of −Z′′ at both low and high frequencies is consistent with previously reported results for similar organic–inorganic hybrid systems.54–56
As depicted in Fig. 9, the conductivity spectra display two distinct regions. In the low-frequency range (1–103 rad s−1), σac remains almost constant, implying minimal frequency dependence. This plateau suggests a thermally activated transport process, likely associated with DC conductivity, which becomes more pronounced with increasing temperature. At higher frequencies, the conductivity increases, forming a dispersive region. This behavior is attributed to localized charge carriers within the grains acquiring enough energy to overcome potential barriers over short distances. The conductivity follows Jonscher's universal power law:34
σac = σdc + Aωs | (5) |
Furthermore, the temperature dependence of the frequency exponent s provides valuable insights into the dominant charge transport mechanism within the material.58–61 By fitting the experimental AC conductivity data to Jonscher's power law, important trends emerge, particularly about how s varies with temperature, as depicted in Fig. 10. The results show a clear decrease in the value of s with increasing temperature, a behavior that is consistent with the correlated barrier hopping (CBH) conduction model.62
![]() | (6) |
![]() | (7) |
Fig. 10 presents as well the linear fitting of the (1 − s) parameter as a function of temperature for the investigated material. From the slope of this fit, the maximum barrier height (WM) is estimated to be approximately 0.32 eV. Notably, the exponent s remains consistently below unity across the entire temperature range, which reinforces the interpretation that charge transport occurs via localized hopping processes. This behavior aligns well with the theoretical framework proposed by Funke, which describes charge carrier dynamics in disordered materials through thermally activated hopping between energetically favorable sites.
![]() | (8) |
M = jωC0Z = M′ + jM′′ | (9) |
The temperature-dependent evolution of the real part of the electrical modulus (M′) across the studied frequency range is illustrated in Fig. 12. At low frequencies, M′ remains close to zero for all temperatures, indicating a negligible contribution from electrode polarization. Conversely, at higher frequencies, M′ increases and eventually tends toward a saturation value M∞. This behavior reflects relaxation phenomena and charge conduction associated with localized, short-range carrier mobility.
Fig. 13 shows the frequency-dependent behavior of the imaginary part of the electrical modulus (M′′) at various temperatures. The results reveal a broad, asymmetric relaxation peak. At low frequencies, the M′′ values approach zero, indicating a negligible contribution from electrode effects.64 Each curve displays a single asymmetric peak, and the position of the peak maximum (ωmax) shifts toward higher frequencies as temperature increases, highlighting the thermally activated nature of the relaxation process.
In the frequency region below ωmax, charge carriers are likely able to move over long distances. In contrast, in the region beyond ωmax, they become confined within potential wells, limiting their mobility to shorter distances.20 The imaginary component M′′(ω) was modeled using the following relation:
![]() | (10) |
The best fits of M′′(ω) across the temperature range are displayed in Fig. 13. The corresponding β values are summarized in Table S4, clearly showing a temperature-dependent trend. All values fall within the interval 0 < β < 1, indicating a non-Debye type relaxation and reflecting the strong coupling between mobile ions involved in the conduction mechanism.
ε(ω) = ε′(ω) + jε′′(ω) | (11) |
Fig. 14(a) and (b) illustrate the frequency dependence of ε′ and ε′′ at various temperatures ranging from 318 K to 363 K. From the data, it is evident that the real part ε′ increases significantly at low frequencies with rising temperature, followed by a sharp decline beyond approximately 102 Hz. This behavior confirms that ε′(ω) is influenced by bound charge polarization, which plays a role in the material's ability to store electric energy.
![]() | ||
Fig. 14 Frequency-dependent (a) real part (ε′) and (b) imaginary part (ε′′) of the dielectric permittivity. |
Similarly, the imaginary part ε′′ also increases with temperature, which reflects the presence of a broad distribution of relaxation times in the studied compound. The imaginary component ε′′ provides insight into the energy dissipation processes. In general, dielectric response is governed by four types of polarization mechanisms: interfacial, dipolar, electronic, and ionic. At low frequencies, ε′ is mainly dominated by interfacial and dipolar polarization, indicating a deviation from ideal Debye relaxation behavior. The pronounced decline of ε′ at higher frequencies is attributed to the gradual loss of space charges, which otherwise enhance the dielectric response.
The dielectric measurements reveal that the investigated compound exhibits a remarkably high real permittivity (ε′), reaching up to 105 at 363 K. To ensure that this exceptionally elevated dielectric constant is not significantly affected by electrode polarization, a detailed analysis was conducted using the electric modulus formalism. As illustrated in Fig. 14 and 15, both the real (M′) and imaginary (M′′) components of the complex modulus approach zero at low frequencies across the entire studied temperature range (318–363 K), indicating negligible electrode polarization effects. This behavior suggests that the observed dielectric response is dominated by intrinsic material properties rather than interfacial electrode contributions. Additionally, impedance spectroscopy data, reinforced by equivalent circuit modeling (inset of Fig. 6), further support the absence of significant electrode-related artifacts, as no low-frequency semicircle typically associated with electrode effects was detected.
To contextualize this result, Table 3 presents a comparison between the dielectric constant of (C6H9N2)3[BiBr6] H2O and those of similar bismuth-based or hybrid compounds. The permittivity observed in our compound significantly surpasses that of other reported materials, such as [C13H16N2]5(BiCl6)3Cl (ε′ ≈ 104),20 (pyrrolidinium)3[Bi2I9] (ε′ ≈ 500),68 and conventional semiconductors like the B4ATCZ crystal (ε′ ≈ 26.2).69 This elevated dielectric response is likely attributed to intrinsic structural features, including interfacial polarization at grain boundaries, strong dipolar interactions, and thermally activated charge hopping mechanisms. Collectively, these characteristics point to the promising potential of (C6H9N2)3[BiBr6] H2O for advanced dielectric applications, such as capacitive energy storage, microelectronics, and photonic technologies.
The variation of the dielectric loss tangent, tg(δ), as a function of frequency at 363 K is displayed in Fig. 15. The dielectric loss factor, tg(δ), is defined by the relation:
![]() | (12) |
At lower frequencies, tg(δ) exhibits a prominent peak before gradually decreasing with increasing frequency. This behavior reflects the presence of two distinct relaxation processes in the material. The first peak appears in the low-frequency range (1–103 rad s−1) and is attributed to space charge polarization. The second peak, located in the mid-frequency domain (103–106 rad s−1), is associated with dipolar polarization. These observations indicate the contribution of multiple polarization mechanisms, including electronic, ionic, orientational, and space charge effects, most notably in the low-frequency region.73,74
As frequency increases, the material's resistivity decreases, enabling easier movement of charge carriers. Consequently, the energy required for their motion diminishes, which explains the reduction in dielectric loss observed at higher frequencies. Furthermore, Fig. 15 demonstrates that tg(δ) values for the investigated compound remain relatively low, ranging approximately from 0.01 to 35. This low dielectric loss is particularly beneficial for optoelectronic device applications.75–77
Given these properties, the (C6H9N2)3[BiBr6]. H2O compound shows great potential for use in optoelectronic systems. It is also important to note that the dielectric loss tangent, tg(δ), plays a crucial role in determining energy efficiency and thermal stability of materials. Higher values of tg(δ) generally correspond to greater energy dissipation and heat generation. In contrast, controlled or minimal dielectric losses, as seen in capacitors, can enhance device efficiency. Thus, tg(δ) significantly influences the performance of technologies related to electronics, telecommunications, and energy storage.
The DC conductivity exhibits Arrhenius-type temperature dependence, with two distinct activation energies, which reflect a change in the conduction mechanism. AC conductivity analysis, along with the temperature-dependent behavior of the frequency exponent s, supports the correlated barrier hopping (CBH) model as the primary conduction process. Additionally, the compound demonstrates significant dielectric permittivity, indicating a strong response to external electric fields, an essential feature for energy storage and conversion applications.
Overall, these results underscore the potential of (C6H9N2)3[BiBr6]·H2O as a promising candidate for future development in optoelectronic devices and advanced energy-related technologies.
CCDC 2371583 contains the supplementary crystallographic data for this paper.78
All data supporting the findings of this study are available within the article and its SI. See DOI: https://doi.org/10.1039/d5ra04097c.
This journal is © The Royal Society of Chemistry 2025 |