Open Access Article
Amin Alibia,
Nour Elleucha,
Sergiu Shovab,
Jerome Lhostec,
Guillaume Duvalc and
Mohamed Boujelbene
*a
aLaboratory of Physico-Chemistry of Solid State, LR11ES51, Sfax Faculty of Sciences, University of Sfax, Sfax 3000, Tunisia
b“Petru Poni” Institute of Macromolecular Chemistry, Alea Grigore Ghica Voda 41-A, 700487 Iasi, Romania
cMMM-UMR 6283 CNRS, Lunam, Faculty of Sciences and Techniques, University of Maine, Avenue Olivier Messiaen, 72085, Le Mans Cedex 9, France
First published on 23rd October 2025
A newly synthesized bismuth-based organic–inorganic hybrid material, (C8H14N2)2[Bi2Br10]·2H2O, was synthesized using a slow evaporation technique and thoroughly characterized to explore its structural, vibrational, optical, and electronic properties. Single-crystal X-ray diffraction confirmed its monoclinic crystal system with a centrosymmetric P21/c space group, featuring edge-sharing [Bi2Br10]4− dimers interconnected via non-covalent interactions. Hirshfeld surface analysis and fingerprint plots revealed dominant H⋯Br and H⋯H interactions, highlighting the role of hydrogen bonding in stabilizing the crystalline architecture. Vibrational studies using FTIR and Raman spectroscopy provided detailed assignments of molecular vibrations, corroborated by density functional theory (DFT) calculations. The optical properties were investigated through UV-vis spectroscopy in solution and diffuse reflectance spectroscopy (DRS) in the solid state, revealing an indirect band gap of 2.9 eV (solid-state) and 3.086 eV (solution), validated by theoretical electronic structure calculations. Photoluminescence (PL) studies demonstrated a strong blue and rose emission, supported by CIE chromaticity analysis, positioning this material as a promising candidate for optoelectronic applications. Additionally, advanced DFT analyses, including electron localization function (ELF), localized orbital locator (LOL), reduced density gradient (RDG), and non-covalent interaction (NCI) analyses, provided deep insights into electronic charge distribution, weak intermolecular interactions, and structural stability. These findings establish (C8H14N2)2[Bi2Br10]·2H2O as a bismuth-based hybrid material with significant potential in optoelectronics, luminescent materials, and functional optical applications.
Beyond structural interest, these compounds exhibit ferroelasticity, semiconductivity,7–10 thermochromism,11–13 and photochromism,14–16 though optimizing multiple properties within a single material remains challenging. The impact of structural phase transitions on electrical properties is still not fully understood, limiting material design predictability. While luminescence studies on polynuclear Bi(III) halides (polyhalide bismuthates, PHBs) have increased,17–19 they remain relatively sparse. Luminescence typically originates from aromatic cations, while PHB crystal packing significantly influences optical properties.20 These hybrid materials, represented by the general formula RaMbX(3b+a) (X = Cl, Br, I; M = Sb(III), Bi(III); R = organic cation), form discrete or polyoctahedral anions in 0D to 3D networks.21,22
This article is divided into two main sections. The first examines the crystal structure and molecular composition of (C8H14N2)2[Bi2Br10]·2H2O using SCXRD, FT-IR, Raman spectroscopy, and Hirshfeld surface analysis to understand molecular geometry, vibrational modes, and intermolecular interactions, with a focus on hydrogen bonding and halogen coordination. The second section explores its optical and luminescence properties, emphasizing material applications, supported by in-depth DFT calculations. The interplay between the organic cation and the Bi–Br framework is analyzed to understand tunable physical properties such as solvatochromism and photoluminescence. This study aims to expand knowledge of bismuth halides for advanced material development.
We present the synthesis, DFT calculations, and in-depth analysis of bis(N,N′-di(4-ethyl aminomethyl)pyridinium) decabromodibismuthate(III) dihydrate, focusing on its structural, vibrational, and optical properties. Key techniques include XRD, Hirshfeld surface analysis, IR and Raman spectroscopy, and optical studies through photoluminescence (PL) and UV-vis spectroscopy in both liquid and solid states. DFT calculations (Gaussian 09W and Multiwfn) provide insights into its electronic structure. This research contributes to the growing understanding of bismuth-based hybrids, addressing structure–performance relationships to facilitate high-efficiency, stable material design.
The optical properties of the synthesized material were investigated by recording the diffuse reflectance spectra (DRS) at room temperature using a PerkinElmer Lambda 35 UV-vis spectrophotometer equipped with an integrating sphere. An 8 mm diameter sample pellet was used, and spectra were captured in the range of 200 to 1100 nm. Furthermore, UV-vis spectra in aqueous solution were obtained using a Cary 5000 UV-vis-NIR spectrophotometer, within the 250–600 nm range. The photoluminescence properties of the compound were assessed at room temperature using a PerkinElmer LS 55 spectrometer. The sample was dissolved in a suitable solvent to prepare a clear solution, which was placed in a quartz cuvette for measurement. The emission spectrum was recorded under ambient conditions using excitation wavelengths of 319 and 350 nm. The relationship between the UV-vis spectra in both liquid and solid states, along with the TD-DFT analysis, was employed to further confirm the energy gap and to ascertain whether the band gap corresponds to a direct or indirect transition.
:
1 molar ratio. Each solution was stirred for about 30 minutes to achieve uniform mixing. After combining the solutions, concentrated hydrobromic acid (HBr, 48% purity) was carefully introduced in three equal portions at intervals of 30 minutes, with the mixture continuously stirred. Following the final addition, the solution was stirred for another 30 minutes, resulting in a total stirring duration of 1.5 hours.
The combined solution was left undisturbed for four days to allow crystal formation through slow evaporation. This process produced yellow plate-shaped crystals with consistent morphology. The crystals were collected through filtration, and a single crystal was selected for detailed structural analysis using single-crystal X-ray diffraction. The synthesis method was straightforward and reproducible, with no difficulties encountered. Furthermore, the compound exhibited good stability under normal storage conditions without requiring special precautions.
The compositional purity and structural integrity of the compound (C8H14N2)2[Bi2Br10]·2H2O were first assessed by energy-dispersive X-ray spectroscopy (EDS) combined with elemental mapping, as presented in the upper part of Fig. 1.S. The EDS spectrum displays exclusively the characteristic peaks of nitrogen, bromine, and bismuth, with no detectable contributions from extraneous elements, thereby confirming the excellent chemical purity of the synthesized material. The corresponding elemental distribution maps further demonstrate the homogeneous dispersion of Br and Bi across the crystalline domains, while the nitrogen signal highlights the uniform incorporation of the organic cations within the hybrid lattice. It is worth noting that carbon, oxygen, and hydrogen atoms are not observed in the EDS results. This limitation is inherent to the technique: hydrogen cannot be detected at all due to its extremely low atomic number, while carbon and oxygen, although present in the organic cations and hydration water, often produce weak or unreliable signals that are absorbed or overlap with background noise. For this reason, EDS predominantly provides information on the heavier inorganic elements, and its results are interpreted accordingly. Importantly, the absence of any additional signals beyond N, Br, and Bi unambiguously confirms the high purity and compositional integrity of the compound.
Complementary structural characterization was performed using powder X-ray diffraction (PXRD), shown in the lower part of Fig. 1.S. The experimental diffractogram (red curve) reveals sharp and well-defined reflections, indicative of a high degree of crystallinity, and matches closely with the simulated pattern calculated from the single-crystal structural data (black curve). The excellent agreement in both peak positions and intensities excludes the presence of secondary phases and validates the proposed structural model of the dimeric bismuth–bromide hybrid material (C8H14N2)2[Bi2Br10]·2H2O. Together, the EDS and PXRD analyses unequivocally confirm both the compositional and structural purity of the material.
| Crystallographic data | |
|---|---|
| Empirical formule | (C8H14N2)2[Bi2Br10]·2H2O |
| Color/shape | Yellow/plate |
| Molar mass (g mol−1) | 1529.51 |
| Diffractometer | XtaLAB synergy, Dualflex, HyPix diffractometer |
| Radiation type | Mo Kα (0.71073 Å) |
| Temperature (K) | 293 |
| Calculated density (Mg m−3) | 2.950 |
| Crystal system | Monoclinic |
| Space group | P21/c |
| Z/Z′ | 2/0.5 |
| Unit cell parameters a (Å) | 10.6409 (2) |
| b (Å) | 12.0862 (2) |
| c (Å) | 14.1356 (4) |
| β (°) | 108.685 (3) |
| Absorption coefficient (mm−1) | 21.83 |
| Number of reflections measured variation of h, k, l | h = −12 → 12, k = −13 → 14, l = −16 → 16 |
| Scanning range of θ (°) | 2.6 < θ < 30.2 |
| F (000) | 1376 |
| Independent parameters | 159 |
| Δρmax/Δρmin (e Å−3) | 1.89/−1.05 |
| (Δ/σ)max | < 0.001 |
| R[F2 > 2σ(F2)] = R1 | 0.023 |
| wR(F2) = wR2 | 0.055 |
| S = GooF | 1.07 |
| CCDC | 2384365 |
To highlight these interactions, 2D fingerprint plots37,38 were generated using Crystal Explorer 21.5 software,39 with the CIF file as input. These plots map the frequency and type of intermolecular contacts. Each point on the Hirshfeld isosurface is defined by two key distances:
di: distance from the surface to the nearest nucleus inside the molecule. de: distance from the surface to the nearest nucleus in neighboring molecules.
The normalized contact distance (dnorm) is computed as:
This analysis identifies interaction regions, with red, white, and blue representing short, moderate, and long contacts, respectively. The 2D fingerprint plots quantify the percentage contributions of different non-covalent interactions, offering a detailed understanding of the crystal packing forces in our hybrid material.
The decision to omit the GENECP option was based on two factors. First, its computational cost is significantly higher than LANL2DZ, making it less efficient. Second, while GENECP is often preferred for heavy metals due to its accuracy, the LANL2DZ results showed minimal deviation, providing sufficiently precise data. Given this close agreement, the more computationally demanding GENECP was deemed unnecessary.
Fig. 1(a), generated using DIAMOND 3, presents both the asymmetric unit and the full formula unit, which consists of two diprotonated organic cations (C8H14N2)2+, one [Bi2Br10]4− dimer, and two H2O molecules. The structure exhibits no atomic disorder or structural defects, ensuring high structural stability, as confirmed by the normal ellipsoid volumes in Fig. 1(a). Additionally, no excess electron density or unresolved disorder was detected. A summary of the lattice parameters, unit cell volume, and crystallographic details is provided in Table 1, while Table 1.S lists the atomic positions and their thermal agitation factors (Ueq).
![]() | ||
| Fig. 1 (a) The formula unit of (C8H14N2)2[Bi2Br10]·2H2O and (b) the optimized geometry of the title compound. | ||
Fig. 1(b) illustrates the optimized geometry of the formula unit, highlighting the calculated bond distances. A comparison between theoretical and experimental bond lengths shows that most optimized bond lengths are slightly longer than their experimental counterparts. This discrepancy is expected, as theoretical calculations consider an isolated molecule, whereas experimental measurements reflect a solid-state environment influenced by intermolecular interactions and crystal packing effects. These findings validate the accuracy of our computational approach, confirming that the B3LYP/LanL2DZ level of theory is well-suited for this system.
Fig. 2 offers valuable insights into the crystalline structure. The projection along the b-axis reveals that the inorganic entities are arranged in parallel associations along the c-axis. The anionic components occupy two distinct positions within the unit cell along the c-axis: one at the corners and the other at the center of the faces parallel to the c-axis. Notably, these anionic components are stabilized by water molecules surrounding the polyhedra and the positive charge from the protonated nitrogen atoms of the organic cations.
![]() | ||
| Fig. 2 Projection of the unit cell of (C8H14N2)2[Bi2Br10]·2H2O structure along the crystallographic b-axis. | ||
Fig. 2.S illustrates notable variations in Bi–Br bond lengths, which range from 2.7271(5) Å to 3.0612(5) Å. The Br–Bi–Br bond angles exhibit a broad distribution, spanning 87.274(13)° to 94.257(15)° for cis configurations and 174.713(14)° to 177.625(14)° for trans configurations. The most significant bond length difference of 0.3341 Å is observed between two opposing halogen atoms (Bi–Br1 and Bi1–Br4i). A comprehensive summary of the geometric parameters, including bond angles and distances for the anionic groups, is provided in Table 2.S.
Furthermore, the presence of hydrogen bonding involving the bromine atoms influences the lone electron pair of the Bi atom, causing a shift in electron density toward the hydrogen atom. This shift reduces the electron density around the bridging Br atom, leading to an elongation of the Bi–Br bonds. As a result, the bond length variations and angular deviations contribute to a slight distortion of the BiBr6 octahedral geometry, quantified by an octahedral distortion index of ID (Bi–Br) = 1.7 × 10−6. This distortion originates from a combination of primary deformations, driven by the stereochemical activity of Bi's lone electron pair, and secondary deformations, induced by hydrogen bonding interactions.
The spatial arrangement of atoms or ions within a crystal lattice plays a crucial role in determining a material's physical properties, including mechanical strength and thermal conductivity. Assessing the efficiency of this arrangement can be achieved through both direct and indirect methods. One of the most straightforward measures is crystal density, where a higher density signifies more efficient atomic or ionic packing. In this study, our hybrid material exhibits a remarkably high density of 2.950 g cm−3, indicating a highly efficient packing arrangement.
An alternative approach to evaluating packing efficiency is void analysis, which quantifies the presence of unoccupied spaces or pores within the crystal structure. The extent of these voids directly impacts the material's packing efficiency and influences its macroscopic properties. As part of our Hirshfeld surface analysis, we conducted a crystal void analysis to gain further insight into the packing efficiency. The results of this analysis were in strong agreement with the density measurement, reinforcing the conclusion that the crystal structure of our material is highly efficient.
Conjugation refers to the extended delocalization of π-electrons across alternating single and multiple bonds through overlapping p-orbitals. The studied organic molecule possesses three conjugated forms, leading to extensive electron delocalization that significantly influences its electronic, optical, and chemical properties. Fig. 3.S illustrates the spatial arrangement of the organic molecules within the unit cell, where two diprotonated organic cations align parallel to the (b,c) plane. Key geometric parameters, including bond angles and interatomic distances of the organic groups, are summarized in Table 3.S, providing a detailed structural analysis.
The crystal's structural integrity is governed by an intricate arrangement of molecules interconnected via hydrogen bonding. Within this framework, three distinct bond types are observed: O–H⋯Br, N–H⋯O, and N–H⋯Br, totaling four hydrogen bonds. The cohesive forces and physical characteristics of the crystal hinge upon these bonding interactions. Specifically, the water molecules engage with both organic cations and nitrogen atoms, fostering cohesion between them and facilitating the formation of a three-dimensional lattice. The strength of hydrogen bonds associated with halogen atoms can be delineated into weak and strong categories based on the Brown criteria,47 where bonds with dD–A distances exceeding 3.19 Å are considered weak, and those below are deemed strong. Applying this criterion, bonds such as O1–H1C⋯Br5ii, N2–H2B⋯Br5iii, and N2–H2A⋯Br2 are classified as weak. For the determination of N–H⋯O hydrogen bond strength, the R. H. Blessing criteria48 prescribe a threshold distance of dD–A = 2.7 Å, categorizing the N1–H1⋯O1iv bond as strong. Fig. 3 illustrates these hydrogen bonds within the asymmetric unit and unit cell, with their replication evident through inversion symmetry at the center. A detailed account of the distances and angles of these bonds is presented in Table 2.
![]() | ||
| Fig. 3 The arrangement of hydrogen bonds between water molecules, organic cations, and inorganic [Bi2Br10]4− ions: (a) in the formula unit and (b) in the unit cell. | ||
| D–H···A | D–H (Å) | H⋯A (Å) | D⋯A (Å) | D–H···A (°) |
|---|---|---|---|---|
| a Symmetry codes: (ii) −x, y − 1/2, −z + 3/2; (iii): −x + 1, −y + 1, −z + 2; (iv): −x + 1, y − 1/2, −z + 3/2. | ||||
| O1–H1C⋯Br5ii | 0.85 | 2.61 | 3.389 (4) | 153 |
| N1–H1⋯O1iii | 0.86 | 1.83 | 2.684 (6) | 170 |
| N2–H2A⋯Br2 | 0.89 | 2.62 | 3.494 (4) | 169 |
| N2–H2B⋯Br5iv | 0.89 | 2.53 | 3.301 (4) | 146 |
A comparative analysis of experimental and theoretical FTIR spectra is illustrated in Fig. 4, while Fig. 5 presents the experimental and theoretical Raman spectra of (C8H14N2)2[Bi2Br10]·2H2O. The vibrational mode assignments, along with their corresponding experimental and calculated wavenumbers, are summarized in Table 4.S, demonstrating a strong correlation between theoretical predictions and observed spectral data. The close agreement between experimental and computed frequencies supports the reliability of the proposed spectral assignments. While periodic calculations could provide an even more precise description of the system, their application is often restricted due to substantial computational demands.
The table further highlights vibrational modes for ethyl groups, such as CH2 asymmetric stretching at 3160 cm−1 and CH3 at 3081 cm−1, with theoretical values at 3169 and 3069 cm−1, respectively. Hydrogen bonding interactions, such as the N–H⋯O stretching mode, appear in the IR spectrum between 1845–2694 cm−1, with a theoretical range of 2280–2323 cm−1, crucial for linking organic cations and water molecules to the inorganic framework.
In the lower wavenumber region, bending and twisting modes are observed, including NH2 bending at 1720 cm−1, water bending at 1651 cm−1, and CH2 deformation at 1508 cm−1. These modes provide insight into molecular flexibility and dynamics within the hybrid structure. The presence of multiple vibrational modes across frequency regions underscores the material's complex bonding interactions, aligning with computational predictions and validating its structural properties.
The high-energy absorption at 258 nm is attributed to localized π → π transitions within the organic moiety, specifically within the aromatic rings of the diprotonated (C8H14N2)2+ cations. The absorption feature at 317 nm corresponds to charge transfer transitions within the [Bi2Br10]4− (LMCT, from Br 4p to Bi 6p), while the most critical absorption peak at 348 nm originates from intramolecular charge transfer transitions within the [Bi2Br10]4− inorganic sublattices and the organic cation, specifically, Bi 6p + Br 4p → N 2p + C 2p electronic excitations, a characteristic feature of bismuth-based halide hybrids that is also reflected in the density of states (DOS and PDOS) calculations (Fig. 17), where the conduction band minimum (CBM) is dominated by N 2p + C 2p states and the valence band maximum (VBM) consists primarily of Bi 6p + Br 4p orbitals.
Fig. 7 presents the Tauc plots used to determine the optical bandgap of (C8H14N2)2[Bi2Br10]·2H2O, distinguishing between direct and indirect electronic transitions. The Tauc method49 estimates bandgap energies by plotting (αhν)n versus photon energy (hν), where α is the absorption coefficient and n depends on the transition type: n = 2 for an indirect bandgap (phonon-assisted transitions) and n = 1/2 for a direct bandgap (phonon-independent). For this hybrid material, the (αhν)1/2 plot confirms an indirect bandgap, as the direct transition model fails to fit the absorption data. The extracted bandgap values are 3.086 eV in solution (UV-vis, Fig. 6) and 2.9 eV in the solid state (DRS, Fig. 10), in excellent agreement with theoretical calculations using the zero-DOS method (Fig. 17(a)), which predicts 2.85 eV. The zero-DOS approach identifies the band edges by determining the precise energy values at which the DOS first becomes strictly zero, accounting for numerical precision. The band gap is thus defined as the interval between the last occupied state and the first unoccupied state where the DOS vanishes completely. In contrast, the conventional DOS method typically estimates the band gap from a broadened DOS spectrum, where small non-zero values may appear within the nominal gap due to smearing functions, limited k-point sampling, or numerical noise.
Fig. 8 presents the photoluminescence (PL) spectrum of (C8H14N2)2[Bi2Br10]·2H2O, offering detailed insights into the material's excited-state electronic behavior, band structure, and charge carrier dynamics. PL spectroscopy, as applied here, complements the UV-vis absorption (Fig. 6), diffuse reflectance (Fig. 10), and Tauc analysis, serving as a powerful probe for exciton recombination and defect-related emissions. The PL spectrum displays two distinct emission features: a broad band centered around 350 nm and a sharp emission at 706 nm, with an intermediate shoulder near 645 nm. The broad peak at 350 nm is attributed to radiative relaxation of high-energy excitons within the [Bi2Br10]4− framework, in line with the charge transfer absorption band at 348 nm. The broadening of this emission, alongside its spectral width, indicates phonon-assisted recombination and multiple non-radiative decay pathways, hallmarks of an indirect bandgap system.
The transition from a broad to a sharp emission as the energy decreases reflects a shift from delocalized to more localized recombination processes. The sharp peak at 706 nm, redshifted from the 645 nm shoulder, confirms a significant Stokes shift, supporting an indirect transition where carriers undergo phonon-mediated energy loss prior to recombination. This sharp emission also points to a highly specific and radiatively efficient decay channel with limited energetic disorder, which aligns with the low Urbach energy observed in Fig. 12(a). The Urbach energy analysis confirms a well-ordered crystalline lattice with minimal band tailing and defect-mediated sub-gap states, further supporting the presence of clean, trap-free recombination channels in the solid state.
Additionally, the PL response is excitation-dependent, with the emission wavelength shifting as the excitation energy changes. This behavior indicates the presence of multiple emissive states and relaxation pathways, commonly observed in low-dimensional bismuth-based hybrids. The underlying mechanism is attributed to self-trapped excitons (STEs) formed within the distorted [Bi2Br10]4− subunits. In such systems, the energy landscape contains shallow and deep potential wells created by lattice distortions and local electronic inhomogeneities, allowing excitons to relax into different emissive channels depending on the excitation energy. The indirect nature of the bandgap further supports phonon-assisted recombination, increasing spectral sensitivity to excitation conditions. Additionally, weak energetic disorder and dynamic coupling between the flexible organic cations and the inorganic framework contribute to the excitation-dependent tunability, as reflected in the shift of emission maxima.
The difference between the bandgap values extracted from liquid-phase UV-vis measurements (3.086 eV, Fig. 6) and solid-state DRS (2.9 eV, Fig. 10) is attributed to solvation effects, molecular rearrangements, and enhanced intermolecular interactions in the crystalline phase. Solvent screening in solution typically leads to an overestimation of the bandgap due to weaker exciton binding and limited inter-framework coupling, whereas solid-state measurements better reflect the condensed, polarizable environment. Furthermore, TD-DFT calculations (Fig. 17) yield a bandgap of 2.85 eV, closely matching the DRS-derived value, and confirming the indirect nature of the transition.
Finally, the strong correlation between the observed PL features and optical constants, including the penetration depth (Fig. 13) and extinction coefficient (Fig. 14), confirms the consistency of the material's optical behavior across techniques. The compound's broad absorption, strong Stokes shift, and stable emission make it a promising candidate for near-infrared emitters, broadband photonic absorbers, and hybrid optoelectronic devices where emission tunability and structural robustness are essential.
Fig. 9 presents the CIE 1976 chromaticity diagram, offering a precise assessment of the emission color properties of (C8H14N2)2[Bi2Br10]·2H2O. This diagram quantitatively represents the compound's photoluminescence (PL) emission in the perceptible color space, highlighting its potential for optoelectronic applications, such as LEDs, laser technologies, and displays. The CIE 1976 (u′, v′) color space ensures a perceptually uniform color distribution, making it the preferred standard for accurate evaluation of chromatic coordinates and emission purity. The diagram reveals two distinct (u′, v′) coordinates for two excitation wavelengths: 319 nm and 350 nm, correlating with the primary PL emission features observed in Fig. 8. The emission under 319 nm excitation is characterized by (u′ = 0.311, v′ = 0.371), placing it in the rose region, while under 350 nm excitation, the coordinates shift to (u′ = 0.164, v′ = 0.344), moving toward the blue-cyan region. This shift highlights a variation in emission behavior depending on excitation energy, linked to differences in radiative recombination pathways and electronic states within the band structure. A key feature of the chromaticity data is its correlation with the PL spectrum, particularly the transformation from a broad 645 nm peak to a sharp 706 nm peak, marking the deep-red emission region. The chromaticity coordinates reinforce the indirect bandgap nature of the material, as phonon-assisted recombination mechanisms lead to longer-wavelength emissions with defined spectral characteristics. The excitation-dependent emission suggests multiple relaxation pathways, likely influenced by exciton localization effects in the [Bi2Br10]4− sublattices.
The narrow spread of chromaticity coordinates indicates a low density of defect states, minimizing non-radiative recombination losses, aligning with the low Urbach energy values (Fig. 12(a)). Additionally, the optical penetration depth (Fig. 13) and extinction coefficient variations (Fig. 14) confirm the efficient light absorption and emission within a well-defined spectral range, demonstrating excitation-dependent tunability and reinforcing its suitability for photonic applications, such as tunable LED phosphors, near-infrared emitters, and high-efficiency displays.
The DRS spectrum (Fig. 10(a)) shows a gradual reflectance decrease from the UV to visible region, with a well-defined edge around 400 nm (Fig. 10(b)), indicating strong absorption onset due to interband electronic transitions. Converting reflectance to an absorption spectrum reveals a pronounced band in the UV region, confirming charge-transfer excitations. The derivative analysis clarifies the absorption edge at 400 nm, suggesting that (C8H14N2)2[Bi2Br10]·2H2O behaves as a semiconductor, absorbing UV and near-visible photons.
The optical bandgap extracted via the Tauc plot method (Fig. 11) confirms an indirect transition with Eg ≈ 2.9 eV, in agreement with DOS calculations (Fig. 17), which predict Eg = 2.85 eV. The discrepancy with the liquid-state bandgap (3.086 eV, Fig. 6) highlights the effects of solid-state interactions, where molecular orbital overlap and crystal packing lead to a narrower bandgap compared to the solvated environment. This is corroborated by PL analysis (Fig. 8), where a broad 645 nm emission evolves into a sharp 706 nm peak, confirming an indirect band structure with phonon-assisted recombination. The DOS and PDOS (Fig. 17) reveal that the valence band maximum (VBM) is composed mainly of Br 4p states, with contributions from Bi 6p orbitals, while the conduction band minimum (CBM) consists of N 2p and C 2p states. The HOMO–LUMO transition involves intermolecular charge transfer from the Bi 6p + Br 4p orbitals to N 2p + C 2p orbitals, aligning with the solid-state absorption spectrum (Fig. 10(b)) and liquid-state UV-vis data (Fig. 8). This confirms the complex light absorption process in this system. The experimental absorption spectrum shows several distinct peaks, notably at 249 nm (4.98 eV), 371 nm (3.34 eV), 417 nm (2.97 eV), and 483 nm (2.57 eV), each corresponding to different types of electronic transitions. The 249 nm peak is primarily attributed to localized electronic excitations within the inorganic [Bi2Br10]4− lattice, given its correspondence with the expected energy range of such transitions in bismuth halide frameworks. However, since π–π* transitions in aromatic organic systems also typically occur in the 200–300 nm range, a partial spectral overlap with the organic cation cannot be definitively excluded. Making the 249 nm band arise predominantly from inorganic-centered excitations, with possible minor contributions from π–π* transitions within the organic component, consistent with the hybrid nature of the material and the potential for orbital coupling at higher excitation energies. The 371 nm peak corresponds to a charge transfer within the inorganic framework. The 417 nm peak represents intermolecular charge transfer (ICT), confirming strong coupling between the organic and inorganic components. The 483 nm peak suggests phonon-assisted transitions typical of indirect bandgap semiconductors, supporting the indirect bandgap nature confirmed by Tauc analysis (Fig. 11(a)) and DOS calculations (Fig. 17(a)). Overall, Fig. 10(a) and (b) provide a comprehensive understanding of the solid-state optical behavior of (C8H14N2)2[Bi2Br10]·2H2O, reaffirming its indirect bandgap nature, strong UV absorption, and potential for optoelectronic applications. The integration of reflectance, absorption analysis, derivative spectral features, DOS calculations, and PL studies highlights its suitability for advanced photonic applications with precise bandgap control and efficient absorption.
![]() | ||
| Fig. 11 (a) Direct versus indirect band gap determination via Tauc plot, and (b) Ln(αhv) vs. Ln(hv − Eg) Plot. | ||
Fig. 11 provides a detailed evaluation of the optical bandgap and transition nature in (C8H14N2)2[Bi2Br10]·2H2O, essential for understanding its fundamental electronic properties. Fig. 11(a) presents the Tauc plots for direct and indirect transitions, where a clear linear extrapolation in the Kubelka–Munk plot confirms the indirect nature of the bandgap, yielding an Eg value of 2.9 eV. This result is consistent with the Density of States (DOS) calculations (Fig. 17(a)), which predict Eg = 2.85 eV, further reinforcing the material's indirect band structure. The insets in Fig. 10(a) and (b), showing the derivatives of the reflectance and absorption spectra, provide an initial approximation of the bandgap at ∼3.0 eV, offering an independent confirmation of the absorption edge position before precise Kubelka–Munk's Tauc extrapolation. The Kubelka–Munk function is defined as follows:50,51
According to inter-band absorption theory, the absorption coefficient near the threshold, in relation to incident energy, can be described by the following equation:52,53
Fig. 11(b) presents the Ln(F(R)hν) vs. Ln(hν − Eg) plot, which determines the exponent n in the Kubelka–Munk function, a critical parameter for identifying whether the transition is allowed direct (n = ½), allowed indirect (n = 2), forbidden direct (n = 2/3), or forbidden indirect (n = 3). The best linear fit to the experimental data confirms n ≈ 2, definitively proving that the bandgap is of an indirect allowed nature, aligning perfectly with the Stokes shift and phonon-assisted recombination processes observed in the PL spectrum (Fig. 8). The indirect nature is also evident in the continuous absorption behavior beyond the band edge in Fig. 10(b), particularly in the small peak at 483 nm, which is characteristic of indirect semiconductors where absorption extends beyond the main threshold due to phonon contributions.
The consistency between the DRS-derived bandgap (Fig. 11(a)), DOS calculations (Fig. 17(a)), PL characteristics (Fig. 8), and indirect transition coefficient (Fig. 11(b)) firmly establishes that (C8H14N2)2[Bi2Br10]·2H2O is a semiconductor with an indirect bandgap, a crucial distinction that influences its charge carrier dynamics, exciton lifetime, and optoelectronic efficiency. These findings reinforce the material's potential for photonic applications, where controlled bandgap engineering and minimal defect-induced electronic disorder are required for high-performance device integration.
Fig. 12(a) presents the Urbach energy (Eu), which quantifies band tail states due to structural disorder, defects, or electron–phonon interactions. A low Eu indicates minimal disorder and a sharp absorption edge, as observed in Fig. 10(a) and (b), aligning with the absence of mid-gap states in the DOS spectrum (Fig. 17(a)). This confirms the material's high electronic stability and supports the indirect nature of the bandgap, as indirect semiconductors typically show extended absorption tails due to phonon-assisted transitions. Fig. 12(b) shows the evolution of (α/λ)2 with 1/λ, revealing how the absorption coefficient varies with wavelength. The trend, correlating with the diffuse reflectance data (Fig. 10(a)) and extinction coefficient (Fig. 14), suggests uniform light absorption across energy ranges, indicating high crystallinity and purity. This smooth, continuous increase in (α/λ)2 further validates the calculated bandgap. The combination of low Urbach energy (Fig. 12(a)), defined absorption characteristics (Fig. 12(b)), and band structure analysis (Fig. 11(a) and (b)) confirms that (C8H14N2)2[Bi2Br10]·2H2O has a stable electronic structure with minimal defect-induced recombination losses, making it suitable for optoelectronic applications requiring strong light absorption and controlled charge carrier dynamics.
Fig. 13 illustrates the penetration depth vs. λ, given by the formula:54
The plot shows three distinct regions:
High-energy UV region (λ < 400 nm) – Minimal penetration depth due to strong absorption from high-energy charge transfer transitions, consistent with the 249 nm absorption peak (Fig. 10(b)).
Mid-visible region (400–800 nm) – increased δ, indicating reduced absorption and stronger light penetration, typical of indirect semiconductors.
Near-infrared region (λ > 800 nm) – further increase in δ, confirming weaker absorption and the absence of mid-gap defect states (Fig. 17(a), DOS analysis).
Fig. 14 shows the optical extinction coefficient (K), derived from:55
Four peaks are identified: 253 nm, 379 nm, 420 nm, and 492 nm. The 253 nm peak corresponds to high-energy charge transfer transitions in the inorganic framework, while the 379 nm and 420 nm peaks reflect intramolecular charge transfer (ICT) between the organic and inorganic sublattices. The 492 nm peak aligns with the continuous absorption tail, confirming phonon-assisted transitions characteristic of indirect semiconductors. The correlation between penetration depth, extinction coefficient, and absorption spectrum establishes (C8H14N2)2[Bi2Br10]·2H2O as a material with well-defined absorption, minimal optical losses, and efficient charge transfer, making it ideal for photonic and optoelectronic applications requiring precise light-matter interaction control.
Fig. 15 illustrates the variation of the refractive index (n) as a function of wavelength across the 200–1100 nm range. Contrary to the classical dispersion behavior typically observed in simple inorganic materials, where n decreases monotonically with increasing wavelength, the present compound reveals a more complex and region-dependent evolution:
In the ultraviolet region (λ < 400 nm), n remains relatively low (∼1.5–2.0), which can be attributed to the proximity of strong absorption edges and the influence of anomalous dispersion effects near the band edge.
Across the visible region (400–800 nm), n increases progressively and more markedly, reflecting enhanced electronic polarization processes and charge-transfer interactions within the hybrid inorganic–organic network.
In the near-infrared region (λ > 800 nm), the rise of n continues but at a diminished rate, with values approaching and slightly exceeding 4.0. This saturation-like behavior is indicative of a high dielectric response combined with minimal absorption losses, underscoring the stability of the optical response at longer wavelengths.
Such non-classical dispersion characteristics are emblematic of bismuth-based organic–inorganic hybrid materials, where the unusually strong polarizability of Bi3+ cations, the stereochemically active 6s2 lone pairs, and the cooperative electronic coupling between organic and inorganic substructures contribute synergistically to complex optical responses. Overall, this distinctive dispersion profile confirms both the strong optical anisotropy and dense electronic structure of the compound, in line with the high polarizability and the indirect band gap nature previously established by DOS and optical analyses.
The optical conductivity σopt can be calculated using the following expression where c is the speed of light in free space:
Fig. 16 presents the optical conductivity (σopt), a key parameter reflecting the material's ability to absorb and transfer photogenerated charge carriers, with four key peaks at 275 nm, 398 nm, 447 nm, and 784 nm. The 275 nm peak corresponds to high-energy charge transfer excitations, while the 398 nm and 447 nm peaks align with ICT processes between the inorganic and organic components, reinforcing the hybrid nature of the electronic transitions. The 784 nm peak, appearing in the near-IR region, suggests long-wavelength residual absorption, likely due to defect-mediated transitions or phonon interactions, a feature common in indirect bandgap semiconductors.
The consistency between the refractive index trends, optical conductivity peaks, and band structure analysis (Fig. 16 and 17(a)) underscores the material's well-defined electronic transitions, low defect density, and high potential for optoelectronic applications requiring precise control over refractive and conductive properties.
![]() | ||
| Fig. 17 (a) Total density of state (DOS), and (b) partial density of state (PDOS) spectra of the compound spectrum and the frontier molecular orbitals of the compound. | ||
The compound (C8H14N2)2[Bi2Br10]·2H2O exhibits markedly distinct optical properties compared to the recently reported antimony-based hybrid (C8H14N2)2[Sb2I10]·2H2O,57 despite both adopting dimeric [M2X10]4− structural motifs. While both structures feature edge-sharing halometalate dimers stabilized by the same organic cation, the substitution of Sb3+ with Bi3+ and the use of Br− instead of I− lead to significantly different electronic environments and resulting optical behavior. Our bismuth-based compound demonstrates a much narrower band gap (2.9 eV) compared to the wide-gap antimony analog (5.27 eV), shifting absorption from the ultraviolet into the visible region. This enhancement is attributed to the stronger spin–orbit coupling (SOC) effects associated with the heavier Bi3+ center, which effectively reduce the band gap by stabilizing the conduction band minimum. Furthermore, the use of bromide (a harder halide than iodide) contributes to increased orbital overlap and a more delocalized electronic structure. These differences translate into more favorable optical characteristics for visible-light-driven applications in our compound, including enhanced light-harvesting capability and emission behavior, as confirmed by PL and TD-DFT analyses.
The RDG scatter plot (Fig. 7.S(b)) quantifies these interactions: blue peaks for attractive hydrogen bonds, green for van der Waals, and red for steric repulsions. The sharp blue peaks confirm dominant hydrogen bonding between the organic cation and the anionic framework, while green peaks show van der Waals interactions, and red peaks reflect steric repulsions.
These visualizations reveal the synergy of hydrogen bonding, van der Waals forces, and steric effects in stabilizing the hybrid framework. Understanding these interactions is crucial for designing materials with optimized properties for catalysis, sensing, or optoelectronics.
Fig. 9.S displays the Electron Localization Function (ELF) mapped across the same three planes, highlighting regions of high electron localization (ELF ≈ 1.0, red) corresponding to lone pairs and covalent bonds, and regions of delocalization (ELF ≈ 0.0, blue) indicative of weak interactions. In the (XY) plane, symmetric electron localization is observed around the Bi–Br bonds, with pronounced lone pair density on Bi3+ centers. The (XZ) and (YZ) planes further emphasize electron-rich regions associated with hydrogen bonding between the halides and coordinated water molecules. These features collectively confirm a mixed bonding character dominated by localized covalent and noncovalent interactions.
Fig. 10.S complements this with Localized Orbital Locator (LOL) maps in the same spatial planes. The 3D surface plots reveal peaks in electron pairing near key bonding sites, while the 2D contour maps highlight anisotropic orbital localization, transitioning from low (blue) to high (red) values. The (XY) plane demonstrates strong localization along covalent Bi–Br bonds, whereas the (XZ) and (YZ) views reinforce the directional nature of electron density across the inorganic dimer and organic interface. Together, ELF and LOL analyses provide consistent and detailed insight into the electronic distribution and bonding topology of the hybrid, underscoring the structural and electronic coherence of the system.
Fig. 12.S illustrates the 2D fingerprint plots, offering a comprehensive representation of total contacts contributing to the Hirshfeld surface. The plot spans a distance range of 0.6 to 2.8 Å, with de and di values plotted along respective axes. The analysis reveals that the dominant interaction is H⋯Br (57.6% of the total surface contacts), with H⋯H contacts contributing 22.5%. The prevalence of hydrogen bonding interactions indicates their importance in the material's stability, impacting its mechanical, thermal, optical, and conductive properties.
Additionally, an assessment of crystal voids was performed to gain insights into packing efficiency and the physicochemical properties of the material. Crystal voids represent unoccupied regions in the solid structure, affecting solubility, density, mechanical strength, and thermal conductivity. These voids also facilitate molecular vibrations, influencing the compound's behavior. Using CrystalExplorer 21.5, the void volume was determined to be 101.58 Å3, representing 5.9% of the unit cell volume, as shown in Fig. 13.S. This suggests a high packing density, contributing to enhanced mechanical robustness and structural integrity. The reduced void space minimizes defect sites, preventing crack formation and propagation, while also improving thermal conductivity by allowing better phonon transport. These properties indicate that the hybrid material is a promising candidate for applications requiring both high optical quality and mechanical stability.
The molecular architecture of (C8H14N2)2[Bi2Br10]·2H2O exemplifies an acceptor–donor–acceptor (A–D–A) system, where the organic cations function as electron-rich donors and the bismuth bromide dimeric anion serves as a robust electron acceptor. This configuration facilitates effective charge transfer between the organic and inorganic sublattices, as supported by the observed optical transitions and electronic structure features. The properties uncovered, such as the indirect band gap, strong refractive response, and prominent charge transfer bands, are indicative of generalizable trends among low-dimensional bismuth-based hybrids. Substitution or structural modification of the donor units (e.g., using more π-conjugated or electron-rich cations) or the inorganic acceptor (via halide tuning) offers a promising route to modulate the electronic and photophysical properties systematically. These findings highlight the compound's relevance within donor–acceptor materials design and suggest its potential utility as a model system for developing next-generation hybrid optoelectronic materials.
CCDC 2384365 contains the supplementary crystallographic data for this paper.58
Supplementary information is available. See DOI: https://doi.org/10.1039/d5ra06614j.
| This journal is © The Royal Society of Chemistry 2025 |