DOI:
10.1039/D5RA05828G
(Paper)
RSC Adv., 2025,
15, 34939-34947
Comprehensive study of organic–inorganic hybrid [N(C2H5)4]2CdBr4: crystal structure, phase transitions, and structural geometry
Received
9th August 2025
, Accepted 10th September 2025
First published on 23rd September 2025
Abstract
The compound [N(C2H5)4]2CdBr4, an organic–inorganic hybrid compound, has garnered attention for its potential applications across various fields. In this study, single crystals of [N(C2H5)4]2CdBr4 were grown, and two distinct phase transition temperatures were identified: approximately 232 K (TC1) and 476 K (TC2). Single-crystal X-ray diffraction analysis at 300 K revealed a tetragonal structure belonging to the space group P
21m. The thermal properties of the material were also briefly examined. Interestingly, while the 1H NMR chemical shifts exhibited minimal variation around TC1, the 13C NMR spectra showed a noticeable change in the number of peaks, suggesting a structural phase transition. These observations indicate that the crystal retains a tetragonal structure above TC1 but transitions to a phase with lower symmetry below this temperature. Furthermore, 113Cd NMR chemical shifts showed a significantly more pronounced change near TC1 compared to the shifts observed for 1H and 13C. This behavior reflects alterations in the local electronic environment around the Cd2+ ions, implying reconfiguration of the surrounding atomic framework. Additionally, the spin–lattice relaxation times for 1H and 13C, which are related to molecular motion, were nearly identical for both CH2 and CH3 groups in [N(C2H5)4] cations. The phase transition observed at TC1 is thus attributed to substantial structural rearrangements, particularly involving changes in atomic coordinates and the rotation of the CdBr42− tetrahedra.
1. Introduction
Hybrid organic–inorganic metal halides have been extensively studied due to their promising applications in photovoltaic and optoelectronic devices. The broad range of applications and the rapid development of these materials have attracted significant attention from researchers.1–12 In these compounds, the optical properties and structural flexibility are primarily governed by the organic cations, whereas the inorganic anions mainly determine their thermal and mechanical behavior. Their unique ability to combine the advantages of both organic and inorganic components makes them highly versatile for designing advanced functional materials.13 Among these, the three-dimensional (3D) perovskite CH3NH3PbI3 has recently received considerable interest due to its exceptional photovoltaic performance. However, its practical application is limited by its sensitivity to humidity and the toxicity associated with lead.14–18 As a result, significant efforts have been directed toward the development of zero-dimensional (0D) hybrid organic–inorganic metal halides as more stable and less toxic alternatives. 0D hybrid metal halides with the general formula A2MeX4 have emerged as excellent candidates for phase-transition materials. In particular, tetraethylammonium-based compounds, [N(C2H5)4]2MeX4 (where Me = Mn, Co, Cu, Zn, Cd and X = Cl, Br), belong to a large family of A2MeX4 crystals known to exhibit complex phase transitions influenced by both the metal (Me) and halide (X) ions.19–53 These transitions are highly sensitive to external conditions due to hydrogen bonding between the molecular cations and the metal halide anions. Each [N(C2H5)4]+ cation resides in a cavity formed by surrounding MeX4 anions, and many of these crystals undergo phase transitions involving rearrangements of both the MeX4 units and the organic groups. The tetraethylammonium group itself can adopt two distinct stable conformations, with its ethyl chains offering greater configurational freedom compared to methyl-substituted analogs. While the nitrogen and terminal carbon atoms of the [N(C2H5)4]+ ion occupy well-defined positions, the central carbon atom can adopt two orientations, suggesting a disordered ionic configuration. More recently, attention has also turned to compounds that incorporate either mixed inorganic or organic components, such as [N(C2H5)4]2CoCl2Br2,54–60 which contains mixed halides, and [N(CH3)4][N(C2H5)4]MeX4,61–64 which features different organic cations. These systems offer further opportunities to explore structure–property relationships and design novel hybrid materials.
The [N(C2H5)4]2CdBr4 compounds, the case where Me = Cd and X = Br, i.e., has been reported in the literature only with respect to a few aspects. The transition temperature at 230 K was first reported by Kahrizi et al.23 through capacitance dilatometry measurements. Vlokh et al.25 investigated the optical birefringence of this single crystal and revealed a second-order phase transition at 118 K. Additionally, Sveleba et al.29 reported a second-order phase transition at 311 K based on birefringence and dielectric experiments. Furthermore, changes in lattice parameters observed via X-ray diffraction suggested first-order phase transitions at 174 and 226 K, while an anomalous change was detected at 293 K.44 In addition, Dekola et al.45 demonstrated, through temperature-dependent heat capacity measurements, the occurrence of a first-order phase transition at 226.5 K. However, as shown in Table 1, these results are inconsistent with one another, and few detailed studies have been published on this compound to date. Although [N(C2H5)4]2CdBr4 crystals have been relatively understudied among A2MeX4-type compounds, they have recently attracted attention due to their promising potential for practical applications.
Table 1 The phase transition temperatures for previously reported [N(C2H5)4]2CdBr4 crystals according to the various experimental methods
In the present work, single crystals of [N(C2H5)4]2CdBr4 were grown, and their phase transition behavior was investigated using differential scanning calorimetry (DSC) and powder X-ray diffraction (PXRD). The thermodynamic properties of the compound were briefly examined using thermogravimetric analysis (TGA) and differential thermal analysis (DTA). In addition, the crystal structure and lattice parameters at 300 K were determined by single-crystal X-ray diffraction (SCXRD) experiments. To examine the roles of [N(C2H5)4]+ cations and CdBr42− anions near the phase transition temperature, temperature-dependent chemical shifts in the 1H, 13C, and 113Cd magic-angle spinning (MAS) nuclear magnetic resonance (NMR) spectra were analyzed. Moreover, spin–lattice relaxation times (T1ρ) obtained from the NMR measurements were used to investigate molecular motion rates.
2. Experimental
2.1. Crystal growth
High-quality single crystals of [N(C2H5)4]2CdBr4 were successfully grown using a slow evaporation method from supersaturated aqueous solutions. The crystals were obtained by dissolving tetraethylammonium bromide (N(C2H5)4Br, Sigma-Aldrich, 99%) and cadmium bromide (CdBr2, Sigma-Aldrich, 97%) in distilled water at a molar ratio of 2
:
1. The solution was thoroughly stirred and gently heated to achieve saturation, followed by gradual evaporation over several weeks in a constant-temperature bath maintained at 300 K. Several colorless and transparent single crystals with dimensions of approximately 5 × 4 × 3 mm were obtained. The resulting single crystals were carefully stored in a desiccator to prevent degradation caused by moisture exposure.
2.2. Characterization
Fourier-transform infrared (FT-IR) spectra were recorded in the range of 4000–500 cm−1 using a PerkinElmer L1600300 spectrometer (Waltham, MA, USA), with samples prepared as compressed KBr pellets. Differential scanning calorimetry (DSC) analyses were performed using a TA Instruments DSC system (Model 25) over the temperature range of 200–550 K, employing a heating rate of 10 K min−1 under a dry nitrogen atmosphere. The morphology of the crystals and their thermal evolution were monitored using a Carl Zeiss optical polarizing microscope equipped with a Linkam THM-600 heating stage. Thermogravimetric analysis (TGA) was carried out from 300 to 873 K at a heating rate of 10 K min−1 under a continuous flow of nitrogen gas.
The crystal structure and lattice parameters at 300 K were determined by single-crystal X-ray diffraction (SCXRD) analysis, conducted at Sookmyung Women's University using the Sookmyung Research Facilities. SCXRD data were obtained using a Bruker SMART CCD diffractometer equipped with a graphite-monochromated Mo-Kα radiation source. Data acquisition and integration were carried out using the SMART APEX3 and SAINT software suites, and absorption corrections were applied via the multiscan method implemented in SADABS.65 Additionally, powder X-ray diffraction (PXRD) analysis was conducted using an Expert Pro-MPD X-ray diffractometer (PANalytical) fitted with Cu-Kα radiation and operated at 40 kV and 40 mA at the Korea Basic Science Institute (KBSI), Metropolitan Seoul Center. This instrumentation enabled the examination of the crystalline structure and phase composition of the samples. Furthermore, high temperature XRD analysis was performed utilizing an Anton Paar HTK 1200N stage coupled with the aforementioned XRD system, allowing for the investigation of structural changes in the samples at elevated temperatures.
Solid-state 1H and 13C NMR spectra of [N(C2H5)4]2CdBr4 were acquired using Bruker NMR spectrometers operating at 400 MHz and 500 MHz (Germany) at the Metropolitan Seoul Center of the KBSI and the NCIRF at Seoul National University, respectively. The 1H magic angle spinning (MAS) NMR experiments were conducted at a Larmor frequency of 400.13 MHz, while the 13C MAS spectra were recorded at 100.61 MHz and 125.77 MHz. Chemical shifts for both nuclei were referenced to tetramethylsilane (TMS). To minimize spinning sidebands, the MAS rate was set to 5 kHz. One-dimensional 1H and 13C spectra were collected with delay times of 5 s and 10 s, respectively. Spin–lattice relaxation times in the rotating frame (T1ρ) were measured using a π/2 pulse followed by a spin-lock pulse of varying duration t. T1ρ values for 1H and 13C were determined with variable delay times ranging from 10 ms to 15 s. Additionally, 113Cd MAS NMR chemical shifts were recorded at a Larmor frequency of 88.75 MHz, with CdCl2O8·6H2O serving as the reference compound. One-dimensional 113Cd spectra were obtained using the one-pulse method with a delay time of 100 s. Temperature-dependent NMR measurements were carried out over the range of 180–430 K, with precise temperature control maintained by regulating the nitrogen gas flow and heating current.
3. Results and discussion
3.1. Fourier-transform infrared spectroscopy
The information related to FT-IR is presented in SI S1.
3.2. Phase transition temperature and thermodynamic properties
To determine the phase transition temperatures of [N(C2H5)4]2CdBr4, single crystals were finely ground into powder using a mortar. Differential scanning calorimetry (DSC) measurements were conducted on approximately 8.5 mg of the sample at controlled heating and cooling rates of 10 K min−1. As shown in Fig. 1, two endothermic peaks appeared during heating: a subtle peak near 232 K and a prominent peak around 476 K, corresponding to enthalpy changes of 246 J mol−1 and 15.15 kJ mol−1, respectively. Upon cooling, a strong exothermic peak was observed at approximately 424 K, with an associated enthalpy change of 18.17 kJ mol−1. These thermal transitions were found to be irreversible.
 |
| Fig. 1 Differential scanning calorimetry curves of [N(C2H5)4]2CdBr4 measured at a heating and cooling rate of 10 K min−1 (inset: single crystal images at (a) 300 K, (b) 513 K, and (c) 583 K). | |
Complementary observations using optical polarizing microscopy revealed no significant morphological changes in the single crystals between 300 K and 513 K (see (a) and (b) within Fig. 1). At temperatures above this range, the initially colorless and transparent crystal became opaque (see (c) within Fig. 1), although no melting was observed up to approximately 600 K. This interpretation is further supported by subsequent powder X-ray diffraction (PXRD) analysis.
To investigate the thermal properties of [N(C2H5)4]2CdBr4, thermogravimetric analysis (TGA) and differential thermal analysis (DTA) were carried out. As with the DSC measurements, the sample was finely ground into powder, and approximately 13.6 mg was used for analysis. The measurements were performed over the temperature range of 300–873 K at a heating rate of 10 K min−1, consistent with the DSC conditions. As illustrated in Fig. 2, the TGA results indicated that the crystal remained thermally stable up to approximately 560 K, showing only a slight weight loss of about 2%. Partial thermal decomposition began at this point. Above the decomposition temperature (Td) of 560 K, a significant weight loss was observed. The DTA curve exhibited two endothermic peaks, one near 476 K and another around 581 K. The peak at 476 K aligns well with that observed in the DSC results. In contrast, the peak at 581 K does not correspond to melting, as confirmed by polarizing microscopy, but rather indicates the onset of substantial sample decomposition. An inflection point observed at 662 K is attributed to the release of 2 equivalents of [N(C2H5)4]Br, accounting for approximately 60% of the initial molecular weight, suggesting major thermal degradation. At temperatures exceeding 854 K, the sample showed nearly complete decomposition, with a total weight loss approaching 100%. These thermal analyses offer valuable insights into the compound's thermal stability, phase transition behavior, and decomposition profile.
 |
| Fig. 2 Thermogravimetry and differential thermal analysis curves of [N(C2H5)4]2CdBr4 measured at a heating rate 10 K min−1. Td is the decomposition temperature. | |
3.3. X-ray diffraction experiment on single-crystal
Single-crystal X-ray diffraction (SCXRD) analysis was carried out on [N(C2H5)4]2CdBr4 at 300 K. The compound crystallizes in the tetragonal system with space group P
21m, and the lattice parameters were determined to be a = b = 13.4605(3) Å, c = 14.4058(4) Å, with α = β = γ = 90°. The number of formula units per unit cell (Z) is 4, as summarized in SI S2. Fig. 3(a) and (b) depict the tetragonal crystal structure and thermal ellipsoids of each atom in [N(C2H5)4]2CdBr4 at 300 K. Within the unit cell, two crystallographically distinct tetraethylammonium cations, denoted as [N(1)(C2H5)4] and [N(2)(C2H5)4], are present. The organic–inorganic framework consists of these cations and CdBr42− anions. A summary of selected bond-lengths and angles is provided in SI S3. The Cd–Br bond-lengths range from 2.582 Å to 2.587 Å, and the Br–Cd–Br bond angles vary between 107.69° and 112.58°, indicating a distorted tetrahedral coordination geometry around the Cd2+ ion. As also shown in SI S3, the distances from N(1) to its four surrounding carbon atoms range from 1.46 Å to 1.49 Å, while all four N(2)–C distances are uniformly 1.50 Å. This suggests that the local symmetry around N(2) is higher than that around N(1). The crystallographic information file (CIF) corresponding to the SCXRD data collected at 300 K is available in SI S4. Additionally, the complete crystallographic dataset has been deposited with the Cambridge Crystallographic Data Centre (CCDC 2477720). And, the morphology of the colorless and transparent single crystal grown in this study is shown in Fig. 3.
 |
| Fig. 3 (a) Tetragonal structure and (b) thermal ellipsoids for two type of [N(C2H5)4] cations and CdBr4 anion of [N(C2H5)4]2CdBr4 crystal at 300 K (inset: morphology of a colorless and transparent single crystal at 300 K) (CCDC no. 2477720). | |
A simulated powder X-ray diffraction (PXRD) pattern generated from the CIF file closely agrees with the experimental PXRD pattern obtained from powdered [N(C2H5)4]2CdBr4 samples, as shown in Fig. 4. Notably, two prominent diffraction peaks were observed at 11.09° and 13.15°, corresponding to the (111) and (200) plane, respectively. Peak indexing was carried out using the Mercury software. Powder X-ray diffraction (PXRD) measurements were conducted on crushed single crystals of [N(C2H5)4]2CdBr4 over a 2θ range of 8–50° at temperatures above 300 K, and the results are presented in Fig. 5. The diffraction patterns recorded below 450 K remained nearly identical, indicating that the crystal structure is stable within this temperature range. In contrast, a clear change in the PXRD pattern was observed at 500 K, which corresponds well with the endothermic peak near 476 K identified in the DSC analysis. Furthermore, the presence of well-defined diffraction peaks at 500 K suggests the formation of a new crystalline phase rather than melting. The PXRD pattern obtained at 300 K after cooling the sample from 500 K (denoted as 300 K (A)) showed a significant difference compared to the original 300 K pattern. This indicates that the structural change is irreversible, in agreement with the DSC results. Based on the combined evidence from DSC and PXRD analyses, the phase transition temperatures of [N(C2H5)4]2CdBr4 are estimated to be approximately 232 K (TC1) and 476 K (TC2).
 |
| Fig. 4 Powder X-ray diffraction (PXRD) patterns with temperature variation and the simulated XRD pattern of [N(C2H5)4]2CdBr4. The 300 K (A) pattern represents the PXRD pattern measured after cooling the single crystal from 500 K to 300 K. | |
 |
| Fig. 5 1H NMR spectra for CH2 and CH3 in [N(C2H5)4]2CdBr4 with increasing temperature (inset: recovery traces measured with the delay time of 10 ms–15 s). | |
3.4. 1H MAS NMR chemical shifts and spin–lattice relaxation times
The 1H MAS NMR spectra of [N(C2H5)4]2CdBr4 single crystals were recorded at a Larmor frequency of 400.13 MHz under a 9.4 T magnetic field to investigate structural changes near TC1. Tetramethylsilane (TMS) was used as an external reference for accurate calibration of chemical shifts. The spectra, acquired at a spinning rate of 5 kHz, are presented in Fig. 5. Peaks marked with asterisks (*) correspond to spinning sidebands, located approximately ±12.5 ppm from the central peak, consistent with the 5 kHz MAS rate. Below TC1, the spectrum exhibits a single resonance, whereas above TC1, two partially overlapping signals become apparent. At 300 K, signals at 1.04 ppm and 2.91 ppm are assigned to the protons of the CH3 and CH2 groups, respectively. The signal at 1.04 ppm shows significantly higher intensity than the one at 2.91 ppm, consistent with the relative number of hydrogen atoms in the CH3 and CH2 groups. Overall, the 1H chemical shifts exhibit minimal variation with temperature, indicating weak temperature dependence.
At 300 K, the spin–lattice relaxation time T1ρ value for 1H in [N(C2H5)4]2CdBr4 was determined, and the recovery trace of nuclear magnetization can be well described by a single exponential function, as expressed by the following equation:66,67
|
P(t) = P(0) exp(−t/T1ρ)
| (1) |
In this context,
P(
t) denotes the magnetization measured after a spin-lock period of duration
t, whereas
P(0) represents the initial magnetization prior to spin-locking. The
1H MAS NMR spectra were acquired at 300 K by varying the spin-lock delay time. The resulting signals, obtained over a delay range of 10 ms to 15 s, are shown in the inset of
Fig. 5, and the decay of magnetization follows the exponential behavior described by
eqn (1). From the slope of the magnetization decay curves, the spin–lattice relaxation time in the rotating frame (
T1ρ) was determined. The
T1ρ values for the CH
3 and CH
2 groups were found to be identical, each measuring 2.68 s.
3.5. 13C MAS NMR chemical shifts and spin–lattice relaxation times
The 13C MAS NMR spectra of [N(C2H5)4]2CdBr4 were recorded at resonance frequencies of 100.61 MHz and 125.77 MHz under magnetic fields of 9.4 T and 11.7 T, respectively. Tetramethylsilane (TMS) was used as the external reference to ensure accurate chemical shift calibration. At 300 K and a spinning speed of 10 kHz, two distinct signals were observed at 54.24 ppm and 10.30 ppm, corresponding to the CH2 and CH3 carbon environments, respectively. The temperature dependence of the 13C chemical shifts was examined at a reduced spinning rate of 5 kHz for both Larmor frequencies, as presented in Fig. 6. The square and triangle symbols in the plot represent data obtained at 100.61 MHz and 125.77 MHz, respectively, and show good agreement. Due to instrumental limitations, low- and high-temperature spectra were not acquired at 10 kHz, as such conditions could impose excessive mechanical stress on the MAS NMR probe. Overall, the 13C chemical shifts exhibited minimal variation with increasing temperature. However, below the TC1 transition temperature, the spectra for both CH3 and CH2 groups displayed splitting into multiple peaks. As the temperature increased and approached TC1 (e.g., at 238 K), the split signals began to merge; above TC1, only one or two signals were observed, as illustrated in the inset of Fig. 6. The peak labeled “x” near 54 ppm is a spinning sideband associated with the CH2 group. At 218 K, well below TC1, the CH3 signal is split into several distinct peaks, indicating reduced symmetry. This spectral behavior suggests that the local structural environment around the 13C nuclei is less symmetric at lower temperatures and becomes more symmetric above TC1.
 |
| Fig. 6 13C NMR chemical shifts for CH2 and CH3 in [N(C2H5)4]2CdBr4 measured at Larmor frequency of 100.61 and 125.77 MHz with increasing temperature (inset: 13C NMR spectra at 218, 238, 253, and 333 K). | |
The spin–lattice relaxation time in the rotating frame (T1ρ) for 13C nuclei in [N(C2H5)4]2CdBr4 at 300 K was determined using eqn (1), following a method analogous to that used for the 1H T1ρ measurements. The 13C MAS NMR spectra were analyzed by tracking changes in signal intensity as a function of delay time. Spectral data collected over a delay range of 10 ms to 15 s are shown in Fig. 7. The T1ρ values were extracted from the slopes of the magnetization recovery curves. As a result, the T1ρ values for the CH2 and CH3 carbon atoms were found to be very similar, measured at 2.13 s and 2.14 s, respectively.
 |
| Fig. 7 13C NMR recovery traces in [N(C2H5)4]2CdBr4 measured with the delay time of 10 ms–15 s at 300 K. | |
3.6. 113Cd MAS NMR chemical shifts
The 113Cd MAS NMR experiments were performed across a temperature range encompassing the phase transition temperature TC1, in order to gain insight into the coordination environment of the Cd2+ ion within the CdBr42− anion. The temperature-dependent 113Cd chemical shifts, measured at a Larmor frequency of 88.75 MHz, are presented in Fig. 8. The chemical shift and line width in the 113Cd NMR spectrum as a function of temperature are shown in detail in Fig. 9. The 113Cd resonance shifts toward lower (more negative) values with increasing temperature, exhibiting an abrupt discontinuity near TC1. Specifically, the chemical shift changes from approximately 392 ppm at 180 K below TC1 to around 362 ppm at 380 K above this transition, suggesting a structural modification in the local environment surrounding the Cd2+ ion. Similarly, the line width follows a temperature-dependent trend analogous to that of the chemical shift. Notably, a pronounced narrowing of the line width occurs near TC1, decreasing sharply from roughly 15 ppm below TC1 to approximately 1.3 ppm at 380 K. The narrowing of the 113Cd line widths with increasing temperature suggests an enhanced mobility of the 113Cd nuclei. This significant reduction implies enhanced dynamic behavior of the Cd2+ ion, which is tetrahedrally coordinated by four Br− ions, at elevated temperatures.
 |
| Fig. 8 113Cd NMR spectra in [N(C2H5)4]2CdBr4 with increasing temperature. | |
 |
| Fig. 9 113Cd NMR chemical shifts and line widths in [N(C2H5)4]2CdBr4 near the phase transition temperature TC1. | |
4. Conclusions
Colorless and transparent single crystals of [N(C2H5)4]2CdBr4 were grown via an aqueous solution method. Differential scanning calorimetry and related analyses revealed two phase transition temperatures at approximately 232 K (TC1) and 476 K (TC2). The compound crystallizes in a tetragonal system with space group P
21m, and demonstrates thermal stability up to approximately 560 K. On the other hand, although the 1H MAS NMR chemical shifts showed minimal variation near TC1, the 13C MAS NMR spectra exhibited notable changes in peak multiplicity across this temperature, indicating a lowering of symmetry below TC1. This suggests that while the crystal maintains tetragonal symmetry above TC1, it transitions to a lower-symmetry phase at lower temperatures. Furthermore, 113Cd MAS NMR spectra revealed more pronounced changes near TC1 compared to those of 1H and 13C. The change in chemical shifts with temperature indicates a change in local electronic configuration at the 113Cd sites, which in turn means that the neighboring atoms change their configurations. The line width behavior is consistent with increased rotational freedom of the CdBr42− tetrahedra upon phase transition; the changing line widths are attributed to various internal molecular motions. Additionally, spin–lattice relaxation times in the rotating frame (T1ρ) for both CH3 and CH2 groups were found to be nearly identical: 2.68 s for 1H and 2.14 s for 13C. This similarity implies comparable molecular mobility for both functional groups across the transition. The phase transition in the [N(C2H5)4]2CdBr4 crystal occurs at TC1 = 232 K and is accompanied by distinct structural changes, primarily due to variations in atomic coordinates and the rotation of the CdBr42− tetrahedra. A detailed investigation of the structural and physical properties of this organic–inorganic hybrid compound will demonstrate its potential for various technological applications in the future.
Author contributions
A. R. Lim performed crystal growth and wrote the manuscript. S. H. Kim and Y.-J. Ko measured NMR experiments, and D. Shin measured X-ray experiments.
Conflicts of interest
There are no conflicts to declare.
Data availability
CCDC 2477720 contains the supplementary crystallographic data for this paper.68
All relevant data are within the manuscript and the SI. See DOI: https://doi.org/10.1039/d5ra05828g.
Acknowledgements
This research was supported by the Regional Innovation System & Education (RISE) program through the Jeonbuk RISE Center, funded by the Ministry of Education (MOE) and the Jeonbuk State, Republic of Korea (2025-RISE-13-JJU).
References
- Y.-F. Gao, T. Zhang, W.-Y. Zhang, Q. Ye and D.-W. Fu, Great advance in high Tc for hybrid photoelectric-switch bulk/film coupled with dielectric and blue-white light, J. Mater. Chem. C, 2019, 7, 9840 RSC.
- S. K. Abdel-Aal, A. S. Abdel-Rahman, W. M. Gamal, M. Abdel-Kader, H. S. Ayoub, A. F. El-Sherif, M. F. Kandeel, S. Bozhko, E. Yakimov and E. B. Yakimov, Crystal structure, vibrational spectroscopy and optical properties of a one-dimensional organic–inorganic hybrid perovskite of [NH3CH2CH(NH3)CH2]BiCl5, Acta Crystallogr., Sect. B, 2019, 75, 880 CrossRef CAS.
- N. Mahfoudh, K. Karoui, F. Jomni and A. B. Rhaiem, Structural phase transition, thermal analysis, and spectroscopic studies in an organic–inorganic hybrid crystal: [(CH3)2NH2]2ZnBr4, Appl. Organomet. Chem., 2020, 34, e5656 CrossRef CAS.
- Y. Xie, Y. Ai, Y.-L. Zeng, W.-H. He, X.-Q. Huang, D.-W. Fu, J.-X. Gao, X.-G. Chen and Y.-Y. Tang, The Soft Molecular Polycrystalline Ferroelectric Realized by the Fluorination Effect, J. Am. Chem. Soc., 2020, 142, 12486 CrossRef CAS PubMed.
- D.-W. Fu, J.-X. Gao, W.-H. He, X.-Q. Huang, Y.-H. Liu and Y. Al, High-Tc Enantiomeric Ferroelectrics Based on Homochiral Dabco-derivatives (Dabco =1,4-Diazabicyclo[2.2.2]octane), Angew. Chem., Int. Ed., 2020, 59, 17477 CrossRef CAS.
- C. Su, M. Lun, Y. Chen, Y. Zhou, Z. Zhang, M. Chen, P. Huang, D. Fu and Y. Zhang, Hybrid Optical-Electrical Perovskite Can Be a Ferroelastic Semiconductor, CCS Chem., 2021, 4, 2009 CrossRef.
- S. K. Abdel-Aal, M. F. Kandeel, A. F. El-Sherif and A. S. Abdel-Rahman, Synthesis, Characterization, and Optical Properties of New Organic–Inorganic Hybrid Perovskites [(NH3)2(CH2)3]CuCl4 and [(NH3)2(CH2)4]CuCl2Br2, Phys. Status Solidi A, 2021, 218, 2100036 CrossRef.
- S. K. Abdel-Aal and A. Ouasri, Crystal structure, Hirshfeld surfaces and vibrational studies of tetrachlorocobaltate hybrid perovskite salts NH3(CH2)nNH3CoCl4 (n = 4, 9), J. Mol. Struct., 2022, 1251, 131997 CrossRef.
- Y.-L. Zeng, X.-Q. Huang, C.-R. Huang, H. Zhang, F. Wang and Z.-X. Wang, Unprecedented 2D homochiral hybrid lead-iodide perovskite thermochromic ferroelectrics with ferroelastic switching, Angew. Chem., Int. Ed., 2021, 60, 10730 CrossRef PubMed.
- J.-Q. Gan, Z.-K. Xu, T. Gan, Y. Qin and Z.-X. Wang, Large phase-transition temperature enhancement achieved in a layered lead iodide hybrid crystal by H/F substitution, Inorg. Chem., 2023, 62, 14469 CrossRef PubMed.
- L. long, Z. Huang, Z.-K. Xu, T. Gan, Y. Qin, Z. Chen and Z.-X. Wang, H/F substitution activating tunable dimensions and dielectric-optical properties in organic lead-bromide hybrids, Inorg. Chem. Front., 2024, 11, 845 RSC.
- J.-Y. Yu, Z.-K. Xu, J.-Q. Gan, Y. Qin, H.-P. Lv and Z.-X. Wang, Beryllium-based ABX3-type organic-inorganic hybrid halide ferroelastic, Inorg. Chem., 2025, 64, 5284 CrossRef PubMed.
- M. F. Mostafa and A. Hassen, Phase transition and electric properties of long chain Cd(II) layered perovskites, Phase Transitions, 2006, 79, 305 CrossRef.
- X. N. Hua, W. Q. Liao, Y. Y. Tang, P.-F. Li, P. P. Shi, D. Zhao and R. G. Xiong, A Room-Temperature Hybrid Lead Iodide Perovskite Ferroelectric, J. Am. Chem. Soc., 2018, 140, 12296 CrossRef PubMed.
- A. Kojima, K. Teshima, Y. Shirai and T. Miyasaka, Organometal halide perovskites as visible-light sensitizers for photovoltaic cells, JACS Commun., 2009, 131, 6050 CrossRef PubMed.
- T. M. Koh, et al., Formamidinium-containing metal-halide: An alternative material for near-IR absorption perovskite solar cells, J. Phys. Chem. C, 2014, 118, 16458 CrossRef.
- Y. H. Khattak, E. Vega, F. Baig and B. M. Soucase, Performance investigation of experimentally fabricated lead iodide perovskite solar cell via numerical analysis, Mater. Res. Bull., 2022, 151, 111802 CrossRef.
- A. Babayight, A. Ethirajan, M. Muller and B. Conings, Toxicity of organometal halide perovskite solar cells, Nat. Mater., 2016, 15, 247 CrossRef PubMed.
- G. D. Stucky, J. B. Folkers and T. J. Kistenmacher, The crystal and molecular structure of tetraethylammonium tetrachloro-nickelate (II), Acta Crystallogr., 1967, 23, 1064 CrossRef CAS.
- T. P. Melia and R. Merrifield, Thermal properties of transition-metal compounds. part II. complexes of manganese, Iron, Cobalt, Nickel, Copper, and Zinc of the type (Net4)2MCl4, J. Chem. Soc. A, 1970, 1166 RSC.
- J. N. Mcelearney, G. E. Shankle, R. W. Schwartz and R. L. Carlin, Low-temperature magnetic characteristics of tetrahedral CoCl42-. II. Nature of the phase transition in [(C2H5)4N]2CoCl4, J. Chem. Phys., 1972, 56, 3755 CrossRef CAS.
- A. J. Wolthuis, W. J. Huiskamp, L. J. de Jongh and R. L. Carlin, Investigation of structural phase transitions in some [(C2H5)4N]2MX4 compounds, with M = Co, Zn, Mn, and X = Cl, Br, Physica B, 1986, 142, 301 CrossRef CAS.
- M. Kahrizi and M. O. Steinitz, Structural phase transitions in ((C2H5)4N)2CdX4 compounds with X = Cl, Br, Solid State Commun., 1989, 70, 599 CrossRef CAS.
- M. Kahrizi and M. O. Steinitz, Thermal expansion and phase transitions in some tetra-alkylammonium metal chlorides, Solid State Commun., 1990, 74, 333 CrossRef CAS.
- O. G. Vlokh, I. I. Polovinko, V. I. Mokryi and S. A. Sveleba, Optical birefringence of single crystals of [N(C2H5)4]2ZnBr4 and [N(C2H5)4]2CdBr4, Sov. Phys. Crystallogr., 1991, 36, 131 Search PubMed.
- M. Iwata and Y. Ishibashi, Dielectric dispersion in [N(C2H5)4]2ZnCl4 single crystal, J. Phys. Soc. Jpn., 1991, 60, 3245 CrossRef CAS.
- Z. Czapla and S. Dacko, Structural phase transitions in [(C2H5)4N]2MX4 crystals, Ferroelectrics, 1992, 125, 17 CrossRef CAS.
- T. Kawata, T. Aoyama and S. Ohba, Tetraethylammonium tetrachlorocuprate (II), [N(C2H5)4]2[CuCl4], Acta Crystallogr., Sect. C: Cryst. Struct. Commun., 1993, 49, 137 CrossRef.
- S. Sveleba, V. Morkyi, I. Polovinko, V. Kapustyanik and Z. Trybula, Birefringent and dielectric properties of [N(C2H5)4]2ZnBr4 and [N(C2H5)4]2CdBr4 crystals, Acta Phys. Pol. A, 1993, 83, 777 CrossRef.
- O. Caetano, M. Lopez, A. Mahoui, J. Lapasset, J. Moret, T. Assih and P. S. Gregoire, Structural instabilities in the (TEA)2 m Cl4 crystalline family: A DSC study, Ferroelectr. Lett., 1995, 19, 69 CrossRef.
- A. Mahoui, J. Lapasset, J. Moret and P. S. Gregoire, Structure of (TEA)2CuCl4 and hydration, Z. Kristallogr., 1995, 210, 125 Search PubMed.
- A. Mahoui, J. Lapasset, J. Moret and P. S. Gregoire, Bis (tetraethylammonium) Tetrachlorometallates, [(C2H5)4N]2[MCl4], where M=Hg, Cd, Zn, Acta Crystallogr., Sect. C: Cryst. Struct. Commun., 1996, 52, 2671 CrossRef.
- A. Mahoui, J. Lapasset, D. G. Sannikov, J. Moret and P. S. Gregoire, On the ordering in [(C2H5)4N]2CuCl4 crystal: an X-ray study and theoretical considerations, Z. Phys. B, 1996, 99, 543 CrossRef.
- Z. Tylczynski and P. Biskupski, Low-temperature dielectric dispersion in [N(C2H5)4]2CuCl4 crystal, Solid State Commun., 1997, 102, 385 CrossRef.
- M. Machida, T. Ishino, Y. Shimoikeda, S. Gondo, N. Kotano and Y. Iwata, NMR and X-ray investigations of phase transition in [N(C2H5)4]2ZnCl4, Ferroelectrics, 1998, 217, 105 CrossRef.
- Z. Tylczynski and P. Biskupski, Thermal properties of [N(C2H5)4]2CuCl4, J. Korean Phys. Soc., 1998, 32, S235 Search PubMed.
- R. Poprawski, A. Liber and E. Malek, Dilatometric investigations of overcritical behavious in [N(C2H5)4]2CuCl4 crystals, Acta Phys. Pol. A, 2000, 98, 61 CrossRef.
- Z. Tylczynski, P. Biskupski and M. Slaboszewska, Dielectric dispersion in [N(C2H5)4]2MeCl4 crystals, Ferroelectrics, 2002, 272, 315 CrossRef.
- M. A. Kandhaswamy and V. Srinivasan, Synthesis and characterization of tetrarthylammonium tetrachlorocobaltate crystals, Bull. Mater. Sci., 2002, 25, 41 CrossRef.
- K. Gesi, Effect of hydrostatic pressure on the phase transitions in [N(C2H5)4]2CuCl4, Ferroelectrics, 2003, 285, 139 CrossRef.
- P. Biskupski, M. Slaboszewska and Z. Tylczynski, Changes in the optical properties at phase transitions in TEA2MeCl4 (Me = Zn, Mn, Hg, Cu) crystals, Phys. B, 2005, 370, 6 CrossRef CAS.
- A. U. Sheleg, A. M. Natumovets, T. I. Dekola and N. P. Tekhanovich, Effect of γ irradiation on the structural and thermal properties of [N(C2H5)4]2ZnBr4 in the vicinity of a first-order phase transition, Phys. Solid State, 2006, 48, 354 CrossRef CAS.
- M. Maczka, A. Cizman, R. Poprawski and J. Hanuza, Temperature-dependent vibrational studies of [N(C2H5)4]2MnCl4, J. Raman Spectrosc., 2007, 38, 1622 CrossRef CAS.
- A. U. Sheleg, E. M. Zub and A. Y. Yachkovskii, Crystallographic characteristics and phase transitions in the [N(C2H5)4]2CdBr4 crystal in the low-temperature range, Phys. Solid State, 2007, 49, 1973 CrossRef CAS.
- T. I. Dekola, A. U. Sheleg and N. P. Tekhanovich, Heat capacity of the [N(C2H5)4]2CdBr4 crystal in the temperature range 80-300 K, Phys. Solid State, 2007, 49, 1766 CrossRef CAS.
- A. R. Lim and K.-Y. Lim, Phase-transition mechanisms of [N(C2H5)4]2BCl4 and [N(CH3)4]2BCl4 (B=63Cu and 67Zn) single crystals studied by proton NMR, Solid State Commun., 2008, 147, 11 CrossRef CAS.
- P. Biskupski and Z. Tylczynski, Structure of TEA2ZnCl4 crystal surfaces studied by AFM, Phase Transitions, 2008, 81, 971 CrossRef CAS.
- A. Ostrowski and A. Cizman, EPR studies of linewidth anomalies at phase transitions in [N(C2H5)4]2MnCl4, Physica B, 2008, 403, 3110 CrossRef CAS.
- A. R. Lim, Study on ethyl groups with two different orientations in [N(C2H5)4]2CuBr4, J. Phys. Chem. Solid., 2016, 93, 59 CrossRef CAS.
- A. R. Lim, Study of the ferroelastic phase transition in the tetraethylammonium compound [N(C2H5)4]2ZnBr4 by magic-angle spinning and static NMR, AIP Adv., 2016, 6, 035307 CrossRef.
- A. R. Lim, M. S. Kim and K.-Y. Lim, Nuclear magnetic resonance study of the ferroelastic phase transition of order-disorder type [N(C2H5)4]2CdCl4, Solid State Sci., 2016, 58, 101 CrossRef CAS.
- A. R. Lim and K.-Y. Lim, Structural changes near phase transition temperatures for the [N(C2H5)4] groups in hydrated [N(C2H5)4]2CuCl4·xH2O, J. Therm. Anal. Calorim., 2017, 130, 879 CrossRef CAS.
- M. B. Bechir and A. B. Rhaiem, Synthesis, Thermal Analysis, Optical, Electric Properties and Conduction Mechanism of Hybrid Halogenometallates: [N(C2H5)4]2CoCl4, J. Phys. Soc. Jpn., 2021, 90, 74709 CrossRef.
- V. Kapustianik, I. Polovinko, Yu. Korchak, S. Sveleba, R. Tchukvinskyi and S. Kaluza, Phase transitions in [N(C2H5)4]2MeCl2Br2 (Me=Zn, Co, Cu) solid solutions, Phys. Status Solidi A, 1997, 161, 515 CrossRef CAS.
- V. Kapustianyk, Y. Shchur, I. Kityk, V. Rudyk, G. Lach, L. Laskowski, S. Tkaczyk, J. Swiatek and V. Davydov, Resonance dielectric dispersion of TEA-CoCl2Br2 nanocrystals incorporated into the PMMA matrix, J. Phys.: Condens. Matter, 2008, 20, 365215 CrossRef.
- V. Kapustianyk, Ya. Shchur, I. Kityk, V. Rudyk, G. Lach, L. Laskowski, S. Tkaczyk, J. Swiatek and V. Davydov, Resonance dielectric dispersion of TEA-CoCl2Br2 nanocrystals incorporated into the PMMA matrix, J. Phys.: Condens. Matter, 2008, 20, 365215 CrossRef.
- V. Kapustianyk, V. Rudyk, P. Yonak and B. Kundys, Magnetic and dielectric properties of [N(C2H5)4]2CoClBr3 solid solution: A new potential multiferroic, Phys. Status Solidi B, 2015, 252, 1778 CrossRef CAS.
- V. Kapustianyk, B. Cristovao, D. Osypiuk, Yu. Eliyashevskyy, Yu. Chornii and B. Sadovyi, Magnetic and ferroelectric properties of new potential magnetic multiferroic [N(C2H5)4]2CoCl2Br2, Acta Phys. Pol. A, 2021, 140, 450 CrossRef.
- V. Kapustianyk, Yu. Eliyashevskyy, U. Mostovoi, S. Semak, R. Tarasenko, V. Tkac, A. Feher and E. Cizmar, The correlation between electric polarization and magnetic properties in [N(C2H5)4]2CoCl2Br2 crystal at low temperatures, Physica B, 2022, 646, 414299 CrossRef.
- Z. Wu, H. Tang, T. Dai, Y. Long, D. Luo, P. Jiang, X. Xiong, Y. Xu, X. Zhang and Q. Hu, Mixed-ligand Engineering to enhance luminescence of Mn2+- based metal halides for wide color gamut display, materials, 2024, 17, 4459 CrossRef PubMed.
- N. Izumi, Dielectric properties of N(CH3)4N(C2H5)4CoBr4 crystal, J. Phys. Soc. Jpn., 1996, 65, 1189 CrossRef.
- F. Hlel, A. Rheim, T. Guerfel and K. Guidara, Synthesis, calorimetric study, infrared spectroscopy and crystal structure investigation of β-[tetraethylammonium tetramethylammonium tetrachlorozincate (II)] [β]-[(C2H5)4N][(CH3)4N]ZnCl4, Z. Naturforsch., 2006, 61b, 1002 CrossRef.
- E. Bozkurt, I. Ucar, I. Kartal, B. Karabulut and Y. Bekdemir, Structural, spectroscopic and EPR studies of tetraethlyammonium tetraethylammonium tetrabromocuprate (II) complex, Z. Kristallogr., 2009, 224, 163 Search PubMed.
- K. Karoui, A. B. Rhaiem, F. Hlel, M. Arous and K. Guidara, Dielectric and electric studies of the [N(CH3)4][N(C2H5)4]ZnCl4compound at low temperature, Mater. Chem. Phys., 2012, 133, 1 CrossRef.
- SMART and SAINT-Plus v6.22, Bruker AXS Inc., Madison, Wisconsin, USA, 2000 Search PubMed.
- J. L. Koenig, Spectroscopy of Polymers, Elsevier, New York, 1999 Search PubMed.
- A. Abragam, The Principles of Nuclear Magnetism, Oxford University Press, 1961 Search PubMed.
- S. H. Kim, D. Shin, Y.-J. Ko and A. R. Lim, CCDC 2477720: Experimental Crystal Structure Determination, 2025, DOI:10.5517/ccdc.csd.cc2p58gm.
|
This journal is © The Royal Society of Chemistry 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.