Next Article in Journal / Special Issue
Numerical Studies of Batch and Inline High Shear Melt Conditioning Technology Using Different Rotors
Previous Article in Journal
A New Correction Theory and Verification on the Reducing Rate Distribution for Seamless Tube Stretch-Reducing Process
Previous Article in Special Issue
Selection Criterion of Stable Dendritic Growth for a Ternary (Multicomponent) Melt with a Forced Convective Flow
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Structural Characterization and Dynamics of a Layered 2D Perovskite [NH3(CH2)5NH3]MnCl4 Crystal near Phase Transition Temperature

1
Graduate School of Carbon Convergence Engineering, Jeonju University, Jeonju 55069, Korea
2
Department of Science Education, Jeonju University, Jeonju 55069, Korea
*
Author to whom correspondence should be addressed.
Crystals 2022, 12(9), 1298; https://doi.org/10.3390/cryst12091298
Submission received: 29 July 2022 / Revised: 9 September 2022 / Accepted: 12 September 2022 / Published: 14 September 2022
(This article belongs to the Special Issue Phase Transition in External Fields)

Abstract

:
[NH3(CH2)5NH3]MnCl4 crystals are grown via slow evaporation, and the crystal undergoes a phase transition at 298 K (TC) according to differential scanning calorimetry, and the structures determined via X-ray diffraction at 173 and 333 K are orthorhombic systems in the space group Imma. These results differed slightly from those previously reported, and the reasons for this are analyzed. The thermal stability is relatively high, with a thermal decomposition temperature of approximately 570 K. The 1H spin-lattice relaxation times t exhibited very large variations, as indicated by the large thermal displacement around the 1H atoms, suggesting energy transfer at ~TC, even if no structural changes occurred. The influences of the chemical shifts of 1H of NH3 and short t of C1 adjacent to NH3 in cation are insignificant, indicating a minor change in the N−H⋯Cl hydrogen bond related to the coordination geometry of the MnCl6 octahedron. These properties will be make it a potential application for eco-friendly solar cells.

1. Introduction

Hybrid perovskite compounds are of scientific interest because of the diversity of their crystal structures, which govern their structural dynamics and ferroelastic and thermodynamic properties. In addition, ferroelasticity is commonly observed in compounds with perovskite crystal structures, and the ferroelastic twin domains in organic–inorganic hybrid perovskites attract much attention [1,2,3,4,5,6]. The organic cation of the hybrid material contributes to properties such as structural flexibility and optical characteristics, whereas the inorganic anion is responsible for the thermal and mechanical properties [7,8]. Moreover, the fabrication of hybrid perovskites was recently reported as a major challenge in the context of developing ferroelastic semiconductors [9]. Furthermore, successful hybrid perovskite ferroelectric performances render hybrid perovskites suitable candidates for use in flexible and wearable devices [10,11]. Additionally, solar cells based on CH3NH3PbX3 (X = Cl, Br, or I) organic–inorganic hybrid compounds recently attracted interest. However, perovskites containing Pb are toxic and decompose in humid air, and thus, developing alternative green hybrid perovskite solar cells is necessary [4,5,6,12,13,14,15,16]. Hence, detailed characterizations of perovskite structures and the dynamics of novel organic–inorganic hybrid compounds [NH3(CH2)nNH3]MX4 (n = 2, 3, 4, …, MII is a transition metal, such as Mn, Fe, or Cu, and X is a halogen ion), which crystallize in perovskite-type layer structures in various configurations, are increasingly necessary due to their potential applications as green alternatives. The diammonium hybrid perovskites [NH3(CH2)nNH3]BX4, with one-dimensional (0D) and two-dimensional (2D) structures, have been extensively investigated [17,18,19,20,21,22,23,24,25,26,27,28,29,30]. In the case of B = Mn, Cu, and Cd, the crystal structures consist of an alternate octahedrally coordinated (BX6)2− and organic cations. In the case of B = Co and Zn, isolated tetrahedral (BX4)2− is coordinated between the organic cations [31,32]. These 2D hybrid perovskite types have various potential applications in electrochemical devices such as chemical sensors, supercapacitors, batteries, and solar cells [17,24,25,33].
Layered 2D hybrid perovskite [NH3(CH2)5NH3]MnCl4 crystals (n = 5, M = Mn, and X = Cl) comprise organic and inorganic ions. The organic [NH3(CH2)5NH3]2+ cations and inorganic [MnCl4]2− anions are alternately stacked along the longest axis, with the inorganic layer extended via corner-shared octahedra. The organic and inorganic layers are interconnected by N−H⋯Cl hydrogen bonds [23]. The phase transition associated with the order–disorder transition between two orthorhombic phases is from Pnma to Imma at 299.6 K (= phase transition temperature TC) [31]. The lattice constants at 298 K reported by Mondal et al. [23] are a = 7.1742 Å, b = 7.3817 Å, c = 23.9650 Å, and Z = 4. The crystal structure at 298 K is shown in Figure 1 (CCDC 1401387) [23], with each layer of alkylenediammonium chains inserted between two infinite sheets of corner-sharing MnCl6 octahedra.
Research regarding [NH3(CH2)5NH3]MnCl4 was initially published by Arend et al. [34,35], mainly as research reports regarding the crystal structure and TC of 299.6 K based on the heat capacity. Based on recent results, Mondal et al. [23] studied the crystallographic characteristics of this crystal, with Lv et al. [36] reporting the dielectric and photoluminescence characteristics.
In this study, [NH3(CH2)5NH3]MnCl4 single crystals were grown using an aqueous solution-based method, and TC was confirmed using differential scanning calorimetry (DSC). In addition, the structures of the crystals below TC and above TC were confirmed using single-crystal X-ray diffraction (XRD). The thermodynamic properties were investigated as a function of temperature. Finally, the structural dynamics of the [NH3(CH2)5NH3]2+ cation at ~TC were analyzed using nuclear magnetic resonance (NMR) chemical shifts and spin-lattice relaxation times t. The physicochemical properties of [NH3(CH2)5NH3]MnCl4 without structural changes at ~TC should render its use in proton conductors viable.

2. Materials and Methods

Single crystals of perovskite-type [NH3(CH2)5NH3]MnCl4 were grown via slow evaporation from an aqueous solution containing NH2(CH2)5NH2·2HCl (98%, Sigma-Aldrich, St. Louis, MO, USA) and MnCl2 (98%, Sigma-Aldrich). The mixture was stirred and heated, the resulting solution was filtered, and light-yellow single crystals were obtained after five weeks in a constant-temperature bath at 300 K.
Fourier transform infrared (FT-IR) spectra were measured between 4000 and 500 cm−1 using an L1600300 spectrometer (PerkinElmer, Waltham, MA, USA) and compressed KBr pellets.
DSC (DSC 25, TA Instruments, New Castle, DE, USA) was performed to observe the structural phase transitions by heating in the temperature range 200–480 K at 10 K min−1 under N2 gas. Thermogravimetric analysis (TGA) was performed using a thermogravimetric analyzer (TA Instruments) in the temperature range 300–870 K at the same heating rate.
The lattice parameters at various temperatures were determined via single-crystal XRD at the Western Seoul Center of the Korea Basic Science Institute. A crystal was lifted in paratone oil and mounted in a D8 Venture diffractometer (Bruker, Billerica, MA, USA) equipped with a Mo-Kα radiation source, PHOTON III M14 detector (Bruker), and a nitrogen cold atmosphere (−50 °C). Data collection and integration were performed using SMART APEX3 (Bruker, 2016) and SAINT (Bruker, 2016), and absorption correction was performed using a multiscan method implemented in SADABS (Bruker, 2002). The structure was analyzed and refined via the full-matrix least-squares method on F2 using SHELXTL (University of Göttingen, Göttingen, Germany).
In order to check whether the peak obtained from the DSC result is TC or a melting temperature Tm, it was observed using a polarizing optical microscope (Carl Zeiss, Oberkochen, Germany) with a THMS600 heating stage (Linkam, Salfords, UK) at an appropriate temperature for a single crystal.
NMR spectroscopy of the [NH3(CH2)5NH3]MnCl4 crystals was conducted using a 400 MHz Avance II+ solid NMR spectrometer (Bruker) with a 4 mm magic angle spinning (MAS) probe (Western Seoul Center, KBSI). 1H and 13C MAS NMR spectra were recorded at Larmor frequencies of 400.13 and 100.61 MHz, respectively. The MAS speed used to minimize the spinning sideband overlap was 10 kHz, and NMR chemical shifts were calibrated using tetramethylsilane (TMS) as the standard. The 1H and 13C t values were obtained via the π/2 − τ sequence method by changing the spin-locking pulses—the π/2 pulse widths for 1H and 13C were ~3.7 μs. The temperature variation was determined by adjusting the heater current and N2 gas flow.

3. Results

3.1. FT-IR Spectroscopy

Figure 2 shows the FT-IR spectrum of the [NH3(CH2)5NH3]MnCl4 crystal at 300 K in the range 4000–500 cm−1. The bands at 3122 and 3043 cm−1 are characteristic of the C–H bonds of the protonated ligand, and the band at 2934 cm−1 suggests the presence of N−H⋯Cl hydrogen bonds. The band at 1568 cm−1 is due to the asymmetric mode of NH3, whereas the strong band at 1488 cm−1 is assigned to the symmetric deformation mode of NH3. Finally, the bands close to 1169 and 980 cm−1 are assigned to the C–N and C–C modes, respectively. The observed FT-IR bands are consistent with those previously reported [23].

3.2. Phase Transition and Crystal Structure

The DSC thermogram measured at a heating rate of 10 K min−1 under an N2 atmosphere is shown in Figure 3. An endothermic peak is observed at 298 K, and the TC of 298 K is consistent with that reported previously [36]. The enthalpy for the phase transition was 689 J/mol.
The structures obtained via single-crystal XRD are identical to the orthorhombic structures below TC and above TC. The lattice constants at 173 K (<TC) are a = 24.1756 Å, b = 7.1535 Å, and c = 7.3314 Å in the space group Imma, whereas those at 333 K (>TC) are a = 23.9162 Å, b = 7.1877 Å, and c = 7.3898 Å in the space group Imma. Table 1 shows the single-crystal data collection and refinement parameters of the [NH3(CH2)5NH3]MnCl4 crystal at 173 and 333 K, and the atomic numbering scheme and thermal ellipsoids of the H atoms are shown in Figure 4. The Mn atom is coordinated to six Cl atoms, forming an almost regular octahedron, MnCl6, and the six N-linked hydrogen atoms in one formula unit form N−H⋯Cl hydrogen bonds. The lattice constants as functions of temperature are shown in Figure 5. The lattice constants do not change at ~TC, and it does not appear to be significantly related to the TC. The detailed results of XRD of the crystal structure are shown in the Supplementary Information S1 and S2.

3.3. Thermodynamic Properties

To determine whether the endothermic peaks were related to phase transitions or decomposition, TGA and differential thermal analysis (DTA) were performed at the same heating rate as that used during DSC. The TGA and DTA thermograms shown in Figure 6 reveal that the crystals are virtually stable up to approximately 570 K. The molecular weight of [NH3(CH2)5NH3]MnCl4 decreases with increasing temperature above 570 K, and the amount of residue based on the total molecular weight is obtained using Equation (1) [37,38]:
1st step
[NH3(CH2)5NH3]MnCl4 (MW: 300.94 g) → [NH2(CH2)5NH2·2HCl]MnCl2
                        → [NH2(CH2)5NH2·HCl]MnCl2 (s) + HCl (g)
Residue:
[NH2(CH2)5NH2·HCl]MnCl2 (s)/[NH3(CH2)5NH3]MnCl4 = 87.88%
2nd step:
[NH3(CH2)5NH3]MnCl4 → [NH2(CH2)5NH2·2HCl]MnCl2
               → [NH2(CH2)5NH2]MnCl2 (s) + 2HCl (g)
Residue:
[NH2(CH2)5NH2]MnCl2 (s)/[NH3(CH2)5NH3]MnCl4 = 75.77%
The temperature at which mass loss commences, based on the TGA thermogram, is approximately 570 K. Therefore, 570 K is the partial thermal decomposition temperature Td. Mass losses of approximately 12% and 24% close to 617 and 630 K may be attributed to the loss of HCl and 2HCl, respectively, as shown in Figure 6. The molecular weight of the crystal decreases sharply between 600 and 700 K, with a mass loss of 50% at approximately 700 K.
To verify the results of TGA and DSC, a single crystal was observed using a polarizing optical microscope while varying the temperature. At 300 K, the crystal is transparent and light yellow, as shown in Figure 6a. The crystal turns slightly opaque at ~580 K due to partial thermal decomposition. Upon heating further to 617 K, HCl is eliminated, the crystal turns brown, and the surface also appears to melt slightly, as shown in Figure 6b. Based on the results of DSC, TGA, and polarizing microscopy, TC = 298 K, as shown in the DSC, whereas Td = 570 K.

3.4. 1H and 13C NMR Chemical Shifts

The 1H MAS NMR spectra of the [NH3(CH2)5NH3]MnCl4 crystals recorded at ~TC are shown in Figure 7. The observed resonance lines at low temperatures are asymmetric because of the overlap of the signals representing NH3 and CH2. The linewidths A and B on the left- and right-hand sides of the half-maximum shown in Figure 7 are not equal. Above 300 K, the NH3 and CH2 signals are resolved, and the respective chemical shifts of the resonance lines of NH3 and CH2 are observed at 9.29 and 2.89 ppm. The spinning sidebands are marked with “+” and “o” to represent 1H in NH3 and CH2, respectively. The 1H chemical shifts of CH2 do not vary significantly at ~TC, whereas changes in the 1H chemical shifts of NH3 are observed at ~TC. The larger changes in the 1H NMR chemical shifts of NH3 compared to those in the 1H NMR chemical shifts of CH2 at ~TC suggest a change in the N−H⋯Cl hydrogen bonding between Cl around Mn and H of NH3.
In addition, the 13C NMR chemical shifts in the MAS NMR spectra of CH2 in [NH3(CH2)5NH3]MnCl4 were recorded at ~TC. The 13C signal of TMS was observed at 38.3 ppm at 300 K, and thus, 38.3 ppm was set as the origin for the 13C chemical shifts. Here, C3 in the [NH3(CH2)5NH3] cation is located at the center of the cation, C1 is located adjacent to the NH3 in the cation, and C2 is located between C1 and C3, as shown in the inset of Figure 8. The respective chemical shifts of C1, C2 and C3 at 300 K are observed at 113.44, 88.92, and 80.96 ppm, as shown in Figure 8. The 13C chemical shifts of C1 do not vary significantly at ~TC, whereas those of C2 and C3 vary at ~TC.

3.5. 1H and 13C NMR Spin-Lattice Relaxation Times

The 1H and 13C MAS NMR method has a very important for understanding the local dynamics. The spin-lattice relaxation times T1ρ for 1H and 13C in the rotating frame are the important experiment for studying the dynamical processes. By studying the relaxation of the nuclei in different environments within the cation, it is possible to obtain a detailed picture of the motions. The T1ρ relaxation parameters are particularly informative since it is directly related to those motions in the low- to mid-kHz frequency range [39,40,41].
The 1H and 13C MAS NMR spectra were acquired at several delay times at each temperature. The relationship between the intensities of the NMR signals and delay time is represented by an exponential function. The magnetization decay rates for protons and carbon are characterized by t as follows [42,43]:
P(τ)/P(0) = exp(−τ/t),
where P(τ) and P(0) are the NMR signal intensities at τ and τ = 0, respectively. The 1H and 13C NMR spectra of [NH3(CH2)5NH3]MnCl4 were recorded with various time delays. The decay curves may be represented by a single exponential function, as shown in Equation (2). However, the 1H t values of NH3 and CH2 at low temperatures may not be distinguished because of the overlapping 1H NMR signals. The 1H t values depend highly on the temperature, as shown in Figure 9. The 1H t values of CH2 and NH2 display significant changes at ~TC, indicating that the 1H energy transfer of CH2 and NH3 changes significantly. Moreover, the 13C t values of C1, C2, and C3 are obtained from the slopes of their recovery traces. The 13C t values of C1, C2, and C3 at ~TC are virtually continuous. Notably, the t values of C1, which are adjacent to NH3, are the shortest. The low t values of C1, which are close to the Mn2+ ions, are related to the magnetic moments of the Mn2+ ions, which are paramagnetic.

4. Discussion

The crystal structures, phase transitions, thermal stabilities, and molecular dynamics of the [NH3(CH2)5NH3]MnCl4 crystals were investigated using XRD, DSC, TGA, and NMR spectroscopy. First, we reconfirmed that the structure was an orthorhombic system in the space group Imma at 173 and 330 K, and Table 2 shows our results and those previously reported. Arend et al. suggested the space group as Ima2 or Imma at ~TC [34,35], and Chhor et al. [31] reported that the structure was orthorhombic in the space groups Pnma below TC and Imma above TC. According to Lv et al. [36], the space groups below TC and above TC were Pnma and Imma, respectively, and Modal et al. [23] reported that the space group at TC, not below TC or above TC, was I212121. In this study, the space group and lattice parameters of the crystal structure differ slightly, and thus, to study the phase transition, an accurate structural analysis is required. Our results are similar to those above TC reported by Lv et al. [36].
Our results showed that the thermal stability was relatively high, with a thermal decomposition temperature of approximately 570 K. The NMR spectra further suggested that the energy transfer of 1H at ~TC was very large, as indicated by the large thermal displacement around the 1H atoms.

5. Conclusions

As previously reported by other groups, the space groups under TC were Pnma or I212121. Thus, we analyzed Pnma and I212121 by XRD results, but determined to be closer to Imma. Based on the results of XRD, the structures below TC and above TC were orthorhombic in the space group Imma. There may be several reasons for the slightly different results. TC is 298 K, and thus, the temperature required to characterize the structure should be accurately determined. The difference in the single crystal structure may be due to the observed temperature difference, and it is thought that it may be slightly different depending on the crystal growth conditions; a single crystal may be grown into a naturally occurring structure or a single crystal having a new structure depending on temperature, which is one of the growth conditions. For example, the direction of the twin domain wall in the case of BiVO4 having a ferroelastic twin structure was different [44]. The phase transition at 298 K is due to the energy transfer of 1H, with no structural change. The influences of the chemical shifts of 1H of NH3 and short molecular weight of C1 adjacent to NH3 in the [NH3(CH2)5NH3]MnCl4 crystal were insignificant, indicating a minor change in the N−H∙∙∙Cl hydrogen bond related to the coordination geometry of the MnCl6 anion. The structural phenomenon revealed by XRD and NMR at ~TC shows the potential for the realization of solar cells for use in various applications.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cryst12091298/s1, Table S1: Crystal data and structure refinement for Imma_a; Table S2: Atomic coordinates (×104) and equivalent isotropic displacement parameters (Å2 × 103) for Imma_a; Table S3: Bond length [Å] and angle [°] for Imma_a; Table S4: Anisotropic displacement parameters (Å2 × 103) for Imma_a; Table S5: Hydrogen coordinates (×104) and isotropic displacement parameters (Å2 × 103) for Imma_a.

Author Contributions

A.R.L. designed the project and performed X-ray and NMR experiments. Y.N. performed DSC, TGA, and optical polarizing microscope experiments. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Basic Science Research Program of the National Research Foundation of Korea, which is funded by the Ministry of Education, Science, and Technology (2018R1D1A1B07041593, 2016R1A6A1A03012069).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Zhang, Z.X.; Su, C.-Y.; Li, J.; Song, X.-J.; Fu, D.-W.; Zhang, Y. Ferroelastic Hybrid Bismuth Bromides with Dual Dielectric Switches. Chem. Mater. 2021, 33, 5790–5799. [Google Scholar] [CrossRef]
  2. Su, C.-Y.; Yao, Y.-F.; Zhang, Z.-X.; Wang, Y.; Chen, M.; Huang, P.-Z.; Zhang, Y.; Qiao, W.-C.; Fu, D.-W. The construction of a two-dimensional organic–inorganic hybrid double perovskite ferroelastic with a high Tc and narrow band gap. Chem. Sci. 2022, 13, 4794–4800. [Google Scholar] [CrossRef] [PubMed]
  3. Shao, T.; Gong, J.M.; Liu, J.; Han, L.J.; Chen, M.; Jia, Q.; Fu, D.W.; Lu, H.-F. 2D lead-free organic–inorganic hybrid exhibiting dielectric and structural phase transition at higher temperatures. CrystEngComm 2022, 24, 4346–4350. [Google Scholar] [CrossRef]
  4. Hermes, I.M.; Bretschneider, S.A.; Bergmann, V.W.; Klasen, D.; Mars, J.; Tremel, W.; Laquai, F.; Butt, H.-J.; Mezger, M.; Berger, R.; et al. Ferroelastic Fingerprints in Methylammonium Lead Iodide Perovskite. J. Phys. Chem. C 2016, 120, 5724–5731. [Google Scholar] [CrossRef]
  5. Strelcov, E.; Dong, Q.; Li, T.; Chae, J.; Shao, Y.; Deng, Y.; Gruverman, A.; Huang, J.; Centrone, A. CH3NH3PbI3 perovskites: Ferroelasticity revealed. Sci. Adv. 2017, 3, e1602165. [Google Scholar] [CrossRef] [PubMed]
  6. Liu, Y.; Collins, L.; Proksch, R.; Kim, S.; Watson, B.R.; Doughty, B.; Calhoun, T.R.; Ahmadi, M.; Ievlev, A.V.; Jesse, S.; et al. Chemical nature of ferroelastic twin domains in CH3NH3PbI3 perovskite. Nat. Mater. 2018, 17, 1013–1019. [Google Scholar] [CrossRef] [PubMed]
  7. Zang, W.; Xiong, R.-G. Ferroelectric Metal–Organic Frameworks. Chem. Rev. 2012, 112, 1163–1195. [Google Scholar] [CrossRef]
  8. Lim, A.R.; Kwac, L.K. Advances in physicochemical characterization of lead-free hybrid perovskite [NH3(CH2)3NH3]CuBr4 crystals. Sci. Rep. 2022, 12, 8769. [Google Scholar] [CrossRef]
  9. Su, C.; Lun, M.; Chen, Y.; Zhou, Y.; Zhang, Z.; Chen, M.; Huang, P.; Fu, D.; Zhang, Y. Hybrid Optical-Electrical Perovskite Can Be a Ferroelastic Semiconductor. CCS Chem. 2021, 4, 2009–2019. [Google Scholar] [CrossRef]
  10. Xie, Y.; Ai, Y.; Zeng, Y.-L.; He, W.-H.; Huang, X.-Q.; Fu, D.-W.; Gao, J.-X.; Chen, X.-G.; Tang, Y.-Y.; Am, J. The Soft Molecular Polycrystalline Ferroelectric Realized by the Fluorination Effect. Chem. Soc. 2020, 142, 12486–12492. [Google Scholar] [CrossRef]
  11. Fu, D.-W.; Gao, J.-X.; He, W.-H.; Huang, X.-Q.; Liu, Y.-H.; Al, Y. High-Tc Enantiomeric Ferroelectrics Based on Homochiral Dabco-derivatives (Dabco=1,4-Diazabicyclo [2.2.2]octane). Angew. Chem. Int. Ed. 2020, 59, 17477–17481. [Google Scholar] [CrossRef] [PubMed]
  12. Knop, O.; Wasylishen, R.E.; White, M.A.; Cameron, T.S.; Van Oort, M.J.M. Alkylammonium lead halides. Part 2. CH3NH3PbX3 (X = Cl, Br, I) perovskites: Cuboctahedral halide cages with isotropic cation reorientation. Can. J. Chem. 1990, 68, 412–422. [Google Scholar] [CrossRef]
  13. Chen, Q.; Marco, N.D.; Yang, Y.; Song, T.-B.; Chen, C.-C.; Zhao, H.; Hong, Z.; Zhou, H.; Yang, Y. Under the spotlight: The organic–inorganic hybrid halide perovskite for optoelectronic applications. Nano Today 2015, 10, 355–396. [Google Scholar] [CrossRef]
  14. Abdel-Aal, S.K.; Abdel-Rahman, A.S.; Kocher-Oberlehner, G.G.; Ionov, A.; Mozhchil, R. Structure, optical studies of two-dimensional hybrid perovskite for photovoltaic applications. Acta Cryst. 2017, 73, C1116. [Google Scholar] [CrossRef]
  15. Mostafa, M.F.; Elkhiyami, S.S.; Alal, S.A. Discontinuous transition from insulator to semiconductor induced by phase change of the new organic-inorganic hybrid [(CH2)7(NH3)2]CoBr4. Mater. Chem. Phys. 2017, 199, 454–463. [Google Scholar] [CrossRef]
  16. Milic, J.V.; Im, J.-H.; Kubicki, D.J.; Ummadisingu, A.; Seo, J.-Y.; Li, Y.; Ruiz-Preciado, M.A.; Dar, M.I.; Zakeeruddin, S.M.; Emsley, L.; et al. Supramolecular Engineering for Formamidinium-Based Layered 2D Perovskite Solar Cells: Structural Complexity and Dynamics Revealed by Solid-State NMR Spectroscopy. Adv. Energy Mater. 2019, 9, 1900284. [Google Scholar] [CrossRef]
  17. Rao, C.N.R.; Cheetham, A.K.; Thirumurugan, A. Hybrid inorganic–organic materials: A new family in condensed matter physics. J. Phys. Condens. Matter 2008, 20, 83202. [Google Scholar] [CrossRef]
  18. Cheng, Z.; Lin, J. Layered organic–inorganic hybrid perovskites: Structure, optical properties, film preparation, patterning and templating engineering. CrystEngComm 2010, 12, 2646–2662. [Google Scholar] [CrossRef]
  19. Mostafa, M.F.; El-khiyami, S.S. Crystal structure and electric properties of the organic–inorganic hybrid: [(CH2)6(NH3)2]ZnCl4. J. Solid State Chem. 2014, 209, 82–88. [Google Scholar] [CrossRef]
  20. Gonzalez-Carrero, S.; Galian, R.E.; Perez-Prieto, J. Organometal Halide Perovskites: Bulk Low-Dimension Materials and Nanoparticles. Part. Syst. Charact. 2015, 32, 709–720. [Google Scholar] [CrossRef]
  21. Abdel-Adal, S.K.; Kocher-Oberlehner, G.; Ionov, A.; Mozhchil, R.N. Effect of organic chain length on structure, electronic composition, lattice potential energy, and optical properties of 2D hybrid perovskites [(NH3)(CH2)n(NH3)]CuCl4, n = 2–9. Appl. Phys. A 2017, 123, 531. [Google Scholar] [CrossRef]
  22. Liu, W.; Xing, J.; Zhao, J.; Wen, X.; Wang, K.; Peixiang, L.; Xiong, Q. Giant Two-Photon Absorption and Its Saturation in 2D Organic–Inorganic Perovskite. Adv. Opt. Mater. 2017, 5, 1601045. [Google Scholar] [CrossRef]
  23. Mondal, P.; Abdel-Aal, S.K.; Das, D.; Manirul Islam, S.K. Catalytic Activity of Crystallographically Characterized Organic–Inorganic Hybrid Containing 1,5-Di-amino-pentane Tetrachloro Manganate with Perovskite Type Structure. Cat. Lett. 2017, 147, 2332–2339. [Google Scholar] [CrossRef]
  24. Elseman, A.M.; Shalan, A.E.; Sajid, S.; Rashad, M.M.; Hassan, A.M.; Li, M. Copper-Substituted Lead Perovskite Materials Constructed with Different Halides for Working (CH3NH3)2CuX4-Based Perovskite Solar Cells from Experimental and Theoretical View. ACS Appl. Mater. Interfaces 2018, 10, 11699–11707. [Google Scholar] [CrossRef]
  25. Aramburu, J.A.; Garcia-Fernandez, P.; Mathiesen, N.R.; Garcia-Lastra, J.M.; Moreno, M. Changing the Usual Interpretation of the Structure and Ground State of Cu2+-Layered Perovskites. J. Phys. Chem. C 2018, 122, 5071–5082. [Google Scholar] [CrossRef]
  26. Abdel-Aal, S.K.; Ouasri, A. Crystal structure, Hirshfeld surfaces and vibrational studies of tetrachlorocobaltate hybrid perovskite salts NH3(CH2)nNH3CoCl4 (n = 4, 9). J. Mol. Struct. 2022, 1251, 131997. [Google Scholar] [CrossRef]
  27. Abdel, S.K.; Abdel-Rahman, A.S.; Gamal, W.M.; Abdel-Kader, M.; Ayoub, H.S.; El-Sherif, A.F.; Kandeel, M.F.; Bozhko, S.; Yakimov, E.E.; Yakimov, E.B. Crystal structure, vibrational spectroscopy and optical properties of a one-dimensional organic–inorganic hybrid perovskite of [NH3CH2CH(NH3)CH2]BiCl5. Acta Cryst. 2019, 75, 880–886. [Google Scholar] [CrossRef]
  28. Guillaume, F.; Sourisseau, C.; Lucazeau, G.; Dianoux, A.J. Reorientational motions and phase transitions in some perovskite type layered compounds [NH3(CH2)nNH3]MnCl4, n = 3, 5, in Dynamics of molecular crystals. In Proceedings of the 41st International Meeting on Physical Chemistry, Grenoble, France, 30 June–4 July 1986; p. 173. [Google Scholar]
  29. Guillaume, F.; Sourisseau, C.; Dianoux, A.J. Molecular motions in perovskite type layered compounds [NH3(CH2)nNH3]MnCl4, with n = 3, 4, 5; an incoherent neutron scattering study. Physica B 1989, 156, 359–362. [Google Scholar] [CrossRef]
  30. Negrier, P.; Couzi, M.; Chanh, N.B.; Hauw, C.; Meresse, A. Structural phase transitions in the perovskite-type layer compound NH3(CH2)5NH3CdCl4. J. Phys. Fr. 1989, 50, 405–430. [Google Scholar] [CrossRef]
  31. Chhor, K.; Abello, L.; Pommier, C.; Sourisseau, C. Reorientational motions in a perovskite-type layer compound [NH3(CH2)5NH3]MnCl4. A calorimetric study. J. Phys. Chem. Solids 1988, 49, 1079–1085. [Google Scholar] [CrossRef]
  32. Lim, A.R.; Kim, S.H. Physicochemical Property Investigations of Perovskite-Type Layer Crystals [NH3(CH2)nNH3]CdCl4 (n = 2, 3, and 4) as a Function of Length n of CH2. ACS Omega 2021, 6, 27568–27577. [Google Scholar] [CrossRef] [PubMed]
  33. Rong, S.-S.; Faheem, M.B.; Li, Y.-B. Perovskite single crystals: Synthesis, properties, and applications. J. Electron. Sci. Technol. 2021, 19, 100081. [Google Scholar] [CrossRef]
  34. Arend, H.; Granicher, H. On phase transitions in chloride perovskite layer structures. Ferroelectrics 1976, 13, 537–539. [Google Scholar] [CrossRef]
  35. Arend, H.; Tichy, K.; Baberschke, K.; Rys, F. Chloride perovskite layer compounds of [NH3-(CH2)n-NH3]MnCl4 formula. Solid State Commun. 1976, 18, 999–1003. [Google Scholar] [CrossRef]
  36. Lv, X.-H.; Liao, W.-Q.; Li, P.-F.; Wang, Z.-X.; Mao, C.-Y.; Zhang, Y. Dielectric and photoluminescence properties of a layered perovskite-type organic–inorganic hybrid phase transition compound: NH3(CH2)5NH3MnCl4. J. Mater. Chem. C 2016, 4, 1881–1885. [Google Scholar] [CrossRef]
  37. Lim, A.R.; Schueneman, G.T.; Novak, B.M. The T spin-lattice relaxation time of interpenetrating networks by solid state NMR. Solid State Commun. 1998, 109, 465. [Google Scholar] [CrossRef]
  38. Lim, A.R. Structural characterization, thermal properties, and molecular motions near the phase transition in hybrid perovskite [(CH2)3(NH3)2]CuCl4 crystals: 1H, 13C, and 14N nuclear magnetic resonance. Sci. Rep. 2020, 10, 20853. [Google Scholar] [CrossRef]
  39. Mcbrierty, V.J.; Packer, K.J. Nuclear Magnetic Resonance in Solid Polymers; Cambridge University Press: Cambridge, UK, 1993. [Google Scholar]
  40. Koenig, J.L. Spectroscopy of Polymers; Elsevier: New York, NY, USA, 1999. [Google Scholar]
  41. Harris, R.K. Nuclear Magnetic Resonance Spectroscopy; Pitman Pub: London, UK, 1983. [Google Scholar]
  42. Lim, A.R.; Kim, S.H.; Joo, Y.L. Physicochemical properties and structural dynamics of organic–inorganic hybrid [NH3(CH2)3NH3]ZnX4 (X = Cl and Br) crystals. Sci. Rep. 2021, 11, 8408. [Google Scholar] [CrossRef]
  43. Abragam, A. The Principles of Nuclear Magnetism; Oxford University Press: Oxford, UK, 1961. [Google Scholar]
  44. Lim, A.R.; Choh, S.H.; Jang, M.S. Prominent ferroelastic domain walls in BiVO4 crystal. J. Phys. Condens. Matter 1995, 7, 7309. [Google Scholar] [CrossRef]
Figure 1. Orthorhombic structure of a [NH3(CH2)5NH3]MnCl4 crystal at 298 K [23].
Figure 1. Orthorhombic structure of a [NH3(CH2)5NH3]MnCl4 crystal at 298 K [23].
Crystals 12 01298 g001
Figure 2. FT–IR spectrum of [NH3(CH2)5NH3]MnCl4 at 300 K.
Figure 2. FT–IR spectrum of [NH3(CH2)5NH3]MnCl4 at 300 K.
Crystals 12 01298 g002
Figure 3. Differential scanning calorimetry curve of [NH3(CH2)5NH3]MnCl4.
Figure 3. Differential scanning calorimetry curve of [NH3(CH2)5NH3]MnCl4.
Crystals 12 01298 g003
Figure 4. Thermal ellipsoid plot (50% probability) for structure of [NH3(CH2)5NH3]MnCl4 at 300 K.
Figure 4. Thermal ellipsoid plot (50% probability) for structure of [NH3(CH2)5NH3]MnCl4 at 300 K.
Crystals 12 01298 g004
Figure 5. Single-crystal XRD parameters of a [NH3(CH2)5NH3]MnCl4 crystal at 173, 280, 300 and 333 K.
Figure 5. Single-crystal XRD parameters of a [NH3(CH2)5NH3]MnCl4 crystal at 173, 280, 300 and 333 K.
Crystals 12 01298 g005
Figure 6. Thermogravimetric analysis and differential thermal analysis curves of [NH3(CH2)5NH3]MnCl4 (Inset: Changes in crystal by optical polarizing microscopy at (a) 300 K and (b) 617 K).
Figure 6. Thermogravimetric analysis and differential thermal analysis curves of [NH3(CH2)5NH3]MnCl4 (Inset: Changes in crystal by optical polarizing microscopy at (a) 300 K and (b) 617 K).
Crystals 12 01298 g006
Figure 7. 1H NMR chemical shifts of [NH3(CH2)5NH3]MnCl4 at 200, 250, 300, 310, 320, and 350 K. + and o are the spinning sidebands.
Figure 7. 1H NMR chemical shifts of [NH3(CH2)5NH3]MnCl4 at 200, 250, 300, 310, 320, and 350 K. + and o are the spinning sidebands.
Crystals 12 01298 g007
Figure 8. 13C NMR chemical shifts of [NH3(CH2)5NH3]MnCl4 at 250, 280, 290, 300, and 350 K. * and o are the spinning sidebands.
Figure 8. 13C NMR chemical shifts of [NH3(CH2)5NH3]MnCl4 at 250, 280, 290, 300, and 350 K. * and o are the spinning sidebands.
Crystals 12 01298 g008
Figure 9. Spin-lattice relaxation times t for 1H and 13C of [NH3(CH2)5NH3]MnCl4 near TC.
Figure 9. Spin-lattice relaxation times t for 1H and 13C of [NH3(CH2)5NH3]MnCl4 near TC.
Crystals 12 01298 g009
Table 1. Crystal data and structure refinement for [NH3(CH2)5NH3]MnCl4 at 173 K and 330 K. The full data are available in the CIF files.
Table 1. Crystal data and structure refinement for [NH3(CH2)5NH3]MnCl4 at 173 K and 330 K. The full data are available in the CIF files.
Chemical FormulaC5H16N2MnCl4C5H16N2MnCl4
Weight300.94300.94
Crystal systemOrthorhombicOrthorhombic
Space groupImmaImma
T (K)173330
a (Å)24.175623.9162
b (Å)7.15357.1877
c (Å)7.33147.3898
Z44
V (Å3)1267.891270.32
Radiation typeMo-KαMo-Kα
Wavelength (Å)0.710730.71073
Reflections collected54185258
Independent reflections867863
Goodness of fit on F21.0701.118
Final R indices [I > 2sigma(I)]R1 = 0.0383, wR2 = 0.1178R1 = 0.0312, wR2 = 0.0957
R indices (all data)R1 = 0.0394, wR2 = 0.1190R1 = 0.0330, wR2 = 0.0974
Table 2. Phase transition temperature TC (K), structure, space group, lattice constants (Å), Z, and measured temperature (K) for [NH3(CH2)5NH3]MnCl4 crystal.
Table 2. Phase transition temperature TC (K), structure, space group, lattice constants (Å), Z, and measured temperature (K) for [NH3(CH2)5NH3]MnCl4 crystal.
Arend et al.Chhor et al.Lv et al.Mondal et al.Present Work
TC301299.6298 298
StructureOrthor.Orthor.Orthor.Orthor.Orthor.Orthor.Orthor.Orthor.
Space groupIma2 or ImmaPnmaImmaPnmaImmaI212121ImmaImma
Lattice constantsa = 7.152 a = 7.149a = 23.94a = 7.1742a = 24.1756a = 23.9162
b = 7.360 b = 24.171b = 7.191b = 7.3817b = 7.1535b = 7.1877
c = 23.986 c = 7.334c = 7.399c = 23.9650c = 7.3314c = 7.3898
Z44244444
MeasuredAt room temp.299.6 < TC299.6 > TC173333298173333
Temperature
Reference[34,35][31][31][36][36][23]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lim, A.R.; Na, Y. Structural Characterization and Dynamics of a Layered 2D Perovskite [NH3(CH2)5NH3]MnCl4 Crystal near Phase Transition Temperature. Crystals 2022, 12, 1298. https://doi.org/10.3390/cryst12091298

AMA Style

Lim AR, Na Y. Structural Characterization and Dynamics of a Layered 2D Perovskite [NH3(CH2)5NH3]MnCl4 Crystal near Phase Transition Temperature. Crystals. 2022; 12(9):1298. https://doi.org/10.3390/cryst12091298

Chicago/Turabian Style

Lim, Ae Ran, and Yeji Na. 2022. "Structural Characterization and Dynamics of a Layered 2D Perovskite [NH3(CH2)5NH3]MnCl4 Crystal near Phase Transition Temperature" Crystals 12, no. 9: 1298. https://doi.org/10.3390/cryst12091298

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop