Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Phase transition of (CH3)2CHNH3CuCl3: crystal growth, crystal structure, coordination geometry, and molecular motion

Hyunje Park b and Ae Ran Lim *ab
aGraduate School of Carbon Convergence Engineering, Jeonju University, Jeonju 55069, Korea
bDepartment of Science Education, Jeonju University, Jeonju 55069, Korea. E-mail: aeranlim@hanmail.net; arlim@jj.ac.kr

Received 24th January 2025 , Accepted 3rd June 2025

First published on 5th June 2025


Abstract

The single crystals of (CH3)2CHNH3CuCl3 were grown, and their phase transition temperature was determined to be 341 K (TC) through differential scanning calorimetry and powder X-ray diffraction. Additionally, the structures of the single crystals were examined via single-crystal X-ray diffraction at 300 K (phase II, below TC) and 350 K (phase I, above TC). The sample underwent a thermochromic transition near the TC, changing from a low-temperature brown material to a high-temperature orange material. The phase II and I systems were triclinic (P[1 with combining macron]) and orthorhombic (Pcan), respectively. The 1H, 13C, and 15N nuclear magnetic resonance (NMR) chemical shifts of the system were monitored to investigate the structural geometry of the (CH3)2CHNH3 cations near phase I and II. Abrupt changes were observed in the chemical shifts near the TC, indicating a first-order phase transition from triclinic to orthorhombic symmetry. The results of this study imply that the structural environments around 1H, 13C, and 15N in the (CH3)2CHNH3 cations change significantly during phase II to I transition. Moreover, energy transfer was discussed based on NMR spin–lattice relaxation time data. This study indicates that the structural and physical properties of (CH3)2CHNH3CuCl3, an organic–inorganic material, make it a promising candidate for a wide range of potential applications.


1. Introduction

The effects of external magnetic fields on low-dimensional magnets have garnered significant attention from both experimental and theoretical perspectives. In particular, the critical properties of alternating spin-1/2 chains in magnetic fields have been extensively studied. Rapid progress in the fabrication of antiferromagnetic–ferromagnetic (AF–F) compounds has fueled this interest further. Owing to a gap in their spin excitation spectrum, AF–F chains exhibit interesting quantum behavior when subjected to a magnetic field.1 A typical example of an AF–F compound comprising alternating ferromagnetic and antiferromagnetic Heisenberg chains with S = 1/2 is organic-inorganic (CH3)2CHNH3CuCl3 (isopropylammonium copper trichloride).2–18 From the perspective of crystal structure, the origin of the spin gap in AF–F chains is attributed to the S = 1/2 ferromagnetic–antiferromagnetic alternating chains along the a-axis in the system. Consequently, AF–F compounds can be described as structures comprising a spin ladder along the c-axis consisting of strongly coupled ferromagnetic rungs and an antiferromagnetic chain with an excitation gap of 13.6 K based on neutron inelastic scattering experiments. Meanwhile, similar findings on crystal growth, density functional theory (DFT), and characterization of organic optical materials have been recently reported.19–21

(CH3)2CHNH3CuCl3 undergoes a thermochromic phase transition at TC = 324 K with a color change from brown at low temperatures to orange at high temperatures.22,23 The most easily detectable solid-state phase transitions are those in which the crystal color changes with temperature. Discontinuous thermochromic phase transitions are characterized by a significant color change at a specific temperature, indicating an abrupt shift in the properties of the crystal at TC. In transition–metal complexes, thermochromism often indicates a change in ligand coordination. These transformations may involve alterations in atomic connectivity or geometric distortions in chemical structures.23

Low-temperature phase II systems adopt a triclinic structure with the space group P[1 with combining macron]. However, the lattice parameters reported by Roberts et al.23 are a = 11.692 Å, b = 7.804 Å, c = 6.106 Å, α = 79.00°, β = 122.60°, and γ = 116.47° with Z = 2, whereas those reported by Manaka et al.2 are α = 97.62°, β = 101.05°, and γ = 67.28°. High-temperature phase I systems are orthorhombic with the space group Pcan, and show the following lattice parameters: a = 17.589 Å, b = 7.296 Å, and c = 6.365 Å with Z = 4. In phase II, each Cu2+ is located almost at the center of the base of a pyramid comprising five Cl ions, and adjacent pyramids are inverted. Thus, the phase II structure consists of bibridged linear chains of Cu2Cl6 dimers, whereas the phase I structure comprises tribridged chains of CuCl3 dimers. The (CH3)2CHNH3 cations are ordered in phase II, but become disordered in phase I.23

The heat transport properties of the spin-gap material (CH3)2CHNH3CuCl3 at ultra-low temperatures and under high magnetic fields are discussed in the literature.14 The temperature dependencies of the gap energies and magnon lifetimes are measured using quasi-one-dimensional S = 1/2 gapped quantum magnets using inelastic neutron scattering.12 Additionally, inelastic neutron scattering and bulk magnetic susceptibility measurements of the quantum S = 1/2 spin ladder system (CH3)2CHNH3CuCl3 have been conducted under hydrostatic pressure.24 Nuclear magnetic resonance (NMR) studies on 1H and 35Cl were conducted at low temperatures below TN = 13.5 K.25–29 The temperature dependence of optical linear birefringence in the AF–F state has also been measured.30 However, studies on the thermodynamic properties and molecular motions of the (CH3)2CHNH3 cations in (CH3)2CHNH3CuCl3 near the phase transition temperature of the system are scarce.

In this study, single crystals of (CH3)2CHNH3CuCl3 were grown and their phase transition temperatures were determined using differential scanning calorimetry (DSC) and X-ray diffraction (XRD) experiments with powder samples. Additionally, the structures of the single crystals above and below the phase transition temperature, corresponding to phases I and II, were examined via single-crystal XRD at 350 and 300 K, respectively. Moreover, the thermal properties of the system are briefly discussed. To investigate the influence of (CH3)2CHNH3 cations near phase I and II, the temperature dependence of the chemical shifts in the 1H, 13C, and 15N NMR spectra of the synthesized system was analyzed. Abrupt shifts were observed in phase I and II near 340 K, indicating a transition from triclinic to orthorhombic symmetry. In addition, the energy transfer, indicated by NMR spin–lattice relaxation time (T) data, was examined to highlight the contribution of (CH3)2CHNH3 cations toward the properties of (CH3)2CHNH3CuCl3. Through DSC, TGA, and NMR measurements, we want to understand the thermal stability and molecular motion of this material as a result of its physicochemical properties, enabling the development of high-efficiency and highly stable materials.

2. Experimental

2.1. Crystal growth

High-quality single crystals of (CH3)2CHNH3CuCl3 were grown through the slow evaporation method using a mixture containing (CH3)2CHNH2·HCl and CuCl2 in a 1[thin space (1/6-em)]:[thin space (1/6-em)]1 ratio of 15 g and 21.10 g, respectively, in ethanol. After stirring and heating to saturation, single crystals were grown from it via gradual evaporation for several weeks in a constant-temperature bath at 297 K. The flat, rectangular, dark-brown crystals with a size of up to 15 × 5 × 1 mm3 were stored in a desiccator to prevent moisture-related degradation.

2.2. Characterization

A TA Instruments DSC (Model 25) system with temperature increment and heating rates of 10 °C min−1 over a temperature range of 200–570 K under a dry nitrogen gas flow was used for DSC measurements. The crystal morphology and its variations with temperature were observed using an optical polarizing microscope (Carl Zeiss) on a hot stage (Linkam THMS 600). In addition, thermogravimetric (TG) analysis was conducted over a temperature range of 300–973 K under a nitrogen atmosphere at a temperature-increment rate of 10 °C min−1. Fourier transform infrared (FT-IR) spectra were measured within 4000–500 cm−1 on a PerkinElmer (L1600300) spectrometer using a compressed KBr pellet.

The structures and lattice parameters of the crystals at 300 and 350 K were determined using single-crystal XRD (SCXRD) experiments conducted at the Korea Basic Science Institute (KBSI), Seoul Western Centre. A Bruker SMART CCD diffractometer utilizing graphite-monochromated Mo-Kα radiation and a nitrogen cold stream at −50 °C was used for SCXRD. SMART APEX3 and SAINT were used for data collection and integration, whereas the multiscan method implemented in SADABS was used for absorption corrections. The crystal structure was solved by direct methods and refined using full-matrix least squares on F2 using the SHELXTL program.31 All non-hydrogen atoms were refined anisotropically, and hydrogen atoms were positioned at their idealized geometric locations. In addition, powder XRD (PXRD) patterns were recorded at several temperatures using an XRD system with a Mo-Kα radiation source, similar to the system used for SCXRD.

The NMR spectra of (CH3)2CHNH3CuCl3 were recorded using a 400 MHz NMR spectrometer (Bruker, Germany) at the KBSI, and T data were recorded on a 500 MHz NMR spectrometer (Bruker, Germany) at the Laboratory of NMR, NCIRF, Seoul National University. 1H magic angle spinning (MAS) NMR experiments were conducted at a Larmor frequency of 400.13 and 500.13 MHz, whereas 13C MAS NMR experiments were conducted at a Larmor frequency of 100.61 and 125.77 MHz. The 1H and 13C chemical shifts were referenced to tetramethylsilane (TMS). 15N MAS NMR spectra were recorded at a Larmor frequency of 40.54 MHz, and NH4NO3 was used as the standard. During experimentation, powdered samples were placed in a 4-mm CP/MAS tube, and magic-angle spinning speeds of 5 and 10 kHz were used to minimize spinning sidebands. The one-dimensional 1H and 13C NMR spectra were recorded at a delay time of 0.5 s. The 1H and 13C T values were measured with delay times ranging from 1 ms to 5 s, and 90° pulses for 1H and 13C were set to 3.2 and 6.5 μs, respectively. NMR experiments were conducted by lowering the temperature from 300 K to 180 K, and then increasing the temperature again from 180 K to 420 K.

3. Results and discussion

3.1. Phase transition temperature and thermodynamic properties

For DSC experiment, powdered (CH3)2CHNH3CuCl3 single crystals, utilizing a sample mass of 7.5 mg were subjected to a temperature increment at the rate of 10 °C min−1. As shown in Fig. 1, a small endothermic peak during the heating step was obtained at 341 K with an enthalpy value of 3.52 kJ mol−1, and a large peak was obtained near 469 K. The small peak observed at 341 K corresponds to the phase transition temperature, and the peak at 469 K indicates the melting point. The phase change occurring at this temperature was later confirmed using SCXRD and PXRD. Here, the temperature range below 341 K was classified as phase II and the region above 341 K was classified as phase I. Changes in the state of the single crystal with temperature, observed using a polarizing microscope, are shown in the inset of Fig. 1. Below 345 K, the single crystal is dark brown (Fig. 1(a)–(c)), and above 345 K, which is near the phase transition temperature, the color of the single crystal is slightly orange (Fig. 1(d)). In addition, at ∼469 K, a small bubble appears on the right side of the single crystal (Fig. 1(e) and (f)), consistent with the DSC results, and the crystal begins melting. Near 490 K (Fig. 1(g)), the crystal melts. Thus, both DSC and polarizing microscopy experiments indicated a thermochromic phase transition from dark-brown to dark-orange at 340 K and single-crystal melting at 469 K. Even though the color of the single crystal changes above TC, the coexistence of orange and black phases is observed. Detailed images of the single crystal are shown in the supplementary data S1 (ESI). Additionally, the single crystal was maintained at 350 K and the change over time was observed. After 1 min, the color immediately changed to dark orange and remained dark orange until 100 min. These results are represented in supplementary data S2 (ESI).
image file: d5ma00063g-f1.tif
Fig. 1 DSC curve of (CH3)2CHNH3CuCl3 during heating at 10 °C min−1 (Inset: Morphology of a single crystal at (a) 300 K, (b) 330 K, (c) 345 K, (d) 360 K, (e) 460 K, (f) 475 K, (g) 490 K, and (f) 500 K, as observed by optical polarizing microscopy).

Similar to DSC, TG experiments were conducted with the sample of 6.7 mg over a temperature range of 300–873 K at a heating rate of 10 °C min−1 (Fig. 2). TG analysis showed that the crystals were thermally stable up to 461 K with a weight loss of 2%; at 461 K, partial decomposition began to occur. In addition, the TG curve decreased significantly over two stages, and only ∼10% of the residue remained at temperatures above 850 K. In particular, the inflection point near 585 K in the TG curve corresponds to a weight loss of 60%, which is due to the decomposition of (CH3)2CHNH3Cl; this can almost entirely be attributed to the decomposition of organic matter.


image file: d5ma00063g-f2.tif
Fig. 2 TG curve of (CH3)2CHNH3CuCl3 during heating at 10 °C min−1. Td indicates the decomposition temperature. (Inset: Morphology of a single crystal and FT-IR spectrum of (CH3)2CHNH3CuCl3 within 4000–500 cm−1 at 300 K).

FT-IR spectrum recorded within 4000–500 cm−1 at room temperature are shown in the inset of Fig. 2. The very strong peak near 3116 cm−1 corresponds to the C–H mode, the weak and broad peak at 3000 cm−1 corresponds to the N–H⋯Cl hydrogen bond, the very small peak at 1683 cm−1 corresponds to NH3, the spectral peaks at 1468 and 1382 cm−1 correspond to CH3, and the peaks near 1202 and 975 cm−1 correspond to the C–N mode.

3.2. X-ray diffraction experiments on single-crystals and powder

The crystal structure and lattice parameters were determined by SCXRD experiments with single crystals of (CH3)2CHNH3CuCl3 at 300 and 350 K. In phase II of 300 K, the crystal structure exhibited a triclinic structure with the space group P[1 with combining macron] and the following lattice parameters: a = 6.1124 Å, b = 7.8109 Å, c = 9.7517 Å, α = 67.3010°, β = 82.3850°, γ = 79.0490°, and Z = 2. In phase I of 350 K, the crystal structure showed orthorhombic symmetry with the space group Pbcn and the following lattice constants: a = 6.3635 Å, b = 17.6055 Å, c = 7.2977 Å, α = β = γ = 90°, and Z = 4. The detailed SCXRD results at 300 and 350 K are listed in Table 1. The triclinic structure and numbering of atoms in phase II at 300 K are shown in Fig. 3(a) and (b), respectively. The orthorhombic structure and numbering of atoms at 350 K in phase I are shown in Fig. 4(a) and (b), respectively. The bond lengths and angles at 300 and 350 K are listed in Tables 2 and 3. The bond-lengths of N–H⋯Cl hydrogen bond at 300 K and 350 K are 3.806 Å and 3.489 Å, respectively. Single crystals of the synthesized sample were confirmed to comprise (CH3)2CHNH3 cations and CuCl3 anions, with Cu atoms surrounded by Cl atoms. Fig. 3 shows that the anionic structure of phase II consists of nearly planar Cu2Cl6 dimers linked by long axial Cu–Cl bonds, forming bibridged linear chains along the c-axis. The Cu is surrounded by five Cl atoms in a nearly square pyramidal arrangement. In the high – temperature phase, the crystal structure changes from a bibridged chain of dimers to a tribridged chain. The crystal structure consists of linear chains of face-covalent CuCl3 linked by hydrogen bonds to (CH3)2CHNH3 cations. During phase II to I transition, the b-axis elongates and CuCl3 is aligned along the c-axis. Notably, the triclinic structure in phase II with C1, C2, and C3 in different environments at 300 K changes to an orthorhombic structure with improved symmetry owing to C1 and C2 in different environments at 350 K in phase I. Therefore, the symmetry of the (CH3)2CHNH3 ions is higher in phase I than in phase II. Crystallographic data at 300 and 350 K, including the CIF files, were deposited at the Cambridge Crystallographic Data Center (CCDC 2419287, 2419289).
Table 1 Crystal data and structure refinement for (CH3)2CHNH3CuCl3 at 300 K and 350 K
Temperature 300 K 350 K
Chemical formula C3H10NCuCl3 C3H10NCuCl3
Weight 230.01 230.01
Crystal system Triclinic Orthorhombic
Space group P[1 with combining macron] Pbcn
a (Å) 6.1124 (2) 6.3635 (2)
b (Å) 7.8109 (3) 17.6055 (9)
c (Å) 9.7517 (4) 7.2977 (3)
α (°) 67.3010 (10) 90
β (°) 82.3850 (10) 90
γ (°) 79.0490 (10) 90
Z 2 4
V3) 420.79 (3) 817.58 (17)
Density (calculated)(Mg m−3) 1.815 1.869
Absorption coefficient (mm−1) 3.457 3.559
F(000) 230 460
Crystal size (mm3) 0.410 × 0.146 × 0.135 0.204 × 0.173 × 0.113
Theta range for data collection (°) 2.860 to 28.314 3.951 to 28.250
Index ranges −8 ≤ h ≤ 8, −10 ≤ k ≤ 10, −13 ≤ l ≤ 13 −9 ≤ h ≤ 9, −23 ≤ k ≤ 23, −8 ≤ l ≤ 8
Completeness to theta = 25.242° (%) 99.9 98.7
Max and min transmission 0.7457 and 0.6480 0.7457 and 0.6396
Data/restraints/parameters 2092/0/76 1011/2/44
Radiation type Mo-Kα Mo-Kα
Wavelength (Å) 0.71073 0.71073
Reflections collected 15[thin space (1/6-em)]685 8240
Independent reflections 2092 (Rint = 0.0294) 1011 (Rint = 0.0208)
Goodness-of-fit on F2 1.102 1.116
Final R indices [I > 2 Sigma(I)] R 1 = 0.0167, wR2 = 0.0430 R 1 = 0.0205, wR2 = 0.0559
R indices (all data) R 1 = 0.0194, wR2 = 0.0441 R 1 = 0.0236, wR2 = 0.0581
Extinction coefficient 0.065(2) n/a
Largest diff. peak and hole (e Å−3) 0.323 and −0.232 0.314 and −0.379



image file: d5ma00063g-f3.tif
Fig. 3 The triclinic structure and numbering of atoms in a (CH3)2CHNH3CuCl3 crystal at 300 K (A. R. Lim CCDC 2419287).

image file: d5ma00063g-f4.tif
Fig. 4 The orthorhombic structure and numbering of atoms in a (CH3)2CHNH3CuCl3 crystal at 350 K (A. R. Lim CCDC 2419289).
Table 2 The bond-lengths (Å) and bond-angles (°) for (CH3)2CHNH3CuCl3 at 300 K
Bond-lengths Bond-angles Bond-angles
Cu(1)–Cl(2) 2.2709(4) Cl(2)–Cu(1)–Cl(3) 93.208(14) N(1)–C(1)–C(2) 108.42(14)
Cu(1)–Cl(3) 2.2722(4) Cl(2)–Cu(1)–Cl(1) 90.925(14) N(1)–C(1)–C(3) 108.73(14)
Cu(1)–Cl(1) 2.3033(4) Cl(3)–Cu(1)–Cl(1) 160.430(18) C(2)–C(1)–C(3) 113.36(16)
Cu(1)–Cl(1)#1 2.3168(4) Cl(2)–Cu(1)–Cl(1)#1 175.085(15) N(1)–C(1)–H(1) 108.8
Cu(1)–Cl(2)#2 2.7040(4) Cl(3)–Cu(1)–Cl(1)#1 90.587(14) C(2)–C(1)–H(1) 108.8
N(1)–C(1) 1.504(2) Cl(1)–Cu(1)–Cl(1)#1 84.469(14) C(3)–C(1)–H(1) 108.8
N(1)–H(1A) 0.8900 Cl(2)–Cu(1)–Cl(2)#2 90.687(13) C(1)–C(2)–H(2A) 109.5
N(1)–H(1B) 0.8900 Cl(3)–Cu(1)–Cl(2)#2 98.506(14) C(1)–C(2)–H(2B) 109.5
N(1)–H(1C) 0.8900 Cl(1)–Cu(1)–Cl(2)#2 100.565(15) H(2A)–C(2)–H(2B) 109.5
C(1)–C(2) 1.508(2) Cl(1)#1–Cu(1)–Cl(2)#2 91.840(15) C(1)–C(2)–H(2C) 109.5
C(1)–C(3) 1.516(2) Cu(1)–Cl(2)–Cu(1)#2 89.313(13) H(2A)–C(2)–H(2C) 109.5
C(1)–H(1) 0.9800 Cu(1)–Cl(1)–Cu(1)#1 95.531(14) H(2B)–C(2)–H(2C) 109.5
C(2)–H(2A) 0.9600 C(1)–N(1)–H(1A) 109.5 C(1)–C(3)–H(3A) 109.5
C(2)–H(2B) 0.9600 C(1)–N(1)–H(1B) 109.5 C(1)–C(3)–H(3B) 109.5
C(2)–H(2C) 0.9600 H(1A)–N(1)–H(1B) 109.5 H(3A)–C(3)–H(3B) 109.5
C(3)–H(3A) 0.9600 C(1)–N(1)–H(1C) 109.5 C(1)–C(3)–H(3C) 109.5
C(3)–H(3B) 0.9600 H(1A)–N(1)–H(1C) 109.5 H(3A)–C(3)–H(3C) 109.5
C(3)–H(3C) 0.9600 H(1B)–N(1)–H(1C) 109.5 H(3B)–C(3)–H(3C) 109.5
H(1A)–Cl(3) 2.916


Table 3 The bond-lengths (Å) and bond-angles (°) for (CH3)2CHNH3CuCl3 at 350 K
Bond-lengths Bond-angles Bond-angles
Cu(1)–Cl(2)#1 2.2915(4) Cl(2)#1–Cu(1)–Cl(2) 180.0 C(2)–C(1)–N(1) 109.6(2)
Cu(1)–Cl(2) 2.2915(4) Cl(2)#1–Cu(1)–Cl(1) 90.878(13) C(2)–C(1)–H(1) 105.3
Cu(1)–Cl(1) 2.3384(4) Cl(2)–Cu(1)–Cl(1) 89.122(13) N(1)–C(1)–H(1) 105.3
Cu(1)–Cl(1)#1 2.3385(4) Cl(2)#1–Cu(1)–Cl(1)#1 89.123(13) C(1)–C(2)–H(2A) 109.5
N(1)–C(1) 1.506(4) Cl(2)–Cu(1)–Cl(1)#1 90.878(13) C(1)–C(2)–H(2B) 109.5
N(1)–H(1A) 0.8900 Cl(1)–Cu(1)–Cl(1)#1 180.0 H(2A)–C(2)–H(2B) 109.5
N(1)–H(1B) 0.8900 Cl(1)–Cu(1)–Cl(1)#3 85.738(18) C(1)–C(2)–H(2C) 109.5
N(1)–H(1C) 0.8900 C(1)–N(1)–H(1A) 109.5 H(2A)–C(2)–H(2C) 109.5
C(1)–C(2) 1.477(3) C(1)–N(1)–H(1B) 109.5 H(2B)–C(2)–H(2C) 109.5
C(1)–H(1) 0.9800 H(1A)–N(1)–H(1B) 109.5
C(2)–H(2A) 0.9600 C(1)–N(1)–H(1C) 109.5
C(2)–H(2B) 0.9600 H(1A)–N(1)–H(1C) 109.5
C(2)–H(2C) 0.9600 H(1B)–N(1)–H(1C) 109.5
H(1C)–Cl(2) 2.599


Simulated PXRD patterns based on the CIF data of the crystal structures at 300 and 350 K are shown in Fig. 5. The patterns corresponding to phase I and II are completely different owing to structural differences in both phases. The strongest peak in phase II is observed at (001), whereas that in phase I is observed at (020). The Mercury software was used for peak reflection. Crushed single crystal samples were used for PXRD experiments over the 2θ range of 8–55° above 300 K including the simulated powder patterns in 300 and 350 K, as shown in Fig. 6. The same peaks were consistently observed in the PXRD patterns recorded below 340 K, but a change in the peak pattern was observed in spectra recorded above 340 K, consistent with the endothermic peak observed at 341 K in the DSC results. Notably, spectra recorded above 460 K did not contain any peaks corresponding to a crystalline structure, consistent with the melting of the single crystal. Consequently, based on DSC, SCXRD, and PXRD analyses, 341 K was identified as the phase transition temperature. On the other hand, the PXRD peaks at 350 K exhibit a coexistence of peaks from both phase I and II, primarily in the low-angle region, as indicated by the ellipses in Fig. 6. A more detailed figure was presented in supplementary data S3 (ESI).


image file: d5ma00063g-f5.tif
Fig. 5 The simulated powder X-ray diffraction patterns of (CH3)2CHNH3CuCl3 crystals in phase II and I at 300 and 350 K, respectively.

image file: d5ma00063g-f6.tif
Fig. 6 X-ray diffraction powder patterns of (CH3)2CHNH3CuCl3 near the phase transition temperature including the simulated powder patterns in 300 and 350 K.

3.3. 1H MAS NMR chemical shifts and spin–lattice relaxation time

The NMR spectra of (CH3)2CHNH3CuCl3 crystals, recorded using MAS NMR at a frequency of 400.13 MHz under a magnetic field of 9.4 T, were analyzed to understand the structural geometry of 1H near the TC. Complex spectra were recorded owing to the chemical shift distribution and anisotropy at numerous nonequivalent proton sites in the (CH3)2CHNH3 group. The 1H NMR spectra of phase II and I below and above the TC, respectively, are shown in Fig. 7. NMR experiments at a spinning speed of 5 kHz indicated a sideband distance of 5 kHz (= 12.5 ppm) around the 1H signal, which is similar to the spinning rate. In Fig. 7, the 1H NMR peaks are indicated by arrows. At 300 K, a 1H NMR chemical shift is observed at 10.8 ppm. Only one peak overlaps with the 1H peaks of CH3, CH, and NH3. However, the peaks corresponding to NH3 and CH3 at 400 K above the TC are separated to 5.78 and 3.85 ppm, respectively. The sidebands of NH3 and CH3 are denoted as x and o, respectively. The 1H NMR chemical shifts with changes in temperature are shown in Fig. 7. With increasing temperature, the 1H NMR chemical shifts show an abrupt change with peak splitting in phase I, indicating significant changes in the structural environment around 1H atoms during phase II to I transition. With increasing temperature, the line widths become narrower above TC, indicating a higher mobility of 1H in phase I than in phase II.
image file: d5ma00063g-f7.tif
Fig. 7 1H NMR spectra of (CH3)2CHNH3CuCl3 with increasing temperature at 400.13 MHz. The x and o symbols indicate the sidebands for NH3 and CH3, respectively (Inset: 1H NMR chemical shifts in phase II and I).

T values, which represent the extent of energy transfer surrounding the 1H in the (CH3)2CHNH3 cation, were estimated to understand the nature of molecular motions in the synthesized system. The intensity changes in the 1H NMR spectra, recorded at various delay times, are characterized by the decay rate of magnetization, which is governed by the T, as follows:32–35

 
f(τ) = f(0)exp(−τ/T),(1)
where f(τ) and f(0) represent the signal intensities at time τ and τ = 0, respectively. The 1H T value, obtained from the linear slope of the intensity vs. delay time plot, determined using eqn (1), was found to be about 8.73 ms at 300 K.

3.4. 13C MAS NMR chemical shifts and spin–lattice relaxation time

13C NMR spectra of (CH3)2CHNH3CuCl3 were recorded using 13C MAS NMR spectroscopy at a frequency of 100.61 MHz and spinning speed of 5 kHz. 13C NMR spectra as a function of temperature are shown in Fig. 8. Using TMS as the standard, the sidebands marked with * are spaced around the 13C signal at 5 kHz (equivalent to 50 ppm), which is consistent with the spinning speed. The 13C NMR peaks are indicated by arrows in Fig. 8(a). Fig. 8(b) shows the temperature dependence of the 13C NMR chemical shifts in phase II and I. Two 13C NMR peaks are observed in both phases, consistent with the presence of 13C in CH and CH3. The 13C chemical shifts corresponding to CH and CH3 at 330 K of phase II are observed at 77.16 and 19.19 ppm, respectively, and at 113.35 and 56.52 ppm, respectively, whereas the 13C chemical shifts recorded for CH and CH3 at 340 K of phase I are obtained at 113.35 and 56.52 ppm, respectively. The discontinuous temperature-dependent changes in 13C chemical shifts near the TC indicate abrupt changes in the structural geometry around these nuclei, implying that the structural environment around 13C changes significantly during phase II to I transition. Fig. 8(b) shows that with increasing temperature, the line widths above the TC reduce rapidly from about 12.60 ppm at 200 K to 2.52 ppm at 400 K, indicating a higher mobility of 13C in phase I than in phase II.
image file: d5ma00063g-f8.tif
Fig. 8 (a) 13C MAS NMR spectra of (CH3)2CHNH3CuCl3 in phase II and I recorded at 100.61 MHz. The * symbol indicates the sidebands for CH and CH3, and (b) 13C NMR chemical shifts and line widths of (CH3)2CHNH3CuCl3 in phase II and I.

The 13C T for the (CH3)2CHNH3 cation was analyzed to understand the nature of molecular motions in (CH3)2CHNH3CuCl3. Variations in the intensity of the 13C NMR spectra were monitored by extending the delay time to 300 K. 13C T values were determined by modifying the spin-locking pulse sequence applied after cross-polarization (CP). Following CP, 13C magnetization was generated by proton spin-locking. The proton field was then deactivated for a variable period τ, while the 13C rf field remained active. Subsequently, the free induction decay of 13C was recorded under high-power proton decoupling and subjected to Fourier transformation. Using the same method for obtaining T in 1H, the values of the 13C T for CH and CH3 were determined to be about 99.8 and 55.9 ms, respectively, at 300 K. This result implies that the C atom of CH is bonded with N and H, whereas the C atom of CH3 is bonded with H; thus, energy transfer is easier in CH3.

3.5. 15N MAS NMR chemical shifts

15N NMR spectra were measured using MAS NMR at a frequency of 40.54 MHz under a magnetic field of 9.4 T. Although the natural abundance of 14N is higher than that of 15N, the frequency of 14N (26.52 MHz) is lower than that of 15N (40.54 MHz), making it relatively easy to acquire 15N NMR spectra. The 15N NMR spectra of (CH3)2CHNH3CuCl3 as a function of temperature are shown in Fig. 9. Recording 15N NMR signals was challenging owing to the low natural abundance of 15N and baseline wiggling. The chemical shifts in this spectrum change abruptly on phase II to I transition. Notably, two signals are observed at 135.84 ppm (marked in blue) and 14.30 ppm (marked in olive) at 330 K, which is slightly below the TC, indicating the coexistence of phase II and I, while only one signal is observed at 17.06 ppm in phase I at 340 K. There seems to be some coexistence of phase I and II near TC; the peaks at 350 K in the PXRD are the coexistence of phase I and II, and similarly, the 15N NMR peaks in phase I and II were also observed simultaneously. The abrupt chemical shifts observed in 15N spectra were attributed to changes in the structural geometry of the system, reflecting alterations in the atomic configuration surrounding the 15N nuclei. This abrupt shift in phase I and II near 340 K indicates a transition from triclinic to orthorhombic symmetry.
image file: d5ma00063g-f9.tif
Fig. 9 15N MAS NMR spectra and chemical shifts of (CH3)2CHNH3CuCl3 in phase II and I at 40.54 MHz. Blue squares and arrows indicate peaks observed in phase II, while red circles and arrows indicate peaks observed in phase I. Here, olive arrows indicate signals suggesting the slight coexistence of the phase II and I states near the TC.

4. Conclusions

In summary, the phase transition temperature and the melting temperature of (CH3)2CHNH3CuCl3 single crystals were determined to be 341 K (TC) and 469 K (Tm), respectively, using DSC and PXRD in this study. Additionally, the structures of the single crystals were examined using SCXRD at 300 and 350 K, corresponding to phase II and I. Low-temperature phase II systems showed a triclinic structure with the space group P[1 with combining macron], whereas high-temperature phase I systems showed an orthorhombic structure with the space group Pcan with thermodynamic stability extending up to ∼461 K. Meanwhile, the single crystal appeared dark brown at room temperature, but near TC, a mixture of dark brown and dark orange was observed. This indicates the coexistence of phase I and II near TC, as confirmed by PXRD experiments and the 15N NMR spectrum. Furthermore, the 1H, 13C, and 15N chemical shifts changed discontinuously near the TC, and the phase transition was confirmed to be first-order based on NMR results. These results imply that the structural environments around 1H, 13C, and 15N in (CH3)2CHNH3 cations change significantly when transitioning from phase II to phase I. Notably, phase transition and 1H and 13C NMR linewidth narrowing occur at the same temperature. The narrowing of linewidths near the TC is consistent with changes in the free rotation of the cation on phase transition. The structural phase transition can be attributed to changes in the positions of CH3, CH, and NH3, which do not affect the magnetic properties. The fundamental mechanism of phase II to I transition in (CH3)2CHNH3CuCl3 crystals indicates that these materials are promising candidates for a wide variety of applications in various fields. Also, understanding their motion through NMR measurements can improve thermal stability and charge transport, enabling the development of high-efficiency and highly stable materials.

Author contributions

A. R. Lim performed the NMR and the x-ray experiments, and wrote the manuscript. H. Park measured the DSC, TGA experiments.

Data availability

Crystallographic data for (CH3)2CHNH3CuCl3 has been deposited at the CCDC 2419287 and 2419289.

Conflicts of interest

There are no conflicts to declare.

References

  1. S. Mahdavifar and A. Akbara, J. Phys. Soc. Jpn., 2008, 77, 024710 CrossRef.
  2. H. Manaka, I. Yamada and K. Yamaguchi, J. Phys. Soc. Jpn., 1997, 66, 564 CrossRef CAS.
  3. H. Manaka and I. Yamada, J. Mag. Mag. Mater., 1998, 177–181, 681 Search PubMed.
  4. H. Manaka, I. Yamada, N. V. Mushnikov and T. Goto, J. Phys. Soc. Jpn., 2000, 69, 675 CrossRef CAS.
  5. T. Nakamura, Prog. Theor. Phys. Suppl., 2002, 145, 353 Search PubMed.
  6. I. Yamada and H. Manaka, J. Magn. Magn. Mater., 2003, 258–259, 114 CrossRef CAS.
  7. T. Sakai and S. Yamamoto, Synth. Met., 2005, 152, 477 Search PubMed.
  8. T. Nakamura, Phys. Rev. B: Condens. Matter Mater. Phys., 2005, 71, 144401 CrossRef.
  9. T. Suzuki, T. Saito, T. Sasaki, A. Oosawa, T. Goto, S. Awaji, K. Watanabe, N. Kobayashi and H. Manaka, J. Phys.: Conf. Ser., 2006, 51, 187 CrossRef CAS.
  10. T. Saito, T. Sasaki, T. Suzuki, A. Oosawa, T. Goto, S. Awaji, K. Watanabe and N. Kobayashi, J. Phys. Soc. Jpn., 2007, 76, 084708 Search PubMed.
  11. V. O. Garlea, A. Zheludev, T. Masuda, H. Manaka, L.-P. Regnault, E. Ressouche, B. Grenier, J.-H. Chung, Y. Qiu, K. Kiefer and M. Boehm, Phys. Rev. Lett., 2007, 98, 167202 CrossRef CAS PubMed.
  12. A. Zheludev, V. O. Garlea, L.-P. Regnault, H. Manaka, A. Tsvelik and J.-H. Chung, Phys. Rev. Lett., 2008, 100, 157204 Search PubMed.
  13. H. Tsujii, Y. H. Kim, Y. Yoshida, Y. Takano, T. P. murphy, K. Kanada, T. Saito, A. Oosawa and T. Goto, J. Phys.: Conf. Ser., 2009, 150, 42217 Search PubMed.
  14. Z. Y. Zhao, B. Tong, X. Zhao, L. M. Chen, J. Shi, F. B. Zhang, J. D. Song, S. J. Li, J. C. Wu, H. S. Xu, X. G. Liu and X. F. Sun, Phys. Rev. B: Condens. Matter Mater. Phys., 2015, 91, 134420 Search PubMed.
  15. T. Saito, A. Oosawa, T. Goto, T. Suzuki and I. Watanabe, Phys. Rev. B: Condens. Matter Mater. Phys., 2006, 74, 134423 Search PubMed.
  16. T. Nakamura, J. Phys. Soc. Jpn., 2003, 72, 789 Search PubMed.
  17. B. Nafradi, T. Keller, H. Manaka, A. Zheludev and B. Keimer, Phys. Rev. Lett., 2011, 106, 177202 CrossRef CAS PubMed.
  18. H. Manaka, I. Yamada, M. Tokunaga, T. Yasuhira and N. Miura, Phys. B, 1998, 246–247, 513 CrossRef CAS.
  19. P. Prabukanthan, V. Bhakyajothi, M. Saravana Kumar, G. Harichandran, K. Dinakaran and P. Seenuvasakumaran, J. Mol. Struct., 2021, 1246, 131172 CrossRef CAS.
  20. P. Prabukanthan, C. Raveendiran, M. Saravana Kumar, G. Harichandran, K. Dinakaran, A. A. Al-Kahtani, M. Ubaidullan, G. Ushanandhini and B. Pandit, Optik, 2022, 270, 170014 Search PubMed.
  21. P. Prabukanthan, V. Bhakyajothi, K. Dinakaran, S. Uthayakumar and A. Younis, Chem. Papers, 2023, 77, 703 CrossRef CAS.
  22. A. Nishimori and M. Sorai, J. Phys. Chem. Solids, 1999, 60, 895 CrossRef CAS.
  23. S. A. Roberts, D. R. Bloomquist, R. D. Willett and H. W. Dodgen, J. Am. Chem. Soc., 1981, 103, 2603 CrossRef CAS.
  24. T. Hong, V. O. Garlea, A. Zheludev and J. A. Fernandez-Beca, Phys. Rev. B: Condens. Matter Mater. Phys., 2008, 78, 224409 CrossRef.
  25. K. Kanada, T. Saito, A. Oosawa, T. Goto, T. Suzuki and H. Manaka, J. Phys. Chem. Solids, 2007, 68, 2191 CrossRef CAS.
  26. H. Manaka, I. Yamada and W. Higemoto, J. Phys. Soc. Jpn., 2007, 76, 014704 Search PubMed.
  27. T. Saito, H. Inoue, J. Tonisi, A. Oosawa, T. Goto, T. Sasaki, N. Kobayasi, S. Awaji and K. Watanabe, J. Phys.: Conf. Ser., 2006, 51, 203 CrossRef CAS.
  28. H. Manaka and I. Yamada, Phys. B, 2000, 284–288, 1607 CrossRef CAS.
  29. T. Kubo and T. Waketa, Phys. B, 2000, 284–288, 1609 CrossRef CAS.
  30. H. Manaka, I. Yamada, T. Kikuchi, K. Morishita and K. Ilo, J. Phys. Soc. Jpn., 2001, 70, 2509 Search PubMed.
  31. SHELXTL v6.10, Bruker AXS, Inc., 2000 Search PubMed.
  32. A. R. Lim, Sci. Rep., 2024, 14, 11808 Search PubMed.
  33. A. Abragam, The Principles of Nuclear Magnetism, Oxford University Press, 1961 Search PubMed.
  34. R. K. Harris, Nuclear Magnetic Resonance Spectroscopy, Pitman Pub, UK, 1983 Search PubMed.
  35. J. L. Koenig, Spectroscopy of Polymers, Elsevier, New York, 1999 Search PubMed.

Footnote

Electronic supplementary information (ESI) available. CCDC 2419287 and 2419289. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d5ma00063g

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.