Open Access Article
Chang-Yuan Su
*ac,
Heng-Guan Yiac,
Hao-Fei Nib,
Guo-Wei Dud,
San-Qiang Xiaac,
Zunqi Liu*e,
Zhi-Xu Zhang
*b and
Da-Wei Fu
*b
aShanxi Key Laboratory for Radiation Safety and Protection, China Institute for Radiation Protection, Taiyuan, 030006, China. E-mail: scywmy@163.com
bInstitute for Science and Applications of Molecular Ferroelectrics, Key Laboratory of the Ministry of Education for Advanced Catalysis Materials, Zhejiang Normal University, Jinhua, 321004, China. E-mail: zhangzhixu@zjnu.edu.cn; dawei@zjnu.edu.cn
cCNNC Key Laboratory for Radiation Protection Technology, China Institute for Radiation Protection, Taiyuan, 030006, China
dJiangsu Key Laboratory of Advanced Laser Materials and Devices, School of Physics and Electronic Engineering, Jiangsu Normal University, Xuzhou, 221116, China
eChemistry and Chemical Engineering College, Xinjiang Agricultural University, Urumqi 830052, China. E-mail: lzq@xjau.edu.cn
First published on 3rd November 2025
Ferroelastics, as a key branch of ferroic materials, are vital for mechanical switches, energy conversion, etc. While research on ferroelastic organic–inorganic hybrid perovskites has laid a preliminary foundation, in-depth exploration is still needed to expand their material system and functional integration. In this work, we constructed two perovskite ferroelastics using cyclobutylmethanaminium (CBA+) as the cationic template. First, the lead-based (CBA)2PbBr4 exhibits a ferroelastic phase transition at Curie temperature (Tc) = 383.4 K (mmmF2/m). To address lead toxicity, we developed the lead-free double perovskite (CBA)4AgBiBr8 (Tc = 346.7 K, 4/mmmF2/m), which not only achieves lead-free design and increases domain states from 2 to 4, but also features a narrower bandgap (2.22 eV vs. 2.97 eV in (CBA)2PbBr4). Most notably, (CBA)4AgBiBr8 shows thermochromic behavior, which is rarely observed in hybrid double perovskites. These findings not only expand the perovskite ferroelastic family but also provide a strategy for integrating lead-free design with functional properties like thermochromism.
The homo-valent and hetero-valent replacement at inorganic sites can be used to construct lead-free perovskites. The homo-valent replacement refers to the use of metal ions with a +2 valence state (Ge2+, Sn2+, Cd2+, Ba2+, etc.),16–20 which largely maintain good physical properties, but their chemical richness and stability are often questioned. In comparison, the hetero-valent replacement adopts ion-splitting (A2BIBIIIX6) and ordered vacancies (A3□BIIIX9, A2□BIVX6, etc., □ refers to the vacancy), wherein BI-site cations mainly include alkali metals and group IB elements, BIII-site cations are abundant and can be located in group A and group B, and X at the corner comprises halogens, CN− and NO3−.21–27 It can not only maintain charge neutrality, but also provide a platform for diverse properties, showing better potential waiting for us to explore.28–33
Ferroelastics, which are the sister ferroic materials of ferroelectrics and ferromagnetics, can exhibit two or more strain states without mechanical stress, and under the application of mechanical stress, the strain state can transform from one to another.34 This material plays an important role in fields such as mechanical switches, shape memory, and the enhancement of key material properties (e.g., piezoelectric properties)—such application prospects have spurred research progress in ferroelastic materials, including organic–inorganic hybrid perovskites.7,21–24,35–39 As a prominent example of the hetero-valent replacement, two-dimensional double perovskites (A4BIBIIIX8) have witnessed significant research expansion in various fields in recent years.40–42 However, there are few reports on the research of two-dimensional double perovskite ferroelastics, especially in their design and construction and multifunctional integration. Considering the demand for such multifunctional integration, thermochromism—as another functional property that could enrich the multifunctional potential of perovskites—is worth attention: in organic–inorganic hybrid perovskites, thermochromic behavior has been reported in multiple systems such as lead-based and copper-based ones (e.g., MHy2PbBr4,12 MHyPbBr3,43 PEA2PbBr4,44 [3,3-difluorocyclobutylammonium]2CuCl4,45 (BEA)2CuCl4,46 etc.), yet for double perovskites, this property is only found in inorganic systems (i.e., Cs2AgBiBr6 and Cs2NaFeCl6) and a single hybrid double perovskite case ((H2MPP)2[BiAgI8]).47–49 This obvious research gap further motivates us to explore thermochromism-integrated 2D double perovskite ferroelastics.
In 2022, our team successfully constructed a hybrid double perovskite ferroelastic (DPA)4AgBiBr8 (DPA+ = 2,2-dimethylpropan-1-aminium) with high Curie temperature (Tc), using DPA+ as a cation and (AgBr6)5− and (BiBr6)3− as two-dimensional inorganic frameworks.50 In this work, taking inspiration from (DPA)4AgBiBr8, two new two-dimensional perovskite ferroelastics (CBA)2PbBr4 (Tc = 383.4 K, mmmF2/m) and (CBA)4AgBiBr8 (Tc = 346.7 K, 4/mmmF2/m) were successfully constructed by selecting CBA+ (CBA+ = cyclobutylmethanaminium) that may rotate and combining it with (PbBr6)4− and (AgBr6)5− & (BiBr6)3− (Scheme 1). It is worth highlighting that the lead-free double perovskite ferroelastic (CBA)4AgBiBr8 not only increases the domain states from 2 to 4 but also exhibits a reduced band gap from 2.97 eV to 2.22 eV. Most notably, (CBA)4AgBiBr8 exhibits thermochromic behavior, which is rarely observed in hybrid double perovskites. The discovery of these two double perovskite ferroelastics expands the ferroelastic family. Meanwhile, it offers a viable approach for screening appropriate organic cations and inorganic components, which can spur the exploration of more perovskite ferroelastic materials with functional properties like thermochromism.
![]() | ||
| Fig. 1 The two-dimensional stacking structures (a and b), DSC profile (c and d), and dielectric constant at 1 MHz (e and f) of (CBA)2PbBr4 and (CBA)4AgBiBr8. | ||
The inorganic frameworks of the two-dimensional hybrid perovskite exhibit the characteristics of an ordered structure and can provide enough space for charge balanced organic cations to realize the possible thermal motion of molecules, which provides the possibility of inducing the order–disorder structural phase transition and leading to the transformation of physical properties. For this, differential scanning calorimetry (DSC) (Fig. 1c and d) and temperature-dependent dielectric measurements (Fig. 1e and f) of (CBA)2PbBr4 and (CBA)4AgBiBr8 were conducted. As shown, (CBA)2PbBr4 and (CBA)4AgBiBr8 exhibit phase transition behavior including reversible thermal and dielectric anomalies at 383.4 K and 346.7 K, respectively. We further calculated N (the ratio of the number of equivalent orientations in the high-temperature phase) for both via the Boltzmann equation ΔS = R
ln
N (R is a gas constant): N = 5.1 for (CBA)2PbBr4, which indicates that CBA+ cations are in a disordered state and the (PbBr6)4− exhibits distortion in the high-temperature phase; in sharp contrast, N = 51.8 for (CBA)4AgBiBr8—this much higher value than that of (CBA)2PbBr4 suggests that its organic cations tend to be in a more disordered state and the (AgBr6)5−/(BiBr6)3− inorganic framework undergoes more significant deformation.
Unlike the space group Cmce (No. 64) of (CBA)2PbBr4 in the high-temperature phase, (CBA)4AgBiBr8 crystallizes on I4/mmm (No. 139) of the tetragonal crystal system with cell parameters a = b = 5.818(2) Å, c = 28.386(15) Å and volume = 960.8 (8) Å3 at 347 K (Table S2). Before the phase transition, the organic part exhibits an ordered molecular configuration and the inorganic skeleton exhibits distortion and deformation, possessing fewer symmetric elements (Fig. 2a–c). After the phase transition, the CBA+ cations are located at symmetric sites featuring 4-fold rotoinversion axes and multiple groups of mirror and glide planes (Fig. 2d). For inorganic frameworks, the Ag+ and Bi3+ in the high temperature-phase cannot be distinguished, manifested in a co-occupied form of Ag+ & Bi3+, with octahedra exhibiting highly symmetric shapes (Fig. 2e, f and Table S6), a feature similar to that in inorganic double perovskites Cs2AgBiBr6 and Cs2AgBiCl6.
According to the space group information in the high-temperature and low-temperature phases, (CBA)2PbBr4 and (CBA)4AgBiBr8 can be classified as ferroelastic phase transitions with the Aizu notations of mmmF2/m and 4/mmmF2/m, respectively. The basic stacking structures of (CBA)2PbBr4 and (CBA)4AgBiBr8 in the ferroelastic and paraelastic phases are depicted in Fig. S2g, h, and 2g and h, respectively. In addition, the cell relationships of (CBA)2PbBr4 and (CBA)4AgBiBr8 in the ferroelastic and paraelastic phases are shown in Fig. S2i and 2i.
As one of the triggers of the ferroelastic phase transition of these two, CBA+ cations exhibiting disordered multi-directional states may undergo rotational motion in the paraelastic phase similar to DPA+ cations in (DPA)4AgBiBr8. Regarding this, the organic cations in the asymmetric units of (DPA)4AgBiBr8, (CBA)2PbBr4 and (CBA)4AgBiBr8 were appropriately rotated and modeled (Fig. S3–S5), and the corresponding rotational energy was calculated (Fig. S6). The results show that the maximum energy required for the rotation of CBA+ cations in (CBA)2PbBr4 is 6.99 eV, which is greater than that of DPA+ (4.92 eV) in (DPA)4AgBiBr8 and CBA+ (3.78 eV) in (CBA)4AgBiBr8. This calculation is consistent with the phase transition temperature shown by DSC and temperature-dependent dielectric measurements, which indirectly reflects the possibility of cation rotation and the rationality of modeling and calculation.
The eight symmetric elements of (CBA)2PbBr4 in the paraelastic phase are E, C2,
,
, i, σh, σv and
, and the four symmetric elements in the ferroelastic phase are E, C2, i and σh. The possible orientation state is q = 8/4, i.e., 2. The factor 2 arises because the lost operations (
,
, σv, and
) form conjugate pairs, breaking the a–c shear-direction freedom. This creates two domains with opposed strains, separated by domain walls in the (010) plane—which contains both a- and c-axes (Fig. 3b). By observing the thin film of (CBA)2PbBr4 in the ferroelastic phase, the striped domain walls can be clearly seen. As the temperature gradually increases and exceeds Tc = 383.4 K, the ferroelastic domains will suddenly disappear, which can be understood as the disappearance of spontaneous strain in the paraelastic phase. But as the temperature decreases, the ferroelastic domains of (CBA)2PbBr4 will gradually appear again (Fig. 3c). Unlike (CBA)2PbBr4, the point group of (CBA)4AgBiBr8 in the paraelastic phase has sixteen symmetric elements: E, 2C4, C2,
,
, i, 2S4, σh, 2σv and 2σd, and the ferroelastic phase includes four symmetric elements containing E, C2, i and σh. The possible orientation state is q = 16/4, that is, 4. This symmetry breaking generates 4 possible orientation states separated by two domain walls: regions with purely horizontal walls (∥b-axis, σv-bound) and regions with purely vertical walls (∥a-axis, σd-bound), meeting only at T-junctions. Horizontal domain walls result from broken σv mirrors that were perpendicular to the a-axis, whereas vertical walls stem from broken σd mirrors that were aligned along [110] and [1
0] directions in the tetragonal system (Fig. 3d). Therefore, in the ferroelastic phase, we observe T-junctioned orthogonal domain walls featuring exclusively T-type intersections without any cross-shaped overlaps as visually distinct horizontal and vertical stripes. Similarly, as the temperature exceeds the Curie temperature of 346.7 K, ferroelastic domains will suddenly disappear and then reappear as the temperature decreases (Fig. 3e).
Notably, the domains separated by domain walls and the “disappearance of strain-related domain contrast” at Tc are direct consequences of spontaneous strain—a core physical quantity for ferroelastics explicitly defined by Aizu.34,51 As Aizu elaborated, it refers to the ferroelastic phase's intrinsic lattice distortion relative to its high-symmetry paraelastic “prototype” phase, distinct from temperature-driven reversible thermal expansion. Spontaneous strain stems from symmetry breaking at Tc (vanishing above Tc when the prototype's symmetry is restored) and is described by a symmetric second-rank tensor (εij) inheriting both phases' point-group symmetries.51,52 This tensor links microscopic distortion to macroscopic domains—for example, (CBA)2PbBr4's striped domains come from spontaneous strain tensor differences between its two orientation states, which create domain boundary-forming strain contrast.
The spontaneous strain tensor can be calculated using the following matrix51,53,54 (1) according to their Aizu notations of mmmF2/m and 4/mmmF2/m from a high-symmetry phase to a low-symmetry phase:
![]() | (1) |
![]() | (2) |
For (CBA)2PbBr4 (mmmF2/m), its paraelastic prototype (388 K, Cmce) has aPP = 29.24 Å, bPP = 8.247 Å, cPP = 8.249 Å, αPP = βPP = γPP = 90°, and Z = 2, matching the ferroelastic phase's Z = 2. We normalize aPP to aPP,norm = aPP/2 = 14.62 Å (aligning with the ferroelastic phase's aFP).55 Key strain components include tensile strains (ε11 ≈ −0.0192, ε22 ≈ 0.0051, and ε33 ≈ 0.0021) and shear strain (ε12 = ε21 ≈ −0.0853), with other components zero (allowed by 2/m symmetry) in the formulae (1).
For (CBA)4AgBiBr8 (4/mmmF2/m), its paraelastic prototype (347 K, I4/mmm) has aPP = bPP = 5.818 Å, cPP = 28.386 Å, αPP = βPP = γPP = 90°, and Z = 1. We scale aPP to
(matching the ferroelastic phase's Z = 2). Strain components include tensile strains (ε11 ≈ −0.0048, ε22 ≈ 0.005, and ε33 ≈ −0.0616) and shear strain (ε13 = ε31 ≈ −0.0698), with other components zero (forbidden by 4/mmmF2/m symmetry breaking) in the formulae (1). Eventually, by substituting the parameters in the ferroelastic and paraelastic phases in formulae (1) and (2) (see the SI for details), the total spontaneous strain εss of (CBA)2PbBr4 and (CBA)4AgBiBr8 is estimated to be 0.124 and 0.131, respectively. For comparison, reported εss values of other two-dimensional organic–inorganic hybrid perovskite ferroelastics fall in a similar range: 0.191 ([C4H9N]2[PbBr4]), 0.16 ((3-FC6H5CH2CH2NH3)2[CdCl4]), 0.156 ((DPA)4AgBiBr8) and 0.134 ([C7H16N]2[SnI4]).50,56–58
In addition, the evolution of ferroelastic domain structures of (CBA)2PbBr4 and (CBA)4AgBiBr8 was recorded under two conditions: first, during heating/cooling cycles. The temperatures at which the domains disappear and recover are completely consistent with the results from DSC, dielectric, and single crystal X-ray diffraction measurements, showing a clear transformation (Videos S1 and S2, SI). Second, under the application of mechanical stress. Application of stress to the green-ellipsed region leads to a distinct change in ferroelastic domains within the red-ellipsed region, indicating a corresponding change in the strain state (Fig. S7 and S8, SI; note that unlike organic materials, these organic–inorganic hybrid perovskites exhibit brittleness similar to inorganic materials, which may cause slight crystal damage in local areas during mechanical stress application, as observed in the green-ellipsed region).
![]() | ||
| Fig. 4 (a) The thermochromic behavior of (CBA)4AgBiBr8 during heating. The temperature-dependent UV-vis absorbance spectra (b) and corresponding bandgap (c) of (CBA)4AgBiBr8. | ||
In order to explore the thermochromism behavior of (CBA)4AgBiBr8, the temperature-dependent UV-vis absorbance spectra were collected over the range of 300–400 K (Fig. 4b and S10). When the temperature increases, the absorption band shows a remarkable red shift from 300 to 400 K: the absorption cut-off edge gradually shifts from 519 nm at 300 K to 537 nm at 400 K. When the temperature is cooled back to 300 K, the absorption cut-off edge recovers to 519 nm, confirming that this hybrid double perovskite exhibits excellent thermochromic reversibility, which is consistent with the observations in Fig. S9. Calculation of the corresponding bandgap values shows that the bandgap narrows gradually with increasing temperature, from 2.22 eV at 300 K to 2.11 eV at 400 K, representing a total bandgap narrowing of 112 meV (Fig. 4c). This redshift is attributed to the reduced tilting and distortion of the inorganic octahedra with increasing temperature. Notably, the bandgap changes more rapidly near the phase transition temperature, with a ∼30 meV decrease occurring in this range that is greater than the ∼20 meV of MHy2PbBr4 and smaller than the ∼55 meV of MHyPbBr3,12,43 which can be attributed to the symmetrization of the structure near Tc.43,45
Furthermore, the Raman spectroscopy measurements of the (CBA)4AgBiBr8 were conducted using a 532 nm laser as the excitation source and covering a variable temperature range from 293 K to 413 K (Fig. S11a and b). At 293 K, within the 100–200 cm−1 wavenumber range, two prominent peaks are observed at 141.6 cm−1 and 162.7 cm−1, both assigned to the stretching vibrations of the inorganic octahedra in the perovskite structure.59–63 As the temperature increases to 413 K, these two main peaks exhibit an abnormal blueshift, shifting to 143 cm−1 and 168.8 cm−1 respectively, which is contrary to the expected redshift that results from thermal expansion weakening chemical bonds (Fig. S11c–f). This phenomenon is mainly due to high-temperature-induced disordering of organic CBA+ cations, which weakens their directional hydrogen bonds with Br− in the inorganic octahedral layer, reducing the degree of octahedral tilting and distortion and thus elevating the phonon frequency.59,63 Meanwhile, this structural regularity is linked to the phase transition-induced statistical co-occupation of Ag+ and Bi3+: as Ag has a much smaller atomic mass than Bi, the substitution of lighter Ag+ in co-occupied octahedra further boosts the phonon vibrational frequency, causing the blueshift.
Using the VASP, the valence band maximum (VBM), the conduction band minimum (CBM) and corresponding partial density of states (PDOS) of (CBA)2PbBr4 and (CBA)4AgBiBr8 were calculated to discuss the electronic structure of Pb-based and Ag&Bi-based perovskites (Fig. 5c–f, S12 and S13). Among them, (CBA)2PbBr4 exhibits a direct bandgap, similar to previously reported Pb-based perovskites, while (CBA)4AgBiBr8 exhibits an indirect bandgap. As shown in Fig. 5e, the bandgap of (CBA)2PbBr4 is mainly contributed by inorganic components, with the valence band maximum (VBM) dominated by Pb 6s and Br 4p orbitals and the conduction band minimum (CBM) dominated by Pb 6p orbitals—consistent with the electronic structure characteristics of Pb-based perovskites summarized earlier. Similarly, the bandgap of (CBA)4AgBiBr8 is mainly derived from its inorganic part: the VBM is primarily composed of Ag 4d and Br 4p orbitals and the CBM is dominated by Bi 6p and Br 4p orbitals (Fig. 5f). In addition, in the newly constructed (CBA)2PbBr4 and (CBA)4AgBiBr8, there is a large overlap between the H 1s, N 2p, and C 2p states of organic CBA+ cations, indicating a strong interaction within the organic component (especially covalent interactions among H 1s, N 2p, and C 2p states). These organic states do not directly participate in the composition of the VBM and CBM, but their interactions with inorganic frameworks can affect the differences in band gaps.
The data supporting this article have been included as part of the supplementary information (SI). Supplementary information: Fig. S1–S14, Tables S1–S7, and Videos S1 and S2. See DOI: https://doi.org/10.1039/d5sc05158d.
gor, D. Stefańska, J. K. Zarȩba and A. Sieradzki, Chem. Mater., 2020, 32, 1667–1673 CrossRef.| This journal is © The Royal Society of Chemistry 2025 |