Open Access Article
Yueyue
He
a,
Shufang
Wu
a,
Xiaofei
Li
a,
Qi
Wang
a,
Ruifang
Zhao
a,
Lin
Pan
a,
Chengbing
Qin
bd,
Xian-Ming
Zhang
*c and
Dongying
Fu
*ab
aInstitute of Crystalline Materials, Shanxi University, Taiyuan, Shanxi 030006, P. R. China. E-mail: dyfu@sxu.edu.cn
bState Key Laboratory of Quantum Optics Technologies and Devices, Shanxi University, Taiyuan, Shanxi 030006, P. R. China
cCollege of Chemistry, Key Laboratory of Interface Science and Engineering in Advanced Material, Ministry of Education, Taiyuan University of Technology, Taiyuan, Shanxi 030024, P. R. China
dInstitute of Laser Spectroscopy, Collaborative Innovation Center of Extreme Optics, Shanxi University, Taiyuan, Shanxi 030006, P. R. China
First published on 2nd October 2025
The advantages of molecular ferroelectrics lie in the “designability” and “multifunctionality”, and the molecular-level regulation ability has opened up a brand-new dimension for ferroelectric materials. Non-covalent interactions play a crucial role in the construction of molecular ferroelectrics. However, there remain significant challenges in balancing the strength and reversibility of non-covalent interactions, as well as achieving long-range ordered arrangements. Therefore, a systematic study of non-covalent interactions in the structure is the key to construct high-performance molecular ferroelectrics. Here, we introduced halogenated amines with large dipole moments into adjacent inorganic layers to regulate the non-covalent interactions in the structure, thereby inducing the generation of ferroelectricity. Through a halogen substitution strategy to introduce chlorine (Cl) atoms on PA+ (n-propylaminium) cations, hybrid perovskite photoferroelectric semiconductor (Cl-PA)2PbBr4 (Cl-PA+ is 3-chloropropylaminium) with large piezoelectric response (d33 = 36 pC/N) and high Curie temperature (Tc = 365 K) was obtained. Compared with non-ferroelectric (PA)2PbBr4 (μPA = 1.2 D), the larger dipole moment (μCl-PA = 3.3 D) and the directional ordered arrangement of Cl-PA+ in (Cl-PA)2PbBr4 synergistically induce its ferroelectricity. More importantly, when Cl replaces H, it affects the hydrogen bond network between the organic cation and the inorganic layer, enhancing the dynamic freedom of the Cl-PA+ cations, making the structure of (Cl-PA)2PbBr4 more prone to phase transitions when the temperature changes. The hydrogen bonding and halogen–halogen interactions in (Cl-PA)2PbBr4 lead to the directional and ordered arrangement of Cl-PA+ cations, breaking the centrosymmetric structure and synergistically promoting the generation of ferroelectricity. This work has confirmed the significance of non-covalent interactions in the construction of ferroelectrics.
N-methyl-N′-diazabicyclo [2.2.2] octonium) is regulated by hydrogen bonding, with a modulation amplitude of up to 10 times.15
Although we are aware of the significance of non-covalent interactions in the construction of molecular ferroelectrics, research in this area is still very scarce. In particular, further research is needed to design high-performance molecular ferroelectrics by utilizing the synergistic effects of multiple non-covalent interactions. Constructing molecular ferroelectrics by leveraging the synergistic effect of non-covalent interactions involves integrating multiple weak interactions to achieve precise control of molecular arrangement, enhancement of polarization, and optimization of dynamic response performance.16,17 The synergistic effect of non-covalent interactions essentially lies in the molecular engineering of “combining rigidity and flexibility”, which can lock the molecular polarity arrangement through highly directional forces (hydrogen bonds or halogen bonds), leading to symmetry breaking.18–21 Meanwhile, the weak interaction of the flexible chain as a dynamic response unit provides a low-energy barrier flipping path, endowing the material with reversibility. However, the main challenge lies in how to balance the intensities and directions of various interactions, and avoid instability of the structure due to mutual competition.22 We can take advantage of the designability of molecular ferroelectrics to modify organic amine units at the molecular level and construct multiple non-covalent interactions that work collaboratively in the system. The in-depth analysis of the selective regulation of non-covalent interactions and the structure–activity relationship is beneficial for us to accurately construct multifunctional molecular ferroelectrics.
Herein, we regulated the non-covalent interactions in the structure by modifying the organic components. Starting from non-ferroelectric (PA)2PbBr4 (PA+ is n-propylaminium), we introduced halogenated amines Cl-PA+ (Cl-PA+ is 3-chloropropylaminium) with large dipole moments into adjacent inorganic layers to regulate non-covalent interactions in the structure, thereby inducing ferroelectricity in (Cl-PA)2PbBr4. After replacing the H atom in the PA+ with a Cl atom, the dynamic response ability of the Cl-PA+ cation and the lattice strain of the inorganic framework were significantly enhanced by increasing the molecular polarity and reconstructing the hydrogen bond network. Under the synergistic effect of hydrogen bonding and halogen–halogen interaction, Cl-PA+ cations undergo directional ordered rearrangement, resulting in symmetry breaking in the structure of (Cl-PA)2PbBr4. The weakening of hydrogen bonds between the inorganic and organic parts from (PA)2PbBr4 to (Cl-PA)2PbBr4 gave Cl-PA+ cations a higher degree of freedom, and the ferroelectricity in (Cl-PA)2PbBr4 relies on the synergistic effect between hydrogen bonds and halogen–halogen interactions. This result confirms that non-covalent interactions can not only adjust the molecular arrangement and promote symmetry breaking in the construction of molecular ferroelectrics, but also regulate the ordered-disordered transition of interlayer organic units, thereby triggering ferroelectric phase transition. Furthermore, possibly due to the existence of halogen–halogen interactions, (Cl-PA)2PbBr4 exhibits significantly better piezoelectric properties than other similar molecular ferroelectrics, with its d33 coefficient reaching up to 36 pC/N.
The stochiometric ratio 2
:
1 of the Cl-PA+ cations and Pb(CH3COO)2·3H2O was used to synthesize (Cl-PA)2PbBr4 in the refrigerator with HBr solution (48%) (Experimental section), and the structure of (PA)2PbBr4 has been reported in previous literature.28Fig. 1 is the structure of (PA)2PbBr4 and (Cl-PA)2PbBr4 at room temperature. The crystal structures of (PA)2PbBr4 and (Cl-PA)2PbBr4 are composed of infinite layers of corner-sharing PbBr6 octahedra, which are separated by bilayers of the organic cations PA+ and Cl-PA+, respectively. This arrangement forms a typical 2D monolayer Ruddlesden–Popper organic–inorganic hybrid perovskite structure.29 However, there is a significant difference in the arrangement of PA+ cations and Cl-PA+ cations between adjacent inorganic layers in the two compounds. In (PA)2PbBr4 (Fig. 1a), the dipole direction of each PA+ cation will exhibit a reverse dipole at the corresponding position of the symmetry center, resulting in a vector sum of zero. The dipole of PA+ cations in one layer is upward, while that of the other layer at the symmetrical position is downward, resulting in the overall net dipole being zero. Meanwhile, the octahedra may also be arranged in an anti-parallel inclined or rotational symmetrical manner, such that the local dipole of each octahedron is counterbalanced by the opposite dipole of the adjacent octahedron. The dipole moment vector of organic cations in the structure and the net dipole moment of the entire inorganic skeleton are both zero. Therefore, in both the organic and inorganic parts, the dipole moments cancel each other out at the microscopic level. Fig. 1a shows that (PA)2PbBr4 crystallizes in the centrosymmetric Pbca space group (point group mmm) at 296 K. In contrast to (PA)2PbBr4, (Cl-PA)2PbBr4 has an asymmetric dipole arrangement in the structure. The polarization generated by the inorganic skeleton and the polarization generated by all organic cations will not cancel each other, resulting in a large net spontaneous polarization (Ps) at the macroscopic level. At 300 K, (Cl-PA)2PbBr4 crystallized in the non-centrosymmetric polarity Cmc21 space group (point group mm2) (Fig. 1b). The generation of net dipole in (Cl-PA)2PbBr4 is achieved through the oriented and ordered arrangement of the Cl-PA+ organic layer and structural distortion of the inorganic layer. Fundamentally, the reason for the structural differences between the two compounds is the substitution of the H atom on PA+ cations by the Cl atom, which affects the non-covalent interactions in the structure. Regarding the analysis of how interlayer molecular modification can regulate non-covalent interactions within the structure to induce ferroelectricity, we will provide a detailed discussion in the following sections.
In addition to constructing polar structures, structural phase transitions serve as a key mechanism for triggering symmetry breaking, thereby inducing the generation of spontaneous polarization.5 Before testing the phase transition of (PA)2PbBr4 and (Cl-PA)2PbBr4, we verified the phase purity of the two compounds through powder X-ray diffraction (PXRD). The experimental PXRD results of (Cl-PA)2PbBr4 and (PA)2PbBr4 are in agreement with the simulated patterns, indicating its good crystallinity and high phase purity (Fig. S1a and b). We also noticed that the change of interlayer cations from PA+ to Cl-PA+ had little effect on the PXRD diffraction peaks of the compounds. This smaller difference is mainly due to larger Cl-PA+ cations occupying more space, which increases the interlayer spacing between inorganic layers and causes the corresponding diffraction peaks to shift towards lower 2θ angles (Fig. S1c). Differential scanning calorimetry (DSC) tests were employed to investigate the phase transition behavior of (PA)2PbBr4 and (Cl-PA)2PbBr4. The DSC thermogram revealed two pairs of reversible peaks during both heating and cooling processes, indicating two reversible phase transitions at temperatures of 348/341 K and 365/357 K in (Cl-PA)2PbBr4 (Fig. 2a). In contrast, there was no phase transition observed in (PA)2PbBr4 with temperature changes (Fig. S2). Subsequently, we also obtained the structure of (Cl-PA)2PbBr4 at different temperatures through variable-temperature single-crystal X-ray diffraction (SCXRD). At room temperature (300 K), (Cl-PA)2PbBr4 shows a polar structure with Cmc21 space group, which belongs to one of 10 polar point groups (mm2), and we mark it as the room temperature phase (RTP) (Fig. 3a). When the temperature is raised to 355 K (Fig. 3b), the crystal structure of (Cl-PA)2PbBr4 remains in the polar Cmc21 space group, and we label this state as the intermediate temperature phase (ITP). From Fig. 3a and b, it can be seen that Cl-PA+ cations are in an ordered state below 365 K. Upon further heating to 375 K, the structure of (Cl-PA)2PbBr4 changes to the centrosymmetric orthorhombic space group of Pbca (Fig. 3c) and the Cl-PA+ cations exhibit obvious disorder, and this is called the high temperature phase (HTP). For the treatment of high-temperature crystal structures of (Cl-PA)2PbBr4, we adopted the same method as other similar perovskites.9,30,31 The other crystallographic data of (Cl-PA)2PbBr4 at different temperatures are all listed in Tables S1–S8. We know that in molecular ferroelectrics, the high-temperature disorder and low-temperature order transition of organic units are crucial for the generation and regulation of ferroelectricity.32–34 At HTP, Cl-PA+ cations are in a dynamic disordered state due to thermal motion, resulting in a highly symmetrical structure of (Cl-PA)2PbBr4 and the random arrangement of local dipoles prevents the formation of macroscopic polarization. When the temperature drops below the Curie temperature (Tc = 365 K), the thermal motion of Cl-PA+ cations is suppressed, and the orientation or position tends towards long-range ordered arrangement. This orderliness breaks the original symmetry (such as changing from centrosymmetry to polar space group) and induces the generation of spontaneous polarization. The symmetry breaking in (Cl-PA)2PbBr4 conforms to Aizu notation of mmmFmm2, which satisfies the requirements of 88 species ferroelectric-type phase transition.35
Furthermore, by measuring the variation of the second-harmonic generation (SHG) signal with temperature, we can also detect the process of symmetry breaking in (Cl-PA)2PbBr4. At temperatures above 365 K, (Cl-PA)2PbBr4 is in the centrosymmetric phase, and the SHG signal should approach zero, which indicates the SHG-OFF state. When the temperature drops below Tc, the material undergoes symmetry breaking and transforms to a polar space group, resulting in a significant enhancement of the SHG signal, which indicates the SHG-ON state (Fig. 2b). We also compared the SHG intensity of (Cl-PA)2PbBr4 and KDP (KH2PO4) at the same particle size, and the results showed that the SHG intensity of (Cl-PA)2PbBr4 was 0.55 times that of KDP (Fig. S3), which is comparable to that of other hybrid perovskites ferroelectrics, such as [CH3(CH2)3NH3]2(CH3NH3)Pb2Br7 (0.4 × KDP),36 (C4H9NH3)2CsPb2Br7 (0.5 × KDP)37 and [(CH3)3NCH2I]PbI3 (0.65 × KDP).38 Dielectric anomaly is a common characteristic of ferroelectric phase transitions. Near the Tc of ferroelectric materials, the dielectric constant reaches its peak and exhibits significant dielectric anomalies. Dielectric measurements of (Cl-PA)2PbBr4 (Fig. 2c) showed changes in the dielectric constant near 348 K and 365 K, which is consistent with the results of variable-temperature SCXRD testing and DSC testing. The observation of thermal hysteresis in both heating and cooling processes suggests that these phase transitions are first-order transitions.5 Through the above structural analysis and the characterization of related physical properties, (Cl-PA)2PbBr4 has already met the conditions of polar structure, reversible polarization path, and dynamic phase transition ability, thus proving that it is a potential ferroelectric. The ferroelectric hysteresis loop is the core experimental curve that characterizes its ferroelectricity, reflecting the nonlinear relationship between polarization intensity (P) and external electric field (E). Here, we measured the P–E hysteresis loops of (Cl-PA)2PbBr4 along the polar c-axis direction of crystal samples at different temperatures. As shown in Fig. 2d, the rectangular P–E hysteresis loops of (Cl-PA)2PbBr4 collected at 330 K afford the Ps values of ∼1.3 μC cm−2, which decreased with increasing temperature, consistent with the transition from the ferroelectric phase to paraelectric phase.
To gain insights into the ferroelectricity of (Cl-PA)2PbBr4, we compared and analyzed the non-covalent interactions in the structure of (PA)2PbBr4 and (Cl-PA)2PbBr4. After the Cl atom replaces the H atom on PA, the Cl-PA has a relatively large dipole moment (μCl-PA = 3.3 D), which is a favorable factor for constructing molecular ferroelectrics. This is because the larger the dipole moment of the molecule, the more significant the charge separation within the molecule, and it will provide a stronger microscopic contribution for macroscopic polarization. Although organic amines have large dipole moments that are advantageous for constructing molecular ferroelectrics, it is still necessary to combine appropriate crystal structure design to enable these dipole moments to be arranged in an orderly manner. It is of great significance that precisely controlling the arrangement of organic amine molecules through non-covalent interactions is the core strategy for designing high-performance molecular ferroelectrics. In the crystal structure, the synergy and competition of weak interactions such as hydrogen bonding, VDW interactions, and halogen–halogen bonds can endow materials with dynamic response, structural flexibility, and functional tunability.
Hirshfeld surface analysis transforms the complex molecular interactions into visual and quantifiable data, providing a unique perspective for the study of non-covalent interactions.39–41 The Hirshfeld surface and related 2D fingerprint plot were obtained based on CIF files through the CrystalExplorer program. Fig. S4a and S5a are the Hirshfeld surface of PA+ and Cl-PA+ cations and the intensive red spots indicate relatively strong interactions (the distance smaller than VDW distance), the blue regions correspond to long contacts (the distance exceeds VDW distance), the white regions correspond to the distance of contacts almost equal to VDW distance.39 Hirshfeld surface and 2D fingerprint analyses reveal that the proportion of H–Br interactions in (PA)2PbBr4 and (Cl-PA)2PbBr4 is 43.1% (Fig. S4b) and 35.7% (Fig. S5b), respectively. Especially, the N–H⋯Br and C–H⋯Br hydrogen bonds in the structure are significantly weakened after halogen substitution (Fig. S6, N–H⋯Br for (PA)2PbBr4 is 52.3% and for (Cl-PA)2PbBr4 it is 48.4%, Fig. S7, C–H⋯Br for (PA)2PbBr4 is 17.5% and for (Cl-PA)2PbBr4 it is 14.3%), indicating that the interaction between the inorganic skeleton and the interlayer organic amine cations in (Cl-PA)2PbBr4 is relatively weak, which gives the Cl-PA+ cation a higher degree of freedom of movement. The strength of hydrogen bonds between interlayer cations and inorganic layers in (PA)2PbBr4 and (Cl-PA)2PbBr4 can also be confirmed from the crystal structure in Fig. 4a–d. The penetration depth of Cl-PA+ cations (0.476 Å) in the adjacent inorganic layer is smaller than that of PA+ cations (0.511 Å), which demonstrated the PA+ cations submerge deeper into the inorganic layer in (PA)2PbBr4. Therefore, PA+ cations are closer to Br in the inorganic layer, forming shorter hydrogen bonds (Fig. 4b) and narrower interlayer space (6.574 Å, Fig. 4a). For N–H⋯Br hydrogen bonds, the shortest H–Br distance in (PA)2PbBr4 is 2.521 Å, corresponding to a ∠N–H⋯Br angle of 170.8°, while the shortest H–Br distance in (Cl-PA)2PbBr4 is 2.575 Å, corresponding to a ∠N–H⋯Br angle of 160.3° (Fig. S8). The strong N–H⋯Br hydrogen bonds in (PA)2PbBr4 may limit the vibrational degrees of freedom of PA+ cations, leading to a weakened structural dynamic response of the material under temperature changes, which also explains why (PA)2PbBr4 did not undergo phase transition during temperature changes. In addition, the strong hydrogen bonds in (PA)2PbBr4 may limit the vibration or distortion of octahedra in the inorganic layer, stabilizing the structure in a more symmetrical state. That's because hydrogen bonding enhances the interaction between the organic and inorganic layers, strengthens interlayer coupling, makes the structure of the inorganic layer more stable, and limits the distortion of octahedra. The degree of distortion of PbBr6 octahedra (Δd) in (PA)2PbBr4 (2.40 × 10−6) is smaller than that of (Cl-PA)2PbBr4 (1.36 × 10−4), indicating the rigid structure in (PA)2PbBr4. The specific calculation formula of Δd is as follows:42
![]() | (1) |
We know that non-covalent interactions are extremely temperature sensitive. In (Cl-PA)2PbBr4, there are fundamental differences in the non-covalent interactions between the RTP (ferroelectric phase) and the HTP (paraelectric phase), which are the driving forces behind ferroelectric phase transitions and spontaneous polarization. In the ferroelectric phase, the system is in an ordered state with the lowest energy. All non-covalent interactions reach their optimal and most stable configurations, and work together to stabilize the ferroelectric state. Moreover, halogen bonds and hydrogen bonds work together to determine the tilt direction and angle of organic cations. The halogen bond acts as a “lock buckle” that holds the other end of the organic cation in place, thereby “tightening” the entire cation together with the hydrogen bond, directing it in a specific direction and generating a net dipole moment. The most prominent feature of the ferroelectric phase is that hydrogen bonds and halogen bonds are no longer independent, but synergistically control the orientation, tilt angle, and position of organic cations, thereby stabilizing polar structures with net dipole moments (Fig. S13a, b and S14a). When the temperature rises and the thermal energy gradually exceeds the energy scale of these non-covalent interactions, the system undergoes a phase transition and enters the disordered paraelectric phase. In the HTP, all non-covalent interactions are dominated by thermal motion. Organic cations undergo rapid disordered motion (rotation and oscillation, Fig. S13c and S14b). Among them, N and H atoms will experience severe orientational disorder, while C atoms will experience conformational disorder. In addition, it can be seen from the significantly increased atomic displacement parameter of halogen atoms that their position in space is very uncertain (Table S9). That is, at high temperatures, halogen atoms are not “stationary in a vague position”, but “rapidly moving between multiple possible positions”. Therefore, the hydrogen bonds and halogen bonds become very weak or only exist instantaneously, which undoubtedly exacerbates the overall disordered movement of cations (Fig. S14b). In summary, the dynamic disorder of hydrogen bonds from low temperature to high temperature and the “formation breakage” process of halogen bonds from low temperature to high temperature are one of the core mechanisms driving the ferroelectric paraelectric phase transition of (Cl-PA)2PbBr4. Moreover, the hydrogen-bond parameters of (PA)2PbBr4 and (Cl-PA)2PbBr4 at different temperatures are listed in Tables S10–S13. Meanwhile, as halogen bonds can enhance the long-range ordering of local dipoles and improve macroscopic polarizability by fixing the relative orientation of organic components, this is positively correlated with the piezoelectric coefficient d33. Therefore, the relatively large d33 of (Cl-PA)2PbBr4 (36 pC/N, Fig. S15) may be related to halogen–halogen interactions, as halogen–halogen interactions can induce non-centrosymmetric structures and dipole ordering to enhance piezoelectric activity.44 The comparison of d33 values between (Cl-PA)2PbBr4 and other hybrid perovskites is shown in Fig. S16.
Based on the semiconductor properties of (Cl-PA)2PbBr4, a stable built-in electric field formed by the spontaneous polarization of the ferroelectric phase can automatically separate electron–hole pairs generated by X-ray excitation, significantly reducing recombination losses. Compared to traditional X-ray detectors that require high voltage to drive carrier migration, this significantly simplifies the structure of the device and reduces energy consumption. Next, the X-ray detection performance of the (Cl-PA)2PbBr4 single crystal was systematically investigated. Fig. 5a shows the absorption coefficients as a function of X-ray photon energy of typical detector materials. The absorption coefficient of (Cl-PA)2PbBr4 is comparable to that of α-Se due to the high-Z element. The X-ray detectors were assembled with a typical photoconductor structure of Au/(Cl-PA)2PbBr4/Au. The mobility-lifetime (μτ) value is the “performance lever” of X-ray detector materials, which is directly related to the core indicators of the detector.45–47 As fitted in Fig. 5b, the (Cl-PA)2PbBr4 single crystal has a μτ product of 1.32 × 10−3 cm2 V−1, which is superior to that of other 2D hybrid perovskites previously reported, such as PA2MHy2Pb3Br10 (MHy is methylhydrazine, 4.72 × 10−4 cm2 V−1)48 and (2IPA)2FAPb2I7 (2IPA = 2-iodopropylammonium, FA is formamidinium, 1.2 × 10−3 cm2 V−1).49 The ratio of the photovoltaic voltage generated by ferroelectrics through the bulk photovoltaic effect to the incident light intensity reflects the intrinsic efficiency of the material in converting light energy into electrical energy. A larger photovoltaic voltage will generate a strong built-in electric field, which enhances the ability of carrier directional migration. As shown in Fig. 5c, the bulk photovoltaics of (Cl-PA)2PbBr4 through current–voltage (I–V) traces under X-ray are estimated to be 1.40 V, which is beneficial for achieving passive X-ray detection. Furthermore, the photocurrent density in Fig. 5d–g and S17 exhibits a good linear relationship with the X-ray dose rate under different biases. The sensitivity increases with the increase of the bias and even under 0 V bias the sensitivity of the device can also reach 127.2 μC Gy−1 cm−2 (Fig. 5g), which is much better than that of the device based on the (isopropylammonium)2PbBr4 ferroelectric (51.4 μC Gy−1 cm−2 at 0 V bias).50 The Limit of Detection (LoD) of an X-ray detector refers to the minimum X-ray dose or photon flux that the detector can reliably identify, usually based on the signal strength corresponding to a Signal to Noise Ratio (SNR) of ≥ 3. The LoD of 52 nGy s−1 is derived from the linear fitting with an SNR of 3 shown in Fig. 5h, which is lower than that of some hybrid perovskite-based X-ray detectors under 0 V bias (Table S14) and is far lower than the dose rate of a regular X-ray medical imaging (5.5 μGy s−1). The photostability of the device is crucial due to the phase transition, ion migration, or interface reaction induced by light waves. We measured the photostability of single crystal devices based on (Cl-PA)2PbBr4 under long-term X-ray detection. In Fig. 5i, the photocurrent exhibits no significant changes after passing through a total amount of 450 mGy X-ray dosage, demonstrating its high radiation stability. These photoelectric test results indicate that the excellent semiconductor properties and unique bulk photovoltaic effect make (Cl-PA)2PbBr4 highly promising for passive X-ray detection.
Supplementary information: structural information, PXRD patterns, synthesis process of (Cl-PA)2PbBr4, Hirshfeld surface and related 2D fingerprint plot and additional information (PDF). See DOI: https://doi.org/10.1039/d5sc05946a.
| This journal is © The Royal Society of Chemistry 2025 |