Enantiomeric hybrid high-temperature multiaxial ferroelectrics with a narrow bandgap and high piezoelectricity

Chang-Feng Wang Na Wang Lang Liu Le-Ping Miao Heng-Yun Ye Yi Zhang Chao Shi

Citation:  Chang-Feng Wang, Na Wang, Lang Liu, Le-Ping Miao, Heng-Yun Ye, Yi Zhang, Chao Shi. Enantiomeric hybrid high-temperature multiaxial ferroelectrics with a narrow bandgap and high piezoelectricity[J]. Chinese Chemical Letters, 2023, 34(8): 108051. doi: 10.1016/j.cclet.2022.108051 shu

Enantiomeric hybrid high-temperature multiaxial ferroelectrics with a narrow bandgap and high piezoelectricity

English

  • Ferroelectric semiconductor materials have attracted tremendous attention for their potential applications, such as in information storage, conversion, and self-powered detections [1,2]. Notably, ferroelectric semiconductors possess a photoelectric conversion mechanism different from that of conventional counterparts. Within ferroelectric semiconductors, the internal electric field induced by spontaneous polarization can separate effectively photogenerated charge carriers, which in turn enhances greatly the photophysical properties [3]. The mainstream of ferroelectric semiconductors is currently still dominated by conventional inorganic oxides. However, they suffer from poor light absorption resulting from a wide bandgap, such as BiFeO3 of 2.7 eV, BaTiO3 of 3.2 eV, and LiNbO3 of 3.6 eV, necessitating the development of new narrow bandgap ferroelectric semiconductors [3,4].

    Organic-inorganic hybrid halide materials have been demonstrated as excellent semiconductor materials, spearheaded by CH3NH3PbI3, due to their desirable optoelectronic performance, such as strong absorption coefficient, long carrier lifetime, and high mobility [5-8]. Moreover, organic-inorganic hybrid halide materials have been proposed as a promising platform for the development of ferroelectric materials due to their structural flexibility and variety [9,10]. Although a mass of hybrid halide ferroelectric semiconductors has been developed with unremitting efforts, most of the reported ones present a wide band gap (> 2 eV) [11-15]. In this context, substantial feasible strategies have been developed to obtain hybrid halide semiconductors with a narrow band gap, such as chemical doping, application of high-pressure, and tailoring perovskite layer thickness [4,16,17]. Among them, chemical doping and high-pressure engineering can efficiently narrow the band gap, but only to a small extent. Excess doping or pressure usually brings about phase transition, in turn losing ferroelectricity, as demonstrated by (CHA)2PbBr4−4xI4x (x = 0–1) [18]. The addition of perovskite layer thickness strategy is limited to two-dimensional (2D) perovskites [19,20]. Recently, iodide management (i.e., incorporation of iodide (I2) or triiodide (I3) into the crystal lattice) has emerged as a desirable design strategy for the achievement of narrow band gap hybrid semiconductor materials, due to their capacities for decreasing of concentration of deep-level defects and improvement of light absorption, as exemplified by (C8H11S)2Pb2I6*I2 (1.89 eV), (Me3S)2Pb5I14×2I2 (1.86 eV), (bis-(2-dimethylaminoethyl)ether)3I3Bi3I14 (1.59 eV), (4-methyl-piperidinium 4·I3·BiI6 (1.58 eV) and (n-propylammonium)4I3BiI6 (1.34 eV) [21-25]. Despite blooming advances, ferroelectric semiconductor incorporating I2 or I3 still remains a blank. For obtaining ferroelectricity, the introduction of chiral cations is one of the ideal strategies, which has been tested [26]. Because so long as the incorporation of chiral organic cations, according to the chirality transfer principle, the crystal must crystallize in the 11 chiral point groups (1, 2, 222, 4, 422, 3, 32, 6, 622, 23 and 432), five of which (1, 2, 4, 3, 6) allow ferroelectricity [26-28]. It is worth mentioning that chirality-induced ferroelectrics possess a high probability of enjoying a satisfactory piezoelectric response [29,30].

    In this work, combined with iodide management and introduction of chirality, we successfully synthesized a pair of hybrid ferroelectric semiconductors, (R-1,2-DAP·I)4·I3·BiI6 and (S-1,2-DAP·I)4·I3·BiI6 (R/S-1,2-DAP = (R/S)-(–)−1,2-diaminopropane), which exhibits a narrow bandgap of 1.56 eV comparable to that of CH3NH3PbI3 (~1.5 eV), a high Curie temperature (Tc) of 405 K, multiaxial feature and high piezoelectric coefficient (d22) of 35 pC/N.

    The single crystals of (S-1,2-DAP·I)4·I3·BiI6, (R-1,2-DAP·I)4·I3·BiI6 and racemic (Rac-1,2-DAP·I)4·I3·BiI6 were obtained by evaporation of HI aqueous solution at 323 K (Fig. S1 in Supporting information). The purity of the bulk phases of those compounds was examined by powder X-ray diffraction (PXRD) measurements (Fig. S2 in Supporting information). We carried out the circular dichroism (CD) spectra measurements in the ultraviolet-visible (UV–vis) absorption range to demonstrate the chirality of (S-1,2-DAP·I)4·I3·BiI6 and (R-1,2-DAP·I)4·I3·BiI6. As depicted in Fig. S3 (Supporting information), (S-1,2-DAP·I)4·I3·BiI6 and (R-1,2-DAP·I)4·I3·BiI6 exhibit obvious CD signals at the same wavelength, located at around 300, 361, 415, 466 nm, corresponding to those in the UV–vis absorption spectra. The CD signal responses of the (S-1,2-DAP·I)4·I3·BiI6 and (R-1,2-DAP·I)4·I3·BiI6 are opposite, revealing the enantiomorphic structure feature.

    To determine Tc of the (S-1,2-DAP·I)4·I3·BiI6 and (R-1,2-DAP·I)4·I3·BiI6, we performed differential scanning calorimetry (DSC) measurements, which is an overwhelmingly reliable method to detect the Tc of the molecular ferroelectrics [31-33]. The results of (S-1,2-DAP·I)4·I3·BiI6 and (R-1,2-DAP·I)4·I3·BiI6 are the same (Figs. S4 and S5a in Supporting information). As shown in Fig. S4, an anomalous endothermic peak appears at around 405 K in the heating run, and the corresponding exothermic peak is determined at around 394 K in the cooling run, indicating that (S-1,2-DAP·I)4·I3·BiI6 undergoes a reversible phase transition. The large thermal hysteresis (11 K) suggests that the phase transition belongs to the first-order ones [34,35]. The phase above and below the Tc is labeled as the high-temperature phase (HTP) and low-temperature phase (LTP), respectively. The enthalpy change (ΔH) and entropy change (ΔS) in the phase transition process are determined as 9.47 J/g and 50 J mol−1 K1, respectively. In light of the Boltzmann equation, ΔS = RlnN, in which R is the gas constant (8.314), and N is the ratio of the numbers of respective geometrically distinguishable orientations, the calculated N value is about 418, disclosing that the phase transition of (S-1,2-DAP·I)4·I3·BiI6 can be classified as the ordered-disordered ones, and the HTP is highly disordered. Significantly, Tc (405 K) of (S-1,2-DAP·I)4·I3·BiI6 is higher than that of BaTiO3 (393 K) and those of recently reported ferroelectrics, such as (4,4-difluorocyclohexylammonium)2PbI4, [(S/R)-N-(1-phenylethyl)ethane-1,2-diaminium]PbI4 and (n-propylammonium)2CsAgBiBr7 [15,36,37]. Similarly, (Rac-1,2-DAP·I)4·I3·BiI6 also exhibits high-temperature reversible phase transitions at 448 K (Fig. S5b in Supporting information).

    We take (S-1,2-DAP·I)4·I3·BiI6 as an example to analyze structure in the LTP and HTP because (S-1,2-DAP·I)4·I3·BiI6 and (R-1,2-DAP·I)4·I3·BiI6 are enantiomorphic each other. To disclose the mechanism of paraelectric-ferroelectric phase transition for (S-1,2-DAP·I)4·I3·BiI6, we performed the variable-temperature single-crystal X-ray diffraction measurements. Racemic (Rac-1,2-DAP·I)4·I3·BiI6 adopts a centrosymmetric space group of P21/c (point group of 2/m). In the LTP (ferroelectric phase), (S-1,2-DAP·I)4·I3·BiI6 adopts a 0D structure, and crystallizes in the monoclinic space group of P21 (point group of 2) with a cell parameter: a = 14.2307(4) Å, b = 14.1214(3) Å, c = 24.3679(6) Å, β = 105.702(3)° and V = 4714.17 Å3 (Table S1 in Supporting information). The asymmetric unit is composed of eight different geometry configurations S-1,2-DAP cations, two I3 anions, eight free I anions and two discrete BiI63– octahedral. As shown in Fig. 1a, the structure of (S-1,2-DAP·I)4·I3·BiI6 can be regarded as a sandwich structure, which consists of organic cation layers, I anion ones and ones composed of alternating I3 anions, and discrete BiI63– octahedral. By weak N–H…I and C–H…I hydrogen bonds, S-1,2-DAP cations act as bridging linkers to bond the I anion layers and layers composed of I3 anions and BiI63– octahedral (Fig. 1b). Polar structure brings about spontaneous polarization (Ps) along the [010]-direction, i.e., b-axis.

    Figure 1

    Figure 1.  The packing structure of (S-1,2-DAP·I)4·I3·BiI6 in the LTP (a) and HTP (c) (H atoms are omitted for clarity). (b) Distribution of hydrogen bonds of S-1,2-DAP cations. (d) Two-fold disordered S-1,2-DAP cations around C2 rotation axis in the HTP. (e) The relationship of the unit cells of (S-1,2-DAP·I)4·I3·BiI6 between the LTP and HTP. The red and blue edges represent the crystal lattice of the LTP and the HTP, respectively. The red arrow represents the spontaneous polarization direction in the LTP. The blue arrows represent the possible equivalent polarization directions. Only the Bi atoms are shown for clarity.

    In the HTP, (paraelectric phase), The space group changes to tetragonal group space of I4122 (point group of 422) with a cell parameter: a = 10.1224(6) Å, c = 47.470(3) Å, and V = 4863.92 Å3 (Table S1). As depicted in Fig. 1c, BiI63– octahedral become more geometrically regular, and the organic cations become disordered around the crystallographic C2 rotation axis parallel to the c-axis (Fig. 1d). The relationship of cell parameters between the HTP and LTP can be described to be bLTPaHTP + bHTP, aLTPbHTP - aHTP and cTP ≈ 1/2aHTP + 1/2cHTP - 1/2bHTP (Fig. 1e). Since the crystallographic symmetry changes from the point group of 422 (E, 2C4, C2, 2C'2, 2C"2) to 2 (E, C2), (R-1,2-DAP·I)4·I3·BiI6 and (S-1,2-DAP·I)4·I3·BiI6 belong to the ferroelectric species with Aizu nation 422F2(s) [31]. In view of the change in the number of symmetry element, the ferroelectric phase possesses four equivalent polarization directions, i.e., (R-1,2-DAP·I)4·I3·BiI6 and (S-1,2-DAP·I)4·I3·BiI6 are of multiaxial ferroelectrics (Fig. 1e).

    To confirm the paraelectric-ferroelectric phase transition of (S-1,2-DAP·I)4·I3·BiI6 and (R-1,2-DAP·I)4·I3·BiI6, we measured the second harmonic generation (SHG) signal as a function of temperature. As shown in Fig. 2a, in the LTP, the SHG signal response of (S-1,2-DAP·I)4·I3·BiI6 exhibits an active state, which is in line with the polar point group of 2. As the temperature rises to around Tc, the SHG signal intensity drops sharply, and decreases to zero above Tc. Chiral point group of 422 are not SHG-active, due to restriction of Kleinman symmetry [32,33]. This variable-temperature SHG signal response agrees with the paraelectric-ferroelectric phase transition for the 422F2(s) type.

    Figure 2

    Figure 2.  The ferroelectricity and related properties of (S-1,2-DAP·I)4·I3·BiI6. (a) The SHG signal intensity a function of temperature. (b) The ε′ in the different temperature at 1 MHz. (c) The PE hysteresis loop at 333 K.

    Usually, paraelectric-ferroelectric phase transitions are characterized by the anomalies of dielectric permittivity (ε = ε′ − iε″, in which ε′ and ε″ represent the real and the imaginary parts in several) at around Tc [34,35]. Then, we measured variable-temperature ε′ of the (S-1,2-DAP·I)4·I3·BiI6 at 1 MHz. As shown in Fig. 2b, the steplike anomaly of ε′ is observed at around 405 and 394 K in the heating and cooling run, respectively, corresponding to the DSC results, indicating that there is a reversible phase transition.

    To confirm the polarization reversal in (S-1,2-DAP·I)4·I3·BiI6, we measured the polarization-electric (PE) field hysteresis loop along the b-axis direction on a single-crystal slice at 333 K with the double-wave method at frequency of 50 Hz (The detailed process for determination of the b-axis direction can be seen in Supporting information). Notably, the measurements were carried out at relatively high temperature to facilitate reversal of Ps, which is common for molecule-based ferroelectrics [36]. As shown in Fig. 2c, the obtained current density-electric field (JE) loop exhibits two opposite peaks in the different directions of the electric field. Such JE response is due to the polarization reversal upon the electric field exceeding a particular value, known as the coercive field (Ec). The PE loop calculated from the JE loop possesses a high rectangularity, which unambiguously proves ferroelectricity of (S-1,2-DAP·I)4·I3·BiI6 (Fig. 2c). The Ps and Ec are determined to be 1.62 µC/cm2 and 11 kV/cm, respectively. The Ps of (S-1,2-DAP·I)4·I3·BiI6 is greater than those recently reported ferroelectrics, such as [(2-aminoethyl)trimethylphosphanium)]2PbBr4 (0.95 µC/cm2), and (4,4-difluorohexahydroazepine)2PbI4 (1.1 µC/cm2), but is smaller than those of (4,4-difluoropiperidinium)2PbI4 (10 µC/cm2), (4-(aminomethyl)piperidinium)PbI4 (9.8 µC/cm2), (S)−3-F-(pyrrolidinium)CdCl3 (5.79 µC/cm2), (R)−3-F-(pyrrolidinium)CdCl3 (5.63 µC/cm2), (isoamylammonium)2(methylammonium)2Pb3Br10 (5 µC/cm2), (C4H9NH3)2CsPb2Br7 (4.2 µC/cm2), and (C4H9NH3)2PbCl4 (2.1 µC/cm2) [14,37-43]. (R-1,2-DAP·I)4·I3·BiI6 possess similar ferroelectricity and related properties (Fig. S6 in Supporting information).

    In a ferroelectric crystal, there are equivalent directions of Ps, and the probability of occurrence of these polarization directions is equal. Therefore, a ferroelectric crystal usually shows a multi-domain structure, which minimizes the energy of the system. Piezoresponse force microscopy (PFM) has been fully proven to be a useful tool for characterizing the domain structure of ferroelectrics at the nanoscale. Vertical and lateral PFM imaging can detect the out-of-plane and in-plane components of polarization, respectively [44]. Each PFM component consists of amplitude and phase signal, which reflect the intensity of piezoelectric response and direction of polarization, respectively. We simultaneously measured the lateral and vertical PFM on a bulk single crystal in the same area (5 × 5 µm2) at room temperature. As depicted in Figs. 3a-e, the observed domains with irregular shapes are irrelevant to the topography, indicating that the signal originates from the polarization state rather than from the interference in topography. For the lateral PFM, the amplitude patterns (Fig. 3a) agree with phase patterns (Fig. 3b), in which domain walls with weak piezoelectric responses separate the domains with strong piezoelectric responses. However, the domain structures of vertical PFM are different from lateral PFM (Figs. 3d and e), implying that there is the existence of non-180° domains [45]. Hence, (S-1,2-DAP·I)4·I3·BiI6 should belong to the multiaxial ferroelectrics, which is consistent with the structural analysis. To confirm (S-1,2-DAP·I)4·I3·BiI6 possesses the switchable Ps, we measured switching spectroscopy lateral PFM at a selected point. The phase-bias loop (red line) and amplitude-bias loop (blue line) were recorded in Fig. 3f. When the applied bias voltage exceeds the coercive voltage, the phase is reversed, demonstrating that the Ps is switchable. The polarization switching behavior with hysteresis gives rise to the amplitude-bias butterfly-like loop. Furthermore, measurements of ferroelectric domains switching for (S-1,2-DAP·I)4·I3·BiI6 were carried out. As shown in Fig. S7a (Supporting information), applying a + 45 V voltage pulse with a duration of 5 s on a selected point (marked by green), the polarization direction appears the reverse (Fig. S7b in Supporting information), confirming its ferroelectricity.

    Figure 3

    Figure 3.  The PFM measurement results of (S-1,2-DAP·I)4·I3·BiI6. The lateral PFM (a) amplitude and (b) phase images. (c) The topographical image. The vertical PFM (d) amplitude and (e) phase images. (f) The signal of the lateral PFM phase (red line) and amplitude (blue line) as functions of tip DC bias voltage at a selected point.

    The quasi-static method was employed to measure the piezoelectric coefficient (d22) on the single-crystal of (S-1,2-DAP·I)4·I3·BiI6 along the b-axis at room temperature [46]. As shown in Fig. S8 (Supporting information), the d22 of (S-1,2-DAP·I)4·I3·BiI6 was determined to be about 35 pC/N, which is greater than those of LiNbO3 (11 pC/N), MDABCO—NH4I3 (14 pC/N), and TGS (triglycine sulfate, 22 pC/N), comparable to that for PVDF (polyvinylidene fluoride, 30 pC/N), but is smaller than those of (4-aminotetrahydropyran)2PbBr4 (76 pC/N), [(CH3)4 N][FeCl4] (80 pC/N), (TMBM)MnBr3 (112 pC/N), (TMCM)CdBr3 (139 pC/N), and (TMCM)MnCl3 (185 pC/N) [36,47].

    To evaluate the bandgap of (S-1,2-DAP·I)4·I3·BiI6 and (R-1,2-DAP·I)4·I3·BiI6, the UV–vis absorption spectrum was recorded at room temperature. As shown in Fig. 4, the (S-1,2-DAP·I)4·I3·BiI6 possesses a gradually decreasing absorption edge in the range from 540 nm to 840 nm, implying an indirect bandgap. Using the Tauc equation (for details process see Supporting information), the value of bandgap was determined to be about 1.56 eV (inset of Fig. 4), which is an abnormally narrow bandgap among reported ferroelectric semiconductors, and even almost comparable to CH3NH3PbI3 (~1.5 eV) (Fig. S9 in Supporting information). (R-1,2-DAP·I)4·I3·BiI6 also possesses the same bandgap, 1.56 eV, (Fig. S10 in Supporting information). The narrow bandgap of (S-1,2-DAP·I)4·I3·BiI6 and (R-1,2-DAP·I)4·I3·BiI6 is ascribed to the introduction of I3 ions, as reported by the previous reports [21-25]. I-5p orbitals of I3 ions lower the bottom of the conduction band.

    Figure 4

    Figure 4.  The UV–vis absorbance spectrum of (S-1,2-DAP·I)4·I3·BiI6. The inset shows the Tauc plot for fitting the indirect band gap of 1.56 eV.

    In summary, we have developed a pair of chiral hybrid compounds, (R-1,2-DAP·I)4·I3·BiI6 and (S-1,2-DAP·I)4·I3·BiI6. Benefited from intercalation of I3 ions, (R-1,2-DAP·I)4·I3·BiI6 and (S-1,2-DAP·I)4·I3·BiI6 exhibit a narrow bandgap of 1.56 eV comparable to that of CH3NH3PbI3, which is confirmed by UV–vis absorbance spectrum. Moreover, by the introduction of chirality, (R-1,2-DAP·I)4·I3·BiI6 and (S-1,2-DAP·I)4·I3·BiI6, possess multiaxial ferroelectricity and efficient piezoelectric response of 35 pC/N. This work provides inspiration for the direct design of molecular ferroelectric with the desired function.

    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.

    This work was supported financially by the National Key Research and Development Program of China (No. 2017YFA0204800), National Natural Science Foundation of China (Nos. 22175079 and 21875093), Natural Science Foundation of Jiangxi Province (Nos. 20204BCJ22015 and 20202ACBL203001), Jiangxi Provincial Department of Education Science and Technology Research Project (No. GJJ210812), Jiangxi Provincial Natural Science Foundation of China (No. 20212BAB214021), Science and Technology Project of Jiangxi Provincial Department of Education (No. GJJ200836).

    Supplementary material associated with this article can be found, in the online version, at doi: 10.1016/j.cclet.2022.108051.


    1. [1]

      Y.Y. Tang, P.F. Li, W.Q. Liao, et al., J. Am. Chem. Soc. 140 (2018) 8051–8059. doi: 10.1021/jacs.8b04600

    2. [2]

      Z. Wu, W. Zhang, H. Ye, et al., J. Am. Chem. Soc. 143 (2021) 7593–7598. doi: 10.1021/jacs.1c00459

    3. [3]

      I. Grinberg, D.V. West, M. Torres, et al., Nature 503 (2013) 509–512. doi: 10.1038/nature12622

    4. [4]

      S. Han, M. Li, Y. Liu, et al., Nat. Commun, 12 (2021) 284. doi: 10.1038/s41467-020-20530-4

    5. [5]

      X. Li, J.M. Hoffman, M.G. Kanatzidis, Chem. Rev. 121 (2021) 2230–2291. doi: 10.1021/acs.chemrev.0c01006

    6. [6]

      S. Liu, L. He, Y. Wang, et al., Chin. Chem. Lett. 33 (2022) 1032–1036. doi: 10.1016/j.cclet.2021.07.039

    7. [7]

      Q. Jia, T. Shao, L. Tong, et al., Chin. Chem. Lett. 34 (2023) 107539. doi: 10.1016/j.cclet.2022.05.053

    8. [8]

      C. Su, Z. Zhang, J. Yao, et al., Chin. Chem. Lett. 34 (2023) 107442. doi: 10.1016/j.cclet.2022.04.040

    9. [9]

      C.F. Wang, H. Li, Q. Ji, et al., Adv. Funct. Mater. 32 (2022) 2205918. doi: 10.1002/adfm.202205918

    10. [10]

      L. Mao, C.C. Stoumpos, M.G. Kanatzidis, J. Am. Chem. Soc. 141 (2018) 1171–1190. doi: 10.1007/s40195-018-0755-z

    11. [11]

      C. Ji, D. Dey, Y. Peng, et al., Angew. Chem. Int. Ed. 59 (2020) 18933–18937. doi: 10.1002/anie.202005092

    12. [12]

      L. Li, X. Liu, C. He, et al., J. Am. Chem. Soc. 142 (2020) 1159–1163. doi: 10.1021/jacs.9b11341

    13. [13]

      W. Zhang, M. Hong, J. Luo, J. Am. Chem. Soc. 143 (2021) 16758–16767. doi: 10.1021/jacs.1c08281

    14. [14]

      H.Y. Zhang, Z.X. Zhang, X.G. Chen, et al., J. Am. Chem. Soc. 143 (2021) 1664–1672. doi: 10.1021/jacs.0c12907

    15. [15]

      Y.L. Zeng, X.Q. Huang, C.R. Huang, et al., Angew. Chem. Int. Ed. 60 (2021) 10730–10735. doi: 10.1002/anie.202102195

    16. [16]

      G. Liu, L. Kong, J. Gong, et al., Adv. Funct. Mater. 27 (2016) 1604208.

    17. [17]

      Y.Y. Chen, C.H. Gao, T. Yang, et al., Chin. J. Struct. Chem. 41 (2022) 2204001–22040011.

    18. [18]

      H.Y. Ye, W.Q. Liao, C.L. Hu, et al., Adv. Mater. 28 (2016) 2579–2586. doi: 10.1002/adma.201505224

    19. [19]

      L. Li, X. Liu, Y. Li, et al., J. Am. Chem. Soc. 141 (2019) 2623–2629. doi: 10.1021/jacs.8b12948

    20. [20]

      L. Li, Z. Sun, P. Wang, et al., Angew. Chem. Int. Ed. 56 (2017) 12150–12154. doi: 10.1002/anie.201705836

    21. [21]

      A. Starkholm, L. Kloo, P.H. Svensson, ACS Appl. Energy Mater. 2 (2018) 477–485.

    22. [22]

      B. Kou, W. Zhang, C. Ji, et al., Chem. Commun. 55 (2019) 14174–14177. doi: 10.1039/c9cc05365d

    23. [23]

      W. Zhang, B. Kou, Y. Peng, et al., J. Mater. Chem. C 6 (2018) 12170–12174. doi: 10.1039/c8tc04372h

    24. [24]

      W. Zhang, X. Liu, L. Li, et al., Chem. Mater. 30 (2018) 4081–4088. doi: 10.1021/acs.chemmater.8b01200

    25. [25]

      W.S. Yang, B.W. Park, E.H. Jung, et al., Science 356 (2017) 1376–1379. doi: 10.1126/science.aan2301

    26. [26]

      H.Y. Zhang, Y.Y. Tang, P.P. Shi, et al., Acc. Chem. Res. 52 (2019) 1928–1938. doi: 10.1021/acs.accounts.8b00677

    27. [27]

      V. Mujica, Nat. Chem. 7 (2015) 543–544. doi: 10.1038/nchem.2294

    28. [28]

      G. Long, R. Sabatini, M.I. Saidaminov, et al., Nat. Rev. Mater. 5 (2020) 423–439. doi: 10.1038/s41578-020-0181-5

    29. [29]

      C.F. Wang, C. Shi, A. Zheng, et al., Mater. Horiz. 9 (2022) 2450–2459. doi: 10.1039/d2mh00698g

    30. [30]

      C. Shi, J.J. Ma, J.Y. Jiang, et al., J. Am. Chem. Soc. 142 (2020) 9634–9641.

    31. [31]

      K. Aizu, Phys. Rev. 146 (1966) 423–429. doi: 10.1103/PhysRev.146.423

    32. [32]

      D.A. Kleinman, Phys. Rev. 126 (1962) 1977–1979. doi: 10.1103/PhysRev.126.1977

    33. [33]

      Y. Xue, Z. Zhang, P. Shi, et al., Chin. Chem. Lett. 32 (2021) 539–542. doi: 10.1016/j.cclet.2020.02.005

    34. [34]

      W.P. Zhao, C. Shi, A. Stroppa, et al., Inorg. Chem. 55 (2016) 10337–10342. doi: 10.1021/acs.inorgchem.6b01545

    35. [35]

      M.E. Kamminga, A. Stroppa, S. Picozzi, et al., Inorg. Chem. 56 (2017) 33–41. doi: 10.1021/acs.inorgchem.6b01699

    36. [36]

      Y.M. You, W.Q. Liao, D. Zhao, et al., Science 357 (2017) 306–309. doi: 10.1126/science.aai8535

    37. [37]

      T.T. Sha, Y.A. Xiong, Q. Pan, et al., Adv. Mater. 31 (2019) 1901843. doi: 10.1002/adma.201901843

    38. [38]

      H.Y. Zhang, X.J. Song, X.G. Chen, et al., J. Am. Chem. Soc. 142 (2020) 4925–4931. doi: 10.1021/jacs.0c00371

    39. [39]

      I.H. Park, Q. Zhang, K.C. Kwon, et al., J. Am. Chem. Soc. 141 (2019) 15972–15976. doi: 10.1021/jacs.9b07776

    40. [40]

      Y.Y. Tang, Y. Ai, W.Q. Liao, et al., Adv. Mater. 31 (2019) 1902163. doi: 10.1002/adma.201902163

    41. [41]

      M. Li, Y. Xu, S. Han, et al., Adv. Mater. 32 (2020) 2002972. doi: 10.1002/adma.202002972

    42. [42]

      Z. Wu, C. Ji, L. Li, et al., Angew. Chem. Int. Ed. 57 (2018) 8140–8143. doi: 10.1002/anie.201803716

    43. [43]

      C. Ji, S. Wang, L. Li, et al., Adv. Funct. Mater. 29 (2019) 1805038. doi: 10.1002/adfm.201805038

    44. [44]

      H. Lu, T. Li, S. Poddar, et al., Adv. Mater. 27 (2015) 7832–7838. doi: 10.1002/adma.201504019

    45. [45]

      X.G. Chen, X.J. Song, Z.X. Zhang, et al., J. Am. Chem. Soc. 142 (2020) 10212–10218. doi: 10.1021/jacs.0c03710

    46. [46]

      H.Y. Ye, Y.Y. Tang, P.F. Li, et al., Science 361 (2018) 151–155. doi: 10.1126/science.aas9330

    47. [47]

      Q. Pan, Y.A. Xiong, T.T. Sha, et al., Mater. Chem. Front. 5 (2021) 44–59. doi: 10.1039/d0qm00288g

  • Figure 1  The packing structure of (S-1,2-DAP·I)4·I3·BiI6 in the LTP (a) and HTP (c) (H atoms are omitted for clarity). (b) Distribution of hydrogen bonds of S-1,2-DAP cations. (d) Two-fold disordered S-1,2-DAP cations around C2 rotation axis in the HTP. (e) The relationship of the unit cells of (S-1,2-DAP·I)4·I3·BiI6 between the LTP and HTP. The red and blue edges represent the crystal lattice of the LTP and the HTP, respectively. The red arrow represents the spontaneous polarization direction in the LTP. The blue arrows represent the possible equivalent polarization directions. Only the Bi atoms are shown for clarity.

    Figure 2  The ferroelectricity and related properties of (S-1,2-DAP·I)4·I3·BiI6. (a) The SHG signal intensity a function of temperature. (b) The ε′ in the different temperature at 1 MHz. (c) The PE hysteresis loop at 333 K.

    Figure 3  The PFM measurement results of (S-1,2-DAP·I)4·I3·BiI6. The lateral PFM (a) amplitude and (b) phase images. (c) The topographical image. The vertical PFM (d) amplitude and (e) phase images. (f) The signal of the lateral PFM phase (red line) and amplitude (blue line) as functions of tip DC bias voltage at a selected point.

    Figure 4  The UV–vis absorbance spectrum of (S-1,2-DAP·I)4·I3·BiI6. The inset shows the Tauc plot for fitting the indirect band gap of 1.56 eV.

  • 加载中
计量
  • PDF下载量:  2
  • 文章访问数:  281
  • HTML全文浏览量:  5
文章相关
  • 发布日期:  2023-08-15
  • 收稿日期:  2022-08-27
  • 接受日期:  2022-12-06
  • 修回日期:  2022-11-10
  • 网络出版日期:  2022-12-08
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章