• Journal of Semiconductors
  • Vol. 46, Issue 5, 051804 (2025)
Junyu Li1,†, Songwei Zhang1,†, Mohd Nazim Mohtar2, Nattha Jindapetch3..., Istvan Csarnovics4, Mehmet Ertugrul5, Zhiwei Zhao1, Jing Chen1,*, Wei Lei1,** and Xiaobao Xu1,***|Show fewer author(s)
Author Affiliations
  • 1Joint International Research Laboratory of Information Display and Visualization, School of Electronic Science and Engineering, Southeast University, Nanjing 211189, China
  • 2MyAgeingS1, University Putra Malaysia, Serdang, Selangor 43400, Malaysia
  • 3Department of Electrical and Biomedical Engineering, Faculty of Engineering, Prince of Songkla University, Hat Yai, Songkhla 90110, Thailand
  • 4Faculty of Science and Technology, University of Debrecen, Debrecen 4026, Hungary
  • 5Department of Metallurgy and Material Science Engineering Composite Materials, Engineering Faculty, Karadeniz Technical University, Trabzon 61080, Turkey
  • show less
    DOI: 10.1088/1674-4926/24100024 Cite this Article
    Junyu Li, Songwei Zhang, Mohd Nazim Mohtar, Nattha Jindapetch, Istvan Csarnovics, Mehmet Ertugrul, Zhiwei Zhao, Jing Chen, Wei Lei, Xiaobao Xu. Advances in multi-phase FAPbI3 perovskite: another perspective on photo-inactive δ-phase[J]. Journal of Semiconductors, 2025, 46(5): 051804 Copy Citation Text show less

    Abstract

    Halide perovskites have attracted great interest as active layers in optoelectronic devices. Among perovskites with diverse compositions, α-FAPbI3 is of utmost importance with great optoelectronic properties and a decent bandgap of 1.48 eV. However, the α-phase suffers an irreversible transition to the photo-inactive δ-phase, whereas the δ-phase is usually regarded as useless phase with poor optoelectronic properties. Therefore, it is commonly accepted that the thermodynamic stable δ-FAPbI3 greatly limits the application of FAPbI3. Every coin has two sides, although the δ-phase is difficult to apply as photoelectrical active layers, it is possible to combine δ-FAPbI3 with α-FAPbI3 to realize functional applications. Firstly, this review analyzes the cause of the contrasting properties between α- and δ-FAPbI3, where the stronger electron?phonon coupling in 1D hexagonal δ-FAPbI3 restricts its internal carrier and phonon transport. Secondly, the factors affecting the phase transitions and strategies to control phase transition between α- and δ-FAPbI3 are presented. Finally, some functional applications of δ-FAPbI3 in combination with α-FAPbI3 are given according to previous reports. By and large, we hope to introduce δ-FAPbI3 from another perspective and give some insights into its unique properties, hopefully providing new strategies for the subsequent advances to FAPbI3.

    Introduction

    Halide perovskites (HPs) have garnered enormous interest in the past decade, owing to the low-cost solution preparation[13], the easily adjustable bandgap[4, 5], and extraordinary optoelectronic properties[611]. HPs have a generic structure of ABX3, whereas A denotes organic ammonium or cesium cation (CH3NH3+, H2NCHNH2+, Cs+), B is divalent metal cation (Pb2+, Sn2+), and X is halide anion (Cl, Br, I). The diversiform compositions lead to easily tailored bandgap and consequently altered properties, giving a broad range of candidates with excellent optical absorption coefficients and optoelectronic properties. It is no exaggeration that the past decade has witnessed the booming development of these promising HPs. The certified energy conversion efficiency of single-junction perovskite-based solar cells has reached 26.7%[12], exhibiting the capability for further advancement to exceed silicon-based solar cells. Besides, the external quantum efficiency (EQE) of perovskite light-emitting diodes exceeds 30%[13], and multifunctional photodetectors have been widely constructed[1416]. All of these applications demonstrate the remarkable potential of HPs as active layers for optoelectronic devices.

    Among diverse A-site and X-site substituted HP systems, formamidinium lead iodide (FAPbI3) is of utmost importance. α-FAPbI3 possesses extraordinary optical and electrical properties[17] with a decent bandgap (1.48 eV) close to the theoretical value of the single-junction solar cell (1.39 eV)[18], exhibiting high efficiency up to 25.6%[19]. Despite its outstanding performance as an active material, α-FAPbI3 is plagued by the issues of the diversiform phases of FAPbI3, including α-, β-, γ-, and δ-phase[20]. As shown in Fig. 1, β- and γ-phase are both low-temperature tetragonal phases, which are presented below 285 and 140 K, respectively[21]. While the δ-phase is thermodynamically stable at room temperature, α-FAPbI3 would convert to the δ-FAPbI3 regardless of the storage environment at room temperature[22]. α-FAPbI3 is a cubic system with corner-sharing octahedrons, whereas δ-FAPbI3 is a hexagonal system with face-sharing octahedrons, the molecular volume increases by only 2.2% from α- to δ-phase transition, resulting in the good coexistence between these two phases at room temperature[23]. Regrettably, due to the non-perovskite 1D hexagonal structure of δ-FAPbI3, the potential barrier for self-trapping in the low-dimension structures is significantly reduced, resulting in self-trapped states under strong electron−phonon coupling[24]. Thus, δ-FAPbI3 is photo-inactive and electronic insulating (bandgap of ~2.88 eV), and the power conversion efficiency of δ-FAPbI3-based solar cells is only 0.2%−0.7%[25, 26].

    (Color online) Schematic diagram of FAPbI3 phases at different temperatures. Reprinted with permission from Ref. [21], Copyright 2018, American Chemical Society.

    Figure 1.(Color online) Schematic diagram of FAPbI3 phases at different temperatures. Reprinted with permission from Ref. [21], Copyright 2018, American Chemical Society.

    As a result of the stark contrast efficiencies of these two phases, the non-functional δ-phase is always treated as a "trashy" parasitic phase during the formation of α-FAPbI3[27]. Thus, the major challenge for FAPbI3 is to suppress the phase transition from the photoactive black α-phase to the yellow δ-phase, and a great number of strategies have been proposed. For solution-processed FAPbI3-based films, A-site incorporation with smaller inorganic Cs+ or organic MA+ and B-site incorporation with Br was widely demonstrated as an effective approach to suppress the transformation from the black phase to the yellow phase at room temperature[28, 29]. This strategy can be attributed to the reduced electron−phonon coupling strength by restricting FA+ cation motion[3032], leading to larger charge carrier mobility, lower non-radiative electron−hole recombination rate, and thus facilitating the charge carrier collection process[31].

    Every coin has two sides, is δ-FAPbI3 perovskite thoroughly impractical? To ponder this impediment from another perspective, could the δ-FAPbI3, with unique electronic insulating properties, act as electrical insulating layers and thus use the phase transition between α- and δ-phase to manipulate electron insulation to conduction? In this review, the drastically distinct structure and associated properties of α- and δ-phase are firstly compared, looking forward to clarifying the origin of the nature of electronic insulating δ-FAPbI3. Secondly, the phase transition from α- to δ-phase is an intricate process, which can be attributed to factors such as temperature, pressure (internal/external pressure), humidity, etc. Thus, different characteristics and mechanisms related to phase transition are systematically presented, giving a comprehensive summary and profound insight into the process. Meanwhile, strategies for promoting and preventing phase transitions utilizing different means are also provided. Finally, combined with some novel δ-phase applications reported so far and the above analysis of properties and phase transition mechanisms, we hope to give some insights into the further functionalized application of δ-FAPbI3. In a word, we are looking forward to giving a roundabout strategy to exploit the "trashy" δ-FAPbI3 and providing more possibilities for the application of FAPbI3-based perovskite.

    Electron−phonon coupling in α- and δ-FAPbI3

    Despite the recent controversy about the distorted tetragonal structure of α-FAPbI3[33], α-FAPbI3 is still widely recognized as a cubic structure with Pm3¯m symmetry, while δ-FAPbI3 is accepted as 1D hexagonal P63mc[22, 34], as can be seen from XRD patterns with different periodic alignments in Fig. 2(a). The phase transition from α- to δ-FAPbI3 is also accompanied by the main Raman peak redshifted from 135 to 111 cm−1 due to the different vibration modes of FA molecules (Fig. 2(b))[17]. It has been discovered that there exist contrasting properties between α- and δ-FAPbI3, for instance, δ-FAPbI3 exhibits thermodynamic stable at room temperature with lager bandgap, but the chain-like structure in δ-FAPbI3 obstructs electron transport compared to α-FAPbI3. The strong electron−phonon coupling in this 1D structure easily lead to self-trapped excitons (STEs)[35, 36], which induces an elastic distortion of the surrounding lattice and further captures the photogenerated electrons. Owing to the more stable self-trapped state, the electron−phonon coupling in δ-FAPbI3 would further deteriorate the electrical properties, which can explain the electronically isolated states in the δ-phase[37].

    (Color online) (a) XRD patterns with different periodic alignments for α- and δ-FAPbI3 single crystals and powder. (b) Different Raman shift for α- and δ-FAPbI3. (a) and (b) Reprinted with permission from Ref. [44], Copyright 2023, American Chemical Society. (c) Temperature-dependent steady-state PL spectrum of α-FAPbI3. (d) The extracted FWHM from steady-state PL spectrum of α-FAPbI3, and the well fitted red line indicates contributions from inhomogeneous broadening and Fröhlich coupling. (c) and (d) Reprinted with permission from Ref. [38], Copyright 2016, published under the terms of the Creative Commons CC BY license.

    Figure 2.(Color online) (a) XRD patterns with different periodic alignments for α- and δ-FAPbI3 single crystals and powder. (b) Different Raman shift for α- and δ-FAPbI3. (a) and (b) Reprinted with permission from Ref. [44], Copyright 2023, American Chemical Society. (c) Temperature-dependent steady-state PL spectrum of α-FAPbI3. (d) The extracted FWHM from steady-state PL spectrum of α-FAPbI3, and the well fitted red line indicates contributions from inhomogeneous broadening and Fröhlich coupling. (c) and (d) Reprinted with permission from Ref. [38], Copyright 2016, published under the terms of the Creative Commons CC BY license.

    Electron−phonon coupling refers to the interaction between charge carriers and lattice vibrations (phonons), where charge carriers dressed by phonons induce the formation of polarons, setting a fundamental limit to charge carrier mobility without extrinsic scattering off impurities or interfaces[38]. Electron−phonon coupling is an aggregation of various mechanisms, where the Fröhlich interaction is expected to dominate in ionic crystals as HPs at room temperature, and the phonon anharmonicity may also form unconventional polaritons, which causes phonon overdamping and may have a strong impact on the physical properties, especially electron−phonon interactions[39]. Fröhlich interaction refers to the long-range interaction of electric fields generated by longitudinal optical (LO) phonons with electrons[40], which is the dominant source of PL linewidth broadening[41, 42]. As presented in Eq. (1):

    Γ(T)=Γ0+ΓLO=Γ0+γLOeELO/kT1,

    where Г0 is a temperature-independent inhomogeneous broadening term as a result of disorder and/or imperfections, ГLO is a homogeneous broadening term attributed to longitudinal optical (Fröhlich) scattering, γLO is coupling strength and ELO is an energy representative of longitudinal optical phonon. The temperature-dependent steady-state PL spectrum and the extracted full width at half-maximum (FWHM) of FAPbI3 (here refers to the alpha phase) are shown in Figs. 2(c) and 2(d), and the well-fitted PL linewidth to Eq. (1) indicates that the existence of Fröhlich coupling in α-FAPbI3 like other HPs, and the dominant contribution to the homogeneous linewidth broadening attributes to the Fröhlich coupling between charge carriers and LO phonons[38]. Meanwhile, it is found the substitution of organic cations in HPs has relatively little effect on the Fröhlich interaction compared to the substitution of inorganic frameworks[43]. Many studies have also found that doping Cs+[30], (CH3)2NH2+ (abbreviated as DMA) cation[31], and Br anion[31, 32] restrict FA+ cation motion by restraining the structural fluctuations of the lattice, exhibiting reduced electron−phonon coupling strength and suppressed ionic migration, which is conducive to larger charge carrier mobility, lower non-radiative recombination rate, and better stability than pure α-FAPbI3.

    However, it is still hard to explain the different performance between α- and δ-FAPbI3 (Fig. 3(a)) with only different Fröhlich coupling, due to the similar phonon densities of states in these FAPbI3 phases (Fig. 3(b)). Thus, a more complex electron−phonon coupling process exists in electronically isolated δ-FAPbI3, and the anharmonic effects in δ-FAPbI3 should be taken into consideration[33]. The 1D distorted face-sharing PbX6 octahedrons could result in anharmonic electron−phonon coupling[35, 45]. The self-trapping is strongly related to the dimensionality of materials, and in three-dimensional materials there is a potential barrier that needs to be overcome to initiate the self-trapping process. However, in low-dimensional materials, the potential barrier for self-trapping is weakened or disappears, and thus STEs are more easily observed, e.g., as the number of quasi-two-dimensional HP layers decreases (i.e., the dimensionality is closer to two-dimensionality), the intensity of STEs increases, as shown in Fig. 3(c)[24]. Therefore, the lower-dimensional 1D structure of δ-FAPbI3 gives rise to the low potential barrier for self-trapping like CsPbI3[45] and TMAPbI3[35] with a similar 1D structure, resulting in strong STEs and poor electrical performance because of anharmonic electron−phonon coupling[24]. As a result of the existence of STEs, the captured photogenerated excitons then release energy in the form of luminescence, and the multiple broadband photoluminescence (PL) emission of δ-FAPbI3 can be seen, as compared with that of α-FAPbI3 in Fig. 3(d)[37]. Based on the above analysis, we believe that the origin of the different properties of α- and δ-FAPbI3 is due to the STEs caused by anharmonic electron−phonon coupling in the low-dimensional δ-FAPbI3.

    (Color online) (a) Different structures of photoactive α-FAPbI3 and photo-inactive δ-FAPbI3. (b) Phonon density of states for the cubic, tetragonal, and hexagonal structures of FAPbI3, where the blue and red peaks represent the vibrations of the PbI3 octahedrons and FA+ cations, respectively. (a) and (b) Reprinted with permission from Ref. [33], Copyright 2022, American Chemical Society. (c) Schematic diagram of energy bands for STEs. Reprinted with permission from Ref. [24], Copyright 2020, Springer Nature. (d) PL spectra of α- and δ-FAPbI3. Reprinted with permission from Ref. [44], Copyright 2023, American Chemical Society.

    Figure 3.(Color online) (a) Different structures of photoactive α-FAPbI3 and photo-inactive δ-FAPbI3. (b) Phonon density of states for the cubic, tetragonal, and hexagonal structures of FAPbI3, where the blue and red peaks represent the vibrations of the PbI3 octahedrons and FA+ cations, respectively. (a) and (b) Reprinted with permission from Ref. [33], Copyright 2022, American Chemical Society. (c) Schematic diagram of energy bands for STEs. Reprinted with permission from Ref. [24], Copyright 2020, Springer Nature. (d) PL spectra of α- and δ-FAPbI3. Reprinted with permission from Ref. [44], Copyright 2023, American Chemical Society.

    Contrasting properties between α- and δ-FAPbI3

    It becomes more reasonable to explain the contrasting properties between α- and δ-FAPbI3 after considering STEs in δ-FAPbI3 due to anharmonic electron−phonon coupling. The different coupling effects between electrons and phonons of these two phases in the microscopic perspective can also be manifested in the macroscopic properties. Owing to the confinement of electrons and phonons in δ-FAPbI3, it exhibits significantly reduced electronic and thermal properties compared to the α-phase. Firstly, relative permittivity is the ratio of a material’s permittivity to vacuum permittivity and is widely used to describe and compare the electronic properties of different materials. Due to the different coupling effects and structural arrangements of α- and δ-FAPbI3, these two phases exhibit greatly different relative permittivity, as shown in Figs. 4(a)−4(c). Regardless of varying temperature cycles, the real part of the permittivity measured at 100 kHz from δ-FAPbI3 to α-FAPbI3 undergoes a rapid increase, approximately a twofold increase from 20 to about 50[46]. Since the electronic mobility at room temperature is typically governed by phonon scattering via electron−phonon coupling[47], the carrier mobility of δ-FAPbI3 is even about 20 times lower than that of α-FAPbI3[5, 48], which also explains the additional electron transport limitations present in δ-FAPbI3, such as the anharmonic electron−phonon coupling mentioned above.

    (Color online) (a)−(c) Real part of the permittivity of FAPbI3 measured at 100 kHz during temperature cycles at different conditions. Reprinted with permission from Ref. [46], Copyright 2019, American Chemical Society. (d) Thermal diffusivity test of α- and δ-FAPbI3. Reprinted with permission from Ref. [44], Copyright 2023, American Chemical Society.

    Figure 4.(Color online) (a)−(c) Real part of the permittivity of FAPbI3 measured at 100 kHz during temperature cycles at different conditions. Reprinted with permission from Ref. [46], Copyright 2019, American Chemical Society. (d) Thermal diffusivity test of α- and δ-FAPbI3. Reprinted with permission from Ref. [44], Copyright 2023, American Chemical Society.

    Meanwhile, the δ-phase also exhibits lower thermal transport properties, since the thermal conductivity is related to the crystal structure through phonon properties. The thermal conductivity (k) consists of two parts: electronic and phononic/lattice contributions, as expressed as follows:

    k=ke+kph=LσT+13CphvgΛ,

    where ke is electron thermal conductivity (according to the Wiedemann−Franz law ke = LσT), L is the Lorenz constant, σ is electrical conductivity, T is temperature, and kph is phonon thermal conductivity, where Cph is the volumetric phonon heat capacity, vg is the average phonon group velocity, and Λ is the mean free path of phonon. Owing to the low carrier concentration of non-metallic HPs, the electronic contribution is negligible, and the phononic contribution dominates the total thermal conductivity, expressed as[49]:

    k=13CphvgΛ.

    The ultralow thermal conductivity (less than 1 W∙K−1∙m−1) of HPs can be related to the enhanced anharmonic phonon−phonon scattering due to the highly overlapped phonon branches[50, 51]. Considering the lifetime of acoustic phonons (~ps) is orders of magnitude larger than the sub-picosecond lifetime of optical phonons, it could be concluded that acoustic phonons dominate the thermal transport, and the mean free path of acoustic phonons is in the range of 20−200 Å[52].

    Microscopically, effective thermal conduction is the transfer of excess kinetic energy of carriers to the surroundings in an irreversible manner, which can be categorized into three stages: the scattering of carrier−phonon (Frӧhlich interaction), the attenuation from high-frequency optical phonons to low-frequency acoustic phonons and the propagation of acoustic phonons to the far-field region[50]. Therefore, the stronger electron−phonon coupling in material severely restricts the thermal conduction and results in lower thermal conductivity, which has been reported in metals[53], metal−nonmetal interfaces[54], and semimetals[5557]. In our previous work, the thermal diffusivity of δ-FAPbI3 (α = k/ρc, where ρ is density and c is specific heat capacity) is about two-thirds of that of α-phase[44]. It is worth mentioning that due to the strict requirements of the test instrument on the sample size and surface flatness, it is not a single crystal in situ test, but a test of pressed crystallite tablets. However, the yellow δ-FAPbI3 will transition to the different phase under pressure. The actual "δ-phase" tablet tested is a mixture of the multi-phases, which can be intuitively seen from the single-crystal and tablet color comparisons in Fig. 4(d), so it is reasonable to believe that the real thermal transport properties of δ-FAPbI3 should be further lower.

    Controlled phase transition between α- and δ-FAPbI3

    Although α-FAPbI3 and δ-FAPbI3 can coexist well at room temperature, the α-phase is metastable and would gradually transition to the thermodynamically stable δ-phase. As shown in Fig. 5(a), the lower Gibbs free energy in δ-FAPbI3 results in the inevitable α-to-δ phase transition, mainly aroused from the entropy contribution of the organic cation[58]. At high temperatures, the FA+ cation has an isotropic orientation with a large entropy value, stabilizing the cubic structure. In contrast, as the temperature decreases, the FA+ cation acquires a strong preferred orientation in the hexagonal phase with a lower entropy value (Fig. 5(b))[59]. Thus, α- to δ-FAPbI3 transition is a Gibbs free energy (chemical potential) reduction process, and the driving force of the α-to-δ phase transition can be attributed to the internal tensile strain in the α-FAPbI3 unit cells. As presented in Fig. 5(c), strain-free α-FAPbI3 films rapidly transform to δ-FAPbI3, while the strained α-FAPbI3 films show striking long-term stability, which is stable for at least 360 days after growth[60]. The driving force of the α-to-δ phase transition is thought to be the internal tensile strain in the unit cell, which induces the formation of vacancies and the subsequent phase transition, whereas, the films under compressive strain effectively neutralize the effect of the internal tensile strain, which is key to the stabilization of α-FAPbI3. What’s more, this strain could even effectively changes the crystal structure, reduces the bandgap and increases the hole mobility of α-FAPbI3. In addition to the internal strain, the external pressure can also affect the process of phase transition as illustrated in Fig. 5(d)[22]. During compression from 0.1 to 6.59 GPa, the structural transition of the cubic α-FAPbI3 follows the sequence Pm3¯m to P4/mbm to Im3¯ to partially amorphous, whereas the δ-FAPbI3 converts to the orthorhombic Cmc21 structure between 1.26 and 1.73 GPa. Conversely, the phase transition can also be recovered during the decompression process.

    (Color online) (a) Schematic diagram of Gibbs free energy for α- and δ-FAPbI3. Reprinted with permission from Ref. [58], Copyright 2024, Elsevier. (b) The kinetic diagram with oriented and isotropic FA+ for cubic α-FAPbI3 and hexagonal δ-FAPbI3. Reprinted with permission from Ref. [59], Copyright 2016, published under the terms of the Creative Commons CC BY−NC license. (c) Phase stability comparison of α-FAPbI3 films with/without internal strain. Reprinted with permission from Ref. [60], Copyright 2020, Springer Nature. (d) Phase transition diagram of α-FAPbI3 during compression and decompression. Reprinted with permission from Ref. [22], Copyright 2018, American Chemical Society.

    Figure 5.(Color online) (a) Schematic diagram of Gibbs free energy for α- and δ-FAPbI3. Reprinted with permission from Ref. [58], Copyright 2024, Elsevier. (b) The kinetic diagram with oriented and isotropic FA+ for cubic α-FAPbI3 and hexagonal δ-FAPbI3. Reprinted with permission from Ref. [59], Copyright 2016, published under the terms of the Creative Commons CC BY−NC license. (c) Phase stability comparison of α-FAPbI3 films with/without internal strain. Reprinted with permission from Ref. [60], Copyright 2020, Springer Nature. (d) Phase transition diagram of α-FAPbI3 during compression and decompression. Reprinted with permission from Ref. [22], Copyright 2018, American Chemical Society.

    To obtain the photoactive α-FAPbI3 from the thermodynamically stable δ-phase, elevating the temperature is the most direct and widely used strategy. As illustrated in Fig. 6(a), at temperatures greater than 150 °C, the δ-phase would completely transition to the black α-FAPbI3 within a few minutes. Therefore, Steele et al.[61] proposed to apply direct-laser-writing on the δ-FAPbI3 surface to realize micro-area δ-to-α transition, utilizing the transient high temperature (>150 °C) of the focused laser beam, as shown in Fig. 6(b). This strategy gives a controlled and convenient way to achieve micro-region phase transitions from δ-FAPbI3 to α-FAPbI3. Provided that the heat generated by the laser beam is sufficient to achieve the phase transition, the higher the laser power the smaller the minimum irradiation time required, and the phase transition can be realized within a few seconds of the focused laser sweep (Fig. 6(c)). When the required laser irradiation time is small enough, a patterned phase change region can be achieved by controlling the laser movement via the direct-laser-writing displacement stage. But too much laser power can also increase the phase transition region, as shown in Fig. 6(d). By choosing an appropriate laser power to balance the phase transition time and line array width, the α-FAPbI3 line array can be constructed on the δ-FAPbI3 host by controlling the focused laser using a displacement stage. However, due to the limitations of laser power and focusing lenses to realize surface temperatures greater than 150 °C, the smallest achievable micro-region width is still a few microns[44, 61]. This customizable micro-region hybrid α-/δ-FAPbI3 can take full advantage of the different properties of these two phases for functional sensing, which will be introduced in the next Section.

    (Color online) (a) Schematic diagram of phase transition between α- and δ-FAPbI3. Reprinted with permission from Ref. [58], Copyright 2024, published under the terms of the Creative Commons CC BY−NC−ND license. (b) Schematic diagram of the micro-area δ-to-α phase transition using direct-laser-writing. (c) Relationship between laser power and irradiation time of direct-laser-writing to realize δ-to-α phase transition. (d) Hybrid α-/δ-FAPbI3 under visible light and UV light constructed by direct-laser-writing, and the relationship between linewidth of the phase transition region and laser power. (b)−(d) Reprinted with permission from Ref. [44], Copyright 2023, American Chemical Society.

    Figure 6.(Color online) (a) Schematic diagram of phase transition between α- and δ-FAPbI3. Reprinted with permission from Ref. [58], Copyright 2024, published under the terms of the Creative Commons CC BY−NC−ND license. (b) Schematic diagram of the micro-area δ-to-α phase transition using direct-laser-writing. (c) Relationship between laser power and irradiation time of direct-laser-writing to realize δ-to-α phase transition. (d) Hybrid α-/δ-FAPbI3 under visible light and UV light constructed by direct-laser-writing, and the relationship between linewidth of the phase transition region and laser power. (b)−(d) Reprinted with permission from Ref. [44], Copyright 2023, American Chemical Society.

    However, whether utilizing pure α-FAPbI3 or hybrid α-/δ-FAPbI3, the stability issue needs to be first considered, both in terms of phase and material stability. The phase stability of different forms of α-FAPbI3 varies significantly. Due to undesirable surface and interfacial defects in spin-coated films, the α-FAPbI3 films would completely transition to δ-phase within a few hours or a few days[6266], thus additives and alloying/doping engineering has been proposed to suppress the α-to-δ phase transition[20]. In contrast, α-FAPbI3 single crystals exhibit greater stability, taking about one month to fully phase transition to the δ-phase, which can be attributed to the well-arranged lattice inside effectively inhibits the entry of water and oxygen molecules. As a result, the α-FAPbI3 line array constructed on δ-FAPbI3 single crystal through direct-laser-writing was maintained for nearly three weeks in an unencapsulated air environment, where the temperature is between 20−25 °C and the relative humidity is between 40%−55%[44]. Although the α-to-δ phase transition is spontaneous, humidity accelerates the process[23], and excessive humidity results in the degradation of HPs, where H2O molecules irreversibly decompose FAPbI3 into precursors accompanied by FA+ molecules dissolved in water or lost as gas[46]. Due to the water absorption of FA+ molecules, α-FAPbI3 and δ-FAPbI3 exhibit close material stability.

    Functional applications of photo-inactive δ-phases

    Because of the poor electrical transmission and narrower light absorption range, δ-FAPbI3 has been treated as "trash", and a variety of strategies have been proposed to suppress the α-to-δ phase transition[26, 29]. Although δ-FAPbI3 is difficult to be applied to optoelectronic devices as an active layer material alone, its special properties can be utilized to assist the photoactive α-FAPbI3 layer for functional applications, and the comparison of α-FAPbI3 and δ-FAPbI3 from phase transition to application is shown in Table 1. In this Section, we will introduce the potential application of α/δ-FAPbI3 based on their special properties, hoping to change the stereotype of uselessness for δ-FAPbI3.

    It was found that the presence of nanoscale inclusions of δ-FAPbI3 in α-FAPbI3 thin films work as barriers, while α-FAPbI3 behave as quantum wells, these hybrid structure leads to quantum confinement effects[67]. As shown in Fig. 7(a), the absorption spectra of FAPbI3 thin films exhibit oscillatory features as a result of the quantum confinement effects, and the length scale of confinement is estimated to be 10−20 nm. Even the width of δ-phase barriers is less than 10% of the potential wells and the height of barriers is less than 100 meV, the periodic δ-phase superlattice could still cause absorption peaks at higher energies. The energies of above-bandgap oscillatory peaks change with temperature according to the inverse square of the intrinsic lattice parameter, and with peak index in a quadratic manner[68]. This phenomenon suggests a controlled strategy to change the confinement potential in FAPbI3 by properly manipulating the phase distribution. Utilizing the instantaneous high temperature of the focused laser, direct-laser-writing can act on the δ-phase to achieve the α-phase transition in specific micro-regions[61]. Therefore, customized hybrid α/δ-FAPbI3 structures can be constructed with suitable laser power and displacement speed. Although there is no quantum confinement effect in this macroscopic hybrid structure, where micron-scale α-FAPbI3 is surrounded by δ-FAPbI3, the interior of δ-FAPbI3 has more restricted carrier and phonon transport than α-FAPbI3 due to stronger electron−phonon coupling and macroscopically exhibits restricted electrical and thermal transport properties. Given this, we propose to utilize hybrid α/δ-FAPbI3 structure constructed by direct-laser-writing to achieve low-crosstalk terahertz sensing. Terahertz is an electromagnetic radiation between infrared and microwave, with a special wavelength (3 mm−30 μm) that gives it many intriguing properties, such as the ability to penetrate the vast majority of materials, non-ionizing photonic properties, and correspondence to many molecular vibrational frequencies[69]. Therefore, terahertz sensing has a wide range of applications including non-destructive imaging, security inspection, molecular identification, etc.[7072]. However, terahertz waves of several meV energy are susceptible to room temperature thermal noise and often requires additional cryogenic devices, and conventional infrared antenna detectors are limited by their reliance on structures with nonlinear optics and materials with nonlinear electronics. To avoid delicate device architecture design[73, 74] or cryogenic cooling systems[75, 76] in detecting weak terahertz photons (0.41 to 41 meV)[77], the photo-thermal effect based on HPs has been widely used in room-temperature terahertz sensing[7881]. This far below-bandgap photo−thermal response could be realized based on the characteristic terahertz absorption due to Pb−X bonds’ vibration and the extremely low thermal transport performance of HPs, which convert terahertz wave into thermal energy, easily accumulate heat to raise the temperature, and the temperature rise alters the material resistance (bolometric effect) and electric potential (Seebeck effect), producing the thermal-induced electrical response. Although a single terahertz detection unit can be realized in this way, wavelength-dependent diffraction of terahertz causes inevitable thermal/electrical crosstalk limiting the improvement of detection accuracy. In general, thermal crosstalk is caused by heat diffusion from adjacent pixels, which further leads to electrical crosstalk in the photo−thermal detector arrays.

    (Color online) (a) Above-gap oscillations in the absorption spectra of FAPbI3 films at different temperatures. The inset illustrates two mechanisms that may result in the oscillations: quantum confinement in deep wells and periodicity of the superlattice confining potential. Reprinted with permission from Ref. [67], Copyright 2020, Springer Nature. (b) Schematic diagram of α-FAPbI3 single-crystal photothermal detector array, which denotes as device-α(SC) array, and the crosstalk of the nearest-neighbor and next-nearest-neighbor detection units in the device-α(SC) array. (c) Schematic diagram of hybrid α/δ-FAPbI3 single-crystal photothermal detector array by direct-laser-writing, which denotes as device-α array, and the crosstalk of the nearest-neighbor and next-nearest-neighbor detection units in the device-α array. (d) Terahertz photothermal proof-of-concept imaging of the device-α(SC) array. (e) Terahertz photothermal proof-of-concept imaging of the device-α array. (b)−(e) Reprinted with permission from Ref. [44], Copyright 2023, American Chemical Society.

    Figure 7.(Color online) (a) Above-gap oscillations in the absorption spectra of FAPbI3 films at different temperatures. The inset illustrates two mechanisms that may result in the oscillations: quantum confinement in deep wells and periodicity of the superlattice confining potential. Reprinted with permission from Ref. [67], Copyright 2020, Springer Nature. (b) Schematic diagram of α-FAPbI3 single-crystal photothermal detector array, which denotes as device-α(SC) array, and the crosstalk of the nearest-neighbor and next-nearest-neighbor detection units in the device-α(SC) array. (c) Schematic diagram of hybrid α/δ-FAPbI3 single-crystal photothermal detector array by direct-laser-writing, which denotes as device-α array, and the crosstalk of the nearest-neighbor and next-nearest-neighbor detection units in the device-α array. (d) Terahertz photothermal proof-of-concept imaging of the device-α(SC) array. (e) Terahertz photothermal proof-of-concept imaging of the device-α array. (b)−(e) Reprinted with permission from Ref. [44], Copyright 2023, American Chemical Society.

    To address the unavoidable thermal crosstalk and subsequent electrical crosstalk as a result of the terahertz diffraction, the hybrid α/δ-FAPbI3 single-crystal detector array was utilized to realize low-crosstalk terahertz detection. α-FAPbI3 single crystal with excellent optoelectronic performance was used as the active layer, while δ-FAPbI3 single crystal was used as the α-phase spacer layer, utilizing its more restricted thermal transport properties and carrier mobility to reduce the thermal and electrical crosstalk. Based on the bolometric effect, the α-FAPbI3 terahertz detector achieves a responsivity of 34.3 μA∙W−1, a noise equivalent power of 5.3 × 10−10 W∙Hz−1/2, and a specific detectivity of 1.4 × 107 Jones to 0.1 THz at 200 V∙mm−1. Since the δ-FAPbI3 spacer layer effectively blocks the thermal and electrical signals to reduce the crosstalk, the crosstalk of the nearest-neighbor and next-nearest-neighbor detection units in the two-phase coexisting α/δ-FAPbI3 is reduced from −5.35 and −12.38 dB to −17.60 and −21.53 dB, respectively (Figs. 7(b) and 7(c)). Therefore, the proof-of-concept imaging to terahertz using the δ-FAPbI3 spacing layer also has sharper contrast, as shown in Figs. 7(d) and 7(e).

    Parametersα-FAPbI3δ-FAPbI3
    StructureCubic structure (Pm3¯m) with corner-sharing octahedronsHexagonal structure (P63mc) with face-sharing octahedrons
    StabilitySensitive to humidity and oxygenSensitive to humidity and oxygen
    Temperature dependenceThermodynamic instable at RT spontaneous transition to the δ-phaseThermodynamically stable at RT transition to the α-phase via heating (about 150 °C )
    Pressure influenceIn compression process, α-FAPbI3 undergoes Pm3¯m → P4/mbm → Im3¯ → partially amorphous from ambient pressure to 6.59 GPa, δ-FAPbI3 converts to the orthorhombic Cmc21 structure between 1.26 and 1.73 GPa (the phase transition can also be recovered during the decompression process)
    Mechanism of phase transitionThe lower Gibbs free energy in δ-FAPbI3 results in the inevitable α-to-δ phase transition at RT
    Kinetics of transitionThe driving force of the phase transition can be attributed to the temperature and internal/external pressure in the FAPbI3 unit cells
    Optoelectronic propertiesExcellentPoor
    ApplicationsActing as active layer materialsAssisting the photoactive α-FAPbI3 layer for functional applications

    Table 1. The comparison of α-FAPbI3 and δ-FAPbI3.

    In addition to reducing photo−thermal crosstalk, the hybrid α/δ-FAPbI3 constructed by direct-laser-writing is also ideally suited to enable polarization detection. Tian et al.[82] etched a micro-grating on α-FAPbI3 film using direct-laser-writing, where uniform grating structure in perovskite films can be obtained without visible damages around the grating. As can be seen in the enlarged SEM image, the perovskite grain size increases in the valley where the laser scanned, which indicates that fs laser may help grain growth with better crystallinity and contribute to carrier transport and optoelectronic efficiency (Fig. 8(a)). The photodetectors achieved polarized responses due to the anisotropy conferred by the triangular grating on the film surface under a rotating linearly polarized light (Fig. 8(b)). It is worth noting that the polarization originates exclusively from the arrayed grating structure, where the ridges and valleys of the grating are composed of homogeneous α-FAPbI3. As a result, the measurement of α-FAPbI3 photodetectors with micro-grating shows that the highest and smallest photocurrent appear at 0° and 90° polarization directions, respectively, exhibiting a linear polarization of 0.97 under light intensity of 37.8 mW∙cm−2, demonstrating that unidirectional perovskite triangular grating arrays with potential for polarization detection (Fig. 8(c)). Linear sensitivity also indicates the increased polarization detection performance with micro-grating, and the linear sensitivity of α-FAPbI3 photodetector with micro-grating (258.4 nA per degree) is superior to that of the pristine α-FAPbI3 photodetector (189.2 nA per degree) under light intensity of 37.8 mW∙cm−2. Although only the change in film morphology of α-FAPbI3 has been utilized in this work to achieve polarization detection, it is reasonable to speculate that if δ-FAPbI3 with different (lower) optoelectronic properties are utilized as spacer grating for α-FAPbI3 as described above, the anisotropic light absorption and electrical transport properties superimposed on the anisotropic structure will further increase the polarization sensitivity. But honestly speaking, the current focused laser spot in direct-laser-writing is still too large, and the width of the realized grating is about 10 μm, which is still far from the wavelength of visible light about hundreds of nanometers. Thus, the polarization effect will be further enhanced if the size of the focused laser spot can be further reduced to construct a nanoscale grating or the detection band can be tuned to the micron-scale.

    (Color online) (a) SEM images of micro-grating on α-FAPbI3 film constructed by direct-laser-writing. (b) Schematic diagram of polarization photodetector with micro-grating. (c) Polarization performance of the α-FAPbI3 photodetector with micro-grating, including angle-dependent photocurrent and linear sensitivity. (a)−(c) Reprinted with permission from Ref. [82], Copyright 2022, John Wiley and Sons.

    Figure 8.(Color online) (a) SEM images of micro-grating on α-FAPbI3 film constructed by direct-laser-writing. (b) Schematic diagram of polarization photodetector with micro-grating. (c) Polarization performance of the α-FAPbI3 photodetector with micro-grating, including angle-dependent photocurrent and linear sensitivity. (a)−(c) Reprinted with permission from Ref. [82], Copyright 2022, John Wiley and Sons.

    Different from the incorporation of other ions into thin films to stabilize FAPbI3, some concepts have also been proposed about the application of δ-FAPbI3 to stabilize α-FAPbI3 phase[8385]. Unlike traditional bypassing of δ-FAPbI3 phase formation by altering the reaction pathway[86], it was found that the existence of δ-FAPbI3 in α-FAPbI3 films can stabilize the α-phase, accompanied by the enhancement and modulation of the NIR emission blueshift to 780 nm[87]. Ma et al. found that with the molar ratio of FAI/PbI2 increases, the colloidal size in precursor solution gradually decreases, and more coordination effects can be aroused from excess FAI in the precursor solution, causing the formation of large colloids. Once the size of the colloidal particles reaches a sufficiently low value, the total free energy of δ-FAPbI3 (including surface and bulk) may be equal to or even higher than that of α-FAPbI3, thus promoting the formation of α-FAPbI3 at lower annealing temperatures and the formation of α/δ phase junctions. Therefore, the α/δ phase junction could be formed through stoichiometrically modified precursors (FAI/PbI2 = 1.2), as shown in Fig. 9(a). High resolution transmission electron microscopy (HRTEM) in Fig. 9(b) clearly displays a junction structure, where the lattice fringes with spacing of 3.17 and 2.38 Å are indexed to the (222) plane of α-FAPbI3 and the (031) plane of δ-FAPbI3, respectively. The well aligned α/δ-FAPbI3 junction has energy levels between those of α-FAPbI3 and δ-FAPbI3 (Fig. 9(c)), and possesses not only structural stability but also long-term stability against humidity. This stable and controllable α/δ phase junction provides an example of how δ-FAPbI3 should be applied[87].

    (Color online) (a) Schematic diagram of the fabrication process of the α/δ-FAPbI3 phase junction. (b) HRTEM of the α/δ-FAPbI3 phase junction. (c) Schematic diagrams of the energy levels of pure α-FAPbI3, pure δ-FAPbI3, and the α/δ-FAPbI3 phase junction. (d) Schematic diagram of the fabrication process of the bilayered δ-FAPbI3/perovskite films. (e) Current density−voltage curves for devices based on pristine perovskite films and bilayered δ-FAPbI3/perovskite films. (f) Statistics of PCEs and Jsc for devices based on pristine perovskite films and bilayered δ-FAPbI3/perovskite films. (g) Evolution of PCEs for devices based on pristine perovskite films and bilayered δ-FAPbI3/perovskite films under 40% ± 5% relative humidity (RH) at room temperature. (a)−(c) Reprinted with permission from Ref. [87], Copyright 2017, published under the terms of the Creative Commons CC BY−NC license. (d)−(g) Reprinted with permission from Ref. [88], Copyright 2022, John Wiley and Sons.

    Figure 9.(Color online) (a) Schematic diagram of the fabrication process of the α/δ-FAPbI3 phase junction. (b) HRTEM of the α/δ-FAPbI3 phase junction. (c) Schematic diagrams of the energy levels of pure α-FAPbI3, pure δ-FAPbI3, and the α/δ-FAPbI3 phase junction. (d) Schematic diagram of the fabrication process of the bilayered δ-FAPbI3/perovskite films. (e) Current density−voltage curves for devices based on pristine perovskite films and bilayered δ-FAPbI3/perovskite films. (f) Statistics of PCEs and Jsc for devices based on pristine perovskite films and bilayered δ-FAPbI3/perovskite films. (g) Evolution of PCEs for devices based on pristine perovskite films and bilayered δ-FAPbI3/perovskite films under 40% ± 5% relative humidity (RH) at room temperature. (a)−(c) Reprinted with permission from Ref. [87], Copyright 2017, published under the terms of the Creative Commons CC BY−NC license. (d)−(g) Reprinted with permission from Ref. [88], Copyright 2022, John Wiley and Sons.

    Moreover, Zhang et al.[88] found that photo-inactive δ-FAPbI3 could formed in situ on the black phase (FAMAPbI3) surface, which serves as a surface water−oxygen barrier and reduces the surface defect density. As illustrated in Fig. 9(d), after the two-step spin-coating of the FAMAPbI3 thin film, FAI/IPA solution was then spin-coated on the surface, where the excess PbI2 on the pristine film would react with FAI to form a δ-FAPbI3 layer. The formed δ-FAPbI3 mainly distributed on the surface and at grain boundaries, which could not only act as a surface barrier to improve the stability of perovskite films, but also reduce the surface defect density to suppress the trap-assisted recombination. What’s more, the introduce of δ-FAPbI3 on FAMAPbI3 surface gives rise to a type-Ⅰ band alignment at the interface, which efficiently hinder the electron transfer to hole transporting layer (HTL) and inhibit the nonradiative recombination. Thus, the performance of solar cells based on this bilayered δ-FAPbI3/perovskite films has greatly improved compared to the pristine perovskite films, with power conversion efficiency (PCE) increased from 21.05% to 23.10%, short-circuit current (Jsc) increased from 25.0 to 25.3 mA∙cm−2, open-circuit voltage (Voc) increased from 1.12 to 1.16 V, as shown in Figs. 9(e) and 9(f). In addition, the evolution of PCE for the device based on bilayered δ-FAPbI3/perovskite films exhibit long-term stability (Fig. 9(g)), attributing to the stable δ-FAPbI3 locates at grain boundaries protecting the perovskite films from the moisture penetration.

    Besides, Mandal et al.[89] used well-crystallised δ-FAPbI3 single crystals as an intermediate phase for the formation of high-quality α-FAPbI3 to prepare high-performance solar cells. As shown in Fig. 10(a), δ-FAPbI3 single crystals are firstly precipitated utilizing anti-solvent, and then dissolved in solution to be spin-coated as films, which would further phase transition during annealing to form a high-quality α-FAPbI3 films. As shown in Fig. 10(b), the 1D hexagonal δ-FAPbI3 single crystals precipitated using anti-solvents exhibit excellent crystalline. Due to the reduced trap states in δ-FAPbI3 single crystals, the solar cells based on intermediate δ-FAPbI3 single crystals has larger Voc and Jsc (1.16 V and 24.62 mA∙cm−2, respectively) with a champion PCE of 22.61%, compared to solar cells’ PCE of 21.58% based on intermediate δ-FAPbI3 powder (Fig. 10(c)). It is noteworthy that such devices prepared by intermediate δ-FAPbI3 single crystals and powder show excellent long-term stability, as shown in Fig. 10(d). Although the δ-phase itself may not directly serve as a good optoelectronic active layer, high-performance devices can still be prepared through phase transition from the δ-phase to the α-phase, and this use of thermodynamic stable δ-FAPbI3 with good crystallinity provides a new strategy for its application.

    (Color online) (a) Schematic diagram of the fabrication process of δ-FAPbI3 crystals, and the fabrication process of α-FAPbI3 thin films using δ-FAPbI3 crystals. (b) SEM image of the synthesized δ-FAPbI3 crystal. (c) Current density−voltage curves for devices based on intermediate δ-FAPbI3 single crystals (target) and powder sample (control). (d) Evolution of PCEs for target and control devices under continuous 1 sun illumination. (a)−(d) Reprinted with permission from Ref. [89], Copyright 2023, John Wiley and Sons.

    Figure 10.(Color online) (a) Schematic diagram of the fabrication process of δ-FAPbI3 crystals, and the fabrication process of α-FAPbI3 thin films using δ-FAPbI3 crystals. (b) SEM image of the synthesized δ-FAPbI3 crystal. (c) Current density−voltage curves for devices based on intermediate δ-FAPbI3 single crystals (target) and powder sample (control). (d) Evolution of PCEs for target and control devices under continuous 1 sun illumination. (a)−(d) Reprinted with permission from Ref. [89], Copyright 2023, John Wiley and Sons.

    The yellow-phase δ-FAPbI3 and black-phase α-FAPbI3 have poor and good electrical transmission characteristics, respectively, and can be used as off and on elements in circuits. Therefore, it is possible to utilize the δ-to-α phase transition induced by an increase in temperature and pressure, and the circuit is set from off to on to realize the detection of temperature and pressure. Without the phase transition triggers, the transitioned α-FAPbI3 (on) still spontaneously transitions to the thermodynamically stable δ-FAPbI3 (off), enabling self-recovering temperature and pressure detection. If issues such as the recovery rate of α-to-δ phase transitions, and reversible phase transitions under pressure can be guaranteed are resolved, α/δ-FAPbI3 will enable promising self-recovering temperature and pressure detection[22]. Also, the poorly electrically conductive δ-FAPbI3 can inhibit the excess carriers in α-FAPbI3 to retard the degradation of FAPbI3[90]. By and large, based on the fact that δ-FAPbI3 is very different from α-FAPbI3 and is easy to transition between each other, there is still much room for exploration from the preparation of the δ-phase to the functional applications.

    Conclusion

    In this review, in response to the problem of thermodynamically stable δ-FAPbI3 at room temperature limiting the development of α-FAPbI3, we present a fresh perspective on the δ-FAPbI3 with worse optoelectronic performance. Although the unavoidable phase transition from α- to δ-FAPbI3 at room temperature makes the δ-phase interferes with the optical response and electrical transport of α-phase active layers, the unique properties of δ-FAPbI3 can also be used to realize functional applications in combination with α-FAPbI3. For example, we have utilized the much lower thermal and electrical transport properties of δ-FAPbI3 as spacer layers for α-FAPbI3 to reduce the crosstalk of photothermal terahertz detection. Therefore, we wish to provide another perspective on δ-FAPbI3 in this work to change the traditional view that δ-FAPbI3 is useless trash, and thus advance the development of FAPbI3 material systems.

    Firstly, the two phases’ internal carrier and phonon dynamics are comprehensive analyses to explain the contrast properties between α- and δ-FAPbI3. α-FAPbI3 is the cubic structure with corner-sharing octahedrons, while δ-FAPbI3 is the 1D hexagonal structure with face-sharing octahedrons. The 1D distorted face-sharing PbX6 octahedrons of δ-FAPbI3 give rise to the anharmonic electron−phonon coupling and leads to STEs under anharmonic electron−phonon coupling because the low-dimensional structure reduces the potential barrier for self-trapping. Therefore, due to the restricted internal carrier and phonon transport, δ-FAPbI3 exhibits significantly reduced electronic and thermal properties compared to the α-phase, where the carrier mobility of δ-phase is about 5% and the thermal diffusion coefficient is two-thirds as compared to the α-phase. Secondly, several strategies for controlled phase transitions between α- and δ-FAPbI3 are presented. Factors affecting the phase transitions include temperature, pressure (internal/external pressure), humidity, etc., some of which can lead to reversible phase transitions. For example, using the instantaneous high temperatures (>150 °C) of the focused laser in direct-laser-writing, the micro-region phase transition from δ-FAPbI3 to α-FAPbI3 can be achieved. Finally, some functional applications of δ-FAPbI3 in combination with α-FAPbI3 are given and analyzed according to previous reports and our experience.

    For the multi-phase FAPbI3, applications at room temperature are still dominated by the α- and δ- phases rather than the low-temperature β- and γ-phase. The internal electron−phonon coupling of the intrinsically low-dimensional δ-FAPbI3 always limits its optoelectronic properties, while α-FAPbI3 with excellent optoelectronic properties is also limited by the spontaneous phase transition due to the thermodynamic instability at room temperature. Therefore, follow-up ideas should still consider the complementary roles of α- and δ-FAPbI3 and utilize the mixed phases for functional applications, especially the phase stabilizing and the electrical and thermal spacing effect of δ-FAPbI3 on α-FAPbI3. In conclusion, we hope to introduce δ-FAPbI3 from another perspective and provide some insights into its unique properties, in the hope of advancing the following application of multi-phase FAPbI3.

    References

    [1] Z G Xiao, C Bi, Y C Shao et al. Efficient, high yield perovskite photovoltaic devices grown by interdiffusion of solution-processed precursor stacking layers. Energy Environ Sci, 7, 2619(2014).

    [2] Y H Deng, E Peng, Y C Shao et al. Scalable fabrication of efficient organolead trihalide perovskite solar cells with doctor-bladed active layers. Energy Environ Sci, 8, 1544(2015).

    [3] Y H Zhang, R J Sun, X Y Ou et al. Metal halide perovskite nanosheet for X-ray high-resolution scintillation imaging screens. ACS Nano, 13, 2520(2019).

    [4] D M Jang, K Park, D H Kim et al. Reversible halide exchange reaction of organometal trihalide perovskite colloidal nanocrystals for full-range band gap tuning. Nano Lett, 15, 5191(2015).

    [5] Y C Liu, Z Yang, D Cui et al. Two-inch-sized perovskite CH3NH3PbX3 (X = Cl, Br, I) crystals: Growth and characterization. Adv Mater, 27, 5176(2015).

    [6] Z A Xie, S R Sun, Y Yan et al. Refractive index and extinction coefficient of NH2CH = NH2PbI3 perovskite photovoltaic material. J Phys: Condens Matter, 29, 245702(2017).

    [7] G C Xing, N Mathews, S Y Sun et al. Long-range balanced electron- and hole-transport lengths in organic-inorganic CH3NH3PbI3. Science, 342, 344(2013).

    [8] S D Stranks, G E Eperon, G Grancini et al. Electron-hole diffusion lengths exceeding 1 micrometer in an organometal trihalide perovskite absorber. Science, 342, 341(2013).

    [9] A A Zhumekenov, M I Saidaminov, M A Haque et al. Formamidinium lead halide perovskite crystals with unprecedented long carrier dynamics and diffusion length. ACS Energy Lett, 1, 32(2016).

    [10] A H Slavney, T Hu, A M Lindenberg et al. A bismuth-halide double perovskite with long carrier recombination lifetime for photovoltaic applications. J Am Chem Soc, 138, 2138(2016).

    [11] K B Zheng, Q S Zhu, M Abdellah et al. Exciton binding energy and the nature of emissive states in organometal halide perovskites. J Phys Chem Lett, 6, 2969(2015).

    [13] W H Bai, T T Xuan, H Y Zhao et al. Perovskite light-emitting diodes with an external quantum efficiency exceeding 30%. Adv Mater, 35, 2302283(2023).

    [14] F Cao, J D Chen, D J Yu et al. Bionic detectors based on low-bandgap inorganic perovskite for selective NIR-I photon detection and imaging. Adv Mater, 32, 1905362(2020).

    [15] D J Yu, P Wang, F Cao et al. Two-dimensional halide perovskite as beta-ray scintillator for nuclear radiation monitoring. Nat Commun, 11, 3395(2020).

    [16] X B Xu, Z Y Han, Y S Zou et al. Miniaturized multispectral detector derived from gradient response units on single MAPbX3 microwire. Adv Mater, 34, 2108408(2022).

    [17] Q F Han, S H Bae, P Y Sun et al. Single crystal formamidinium lead iodide (FAPbI3): Insight into the structural, optical, and electrical properties. Adv Mater, 28, 2253(2016).

    [18] T Zdanowicz, T Rodziewicz, M Zabkowska-Waclawek. Theoretical analysis of the optimum energy band gap of semiconductors for fabrication of solar cells for applications in higher latitudes locations. Sol Energy Mater Sol Cells, 87, 757(2005).

    [19] Y Zhao, F Ma, Z H Qu et al. Inactive (PbI2)2RbCl stabilizes perovskite films for efficient solar cells. Science, 377, 531(2022).

    [20] H R Chen, Y T Chen, T Y Zhang et al. Advances to high-performance black-phase FAPbI3 perovskite for efficient and stable photovoltaics. Small Struct, 2, 2000130(2021).

    [21] O J Weber, D Ghosh, S Gaines et al. Phase behavior and polymorphism of formamidinium lead iodide. Chem Mater, 30, 3768(2018).

    [22] S J Jiang, Y L Luan, J I Jang et al. Phase transitions of formamidinium lead iodide perovskite under pressure. J Am Chem Soc, 140, 13952(2018).

    [23] X Y Chen, Y K Xia, Z W Zheng et al. In situ formation of δ-FAPbI3 at the perovskite/carbon interface for enhanced photovoltage of printable mesoscopic perovskite solar cells. Chem Mater, 34, 728(2022).

    [24] J Z Li, H Wang, D H Li. Self-trapped excitons in two-dimensional perovskites. Front Optoelectron, 13, 225-234(2020).

    [25] Z W Wang, Y Y Zhou, S P Pang et al. Additive-modulated evolution of HC(NH2)2PbI3 black polymorph for mesoscopic perovskite solar cells. Chem Mater, 27, 7149(2015).

    [26] T K Zhang, M Z Long, M C Qin et al. Stable and efficient 3D-2D perovskite-perovskite planar heterojunction solar cell without organic hole transport layer. Joule, 2, 2706(2018).

    [27] Y Zhang, Z M Zhou, F X Ji et al. Trash into treasure: δ-FAPbI3 polymorph stabilized MAPbI3 perovskite with power conversion efficiency beyond 21%. Adv Mater, 30, 1707143(2018).

    [28] H T Wei, S S Chen, J J Zhao et al. Is formamidinium always more stable than methylammonium. Chem Mater, 32, 2501(2020).

    [29] M Roß, S Severin, M B Stutz et al. Co-evaporated formamidinium lead iodide based perovskites with 1000 h constant stability for fully textured monolithic perovskite/silicon tandem solar cells. Adv Energy Mater, 11, 2101460(2021).

    [30] Y Wang, Y X Zhao, F Gao et al. Sequential cesium incorporation for highly efficient formamidinium cesium perovskite solar cells. ChemRxiv(2022).

    [31] W C Qiao, J Q Liang, W Dong et al. Formamidinium lead triiodide perovskites with improved structural stabilities and photovoltaic properties obtained by ultratrace dimethylamine substitution. NPG Asia Mater, 14, 49(2022).

    [32] M Pratheek, P Sujith, T A Shahul Hameed et al. Effect of bromide doping on the phase stability and shelf life of the triple cation mixed halide perovskite single-crystals. Mater Lett, 326, 132903(2022).

    [33] J W Yin, G Teobaldi, L M Liu. The role of thermal fluctuations and vibrational entropy: A theoretical insight into the δ-to-α transition of FAPbI3. J Phys Chem Lett, 13, 3089(2022).

    [34] J Ibaceta-Jaña, R Muydinov, P Rosado et al. Hidden polymorphism of FAPbI3 discovered by Raman spectroscopy. Phys Chem Chem Phys, 23, 9476(2021).

    [35] Y P Du, Z G Yan, J W Xiao et al. Temperature-dependent luminescence and anisotropic optical properties of centimeter-sized one-dimensional perovskite trimethylammonium lead iodide single crystals. J Phys Chem Lett, 13, 5451(2022).

    [36] K M McCall, C C Stoumpos, S S Kostina et al. Strong electron–phonon coupling and self-trapped excitons in the defect halide perovskites A3M2I9 (a = Cs, Rb; M = Bi, Sb). Chem Mater, 29, 4129(2017).

    [37] L Piveteau, V Morad, M V Kovalenko. Solid-state NMR and NQR spectroscopy of lead-halide perovskite materials. J Am Chem Soc, 142, 19413(2020).

    [38] A D Wright, C Verdi, R L Milot et al. Electron-phonon coupling in hybrid lead halide perovskites. Nat Commun, 7, 11755(2016).

    [39] Y Yamada, Y Kanemitsu. Electron-phonon interactions in halide perovskites. NPG Asia Mater, 14, 48(2022).

    [40] S Dubey, A Paliwal, S Ghosh. Frohlich interaction in compound semiconductors: A comparative study. Adv Mater Res, 1141, 44(2016).

    [41] M Fu, P Tamarat, J B Trebbia et al. Unraveling exciton-phonon coupling in individual FAPbI3 nanocrystals emitting near-infrared single photons. Nat Commun, 9, 3318(2018).

    [42] K Miyata, D Meggiolaro, M T Trinh et al. Large polarons in lead halide perovskites. Sci Adv, 3, e1701217(2017).

    [43] Y P Fu, M P Hautzinger, Z Y Luo et al. Incorporating large A cations into lead iodide perovskite cages: Relaxed Goldschmidt tolerance factor and impact on exciton–phonon interaction. ACS Cent Sci, 5, 1377(2019).

    [44] J Y Li, X Z Zeng, T Ma et al. Direct-laser-writing manipulating α/δ-FAPbI3 array enables low-crosstalk terahertz sensing and imaging. ACS Photonics, 10, 2755(2023).

    [45] M Aebli, B M Benin, K M McCall et al. White CsPbBr3: Characterizing the one-dimensional cesium lead bromide polymorph. Helv Chim Acta, 103, e2000080(2020).

    [46] F Cordero, F Craciun, F Trequattrini et al. Stability of cubic FAPbI3 from X-ray diffraction, anelastic, and dielectric measurements. J Phys Chem Lett, 10, 2463(2019).

    [47] A C Ferreira, A Létoublon, S Paofai et al. Elastic softness of hybrid lead halide perovskites. Phys Rev Lett, 121, 085502(2018).

    [48] M I Saidaminov, A L Abdelhady, B Murali et al. High-quality bulk hybrid perovskite single crystals within minutes by inverse temperature crystallization. Nat Commun, 6, 7586(2015).

    [49] G A Elbaz, W L Ong, E A Doud et al. Phonon speed, not scattering, differentiates thermal transport in lead halide perovskites. Nano Lett, 17, 5734(2017).

    [50] J F Yang, X M Wen, H Z Xia et al. Acoustic-optical phonon up-conversion and hot-phonon bottleneck in lead-halide perovskites. Nat Commun, 8, 14120(2017).

    [51] Y Jung, W Lee, S Han et al. Thermal transport properties of phonons in halide perovskites. Adv Mater, 35, 2204872(2023).

    [52] B Liu, C M M Soe, C C Stoumpos et al. Optical properties and modeling of 2D perovskite solar cells. Sol RRL, 1, 1700062(2017).

    [53] Z Tong, S H Li, X L Ruan et al. Comprehensive first-principles analysis of phonon thermal conductivity and electron-phonon coupling in different metals. Phys Rev B, 100, 144306(2019).

    [54] A Majumdar, P Reddy. Role of electron-phonon coupling in thermal conductance of metal-nonmetal interfaces. Appl Phys Lett, 84, 4768(2004).

    [55] S X Lin, X J Tan, H Z Shao et al. Ultralow lattice thermal conductivity in SnTe by manipulating the electron–phonon coupling. J Phys Chem C, 123, 15996(2019).

    [56] B L Liao, B Qiu, J W Zhou et al. Significant reduction of lattice thermal conductivity by the electron-phonon interaction in silicon with high carrier concentrations: A first-principles study. Phys Rev Lett, 114, 115901(2015).

    [57] J W Zhou, H D Shin, K Chen et al. Direct observation of large electron-phonon interaction effect on phonon heat transport. Nat Commun, 11, 6040(2020).

    [58] S Y Hu, A R Tapa, X C Zhou et al. Formation and stabilization of metastable halide perovskite phases for photovoltaics. Cell Rep Phys Sci, 5, 101825(2024).

    [59] T R Chen, B J Foley, C Park et al. Entropy-driven structural transition and kinetic trapping in formamidinium lead iodide perovskite. Sci Adv, 2, e1601650(2016).

    [60] Y M Chen, Y S Lei, Y H Li et al. Strain engineering and epitaxial stabilization of halide perovskites. Nature, 577, 209(2020).

    [61] J A Steele, H F Yuan, C Y X Tan et al. Direct laser writing of δ- to α-phase transformation in formamidinium lead iodide. ACS Nano, 11, 8072(2017).

    [62] Z Li, M J Yang, J S Park et al. Stabilizing perovskite structures by tuning tolerance factor: Formation of formamidinium and cesium lead iodide solid-state alloys. Chem Mater, 28, 284(2016).

    [63] X J Zheng, C C Wu, S K Jha et al. Improved phase stability of formamidinium lead triiodide perovskite by strain relaxation. ACS Energy Lett, 1, 1014(2016).

    [64] J W Lee, D H Kim, H S Kim et al. Formamidinium and cesium hybridization for photo- and moisture-stable perovskite solar cell. Adv Energy Mater, 5, 1501310(2015).

    [65] E Smecca, Y Numata, I Deretzis et al. Stability of solution-processed MAPbI3 and FAPbI3 layers. Phys Chem Chem Phys, 18, 13413(2016).

    [66] N J Jeon, J H Noh, W S Yang et al. Compositional engineering of perovskite materials for high-performance solar cells. Nature, 517, 476(2015).

    [67] A R Goñi. Echoes from quantum confinement. Nat Mater, 19, 1138(2020).

    [68] A D Wright, G Volonakis, J Borchert et al. Intrinsic quantum confinement in formamidinium lead triiodide perovskite. Nat Mater, 19, 1201(2020).

    [69] M Tonouchi. Cutting-edge terahertz technology. Nature Photon, 1, 97(2007).

    [70] H Cheon, H J Yang, S H Lee et al. Terahertz molecular resonance of cancer DNA. Sci Rep, 6, 37103(2016).

    [71] S Koenig, D Lopez-Diaz, J Antes et al. Wireless sub-THz communication system with high data rate. Nat Photonics, 7, 977(2013).

    [72] W C Shui, J M Li, H Wang et al. Ti3C2Tx MXene sponge composite as broadband terahertz absorber. Adv Opt Mater, 8, 2001120(2020).

    [73] Y Salamin, I C Benea-Chelmus, Y Fedoryshyn et al. Compact and ultra-efficient broadband plasmonic terahertz field detector. Nat Commun, 10, 5550(2019).

    [74] M Mittendorff, S Winnerl, J Kamann et al. Ultrafast graphene-based broadband THz detector. Appl Phys Lett, 103, 021113(2013).

    [75] F H L Koppens, T Mueller, P Avouris et al. Photodetectors based on graphene, other two-dimensional materials and hybrid systems. Nat Nanotechnol, 9, 780(2014).

    [76] X W Lu, L Sun, P Jiang et al. Progress of photodetectors based on the photothermoelectric effect. Adv Mater, 31, 1902044(2019).

    [77] M S Long, P Wang, H H Fang et al. Progress, challenges, and opportunities for 2D material based photodetectors. Adv Funct Mater, 29, 1803807(2019).

    [78] Y F Li, Y T Zhang, T T Li et al. Ultrabroadband, ultraviolet to terahertz, and high sensitivity CH3NH3PbI3 perovskite photodetectors. Nano Lett, 20, 5646(2020).

    [79] J Y Li, Y S Zou, D W Hu et al. Enhanced room-temperature terahertz detection and imaging derived from anti-reflection 2D perovskite layer on MAPbI3 single crystals. Nanoscale, 14, 6109(2022).

    [80] Y F Li, Y T Zhang, T T Li et al. A fast response, self-powered and room temperature near infrared-terahertz photodetector based on a MAPbI3/PEDOT: PSS composite. J Mater Chem C, 8, 12148(2020).

    [81] S J Park, A R Kim, J T Hong et al. Crystallization kinetics of lead halide perovskite film monitored by in situ terahertz spectroscopy. J Phys Chem Lett, 8, 401(2017).

    [82] X Y Tian, R N Wang, Y L Xu et al. Triangular micro-grating via femtosecond laser direct writing toward high-performance polarization-sensitive perovskite photodetectors. Adv Optical Mater, 10, 2200856(2022).

    [83] H R Chen, Y Wang, Y P Fan et al. Decoupling engineering of formamidinium-cesium perovskites for efficient photovoltaics. Natl Sci Rev, 9, nwac127(2022).

    [84] F Y Cheng, P D Wang, C Z Xu et al. The dynamic surface evolution of halide perovskites induced by external energy stimulation. Natl Sci Rev, 11, nwae042(2024).

    [85] T Zhang, K Xu, J Li et al. Ferroelectric hybrid organic-inorganic perovskites and their structural and functional diversity. Natl Sci Rev, 10, nwac240(2022).

    [86] S H Wang, Z P Miao, J Yang et al. Lead-chelating intermediate for air-processed phase-pure FAPbI3 perovskite solar cells. Angew Chem Int Ed, 63, e202407192(2024).

    [87] F S Ma, J W Li, W Z Li et al. Stable α/δ phase junction of formamidinium lead iodide perovskites for enhanced near-infrared emission. Chem Sci, 8, 800(2017).

    [88] J F Zhang, X Q Jiang, X T Liu et al. Maximizing merits of undesirable δ-fapbi3 by constructing yellow/black heterophase bilayer for efficient and stable perovskite photovoltaics. Adv Funct Mater, 32, 2204642(2022).

    [89] T N Mandal, J H Heo, S H Im et al. Green method to prepare pure δ-FAPbI3 crystals for fabrication of highly efficient perovskite solar cells. Sol RRL, 7, 2300496(2023).

    [90] S R Lee, D Lee, S G Choi et al. Accelerated degradation of FAPbI3 perovskite by excess charge carriers and humidity. Sol RRL, 8, 2300958(2024).

    Junyu Li, Songwei Zhang, Mohd Nazim Mohtar, Nattha Jindapetch, Istvan Csarnovics, Mehmet Ertugrul, Zhiwei Zhao, Jing Chen, Wei Lei, Xiaobao Xu. Advances in multi-phase FAPbI3 perovskite: another perspective on photo-inactive δ-phase[J]. Journal of Semiconductors, 2025, 46(5): 051804
    Download Citation