Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2025 May 22;10(6):2906–2912. doi: 10.1021/acsenergylett.5c00576

3D Chiral Metal Halide Semiconductors

Marco Moroni , Luca Gregori ‡,, Clarissa Coccia , Massimo Boiocchi §, Marta Morana , Doretta Capsoni , Andrea Olivati ⊥,#, Antonella Treglia , Giulia Folpini ⊥,, Maddalena Patrini , Isabel Goncalves , Heyong Wang , Chiara Milanese , Annamaria Petrozza , Edoardo Mosconi ●,*, Filippo De Angelis ‡,●,, Lorenzo Malavasi †,*
PMCID: PMC12172314  PMID: 40535462

Abstract

Chiral metal halides are promising materials for nonlinear optics and spin-selective devices. Typically, chirality is introduced via large chiral organic cations, leading to low-dimensional structures and limitations in charge transport. Here, we design a family of chiral metal halides based on the relatively small ditopic R/S-3-aminoquinuclidine (3-AQ) cation, forming an (R/S-3AQ)­Pb2Br6 structure closely related to the 3D corner-sharing octahedral network of perovskites. The resulting material exhibits a direct bandgap, isotropic band structure, and fully 3D photoexcitation. Circular dichroism confirms a chiral anisotropy factor consistent with theoretical predictions. Moreover, the material displays a Rashba effect in the conduction band, which is attributed to spin–orbit coupling and the lack of inversion symmetry. Offering rich chemical tunability and efficient 3D charge transport, this new class of chiral semiconductors provides a promising platform for advancing nonlinear optoelectronic and spintronic devices.


graphic file with name nz5c00576_0005.jpg


graphic file with name nz5c00576_0004.jpg


Metal halide perovskites constitute a class of wondrous materials, characterized by huge structural diversity associated with variations in both the inorganic (metal, halide) and organic (A-site cations) sublattices. Besides their success in solar cells and LEDs, such chemical richness can be exploited to engineer a variety of chiral compounds with associated spin-dependent chiroptical electronic properties. Potential technological applications of chiral perovskites range from circularly polarized light emission and detection, chiral sensing, and chiral-induced spin selectivity up to enantioselective synthesis and photocatalysis. Thus, the broad tunability of these materials arises from the interplay of organic and inorganic moieties through the chirality transfer mechanism, whereby chiral A-site organic cations impart chiroptical response to the optically active inorganic framework. The family of chiral 2D perovskites and low-dimensional metal halides (0D and 1D systems), first reported in 2003, , has grown impressively by exploring various organic cations and different metals. These low-dimensional structures were demonstrated to exhibit chiroptical properties as well as spin-polarized absorption, spin-polarized photoluminescence, and second harmonic generation. The inherent carrier and excitonic confinement of such low-dimensional materials and the associated anisotropic transport remain a challenge for their effective integration in devices.

The synthesis of 3D chiral perovskites would represent a significant advancement, lifting the above-mentioned limitations. It is now clear, however, that only small organic cations, such as methylammonium (CH3NH3 +) or formamidinium (NH2CHNH2 +)), can give rise to 3D perovskites which clearly have no chiral-templating nature. , A different approach was proposed by Chen and co-workers through the growth of MAPbBr3 single crystals with the addition of micro- or nanoparticles as nucleating agents. Chirality, in this case, emerges from chiral orientation patterns of the incorporated A-site cations through the formation of chiral supercells, inducing chiroptical activity. The heterogeneous chiral nucleation proposed, however, assures only a partial enantiomeric excess (i.e., the ratio of the two chiral enantiomers) so that control over the handedness (R/L) selectivity remains a major isssue. In addition, this synthetic process is limited to single crystals whereby thin films are the optimal choice for their embodiment in devices.

Here we face the challenge of achieving 3D chiral metal halide materials by making use of enantiopure cage-like ditopic amines, specifically exploiting R/S-3-aminoquinuclidine (3-AQ) (Figure a). The peculiar steric hindrance of this amine and its relatively small size allow the formation of a 3D chiral metal halide of formula (R/S-3AQ)­Pb2Br6 composed exclusively of PbBr6 corner-sharing octahedra, similar to the 3D connectivity of prototypical perovskite structures. The new compound presents a direct band gap with chiroptical properties in the circular dichroism response while clearly maintaining signatures typical of 3D semiconductors, as shown by its band dispersion and reduced exciton binding energy and band gap compared to its 2D perovskite counterpart, (R/S-3-AQ)2PbBr4·2Br. In line with such properties coupled to the noncentrosymmetric crystal structure, the 3D (R/S-3AQ)­Pb2Br6 chiral system is predicted to show a significant Rashba conduction band splitting that makes it potentially amenable for spintronics.

1.

1

(a) Chemical formula of 3-aminoquinuclidine (3-AQ). The asterisk marks the chiral center. (b–d) Sketches of the crystal structures of (R/S-3-AQ)­Pb2Br6 along the a, c, and b axes, respectively. The two crystallographically independent Pb ions are represented with their coordination octahedra in dark green (Pb1) and light green (Pb2), respectively. (e) Highlight of the (R-3-AQ)­Pb2Br6 N–H···Br hydrogen bonds interactions, represented as blue dashed lines.

Samples of (R/S-3AQ)­Pb2Br6 and of (R/S-3-AQ)2PbBr4·2Br have been synthesized as reported in Methods in the Supporting Information (SI). For (R-3AQ)­Pb2Br6 we grow high-quality single crystals which were used to solve the crystal structure by X-ray diffraction, and the main crystallographic data are reported in Table . For the S-enantiomer, however, this was hardly possible, and therefore, the crystal structure was Rietveld-refined starting from the single-crystal data of the R-enantiomer.

1. Crystal Structure Data for [(R/S-3AQ)2PbBr4]·2Br and (R/S-3AQ)­Pb2Br6 .

  [( R -3AQ)2PbBr4]·2Br [( S -3AQ)2PbBr4]·2Br ( R -3AQ)Pb2Br6 ( S -3AQ)Pb2Br6
Empirical formula C14H32N4Br6Pb C14H32N4Br6Pb C7H16N2Br6Pb2 C7H16N2Br6Pb2
Formula weight 943.05 943.05 1022.04 1022.04
Temperature (K) 298 298 298 298
Wavelength (Å) 0.7107 0.7107 0.7107 1.5406
Crystal system Tetragonal Tetragonal Orthorhombic Orthorhombic
Space group P41212 P43212 P212121 P212121
Lattice parameters (Å) a = 6.58258(8) a = 6.58253(5) a = 10.9133(2) a = 10.9184(2)
  b = 6.58258(8) b = 6.58253(5) b = 11.0771(2) b = 11.0820(2)
  c = 56.6268(14) c = 56.6576(8) c = 15.5426(3) c = 15.5470(3)
Lattice Volume (Å 3 ) 2453.66(9) 2454.96(5) 1878.91(6) 1881.14(6)
Z 4 4 4 4
CCDC code 2416647 2416648 2416219
a

For (S-3AQ)­Pb2Br6 we report the Rietveld refinement of powder diffraction pattern.

(R/S-3AQ)­Pb2Br6 crystallizes in the orthorhombic chiral space group P212121 (No. 19) at 293 K, providing therefore a noncentrosymmetric structure. The crystal structure of (R/S-3AQ)­Pb2Br6 is peculiar and, to our knowledge, not yet reported in the literature for any other metal halide: it is composed of two independent Pb sites (Pb1 and Pb2) forming two distinct types of octahedra with Br ions (see Figure b–d). Notably, the reported crystal structure presents an AB2X6 chemical formula analogous to that of typical perovskitoids.

From Figure one may notice the peculiar arrangement of the PbBr6 octahedra: [Pb1Br6] octahedra are corner-sharing arranged along the b-direction while [Pb2Br6] octahedra form an analogous corner-sharing pattern along the a-direction. These two perpendicular rows are then again corner-sharing bonded along the c-direction creating central voids along the a and b directions, which host the 3-AQ moieties. N–H···Br hydrogen bond interactions are established for both N atoms of 3-AQ, namely, the tertiary amine of the quinuclidine ring and the primary amine in position 3 of the ring, with H···Br distances in the 2.36–2.99 Å range (Figure e).

The PbBr6 octahedra present six different Pb–Br bond-lengths; these are, however, very close to each other for both [Pb1Br6] and [Pb2Br6], lying in the range of 2.93–3.09 Å. Interestingly, the average bond length for [Pb1Br6] is 3.0103 Å and 3.0077 Å for [Pb2Br6] octahedra which are only slightly longer than the Pb–Br bond length found in cubic MAPbBr3 (2.964 Å). Indeed, the Robinson distortion parameter, D, is very low for both octahedra types, 0.019, while the bond angle variance (σ2) is 42.7 and 51.9 deg2 for [Pb1Br6] and [Pb2Br6], respectively (see the list of structural parameters in Table S1, Supporting Information).

To note the difference between the (R-3-AQ)­Pb2Br6 3D compound and the typical chiral 2D perovskites, we synthesized also the (R/S-3-AQ)2PbBr4·2Br compound and solved the crystal structure by SC-XRD (Table ). The crystal symmetry of both R/S enantiomers is tetragonal and belongs to the two chiral Sohncke space groups P41212 and P43212, for the R and S sample, respectively, which are noncentrosymmetric and nonpolar space groups. A sketch of the crystal structure for the two samples is reported in Figure S1 (Supporting Information). Both enantiomers present the typical arrangement of Ruddlesden–Popper 2D perovskites characterized by layers of corner-sharing PbBr6 octahedra separated by a bilayer of 3-AQ molecules including an additional layer of bromide anions to ensure electroneutrality. The six Pb–Br bond lengths show a significant distribution from ∼2.78 to ∼ 3.81 Å for both samples, with an average bond length of about 3.20 Å (see also Table S1). The octahedra distortion parameters for (R-3-AQ)2PbBr4·2Br and (S-3-AQ)2PbBr4·2Br are D = 0.128 and σ2 = 40.77 deg2; thus, an increased octahedra distortion is found compared to the 3D system (0.019 vs 0.128).

For both systems we also afforded the preparation of the racemic samples in the form of powders. The X-ray diffraction patterns have been indexed and refined showing an analogous structural arrangement as the chiral compounds but with centrosymmetric space groups, namely Pmma for (rac-3-AQ)2PbBr4·2Br and C2 (rac-3-AQ)2PbBr4·2Br. The refined patterns are reported in Figure S2, while Table S2 lists the corresponding lattice parameters.

Thermal stability of the R-samples has been determined by thermogravimetric analysis in the 30–750 °C temperature range and by differential scanning calorimetry (DSC) in the range from −165 to 65 °C. The results, reported in Figure S3, show good stability of the samples up to 300 °C when the first weight loss occurs. Notably, DSC shows the absence of phase transitions in the considered temperature range, at variance with the typical 3D metal-halide perovskites showing the formation of high-temperature cubic phases. The phase stability of (R-3-AQ)2PbBr4·2Br and (R-3-AQ)­Pb2Br6 has been further explored by collecting XRD patterns of thin films prepared by spin coating (see Methods in the Supporting Information) left in open air in the laboratory environment (T ≈ 22 °C, RH ≈ 40%) as a function of time for up to 47 days, showing significant air and moisture stability (see Figure S4).

To gain insight into the electronic structure of the 3D and 2D samples, we have performed density functional theory (DFT) calculations; see the Supporting Information for computational details. The structures were optimized starting from the crystallographic data at the scalar relativistic PBE level, followed by higher-level HSE06 (α = 0.43) calculations including spin–orbit coupling (SOC). Direct band gap values of 3.46 eV for (R/S-3-AQ)­Pb2Br6 and 3.82 eV for (R/S-3-AQ)2PbBr4·2Br are calculated, at Γ and M points, respectively (see Figure and Table S3).

2.

2

(a, c) Electronic band structures and projected density of states (PDOS) for the 3D and 2D materials, respectively. The PDOS was computed using the HSE06-SOC level of theory, whereas the band structures were initially obtained with PBE-SOC and subsequently corrected to align with the HSE06-SOC band gap. The insets in the PDOS panels illustrate the corresponding Brillouin zone representations. (b, d) Rashba splitting of the valence and conduction bands for the respective materials, highlighting spin–orbit coupling effects. The energy zero reference is set to the valence band maximum (VBM) and conduction band minimum (CBM) to enhance the visualization of the splitting.

The projected density of states (PDOS) of the R-isomers, Figure a–c, shows that the band edges in both 3D and 2D materials are dominated by Pb and Br contributions, while the C, N, and H atoms from the organic molecules contribute to states far away from the band edges, similar to typical lead-halide perovskites. The k-point path used for the band structure calculations, shown in the inset in Figure , was selected according to standardized high-symmetry paths generated by SeeK-path and validated with the Bilbao Crystallographic Server, ensuring an accurate and complete representation of the electronic structure starting from the crystallographic information obtained from the experiments. , Figure a–c shows a direct band gap for both structural arrangements. Importantly, the 2D band structure shows a flat band dispersion along the direction perpendicular to the inorganic perovskite layers, i.e., M → A (see Figure S11 in the Supporting Information), while the (R/S-3-AQ)­Pb2Br6 system has a clearly isotropic band structure, indicating that the charge carriers can be transported along the 3D octahedra network. Carrier effective masses for (R/S-3-AQ)­Pb2Br6 were determined by parabolic fitting along the Γ → Y/Γ → Z in the band gap region, finding m h* and m e* values of 0.54/0.66 and 0.34/0.53 m 0, respectively, where m 0 is the electron mass. Effective masses along other crystallographic directions are listed in Table S3. An exciton reduced mass of 0.21/0.29 m 0 is thus estimated. It is interesting to compare calculated effective masses for (R/S-3-AQ)­Pb2Br6 to those of typical 3D perovskites, e.g., MAPbBr3. Although the values obtained for our system are a factor ∼1.5 times higher than those for MAPbBr3 (m h* and m e* are 0.31 and 0.27 m0, respectively), a similar trend is observed, with electrons being more mobile than holes.

Notably, both 3D and 2D systems display a Rashba/Dresselhaus spin splitting of the electronic band structure (Figure b–d), a characteristic that can be exploited in spintronic applications. The Rashba splitting parameters, including the energy splitting ε c/v , moment offset (Δk), and Rashba coefficient (α), are reported in Tables S4 and S5. For (R/S-3-AQ)­Pb2Br6 a significant Rashba effect is observed in the conduction band with α = 2.22 eV Å. In contrast, the valence band shows weaker Rashba splitting, indicating minimal spin–orbit interactions in this band. On the other hand, high values of α were found for the 2D (R/S-3-AQ)2PbBr4·2Br (7.12 and 8.19 for the valence and conduction band, respectively) in line with a higher octahedra distortion for the 2D species with respect to the 3D (see Table S1).

Structural and electronic structure calculations confirm a 3D character for the novel (R/S-3-AQ)­Pb2Br6 chiral compound. The optical properties of the 3D and 2D systems have thus been measured by UV–vis and PL spectroscopy and by circular dichroism (CD). The absorption measurements at 300 K are shown in Figure a,b for the four samples as Tauc plots, while Figures S5 and S6 report a comparison of room-temperature data with those recorded at 77 K. We observe absorption features at 3.39 and 3.64 eV for (R/S-3-AQ)­Pb2Br6 and for (R/S-3-AQ)2PbBr4·2Br, respectively, compatible with excitonic absorption. The band gap, as identified by fitting the subsequent absorption rise, is found at 3.72 and 4.05 eV for the 3D and 2D structures, respectively. Both data are in good agreement with calculated values; in particular, the band gap increases moving from 3D to 2D structures (0.33 vs 0.36 eV). Thus, for both the excitonic peak and band gap, a noticeable blue shift is observed when moving to the 2D structural arrangement, consistent with a more delocalized excited state and band dispersion in the 3D case. For the (R/S-3-AQ)­Pb2Br6, the exciton binding energy, E b, is 480 meV, consistent with the presence of a stable excitonic population at room temperature (Figure a). For (R/S-3-AQ)2PbBr4·2Br, the exciton binding energy is 500–540 meV for the R and S enantiomers, respectively, consistent with a higher degree of confinement in the 2D arrangement.

3.

3

Tauc plot for (a) (R/S-3AQ)­Pb2Br6 and (b) (R/S-3-AQ)2PbBr4·2Br at 300 K; CD spectra for (c) (R/S-3AQ)­Pb2Br6 and (d) (R/S-3-AQ)2PbBr4·2Br.

The PL emission of (R/S-3AQ)­Pb2Br6, Figure S7, is centered at 1.7 eV, with a significant Stokes shift and a bandwidth of about 500 meV. For the 2D system the photoluminescence spectra are even more broadband, peaking at about 2.3 eV, with a bandwidth up to 800 meV, consistent with the high distortion of the octahedral layers (see Figure S7 and Table S1).

Lastly, we characterized the CD response of thin films of 300 nm average thickness; see Methods in the Supporting Information. The XRD patterns of the thin films are shown in Figure S8, indicating a single-phase nature. Their morphology has been investigated by atomic force microscopy, and representative images are reported in Figure S9. The CD spectra of (R/S-3-AQ)­Pb2Br6, Figure c, present distinct peaks in the range of 315–400 nm, with the one centered at a higher wavelength corresponding to the absorption edge. The presence of an opposite sign in the CD peaks for the R and S samples confirms the chirality transfer from the ditopic 3-AQ chiral cation to the Pb–Br inorganic framework. The chiral anisotropy factor, g CD, has been calculated according to the following equation:

gCD=ALAR2(AL+AR) 1

where A L and A R are the absorbance of the sample for left- and right-handed circularly polarized light, respectively, and resulted in 6 × 10–5 for (R-3-AQ)­Pb2Br6 and −9 × 10–5 for (S-3-AQ)­Pb2Br6. The CD spectra for the 2D perovskites are shown in Figure d. A single peak with opposite sign around 345–350 nm (again in line with the absorption edge, cfr. Figure S6) can be observed, with a slight blue-shift with respect to the 3D systems. g CD values are 4 × 10–4 for (R-3-AQ)2PbBr4·2Br and −3 × 10–4 for (S-3-AQ)2PbBr4·2Br. Interestingly, the chiral anisotropy factor is 1 order of magnitude larger for the 2D perovskite with respect to the 3D compound. Such a result further supports the correlation between the structural dimensionality and g CD values. Indeed, it has been recently shown that there is a progressive reduction of roughly 1 order of magnitude of the anisotropy factor passing from 0D to 1D to 2D chiral systems. ,, Thus, the first chiral 3D system characterized by only corner-sharing octahedra allows extending the correlation, further confirming such a progressive reduction which seems closely related to the higher amount of chiral ligand in lower-dimensional systems and the reduced lattice rigidity. , CPL measurements at 77 K (Figure S10) further confirm the trend of dimensionality also on emission properties. The 2D system has a detectable CPL with g CPL of about 0.2%, while in contrast, the 3D system does not show appreciable CPL.

In summary, a new chiral metal halide semiconductor with a 3D structure, namely, (R/S-3-AQ)­Pb2Br6, has been synthesized by using a small and sterically hindered ditopic cation, providing the first chiral halide structural network composed of corner-shared octahedra only. The calculation of the electronic band structure for (R/S-3-AQ)­Pb2Br6 demonstrates an improved isotropic carrier mobility with respect to the (R/S-3-AQ)2PbBr4·2Br 2D system, which shows transport confined in the plane of the inorganic layers. These findings are corroborated by the photophysical properties of the semiconductor which shows a lower degree of localization of the photoexcitation, and of excitation nature, with respect to its 2D counterpart. The novel 3D system reported has a clear chiroptical response in the CD spectra which scales with the structural dimensionality and shows a sizable Rashba splitting in the conduction band.

Overall, this novel family of chiral 3D hybrid metal halide semiconductors has the potential to overcome the limitation of anisotropic transport of actual chiral systems, paving the way for the development of chiroptical and spintronic devices. Moreover, the well-recognized properties of tunability of hybrid metal halides through chemical alloying could be applied to (R/S-3-AQ)­Pb2Br6 to further manipulate its chiroptical and transport properties, thus creating a library of novel chiral 3D systems.

Supplementary Material

Acknowledgments

L.M. acknowledges support from the Ministero dell’Università e della Ricerca (MUR) and the University of Pavia through the program “Dipartimenti di Eccellenza 2023–2027”. We acknowledge support from the Italian Ministry of Research under the PRIN 2022 Grant No 2022F2K7J5 with the title “Two-dimensional chiral hybrid organic–inorganic perovskites for chiroptoelectronics” PE 3 funded by PNRR Mission 4 Istruzione e Ricerca - Component C2 - Investimento 1.1, Fondo per il Programma Nazionale di Ricerca e Progetti di Rilevante Interesse Nazionale PRIN 2022 - CUP B53D23004130006. G.F. acknowledges financial support from the European Union’s NextGenerationEU Programme with the I-PHOQS Infrastructure [IR0000016, ID D2B8D520, CUP B53C22001750006]. We acknowledge Mireille Thimon of KEP Technologies EMEA for DSC measurements. H.W. is a Marie Skłodowska-Curie Fellow (No. 101105123). L.M. and M.M. acknowledge Fondazione Cariplo through the program “Young Researchers”, Grant No. 2023-1246. AO, AP, AT acknowledge the support by the Italian Ministry of Environment and Energy Security in the framework of the Project GoPV (CSEAA_00011) for Research on the Electric System. FDA, AP and EM acknowledge the support from Horizon Europe Research and Innovation Action programme for VALHALLA project under Grant Agreement no. 101082176

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsenergylett.5c00576.

  • Methods; structural and computational details; and additional figures presenting structural, thermal, optical, and morphological properties (PDF)

The authors declare no competing financial interest.

References

  1. Dang Y., Liu X., Cao B., Tao X.. Chiral Halide Perovskite Crystals for Optoelectronic Applications. Matter. 2021;4(3):794–820. doi: 10.1016/j.matt.2020.12.018. [DOI] [Google Scholar]
  2. Long G., Sabatini R., Saidaminov M. I., Lakhwani G., Rasmita A., Liu X., Sargent E. H., Gao W.. Chiral-Perovskite Optoelectronics. Nat. Rev. Mater. 2020;5(6):423–439. doi: 10.1038/s41578-020-0181-5. [DOI] [Google Scholar]
  3. Malavasi, L. ; Moroni, M. ; Coccia, C. . Chiral Metal Halide Perovskites: Focus on Lead-Free Materials and Structure-Property Correlations; preprint; Chemistry and Materials Science, 2023. 10.20944/preprints202307.0836.v1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Min J., Choi Y., Kim D., Park T.. Beyond Imperfections: Exploring Defects for Breakthroughs in Perovskite Solar Cell Research. Adv. Energy Mater. 2024;14(6):2302659. doi: 10.1002/aenm.202302659. [DOI] [Google Scholar]
  5. Wang H., Treglia A., Albaqami M. D., Gao F., Petrozza A.. Tin-Halide Perovskites for Near-Infrared Light-Emitting Diodes. ACS Energy Lett. 2024;9(6):2500–2507. doi: 10.1021/acsenergylett.4c00470. [DOI] [Google Scholar]
  6. Yuan F., Folpini G., Liu T., Singh U., Treglia A., Lim J. W. M., Klarbring J., Simak S. I., Abrikosov I. A., Sum T. C., Petrozza A., Gao F.. Bright and Stable Near-Infrared Lead-Free Perovskite Light-Emitting Diodes. Nat. Photonics. 2024;18(2):170–176. doi: 10.1038/s41566-023-01351-5. [DOI] [Google Scholar]
  7. Luo X., Liu X., Lin X., Wu T., Wang Y., Han Q., Wu Y., Segawa H., Han L.. Recent Advances of Inverted Perovskite Solar Cells. ACS Energy Lett. 2024;9(4):1487–1506. doi: 10.1021/acsenergylett.4c00140. [DOI] [Google Scholar]
  8. Xing S., Yuan Y., Zhang G., Zhang S., Lian Y., Tang W., Zhou K., Liu S., Gao Y., Ren Z., Zhang G., Sun T., Zhao B., Di D.. Energy-Efficient Perovskite LEDs with Rec. 2020 Compliance. ACS Energy Lett. 2024;9(8):3643–3651. doi: 10.1021/acsenergylett.4c01501. [DOI] [Google Scholar]
  9. Chen C., Gao L., Gao W., Ge C., Du X., Li Z., Yang Y., Niu G., Tang J.. Circularly Polarized Light Detection Using Chiral Hybrid Perovskite. Nat. Commun. 2019;10(1):1927. doi: 10.1038/s41467-019-09942-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Lee C. U., Lee H., Jeong C.-S., Ma S., Jang G., Park Y. S., Yun J., Lee J., Son J., Jeong W., Yang S., Park J. H., Woo K., Moon J.. Enhanced Stability of Spin-Dependent Chiral 2D Perovskite Embedded PV-Biased Anode via Cross-Linking Strategy. ACS Energy Lett. 2024;9(8):4032–4043. doi: 10.1021/acsenergylett.4c01721. [DOI] [Google Scholar]
  11. Chen D., Tang B., Sergeev A. A., Wu Y., Liu H., Zhu D., Hu S., Wong K. S., Yip H.-L., Rogach A. L.. Green Spin Light-Emitting Diodes Enabled by Perovskite Nanocrystals in Situ Modified with Chiral Ligands. ACS Energy Lett. 2025;10(2):815–821. doi: 10.1021/acsenergylett.4c03515. [DOI] [Google Scholar]
  12. Ma S., Ahn J., Moon J.. Chiral Perovskites for Next-Generation Photonics: From Chirality Transfer to Chiroptical Activity. Adv. Mater. 2021;33(47):2005760. doi: 10.1002/adma.202005760. [DOI] [PubMed] [Google Scholar]
  13. Billing D. G., Lemmerer A.. Bis­[(S)-β-Phenethylammonium] Tribromoplumbate­(II) Acta Crystallogr. E Struct Rep. Online. 2003;59(6):m381–m383. doi: 10.1107/S1600536803010985. [DOI] [Google Scholar]
  14. Billing D. G., Lemmerer A.. Synthesis, Characterization and Phase Transitions in the Inorganic-Organic Layered Perovskite-Type Hybrids [(C n H 2 n+1 NH 3) 2 PbI 4], n = 4, 5 and 6. Acta Crystallogr. B Struct Sci. 2007;63(5):735–747. doi: 10.1107/S0108768107031758. [DOI] [PubMed] [Google Scholar]
  15. Coccia C., Morana M., Mahata A., Kaiser W., Moroni M., Albini B., Galinetto P., Folpini G., Milanese C., Porta A., Mosconi E., Petrozza A., De Angelis F., Malavasi L.. Ligand-Induced Chirality in ClMBA 2 SnI 4 2D Perovskite**. Angew. Chem. Int. Ed. 2024;63:e202318557. doi: 10.1002/anie.202318557. [DOI] [PubMed] [Google Scholar]
  16. Wang H., Li J., Lu H., Gull S., Shao T., Zhang Y., He T., Chen Y., He T., Long G.. Chiral Hybrid Germanium­(II) Halide with Strong Nonlinear Chiroptical Properties. Angew. Chem. Int. Ed. 2023;62(41):e202309600. doi: 10.1002/anie.202309600. [DOI] [PubMed] [Google Scholar]
  17. Asensio Y., Bahmani Jalali H., Marras S., Gobbi M., Casanova F., Mateo-Alonso A., Di Stasio F., Rivilla I., Hueso L. E., Martín-García B.. Circularly Polarized Photoluminescence in Chiral Hybrid Organic-Inorganic Manganese Halide Perovskites: From Bulk Materials to Exfoliated Flakes. Advanced Optical Materials. 2024;12:2400554. doi: 10.1002/adom.202400554. [DOI] [Google Scholar]
  18. Dibenedetto A., Coccia C., Boiocchi M., Moroni M., Milanese C., Malavasi L.. Synthesis and Characterization of Cu-Containing Chiral Metal Halides and Role of Halogenation of the Organic Ligand. J. Phys. Chem. C. 2024;128:4803. doi: 10.1021/acs.jpcc.4c00150. [DOI] [Google Scholar]
  19. Hao J., Lu H., Mao L., Chen X., Beard M. C., Blackburn J. L.. Direct Detection of Circularly Polarized Light Using Chiral Copper Chloride-Carbon Nanotube Heterostructures. ACS Nano. 2021;15(4):7608–7617. doi: 10.1021/acsnano.1c01134. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Peng H., Liu Q., Lu Y.-Z., Yang S.-J., Qi J.-C., Chen X.-G., Liao W.-Q.. A Chiral Two-Dimensional Perovskite-like Lead-Free Bismuth­(iii) Iodide Hybrid with High Phase Transition Temperature. Chem. Commun. 2023;59(68):10295–10298. doi: 10.1039/D3CC02798H. [DOI] [PubMed] [Google Scholar]
  21. Wu P., Pietropaolo A., Fortino M., Bando M., Maeda K., Nishimura T., Shimoda S., Sato H., Naga N., Nakano T.. Amplified Chirality Transfer to Aromatic Molecules through Non-specific Inclusion by Amorphous, Hyperbranched Poly­(Fluorenevinylene) Derivatives. Angew. Chem. Int. Ed. 2023;62(29):e202305747. doi: 10.1002/anie.202305747. [DOI] [PubMed] [Google Scholar]
  22. Moroni M., Coccia C., Malavasi L.. Chiral 2D and Quasi-2D Hybrid Organic Inorganic Perovskites: From Fundamentals to Applications. Chem. Commun. 2024;60:9310. doi: 10.1039/D4CC03314K. [DOI] [PubMed] [Google Scholar]
  23. Long G., Jiang C., Sabatini R., Yang Z., Wei M., Quan L. N., Liang Q., Rasmita A., Askerka M., Walters G., Gong X., Xing J., Wen X., Quintero-Bermudez R., Yuan H., Xing G., Wang X. R., Song D., Voznyy O., Zhang M., Hoogland S., Gao W., Xiong Q., Sargent E. H.. Spin Control in Reduced-Dimensional Chiral Perovskites. Nature Photon. 2018;12(9):528–533. doi: 10.1038/s41566-018-0220-6. [DOI] [Google Scholar]
  24. Ma J., Fang C., Chen C., Jin L., Wang J., Wang S., Tang J., Li D.. Chiral 2D Perovskites with a High Degree of Circularly Polarized Photoluminescence. ACS Nano. 2019;13(3):3659–3665. doi: 10.1021/acsnano.9b00302. [DOI] [PubMed] [Google Scholar]
  25. Li Z., Hong E., Zhang X., Deng M., Fang X.. Perovskite-Type 2D Materials for High-Performance Photodetectors. J. Phys. Chem. Lett. 2022;13(5):1215–1225. doi: 10.1021/acs.jpclett.1c04225. [DOI] [PubMed] [Google Scholar]
  26. Saparov B., Mitzi D. B.. Organic-Inorganic Perovskites: Structural Versatility for Functional Materials Design. Chem. Rev. 2016;116(7):4558–4596. doi: 10.1021/acs.chemrev.5b00715. [DOI] [PubMed] [Google Scholar]
  27. Mao L., Stoumpos C. C., Kanatzidis M. G.. Two-Dimensional Hybrid Halide Perovskites: Principles and Promises. J. Am. Chem. Soc. 2019;141(3):1171–1190. doi: 10.1021/jacs.8b10851. [DOI] [PubMed] [Google Scholar]
  28. Chen G., Liu X., An J., Wang S., Zhao X., Gu Z., Yuan C., Xu X., Bao J., Hu H.-S., Li J., Wang X.. Nucleation-Mediated Growth of Chiral 3D Organic-Inorganic Perovskite Single Crystals. Nat. Chem. 2023;15(11):1581–1590. doi: 10.1038/s41557-023-01290-2. [DOI] [PubMed] [Google Scholar]
  29. Lesne E., Fu Y., Oyarzun S., Rojas-Sánchez J. C., Vaz D. C., Naganuma H., Sicoli G., Attané J.-P., Jamet M., Jacquet E., George J.-M., Barthélémy A., Jaffrès H., Fert A., Bibes M., Vila L.. Highly Efficient and Tunable Spin-to-Charge Conversion through Rashba Coupling at Oxide Interfaces. Nat. Mater. 2016;15(12):1261–1266. doi: 10.1038/nmat4726. [DOI] [PubMed] [Google Scholar]
  30. Liang A., Gonzalez-Platas J., Turnbull R., Popescu C., Fernandez-Guillen I., Abargues R., Boix P. P., Shi L.-T., Errandonea D.. Reassigning the Pressure-Induced Phase Transitions of Methylammonium Lead Bromide Perovskite. J. Am. Chem. Soc. 2022;144(43):20099–20108. doi: 10.1021/jacs.2c09457. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Robinson K., Gibbs G. V., Ribbe P. H.. Quadratic Elongation: A Quantitative Measure of Distortion in Coordination Polyhedra. Science. 1971;172(3983):567–570. doi: 10.1126/science.172.3983.567. [DOI] [PubMed] [Google Scholar]
  32. Poglitsch A., Weber D.. Dynamic Disorder in Methylammoniumtrihalogenoplumbates (II) Observed by Millimeter-Wave Spectroscopy. J. Chem. Phys. 1987;87(11):6373–6378. doi: 10.1063/1.453467. [DOI] [Google Scholar]
  33. Gregori L., Meggiolaro D., De Angelis F.. Quantifying the Effect of Interfacial Dipoles on the Energy Level Alignment of Metal-Halide Perovskites. ACS Energy Lett. 2024;9(11):5329–5333. doi: 10.1021/acsenergylett.4c01876. [DOI] [Google Scholar]
  34. Marchal N., Mosconi E., García-Espejo G., Almutairi T. M., Quarti C., Beljonne D., De Angelis F.. Cation Engineering for Resonant Energy Level Alignment in Two-Dimensional Lead Halide Perovskites. J. Phys. Chem. Lett. 2021;12(10):2528–2535. doi: 10.1021/acs.jpclett.0c03843. [DOI] [PubMed] [Google Scholar]
  35. Meggiolaro D., Mosconi E., Proppe A. H., Quintero-Bermudez R., Kelley S. O., Sargent E. H., De Angelis F.. Energy Level Tuning at the MAPbI3 Perovskite/Contact Interface Using Chemical Treatment. ACS Energy Lett. 2019;4(9):2181–2184. doi: 10.1021/acsenergylett.9b01584. [DOI] [Google Scholar]
  36. Umari P., Mosconi E., De Angelis F.. Relativistic GW Calculations on CH3NH3PbI3 and CH3NH3SnI3 Perovskites for Solar Cell Applications. Sci. Rep. 2014;4(1):4467. doi: 10.1038/srep04467. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Hinuma Y., Pizzi G., Kumagai Y., Oba F., Tanaka I.. Band Structure Diagram Paths Based on Crystallography. Comput. Mater. Sci. 2017;128:140–184. doi: 10.1016/j.commatsci.2016.10.015. [DOI] [Google Scholar]
  38. Aroyo M. I., Kirov A., Capillas C., Perez-Mato J. M., Wondratschek H.. Bilbao Crystallographic Server. II. Representations of Crystallographic Point Groups and Space Groups. Acta Crystallogr. A Found Crystallogr. 2006;62(2):115–128. doi: 10.1107/S0108767305040286. [DOI] [PubMed] [Google Scholar]
  39. Mosconi E., Umari P., De Angelis F.. Electronic and Optical Properties of MAPbX3 Perovskites (X = I, Br, Cl): A Unified DFT and GW Theoretical Analysis. Phys. Chem. Chem. Phys. 2016;18(39):27158–27164. doi: 10.1039/C6CP03969C. [DOI] [PubMed] [Google Scholar]
  40. Mosconi E., Etienne T., De Angelis F.. Rashba Band Splitting in Organohalide Lead Perovskites: Bulk and Surface Effects. J. Phys. Chem. Lett. 2017;8(10):2247–2252. doi: 10.1021/acs.jpclett.7b00328. [DOI] [PubMed] [Google Scholar]
  41. Etienne T., Mosconi E., De Angelis F.. Dynamical Origin of the Rashba Effect in Organohalide Lead Perovskites: A Key to Suppressed Carrier Recombination in Perovskite Solar Cells? J. Phys. Chem. Lett. 2016;7(9):1638–1645. doi: 10.1021/acs.jpclett.6b00564. [DOI] [PubMed] [Google Scholar]
  42. Jana M. K., Song R., Liu H., Khanal D. R., Janke S. M., Zhao R., Liu C., Valy Vardeny Z., Blum V., Mitzi D. B.. Organic-to-Inorganic Structural Chirality Transfer in a 2D Hybrid Perovskite and Impact on Rashba-Dresselhaus Spin-Orbit Coupling. Nat. Commun. 2020;11(1):4699. doi: 10.1038/s41467-020-18485-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Coccia C., Moroni M., Treglia A., Boiocchi M., Yang Y., Milanese C., Morana M., Capsoni D., Porta A., Petrozza A., Stroppa A., Malavasi L.. Unraveling the Role of Structural Topology on Chirality Transfer and Chiroptical Properties in Chiral Germanium Iodides. J. Am. Chem. Soc. 2024;146:24377–24388. doi: 10.1021/jacs.4c05992. [DOI] [PubMed] [Google Scholar]
  44. Moroni M., Coccia C., Malavasi L.. Chiral 2D and Quasi-2D Hybrid Organic Inorganic Perovskites: From Fundamentals to Applications. Chem. Commun. 2024;60(70):9310–9327. doi: 10.1039/D4CC03314K. [DOI] [PubMed] [Google Scholar]
  45. Das R., Hossain M., Mahata A., Swain D., De Angelis F., Santra P. K., Sarma D. D.. Unique Chiro-Optical Properties of the Weakly-2D (R-/S-MBA) 2 CuBr 4 Hybrid Material. ACS Materials Lett. 2023;5:1556–1564. doi: 10.1021/acsmaterialslett.3c00268. [DOI] [Google Scholar]
  46. Zhang Z., Wang Z., Sung H. H.-Y., Williams I. D., Yu Z.-G., Lu H.. Revealing the Intrinsic Chiroptical Activity in Chiral Metal-Halide Semiconductors. J. Am. Chem. Soc. 2022;144(48):22242–22250. doi: 10.1021/jacs.2c10309. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials


Articles from ACS Energy Letters are provided here courtesy of American Chemical Society

RESOURCES