Open Access Article
Cássio C. S.
Soares
ab,
Carlos Mera
Acosta
a,
Fábio F.
Ferreira
a,
Aryane
Tofanello
a,
Mayra A. P.
Gómez
b,
Alejandro P.
Ayala
b,
J. R.
Toledo
c,
Yara Galvão
Gobato
c,
Maykon A.
Lemes
a,
Carlos W. A.
Paschoal
b and
José A.
Souza
*a
aCenter for Human and Natural Sciences, Federal University of ABC, Santo André, Brazil. E-mail: joseantonio.souza@ufabc.edu.br
bDepartment of Physics, Federal University of Ceará, Fortaleza, Brazil
cFederal University of São Carlos, Department of Physics, São Carlos, Brazil
First published on 14th October 2025
Incorporating chiral organic cations into hybrid organic–inorganic perovskites (HOIPs) offers a unique strategy for tailoring structural dimensionality and modulating photophysical properties. Here, we report that lead bromide perovskites synthesized with chiral ligands R- and S-α-methylbenzylammonium (MBA) lead to the formation of two-dimensional (2D) Ruddlesden–Popper structures, while the racemic mixture induces the formation of a hydrated 1D phase, (Rac-MBA)3PbBr5·H2O. DFT calculations show that this transition is driven by electrostatic strain from opposing chiral distortions, which favor octahedral connections with compensated dipoles. Water molecules occupy sites between inorganic chains and mediate hydrogen bonds, stabilizing the 1D motif and relieving strain from racemic reordering. These waters form symmetry-related interstitial networks that mirror the chiral organization, acting as electrostatic compensators in the absence of lateral Pb–Br bonding. Theoretical and experimental characterization studies show that the hydrated structure is energetically stable only when four symmetry-related 1D chains and water are present; otherwise, it becomes metallic or unstable. Electronic calculations reveal Rashba-type spin splitting in chiral phases, and the formation of Br-derived mid-gap states and a Fermi level near the conduction band in the racemic compound. This hydrated phase exhibits pronounced broadband photoluminescence attributed to self-trapped excitons (STEs), and temperature-dependent optical and structural studies reveal that exciton localization and emission dynamics are closely related to octahedral distortions and hydration effects. Our findings demonstrate that stereochemical engineering can offer a robust strategy to modulate dimensionality and optical functionality, with implications for broadband light-emitting applications.
The incorporation of chiral organic cations introduces new degrees of freedom in perovskite physics.12–15 Beyond structural templating, chirality can be transferred from the organic to the inorganic framework, potentially leading to novel optical activity, spin-selective transport, and enhanced light–matter interactions.16–20 Enantiopure organic cations favor the formation of well-ordered 2D Ruddlesden–Popper structures. At the same time, racemic mixtures can often result in disordered assemblies introducing packing frustration and symmetry breaking, thereby leading to different phases and new photophysical behavior.21,22 However, the mechanisms by which racemic and enantiopure cations affect the structural dimensionality and enable broadband photoluminescence, particularly through chiral strain and hydration effects, remain to be elucidated.
Here, we present a comparative study of lead bromide perovskites incorporating R-, S-, and racemic α-methylbenzylammonium (MBA) cations. Our results demonstrate that while enantiopure MBA leads to the formation of highly ordered 2D perovskites, the racemic mixture promotes the formation of a 1D hydrated phase, (Rac-MBA)3PbBr5·H2O. This effect is accompanied by the emergence of broadband self-trapped exciton (STE) emission, in contrast to the narrow-band excitonic emission observed in the 2D chiral counterparts. This behavior arises from the combined effects of molecular chirality, packing frustration, exciton–phonon coupling, and structural and electronic responses to hydration. Density functional theory calculations further reveal Rashba-type spin splitting in the chiral phases and the formation of mid-gap Br-derived states in the 1D racemic structure. These states are stabilized electrostatically by symmetry-related interstitial water molecules, which mediate hydrogen bonding and enforce local dipole compensation. Structural and temperature-dependent optical studies reveal that racemic distribution and hydration play a key role not only in stabilizing the 1D framework, but also in modulating the bandgap and enabling phonon-assisted recombination. Our results establish a direct link between molecular stereochemistry, dimensional control, and electronic band structure. These findings highlight the role of stereochemical engineering in controlling both the dimensionality and optoelectronic properties of HOIPs, offering new design principles for devices based on chiral and racemic molecular components.
:
2 ratio as described in the Experimental section. Fig. 1a and b shows the scanning electron microscopy (SEM) images of the obtained samples. The R-/S-MBA based perovskites were in the form of flat needle-like crystals ranging from 50 to 150 µm with lamellar conformation. In contrast, the use of the racemic mixture yielded prolonged block-shaped crystals with axes ranging from 50 to 100 µm with three well-defined dimensions. Fig. 1c and d shows the room-temperature XRD pattern along with Rietveld refinement of the as-prepared (R-/S-MBA) sample. Both compounds crystallize as mirror images of 2D Ruddlesden–Popper perovskites with an orthorhombic crystal structure and a Sohncke P212121 space group,33,34 revealing inorganic layers of corner-sharing PbBr6 octahedra in the (0 0 2l) planes (where l is an integer) separated by bilayers of R-/S-MBA cations along the c axis as shown in Fig. 2a. Weak van der Waals interactions stabilize the organic layers. In contrast, the bonding between the organic and inorganic layers is mediated by hydrogen bonding. The powder XRD also detected a very small amount of the (R-/S-MBA)PbBr3 1D phase. The experimental PXRD patterns of (R-/S-MBA)2PbBr4, along with the simulated single-crystal XRD patterns of both the 2D phase and the (R-/S-MBA)PbBr3 phases, are shown in Fig. S2 for comparison.
Although the same synthesis procedure was adopted to synthesize 2D perovskites with the racemic mixture in the organic cation site, the Bragg reflections do not belong to the 2D orthorhombic crystal structure as observed for chiral molecules. On the other hand, single-crystal X-ray diffraction (SCXRD) of the obtained microcrystals revealed the formation of a completely different crystal phase—a monoclinic monohydrated 1D (Rac-MBA)3PbBr5·H2O perovskite belonging to the C2/c space group that was first reported by Billing and Lemmerer,21,35 as shown in Fig. 1e. This MBA-based perovskite has chains of corner-sharing PbBr6 octahedra running along the b axis in a zig–zag pattern. Organic cations surround each chain arranged radially, with their ammonium heads directed toward the chains to form hydrogen bonds with water molecules positioned between adjacent chains, as illustrated in Fig. 2b. Comparative lattice parameters of the (R-/S-MBA)2PbBr4 and (Rac-MBA)3PbBr5·H2O perovskites are displayed in Table S1, and SCXRD data together with refinement parameters are available in Table S2.
Although there are some examples of chiral halide perovskites where structures with different dimensionalities appear upon incorporating racemic cations,36,37 the underlying mechanism of such phenomena remains not fully defined. The formation of the 1D (Rac-MBA)3PbBr5·H2O structure, instead of the 2D phases observed for enantiopure (R-/S-MBA)2PbBr4 crystals, could be attributed to a combination of electrostatic repulsion and molecular distortions introduced by the racemic mixture. While enantiopure MBA cations adopt a uniform orientation that enables ordered bilayer formation through directional hydrogen bonding and efficient packing, as shown in Fig. 2a, the presence of both R- and S-enantiomers disrupts this supramolecular order. In racemic systems, the coexistence of oppositely handed molecules leads to local dipole cancellation, introducing electrostatic inhomogeneity and long-range Coulombic repulsion between adjacent organic layers. Such electrostatic frustration, coupled with local steric mismatch, can distort the inorganic layers in such a way that inhibits the formation of periodic 2D hydrogen-bonded networks, leading to the reorganization into a lower-dimensional framework of corner-sharing PbBr6 octahedra surrounded by radially reorganized MBA cations, as illustrated in Fig. 2b.38,39
To investigate the thermal stability of (Rac-MBA)3PbBr5·H2O, we have performed differential scanning calorimetry (DSC) measurements on the hydrated perovskite in the temperature range of 65–180 °C, as shown in Fig. 3a (red and blue curves correspond to the heating and cooling processes, respectively). In the first heating cycle, a weak endothermic peak is observed beginning around 85 °C, likely associated with residual solution that remained even after cleaning. In addition, another endothermic event is observed starting around 110 °C, which may be associated with the removal of water from the monohydrated 1D perovskite structure as previously reported by Zhou et al. through thermogravimetric analysis (TGA).40 A third endothermic peak appears at 150 °C during heating, followed by an exothermic event at 130 °C during cooling. In the second cycle, performed on the same sample, only the endothermic peak at 150 °C and the exothermic event at 130 °C are present during the heating and cooling processes, respectively, suggesting that an irreversible structural transformation of (Rac-MBA)3PbBr5·H2O occurred due to the release of water. To confirm this phase transformation, we compared the room-temperature XRD patterns of the as-grown perovskite and the same sample after thermal treatment at 120 °C for 1 hour. As can be seen from Fig. 3b, the XRD patterns were markedly different, confirming the irreversible phase transformation indicated by the DSC data. We compared the XRD pattern of the new phase with those of the precursors (Rac-MBA)Br and PbBr2, as well as the possible 1D and 2D phases of the MBA enantiomer-based perovskites, as shown in Fig. S3, further suggesting a new phase driven by water removal that will be determined elsewhere.
To understand the influence of molecular chirality and dimensionality on the electronic structure of these materials, we have performed DFT calculations for both the enantiopure and racemic systems. As expected from symmetry considerations, the R- and S-enantiomers yield identical band structures, with no discernible differences in electronic dispersion. The valence band maximum (VBM) primarily comprises Br p-orbitals in both cases. In contrast, the conduction band minimum (CBM) is dominated by Pb p-states, which agrees with previous reports on halide perovskites.41,42 We corroborate this by k-resolved band-character projections, consistent with type-I band-edge ordering on the inorganic Pb–Br network. The calculated bandgap is direct and located at the Γ-point, with a value of approximately 2.8 eV using the PBE functional as shown in Fig. 4a. Although this functional is known to underestimate the absolute bandgap by about 1 eV,43–45 the relative trends remain reliable. They will be discussed further in the next section. Notably, the inclusion of spin–orbit coupling (SOC) reveals a Rashba-like spin splitting at the conduction band minima, with a momentum offset and energy splitting on the order of 10 meV (Fig. 4b), as expected by the chiral-induced deformation in polar materials.46 Polarization-resolved spectroscopic measurements would be required to isolate the microscopic origin of this splitting and to disentangle SOC-driven effects from possible optical phenomena such as photon recycling or bound exciton emissions.47,48
In contrast, the racemic compound (Rac-MBA)3PbBr5·H2O exhibits a significantly reduced bandgap relative to its chiral counterpart. This narrowing originates from the mid-gap states derived from Br atoms that lack full coordination due to the disrupted octahedral connectivity in the one-dimensional structure. Although partially stabilized by hydrogen bonding with water molecules, these undercoordinated Br orbitals introduce electronic states deep within the gap region, as revealed in the unfolded band structure (Fig. 4c). Furthermore, the Fermi level lies very close to the conduction band edge, indicating that the hydrated racemic phase behaves as an extrinsic semiconductor—likely a consequence of charge compensation mechanisms facilitated by interstitial water molecules.
To assess the energetic role of the one-dimensional motif, we simulated a configuration where only a single PbBr6 chain per unit cell is retained while maintaining water molecules in the lattice. We find that the total energy per atom increases by approximately 3 eV compared to the system containing the four symmetry-related chains with alternating chirality and polarization. This large energy penalty reflects the emergence of electrostatic instability due to the absence of inter-chain dipole compensation. In this reduced symmetry case, the system spontaneously becomes metallic to screen internal fields, a phenomenon reminiscent of surface electrical stabilization in ferroelectric materials49 (Fig. 4d). Furthermore, we explored the effect of water removal by eliminating interstitial H2O molecules from the racemic hydrated structure. In all configurations examined, the absence of water leads to a drastic increase in internal energy—by nearly an order of magnitude—indicating that hydration is critical to the structural and electronic stability of the system. These results suggest that upon thermal desorption of water, the material would likely undergo substantial atomic rearrangement to restore electronic balance and minimize electrostatic energy.
Fig. 5a presents the room-temperature UV-visible absorption spectra of (R-/S-MBA)2PbBr4 and (Rac-MBA)3PbBr5·H2O crystals. The absorption edges are located at approximately 345 nm for the enantiomer-based compounds, consistent with previously reported absorption spectra of powder samples of these perovskites.50 For the hydrated 1D phase, the absorption spectrum closely resembles that reported by Dang et al.,21 featuring a small peak centered around 400 nm, which can be attributed to defect-induced mid-gap states, and a strong peak at 350 nm, commonly assigned to excitonic absorption. The direct bandgap energies were determined via the Tauc plot method to be 3.74 eV and 3.47 eV for the 2D and 1D perovskites, respectively. To confirm whether the incorporation of chiral ligands resulted in chiral perovskites, we performed circular dichroism (CD) spectroscopy on thin films obtained by dissolving the crystals in DMF and spin-coating them onto quartz substrates. This approach was necessary since the as-synthesized crystals were too thick to yield a reasonable CD response. As shown in Fig. 5b, CD signals were observed for (R-/S-MBA)2PbBr4 around 350 nm with oppositely signed values near the absorption edge for each sample, which can be attributed to the Cotton effect.34,51 This proves that chirality was successfully transferred from the chiral ligands to the inorganic sublattices.
Steady-state photoluminescence (PL) spectra of the samples are shown in Fig. 5c. For the (R-/S-MBA)2PbBr4 compounds, two distinct narrow emissions were detected at 362 nm and 382 nm, both near the absorption edge. The higher-energy PL peak is widely attributed to free exciton emission in 2D halide perovskites,52,53 whereas the origin of the lower-energy emission remains under debate. Possible explanations include photon recycling mechanisms, in which a portion of the high-energy emission becomes strongly confined within the crystal, aided by the 2D structural configuration. This confined light can propagate over long distances through total internal reflection, thereby enhancing photon recycling during emission transmission and resulting in a subsequent red-shifted emission, as previously reported for the 2D perovskites (BA)2PbBr4 and (PEA)2PbBr4.48 Another possible mechanism involves spin splitting induced by SOC. Rashba-type splitting has been previously reported as the origin of the double emission feature in the 2D perovskite (PEA)2PbI4 and in other 3D hybrid perovskites.54 As discussed earlier, the inclusion of SOC revealed a Rashba-like spin splitting at the conduction band minimum in (S-MBA)2PbBr4. However, the calculated energy splitting (∼10 meV) is significantly smaller than the experimentally observed separation between the double-emission peaks (∼180 meV), suggesting that additional mechanisms may also be involved. Nevertheless, the definitive assignment of the PL peaks in the (R-/S-MBA)2PbBr4 compounds would require further experimental analyses, such as power-dependent and time-resolved PL measurements.
An intriguing photoluminescence behavior is observed with the use of the racemic mixture. The (Rac-MBA)3PbBr5·H2O 1D perovskite exhibits a broadband emission (BE) centered around 600 nm, commonly associated with self-trapped exciton (STE) emission in halide perovskites.55,56 Similar broadband emissions are observed in the enantiomer-based 1D perovskites (R-/S-MBA)PbBr3;21 however, in this case the emission is redshifted compared to the racemic hydrated phase, which can be related to the differences in the octahedral connectivity of the PbBr6 octahedra from the corner-sharing type for (Rac-MBA)3PbBr5·H2O and the edge-sharing case for (R-/S-MBA)PbBr3. This difference in connectivity can modify the orbital overlap of Pb–Br bonds and alter the conduction/valence bandwidths and effective mass, which in turn can affect how deeply an exciton could be localized in a STE state. The BE indicates STE recombination, facilitated by lattice softness, reduced dimensionality, and hydrated interchain environments. The emergence of STE luminescence in the racemic compound underscores the profound effect of dimensionality and molecular packing on exciton dynamics. Remarkably, an intriguing photophysical behavior emerges with the use of the racemic cations, which results in the formation of a 1D perovskite structure, (Rac-MBA)3PbBr5·H2O, instead of the 2D phase typically observed with the chiral (R- or S-MBA) analogs. The origin of this broadband emission can be attributed to a confluence of structural and electronic factors.57,58 The use of a racemic mixture likely introduces local packing disorder and reduces overall lattice symmetry, which are known to promote lattice deformation and exciton self-trapping. In low-dimensional halide perovskites, such distortions are particularly impactful, as they can localize photoexcited carriers via enhanced exciton–phonon coupling mechanisms.57,58 Furthermore, the dimensional transition from 2D to 1D significantly enhances exciton–phonon interactions due to reduced dielectric screening and stronger quantum confinement. This effect is further reinforced by interstitial water molecules, which are stabilized between the inorganic chains in the 1D structure. These water molecules can interact dynamically with the organic cations via hydrogen bonding, further softening the lattice and facilitating vibrational modes that support STE formation. The reduced bandgap observed for the 1D Rac-MBA perovskite (3.47 eV) compared to its 2D counterparts (∼3.74 eV) also suggests enhanced electronic delocalization along the chain axis, which may aid in the relaxation of excitons to lower-energy self-trapped states.
Temperature-dependent photoluminescence (PL) measurements were performed to investigate the emission behavior of the (Rac-MBA)3PbBr5·H2O sample, as shown in Fig. 6a. The BE observed at room temperature blueshifts and its intensity rapidly increases upon cooling up to 110 K and reaches saturation around 80–100 K. Below 80 K, the BE redshifts and its intensity decreases followed by the emergence of a narrow-band emission (NE) centered at 410 nm with minor emissions around 350–375 nm that exhibits a relatively large Stokes shift compared to the excitonic peak at 350 nm, but it correlates closely with the absorption feature near 400 nm as shown in Fig. 5a, suggesting that it originates from bound exciton emission associated with mid-gap defect states. The BE profile resembles strong overlapping emission bands as the temperature decreases. Although multiple-peak fitting was attempted, the uncertainty in the number of contributing peaks and their irregular shifts with temperature rendered the fitting results unreliable. As we have not observed well-resolved bands in the range of the NE emission, we have performed the analysis by considering the entire spectral range of the BE (450–700 nm) and NE (350–450 nm) regions to evaluate the following results.
![]() | ||
| Fig. 6 (a) PL temperature dependence of (Rac-MBA)3PbBr5·H2O perovskites. (b) Normalized integrated PL intensity and (c) the FWHM of BE dependence on temperature. | ||
The energy barrier that activates nonradiative processes, such as exciton trapping or nonradiative recombination, is referred to as the activation energy Ea and can be estimated by using the Arrhenius equation59,60 applied to the integrated PL intensity IPL(T) dependence on temperature T according to eqn (1):
![]() | (1) |
As before, the BE is attributed to STEs, which are closely related to the interactions between charge carriers and lattice vibrations. To assess the strength of the exciton–phonon coupling, we analyzed the temperature dependence of the BE linewidth Γ(T), as shown in Fig. 6c. This dependence can be modeled using Toyozawa's theory,62 which employs a configuration coordinate framework where the average number of phonons with effective energy Eph emitted during exciton recombination is described using the Huang–Rhys factor S,63 in the form of eqn (2):
![]() | (2) |
In order to determine whether the changes observed in the photoluminescence were related to structural origins, temperature-dependent XRD measurements of (Rac-MBA)3PbBr5·H2O powder were conducted, as shown in Fig. S6. Rietveld refinement of the diffractograms revealed a continuous decrease of the lattice parameters and unit cell volume from 380 K to 80 K, while the monoclinic crystal structure remained unchanged, as illustrated in Fig. 7a and b (see some selected temperatures in Fig. S7). These results exclude the possibility of structural phase transitions at low temperatures that could modify the electronic structure of the 1D perovskites and lead to different emission behavior.
To gain insights into the temperature behavior of the inorganic framework, the octahedral distortions were further analyzed using the distortion index
, where di is the ith Pb–Br bond length in an octahedron and d is the average bond length, and the bond angle variance
, where θi represents the ith Br–Pb–Br bond angle in an octahedron. As shown in Fig. 7c, the octahedral distortions slowly change above 120 K, which indicates the persistence of the octahedral configuration. However, the opposite trend is observed in the range of 120 K to 80 K, where octahedral distortions decrease with temperature. Octahedral distortions are known to contribute to the formation of self-trapped excitons (STEs), where more distorted environments are prone to exciton trapping.55,66,67 However, lowering the temperature or applying pressure usually promotes exciton self-trapping by increasing octahedral distortions.68,69 In contrast, we report an unusual case for (Rac-MBA)3PbBr5·H2O, where lower temperature reduces octahedral distortions, and we propose that these distortions play a central role in governing the photoluminescence behavior of the hydrated 1D phase at low temperatures. Reduced octahedral distortions create a less favorable environment for exciton trapping, enabling the appearance of a high-energy narrow emission alongside the STE band, as not all excitons can be trapped within the more regular octahedral framework. Similar behavior has previously been observed in layered bromide perovskites under pressure.70
CCDC 2466728 contains the supplementary crystallographic data for this paper.71
| This journal is © The Royal Society of Chemistry 2025 |