Yaiza
Asensio
ab,
Lucía
Olano-Vegas
ab,
Samuele
Mattioni
ab,
Marco
Gobbi
cd,
Fèlix
Casanova
ad,
Luis E.
Hueso
*ad and
Beatriz
Martín-García
*ad
aCIC nanoGUNE BRTA, 20018 Donostia-San Sebastián, Basque Country, Spain. E-mail: l.hueso@nanogune.eu; b.martingarcia@nanogune.eu
bDepartamento de Polímeros y Materiales Avanzados: Física, Química y Tecnología, University of the Basque Country (UPV/EHU), Donostia-San Sebastian, 20018, Spain
cMaterials Physics Center CSIC-UPV/EHU, 20018 Donostia-San Sebastián, Spain
dIKERBASQUE, Basque Foundation for Science, 48009 Bilbao, Spain
First published on 25th February 2025
The chemical and structural flexibility of hybrid organic–inorganic metal halide perovskites (HOIPs) provides an ideal platform for engineering not only their well-studied optical properties, but also their magnetic ones. In this review we present HOIPs from a new perspective, turning the attention to their magnetic properties and their potential as a new class of on-demand low-dimensional magnetic materials. Focusing on HOIPs containing transition metals, we comprehensively present the progress that has been made in preparing, understanding and exploring magnetic HOIPs. First, we briefly introduce HOIPs in terms of composition and crystal structure and examine the synthesis protocols commonly used to prepare those showing magnetic properties. Then, we present their rich magnetic behavior and phenomenology; discuss their origin and guidelines for tuning them by changing the perovskite phase, chemical composition and dimensionality; and showcase their potential application in magneto-optoelectronics and spintronics. Finally, we describe the current challenges in the field, such as their integration into devices, as well as the emerging possibilities of moving from magnetic doping to pure transition metal-based HOIPs, which will motivate further studies in the future.
Wider impactThis review presents our current understanding of the magnetic properties of hybrid organic–inorganic transition metal halide perovskites, as well as their modulation using crystal structure and composition as tuning knobs. Furthermore, their potential applications in a variety of fields such as magneto-optics, sensors and spin filters are described. Both these aspects are crucial to make these materials key for the future development of magnetically controlled optical and electronic devices and spin-based multifunctional technologies. |
Metal halide perovskites are compounds that have the general formula ABX3 for the three-dimensional (3D) arrangement of [BX6]4− octahedra, with the peculiarity of an A-site entirely or partially occupied by small organic molecules leading to HOIPs, the focus of this review, although inorganic cations such as Cs+ or Rb+ can also be present. The B-site is filled by metals and X-site by halogens.10 These structures impose strict size constraints on their elements to fit within the perovskite framework. However, the dimensionality, defined as the number of spatial directions in which the octahedra formed by [BX6]4− are connected, can vary from the 3D case to the two-dimensional (2D) to one-dimensional (1D) to the zero-dimensional (0D) ones.10 As the dimensionality decreases, the A-site's size restrictions are progressively reduced, leading to a wide range of possible combinations. This broad structural and chemical flexibility of HOIPs offers substantial opportunities for fine-tuning their physical properties through simple chemical modifications and to obtain materials with several functionalities.4,11 Among the properties that can be controlled, the electrical, optical and magnetic ones stand out as the most promising, mainly due to their potential for technological advancements. In particular, the magnetic behavior of HOIPs is strongly influenced by their dimensionality, phase and composition, and even minor modifications can lead to changes in their exchange mechanism and anisotropy, giving rise to different properties.12–14 In addition to their structural and compositional diversity, HOIPs offer another key advantage over traditional magnetic semiconductors or metals; their ease of synthesis based on solution processes that are low cost, versatile and do not require very high temperatures or pressures.15–19
Research on the magnetic properties of HOIPs dates back to the 1960s when de Jongh et al. studied numerous materials and established the relationship between their structure and magnetic properties.20–24 Since then, a wide array of magnetic HOIPs5,20–62 have been reported and, recent works have shown renewed interest in this field, highlighting their rich magnetic phenomena and their potential for exploring the interplay between magnetism and optoelectronic properties.63 This opens exciting possibilities for developing multifunctional materials that can be employed in spintronics, magneto-optics, magnetic sensing or data storage devices, among others.63–66
Despite the growing interest in this area, there has been no comprehensive review dedicated to understanding and summarizing the developments related to magnetism in HOIPs. This review aims to fill this gap by providing a detailed examination of the origin of magnetic properties in these materials and how they can be tuned through dimensionality and structural modifications, the methodologies employed to synthesize magnetic HOIPs, and the strategies to exploit these materials for practical applications. We will also discuss the challenges and opportunities in this emerging field, offering insights into potential future directions and the broader impact of magnetic HOIPs in advanced technologies.
Regarding the remaining dimensionalities, restrictions progressively relax until they are not applicable altogether. In 1D HOIPs, the metal halide octahedra are connected in a linear fashion, forming a one-dimensional nanowire. These nanowires are surrounded by organic cations which can adopt either a straight or zigzag configuration, depending on whether the connection between the inorganic octahedra is face-sharing or edge-sharing, respectively. Their general chemical formula is A2BX5. Finally, in 0D HOIPs, the metal halide octahedra are entirely isolated by the surrounding organic cations, resulting in discrete, non-connected entities distributed throughout the crystal lattice without direct connection to one another. The general chemical formula for these materials is A4BX6.83,84
In general, the metal cation B source and the molecule must be dissolved into an appropriate solvent to obtain a solution from which the single crystal can grow. Commonly, metal salts, such as metal halides (e.g. CuCl2, MnCl2, FeCl2, etc.), are used as precursors of the metal cation, while in the case of the organic cations, the molecule or the corresponding organic halide salts can be added directly. Several potential solvents can be used, either hydrohalic acids or organic solvents. In the case of the hydrohalic acids (HX = HI, HBr and HCl), they can act as a solvent as well as a halide source and thus must match the type of halide in the desired perovskite. On the other hand, organic solvents, such as ethanol/methanol, N,N-dimethylformamide (DMF), dimethyl sulfoxide (DMSO), or mixtures of them, can be used. However, these solvents do not provide excess halide like HX; thus, salts with the same halide as the desired HOIPs or an external halide source must be added.87,88
The crystallization from solution can be thought of as a two-step process. The first step is the phase separation, or ‘birth’ of the new crystal, and it is called ‘nucleation’. A supersaturated solution is required for crystallization to occur. Since a supersaturated solution is not at equilibrium, the formation of nuclei from the solution can relieve the supersaturation and move it towards equilibrium. Once crystallization starts, however, supersaturation can be relieved by a combination of nucleation and crystal growth. In the second stage, the nucleated nuclei grow larger by the addition of solute molecules from the supersaturated solution.88,89
In Fig. 2, sketches of the main solution-based techniques to synthesize magnetic HOIPs are shown.
![]() | ||
Fig. 2 Schematics of the main solution-based process for magnetic hybrid organic–inorganic perovskite synthesis: (a) solution temperature lowering, (b) slow evaporation and (c) solvothermal methods. |
Other less common approaches that one can find in the literature for the synthesis of magnetic HOIPs are the antisolvent vapor diffusion crystallization (AVC),105,106 where a ‘poor solvent’ diffuses into a ‘good solvent’, changing the solubility of the compound and inducing crystallization; and the Solvent Acidolysis Crystallization (SAC) generating the organic cation in the growth medium.107 For specific systems, e.g. doped Pb perovskites, supersaturation can also be reached by an increase in the temperature (ITC – inverse temperature crystallization).108,109 Additionally, a few reports can be found on solid-state synthesis of magnetic HOIPs, where the precursors are ground to form a mixture that is then heated in vacuum overnight and later cooled down at a fixed rate.105,110
According to the Goodenough–Kanamori rules,111–113 antiferromagnetic and ferromagnetic orders can be predicted from the geometric arrangement of [BX6]4− octahedra. In pure Cu2+ or Cr2+-based HOIPs, their octahedral frame displays a Jahn–Teller distortion due to a reduction of orbital degeneracies and crystal field effects, leading to alternatively compressed and elongated octahedral cages as shown in Fig. 3a. This configuration results in a ferromagnetic B2+–X−–B2+ superexchange pathway through different orthogonal d orbitals. Conversely, in non-active Jahn–Teller HOIPs, such as those containing Mn2+ or Fe2+, their linear and symmetrical B–X–B will lead to antiferromagnetic interactions following a B2+–X−–B2+ pathway (Fig. 3b).12
![]() | ||
Fig. 3 Schematic diagrams showing the presence or the absence of superexchange pathways in HOIPs. (a) and (b) Illustrations exhibiting short-range superexchange pathways for active and non-active Jahn–Teller compounds, respectively. In the case of double HOIPs, long range superexchange interactions also take place as shown in (c) for (B1+–B23+)-based HOIPs and in (d) for (B12+–B22+)-based HOIPs. (e) Showing the absence of superexchange interactions in non-magnetic HOIPs doped with magnetic elements. Adapted with permission from ref. 7. Copyright © 2022, American Chemical Society. |
Double HOIPs involving two metal cations (B1+–B23+ or B12+–B22+)6,7,13 also exhibit these superexchange interactions. However, they are now typically long-range due to an increased distance between the magnetic sites, leading to weaker magnetic exchange compared to HOIPs with a single metal cation. For double HOIPs containing a diamagnetic ion B1+ combined with a transition metal ion B23+, their superexchange interactions occur via B23+–X−–B1+–X−–B23+ pathways (Fig. 3c).13 Additionally, when two transition metals, B12+–B22+, are combined, these long-range superexchange interactions could proceed along either B12+–X−–B22+–X−–B12+ or B22+–X−–B12+–X−–B22+ pathways (Fig. 3d). Short-range interactions via B12+–X−–B22+ pathways could also be significant, especially if there is a substantial orbital overlap between the two different metal cations, although this situation is uncommon as previously observed in double perovskites combining a transition metal from the 3d series with another from the 4d/5d one.114,115 Furthermore, the introduction of magnetic elements as dopants into HOIPs can induce magnetism, (Fig. 3e). This has been explored in Pb-based HOIPs acting as a host with Mn2+ doping. Nevertheless, the small doping concentrations usually achieved cannot induce long-range magnetic ordering in the host comparable to that observed in the previous cases.63,64,116
Finally, the dimensionality of HOIPs also plays a significant role regarding their magnetic properties, as its reduction implies a decrease in the number of nearest neighbors for each magnetic ion.24 This can result in a weakening of the magnetic interactions, an enhancement of the magnetic anisotropy (which makes the material more sensitive to the direction of the applied field) or even in frustrated magnetism due to competing interactions. The combination of all these effects can lead to complex magnetic behaviors. Therefore, understanding both these factors and their interplay is crucial for tailoring magnetic properties of HOIPs.
In the following sections, we will discuss in detail how magnetism varies among different HOIPs according to their dimensionality and the type of metal cations they contain, as well as how these properties can be modulated by replacing individual components within the HOIP structure.
• Perovskite phase: each phase (RP or DJ) presents different organic spacer arrangements and octahedral alignments between layers, impacting the magnetic exchange.
• Organic cations: these molecules hydrogen bond with the inorganic framework in a particular way, leading to a determined orientation and conformation of the whole structure, and to a particular interlayer spacing, which in turn affects the magnetic interactions.
• Metal: as explained in Section 4.1, the transition metal not only introduces magnetism but also influences the rigidity of the inorganic framework, impacting the previous hydrogen bonding schemes.
• Halogen: this element acts as the bridge for superexchange interactions, and any change of it will affect the overall magnetism.
Therefore, understanding these factors and their interplay is crucial for controlling the magnetic properties of 2D HOIPs. Below, we explain more in detail the role of each component in determining this behavior.
Additionally, in these layered systems, the interlayer distance plays a crucial role because a larger separation weakens the exchange interactions between adjacent layers. However, the position of the metal–halide octahedra is also relevant as it can lead to different interlayer superexchange pathways. In particular, due to their in-line stacking, DJ HOIPs present a nearly linear two-halide B2+–X−⋯X−B2+ superexchange pathway between the layers, which facilitates stronger magnetic coupling. In contrast, compounds with the RP phase typically have a non-linear pathway with larger metal–metal distances for similar interlayer spacing. As a result, HOIPs with the same interlayer distances but different phases can exhibit different magnetic phenomena.
One example of this concept appears in one of our works.5 Among the series of HOIPs studied, we analyze two compounds, (C2H3NH3)2CuCl4 (ethylammonium – EA – C2H3NH3+) (RP) and (NH3C6H4NH3)CuCl4 (phenylenediammonium – PEAA – (NH3C6H4NH3)2+) (DJ), with similar interlayer distances (10.32 Å and 9.68 Å, respectively) but distinct perovskite phases (Fig. 4a and b). This difference leads to variations in the inorganic octahedra distortions, evidenced by the angles formed by the Cu–Cl–Cu atoms (∼170° for EA2CuCl4 and ∼160° for PEAACuCl4), which translates to a higher magnetic anisotropy in the DJ phase. These structural variations have a significant impact on the magnetic behavior, that can be detected from their magnetization (M) versus temperature (T) curves for both in-plane and out-of-plane magnetic fields (Fig. 4c–f). EA2CuCl4 exhibits a Curie temperature (TC) of 10 K with the presence of a small kink for the two configurations of the field. The analysis of the exchange interaction constants reveals ferromagnetic intralayer interactions (J ∼ 10 K) and much weaker antiferromagnetic interlayer interactions (J′ ∼ −10−3 K), to which the emergence of the kink could be ascribed. In contrast, PEAACuCl4 displays a much weaker temperature dependence of magnetization (one order of magnitude lower) with a TC around 13 K, higher than the RP-phase crystal. Additionally, the peak in the M vs. T curve is much broader. While the intralayer exchange interaction remains similar to the previous compound, the interlayer interaction is significantly stronger, around −10−2 K. This suggests that the antiferromagnetic interlayer interactions play a more dominant role in the DJ phase, influencing its overall magnetic behavior. Furthermore, from the magnetization dependance with the magnetic field (H), the DJ compound appears to achieve a saturated ferromagnetic state at higher applied fields compared to the RP phase. In summary, this example demonstrates how the structural variations between RP and DJ phases in 2D HOIPs significantly impact their magnetic behavior.
![]() | ||
Fig. 4 Crystal structure of (a) EA2CuCl4 and (b) PEAACuCl4 drawn with VESTA software. Magnetization (M) versus temperature (T) at 500 Oe parallel (‖, red lines) and perpendicular (⊥, blue lines) to the [CuCl6]4− octahedra layers for (c) EA2CuCl4 and (e) PEAACuCl4. Magnetization (M) versus applied field (H) at 5 K parallel (‖, red lines) and perpendicular (⊥, blue lines) to the [CuCl6]4− octahedra layers for (d) EA2CuCl4 and (f) PEAACuCl4. Ref. 5 © 2022 Wiley-VCH GmbH. |
• Organic chain length. The length of the organic molecule directly impacts the spacing between the inorganic lattice sheets. In turn, this influences the effective dimensionality of the material with respect to its magnetic properties.
• Hydrogen bonding. The specific hydrogen bonding between the organic cations and halides influences the structural arrangement of all components within 2D HOIPs, which are essential in determining the overall magnetic behavior of the material.
![]() | ||
Fig. 5 Crystal structure of (a) (NH3(CH2)xNH3)CuCl4 with organic molecules ranging from (b) x = 2 to x = 9 (ref. 117, copyright © 2017, Springer-Verlag GmbH Germany) and (c) absolute value of the ratio between their exchange constants |J′/J| versus x. (d) Crystal structures of PA2MnCl4, PMA2MnCl4, PEA2MnCl4 and PPA2MnCl4 at 100 K and their field-dependent magnetization measured at 10 K in high and low magnetic fields along the out-of-plane and in-plane directions. Ref. 48, copyright © 2021 The Authors. |
Regarding their intralayer interactions, all the HOIPs within this series exhibit ferromagnetic behavior due to the Jahn–Teller distortion of the [CuCl6]4− octahedra. The intralayer exchange constants J maintain similar values across the series, ranging from 13 to 23 K, indicating a minimal impact of the organic chain length on the intralayer magnetic coupling. In contrast, their interlayer magnetic coupling is antiferromagnetic and presents significant variations across the series. In particular, the interlayer exchange constant J′ passes from values of the order of 10−3 for the highest interlayer distance (x = 9), to values of the order of 10 for x = 2. The ratio between the exchange constants for all the range of studied HOIPs appears in Fig. 5c. It shows how, for larger interlayer distances (higher x), the intralayer interactions are much higher than the interlayer ones, presenting 2D magnetic ordering. However, the interlayer interactions become important as the interlayer distance is reduced (x decreases), achieving the behavior of a 3D magnet for x = 2.
This behavior is applicable for the same series of organics but incorporating Br as a halide,29,32,33 and also in compounds presenting a RP phase. For instance, the (CxH2x+1NH3)2CuX4 (X = Cl, Br) series5,20,22–24 exhibit a similar reduction in the ratio J′/J while decreasing the interlayer distance (see Table 120,22–34), suggesting a general trend in 2D HOIPs. However, the J′ are weaker due to the staggered arrangement of the inorganic layers. Additionally, in the case of (C6H5(CH2)xNH3)2CuX4 series with X = Cl, Br5,49,100,118,119 only the intralayer exchange constants are reported, showing again very low variations when changing the x. This is also revealed by Cr and Mn-based HOIPs (Table 235–45 and Table 3,5,21,46–55 respectively), in particular in compounds mixed with Cl and CxH2x+1NH3+5,21,35,41,46,47,51,52 and C6H5(CH2)xNH3+
5,35,38,48–50 monoammonium molecules, confirming that organics do not have a notable influence on the intralayer interactions of 2D HOIPs.
HOIP | Organic family | n | Halogen | Perovskite phase | T C/TN (K) | J/kB (K) | J′/kB (K) | |J′/J| | Interlayer distance (Å) | Ref. |
---|---|---|---|---|---|---|---|---|---|---|
(CH3NH3)2CuCl4 | CnH2n+1NH3 | 1 | Cl | RP | 8.91 | 18.20 | 0.0006 | 9.99 | 22–24 | |
(C2H5NH3)2CuCl4 | 2 | Cl | 10.20 | 17.20 | −0.003 | 0.0002 | 10.32 | 5, 20 and 22–24 | ||
(C3H7NH3)2CuCl4 | 3 | Cl | 7.61 | 16.00 | 0.0001 | 12.89 | 22–24 | |||
(C4H9NH3)2CuCl4 | 4 | Cl | 7.27 | 15.40 | 0.0001 | 15.83 | 23 and 24 | |||
(C5H11NH3)2CuCl4 | 5 | Cl | 7.26 | 15.90 | 0.0001 | 17.80 | 23 and 24 | |||
(C6H13NH3)2CuCl4 | 6 | Cl | 7.65 | 17.10 | 0.0001 | 23 and 24 | ||||
(C10H21NH3)2CuCl5 | 7 | Cl | 7.91 | 17.90 | <0.00001 | 25.78 | 22 and 24 | |||
(CH3NH3)2CuBr4 | 1 | Br | 15.80 | 10.31 | 24 | |||||
(C2H5NH3)2CuBr4 | 2 | Br | 10.85 | 19.00 | 0.0020 | 11.44 | 23 and 24 | |||
(C3H7NH3)2CuBr4 | 3 | Br | 10.50 | 21.30 | 0.0020 | 12.78 | 23 and 24 | |||
(C4H9NH3)2CuBr4 | 4 | Br | 11.33 | 21.90 | 0.0010 | 14.82 | 24 | |||
(C5H11NH3)2CuBr4 | 5 | Br | 11.40 | 22.00 | 0.0001 | 24 | ||||
(C6H5CH2NH3)2CuCl4 | C6H5(CH2)nNH3 | 1 | Cl | RP | 12.00 | 18.20 | 0.001 | <0.0001 | 18.87 | 5, 49, 100 and 119 |
(C6H5(CH2)2NH3)2CuCl4 | 2 | Cl | 9.00 | 18.80 | 19.69 | 100 and 119 | ||||
(C6H5(CH2)3NH3)2CuCl4 | 3 | Cl | 7.00 | 16.70 | 20.44 | 100 and 119 | ||||
(C6H5CH2NH3)2CuBr4 | 1 | Br | 12.81 | 25.00 | <1.000 | <0.0400 | 118 | |||
(C6H5(CH2)2NH3)2CuBr4 | 2 | Br | 12.85 | 22.70 | <0.500 | <0.0200 | 118 | |||
(C6H5(CH2)3NH3)2CuBr4 | 3 | Br | 10.02 | 21.70 | <0.300 | <0.0100 | 118 | |||
(NH3C2H4COOH)2CuCl4 | Others | Cl | RP | 13.80 | 25 | |||||
(NH4)2CuCl4 | Cl | 11.20 | 17.00 | 0.0032 | 8.91 | 24 | ||||
((CH3)2CHCH2NH3)2CuCl4 | Cl | 6.50 | 14.58 | 120 | ||||||
R/S-(C6H5CHCH3CH2NH2)2CuCl4 | Cl | 5.00 | 121 | |||||||
((CH3)2CHCH2NH3)2CuCl4 | Br | 12.20 | 21.40 | −0.050 | 120 | |||||
(NH3(CH2)2NH3)CuCl4 | NH3(CH2)nNH3 | 2 | Cl | DJ | 31.50 | 23.00 | −13.700 | 0.5950 | 8.11 | 26 |
(NH3(CH2)3NH3)CuCl4 | 3 | Cl | 14.90 | 16.50 | −1.700 | 0.1000 | 9.12 | 27 and 30 | ||
(NH3(CH2)4NH3)CuCl4 | 4 | Cl | 8.90 | 13.00 | −0.160 | 0.0100 | 9.09 | 31 | ||
(NH3(CH2)5NH3)CuCl4 | 5 | Cl | 7.60 | 14.10 | −0.040 | 0.0030 | 11.94 | 30 | ||
(NH3(CH2)6NH3)CuCl4 | 6 | Cl | 9.30 | 15.30 | −0.020 | 0.0010 | 29 | |||
(NH3(CH2)7NH3)CuCl4 | 7 | Cl | 8.30 | 15.20 | −0.010 | 0.0007 | 29 | |||
(NH3(CH2)8NH3)CuCl4 | 8 | Cl | 8.20 | 15.80 | −0.005 | 0.0003 | 29 | |||
(NH3(CH2)9NH3)CuCl4 | 9 | Cl | 6.00 | 16.10 | −0.003 | 0.0002 | 29 | |||
(NH3(CH2)10NH3)CuCl4 | 10 | Cl | 7.00 | 15.60 | <1.000 | <0.0200 | 29 | |||
(NH3(CH2)2NH3)CuBr4 | 2 | Br | 72.00 | 38.20 | −68.400 | 1.8000 | 29 and 33 | |||
(NH3(CH2)3NH3)CuBr4 | 3 | Br | 42.00 | 26.00 | −26.000 | 1.0000 | 8.60 | 32 and 33 | ||
(NH3(CH2)4NH3)CuBr4 | 4 | Br | 19.00 | 29.00 | −5.000 | 0.1700 | 8.92 | 32 and 33 | ||
(NH3(CH2)5NH3)CuBr4 | 5 | Br | 12.20 | 23.10 | −2.000 | 0.0900 | 29 | |||
(NH3(CH2)6NH3)CuBr4 | 6 | Br | 13.00 | 20.30 | −0.100 | 0.0050 | 29 | |||
(NH3(CH2)7NH3)CuBr4 | 7 | Br | 10.20 | 23.00 | <1.000 | <0.0500 | 29 | |||
(NH3(CH2)8NH3)CuBr4 | 8 | Br | 12.60 | 22.10 | −0.050 | 0.0020 | 29 | |||
(NH3(CH2)9NH3)CuBr4 | 9 | Br | 10.00 | 21.20 | <1.000 | <0.0500 | 29 | |||
(NH3(CH2)10NH3)CuBr4 | 10 | Br | 10.00 | 16.80 | <1.000 | <0.0500 | 29 | |||
(NH3(CH2)2NH3)CuCl2Br2 | 2 | Cl/Br | 45.00 | 15.00 | −31.000 | 8.30 | 34 | |||
((NH3CH2CH2)NH2)CuCl4Cl | Others | Cl | DJ | 11.80 | 18.70 | 0.0028 | 11.80 | 28 | ||
(C6N2H10)CuCl4 | Cl | 5.70 | 8.13 | −1.220 | 0.1500 | 122 | ||||
(NH3C6H4NH3)CuCl4 | Cl | 13.00 | 12.00 | −0.025 | 0.0020 | 9.68 | 5 | |||
(NH3C2H2NH3)CuCl4 | Cl | 36.00 | 1.30 | −1.000 | 0.7900 | 7.70 | 5 | |||
(C6N2H10)CuBr4 | Br | 18.90 | 21.30 | −5.870 | 0.2800 | 122 |
HOIP | Organic family | n | Halogen | Perovskite phase | T C/TN (K) | J/kB (K) | Interlayer distance (Å) | Ref. |
---|---|---|---|---|---|---|---|---|
(CH3NH3)2CrCl4 | CnH2n+1NH3 | 1 | Cl | RP | 42.00 | 13.00 | 9.44 | 35 and 41 |
(C2H5NH3)2CrCl4 | 2 | Cl | 41.00 | 10.10 | 10.71 | 35 | ||
(C3H7NH3)2CrCl4 | 3 | Cl | 39.50 | 9.30 | 12.35 | 35 | ||
(C5H11NH3)2CrCl4 | 5 | Cl | 9.30 | 17.81 | 35 | |||
(CH3NH3)2CrBr4 | 1 | Br | 11.60 | 44 | ||||
(C2H5NH3)2CrBr4 | 2 | Br | 10.00 | 44 | ||||
(C3H7NH3)2CrBr4 | 3 | Br | 10.40 | 44 | ||||
(C5H11NH3)2CrBr4 | 5 | Br | 10.70 | 44 | ||||
(C8H17NH3)2CrBr4 | 8 | Br | 10.30 | 44 | ||||
(C12H25NH3)2CrBr4 | 12 | Br | 9.80 | 44 | ||||
(C6H5CH2NH3)2CrCl4 | C6H5(CH2)nNH3 | 1 | Cl | RP | 37.00 | 10.60 | 15.71 | 35 and 38 |
(C6H5CH2NH3)2CrBr4 | 1 | Br | 52.00 | 14.70 | 16.24 | 35, 37 and 45 | ||
(C6H5CH2NH3)2CrBr3.3Cl0.7 | 1 | Br/Cl | 49.00 | 12.50 | 16.10 | 35 and 36 | ||
(C6H5CH2NH3)2CrBr3Cl | 1 | Br/Cl | 13.10 | 45 | ||||
(C6H5CH2NH3)2CrBr1.8Cl2.2 | 1 | Br/Cl | 12.70 | 45 | ||||
(C6H5CH2NH3)2CrBr0.7Cl3.3 | 1 | Br/Cl | 12.30 | 45 | ||||
(C6H5CH2NH3)2CrBr3.3I0.7 | 1 | Br/I | 51.00 | 11.00 | 16.15 | 35 and 39 | ||
(H3N(CH2)3NH3)2CrCl4 | Others | Cl | RP | 10.60 | 42 | |||
(CH3(CH2)11NH3)2CrCl4 | Cl | 9.60 | 31.00 | 40 | ||||
(NH4)2CrBr4 | Br | −5.90 | 44 | |||||
(CH3(CH2)4NH3)CrCl4 | Others | Cl | DJ | 9.30 | 17.81 | 43 |
HOIP | Organic family | n | Halogen | Perovskite phase | T C/TN (K) | J/kB (K) | Interlayer distance (Å) | Spin canting | Canting degree (°) | Metamagnetism | SP transition | Ref. |
---|---|---|---|---|---|---|---|---|---|---|---|---|
(CH3NH3)2MnCl4 | CnH2n+1NH3 | 1 | Cl | RP | 45.30 | −5.00 | No | Yes | 21, 46 and 51 | |||
(C2H5NH3)2MnCl4 | 2 | Cl | 43.10 | −4.60 | 10.77 | Yes | 0.030 | Yes | Yes | 5, 47 and 51 | ||
(C3H7NH3)2MnCl4 | 3 | Cl | 39.20 | −4.45 | Yes | 0.050 | No | Yes | 47 and 51 | |||
(C5H11NH3)2MnCl4 | 5 | Cl | 37.00 | ∼4.00 | 52 | |||||||
(C7H15NH3)2MnCl4 | 7 | Cl | 39.00 | ∼4.00 | 52 | |||||||
(C9H19NH3)2MnCl4 | 9 | Cl | 40.00 | ∼4.00 | 52 | |||||||
(C3H7NH3)2MnBr4 | 1 | Br | 47.00 | −4.50 | No | No | 47 | |||||
(C6H5NH3)2MnCl4 | C6H5(CH2)nNH3 | 0 | Cl | RP | 42.60 | −6.19 | Yes | 0.036 | 48 | |||
(C6H5CH2NH3)2MnCl4 | 1 | Cl | 44.60 | −7.47 | No | 48 | ||||||
(C6H5C2H4NH3)2MnCl4 | 2 | Cl | 45.70 | −7.65 | 19.46 | Yes | 0.060 | Yes | 5 and 48–50 | |||
(C6H5C3H6NH3)2MnCl4 | 3 | Cl | 41.80 | −7.23 | Yes | 0.029 | 48 | |||||
R/S-(C6H5CHCH3CH2NH2)2MnCl4 | Others | Br | 3.20 | −3.64 | 17.65 | Yes | 55 | |||||
(NH3(CH2)2NH3)MnCl4 | NH3(CH2)nNH3 | 2 | Cl | DJ | 8.61 | 54 | ||||||
(NH3(CH2)3NH3)MnCl4 | 3 | Cl | 9.50 | 53 and 54 | ||||||||
(NH3(CH2)4NH3)MnCl4 | 4 | Cl | 8.85 | 13.00 | 10.77 | 54 | ||||||
(NH3(CH2)5NH3)MnCl4 | 5 | Cl | 12.00 | 54 | ||||||||
(NH3C2H2NH3)MnCl4 | Others | Cl | DJ | 80.00 | −10.20 | 8.30 | Yes | 5 |
These results suggest that the organic chain length mainly influences the interlayer magnetic coupling in 2D HOIPs, whereas it does not have a notable impact on the intralayer interactions.
For instance, (C6H5(CH2)xNH3)2MnCl4 series with arylamines being x = 0–3 (x = 0 – phenylammonium – PA, x = 1 – phenylmethylammonium – PMA, x = 2 PEA – phenylethylammonium, x = 3 PPA – phenylpropylammonium) (Fig. 5d) reported by Septiany et al.,48 exhibit different magnetic properties depending on the organic cation. Their M vs. out-of-plane H measurements (Fig. 5e) reveal a SP transition only for the PEA-based HOIP, characterized by a change in slope around 25 kOe. Additionally, hysteresis loops observed at low fields in both this characterization and in M versus in-plane H measurements suggest spin canting in compounds with PA, PEA and PPA organic cations. Spin canting is a magnetic phenomenon in which spins are tilted a few degrees angle from their axis, known as the canting angle. This phenomenon arises from the antisymmetric Dzyaloshinsky–Moriya (DM) interactions, caused by the tilting of the [MnCl6]4− octahedra due to H-bonding between the inorganic layer and the organic part. Spin canting prevents the perfect antiparallel alignment of the spins on neighboring metal ions within an AFM layer, generating residual spins along the canted direction and inducing weak FM.123 Their calculated canting angles are 0.036, 0.060 and 0.029 degrees, respectively, showing no correlation between the length of the organic cation.
The series of HOIPs (CxH2x+1NH3)2MnCl4 with aliphatic chains being x = 1–35,21,46,47,51 also demonstrates the influence of the hydrogen bonding on the magnetic behavior. In this case, all the compounds exhibit SF transitions. This phenomenon appears in AFM systems with weak anisotropy when a magnetic field is applied parallel to the preferred axis of sublattice magnetization. Under these conditions, a competition arises between the strength of the external field and the internal exchange field. In such systems, when the field reaches a critical value, the antiparallel magnetizations (spins) of the two sublattices change their orientation from the easy axis to a direction perpendicular to it, while maintaining their antiparallel alignment. This transition is referred to as SF.123 In addition, x = 1 and x = 2 perovskites have spin canting, with degrees in the same range as the previous ones: 0.030 and 0.050 degrees, respectively. Notably, x = 2 also presents a metamagnetic behavior, where one of the spin sublattices rotates to align parallel to the other when a strong enough magnetic field is applied along the preferred axis, overcoming the internal exchange interactions and inducing FM.123
However, these emergent magnetic properties can vary when the metal cation is modified. For example, when Fe2+ replaces Mn2+ in the previous HOIP family with aliphatic amines, CxH2x+1NH3+,56,57,59,61 spin canting is again reported for x = 1, 2 but with much larger canting angles: 1.4 and 0.63 degrees, respectively (Table 456–62). In contrast, neither of them exhibits now a SF transition and the metamagnetism is now present in x = 1 instead of in x = 2 as in the Mn-based compound. Another Fe-based HOIP that maintains its spin canting is the one containing PEA molecules,62 which is again stronger than its Mn counterpart (canting angle of 0.53 degrees). Interestingly, this perovskite's ferroelastic nature can change its structure in response to temperature or mechanical stress, what in turn can affect the magnetic properties. In particular, uniaxial stress in this material can switch the spin canting direction. The pristine crystal exhibits a magnetic hysteresis loop along the a-axis, but this vanishes when applying uniaxial mechanical stress along the b-axis due to axis swapping, revealing a coupling between the antisymmetric Dzyaloshinskii–Moriya interaction and ferroelasticity in PEA2FeCl4.
HOIP | Organic family | n | Halogen | Perovskite phase | T C/TN (K) | J/kB (K) | Interlayer distance (Å) | Spin canting | Canting degree (°) | Metamagnetism | SP transition | Ref. |
---|---|---|---|---|---|---|---|---|---|---|---|---|
(CH3NH3)2FeCl4 | CnH2n+1NH3 | 1 | Cl | RP | 95.00 | 9.50 | Hidden | 1.4 | Yes | No | 56, 57, 59 and 61 | |
(C2H5NH3)2FeCl4 | 2 | Cl | 110.00 | 10.50 | Yes | 0.63 | No | No | 57 and 61 | |||
(C3H7NH3)2FeCl4 | 3 | Cl | 90.00 | 58 and 61 | ||||||||
(C4H9NH3)2FeCl4 | 4 | Cl | 90.00 | Yes | 58 | |||||||
(C6H5CH2NH3)2FeCl4 | C6H5(CH2)nNH3 | 1 | Cl | RP | 73.00 | 78 | 56 and 61 | |||||
(C6H5C2H4NH3)2FeCl4 | 2 | Cl | 98.00 | 19.50 | Yes | 0.53 | 62 | |||||
(CH3NH3)2FeBr2Cl2 | Others | Cl | RP | 100.00 | Yes | No | Yes | 60 | ||||
(NH3(CH2)2NH3)FeCl4 | NH3(CH2)nNH3 | 2 | Cl | DJ | 53 |
Regarding 2D HOIPs presenting a DJ phase, limited research has been conducted to gain insight into hydrogen bonding in diammonium molecules. Only (NH3(CH2)2NH3)MnCl4 (ethylenediammonium – EDA – (NH3(CH2)2NH3)2+) is reported to show a SF transition less pronounced compared to those observed in RP HOIPs.5
All these findings suggest that the unique hydrogen bonding patterns for each molecule can impact the magnetic behavior. Nevertheless, a direct correlation between specific magnetic phenomena and particular features of the organic molecules is not evident.
Most of the previously discussed Cu-based HOIPs incorporating different organic series were synthesized using Cl and Br as halogens. Among the perovskites with the RP phase, it can be observed in Table 1 how compounds with CxH2x+1NH3+5,20,22–24 (aliphatic) and C6H5(CH2)xNH3+
5,49,100,118,119 (aryl) ammonium cations combined with Br exhibit slightly larger TC and J, as well as a lower ratio between interlayer and intralayer interactions, indicating higher J′ values compared to the Cl-based counterparts. However, more notable differences emerge in 2D HOIPs with the DJ phase, especially within the (NH3(CH2)xNH3)2+ series.26,27,29–33 When this 2D HOIP is x = 2 and contains Cl, its intralayer and interlayer exchange constants are 23.0 K and −13.7 K, respectively. Conversely, when the halogen used is Br, these two values increase to 38.2 K and −68.4 K, resulting in unusual stronger interlayer interactions compared to the intralayer ones. Additionally, TC also rises from 31.5 K to 72.0 K. Also, a compound mixing both halogens is reported,34 (NH3(CH2)2NH3)CuCl2Br2, presenting intermediate TC and J′ values (45 K and −31 K) but, surprisingly, having a value of 15 K for the interlayer interactions.
Nevertheless, when the metal is Cr, some discrepancies arise. For instance, compounds with the CxH2x+1NH3+ cation series with x = 135,41 mixed with Cl halogen exhibit a slightly higher value for J (13.0 K) than the Br analogue (11.6 K).44 Moreover, for this metal combined with C6H5(CH2)xNH3+ (x = 1), six compounds varying the Cl:
Br ratio have been reported.35–38,45 Generally, as expected, increasing the amount of Br (and reducing the amount of Cl) leads to a rise in the intralayer exchange constant, ranging from 10.6 K for pure Cl-based to 14.7 K for pure Br-based HOIP (Table 3). However, (C6H5CH2NH3)2CrBr3.3Cl0.7 deviates from this trend, presenting a lower J value (12.5 K) compared to (C6H5CH2NH3)2CrBr3Cl (13.1 K). Regarding their TC, only values for the pure 2D HOIPs and (C6H5CH2NH3)2CrBr3.3Cl0.7 are reported, following the same trend as Cu-based HOIPs. The compound containing only Br exhibits the highest TC (52 K), followed by the Br/Cl mixture (49 K) and finally, pure Cl-based HOIP presents the lowest one (37 K). Additionally, another work reported this 2D HOIP with the presence of Br and I,35,39 (C6H5CH2NH3)2CrBr3.3I0.7. Due to the larger atomic radius of I, one could expect larger TC and J values for this compound than for (C6H5CH2NH3)2CrBr4, however, it presents a much lower J value, 11.0 K and a similar TC, 51.0 K, highlighting the complex interplay of factors influencing magnetic properties in these materials.
Regarding 3D magnetic HOIPs, Pb-free (CH3NH3)FeCl3 show a typical paramagnetic behavior, different from the expected behavior of 2D Fe2+ based HOIPs.126 Numerous reports of 3D HOIPs can be found in literature related to Mn or Fe doped Pb halide perovskites, which will be detailed in the next section.
Moving toward 1D HOIPs, Lee et al.105,106,110 reported the magnetic properties of ABCl3 systems, studying the different effects of changing the organic cation A in a quasi 1D perovskite chains system containing Ni2+, Fe2+ or Co2+, with face-sharing octahedra. ANiCl3 compounds106 (A = N(CH3)4+, CH3NH3+, (CH3)2NH2+, C(NH2)3+, and CH(NH2)2+) follow the superexchange model, and the magnetic correlations within individual Ni–Cl chains turn from AFM to FM at an average Ni–Cl–Ni angle of approximately 78°. In contrast, the 1D chains of ACoCl3 compounds105 (A = CH3NH3+, CH(NH2)2+, C(NH2)3+), independently on the choice of the molecule, couple antiferromagnetically, but as evidenced by the lack of a relationship between the Co–Cl–Co angle and the corresponding intrachain coupling constant, the strength of the intrachain coupling cannot be forecasted by the superexchange model. However, there is a significant dependence on the interchain Co–Co distance, which directly affects the temperature at which long-range magnetic order is exhibited. While the AFeCl3 perovskites110 (A = CH3NH3+, CH(NH2)2+, C(NH2)3+, C3H5N2+) display intrachain ferromagnetic interactions without a direct trend between the Fe–Cl–Fe angle and the resulting coupling strength, the AFeCl3 perovskites110 (A = CH3NH3+, CH(NH2)2+, C(NH2)3+, C3H5N2+) display intrachain ferromagnetic interactions without a direct trend between the Fe–Cl–Fe angle and the resulting coupling strength.
1D perovskite of trimethylchloromethylammonium (TMCM–ClCH2NH3+) chromium chloride (TMCM–CrCl3), with infinite linear chains of face-sharing CrCl6 octahedra, separated by TMCM cations have been studied by Ai et al.127 The material exhibits a spin-canted antiferromagnetic behaviour with strong antiferromagnetic coupling and TN at 4.8 K. At the same time, 1D Cu2+ based HOIPs have been studied as well. The magnetic properties of (C5H8N3)CuCl3 (being C5H8N3 = 2-amino-4-methylpyrimidinium), can be described by the model of S = ½ antiferromagnetic dimers with J/kB = −122.7 K,128 where the exchange interactions can be evaluated using the results of Marsh et al.129 While, on the opposite side, 1D hybrids such as (C5H14N2)CuCl4 (hexahydro-1,4-diazepinium as organic cation) exhibit paramagnetic behavior due to the cancellation of ferro and anti-ferromagnetic components of superexchange interactions between the magnetic centers in the compound.130
Looking at 1D Mn-based HOIPs, Taniguchi et al. explored [(S)/(R)-MPEA]2[MnCl4(H2O)]55 with chiral R-/S-β-methylphenethylammonium (MPEA) as organic cations, showing that these compounds present antiferromagnetic interactions through the bridging single Cl ion, along individual chains of the Mn2+ ions. Additionally, spontaneous weak FM below 3.2 K was observed and ascribed to the canting of the antiferromagnetic spins induced by the Dzyaloshinskii–Moriya interactions.
Moving from 1D to 0D HOIPs, the magnetic properties of (H2DABCO)MX4·cH2O (DABCO = 1,4-diazabicyclo[2.2.2]octane), M = Mn and Cu; X = Cl and Br; (c = 0, 1, and 4) has been reported by Panda et al.131 All the compounds, showing a structure with isolated octahedra, are paramagnetic in nature, except for the Cu-based halides that exhibit dominant ferromagnetic interactions. Additional low-dimensional Ru based compounds studied by Vishnoi et al.,102e.g. (CH3NH3)2RuCl6, display 0D structure with [RuCl6]2− octahedra, showing a typical single-ion behavior without the influence of exchange interactions.
Finally, Zheng et al.132 have shown the effect of the dimensionality on the magnetic properties of divalent Mn-based metal halides. Compared to the canted antiferromagnetic behavior with Néel temperature (TN) at 45 K of the 2D compounds, the 1D system displays antiferromagnetic behavior with much lower TN at 5 K and absence of hysteresis in the M–H loop. Oppositely, the 0D perovskite presents the typical paramagnetic nature due to the isolated 0D [MnCl4]2− tetrahedral framework. Additionally, the halide Mn hybrids show nanosecond-scale spin coherence times that satisfy the relationship 2D > 1D > 0D at a high-temperature regime.
These results confirm how the different dimensionality of HOIP can play a crucial role in determining the final magnetic properties of the material, and how within each dimensionality, the magnetism is affected by the chemical design.
Regarding B1+–B23+ based double HOIPs, several studies report changes in their magnetic behaviour depending on the choice of the diamagnetic B1+, which affects the long-range superexchange interaction B23+–X−–B1+–X−–B23+. This was demonstrated in a study by Xue et al.,7 where they investigated the magnetism of 2D Fe–Cl based double perovskites [A4B1B2Cl8, B1+ = Ag+/Na+; B23+ = Fe3+/In3+], reported in Fig. 6a. They demonstrate that the nearest Fe3+−Fe3+ centers couple antiferromagnetically and the magnetic coupling strength can readily be tuned by the bridging diamagnetic B+. The studied (C6H5(CH2)2NH3)4AgFeCl8 compared to the (C6H5(CH2)2NH3)4NaFeCl8, presents higher magnetic coupling as a result of the differences in orbital hybridization, where Ag+ 4d orbitals strongly hybridize with Cl 3p orbitals, promoting the long-range superexchange interactions. Additionally, when organic cations are altered in the series of Ag–Fe–Cl, θCW become less negative from (C6H5(CH2)2NH3+, PEA) (−10.7 K) to (NH3(CH2)4NH3)2+ (1,4-butanediammonium, BDA) (−3.31 K) and to chiral R-MPEA (−1.37 K), suggesting a reduced antiferromagnetic coupling, as a result of the structural distortion in the inorganic frameworks due to the particular H-bond between each molecule and the inorganic framework (dFe−Fe: PEA < BDA < R-MPEA).
![]() | ||
Fig. 6 (a) Temperature dependence of the inverse magnetic susceptibility of PEA4AgFeCl8, PEA4NaFeCl8, BDA2AgFeCl8, and (R-MPA)4AgFeCl8. Reprinted with permission from ref. 7, copyright © 2022 American Chemical Society. (b) M–T measurement data at fixed applied field of 400 Oe for (C12H25NH3)2Cu1yMnyCl4 (y = 0, 0.5 and 0.75). Ref. 94 © 2019 Elsevier Inc. All rights reserved. |
Magnetic systems with different transition metal B13+ can be found in the literature. Binwal et al.104 reported new chloride double perovskites with Na+/Ag+ and Mo3+ metal ions and different dimensionality. Similarly to the Fe-based perovskites, the magnetic coupling of these systems (antiferromagnetic) primarily takes place through the –Mo3+–Cl–M+–Cl–Mo3+– pathway, the strength of which depends not only on the nearest-neighbor Mo/Mo distance, but also on the geometry of the magnetic sub-lattice. The 1D compound (CH3NH3)2AgMoCl6 features an antiferromagnetic transition temperature TN of 2.35 K, while the calculated frustration index of 1.2 indicates that this 1D system is magnetically non-frustrating. Compared to the 2D system (1,4-(NH3(CH2)4NH3))2AgMoCl8, it displays higher antiferromagnetic exchange due to reduced distance between neighbor-spins. Similarly, the compounds containing Na+–Mo3+ are ordered antiferromagnetically below TN (5.2–6.8 K). In the (CH3NH3)2NaMoCl(6−x)Brx series of compounds,134 the strength of exchange interaction increases with increasing the Br content x.
The weak antiferromagnetic coupling between Mo3+ ions was further confirmed by Vishnoi et al.6 The group additionally studied Ru3+ based double perovskites. Differently, the magnetic behavior of these compounds is more complex due to the unquenched orbital angular momentum of the low spin t2g5 of Ru3+. The 1D (CH3NH3)2NaRuCl6 does not follow the behavior described by Kotani in which compounds showing isolated and undistorted [RuCl6]3− octahedra exhibit a single ion behavior without the influence of exchange interaction. The lower moment of the double HOIPs in comparison with the ideal Kotani prediction might arise from increased exchange interactions resulting from the higher connectivity between RuCl6 within the chains.102
Few studies are present for double perovskite combining metal cations with the same oxidation state B12+–B22+. 2D layered (CH3(CH2)11NH3)2Cu1−yMnyCl4 hybrid systems at different compositions (y = 0.0, 0.5, 0.75 and 1.0) have been reported by Bochalya et al.94 The magnetic M–T curves are shown in Fig. 6b. In this study, while crystals with y = 0, 0.5 and 0.75 show typical paramagnetic to ferromagnetic transition at Tc ∼ 10 K, compounds with y = 1, present the usual antiferromagnetic interactions. The system with composition y = 0.75, i.e., (CH3(CH2)11NH3)2Cu0.25Mn0.75Cl4, shows two transitions at temperatures of ∼9 K and ∼41.5 K. Although this suggests the coexistence of ferromagnetic and antiferromagnetic orders in y = 0.75, what happens is that both the (CH3(CH2)11NH3)2CuCl4 and (CH3(CH2)11NH3)2MnCl4 crystalline phases are present, and thus, their magnetic behavior.
More complex systems containing three metal cations with different oxidation states have also been reported. Connor et al.135 synthesized a double layered perovskite Cu+–In3+ with half of the metal sites replaced by Cu2+. The alloyed perovskite shows weak ferromagnetic interactions between local Cu2+ moments with no long-range order down to T = 2 K, without the formation of any Cu2+ cluster.
Rajamanickam et al.116 reported how the magnetic properties change at different Mn amounts for the (CH3NH3)Pb1−yMnyI3 films at room temperature. The magnetization as a function of magnetic field (Fig. 7a) shows that the pristine (y = 0) and Mn doped films (y = 0.01) exhibit negligible M; nevertheless, M emerges for films with y > 0.03, showing ferromagnetic behavior. A drastic enhancement in the saturation magnetization was found for y = 0.15 (Fig. 7b). The likely presence of Mn2+–I−–Mn3+ motifs in the sample suggests a plausible explanation for the ferromagnetic coupling observed between Mn ions. That is, the double exchange mechanism proposed by Zener136 was based on FM arising from the indirect interaction between ions with multiple oxidation states, and mediated by the halogen. However, these (CH3NH3)Pb1−yMnyI3 films should be considered as a particular case. The effect of Fe2+ on the magnetism of Pb-based HOIPs has also been reported in the literature both from theoretical137 and experimental points of view.98,108,109
![]() | ||
Fig. 7 (a) SQUID M–H curves for the MAPb1−xMnxI3 films. (b) Dependence of saturation magnetization (Ms) on Mn doping concentration in 450 nm thick MAPb1−xMnxI3 films. Reprinted with permission from ref. 116. Copyright © 2021 American Chemical Society. |
Fu et al.109 showed that depending on the composition, different magnetic behavior can be expected. (CH3NH3)(Pb0.8Fe0.2)I3 is paramagnetic, while (CH3NH3)(Pb0.4Fe0.6)I3 is a low-temperature antiferromagnetic with a TN of 23 K. Paramagnetic behavior has been reported as well by Bonadio et al.98 for Fe-doped microwires halide perovskites. In contrast, Pb-based HOIPS, e.g. (CH3NH3)PbCl3, co-doped with magnetic (Fe2+) and aliovalent (Bi3+) metal ions, can present ferromagnetic interactions below 12 K when the concentration of Fe2+ and Bi3+ reaches respectively 3% and 2.9%.108
Therefore, these results show how it is possible to take advantage of the introduction of different metal cations in the HOIPs for tuning their magnetic properties.
Magnetic applications of HOIPs containing transition metal cations are still in the early stages and most studies focus on doping well-established Pb-based HOIPs, like (CH3NH3)PbI3 and (C6H5(CH2)2NH3)2PbX4 with the purpose of comparing their behavior before and after the inclusion of magnetic counterparts. Nonetheless, there are also few reports covering pure transition-metal-based HOIPs, such as (MBA)2CuCl4 (MBA = methylbenzylammonium) and (CH3NH3)2MnCl4.
This section explores applications that rely directly on the magnetic properties of HOIPs, from magneto-optoelectronic devices to spintronics, demonstrating the transformative impact of transition metal doping on HOIP functionality.
![]() | ||
Fig. 8 (a) Schematic diagram of the photovoltaic device fabricated to study the magnetic dependence of the photocurrent on MAPb1−xMnxI3 (x = 0, 0.05, 0.1, 0.2). (b) Relative change of short-circuit photocurrent in time under different external magnetic field values, and comparison of the behaviour between the ferromagnetic MAPb1−xMnxI3 and the non-magnetic MAPbI3. Reprinted with permission from ref. 136. Copyright © 2020, American Chemical Society. (c) Spin susceptibility over a broad range of illumination intensities. The ferromagnetic order decreases monotonically after a threshold, leading to the melting of the ferromagnetism. The upper inset represents the working principle of a RAM based on this phenomenon. The lower inset depicts the crystal structure of MAPbI3. Ref. 63 copyright © 2016, The Author(s). (d) Schematic diagram of the magnetic conductive-probe atomic force microscopy set-up used for the study of spin filtering along the c-axis of (R/S-MBA)2CuCl4. (e) and (f) Current intensity as a function of applied voltage for each enantiomer, under the probe magnetized in the upward (blue) and downward (red) directions, showing the opposite response. The inset in (f) represents the crystal structure of (R/S-MBA)2CuCl4. Ref. 65 © 2021 Wiley-VCH GmbH. |
To investigate the influence of magnetic fields on photoelectric performance, (CH3NH3)Pb0.8Mn0.2I3 was chosen and compared to undoped (CH3NH3)PbI3. The relative change in short-circuit photocurrent (MIP) was defined as the difference between the photocurrent measured at a specific magnetic field (IP(H)) and that at zero magnetic field (IP(0)), divided by the photocurrent at zero magnetic field (I(0)), as illustrated in Fig. 8b. This relationship is expressed by the equation:
In experiments, an increase in MIP of 0.5% was observed in the Mn-doped perovskite, while the undoped MAPbI3 exhibited a 3.3% decrease. This enhancement in photocurrent in the doped perovskites is attributed to a reduction in photoresistance, a result of Mn ions’ spin alignment under the magnetic field, which intensifies the d5-eg double exchange and facilitates electron hopping.
The RKKY interaction, a phenomenon where magnetic moments in a material are indirectly coupled through conduction electrons, can create oscillating magnetic interactions that vary in strength and orientation depending on electron density.
This effect is promising for magnetic memory applications, allowing magnetic bits to be written and rewritten via light illumination. By disrupting the existing magnetic order with light, a small, localized field can then be applied to establish a new magnetic moment direction (inset of Fig. 8c). This mechanism paves the way for rapid, low-power optical control of spin states.
This spin filtering effect is attributed to chiral-induced spin selectivity (CISS), which arises from the chiral structure of the HOIPs combined with a high degree of spin polarization. The results align with CISS theory, which suggests that the chiral molecular structure can influence electron spin orientation, thereby enabling selective spin filtering.
![]() | ||
Fig. 9 (a) and (b) Low-temperature (4 K) PL spectra of Mn: PEA2PbI4 under a negative and a positive magnetic field respectively, showing opposite responses between the spectra collected for right-handed and left-handed detection for each case. (c) Behavior of the degree of CPL with magnetic field as a function of the dopant content. The degree of CPL increases with increasing Mn content, while showing no variation for undoped PEA2PbI4. Ref. 64 copyright © 2021, The Author(s). (d) Representation of the crystal structure of (CH3NH3)2MnCl4. (e) Temperature dependence of the PL intensity on heating and cooling, showing the different behavior of PL below and above 47 K. (f) Change of PL with magnetic field at 4.2 K for in-plane and out-of-plane configurations, highlighting the AFM to SF and canted FM transitions. Ref. 66 © 2024 Wiley-VCH GmbH. |
Two mechanisms are proposed to explain this CPL behavior. First, variations in carrier excitation rates and dark exciton formation may differ across spin states, leading to a spin population imbalance during recombination that manifests as CPL. Alternatively, Mn dopants might introduce an additional radiative recombination pathway. In this pathway, a normally forbidden transition becomes accessible if Mn spin flips simultaneously with exciton recombination. This pathway shows increased efficiency for the exciton spin state aligned with Mn spins under an external magnetic field, resulting in CPL through a similar spin population imbalance.
In a complementary study, the modulation of photoluminescence (PL) in the 2D HOIP (CH3NH3)2MnCl4 under varying magnetic fields and temperatures has been systematically studied by Zhang et al.66 using detailed PL spectroscopy measurements. They demonstrated that the PL characteristics of (CH3NH3)2MnCl4 are highly dependent on the magnetic ordering of Mn2+ ions, which undergo transitions between antiferromagnetic canted ferromagnetic and SF states as temperature and external magnetic fields vary.
Below 47 K, the Mn2+ spins exhibit antiferromagnetic ordering that suppresses PL by facilitating energy transfer to non-radiative states through superexchange interactions (Fig. 9d). In contrast, as an external magnetic field is applied, antiferromagnetic ordering is progressively disrupted, and a SF transition occurs at approximately 3.5 T, leading to partial ferromagnetic alignment and a pronounced increase in PL intensity (Fig. 9e). This enhancement is attributed to the suppression of non-radiative energy transfer, allowing more efficient radiative recombination. Moreover, at very high magnetic fields, the complete alignment of Mn2+ spins into a ferromagnetic state further increases PL intensity and induces a redshift in the PL peak energy.
Two key mechanisms are identified to explain the observed magneto-optical effects. First, antiferromagnetic ordering favors non-radiative energy loss pathways, reducing PL intensity, while ferromagnetic ordering mitigates these losses, enhancing optical transitions. Second, the external magnetic field alters the selection rules governing d–d transitions in Mn2+ ions, facilitating normally forbidden transitions by breaking spin degeneracy. As the magnetic field increases, energy levels split further, leading to shifts in the emission spectrum and different trends in the PL intensity, which vary depending on the temperature. For example, in the range 60–80 K the intensity of PL reduces as the magnetic field increases, while for temperatures lower than 30 K, the PL intensity is enhanced. It is the competitive interplay between antiferromagnetic and ferromagnetic orderings, together with electron–phonon interactions, what dictates the temperature and field-dependent PL properties of (CH3NH3)2MnCl4.
This work highlights the potential of magnetic-field-modulated PL in Mn2+-based perovskites for applications in magneto-optical and spin-photonic devices, where fine-tuning of emission properties is critical.
Regarding magneto-optoelectronics, Zhang et al.138 synthesized (C4H10N)MnBr3 (pyrrolidinium as the organic cation) and studied its electrical, magnetic and optical properties. They found that this HOIP is ferroelectric, shows weak FM and exhibits intense red PL under UV excitation. The combination of multiferroicity and PL emission opens the door to fabricating multifunctional magneto-optoelectronic devices.
In the field of magnetocaloric materials for cooling devices, Septiany et al.139 investigated the magnetocaloric properties of the ferromagnetic (PMA)2CuCl4. They determined the performance of the material via the relative cooling power (RCP), which can be obtained by determining the magnetic entropy change as a function of temperature. They found that under a magnetic field of 7 T the RCP reaches 47.8 J kg−1, a value large enough to consider its application of cooling devices. A similar study was performed by Ma et al.140 on (CH3NH3)2CuCl4 showing a large magnetocaloric effect as well. Other studies focus on the HOIPs (CH3(CH2)11NH3)2Cu(Br1−xClx)4, (C6H9(CH2)2NH3)2Cu(Br1−xClx)4141 and (CH3CH2NH3)2CoCl4.142
As for multifunctional devices, several research groups have demonstrated that 2D HOIPs can exhibit diverse properties, making them suitable for a variety of applications. For example, Sun et al.143 synthesized and studied the 2D HOIP (PED)CuCl4 and (BED)2CuCl6 (being PED – N-phenylethylenediammonium and BED – N-benzylethylenediamine), which show strong FM below TC = 13 K. They also found reversible thermochromism, which shows applications in thermal sensors; and a six-order-of-magnitude conductivity change in (BED)2CuCl6 upon temperature change, opening the door to designing multifunctional devices.
Likewise, Xiong et al.,144 using the chiral molecule C6H15ClNO (CTA = 3-chloro-2-hydroxypropyltrimethylammonium), synthesized the chiral HOIPs (S-CTA)2CuCl4 and (R-CTA)2CuCl4. Their extensive investigation reveals properties such as ferroelasticity, thermochromism, and chirality-induced effects, along with interactions among these characteristics. Notably, they identify seven physical channel switches, indicating the potential to induce transitions or changes in the properties of these perovskites through external stimuli. While their study does not explicitly address the magnetic properties of these materials, the presence of Cu suggests that they could exhibit magnetic behavior, thereby opening avenues for further exploration in applications like spintronic devices, memory storage, sensors, and smart windows, all of which could benefit from magnetism.
Another example is the 2D polar ferromagnets (3-ClbaH)2CuCl4, (4-ClbaH)2CuCl4 and (2-ClbaH)2CuCl4 (ClbaH+ = chlorobenzylammonium) reported by Han et al.145 These HOIPs not only present FM, but also second harmonic generation and polarity, so they are candidates for multifunctional devices. Other perovskites that are proved, experimentally and/or theoretically, to have multiferroicity are (C6H5CH2NH3)CuCl4,146 (CH2NH3)2[FeCl4], (CH3NH3)2[FeCl4]92 and (CH3CH2CH2NH3)2FeCl4.58
In summary, magnetic HOIPs present exciting opportunities for advancing applications in spintronics, magneto-optoelectronics, and multifunctional devices due to their unique magnetic properties and tunability. Transition metal doping, in particular, has enabled promising proof-of-concept applications that leverage magnetism within perovskite frameworks, demonstrating the potential for high-impact innovations. Notably, as discussed in Section 4.1, pure transition-metal-based HOIPs exhibit stronger magnetic interactions compared to doped perovskites, due to superexchange mechanisms. This distinction could be especially relevant for future applications that require more robust magnetic coupling. However, despite their potential, most application-focused research has centred on doped HOIPs. Additionally, research on applications is in the early stages and many challenges need to be addressed to transition these materials from the laboratory to real-world devices. Continued investigation into their magnetic behaviours and interaction mechanisms will be critical for realizing the full potential of these materials in future technological applications.
• Stability enhancement: the ambient instability of HOIPs poses a significant barrier to their use in devices, and several solutions have been proposed namely for the most studied Pb ones such as surface modifications, encapsulation, or lattice substitutions. But these approaches should be explored in magnetic HOIPs to make them viable for long-term applications. This is particularly important for devices that require consistent performance in diverse environments.
• Fine-tuning magnetic properties: while the magnetism of Cu-based HOIPs has been deeply investigated, this review highlights that Mn, Cr, and Fe-based HOIPs also show promising, tunable characteristics, which yet remain underexplored. Mn-based HOIPs, for example, offer multifunctional potential since Mn in octahedral coordination can produce red photoluminescence, which can be chirality-dependent, particularly with the inclusion of chiral organic molecules. Likewise, Fe-based HOIPs present further possibilities, such as incorporating ferroelectric properties. Additionally, double perovskites increase the opportunities for tuning the properties of the HOIP by e.g. varying each metal cation differently. Further investigation into the effects of dimensionality, organic cation selection, halide composition, as well as expanding research on these underexplored systems, could enable on-demand magnetic characteristics for specific device requirements, facilitating more versatile applications in spintronics, magneto-optics, and optoelectronics. In this direction, the increase of the Curie temperature and the study of new phenomena already highlighted in metal halide 2D magnets such as spin dynamics and topological features are two key points to explore in the future.
• Advanced synthesis techniques: here we focus on synthesis protocols for obtaining single bulk crystals, but integrating these materials in current technologies and devices demands large-scale, reproducible, and efficient methods that are able to produce good quality layers over defined areas. Improvements in solution and vapor-phase approaches are key to achieving commercially feasible, high-quality and crystal-oriented films at a lower cost, making HOIPs a feasible option for device integration.
• From an application perspective, magnetic HOIPs offer immense potential across multiple domains, such as magnetic data storage, magneto-optics, sensors, spin filters, and spatial light modulators.10,146 However, many current studies focus on the fundamental properties of HOIPs, with relatively few reporting fully realized devices. To bridge this gap, more research on lab-scale prototypes is needed, as well as deeper insights into the underlying physics, such as spin-dependent photo-physics and exciton polarization control.64 By addressing these scientific and engineering challenges, HOIPs could become key materials for low-power, magnetically controlled devices, marking a significant advancement in spin-based and multifunctional technologies.
This journal is © The Royal Society of Chemistry 2025 |