Next Article in Journal
Magnetic Properties and Coercivity Mechanism of Nanocrystalline Rare-Earth-Free Co74Zr16Mo4Si3B3 Alloys
Previous Article in Journal
Advances in Functional Magnetic Nanomaterials for Water Pollution Control
Previous Article in Special Issue
A Dy2 Complex Constructed by TCNQ·− Radical Anions with Slow Magnetic Relaxation Behavior
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Molecular Nanomagnets with Photomagnetic Properties: Design Strategies and Recent Advances

1
Frontiers Science Center for New Organic Matter, State Key Laboratory of Advanced Chemical Power Sources and Department of Chemistry, College of Chemistry, Nankai University, Tianjin 300071, China
2
i-Lab, Suzhou Institute of Nano-Tech and Nano-Bionics (SINANO), Chinese Academy of Sciences (CAS), Suzhou 215123, China
*
Authors to whom correspondence should be addressed.
Magnetochemistry 2025, 11(9), 77; https://doi.org/10.3390/magnetochemistry11090077
Submission received: 21 July 2025 / Revised: 27 August 2025 / Accepted: 28 August 2025 / Published: 31 August 2025

Abstract

The magnetic properties of molecular nanomagnets can be finely modulated by light, which provides great potential in optical switches, smart sensors, and data storage devices. Light-induced spin transition, structure changes, and radical formation could tune the static and dynamic magnetic properties of molecular nanomagnets with high spatial and temporal resolutions. Herein, we summarize the design strategies of photoresponsive molecular nanomagnets and review the recent advances in transition metal/lanthanide molecular nanomagnets with photomagnetic properties. The photoresponsive mechanism based on spin transition, photocyclization, and photogenerated radicals is discussed in detail, providing insights into the photomagnetic properties of molecular nanomagnets for advanced photoresponsive materials.

1. Introduction

Stimuli-responsive materials have recently garnered great attention due to their physical and chemical properties being modulated by external stimuli, such as temperature, electric field, and light [1,2,3,4]. Among the diverse stimuli-responsive materials, photoresponsive molecular nanomagnets are particularly intriguing candidates, showing great potential toward applications such as high-density information storage, optical switching devices, smart sensors, and spintronic devices [5,6,7,8].
Molecular nanomagnets, including single-molecule magnets (SMMs) and single-chain magnets (SCMs), are usually constructed by magnetically anisotropic transition or lanthanide metal centers and organic ligands [9,10,11]. The electronic state of metal ions that is defined by the coordination sphere is crucial for the rational design of advanced molecular nanomagnets possessing tailored photomagnetic properties [12,13,14]. For example, photomagnetic properties were observed in cyanide-bridged spin-crossover (SCO) complexes with the light-induced excited spin state trapping (LIESST) effect. The LIESST effect was also applied to modulate the bistable states of molecular nanomagnets, such as {[W(CN)8][(FeII)(bib)2](bibH)}·2CH3OH (bib = 1,4-bis(1H-imidazol-1-yl)benzene), representing an example of reversible switching between SCM and SMM behaviors via light [15]. On the other hand, lanthanide SMMs have been extensively studied, achieving groundbreaking progress, especially in effective energy barrier (Ueff) and blocking temperature (TB) [16,17,18,19,20,21,22,23]. Early studies have demonstrated that light irradiation can effectively modulate the static magnetic susceptibility of lanthanide SMMs; however, regulating their dynamic magnetic susceptibility remains challenging [24,25,26]. Selected examples of photoresponsive lanthanide SMMs generally have photoresponsive cyclic units, such as 1,2-bis(4-pyridyl)ethylene [27], 9-diethylphosphorylmethylanthracene [28], and dipyridyldithienylethene [29], as well as photoresponsive radical precursors [30]. To gain insights into the mechanism of photomagnetic properties of molecular nanomagnets, we summarize the advancements of transition metal and lanthanide molecular nanomagnets with static and dynamic magnetic properties modulated by light (Figure 1). The roles of selecting organic ligands that can undergo light-induced ring-opening/ring-closing transitions or generate stable radicals under illumination are discussed, providing key structural factors for the rational design and synthesis of high-performance lanthanide molecular nanomagnets with photoinduced magnetic prop-erties.

2. Design Strategies for Molecular Nanomagnets with Photomagnetic Properties

Molecular nanomagnets with photomagnetic properties are primarily synthesized by introducing photoresponsive components, notably SCO units, photo-isomerizing organics, or radical-forming moieties. In transition metal molecular nanomagnets, incorporation of SCO units could facilitate optical switching of spin states to modulate the magnetic properties of SMMs and SCMs [31,32,33,34]. For lanthanide molecular nanomagnets, light-induced ligand transformations such as photocyclization or radical formation could modify the ligand field of the metal center and introduce additional spin carriers, thereby tuning room-temperature magnetic moments and/or dynamic magnetic susceptibilities [27,28,29,30].

2.1. Photoresponsive Spin-Crossover Units

The photoinduced electron transfer is fundamental in photophysics and photochemistry, with extensive applications in artificial light sources, solar energy conversion, and photocatalysis [35,36,37,38,39,40,41,42,43,44,45,46,47,48,49], enabling the regulation of properties at a molecular level [50,51,52]. SCO complexes undergo spin transition at the orbitals of metal sites upon illumination, leading to a change in the spin state [53,54,55,56,57,58,59,60,61,62,63,64]. However, assembling SCO units with coordination bonds to form photoresponsive compounds remains highly challenging because minor changes in the coordination environment of metal sites may influence the spin transition to some extent. Through the rational design of molecular structures based on the SCO unit, it is desired to achieve multifunctional molecular materials with controllable photomagnetic properties.

2.2. Photoinduced Structural Change in Organic Molecules

Photochromism refers to the reversible transformation of a chemical substance be-tween two or more forms that exhibit distinct absorption to light [65,66]. Photochromic systems can undergo reversible transformations between distinct states with varying properties, making them widely applicable in data storage, sensors, bioimaging, and nanomachines [67,68,69,70,71,72,73,74,75].
Typical photochromic ligands, for example, are shown in Figure 2. Upon light illumination, the structural changes most commonly observed are trans-to-cis isomerization and ring-opening/ring-closing transitions [75,76,77]. Under ultraviolet light, azobenzene can undergo trans-to-cis isomerization, and the cis isomer can be reverted to the trans form through exposure to visible light or heat [76,77]. For diarylethene, a [2 + 2] photocycloaddition can occur when the distance between parallelly arranged alkene bonds is appropriate [78,79,80,81,82]. Similarly, 9-anthrylmethylphosphonic acid can undergo a [4 + 4] cycloaddition under ultraviolet light when the spatial requirements are satisfied [83,84]. Upon ultraviolet irradiation, dithienylethene can transform from a colorless open-ring state to a colored closed-ring state and can be switched back to its original state via visible light irradiation [85,86]. In the case of spiropyran, ultraviolet irradiation can result in the photochemical cleavage of the C–O bond, resulting in its conversion from the colorless spiropyran state to the colored merocyanine state. Upon turning off the ultraviolet light, it can revert to its initial colorless state [87,88]. Fulgides can transform from a colorless open-ring state to a colored closed-ring state under ultraviolet light and revert to the initial state through a ring-opening reaction when exposed to visible light [89,90].

2.3. Photogenerated Organic Radicals

In addition to photochromic ligands, another important category of photoresponsive ligands is photogenerated organic radicals. These compounds have been used for various applications, such as smart windows, erasable and inkless printing, catalysis, sensing, artificial photosynthesis, and solar cells [91,92,93,94,95].
As shown in Figure 3, pyridine derivatives, viologens, triazoles, 9-anthracene carboxylic acids, and naphthalenediimides can be converted to radicals upon light illumination through electronic transition [96,97,98]. In recent years, there has been a growing interest in the study of not only molecular nanomagnets but also multifunctional metal–organic frameworks with photogenerated radicals [99,100,101].

3. Photomagnetic Properties of Transition Metal Molecular Nanomagnets

The photomagnetic properties of transition metal molecular nanomagnets primarily arise from the alteration of their electronic structures and/or spin states. Light-induced spin state changes could serve as the fundamental mechanism behind photomagnetic properties. For example, in the six-coordination Fe(II) complexes, the transition from a low-spin state to a high-spin state could be induced by light illumination, thereby influencing the magnetic properties of the complexes. The SCO units have been assembled with cyanide to form chains, wherein the spin state, anisotropy, and magnetic interaction were photoswitched to tune the magnetic properties [15,31,32,33,34,102,103]. Additionally, transition metal molecular nanomagnets with photoinduced metal-to-metal electron transfer (MMET) and photoinduced structural changes in the ligands were also reported [104,105]. Reliable experimental characterization is required for understanding the photoinduced mechanism, which is typically accomplished through advanced characterization techniques. The superconducting quantum interference device (SQUID), physical property measurement system (PPMS), and pulsed electron paramagnetic resonance (EPR) spectrometer are commonly employed to determine the photomagnetic properties of molecular nanomagnets [106,107,108,109].

3.1. Photomagnetic Properties of SCO-Type Transition Metal Molecular Nanomagnets

In 2013, Long’s group reported an iron(II) molecular nanomagnet with photomagnetic properties, [Fe(1-propyltetrazole)6](BF4)2 (Figure 4a) [31]. Under an external magnetic field, this complex exhibited fully reversible switching between the S = 0 and S = 2 states. After irradiation by a 505 nm light for 7 h at 10 K, an LIESST effect was observed. The high-spin state showed no significant decay after the light was turned off. An alternating current (AC) magnetic susceptibility measurement was performed using a SQUID magnetometer, which revealed slow magnetic relaxation behavior in the high-spin Fe(II) state under an external magnetic field (Hdc = 2000 Oe) in the temperature range of 1.9–5.0 K, with an effective barrier of 22 K. An analysis of EPR spectra demonstrated that the anisotropy parameters of the light-induced phase of [Fe(1-propyltetrazole)6](BF4)2 agree with the SMM behavior, with a total spin reversal barrier of 72 K. The effective barrier of 22 K from ac magnetic susceptibility is lower than that of the total spin reversal barrier due to the presence of vibronic coupling, typically observed in mononuclear SMMs. In the same year, Smith’s group reported another iron(II) molecular nanomagnet with photomagnetic properties, PhB(MesIm)3FeII-N = PPh3 (Figure 4b). After 18 h of white light irradiation at 10 K, an LIESST effect was observed. Magnetic susceptibility measurements under light irradiation indicated a transition from the low-spin state to the high-spin state. The high-spin complex exhibited slow magnetic relaxation behavior under an external magnetic field (Hdc = 1000 Oe) in the temperature range of 1.8–4.6 K, with an effective barrier of 22 K [32].
Cyanide-bridged SCO complexes have shown intriguing photoinduced magnetic properties. These complexes generally consist of divalent or trivalent iron/cobalt ions linked by cyanide ligands, resulting in either chain or three-dimensional framework structures. From 2013 onwards, Liu’s group reported a series of molecular nanomagnets with photomagnetic properties, including {[FeIII(Tp*)(CN)3]2FeII(bpmh)}·2H2O, (bpmh, N,N′-bis-pyridin-4ylmethylene hydrazine), {[FeIII(bpy)(CN)4]2[CoII(phpy)2]}·2H2O, (bpy, 2,2′-bipyridine, phpy, 4-phenylpyridine), {[W(CN)8][(FeII)(bib)2](bibH)}·2CH3OH, {[(pzTp)FeIII(CN)3]4FeII2(Pmat)4}n·12H2O, {[pzTpFeII(CN)3]4CoII2(Bib)4}·3H2O, and {[pzTpFeII(CN)3]2CoII(Bpi)2}·CH3CN·4H2O [15,33,34,102,103] (Figure 4c–g). {[FeIII(Tp*)(CN)3]2FeII(bpmh)}·2H2O is a double zigzag chain that includes Fe(II) spin-crossover units and well-isolated paramagnetic Fe(III) ions. This chain exhibited thermally reversible SCO at the Fe(II) sites. Following 12 h of irradiation with 473 nm blue light at 5 K, an SCM behavior was triggered, accompanied by the LIESST effect at the Fe(II) sites and synergistic ferromagnetic coupling between the photoinduced high-spin Fe(II) and low-spin Fe(III) ions [33] (Figure 5).
In 2021, Liu’s group reported a molecular nanomagnet with photomagnetic properties in the form of {[W(CN)8][(FeII)(bib)2](bibH)}·2CH3OH, which exhibited a reversible LIESST effect when subjected to alternating infrared (808 nm) and blue (473 nm) light, enabling reversible switching between paramagnetic high-spin and diamagnetic low-spin states of the Fe(II) center [15]. At 1.8 K, irradiation with infrared light generated SCM behavior due to the magnetic interactions between the photogenerated high-spin Fe(II) centers and tungsten(V) centers, accompanied by wide magnetic hysteresis loops and a large coercive field of 19 kOe (Figure 6a). Irradiation by blue light resulted in isolated Fe(II)–W(V) species exhibiting SMM behavior. Magnetic hysteresis loops measured at 3.3 K showed that the high-spin Fe(II) state under infrared light maintained a relatively narrow hysteresis loop, while the low-spin Fe(II) state under blue light led to the closure of the hysteresis loop (Figure 6b). This represents an interesting example of reversible switching between SCM and SMM accompanied by light-controlled magnetic hysteresis loop switching.

3.2. Photomagnetic Properties of Other Transition Metal Molecular Nanomagnets

Dithienylethene is a photochromic molecule that undergoes reversible ring-opening and ring-closing reactions when exposed to light of different wavelengths, resulting in significant alterations to its electronic and optical properties. In 2009, Irie’s group reported a chain-like molecular nanomagnet based on closed-ring dithienylethene, [Mn4(hmp)6(dae-c)2(H2O)2](ClO4)2·CH3CN·4H2O (Mn4-c; Hhmp = 2-hydroxymethylpyridine, H2dae = 1,2-bis(5-carboxyl-2-methyl-3-thienyl)perfluorocyclopentene) [104]. Upon irradiation with visible light (λ > 480 nm), the closed-ring Mn4-c underwent an open-ring reaction, which enhanced the magnetic interactions between the [Mn4] units and altered its slow magnetic relaxation behavior. In 2019, Yamashita’s group reported zero-dimensional molecular nanomagnets with open-ring and closed-ring dithienylethene, [Mn2(saltmen)2(dae-o)] (Mn2-o) and [Mn(saltmen)(dae-c)]·H2O·Et3N (Mn-c) (H2saltmen = 2,2′-((1E,1′E)-((2,3-dimethylbutane-2,3-diyl)bis(azaneylylidene))bis(methaneylylidene))diphenol, H2dae = 1,2-bis(5-carboxyl-2-methyl-3-thienyl)perfluorocyclopentene) [105]. Under an external magnetic field, the photoproducts Mn2-o-UV and Mn-c-vis exhibited a change in AC magnetic susceptibilities compared to Mn2-o and Mn-c. The effective energy barriers were 15.89, 18.10, 13.96, and 19.91 K for Mn2-o, Mn2-o-UV, Mn-c, and Mn-c-vis, respectively.
Cyano-bridged heterometallic {FeCo} chain systems provide the opportunity to achieve photoswitchable SCMs via light-induced MMET [110,111]. A cyanide-bridged chain {[(Tp)Fe(CN)3]2Co(BIT)}·2CH3OH (Tp = hydrotris(pyrazolyl)borate, BIT = 3,4-bis-(1H-imidazol-1-yl)thiophene) exhibits reversible, multi-phase transitions driven by MMET [110]. Light-irradiated single-crystal X-ray diffraction reveals that MMET occurs exclusively between the Co center and one Fe site, switching the chain between diamagnetic [(Fe1)ᴵᴵLS–CoᴵᴵᴵLS–(Fe2)ᴵᴵᴵLS] and metastable ferromagnetic [(Fe1)ᴵᴵᴵLS–CoᴵᴵHS–(Fe2)ᴵᴵᴵLS] properties at low temperatures. At 10 K, alternating 946 and 532 nm illumination enables the reversible on–off switching of SCM behavior. Moreover, rapid thermal quenching traps a distinct supercooled phase (HTsc, P212121), which differs from both the low-temperature phase (LT, Pnma) and the light-induced high-temperature phase (HT*, P21/n). In addition, removal of lattice methanol upon desolvation suppresses MMET and SCM behavior, underscoring the essential role of hydrogen bonding in tuning redox potentials. These findings establish {[(Tp)Fe(CN)3]2Co(BIT)}·2CH3OH as an uncommon platform that integrates light-, thermal-, and solvent-responsive SCM.

4. Photomagnetic Properties of Lanthanide Molecular Nanomagnets

The combination of photoresponsive ligands with lanthanide ions could result in photoresponsive molecular nanomagnets, which have important applications for high-density data storage, quantum information processing, and optical switches. This strategy combines the strong magnetic anisotropy of lanthanide ions with the dynamic characteristics of photoresponsive ligands. In recent years, researchers have assembled lanthanide ions with photocyclization organic compounds and radical precursors as photoresponsive units to construct various lanthanide molecular nanomagnets with photomagnetic properties [27,28,29,30,112,113,114,115,116,117,118,119,120,121,122] (Figure 7 and Table 1).

4.1. Photomagnetic Properties of Photocyclization-Type Lanthanide Molecular Nanomagnets

Photocyclization-type organic ligands include diarylethene, 9-anthrylmethylphosphonic acids, dithienylethenes, spiropyrans, fulgides, and so on [123,124,125,126,127], which have been employed as photoresponsive units in lanthanide molecular nanomagnets. Diarylethenes and 9-anthrylmethylphosphonic ligands can undergo [2 + 2] and [4 + 4] cycloaddition reactions, respectively, under ultraviolet light [27,28,112,113]. Dithienylethene can undergo cycloaddition reaction when exposed to the light of a specific wavelength [29]. It is worth noting that 9-diethylphosphonomethylanthracene (depma) can undergo photoinduced or thermally induced ring-opening reactions, enabling reversible structural conversions [28,113].
To date, the photomagnetic properties of photocyclization-type lanthanide molecular nanomagnets are primarily focused on [2 + 2] and [4 + 4] cycloaddition reactions induced by light based on the alterations of the coordination environment of the metal centers [27,28,112,113].
The photoinduced cyclization process of 1,2-bis(4-pyridyl)ethylene (bpe) primarily occurs through a [2 + 2] cycloaddition reaction. Under illumination, the double bonds of bpe interact with another molecule, leading to the formation of a cyclobutane structure. In 2015, Tong’s group reported a mononuclear Dy(III) complex with bpe, [Dy(bpe)(H2O)4(NO3)2](NO3)·2bpe [27], where the alkene bonds were arranged parallel at a distance of 3.67 Å. Upon UV irradiation, a [2 + 2] cycloaddition reaction occurred, yielding a binuclear Dy(III) complex, [Dy2(tpcb)(H2O)8(NO3)4](NO3)2·2bpe·tpcb (tpcb = tetrakis(4-pyridyl)cyclobutane) (Figure 8a,b). Under a zero-dc field, the mononuclear complex exhibited slow magnetic relaxation, while the photoproduced binuclear complex displayed only a very weak temperature-dependent magnetic relaxation (Figure 8c,d).
In 2017, another mononuclear Dy(III) complex (Hbpe)2[Dy(bpe)(H2O)(4-pyO)(NO3)(SCN)3]SCN with bpe [112] was reported, where the alkene bonds were parallel with a distance of 3.44 Å. UV irradiation induced a [2 + 2] cycloaddition reaction, yielding a binuclear Dy(III) complex (Hbpe)2(H2tpcb)[Dy2(tpcb)2(H2O)2(4-pyO)2(NO3)2(SCN)6](SCN)2 (tpcb = tetrakis(4-pyridyl)cyclobutane, 4-pyO = 4-(1H)-pyridone) (Figure 8e,f). Subtle changes in the coordination environment around the Dy(III) ions increased the effective energy barrier from 153.8 K (Hdc = 0 Oe) and 201.9 K (Hdc = 1500 Oe) in the mononuclear complex to 205.5 K (Hdc = 0 Oe) and 234.5 K (Hdc = 1500 Oe) in the binuclear complex (Figure 8g,h) [128]. However, in the temperature range dominated by the Raman process, the magnetic relaxation time of the binuclear complex under the 1500 Oe field was significantly shorter compared to the mononuclear complex.
Similarly to the [2,2] cycloaddition of bpe, 9-diethylphosphonomethylanthracene (depma) can undergo a [4,4] cycloaddition reaction upon illumination. Under ultraviolet light, the π bonds of two anthracene rings interacted to form a dimer with an eight-membered ring structure. In 2018, Zheng’s group reported a mononuclear Dy(III) complex, DyIII(depma)(NO3)3(hmpa)2 (hmpa = hexamethylphosphoramide) [28], where the anthracene rings were arranged parallel to each other with an interplanar distance of 3.44 Å. UV irradiation at 365 nm induced a [4 + 4] cycloaddition reaction, forming a binuclear DyIII2(depma2)(NO3)6(hmpa)4 (Figure 9a,b). Under a 500 Oe field, the effective energy barriers of 20.4 K for the mononuclear complex and 43.2 K for the binuclear complex were obtained (Figure 9c,d). Exposure to 254 nm of UV irradiation or thermal treatment reverted the binuclear complex to the mononuclear complex, leading to a reduced effective energy barrier of 18.0 K. This light-driven reversibility highlights the wavelength-dependent switching capabilities.
In 2020, another depma-based mononuclear Dy(III) complex, DyIII(SCN)3(depma)2(4-hpy)2 (1; 4-hpy = 4-hydroxypyridine) [113], was reported, where the anthracene rings were arranged parallel to each other with an interplanar distance of 3.49 Å at 300 K (1RT) and 3.52/3.56 Å at 193 K (1LT). UV irradiation at 365 nm induced a [4 + 4] cycloaddition reaction, leading to a one-dimensional Dy(III) complex 1UV. 1UV reverted to 1 upon heating at 120 °C (Figure 9e). The one-dimensional Dy(III) complex had a different dc magnetic susceptibility compared to the mononuclear complex, and heating at 1UV yielded 1R, which exhibited the same DC magnetic susceptibility as 1 (Figure 9f). Under zero field, the effective energy barriers of 1, 1UV, and 1R were 141, 101, and 153 K, respectively (Figure 9g–i). This work demonstrates the feasibility of switching magnetic properties via light and heat.
Dithienylethene is another photoresponsive ligand. The double bond of ethylene could react with two thiophene groups to form a closed-loop structure under ultraviolet irradiation. Upon exposure to visible light or heating, the closed-loop structure can revert to the open-loop structure, which is a reversible process. In 2020, Bernot’s group reported a one-dimensional Dy(III) complex [Dy(Tppy)F(Lc)]PF6 (1c; Tppy = tris(3-(2-pyridyl)pyrazolyl)hydroborate, L = bispyridyl dithienylethene) with a bispyridyl dithienylethene ligand [29]. Upon irradiation using 532 nm green light, the ligand underwent a transformation from a closed-ring to an open-ring structure, resulting in the formation of the open-ring one-dimensional complex 1o (Figure 10a,b). An in-depth study on the multiple relaxation processes of molecular nanomagnets is crucial for clarifying their dominant mechanisms across different temperature ranges. 1o and 1c exhibited Raman, Orbach, and quantum tunneling relaxation processes with the same effective energy barrier of 225.9 K. Here, τ−1 = τQTM−1 + τ0−1exp(−Ueff/kBT) + BTn was used to fit the relaxation data [19]. 1o and 1c showed the Orbach process at a high temperature range and QTM at a low temperature range. However, 1o (τQTM = 0.0015 s) demonstrated a faster quantum tunneling rate compared to 1c (τQTM = 0.0337 s), and exhibited narrower hysteresis loops (Figure 10c,d). Ab initio calculations indicated that the photomodulation of the magnetic properties originated from the change in crystal packing rather than ligand isomerization-induced crystal field splitting.

4.2. Photomagnetic Properties of Photogenerated Radical-Type Lanthanide Molecular Nanomagnets

Compared to photocyclization-type systems, photogenerated radical families usually exhibit faster response and minimal structural changes. For example, pyridyls and 9-anthracene carboxylic acids are used for lanthanide molecular nanomagnets with photogenerated radicals. If a system employs pyridyl derivatives as electron acceptors and hydroxyethylidene diphosphonate (HEDP) as electron donors, light irradiation induces electron transfer between them, resulting in the generation of stable radicals [30,115]. The aromatic structure could stabilize photogenerated radicals as well. These radicals have the capacity to modulate the coordination field strength of metal centers and/or the magnetic interaction between the metal centers and the radicals.
Wang’s group reported a Dy(III) molecular nanomagnet, [Dy3(H-HEDP)3(H2-HEDP)3]·2H3-TPT·H4-HEDP·10H2O (QDU-1; TPT = 2,4,6-tri(4-pyridyl)-1,3,5-triazine) [30] (Figure 11a). At room temperature and under UV light, colorless QDU-1 transformed into blue QDU-1a, simultaneously generating stable radicals, with a sharp EPR signal at g = 2.0004. Heating at 100 °C for 30 min restored QDU-1 to its colorless state (QDU-1b). Before irradiation, QDU-1 was in a paramagnetic state; after irradiation, the stable radicals induced ferromagnetic interactions with Dy(III) ions at low temperatures, and QDU-1a exhibited SMM behavior (Figure 11b–d), with an effective energy barrier of 108.1 K. Through heating, QDU-1a can be reverted to its initial state. This study represents the observation of room-temperature-reversible photo-thermochromic SMM on/off properties in lanthanide complexes.
In 2021, two photoresponsive lanthanide complexes, [Ln3(H-HEDP)2(H2-HEDP)2(H3-HEDP)2]·H3-TPP·11H2O (Ln = Dy (1-Dy), Tb (2-Tb) and TPP = 2,4,6-tris(4-pyridyl)pyridine) [115], were reported (Figure 11e). At room temperature and under irradiation with a xenon lamp for 40 min, yellow 1-Dy and 2-Tb were transformed to purple 1a-Dy and 2a-Tb with EPR signals at g = 2.014 and 1.996, respectively, from the generation of stable radicals. At low temperatures, antiferromagnetic coupling existed between the lanthanide ions and the photogenerated radicals (Figure 11f). 1-Dy and 2-Tb exhibited SMM behavior, while 1a-Dy and 2a-Tb were paramagnetic (Figure 11g–j). Reversible photo-thermochromic properties and controlled quenching and recovery of SMM behaviors were observed.
The photogenerated crown radicals were proven to be a means to modulate the magnetization of lanthanide complexes. Guo’s group reported a crown-based complex, [DyIII(18C6)(H2O)3]FeIII(CN)6·2H2O (DyFe18C6) [116] (Figure 12a). At room temperature and under UV irradiation, yellow DyFe18C6 transformed into orange DyFe18C6-UV, accompanied by the generation of radicals. Heating at 75 °C for 2 h or storing in darkness for 7 days restored the complex to its initial state. The obtained compound exhibited photochromic and photomagnetic behaviors at room temperature, demonstrating a transition from paramagnetic to ferromagnetic states and a remarkable enhancement of χMT by 20.9% (Figure 12b). DyFe18C6 and its photoproducts had effective energy barriers of 18.4 and 17.0 K, respectively (Figure 12c,d).
Due to the advantages of combining the intrinsic properties of photochromic molecules with those of coordination polymers (CPs)/metal–organic frameworks (MOFs), such as modularity, porosity, and structural adjustability, the application of photoresponsive CPs/MOFs in gas storage and separation has been studied [129,130,131,132,133,134,135]. Since 2013, we have reported a series of SMM-based CPs/MOFs [11,120,121,122,136,137,138,139,140,141,142,143,144,145,146,147,148,149]. However, studies on photoresponsive SMMs in CPs/MOFs remain rare.
In 2023, two Er(III) CPs [Er(CA)1.5(bpy)(DMF)]n (Er(bpy)) and [Er(CA)1.5(phen)(DMF)]n (Er(phen)) (H2CA = 2,5-dichloro-3,6-dihydroxy-p-quinone, bpy = 2,2′-bipyridine, phen = 1,10-phenanthroline) and their light-converted products Er(bpy)-UV and Er(phen)-UV, exhibiting field-induced magnetization dynamics, were reported [120]. The structure of Er(phen) is shown in Figure 13a. The room temperature χMT of Er(phen)-UV increased 25.7% compared with that of Er(phen) (Figure 13b). The EPR signal at g = 2.002 further supported the generation of radicals, influencing the magnetization dynamics. The magnetization dynamics of Er(phen) and Er(phen)-UV were dominated by both Raman and Orbach processes, and Er(phen)-UV showed an enhanced effective energy barrier value compared with that of Er(phen) (Figure 13c,d). It represents an example where both static and dynamic magnetizations of an SMM-based CP were enhanced by UV illumination.
Recently, a SMM-MOF {[Dy1.5(OPh)2Cl(BPy)3(THF)1.5][(BPh4)1.5]·0.5THF}n (Dy(BPy), BPy = 4,4′-bipyridine, OPh = phenoxide, BPh4 = tetraphenylborate) was reported, which features a pentagonal bipyramid SMM as the node and the photosensitive ligand BPy as the linker [121] (Figure 14a–c). The precise synthesis ensured the equatorial arrangement of BPy and the axial coordination of PhO, thereby enhancing the vertical orientation of the main magnetic axis of DyIII ions across all Kagomé layers. Dy(BPy) displayed photochromic properties when irradiated with UV light at room temperature, generating radicals, with an EPR signal at g = 2.0000. In comparison to Dy(BPy), Dy(BPy)-UV demonstrates faster relaxation kinetics in the temperature range of 12–70 K and a lower divergence temperature of 6 K in the field cooling and zero-field cooling curves (9 K for Dy(BPy)). The energy barrier of Dy(BPy)-UV was 641 K, which is lower than that of 822 K of Dy(BPy) in the low-temperature domain. The coercivity field sharply decreases from 4500 Oe for Dy(BPy) to 1300 Oe for Dy(BPy)-UV at 2 K, while the hysteresis loop opening temperature diminishes from 20 K for Dy(BPy) to 14 K for Dy(BPy)-UV(Figure 14d,e). A study of the mechanism revealed that the modulated magnetic properties are attributed to the formation of photogenerated radicals and alterations in crystal packing.
9-anthracene carboxylic acid (HACA) serves as a valuable unit for the construction of molecular switches, attributed to its ability to transform into a photogenerated radical. Stable two-dimensional Yb(III) CPs (Yb(ACA)) and their Y(III)-diluted analog Yb@Y(ACA), along with their photogenerated radical forms, Yb(ACA)-UV and Yb@Y(ACA)-UV, were reported [122]. For Yb(ACA) (Figure 15a,b), the formation of the radical leads to an 82.3% enhancement of the static magnetic susceptibilities of Yb(ACA)-UV at room temperature compared to those of Yb(ACA), supported by high-level calculation results, which are among the highest values reported so far (Figure 15c). Unexpected increases in the effective energy barrier for magnetization reversal of Yb(ACA)-UV (111 K) compared with Yb(ACA) (94 K) were observed. This study not only establishes a strategy for designing photomagnetic materials via SMM-based CP platform but also reveals the influence of photogenerated radicals on the static and dynamic magnetizations of lanthanide CPs.

5. Conclusions and Outlook

The recent advances of molecular nanomagnets with photomagnetic properties are summarized. Researchers have applied various strategies to introduce photoresponsive groups into molecular nanomagnets featuring magnetic bistability. Although molecular nanomagnets with photomagnetic properties have received increasing attention in recent years, this field is still in its early stages. Challenges primarily include two aspects: (i) how to rationally synthesize photoresponsive molecular nanomagnets with modulated static and dynamic magnetic properties and (ii) how to achieve reversible switching of the magnetic properties using light. To date, the photomagnetic properties of transition metal molecular nanomagnets are limited mainly to SCO complexes, and the lanthanide molecular nanomagnets are usually connected with photocyclization ligands and photogenerated radicals. Generally, the cycloaddition reaction is restricted by the crystal lattice, leading to slow or even irreversible reactions; photogenerated radicals may not be highly stable, making it challenging to observe the influence of these radicals on magnetic properties quantitatively.
Currently, the static and dynamic magnetic properties of most molecular nanomagnets induced by light decrease in terms of the energy barrier and hysteresis window, and only a limited number of examples demonstrate reversible capability. In addition to the discrete system, CPs and MOFs are a promising platform for the study of molecular nanomagnets with photomagnetic properties, which have the advantages of high stability and the structural flexibility of metal nodes, organic linkers, and guest molecules. Based on the fast development of coordination chemistry and MOF chemistry, research into photoresponsive magnetic CPs and MOFs may develop at a fast pace in the future.

Author Contributions

Conceptualization, W.S., P.C., and X.G.; writing—original draft preparation, X.G.; writing—review and editing, P.C., W.S., X.G., and X.S.; visualization, X.G. and W.S.; funding acquisition, P.C., W.S., and X.G. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (Grant Nos. 22435002 and 62401563).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Ferreira, F.; Moreira, S.; Zhao, M.; Barriga, E.H. Stretch-induced endogenous electric fields drive directed collective cell migration in vivo. Nat. Mater. 2025, 24, 462–470. [Google Scholar] [CrossRef]
  2. Ghitani, N.; von Buchholtz, L.J.; MacDonald, D.I.; Falgairolle, M.; Nguyen, M.Q.; Licholai, J.A.; Ryba, N.J.P.; Chesler, A.T. A distributed coding logic for thermosensation and inflammatory pain. Nature 2025, 642, 1016–1023. [Google Scholar] [CrossRef] [PubMed]
  3. Li, Z.; Wu, Q.-Q.; Chen, M.-H.; Xu, J.-J.; Chen, H.-Y.; Zhao, W.-W. Organic photoelectrochemical memtransistor. eScience 2025, 5, 100374. [Google Scholar] [CrossRef]
  4. Wang, Y.-X.; Su, D.; Ma, Y.; Sun, Y.; Cheng, P. Electrical detection and modulation of magnetism in a Dy-based ferroelectric single-molecule magnet. Nat. Commun. 2023, 14, 7901. [Google Scholar] [CrossRef]
  5. Wang, Y.-H.; Yan, S.-K.; Liang, S.; Zhao, Y.-R.; Zhang, J.; Wang, G.-M.; Hu, J.-X. Synergistic photomagnetic and photomechanical dynamics in a dysprosium-based smart molecule. Adv. Sci. 2025, e09088. [Google Scholar] [CrossRef] [PubMed]
  6. Raju, M.S.; Paillot, K.; Breslavetz, I.; Novitchi, G.; Rikken, G.L.J.A.; Train, C.; Atzori, M. Optical readout of single-molecule magnets magnetic memories with unpolarized light. J. Am. Chem. Soc. 2024, 146, 23616–23624. [Google Scholar] [CrossRef]
  7. Wang, J.; Zakrzewski, J.J.; Zychowicz, M.; Xin, Y.; Tokoro, H.; Chorazy, S.; Ohkoshi, S. Desolvation-induced highly symmetrical terbium(III) single-molecule magnet exhibiting luminescent self-monitoring of temperature. Angew. Chem. Int. Ed. 2023, 62, e202306372. [Google Scholar] [CrossRef]
  8. Magott, M.; Brzozowska, M.; Baran, S.; Vieru, V.; Pinkowicz, D. An intermetallic molecular nanomagnet with the lanthanide coordinated only by transition metals. Nat. Commun. 2022, 13, 2014. [Google Scholar] [CrossRef]
  9. Gatteschi, D.; Sessoli, R.; Villain, J. Molecular Nanomagnets; Oxford University Press: Oxford, UK, 2006. [Google Scholar]
  10. Rodolphe, C.; Hitoshi, M.; Masahiro, Y.; Claude, C. Evidence for single-chain magnet behavior in a Mn(III)-Ni(II) chain designed with high spin magnetic units: A route to high temperature metastable magnets. J. Am. Chem. Soc. 2002, 124, 12837–12844. [Google Scholar] [CrossRef]
  11. Liu, X.; Feng, X.; Meihaus, K.R.; Meng, X.; Zhang, Y.; Li, L.; Liu, J.L.; Pedersen, K.S.; Keller, L.; Shi, W.; et al. Coercive fields above 6T in two cobalt(II)-radical chain compounds. Angew. Chem. Int. Ed. 2020, 59, 10610–10618. [Google Scholar] [CrossRef]
  12. Wu, Y.; Ruan, Z.-Y.; Zhang, C.; Chen, J.-N.; Wang, X.-Y.; Chen, Z.; Gou, X.; Lan, W.; Kong, X.-J.; Liu, J.-L.; et al. Strong and switchable magneto-optical effect in air-stable chiral DyIII complexes with magnetic anisotropy. J. Am. Chem. Soc. 2025, 147, 20799–20806. [Google Scholar] [CrossRef] [PubMed]
  13. Marin, R.; Brunet, G.; Murugesu, M. Shining new light on multifunctional lanthanide single-molecule magnets. Angew. Chem. Int. Ed. 2021, 60, 1728–1746. [Google Scholar] [CrossRef] [PubMed]
  14. Gou, X.; Wu, Y.; Wang, M.; Liu, N.; Lan, W.; Zhang, Y.-Q.; Shi, W.; Cheng, P. The influence of light on the field-induced magnetization dynamics of two Er(III) coordination polymers with different halogen substituents. Dalton Trans. 2024, 53, 148–152. [Google Scholar] [CrossRef]
  15. Zhao, L.; Meng, Y.S.; Liu, Q.; Sato, O.; Shi, Q.; Oshio, H.; Liu, T. Switching the magnetic hysteresis of an [Fe(II)-NC-W(V)]-based coordination polymer by photoinduced reversible spin crossover. Nat. Chem. 2021, 13, 698–704. [Google Scholar] [CrossRef]
  16. Errulat, D.; Harriman, K.L.M.; Gálico, D.A.; Kitos, A.A.; Mansikkamäki, A.; Murugesu, M. A trivalent 4f complex with two bis-silylamide ligands displaying slow magnetic relaxation. Nat. Chem. 2023, 15, 1100–1107. [Google Scholar] [CrossRef]
  17. Meihaus, K.R.; Minasian, S.G.; Lukens, W.W., Jr.; Kozimor, S.A.; Shuh, D.K.; Tyliszczak, T.; Long, J.R. Influence of pyrazolate vs N-heterocyclic carbene ligands on the slow magnetic relaxation of homoleptic trischelate lanthanide(III) and uranium(III) complexes. J. Am. Chem. Soc. 2014, 136, 6056–6068. [Google Scholar] [CrossRef]
  18. Gao, C.; Genoni, A.; Gao, S.; Jiang, S.; Soncini, A.; Overgaard, J. Observation of the asphericity of 4f-electron density and its relation to the magnetic anisotropy axis in single-molecule magnets. Nat. Chem. 2019, 12, 213–219. [Google Scholar] [CrossRef]
  19. Guo, F.S.; Day, B.M.; Chen, Y.C.; Tong, M.L.; Mansikkamäki, A.; Layfield, R.A. Magnetic hysteresis up to 80 kelvin in a dysprosium metallocene single-molecule magnet. Science 2018, 362, 1400–1403. [Google Scholar] [CrossRef]
  20. Goodwin, C.A.P.; Ortu, F.; Reta, D.; Chilton, N.F.; Mills, D.P. Molecular magnetic hysteresis at 60 kelvin in dysprosocenium. Nature 2017, 548, 439–442. [Google Scholar] [CrossRef] [PubMed]
  21. Li, L.; Fang, Y.; Liu, S.; Hu, M.; Wang, W. Slow magnetic relaxation in a 3D dysprosium(III)-fluoro-oxalate framework containing zig-zag [Dy-F] chains. J. Rare Earths 2023, 41, 100–107. [Google Scholar] [CrossRef]
  22. Liao, X.; Chen, Y.; Xie, T.; Sun, R.J.; Yang, L.Z.; Liu, C.F.; Wang, R.; Klyatskaya, S.; Ruben, M.; Zhang, W.; et al. Spatially dependent f-pi exchange interaction within a single-molecule magnet TbPc2. Nat. Commun. 2025, 16, 6263. [Google Scholar] [CrossRef]
  23. Liu, Y.; Luo, Q.; Jin, P.; Zhai, Y.; Zheng, Y. Fine tuning dynamic magnetism of dysprosiacarboranyl sandwiches. J. Rare Earths 2025, 43, 1205–1212. [Google Scholar] [CrossRef]
  24. Khariushin, I.V.; Ovsyannikov, A.S.; Islamov, D.R.; Samigullina, A.I.; Solovieva, S.E.; Zakrzewski, J.J.; Chorazy, S.; Ferlay, S. Tuning crystal packing and magnetic properties in a series of [Dy12] metallocubanes based on azobenzene derivatives of salicylic acid. Inorg. Chem. 2023, 62, 10548–10558. [Google Scholar] [CrossRef] [PubMed]
  25. Li, G.-M.; Xu, F.; Han, S.-D.; Pan, J.; Wang, G.-M. Hybrid photochromic lanthanide phosphonate with multiple photoresponsive functionalities. Inorg. Chem. 2022, 61, 8379–8385. [Google Scholar] [CrossRef]
  26. Bartolomé, E.; Arauzo, A.; Luzón, J.; Gasque, L. Light-induced magnetic switching in a coumarin-based Tb single molecule magnet. J. Mater. Chem. C 2025, 13, 831–841. [Google Scholar] [CrossRef]
  27. Wang, L.-F.; Qiu, J.-Z.; Liu, J.-L.; Chen, Y.-C.; Jia, J.-H.; Jover, J.; Ruiz, E.; Tong, M.-L. Modulation of single-molecule magnet behaviour via photochemical [2 + 2] cycloaddition. Chem. Commun. 2015, 51, 15358–15361. [Google Scholar] [CrossRef]
  28. Huang, X.-D.; Xu, Y.; Fan, K.; Bao, S.-S.; Kurmoo, M.; Zheng, L.-M. Reversible SC-SC Transformation involving [4 + 4] cycloaddition of anthracene: A single-ion to single-molecule magnet and yellow-green to blue-white emission. Angew. Chem. Int. Ed. 2018, 57, 8577–8581. [Google Scholar] [CrossRef]
  29. Hojorat, M.; Al Sabea, H.; Norel, L.; Bernot, K.; Roisnel, T.; Gendron, F.; Guennic, B.L.; Trzop, E.; Collet, E.; Long, J.R.; et al. Hysteresis photomodulation via single-crystal-to-single-crystal isomerization of a photochromic chain of dysprosium single-molecule magnets. J. Am. Chem. Soc. 2020, 142, 931–936. [Google Scholar] [CrossRef] [PubMed]
  30. Ma, Y.-J.; Hu, J.-X.; Han, S.-D.; Pan, J.; Li, J.-H.; Wang, G.-M. Manipulating on/off single-molecule magnet behavior in a Dy(III)-based photochromic complex. J. Am. Chem. Soc. 2020, 142, 2682–2689. [Google Scholar] [CrossRef]
  31. Feng, X.; Mathoniere, C.; Jeon Ie, R.; Rouzieres, M.; Ozarowski, A.; Aubrey, M.L.; Gonzalez, M.I.; Clerac, R.; Long, J.R. Tristability in a light-actuated single-molecule magnet. J. Am. Chem. Soc. 2013, 135, 15880–15884. [Google Scholar] [CrossRef] [PubMed]
  32. Mathonière, C.; Lin, H.J.; Siretanu, D.; Clerac, R.; Smith, J.M. Photoinduced single-molecule magnet properties in a four-coordinate iron(II) spin crossover complex. J. Am. Chem. Soc. 2013, 135, 19083–19086. [Google Scholar] [CrossRef]
  33. Liu, T.; Zheng, H.; Kang, S.; Shiota, Y.; Hayami, S.; Mito, M.; Sato, O.; Yoshizawa, K.; Kanegawa, S.; Duan, C. A light-induced spin crossover actuated single-chain magnet. Nat. Commun. 2013, 4, 2826. [Google Scholar] [CrossRef]
  34. Jiang, W.; Jiao, C.; Meng, Y.; Zhao, L.; Liu, Q.; Liu, T. Switching single chain magnet behavior via photoinduced bidirectional metal-to-metal charge transfer. Chem. Sci. 2018, 9, 617–622. [Google Scholar] [CrossRef]
  35. Zhou, X.; Yan, F.; Lyubartsev, A.; Shen, B.; Zhai, J.; Conesa, J.C.; Hedin, N. Efficient production of solar hydrogen peroxide using piezoelectric polarization and photoinduced charge transfer of nanopiezoelectrics sensitized by carbon quantum dots. Adv. Sci. 2022, 9, 2105792. [Google Scholar] [CrossRef] [PubMed]
  36. Homer, M.K.; Kuo, D.-Y.; Dou, F.Y.; Cossairt, B.M. Photoinduced charge transfer from quantum dots measured by cyclic voltammetry. J. Am. Chem. Soc. 2022, 144, 14226–14234. [Google Scholar] [CrossRef] [PubMed]
  37. Lu, Z.; Wu, X.; Chen, N.; Cao, M.; Sartin, M.M.; Ren, B. Photoinduced charge transfer from a semiconductor to a metal probed at the single-nanoparticle level. ACS Energy Lett. 2021, 6, 3473–3480. [Google Scholar] [CrossRef]
  38. Li, X.; Yu, J.; Lu, Z.; Duan, J.; Fry, H.C.; Gosztola, D.J.; Maindan, K.; Rajasree, S.S.; Deria, P. Photoinduced charge transfer with a small driving force facilitated by exciplex-like complex formation in metal-organic frameworks. J. Am. Chem. Soc. 2021, 143, 15286–15297. [Google Scholar] [CrossRef]
  39. Cammarata, M.; Zerdane, S.; Balducci, L.; Azzolina, G.; Mazerat, S.; Exertier, C.; Trabuco, M.; Levantino, M.; Alonso-Mori, R.; Glownia, J.M.; et al. Charge transfer driven by ultrafast spin transition in a CoFe Prussian blue analogue. Nat. Chem. 2020, 13, 10–14. [Google Scholar] [CrossRef]
  40. Ahmad, I.; Akhtar, M.S.; Ahmed, E.; Ahmad, M. Aluminium and cerium co-doped ZnO nanoparticles: Facile and inexpensive synthesis and visible light photocatalytic performances. J. Rare Earths 2021, 39, 151–159. [Google Scholar] [CrossRef]
  41. Ishikawa, T.; Hayes, S.A.; Keskin, S.; Corthey, G.; Hada, M.; Pichugin, K.; Marx, A.; Hirscht, J.; Shionuma, K.; Onda, K.; et al. Direct observation of collective modes coupled to molecular orbital-driven charge transfer. Science 2015, 350, 1501–1505. [Google Scholar] [CrossRef]
  42. Kaneva, N.; Bachvarova-Nedelcheva, A. Photocatalytic efficiency of pure and palladium Co-catalytic modified binary system. Inorganics 2025, 13, 161. [Google Scholar] [CrossRef]
  43. Zoppellaro, G.; Medved, M.; Hrubý, V.; Zbořil, R.; Otyepka, M.; Lazar, P. Solvent controlled generation of spin active polarons in two-dimensional material under UV light irradiation. J. Am. Chem. Soc. 2024, 146, 15010–15018. [Google Scholar] [CrossRef] [PubMed]
  44. Widel, Z.X.W.; Alatis, J.A.; Stephenson, R.H.; Mastrocinque, F.; Wilcox, A.C.; Bullard, G.; Olivier, J.-H.; Bai, Y.; Zhang, P.; Beratan, D.N.; et al. Fluence-dependent photoinduced charge transfer dynamics in polymer-wrapped semiconducting single-walled carbon nanotubes. J. Am. Chem. Soc. 2024, 146, 31169–31176. [Google Scholar] [CrossRef] [PubMed]
  45. Shu, A.; Qin, C.; Li, M.; Zhao, L.; Shangguan, Z.; Shu, Z.; Yuan, X.; Zhu, M.; Wu, Y.; Wang, H. Electric effects reinforce charge carrier behaviour for photocatalysis. Energy Environ. Sci. 2024, 17, 4907–4928. [Google Scholar] [CrossRef]
  46. Royakkers, J.; Yang, H.; Gillett, A.J.; Eisner, F.; Ghosh, P.; Congrave, D.G.; Azzouzi, M.; Andaji-Garmaroudi, Z.; Leventis, A.; Rao, A.; et al. Synthesis of model heterojunction interfaces reveals molecular-configuration-dependent photoinduced charge transfer. Nat. Chem. 2024, 16, 1453–1461. [Google Scholar] [CrossRef] [PubMed]
  47. Pearce, N.; Reynolds, K.E.A.; Kayal, S.; Sun, X.Z.; Davies, E.S.; Malagreca, F.; Schürmann, C.J.; Ito, S.; Yamano, A.; Argent, S.P.; et al. Selective photoinduced charge separation in perylenediimide-pillar[5]arene rotaxanes. Nat. Commun. 2022, 13, 415. [Google Scholar] [CrossRef]
  48. Zheng, F.; Dong, F.; Zhou, L.; Yu, J.; Luo, X.; Zhang, X.; Lv, Z.; Jiang, L.; Chen, Y.; Liu, M. Cerium and carbon-sulfur codoped mesoporous TiO2 nanocomposites for boosting visible light photocatalytic activity. J. Rare Earths 2023, 41, 539–549. [Google Scholar] [CrossRef]
  49. Pajuelo-Corral, O.; García, J.A.; Castillo, O.; Luque, A.; Rodríguez-Diéguez, A.; Cepeda, J. Single-ion magnet and photoluminescence properties of lanthanide(III) coordination polymers based on pyrimidine-4,6-dicarboxylate. Magnetochemistry 2021, 7, 8. [Google Scholar] [CrossRef]
  50. Fu, C.; Li, F.; Zhang, J.; Li, D.; Qian, K.; Liu, Y.; Tang, J.; Fan, F.; Zhang, Q.; Gong, X.Q.; et al. Site sensitivity of interfacial charge transfer and photocatalytic efficiency in photocatalysis: Methanol oxidation on anatase TiO2 nanocrystals. Angew. Chem. Int. Ed. 2021, 60, 6160–6169. [Google Scholar] [CrossRef] [PubMed]
  51. Graf, M.; Vonbun-Feldbauer, G.B.; Koper, M.T.M. Direct and broadband plasmonic charge transfer to enhance water oxidation on a gold electrode. ACS Nano 2021, 15, 3188–3200. [Google Scholar] [CrossRef]
  52. Wang, M.; Zhang, G.; Guan, Z.; Yang, J.; Li, Q. Spatially separating redox centers and photothermal effect synergistically boosting the photocatalytic hydrogen evolution of ZnIn2S4 Nanosheets. Small 2021, 17, 2006952. [Google Scholar] [CrossRef]
  53. Delgado, T.; Pelé, A.-L. Measurement of reverse-light-induced excited spin state trapping in spin crossover systems: A study case with Zn1−xFex(6-mepy)3tren(PF6)2·CH3CN; x = 0.5%. Crystals 2024, 14, 210. [Google Scholar] [CrossRef]
  54. Zhu, H.-L.; Meng, Y.-S.; Hu, J.-X.; Oshio, H.; Liu, T. Photoinduced magnetic hysteresis in a cyanide-bridged two-dimensional [Mn2W] coordination polymer. Inorg. Chem. Front. 2022, 9, 4974–4981. [Google Scholar] [CrossRef]
  55. Su, S.-Q.; Wu, S.-Q.; Huang, Y.-B.; Xu, W.-H.; Gao, K.-G.; Okazawa, A.; Okajima, H.; Sakamoto, A.; Kanegawa, S.; Sato, O. Photoinduced persistent polarization change in a spin transition crystal. Angew. Chem. Int. Ed. 2022, 61, e202208771. [Google Scholar] [CrossRef] [PubMed]
  56. Nadeem, M.; Cruddas, J.; Ruzzi, G.; Powell, B.J. Toward high-temperature light-induced spin-state trapping in spin-crossover materials: The interplay of collective and molecular effects. J. Am. Chem. Soc. 2022, 144, 9138–9148. [Google Scholar] [CrossRef] [PubMed]
  57. Benchohra, A.; Li, Y.; Chamoreau, L.M.; Baptiste, B.; Elkaïm, E.; Guillou, N.; Kreher, D.; Lescouëzec, R. The atypical hysteresis of [Fe(C6F5Tp)2]: Overlay of spin-crossovers and symmetry-breaking phase transition. Angew. Chem. Int. Ed. 2021, 60, 8803–8807. [Google Scholar] [CrossRef]
  58. Wen, W.; Meng, Y.-S.; Jiao, C.-Q.; Liu, Q.; Zhu, H.-L.; Li, Y.-M.; Oshio, H.; Liu, T. A mixed-valence {Fe13} cluster exhibiting metal-to-metal charge-transfer-switched spin crossover. Angew. Chem. Int. Ed. 2020, 59, 16393–16397. [Google Scholar] [CrossRef]
  59. Ossinger, S.; Näther, C.; Buchholz, A.; Schmidtmann, M.; Mangelsen, S.; Beckhaus, R.; Plass, W.; Tuczek, F. Spin transition of an iron(II) organoborate complex in different polymorphs and in vacuum-deposited thin films: Influence of cooperativity. Inorg. Chem. 2020, 59, 7966–7979. [Google Scholar] [CrossRef] [PubMed]
  60. Zhang, K.; Ash, R.; Girolami, G.S.; Vura-Weis, J. Tracking the metal-centered triplet in photoinduced spin crossover of fe(phen)32+ with tabletop femtosecond M-Edge X-ray absorption Near-Edge structure spectroscopy. J. Am. Chem. Soc. 2019, 141, 17180–17188. [Google Scholar] [CrossRef]
  61. Fatur, S.M.; Shepard, S.G.; Higgins, R.F.; Shores, M.P.; Damrauer, N.H. A synthetically tunable system to control MLCT excited-state lifetimes and spin states in iron(II) polypyridines. J. Am. Chem. Soc. 2017, 139, 4493–4505. [Google Scholar] [CrossRef]
  62. Wu, D.-Y.; Sato, O.; Einaga, Y.; Duan, C.-Y. A spin-crossover cluster of iron(II) exhibiting a mixed-spin structure and synergy between spin transition and magnetic interaction. Angew. Chem. Int. Ed. 2009, 48, 1475–1478. [Google Scholar] [CrossRef]
  63. Halder, G.J.; Chapman, K.W.; Neville, S.M.; Moubaraki, B.; Murray, K.S.; Létard, J.F.; Kepert, C.J. Elucidating the mechanism of a two-step spin transition in a nanoporous metal-organic framework. J. Am. Chem. Soc. 2009, 130, 17552–17562. [Google Scholar] [CrossRef] [PubMed]
  64. Agustí, G.; Ohtani, R.; Yoneda, K.; Gaspar, A.B.; Ohba, M.; Sánchez-Royo, J.F.; Muñoz, M.C.; Kitagawa, S.; Real, J.A. Oxidative addition of halogens on open metal sites in a microporous spin-crossover coordination polymer. Angew. Chem. Int. Ed. 2009, 48, 8944–8947. [Google Scholar] [CrossRef] [PubMed]
  65. Chen, T.; Yang, Q.; Fang, C.; Deng, S.; Xu, B. Advanced design for stimuli-reversible chromic wearables with customizable functionalities. Adv. Mater. 2024, 37, 2413665. [Google Scholar] [CrossRef]
  66. He, Y.; Dang, T.; Leach, A.G.; Zhang, Z.-Y.; Li, T. Photoswitchable azobispyrazole crystals achieving near-quantitative crystalline-state bidirectional EZ conversions. J. Am. Chem. Soc. 2024, 146, 29237–29244. [Google Scholar] [CrossRef]
  67. Zhao, H.; Liao, J.; Fu, S.; Zi, Y.; Bai, X.; Ci, Y.; Zhang, Y.; Cai, X.; Li, Y.; Cun, Y.; et al. Direct 3D lithography of reversible photochromic patterns with tunable luminescence in amorphous transparent media. ACS Energy Lett. 2025, 10, 1235–1244. [Google Scholar] [CrossRef]
  68. Dong, G.; Lin, H.; Yang, H.; Hu, T.; Huang, Q.; Han, X.; Li, X.; Wang, P.; Xu, J.; Cheng, Y.; et al. Twin-perovskite structured Ba8Ti3Nb4O24: Pr3+: A fast photochromics with millisecond-level response speed for optical storage. ACS Energy Lett. 2024, 10, 484–495. [Google Scholar] [CrossRef]
  69. Hou, J.; Long, G.; Zhao, W.; Zhou, G.; Liu, D.; Broer, D.J.; Feringa, B.L.; Chen, J. Phototriggered complex motion by programmable construction of light-driven molecular motors in liquid crystal networks. J. Am. Chem. Soc. 2022, 144, 6851–6860. [Google Scholar] [CrossRef]
  70. Song, Y.-P.; Zhang, J.-N.; Wang, J.-R.; Li, K.; Yuan, Y.-X.; Li, B.; Zang, S.-Q. A fast responsive photochromic SCC-MOF for photoswitching and information encryption. Sci. China Mater. 2024, 67, 698–704. [Google Scholar] [CrossRef]
  71. Lu, H.; Xie, J.; Wang, X.-Y.; Wang, Y.; Li, Z.-J.; Diefenbach, K.; Pan, Q.-J.; Qian, Y.; Wang, J.-Q.; Wang, S.; et al. Visible colorimetric dosimetry of UV and ionizing radiations by a dual-module photochromic nanocluster. Nat. Commun. 2021, 12, 2798. [Google Scholar] [CrossRef]
  72. Roubinet, B.; Weber, M.; Shojaei, H.; Bates, M.; Bossi, M.L.; Belov, V.N.; Irie, M.; Hell, S.W. Fluorescent photoswitchable diarylethenes for biolabeling and single-molecule localization microscopies with optical superresolution. J. Am. Chem. Soc. 2017, 139, 6611–6620. [Google Scholar] [CrossRef]
  73. Zhang, X.; Hou, L.; Samorì, P. Coupling carbon nanomaterials with photochromic molecules for the generation of optically responsive materials. Nat. Commun. 2016, 7, 11118. [Google Scholar] [CrossRef] [PubMed]
  74. Babii, O.; Afonin, S.; Garmanchuk, L.V.; Nikulina, V.V.; Nikolaienko, T.V.; Storozhuk, O.V.; Shelest, D.V.; Dasyukevich, O.I.; Ostapchenko, L.I.; Iurchenko, V.; et al. Direct photocontrol of peptidomimetics: An alternative to oxygen-dependent photodynamic cancer therapy. Angew. Chem. Int. Ed. 2016, 55, 5493–5496. [Google Scholar] [CrossRef] [PubMed]
  75. Rice, A.M.; Martin, C.R.; Galitskiy, V.A.; Berseneva, A.A.; Leith, G.A.; Shustova, N.B. Photophysics modulation in photoswitchable metal-organic frameworks. Chem. Rev. 2019, 120, 8790–8813. [Google Scholar] [CrossRef]
  76. Bisoyi, H.K.; Li, Q. Light-directing chiral liquid crystal nanostructures: From 1D to 3D. Acc. Chem. Res. 2014, 47, 3184–3195. [Google Scholar] [CrossRef]
  77. Kole, G.K.; Vittal, J.J. Solid-state reactivity and structural transformations involving coordination polymers. Chem. Soc. Rev. 2013, 42, 1755–1775. [Google Scholar] [CrossRef]
  78. Cao, L.-H.; Xu, X.-Q.; Tang, X.-H.; Yang, Y.; Liu, J.; Yin, Z.; Zang, S.-Q.; Ma, Y.-M. Controllable strategy for metal-organic framework light-driven [2 + 2] cycloaddition reactions via solvent-assisted linker exchange. Inorg. Chem. 2021, 60, 2117–2121. [Google Scholar] [CrossRef]
  79. Ou, Y.-C.; Liu, W.-T.; Li, J.-Y.; Zhang, G.-G.; Wang, J.; Tong, M.-L. Solvochromic and photodimerization behaviour of 1D coordination polymer via single-crystal-to-single-crystal transformation. Chem. Commun. 2011, 47, 9384–9386. [Google Scholar] [CrossRef]
  80. Wang, X.-Y.; Wang, Z.-M.; Gao, S. A pillared layer MOF with anion-tunable magnetic properties and photochemical [2 + 2] cycloaddition. Chem. Commun. 2007, 1127–1129. [Google Scholar] [CrossRef] [PubMed]
  81. Hutchins, K.M.; Rupasinghe, T.P.; Ditzler, L.R.; Swenson, D.C.; Sander, J.R.G.; Baltrusaitis, J.; Tivanski, A.V.; MacGillivray, L.R. Nanocrystals of a metal-organic complex exhibit remarkably high conductivity that increases in a single-crystal-to-single-crystal transformation. J. Am. Chem. Soc. 2014, 136, 6778–6781. [Google Scholar] [CrossRef]
  82. Park, I.H.; Medishetty, R.; Kim, J.Y.; Lee, S.S.; Vittal, J.J. Distortional supramolecular isomers of polyrotaxane coordination polymers: Photoreactivity and sensing of nitro compounds. Angew. Chem. Int. Ed. 2014, 53, 5591–5595. [Google Scholar] [CrossRef]
  83. Zouev, I.; Cao, D.K.; Sreevidya, T.V.; Botoshansky, M.; Kaftory, M. Photodimerization of anthracene derivatives in their neat solid state and in solid molecular compounds. CrystEngComm 2011, 13, 4376–4381. [Google Scholar] [CrossRef]
  84. Becker, H.D. Unimolecular photochemistry of anthracenes. Chem. Rev. 1993, 93, 145–172. [Google Scholar] [CrossRef]
  85. Furlong, B.J.; Katz, M.J. Bistable dithienylethene-based metal-organic framework illustrating optically induced changes in chemical separations. J. Am. Chem. Soc. 2017, 139, 13280–13283. [Google Scholar] [CrossRef]
  86. Kobatake, S.; Takami, S.; Muto, H.; Ishikawa, T.; Irie, M. Rapid and reversible shape changes of molecular crystals on photoirradiation. Nature 2007, 446, 778–781. [Google Scholar] [CrossRef] [PubMed]
  87. Yang, X.; Du, M.; Chu, Z.; Li, C. Synchronizing multicolor changes and shape deformation into structurally homogeneous hydrogels via a single photochromophore. Adv. Mater. 2025, 37, 2500857. [Google Scholar] [CrossRef]
  88. Williams, D.E.; Martin, C.R.; Dolgopolova, E.A.; Swifton, A.; Godfrey, D.C.; Ejegbavwo, O.A.; Pellechia, P.J.; Smith, M.D.; Shustova, N.B. Flipping the switch: Fast photoisomerization in a confined environment. J. Am. Chem. Soc. 2018, 140, 7611–7622. [Google Scholar] [CrossRef] [PubMed]
  89. Du, Y.; Liao, W.-Q.; Li, Y.; Huang, C.-R.; Gan, T.; Chen, X.-G.; Lv, H.-P.; Song, X.-J.; Xiong, R.-G.; Wang, Z.-X. A homochiral fulgide organic ferroelectric crystal with photoinduced molecular orbital breaking. Angew. Chem. Int. Ed. 2023, 62, e202315189. [Google Scholar] [CrossRef]
  90. Du, Y.; Huang, C.-R.; Xu, Z.-K.; Hu, W.; Li, P.-F.; Xiong, R.-G.; Wang, Z.-X. Photochromic single-component organic fulgide ferroelectric with photo-triggered polarization response. JACS Au 2023, 3, 1464–1471. [Google Scholar] [CrossRef] [PubMed]
  91. Wu, Y.; Schneider, S.; Yuan, Y.; Young, R.M.; Francese, T.; Mansoor, I.F.; Dudenas, P.J.; Lei, Y.; Gomez, E.D.; DeLongchamp, D.M.; et al. Twisted A-D-A type acceptors with thermally-activated delayed crystallization behavior for efficient nonfullerene organic solar cells. Adv. Energy Mater. 2022, 12, 2103957. [Google Scholar] [CrossRef]
  92. Pugliese, E.; Vo, N.T.; Boussac, A.; Banse, F.; Mekmouche, Y.; Simaan, J.; Tron, T.; Gotico, P.; Sircoglou, M.; Halime, Z.; et al. Photocatalytic generation of a non-heme Fe(III)-hydroperoxo species with O2 in water for the oxygen atom transfer reaction. Chem. Sci. 2022, 13, 12332–12339. [Google Scholar] [CrossRef]
  93. Hein, R.; Docker, A.; Davis, J.J.; Beer, P.D. Redox-switchable chalcogen bonding for anion recognition and sensing. J. Am. Chem. Soc. 2022, 144, 8827–8836. [Google Scholar] [CrossRef]
  94. Min, H.; Han, Z.; Sun, T.; Wang, K.; Xu, J.; Yao, P.; Yang, S.; Cheng, P.; Shi, W. Dynamic-static coupled sensing of trace biomarkers by molecularly imprinted metal-organic frameworks. Sci. China Chem. 2023, 66, 3511–3517. [Google Scholar] [CrossRef]
  95. Han, Z.; Wang, K.; Chen, Y.; Li, J.; Teat, S.J.; Yang, S.; Shi, W.; Cheng, P. A multicenter metal-organic framework for quantitative detection of multicomponent organic mixtures. CCS Chem. 2022, 4, 3238–3245. [Google Scholar] [CrossRef]
  96. Burton, S.T.; Lee, G.; Moore, C.E.; Sevov, C.S.; Turro, C. Cyclometallated Co(III) complexes with lowest-energy charge transfer excited states accessible with visible light. J. Am. Chem. Soc. 2025, 147, 13315–13327. [Google Scholar] [CrossRef]
  97. Martin, C.R.; Kittikhunnatham, P.; Leith, G.A.; Berseneva, A.A.; Park, K.C.; Greytak, A.B.; Shustova, N.B. Let the light be a guide: Chromophore communication in metal-organic frameworks. Nano Res. 2020, 14, 338–354. [Google Scholar] [CrossRef]
  98. Wang, S.; Sun, H.; Erdmann, T.; Wang, G.; Fazzi, D.; Lappan, U.; Puttisong, Y.; Chen, Z.; Berggren, M.; Crispin, X.; et al. A chemically doped naphthalenediimide-bithiazole polymer for n-type organic thermoelectrics. Adv. Mater. 2018, 30, 1801898. [Google Scholar] [CrossRef]
  99. Xiong, Z.; Li, Y.; Yuan, Z.; Liang, J.; Wang, S.; Yang, X.; Xiang, S.; Lv, Y.; Chen, B.; Zhang, Z. Switchable anisotropic/isotropic photon transport in a double-dipole metal-organic framework via radical-controlled energy transfer. Adv. Mater. 2024, 36, 2314005. [Google Scholar] [CrossRef]
  100. Liao, P.Y.; Li, J.X.; Liu, J.C.; Xiong, Q.; Ruan, Z.Y.; Li, T.; Deng, W.; Jiang, S.D.; Jia, J.H.; Tong, M.L. Radical-induced photochromic silver(I) metal-organic frameworks: Alternative topology, dynamic photoluminescence and efficient photothermal conversion modulated by anionic guests. Angew. Chem. Int. Ed. 2024, 63, e202401448. [Google Scholar] [CrossRef]
  101. Gong, T.; Sui, Q.; Li, P.; Meng, X.F.; Zhou, L.J.; Chen, J.; Xu, J.; Wang, L.; Gao, E.-Q. Versatile and switchable responsive properties of a lanthanide-viologen metal-organic framework. Small 2019, 15, 1803468. [Google Scholar] [CrossRef] [PubMed]
  102. Liu, Q.; Hu, J.-X.; Meng, Y.-S.; Jiang, W.-J.; Wang, J.-L.; Wen, W.; Wu, Q.; Zhu, H.-L.; Zhao, L.; Liu, T. Asymmetric coordination toward a photoinduced single-chain magnet showing high coercivity values. Angew. Chem. Int. Ed. 2021, 60, 10537–10541. [Google Scholar] [CrossRef]
  103. Hu, J.-X.; Zhu, H.-L.; Meng, Y.-S.; Pang, J.; Li, N.; Liu, T.; Bu, X.-H. Ligand modified and light switched on/off single-chain magnets of {Fe2Co} coordination polymers via metal-to-metal charge transfer. CCS Chem. 2023, 5, 865–875. [Google Scholar] [CrossRef]
  104. Morimoto, M.; Miyasaka, H.; Yamashita, M.; Irie, M. Coordination assemblies of [Mn4] single-molecule magnets linked by photochromic ligands: Photochemical control of the magnetic properties. J. Am. Chem. Soc. 2009, 131, 9823–9835. [Google Scholar] [CrossRef] [PubMed]
  105. Fetoh, A.; Cosquer, G.; Morimoto, M.; Irie, M.; El-Gammal, O.; Abu El-Reash, G.M.; Breedlove, B.K.; Yamashita, M. Synthesis, structures, and magnetic properties of two coordination assemblies of Mn(III) single molecule magnets bridged via photochromic diarylethene ligands. Inorg. Chem. 2019, 58, 2307–2314. [Google Scholar] [CrossRef]
  106. Jiao, T.; Wu, C.-H.; Zhang, Y.-S.; Miao, X.; Wu, S.; Jiang, S.-D.; Wu, J. Solution-phase synthesis of Clar’s goblet and elucidation of its spin properties. Nat. Chem. 2025, 17, 924–932. [Google Scholar] [CrossRef]
  107. Eaton, S.S.; Yamabayashi, T.; Horii, Y.; Yamashita, M.; Eaton, G.R. Anisotropy of spin-lattice relaxation time (T1) for oxo-vanadium(IV) and nitrido chromium(V) porphyrins. J. Am. Chem. Soc. 2025, 147, 13815–13823. [Google Scholar] [CrossRef] [PubMed]
  108. Graham, M.J.; Krzyaniak, M.D.; Wasielewski, M.R.; Freedman, D.E. Probing nuclear spin effects on electronic spin coherence via EPR measurements of vanadium(IV) complexes. Inorg. Chem. 2017, 56, 8106–8113. [Google Scholar] [CrossRef]
  109. Tesi, L.; Lucaccini, E.; Cimatti, I.; Perfetti, M.; Mannini, M.; Atzori, M.; Morra, E.; Chiesa, M.; Caneschi, A.; Sorace, L.; et al. Quantum coherence in a processable vanadyl complex: New tools for the search of molecular spin qubits. Chem. Sci. 2016, 7, 2074–2083. [Google Scholar] [CrossRef] [PubMed]
  110. Yao, N.-T.; Zhao, L.; Yi, C.; Liu, Q.; Li, Y.-M.; Meng, Y.-S.; Oshio, H.; Liu, T. Manipulating selective metal-to-metal electron transfer to achieve multi-phase transitions in an asymmetric [Fe2Co]-assembled mixed-valence chain. Angew. Chem. Int. Ed. 2022, 61, e202115367. [Google Scholar] [CrossRef]
  111. Shang, M.-J.; Lu, H.-H.; Liu, Q.; Zhou, R.-H.; Sun, H.-Y.; Shao, Z.; Zhao, L.; Meng, Y.-S.; Liu, T. Reversible on–off switching of a single-chain magnet via single-crystal-to-single-crystal transition and light-induced metal-to-metal electron transfer. Inorg. Chem. Front. 2024, 11, 8704–8714. [Google Scholar] [CrossRef]
  112. Wang, L.-F.; Qiu, J.-Z.; Chen, Y.-C.; Liu, J.-L.; Li, Q.-W.; Jia, J.-H.; Tong, M.-L. [2 + 2] Photochemical modulation of the Dy(III) single-molecule magnet: Opposite influence on the energy barrier and relaxation time. Inorg. Chem. Front. 2017, 4, 1311–1318. [Google Scholar] [CrossRef]
  113. Huang, X.-D.; Wen, G.-H.; Bao, S.-S.; Jia, J.-G.; Zheng, L.-M. Thermo- and light-triggered reversible interconversion of dysprosium-anthracene complexes and their responsive optical, magnetic and dielectric properties. Chem. Sci. 2021, 12, 929–937. [Google Scholar] [CrossRef]
  114. Li, J.; Kong, M.; Yin, L.; Zhang, J.; Yu, F.; Ouyang, Z.-W.; Wang, Z.; Zhang, Y.-Q.; Song, Y. Photochemically tuned magnetic properties in an erbium(III)-based easy-plane single-molecule magnet. Inorg. Chem. 2019, 58, 14440–14448. [Google Scholar] [CrossRef]
  115. Zhang, Q.; Han, S.-D.; Li, Q.; Hu, J.-X.; Wang, G.-M. Electron transfer photochromism of Ln-based (Ln = Dy, Tb) coordinated polymers for reversibly switching off/on single-molecule magnetic behavior. Sci. China Mater. 2022, 65, 788–794. [Google Scholar] [CrossRef]
  116. Cai, L.-Z.; Guo, P.-Y.; Wang, M.-S.; Guo, G.-C. Photoinduced magnetic phase transition and remarkable enhancement of magnetization for a photochromic single-molecule magnet. J. Mater. Chem. C 2021, 9, 2231–2235. [Google Scholar] [CrossRef]
  117. Tian, H.; Su, J.-B.; Bao, S.-S.; Kurmoo, M.; Huang, X.-D.; Zhang, Y.-Q.; Zheng, L.-M. Reversible ON-OFF switching of single-molecule-magnetism associated with single-crystal-to-single-crystal structural transformation of a decanuclear dysprosium phosphonate. Chem. Sci. 2018, 9, 6424–6433. [Google Scholar] [CrossRef] [PubMed]
  118. Selvanathan, P.; Dorcet, V.; Roisnel, T.; Bernot, K.; Huang, G.; Le Guennic, B.; Norel, L.; Rigaut, S. Trans to cis photo-isomerization in merocyanine dysprosium and yttrium complexes. Dalton. Trans. 2018, 47, 4139–4148. [Google Scholar] [CrossRef] [PubMed]
  119. Cosquer, G.; Kamila, M.; Li, Z.-Y.; Breedlove, B.; Yamashita, M. Photo-modulation of single-molecule magnetic dynamics of a dysprosium dinuclear complex via a diarylethene bridge. Inorganics 2018, 6, 9. [Google Scholar] [CrossRef]
  120. Gou, X.; Chen, Z.; Jiang, J.; Liu, N.; Lan, W.; Wu, Y.; Cheng, P.; Shi, W. Influence of chelating ligands on photomagnetic properties of two erbium(III) coordination polymers. J. Rare Earths 2025, 43, 552–555. [Google Scholar] [CrossRef]
  121. Wang, X.-Q.; Geng, Y.-W.; Wang, Z.; Xie, C.; Han, T.; Cheng, P. Two-dimensional metal-organic framework with high-performance single-molecule magnets as nodes showing magnetic coercivity photomodulation. J. Am. Chem. Soc. 2025, 147, 18044–18053. [Google Scholar] [CrossRef]
  122. Gou, X.; Wu, Y.; Wen, H.; Li, L.; Lan, W.; Liu, N.; Zhang, S.; Ungur, L.; Cheng, P.; Shi, W. Modulating static and dynamic magnetizations of ytterbium(III) coordination polymers by light-induced radicals. CCS Chem. 2025, 1–9. [Google Scholar] [CrossRef]
  123. Zhang, P.; Zhang, Y.; Wu, F.; Xiao, W.; Hua, W.; Tang, Z.; Liu, W.; Chen, S.; Wang, Y.; Wu, W.; et al. Photoisomerization-mediated tunable pore size in metal organic frameworks for U(VI)/V(V) selective separation. Nat. Commun. 2025, 16, 2361. [Google Scholar] [CrossRef]
  124. Wu, L.; Liu, Y.; Yang, W.; Liu, Z.; Liu, C.; Yuan, X.; Zhang, L.; Ju, J.; Yao, X. In situ stimuli transfer in multi-environment shape-morphing hydrogels based on the copolymer between spiropyran and acrylic acid. Adv. Sci. 2025, 12, 2416173. [Google Scholar] [CrossRef]
  125. Tan, X.; Wang, Y.; Li, H.; Duan, Y.; Wen, B.; Zhao, J.; Kim, H.; Lee, J.Y.; Zhou, L.; Cheng, H.-B.; et al. Supramolecular synthesis of dithienylethene-albumin complexes for enhanced photoswitching in photoacoustic imaging-guided near-infrared photothermal therapy. Small 2025, 21, 2409027. [Google Scholar] [CrossRef]
  126. Jin, T.; Zhang, Z.; He, S.; Kaledin, A.L.; Xu, Z.; Liu, Y.; Zhang, P.; Beratan, D.N.; Lian, T. Shell thickness and heterogeneity dependence of triplet energy transfer between core-shell quantum dots and adsorbed molecules. J. Am. Chem. Soc. 2025, 147, 16282–16292. [Google Scholar] [CrossRef]
  127. Malkmus, S.; Koller, F.O.; Draxler, S.; Schrader, T.E.; Schreier, W.J.; Brust, T.; DiGirolamo, J.A.; Lees, W.J.; Zinth, W.; Braun, M. All-optical operation cycle on molecular bits with 250-GHz clock-rate based on photochromic fulgides. Adv. Funct. Mater. 2007, 17, 3657–3662. [Google Scholar] [CrossRef]
  128. Guo, F.-S.; Benjamin, M.D.; Chen, Y.-C.; Tong, M.-L.; Mansikkamäki, A.; Richard, A.L. A Dysprosium metallocene single-molecule magnet functioning at the axial limit. Angew. Chem. Int. Ed. 2017, 56, 11445–11449. [Google Scholar] [CrossRef]
  129. Ou, R.; Zhang, H.; Zhao, C.; Hegab, H.M.; Jiang, L.; Truong, V.X.; Wang, H. Photoresponsive styrylpyrene-modified MOFs for gated loading and release of cargo molecules. Chem. Mater. 2020, 32, 10621–10627. [Google Scholar] [CrossRef]
  130. Arima, H.; Hiraide, S.; Nagano, H.; Abylgazina, L.; Senkovska, I.; Auernhammer, G.K.; Fery, A.; Kaskel, S.; Watanabe, S. Atomic force microscopy strategies for capturing guest-induced structural transitions in single flexible metal-organic framework particles. J. Am. Chem. Soc. 2025, 147, 14491–14503. [Google Scholar] [CrossRef] [PubMed]
  131. Han, Z.; Zhang, R.; Jiang, J.; Chen, Z.; Ni, Y.; Xie, W.; Xu, J.; Zhou, Z.; Chen, J.; Cheng, P.; et al. High-efficiency lithium-ion transport in a porous coordination chain-based hydrogen-bonded framework. J. Am. Chem. Soc. 2023, 145, 10149–10158. [Google Scholar] [CrossRef] [PubMed]
  132. Han, Z.; Li, J.; Lu, W.; Wang, K.; Chen, Y.; Zhang, X.; Lin, L.; Han, X.; Teat, S.J.; Frogley, M.D.; et al. A {Ni12}-wheel-based metal-organic framework for coordinative binding of sulphur dioxide and nitrogen dioxide. Angew. Chem. Int. Ed. 2022, 61, e202115585. [Google Scholar] [CrossRef]
  133. Sun, T.; Min, H.; Han, Z.; Wang, L.; Cheng, P.; Shi, W. Rapid detection of nanoplastic particles by a luminescent Tb-based coordination polymer. Chin. Chem. Lett. 2024, 35, 108718. [Google Scholar] [CrossRef]
  134. Han, Z.; Yan, Z.; Wang, K.; Kang, X.; Lv, K.; Zhang, X.; Zhou, Z.; Yang, S.; Shi, W.; Cheng, P. Observation of oxygen evolution over a {Ni12}-cluster-based metal-organic framework. Sci. China Chem. 2022, 65, 1088–1093. [Google Scholar] [CrossRef]
  135. Jing, P.; Wu, B.; Han, Z.; Shi, W.; Cheng, P. An efficient Ag/MIL-100(Fe) catalyst for photothermal conversion of CO2 at ambient temperature. Chin. Chem. Lett. 2021, 32, 3505–3508. [Google Scholar] [CrossRef]
  136. Gou, X.; Liu, N.; Wu, Y.; Lan, W.; Wang, M.; Shi, W.; Cheng, P. Modulation of magnetization dynamics of an Er(III) coordination polymer by the conversion of a ligand to a radical using UV light. Dalton Trans. 2023, 52, 10372–10377. [Google Scholar] [CrossRef] [PubMed]
  137. Liu, X.; Wang, Y.-X.; Han, Z.-S.; Han, T.; Shi, W.; Cheng, P. Tuning the magnetization dynamics of TbIII-based single-chain magnets through substitution on the nitronyl nitroxide radical. Dalton Trans. 2019, 48, 8989–8994. [Google Scholar] [CrossRef] [PubMed]
  138. Zhang, X.; Xu, N.; Shi, W.; Wang, B.-W.; Cheng, P. The influence of an external magnetic field and magnetic-site dilution on the magnetization dynamics of a coordination network based on ferromagnetic coupled dinuclear dysprosium(III) units. Inorg. Chem. Front. 2018, 5, 432–437. [Google Scholar] [CrossRef]
  139. Wang, M.; Meng, X.; Song, F.; He, Y.; Shi, W.; Gao, H.; Tang, J.; Cheng, P. Reversible structural transformation induced switchable single-molecule magnet behavior in lanthanide metal-organic frameworks. Chem. Commun. 2018, 54, 10183–10186. [Google Scholar] [CrossRef]
  140. Liu, X.; Zhang, Y.; Shi, W.; Cheng, P. Rational design and synthesis of a chiral lanthanide-radical single-chain magnet. Inorg. Chem. 2018, 57, 13409–13414. [Google Scholar] [CrossRef]
  141. Song, F.; Zhang, Y.; Chen, J.; Han, T.; Cheng, P. Syntheses, crystal structures and magnetic properties of three lanthanide-nitronyl nitroxide complexes. J. Rare Earth 2017, 35, 24–27. [Google Scholar] [CrossRef]
  142. Zhang, S.; Li, H.; Duan, E.; Han, Z.; Li, L.; Tang, J.; Shi, W.; Cheng, P. A 3D heterometallic coordination polymer constructed by trimeric {NiDy2} single-molecule magnet units. Inorg. Chem. 2016, 55, 1202–1207. [Google Scholar] [CrossRef]
  143. Zhang, X.; Vieru, V.; Feng, X.; Liu, J.L.; Zhang, Z.; Na, B.; Shi, W.; Wang, B.-W.; Powell, A.K.; Chibotaru, L.F.; et al. Influence of guest exchange on the magnetization dynamics of dilanthanide single-molecule-magnet nodes within a metal-organic framework. Angew. Chem. Int. Ed. 2015, 54, 9861–9865. [Google Scholar] [CrossRef]
  144. Liu, K.; Li, H.; Zhang, X.; Shi, W.; Cheng, P. Constraining and tuning the coordination geometry of a lanthanide ion in metal-organic frameworks: Approach toward a single-molecule magnet. Inorg. Chem. 2015, 54, 10224–10231. [Google Scholar] [CrossRef]
  145. Li, L.; Liu, S.; Li, H.; Shi, W.; Cheng, P. Influence of external magnetic field and magnetic-site dilution on the magnetic dynamics of a one-dimensional Tb(III)-radical complex. Chem. Commun. 2015, 51, 10933–10936. [Google Scholar] [CrossRef]
  146. Chen, D.-M.; Ma, X.-Z.; Zhang, X.-J.; Xu, N.; Cheng, P. Switching a 2D Co(II) layer to a 3D Co7-cluster-based metal-organic framework: Syntheses, crystal structures, and magnetic properties. Inorg. Chem. 2015, 54, 2976–2982. [Google Scholar] [CrossRef]
  147. Zou, J.-Y.; Shi, W.; Gao, H.-L.; Cui, J.-Z.; Cheng, P. Spin canting and metamagnetism in 3D pillared-layer homospin cobalt(II) molecular magnetic materials constructed via a mixed ligands approach. Inorg. Chem. Front. 2014, 1, 242–248. [Google Scholar] [CrossRef]
  148. Han, T.; Shi, W.; Niu, Z.; Na, B.; Cheng, P. Magnetic blocking from exchange interactions: Slow relaxation of the magnetization and hysteresis loop observed in a dysprosium-nitronyl nitroxide chain compound with an antiferromagnetic ground state. Chem. Eur. J. 2013, 19, 994–1001. [Google Scholar] [CrossRef] [PubMed]
  149. Zhang, S.; Wang, B.; Xu, N.; Shi, W.; Gao, S.; Cheng, P. Ferromagnetic interactions in EO-azido-bridged binuclear transition metal(II) systems: Syntheses, crystal structures and magnetostructural correlations. Sci. China Chem. 2012, 55, 942–950. [Google Scholar] [CrossRef]
Figure 1. Molecular nanomagnets with photomagnetic properties.
Figure 1. Molecular nanomagnets with photomagnetic properties.
Magnetochemistry 11 00077 g001
Figure 2. Selected examples of organic molecules with photoinduced structural change groups.
Figure 2. Selected examples of organic molecules with photoinduced structural change groups.
Magnetochemistry 11 00077 g002
Figure 3. Selected examples of photoactive organic molecules that can generate stable radicals after illumination.
Figure 3. Selected examples of photoactive organic molecules that can generate stable radicals after illumination.
Magnetochemistry 11 00077 g003
Figure 4. Transition metal molecular nanomagnets with photomagnetic properties. (a) [FeII(1-propyltetrazole)6](BF4)2; (b) PhB(MesIm)3FeII-N = PPh3; (c) {[FeIII(Tp*)(CN)3]2FeII(bpmh)}·2H2O; (d) {[FeIII(bpy)(CN)4]2[CoII(phpy)2]}·2H2O; (e) {[W(CN)8][(FeII)(bib)2](bibH)}·2CH3OH; (f) {[(pzTp)FeIII(CN)3]4FeII2(Pmat)4}n·12H2O; (g) {[pzTpFeII(CN)3]4CoII2(Bib)4}·3H2O and {[pzTpFeII(CN)3]2CoII(Bpi)2}·CH3CN·4H2O. Reprinted with permission from Refs. [15,31,32,102] Copyright 2021 Nature Chemistry, 2013 American Chemical Society, 2013 American Chemical Society, and 2021 Angewandte Chemie International Edition. Reprinted from Refs. [33,34,103].
Figure 4. Transition metal molecular nanomagnets with photomagnetic properties. (a) [FeII(1-propyltetrazole)6](BF4)2; (b) PhB(MesIm)3FeII-N = PPh3; (c) {[FeIII(Tp*)(CN)3]2FeII(bpmh)}·2H2O; (d) {[FeIII(bpy)(CN)4]2[CoII(phpy)2]}·2H2O; (e) {[W(CN)8][(FeII)(bib)2](bibH)}·2CH3OH; (f) {[(pzTp)FeIII(CN)3]4FeII2(Pmat)4}n·12H2O; (g) {[pzTpFeII(CN)3]4CoII2(Bib)4}·3H2O and {[pzTpFeII(CN)3]2CoII(Bpi)2}·CH3CN·4H2O. Reprinted with permission from Refs. [15,31,32,102] Copyright 2021 Nature Chemistry, 2013 American Chemical Society, 2013 American Chemical Society, and 2021 Angewandte Chemie International Edition. Reprinted from Refs. [33,34,103].
Magnetochemistry 11 00077 g004
Figure 5. The χTT (a) and χ″–T (b) of {[FeIII(Tp*)(CN)3]2FeII(bpmh)}·2H2O. Reprinted from Ref. [33].
Figure 5. The χTT (a) and χ″–T (b) of {[FeIII(Tp*)(CN)3]2FeII(bpmh)}·2H2O. Reprinted from Ref. [33].
Magnetochemistry 11 00077 g005
Figure 6. Magnetic hysteresis loops of {[W(CN)8][(FeII)(bib)2](bibH)}·2CH3OH at 1.8 K (a) and 3.3 K (b) after irradiation. Reprinted with permission from Ref. [15]. Copyright 2021 Nature Chemistry.
Figure 6. Magnetic hysteresis loops of {[W(CN)8][(FeII)(bib)2](bibH)}·2CH3OH at 1.8 K (a) and 3.3 K (b) after irradiation. Reprinted with permission from Ref. [15]. Copyright 2021 Nature Chemistry.
Magnetochemistry 11 00077 g006
Figure 7. Selected examples of lanthanide molecular nanomagnets with photomagnetic properties. Reprinted with permission from Refs. [27,28,29,30,112,113,116,120,121]. Copyright 2015 Chemical Communications, 2018 Angewandte Chemie International Edition, 2019 American Chemical Society, 2020 American Chemical Society, 2017 Inorganic Chemistry Frontiers, 2021 Chemical Science, 2022 Journal of Materials Chemistry C, 2025 Journal of Rare Earths, 2025 American Chemical Society. Reprinted from Ref. [122].
Figure 7. Selected examples of lanthanide molecular nanomagnets with photomagnetic properties. Reprinted with permission from Refs. [27,28,29,30,112,113,116,120,121]. Copyright 2015 Chemical Communications, 2018 Angewandte Chemie International Edition, 2019 American Chemical Society, 2020 American Chemical Society, 2017 Inorganic Chemistry Frontiers, 2021 Chemical Science, 2022 Journal of Materials Chemistry C, 2025 Journal of Rare Earths, 2025 American Chemical Society. Reprinted from Ref. [122].
Magnetochemistry 11 00077 g007
Figure 8. The photochemical reaction from [Dy(bpe)(H2O)4(NO3)2](NO3)·2bpe (a) to [Dy2(tpcb)(H2O)8(NO3)4](NO3)2·2bpe·tpcb (b) via [2,2] photocyclization and their frequency-dependent ac susceptibilities (c,d). Reprinted with permission from Ref. [27]. Copyright 2015 Chemical Communications. The photochemical reaction from (Hbpe)2[Dy(bpe)(H2O)(4-pyO)(NO3)(SCN)3]SCN (e) to (Hbpe)2(H2tpcb)[Dy2(tpcb)2(H2O)2(4-pyO)2(NO3)2(SCN)6](SCN)2 (f) via a [2,2] photocyclization and their ln(τ) vs. T−1 plots (g,h). Reprinted with permission from Ref. [112]. Copyright 2017 Inorganic Chemistry Frontiers.
Figure 8. The photochemical reaction from [Dy(bpe)(H2O)4(NO3)2](NO3)·2bpe (a) to [Dy2(tpcb)(H2O)8(NO3)4](NO3)2·2bpe·tpcb (b) via [2,2] photocyclization and their frequency-dependent ac susceptibilities (c,d). Reprinted with permission from Ref. [27]. Copyright 2015 Chemical Communications. The photochemical reaction from (Hbpe)2[Dy(bpe)(H2O)(4-pyO)(NO3)(SCN)3]SCN (e) to (Hbpe)2(H2tpcb)[Dy2(tpcb)2(H2O)2(4-pyO)2(NO3)2(SCN)6](SCN)2 (f) via a [2,2] photocyclization and their ln(τ) vs. T−1 plots (g,h). Reprinted with permission from Ref. [112]. Copyright 2017 Inorganic Chemistry Frontiers.
Magnetochemistry 11 00077 g008
Figure 9. The photochemical reaction from DyIII(depma)(NO3)3(hmpa)2 (a) to DyIII2(depma2)(NO3)6(hmpa)4 (b) via [4,4] photocyclization and their ln(τ) vs. T−1 plots (c,d). Reprinted with permission from Ref. [28]. Copyright 2018 Angewandte Chemie International Edition. Structural transformation among 1RT, 1LT, and 1UV (e). The χMT vs. T (f) and ln(τ) vs. T−1 plots (gi) for 1, 1UV, and 1R. Reprinted with permission from Ref. [113]. Copyright 2021 Chemical Science.
Figure 9. The photochemical reaction from DyIII(depma)(NO3)3(hmpa)2 (a) to DyIII2(depma2)(NO3)6(hmpa)4 (b) via [4,4] photocyclization and their ln(τ) vs. T−1 plots (c,d). Reprinted with permission from Ref. [28]. Copyright 2018 Angewandte Chemie International Edition. Structural transformation among 1RT, 1LT, and 1UV (e). The χMT vs. T (f) and ln(τ) vs. T−1 plots (gi) for 1, 1UV, and 1R. Reprinted with permission from Ref. [113]. Copyright 2021 Chemical Science.
Magnetochemistry 11 00077 g009
Figure 10. The photochemical reaction from 1c (a) to 1o (b) via photocyclization and their τT (c) and magnetic hysteresis loops (d). Reprinted with permission from Ref. [29]. Copyright 2019 American Chemical Society.
Figure 10. The photochemical reaction from 1c (a) to 1o (b) via photocyclization and their τT (c) and magnetic hysteresis loops (d). Reprinted with permission from Ref. [29]. Copyright 2019 American Chemical Society.
Magnetochemistry 11 00077 g010
Figure 11. The structure of QDU-1 (a). The χMTT (b) and ac susceptibilities (c,d) of QDU-1, QDU-1a, and QDU-1b. Reprinted with permission from Ref. [30]. Copyright 2020 American Chemical Society. The structure of 1-Ln (e). The χMTT (f) and ac susceptibilities (gj) of 1-Dy, 1a-Dy, 2-Tb, and 2a-Tb. Reprinted with permission from Ref. [115]. Copyright 2022 Science China Materials.
Figure 11. The structure of QDU-1 (a). The χMTT (b) and ac susceptibilities (c,d) of QDU-1, QDU-1a, and QDU-1b. Reprinted with permission from Ref. [30]. Copyright 2020 American Chemical Society. The structure of 1-Ln (e). The χMTT (f) and ac susceptibilities (gj) of 1-Dy, 1a-Dy, 2-Tb, and 2a-Tb. Reprinted with permission from Ref. [115]. Copyright 2022 Science China Materials.
Magnetochemistry 11 00077 g011
Figure 12. Crystal structure of DyFe18C6 (a). The χMTT (b) and ac susceptibilities (c,d) of DyFe18C6 and DyFe18C6-UV. Reprinted with permission from Ref. [116]. Copyright 2022 Journal of Materials Chemistry C.
Figure 12. Crystal structure of DyFe18C6 (a). The χMTT (b) and ac susceptibilities (c,d) of DyFe18C6 and DyFe18C6-UV. Reprinted with permission from Ref. [116]. Copyright 2022 Journal of Materials Chemistry C.
Magnetochemistry 11 00077 g012
Figure 13. The structure of Er(phen) (a). The χMTT (b) and ln(τ)–T−1 plots (c,d) for Er(phen) and Er(phen)-UV. Reprinted with permission from Ref. [120]. Copyright 2025 Journal of Rare Earths.
Figure 13. The structure of Er(phen) (a). The χMTT (b) and ln(τ)–T−1 plots (c,d) for Er(phen) and Er(phen)-UV. Reprinted with permission from Ref. [120]. Copyright 2025 Journal of Rare Earths.
Magnetochemistry 11 00077 g013
Figure 14. Structure of [Dy(OPh)X(BPy)4(THF)]+ (X = Cl/PhO) cation in Dy(BPy) viewed from the side (a) and from above (b). Kagomé layer of Dy(BPy) (c). The magnetic hysteresis loops of Dy(BPy) (d) and Dy(BPy)-UV (e). Reprinted with permission from Ref. [121]. Copyright 2025 American Chemical Society.
Figure 14. Structure of [Dy(OPh)X(BPy)4(THF)]+ (X = Cl/PhO) cation in Dy(BPy) viewed from the side (a) and from above (b). Kagomé layer of Dy(BPy) (c). The magnetic hysteresis loops of Dy(BPy) (d) and Dy(BPy)-UV (e). Reprinted with permission from Ref. [121]. Copyright 2025 American Chemical Society.
Magnetochemistry 11 00077 g014
Figure 15. The photochemical reaction from Yb(ACA) to Yb(ACA)-UV (a). EPR spectra of Yb(ACA) and Yb(ACA)-UV (b). The χMTT (c) and ln(τ)–T−1 plots (d,e) of Yb(ACA) and Yb(ACA)-UV. Reprinted from Ref. [122].
Figure 15. The photochemical reaction from Yb(ACA) to Yb(ACA)-UV (a). EPR spectra of Yb(ACA) and Yb(ACA)-UV (b). The χMTT (c) and ln(τ)–T−1 plots (d,e) of Yb(ACA) and Yb(ACA)-UV. Reprinted from Ref. [122].
Magnetochemistry 11 00077 g015
Table 1. The static and dynamic magnetic properties of selected photoresponsive lanthanide molecular nanomagnets discussed in this review.
Table 1. The static and dynamic magnetic properties of selected photoresponsive lanthanide molecular nanomagnets discussed in this review.
Complex/
Photoproduct
Photochemical
Reaction
χMT
at RT
cm3 mol−1 K
Hysteresis
Loop at 2 K
Hc (Oe)
Ueff (K)
Hdc (Oe)
Ref.
[Dy(bpe)(H2O)4(NO3)2](NO3)·2bpe[2,2] cycloaddition13.9--
(slow magnetization relaxation)
[27]
[Dy2(tpcb)(H2O)8(NO3)4](NO3)2·2bpe·tpcb14.2--
(Hbpe)2[Dy(bpe)(H2O)(4-pyO)(NO3)(SCN)3]SCN with bpe[2,2] cycloaddition13.9-
(butterfly-shaped)
153.8 (Hdc = 0)
201.9 (Hdc = 1500)
[112]
(Hbpe)2(H2tpcb)[Dy2(tpcb)2(H2O)2(4-pyO)2(NO3)2(SCN)6](SCN)213.9-
(smaller butterfly-shaped)
205.5 (Hdc = 0)
234.5 (Hdc = 1500)
DyIII(depma)(NO3)3(hmpa)2[4 + 4] cycloaddition13.8-20.4 (Hdc = 500)[28]
DyIII2(depma2)(NO3)6(hmpa)414.1-43.2 (Hdc = 500)
DyIII(SCN)3(depma)2(4-hpy)2 (1)[4 + 4] cycloaddition14.1-
(butterfly-shaped)
141 (Hdc = 0)[113]
1UV14.0-101 (Hdc = 0)
1R14.1-
(butterfly-shaped)
153 (Hdc = 0)
[Dy(Tppy)F(Lc)]PF6 (1c)ring opening13.9-
(butterfly-shaped)
225.9 (Hdc = 0)[29]
1o13.8-
(smaller butterfly-shaped)
225.9 (Hdc = 0)
[Dy3(H-HEDP)3(H2-HEDP)3]·2H3-TPT·H4-HEDP·10H2O (QDU-1)photogenerated radical42.7--(SMM off)[30]
QDU-1a43.7-108.1 (Hdc = 0)
QDU-1b42.7--(SMM off)
[Ln3(H-HEDP)2(H2-HEDP)2(H3-HEDP)2]·H3-TPP·11H2O (Ln = Dy (1-Dy))photogenerated radical42.6-16.9 (Hdc = 2000)[115]
1a-Dy41.8--(SMM off)
[Ln3(H-HEDP)2(H2-HEDP)2(H3-HEDP)2]·H3-TPP·11H2O (Ln = Tb (2-Tb))35.2-12.1 (Hdc = 2000)
2a-Tb33.3--(SMM off)
[DyIII(18C6)(H2O)3]FeIII(CN)6·2H2O (DyFe18C6)photogenerated radical14.5-18.4 (Hdc = 800)[116]
DyFe18C6-UV17.1-17.0 (Hdc = 800)
[Er(CA)1.5(phen)(DMF)]n (Er(phen))photogenerated radical11.5-66.4 (Hdc = 600)[120]
Er(phen)-UV14.4-72.7 (Hdc = 600)
{[Dy1.5(OPh)2Cl(BPy)3(THF)1.5][(BPh4)1.5]·0.5THF}n ((Dy(BPy))photogenerated radical20.57Hc = 4500 Oe822 (Hdc = 0, T = 12–40 K)
1048 (Hdc = 0, T = 53–71 K)
[121]
Dy(BPy)-UV20.63Hc = 1300 Oe641 (Hdc = 0, T = 12–40 K)
1000 (Hdc = 0, T = 53–71 K)
[Yb(CA)(ACA)(DMF)(H2O)]n (Yb(ACA))photogenerated radical2.65-94 (Hdc = 800)[122]
Yb@Y(ACA)-UV4.83-111 (Hdc = 800)
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gou, X.; Sun, X.; Cheng, P.; Shi, W. Molecular Nanomagnets with Photomagnetic Properties: Design Strategies and Recent Advances. Magnetochemistry 2025, 11, 77. https://doi.org/10.3390/magnetochemistry11090077

AMA Style

Gou X, Sun X, Cheng P, Shi W. Molecular Nanomagnets with Photomagnetic Properties: Design Strategies and Recent Advances. Magnetochemistry. 2025; 11(9):77. https://doi.org/10.3390/magnetochemistry11090077

Chicago/Turabian Style

Gou, Xiaoshuang, Xinyu Sun, Peng Cheng, and Wei Shi. 2025. "Molecular Nanomagnets with Photomagnetic Properties: Design Strategies and Recent Advances" Magnetochemistry 11, no. 9: 77. https://doi.org/10.3390/magnetochemistry11090077

APA Style

Gou, X., Sun, X., Cheng, P., & Shi, W. (2025). Molecular Nanomagnets with Photomagnetic Properties: Design Strategies and Recent Advances. Magnetochemistry, 11(9), 77. https://doi.org/10.3390/magnetochemistry11090077

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop