DOI:
10.1039/D4CC05673F
(Feature Article)
Chem. Commun., 2025,
61, 627-638
Valence tautomerism, non-innocence, and emergent magnetic phenomena in lanthanide-organic tessellations
Received
24th October 2024
, Accepted 26th November 2024
First published on 28th November 2024
Abstract
Coordination networks based on lanthanide ions entangle collective magnetic phenomena, otherwise only observed in inorganic 4f materials, and the tunable spatial and electronic structure engineering intrinsic to coordination chemistry. In this review, we discuss the use of 2D-structure-directing linear {LnII/IIII2} nodes to direct the formation of polymeric coordination networks. The equatorial coordination plasticity of {LnII/IIII2} results in broad structural diversity, including previously unobtainable tessellations containing motifs observed in quasicrystalline tilings. The new phases host also magnetic frustration, which is at the origin of enhanced magnetic refrigeration potential. Finally, careful redox matching of Ln node and frontier orbitals of the ligand scaffold has culminated in the discovery of quantitative valence tautomeric conversion in a molecule-based Ln material, opening up new avenues for combining exotic magnetic phenomena with an encoded switch.
Maja A. Dunstan
| Maja A. Dunstan is from Melbourne, Australia, and completed her PhD in the group of Colette Boskovic at the University of Melbourne. In 2022, she joined Kasper S. Pedersen at the Technical University of Copenhagen as a postdoctoral researcher. Her research interests include single-molecule magnetism, neutron spectroscopy, and switchable magnetic materials. Currently, she is working in the field of low-valent lanthanide coordination networks. |
Kasper S. Pedersen
| Kasper S. Pedersen was born in Allerød, Denmark, and received his PhD from the University of Copenhagen in 2014. After postdoctoral stints in Bordeaux and Montreal, he joined the faculty at the Technical University of Denmark in 2017, and was appointed full professor in 2022. His research spans synthetic and structural inorganic chemistry as well as advanced properties characterization. His current focus is directed toward the discovery of novel metal–organic framework materials with hitherto unknown structural and physical properties, and the development of the field of low/zero-valent metal cluster-based MOFs and the application of 3D electron diffraction techniques. |
1. Introduction
Switchable materials, which exhibit disparate chemical and physical properties across multiple states, are necessary for applications that leverage changes in electronic, magnetic, or optical characteristics. Many such materials exist, both molecule-based and purely inorganic, including those where the electronic state can be manipulated and controlled by a stimulus such as temperature or pressure.1 These encompass phenomena such as spin-crossover (SCO), between high-spin (hs) and low-spin (ls) tautomers of a transition metal ion complex;2 stimulated intramolecular electron transfer between two metal ions; and valence tautomerism (VT), involving stimulated intramolecular electron transfer between a redox-active metal ion and a redox-active ligand (Fig. 1).3,4 For SCO and VT materials, the vast majority of research has focused on coordination complexes of the 3d metal ions, for which there are known sets of metal ions and ligand scaffolds leading to facile switching by mild stimuli, such as temperature, pressure, and light.
|
| Fig. 1 Valence tautomeric process in an octahedral Co-dioxolene species, with a transition between a CoIII-catecholate and a CoII-semiquinonate. Occupation of metal 3d (teal/blue) and ligand (pink) valence orbitals shown. | |
VT transitions have been observed in coordination compounds of Mn,5 Fe,6,7 Co,8–10 Cu,11 Sn,12 V,13 amongst others, although most widely researched is the Co-dioxolene family of compounds (Fig. 1). Buchanan and Pierpont first observed the molecule-based VT phenomenon in 1980 in [Co(2,2′-bpy)(dbsq)(dbdiox)] (3,5-dbsq = 3,5-di-tert-butyl semiquinone, 3,5-dbdiox = 3,5-di-tert-butyl dioxolene, 2,2′-bpy = 2,2′-bipyridine), which hosts a thermally driven transition between a high-temperature (HT) CoII-dbsq-dbsq tautomer and a low-temperature (LT) CoIII-dbsq-dbq (dbq = 3,5-di-tert-butyl quinone) tautomer.14 Progress in the proceeding forty years has led to the observation of VT switching induced by various additional stimuli such as light,15–17 pressure,18,19 and X-rays.16,20 Immobilization of VT complexes on surfaces21 or on gold nanoparticles22 has also been achieved for Co-dioxolene systems, with the possibility of now isolating single molecules as molecular switches. For many applications, cooperativity is key, leading to an abrupt SCO or VT transition upon an applied stimulus. In molecular materials, cooperativity in the solid-state bulk material is dictated by the crystal packing, which is difficult or impossible to design for molecular complexes. Instead, incorporation of switchable modules into coordination polymers or metal–organic frameworks (MOFs) is a tactic for engendering cooperativity. For SCO, multidimensional coordination networks have been widely explored since the 1990s, when Kahn et al. demonstrated cooperative, hysteretic SCO switching in FeII 1D coordination polymers at room temperature.23,24 Aside from the possibility of tuning both transition temperature and cooperativity, SCO coordination networks can also host multi-step transitions, where multiple states are accessible.25,26 These materials also exhibit additional advantages over molecular complexes, for example where guest molecule–framework interactions induce a SCO transition.27–29 In contrast, examples of VT coordination networks incorporating Co-dioxolene units typically show gradual transitions, with a strong sensitivity to desolvation30–32 and guest molecule interactions.33 One notable exception is the 2D coordination network [Mn2(NITIm)3]ClO4 (NITImH = 2-(2-imidazolyl)-4,4,5,5-tetramethyl-4,5-dihydro-1H-3-oxide-1-oxyl), which shows a sharp thermal VT transition at room temperature with a thermal hysteresis of 20 K.34 This observation hints that with a suitable combination of metal ion and ligand scaffold, extended networks exhibiting VT transitions could parallel the robust bistability found in SCO materials. However, for VT materials, unlike SCO materials, the concerted change in electronic structure of both the metal ion and the ligand scaffold may result in a highly correlated state existing in close energetic proximity to a diamagnetic or localised magnetic state. In extended networks, such correlations extend beyond the length scale of the molecule and will be at the origin of novel electronic and magnetic phases.
The VT transition is an entropically driven process. In Co-dioxolene compounds specifically, the entropy gain in the paramagnetic, HT CoII-sq˙− form is due to the larger spin state degeneracy leading to a gain in electronic entropy, as well as the longer metal–ligand bond lengths, leading to a higher density of vibrational states and a larger vibrational entropy. The thermal equilibrium is therefore determined by the Gibbs free energy, ΔG = ΔH − TΔS. According to this, the transition will occur when ΔG = 0 and ΔH = TcΔS, which also defines the critical temperature, Tc.35 Substitution of the dioxolene ligands has allowed for controlled tuning of the ligand redox potential, while careful choice in auxiliary ligand allows for tuning of the CoII/CoIII redox couple, both directly modifying ΔH.36 In addition, both solvent effects in solution state37 and crystal packing effects38 in the solid state affect the observed transition.
In stark contrast to transition metal ions, the lanthanide (Ln) ions are almost exclusively found as trivalent ions with 4fn electronic configurations and are typically considered redox innocent. For most LnIII ions, unquenched orbital angular momentum and significant spin–orbit coupling dominate the electronic properties, with the crystal field effect of the ligand scaffold being only weak. One consequence of the unique electronic structure of the LnIII ions is often large magnetic moments with strong magnetic anisotropy, leading to widespread technological applications in hard magnets (e.g., SmCo5, Nd2Fe14B39) or potential future applications of single-molecule magnets.40–43 Even in the case of negligible magnetic anisotropy, such as for 4f7 ions, e.g., GdIII, the large magnetic moment has led to its use in MRI contrast agents44 and in low-temperature magnetocaloric materials.45 On the other hand, the weak crystal field interaction with the coordination environment gives rise to sharp, long-lived luminescence, which is used in, e.g., phosphors, sensors, and luminescent thermometers.46 As these phenomena are intimately linked to the electronic configuration of the LnIII ions, any change in oxidation state necessarily leads to a dramatic change in properties. In other words, manipulation of the oxidation state by a VT transition could allow for any of these properties to be switchable by an external stimulus.
In this Feature Article, we discuss the emergence of VT transitions in extended coordination networks based on Ln ions. First, the known redox properties of Ln species are briefly discussed, which are central to fluxional valence phenomena in lanthanide inorganic solid-state materials, as well as VT and intermediate valence tautomerism in molecular Ln materials. We then outline the work we have recently undertaken in the field of Ln-based coordination networks, which includes observation of novel geometric tessellations, slow magnetic relaxation, and enhanced sub-Kelvin magnetic refrigeration. This work culminates in the recent discovery of a samarium–pyrazine framework featuring a thermally activated and complete VT transition.
2. Frozen oxidation states
The lanthanide ions are by far most common in their trivalent state, with notable exceptions to this rule exhibiting either a closed-shell or half-filled 4f configuration: CeIV (4f0), EuII (4f7), and YbII (4f14). Despite this redox-inactivity, synthetic efforts, primarily in the last decade, have now resulted in the isolation of LnII compounds for all Ln except Pm,47–50 and LnIV coordination compounds for Ce,51–53 Pr,54,55 and Tb.56,57 The implications for coordination chemistry utilising the LnIII/LnII and LnIV/LnIII redox couples have been reviewed by others.58,59 Of the divalent Ln ions, Sm, Eu, and Yb are most stable towards oxidation to their trivalent state (Table 1), and simple precursors are commercially available. EuII, in particular, is widely used in inorganic phosphors, exhibiting tunable 4f7 ← 4f65d emission in the visible spectrum.60 In the field of molecular magnetism, organometallic complexes of TbII and DyII have shown promise as improved single-molecule magnets.61 Synthetically, LnII ions are widely used as one-electron reducing agents62 or more recently, as catalysts,63,64 leveraging their large and negative reduction potentials. SmI2, specifically, is widely used as a versatile, strong one-electron reductant for diverse organic transformations.65,66
Table 1 Standard reduction potentials (E1/2vs. NHE) of LnIII/LnII redox couples for selected Ln67
|
Nd |
Sm |
Eu |
Dy |
Tm |
Yb |
E
1/2(LnIII/LnII)/V |
−2.6 |
−1.5 |
−0.34 |
−2.6 |
−2.2 |
−1.2 |
The possibility to exploit different redox states in molecular systems containing Ln ions is not necessarily limited to the metal centre itself, but could involve the ligand scaffold – a situation of great interest in the field of molecular magnetism. Such resulting radical ligand systems have been a synthetic target in the field of single-molecule magnets (SMMs), compounds which exhibit magnetic bistability at low temperatures.68 In such systems, a strong magnetic anisotropy leads to an energy barrier to magnetization reversal and a wide magnetic hysteresis. This field was dominated by 3d-based polynuclear complexes, until Ishikawa et al. reported that the Ln complex [TbIII(Pc)2]− (Pc = phthalocyaninate) exhibited the same slow magnetic relaxation as the polynuclear complex-based SMMs.69 In the proceeding 20 years, this has led to the discovery of SMMs with record breaking working temperatures.40–43
Further improving the performance of Ln-SMMs via coupling of electronic spins in polynuclear assemblies, akin to the strategy pursued for the 3d metal ion complexes, has been of limited success. The problem is rooted in the shielded nature of the 4f orbitals, that hinders any strong magnetic interaction between Ln ions in a given material. Ligand-based radicals often possess significant electronic spin densities at the ligating atom, thereby promoting the strong magnetic interactions in Ln-radical systems. Notable examples include the N23−-bridged Ln dinuclear complexes, which show some of the highest hysteresis temperatures and, notably, extremely large magnetic coercivities.70–72 The group of Murugesu recently published tetranuclear complexes, [(Cp*2LnIII)4(tz˙−)4] (Cp* = pentamethylcyclopentadienide), consisting of Cp*2LnIII units bridged by tetrazine (tz˙−) anion radical ligands. Here, strong magnetic exchange coupling leads to a large coercive field of Hcoer = 3 T at 1.8 K.73 Switching to a pyrazine anion radical ligand (pyz˙−), the related family of compounds [(Cp*2LnIII)4(pyz˙−)4] could be obtained.74 Here, a coercive field of 6.5 T is observed for the Dy analogue. The extraordinary properties stem from strong Ln-radical magnetic exchange coupling, combined with a strongly axial crystal field at each DyIII ion. Despite the potential of such scenarios for generating novel magnetic phases in coordination networks, these are yet to be discovered.
3. Redox isomerism and related phenomena in lanthanide materials
3.1 The dawn of VT transitions in lanthanide materials
Divalent Ln ions were first widely observed in inorganic solids. Perhaps unsurprisingly, valence change transitions have therefore been observed in intermetallics, chalcogenides, and oxides for all the common LnII ions. These transitions are typically characterised by (partial) electron transfer from the localised 4f orbitals to a conduction band, formally corresponding to an increase in the Ln oxidation state. Conceptually, this situation is analogous to the previously discussed VT transition, while commonly referred to as a valence change transition or valence instability. The first observation in 1912 of such a transition was in metallic Ce.75 Herein, Ce exhibits a temperature- and pressure-dependent change in oxidation state from CeIII to an intermediate valence Ce3.6+ phase, due to the promotion of 4f electrons to the s–d conduction band.75–78 Alloying with Th yields Ce1−xThx phases exhibiting thermally driven VT transitions, allowing for tuning of the transition temperature and pressure, with a cooperativity governed by the Th content (Fig. 2(a)).79 Similarly, in samarium monosulfide, a pressure-induced (0.65 GPa) VT transition promotes an electron from SmII to the conduction band, giving a formal oxidation state of near 2.6+ in the “golden” intermediate-valence phase (Fig. 2(b)).80,81 This transition is accompanied by a dramatic change in the physical properties, reflected in the semiconductor-to-metal transition. Even larger pressures lead to a complete transition to SmIII, resulting in a metallic, correlated heavy fermion phase.82 Chemical alloying of SmS with Y, Eu, Gd, or Yb leads to a shift in the temperature and pressure at which the valence transition occurs, analogous to the shift in the thermally induced transition in Ce1−xThx alloys.83–85 For example, Sm1−xYxS stabilises the golden phase due to the “chemical pressure” from the dopant, and by x = 0.2, the intermediate valence state is observed at ambient pressure and room temperature.86
|
| Fig. 2 (a) Temperature-dependent magnetic susceptibility (χ) in Ce1−xThx alloys exhibiting thermally induced valence transitions. Reproduced with permission from EDP Sciences (1976).79 (b) Phase diagram of SmS, showing temperature- and pressure-dependent valence change transitions. Reproduced from Springer Nature (2017).87 (c) Temperature dependence of the unit cell volume of Sm2.75C60, demonstrating the large negative thermal expansion driven by the valence change transition. Reproduced with permission from Springer Nature (2003).88 | |
Lanthanide fullerides similarly exhibit VT transitions stimulated by temperature or pressure. In these compounds, Ln ions are intercalated between fulleride anions. Due to the molecular nature of the fullerides, the electron-accepting state is not necessarily the conduction-band, but could be of isolated, molecular origin. At room temperature, Sm2.75C60 is intermediate valence, with a Sm valence of 2.3+.88 Upon decreasing temperature, a VT transition at 32 K leads to a localised SmII state, accompanied by a negative thermal expansion due to the increased ionic radius (Fig. 2(c)). Here, the electronic driving force for the VT transition is the thermal population of the low-lying t1u level of C60 upon heating, leading to an overall decrease in occupation of the Sm 4f orbitals. An analogous VT transition is observed in the Yb congener, Yb2.75C60, with a higher transition temperature of 60 K.89 A room-temperature, pressure-induced (4 GPa) VT transition is also observed for Sm2.75C60, which leads to the more intuitively expected increase in valency to SmIII, driven by the smaller radius of the higher oxidation state ion.90 Indeed, VT transitions are now widespread in inorganic solids, in particular in intermetallics.91–95 Herein, the VT transitions are accompanied by stark changes in structural, electronic, magnetic, and optical properties.
3.2 Valence instabilities and nebulous oxidation states
In molecular Ln materials, observation of VT transitions is extremely limited. In 2009, Fedushkin et al. reported the observation of an in-solution VT transition in [(dpp-bian)Yb(dme)(μ-Br)]2 (dpp-bian = 1,2-bis[(2,6-diisopropylphenyl)imino]-acenaphthene; dme = dimethoxyethane).96 Here, a HT tautomer was observed to interconvert with a LT tautomer above room temperature. This was followed in 2012 by the observation of a VT transition in the compound [(dpp-bian)Yb(dme)(μ-Cl)]2 (Fig. 3(a)) in the solid state.97 Single-crystal X-ray diffraction allowed the authors to conclude that one half of the dimer undergoes a VT transition–i.e. a LT-{YbIII2(bian2−)2} ⇄ HT-{YbIIIYbII(bian2−)(bian˙−)} conversion (Fig. 3(b)). For this compound, only a single polymorph undergoes the VT transition, which hints that the tautomerism is particularly dependent on steric effects. This year, a gradual VT transition in the related, monomeric complex [(ArBIG-bian)Yb{(C(S)NMe2)(dme)] (ArBIG-bian = 1,2-bis[(2,6-dibenzhydryl-4-methylphenyl)imino]acenaphthene) was reported.98 Herein, the HT phase (350 K) of a 1:1 mixture of {YbIII(bian2−)} and YbII(bian˙−)}, gradually transitions to a 3:1 mixture in the LT phase (100 K). The presence of four crystallographically independent molecules in the unit cell, only one of which transitions, showcases, again, the sensitivity of the VT transition to minute structural differences.
|
| Fig. 3 (a) Molecular complex of [{(dpp-bian)Yb(dme)(μ-Cl)}2].97 Isopropyl groups and H-atoms are omitted for clarity. (b) Magnetic moment with temperature for several single-crystals of [{(dpp-bian)Yb(dme)(μ-Cl)}2]. Reproduced with permission from Wiley (2012).97 | |
Changes in local coordination environment due to chemical pressure can also induce VT transitions in Ln molecular complexes. When functionalised Ln(Pc)2 complexes, Ce{(15C5)4Pc}299 and Eu{(BuO)8Pc}2100 are introduced into a Langmuir air/water monolayer, they undergo an electron transfer from one ligand to the metal, undergoing CeIV{Pc2−}2 ⇄ CeIII{(Pc2−)(Pc˙−)} and EuIII{(Pc2−)(Pc˙−)} ⇄ EuII{Pc˙−}2 transitions, respectively. For Ce{(15C5)4Pc}2, the surface pressure could be tuned to yield different tautomers. In Ce(Pc)2, the same tautomerism could be induced by deposition on a gold surface, which induces a 45° twist between the Pc2−/˙− ligands, triggering the transition.101
Pairing Ln ions with other redox-active ligands such as 2,2′-bipyridine (2,2′-bpy) ligands has led to the observation of intermediate valency.102 Here, close energetic proximity of ligand and Ln valence states lead to multi-configurational ground states, where the formal valence of both the Ln ion and the ligand are non-integer. The Yb(Cp*)2L systems, where L are derivatives of 2,2′-bpy, are particularly well studied examples of intermediate valency in molecules. Within this family, a phenomenon termed intermediate valence tautomerism (IVT) is exhibited by (Cp*)2Yb(4,4′-Me2-2,2′-bpy) (Fig. 4(a)).103,104 In this compound, magnetometry and Yb L3-edge X-ray absorption near-edge structure (XANES) spectra are consistent with a hysteretic valence transition between two distinct intermediate valence tautomers (Fig. 4(b)). Intermediate valence tautomerism is now well established in both purely inorganic solids and molecular Ln complexes, whilst remaining elusive in coordination networks of any kind.
|
| Fig. 4 (a) Structure of (Cp*)2Yb(4,4′-Me2-2,2′-bpy).104 (b) f-hole occupancy (nf) determined by fitting of XANES spectra across the thermal intermediate valence transition. Reprinted with permission from ref. 103. Copyright 2010 American Chemical Society. | |
4. Tiling valence, tautomerism & magnetism
4.1 Inspiration from transition metal tessellations
Coordination networks extending into one, two, or three dimensions conceptually reside between extended inorganic solids and molecular coordination complexes, thereby benefitting from the designable molecular properties and by physical properties tied to long-range correlations. This is well exemplified by the family of compounds spurred from the discovery of the 2D network CrCl2(pyz)2 (Fig. 5). The reducing nature of CrII leads to an electronic ground state approximated by the formulation CrIII{(pyz0)(pyz˙−)}.105 Strong metal–radical magnetic interactions and significant π–d conjugation lead to ferrimagnetic ordering at relatively high temperatures (55 K) and one of the highest observed electrical conductivities in a coordination compound.
|
| Fig. 5 (a) 2D square tilings formed by MCl2(pyz)2. (b) Local coordination environment of M ions. (c) Packing of 2D sheets. Hydrogen atoms are omitted for clarity.105 | |
Modulation of the CrII/CrIII redox potential through ligand substitution of the trans-chlorido ligands for methanesulfonate yields instead the CrII(pyz0)2 redox isomer, which is electrically insulating and exhibits an antiferromagnetic ground state.106 A postsynthetic reduction of CrCl2(pyz)2 leads to formation of Li0.7Cl0.7Cr(pyz)2·0.25 thf (thf = tetrahydrofuran), consisting of layers of CrII(pyz˙−)2. Here, the strong metal–radical interactions lead to a massive 0.75 T coercivity at room temperature, rivalling rare-earth based hard magnets.107 Similar layered 2D sheets to CrCl2(pyz)2 can be obtained with Ti and V.108 TiCl2(pyz)2, approximated as TiIII{(pyz0)(pyz˙−)}, displays strongly correlated Fermi liquid behaviour, leading to the highest room-temperature electronic conductivity for any metal–organic solid with octahedrally coordinated metal ions. In contrast, the antiferromagnetic insulator VCl2(pyz)2 exhibits no electron transfer from the metal to the ligand, leading to a VII(pyz0)2 redox isomer. These highly correlated, extended materials stand in stark contrast to the vast majority of typical framework materials, which are typically constructed from redox-inactive linkers and metal nodes.
4.2 Innocent Archimedean tessellations
Akin to the MCl2(pyz)2 series, trans-dihalido Ln nodes could be similarly expected to structurally direct 2D sheet-like tessellations. Labile solvates of the trans-{LnIII2} moieties are readily formed by dissolution of LnI2 forming what is posited to be trans-LnI2(CH3CN)5 (Ln = Sm, Eu, Yb)109 in acetonitrile, and trans-LnI2(thf)5 (Ln = Nd,110 Sm,111 Eu112) or trans-LnI2(thf)4 (Ln = Yb113) in thf (Fig. 6). Unlike the MCl2(pyz)2 tessellations, where octahedral nodes lead to square tilings (Fig. 5), the {LnIII2} nodes are pseudo-pentagonal bipyramidal, and as such should engender 2D tilings with five-fold nodes. The proof of principle was achieved through the combination of {YbIII2} nodes with ditopic 4,4′-bipyridine (bpy), leading to either 2D sheets, YbI2(bpy)5/2, or pseudo-1D ladders, YbI2(bpy)2(thf), depending on the reaction solvent (Fig. 7(a) and (b)).114 In YbI2(bpy)5/2, irrespective of the close resemblance of the first coordination sphere of Yb to D5h symmetry–incompatible with periodicity–the flexibility of the bpy ligands results in an elusive, idealized elongated triangular tessellation, consisting of triangular and quadratic motifs. Despite the large and negative reduction potential of the YbIII/YbII redox couple (Table 1), no full or partial electron transfer occurs between YbII and bpy, contrasting the observation in, e.g., CrCl2(pyz)2 (E1/2(CrIII/CrII) = −0.42 V). Confirmation of the YbII oxidation state in YbI2(bpy)5/2 and YbI2(bpy)2(thf) is given by their diamagnetism in the entire temperature range of 1.7–273 K. This assignment is further supported by bond length analysis, where the Yb–I bond length is a sensitive measure of the Yb oxidation state (Fig. 8).
|
| Fig. 6 Left to right: LnII complexes formed in solution: LnI2(thf)5,111 LnI2(CH3CN)5,109 and a LnIII complex, LnI2(thf)5+.115 Selected angles shown illustrate the pseudo-pentagonal bipyramidal coordination sphere. Sm–I bond lengths are highlighted to indicate the ca. 6% decrease upon one-electron oxidation. | |
|
| Fig. 7 Representation of tessellations formed through the combination of trans-{LnI2} nodes with ditopic pyz and bpy ligands. (a) Elongated triangular net LnI2(bpy)5/2.114,116,117 (b) Pseudo-1D chains of LnI2(bpy)2(thf).114 (c) Snub square net of LnI2(bpy)5/2.116 (d) Elongated triangular net of LnI2(pyz)5/2.117 (e) Pseudo-1D chains of LnI2(pyz)3.118 (f) Triangular net of Ln@BaI2(pyz)3.119 | |
|
| Fig. 8 Local coordination sphere of LnI2(bpy)x, LnI2(pyz)x (left), highlighting typical bond angles. Plot of Ln–I bond lengths across the Ln series (right), with groups of LnII and LnIII oxidation states highlighted. For SmI2(pyz)3, a HT and LT valence tautomer can be identified. | |
A similar Archimedean tessellation, i.e., a tiling of different polygons, to that of YbI2(bpy)5/2 has previously been observed by scanning tunnelling microscopy in on-surface self-assembled structures of Eu atoms and para-quaterphenyldicarbonitrile.120 Herein, several metal–organic Archimedean tessellations were shown to form under near identical conditions suggesting little energetic difference. Notably, the random tessellation of triangles and squares led to the realization of a dodecagonal quasicrystalline phase. The observation of the periodic YbI2(bpy)5/2 quasicrystal approximant phase suggests that similar 2D quasicrystalline materials could be designed in bulk, crystalline materials.121,122 On a separate note, coordination networks incorporating divalent Ln ions instead of the much more common trivalent ions,123–125 are rare. The spontaneous self-assembly of YbI2(bpy)5/2 and YbI2(bpy)2(thf) suggests that a wealth of, in coordination chemistry, hitherto unknown tessellations may be readily obtainable.
4.3 Guilty magnetic tilings
In order to leverage the magnetic properties of Ln-based materials, paramagnetic Ln ions were sought. Further, inspired by the burgeoning literature on Ln–radical complexes and their extraordinary magnetic properties, actuation of ligand scaffold non-innocence was pursued. Compared to the less reducing YbII, the largely negative reduction potential of, e.g., DyII (Table 1) is such that even many common solvents are reduced, making exotic reagents like DyI2 impractical. Instead of relying on an in situ reduction of the ligand by the LnII ion, the solution reaction of trivalent LnI3 (Ln = Gd, Dy) with a mixture of bpy0 and bpy˙−, formed simply by Na reduction, yields 2D networks of formula LnI2(bpy)5/2, analogous to the previously discussed YbI2(bpy)5/2.116 However, for Ln = Gd, Dy, a snub square tiling structure is obtained (Fig. 7(c)), as well as a side-phase of the elongated triangular tiling for Gd (Fig. 7(a)). For both DyI2(bpy)5/2 and GdI2(bpy)5/2, the Ln–I bond length analyses are consistent with the presence of LnIII ions (Fig. 8), implying a radical nature of the bpy scaffold and an approximate electronic configuration of LnIII{(bpy˙−)(bpy0)3/2}. The ligand scaffold mixed-valency is reminiscent of the situation encountered in CrCl2(pyz)2, however, the absence of any intervalence charge transfer transitions in GdI2(bpy)5/2 suggests the presence of weakly electronically coupled bpy fragments and an anticipated poor electrical conductivity.
The magnetometric data for GdI2(bpy)5/2 suggest only weak antiferromagnetic GdIII–bpy˙− exchange interactions, on the order of J = 0.07 cm−1 (JŜ1Ŝ2 convention), as often encountered in Ln–radical systems. This number, however, is significantly smaller than those found in the previous discussed molecular complexes.70–74 In the case of DyI2(bpy)5/2, the axial ligand field induced by the anionic trans-iodido ligands combined with a pseudo-D5d symmetry at the DyIII centre allows for the observation of slow paramagnetic relaxation in zero applied magnetic field (Fig. 9). However, despite the presence of a small DyIII–bpy˙− exchange bias, rapid quantum tunnelling of magnetisation is observed.
|
| Fig. 9 In-phase (χ′, left) and out-of-phase (χ′′, right) ac susceptibility data for polycrystalline DyI2(bpy)5/2 obtained at selected temperatures and in the absence of a static (dc) magnetic field. Inset: Temperature dependence of the paramagnetic relaxation rate, τ−1, vs. temperature. The solid lines are simulations, yielding an effective energy barrier to magnetization reversal of 63 cm−1 when measured in Hdc = 3 T. Reprinted with permission from ref. 116. Copyright 2021 American Chemical Society. | |
The isolation of a snub square tiling in a coordination network is exceedingly rare, with only one prior example known in the bulk phase.122 The coexistence of both the elongated triangular and snub square tessellation in the case of GdI2(bpy)5/2 may suggest similar energies of formation and, thus, provide perspectives for the isolation of intergrown motifs required for quasicrystallinity.
4.4 Frosty frustrated triangles
The formation of Archimedean tessellations is a robust outcome of the assembly of {LnI2}0/+ units with also different ditopic ligands. This is illustrated by the isoreticular EuI2(bpy)5/2 and EuI2(pyz)5/2, both forming the elongated triangular tessellation (Fig. 7(a) and (d)).117 Switching the long bpy ligand for shorter pyz leads to essentially identical coordination at the Eu centre, although modulates the Eu⋯Eu distances, and thus directly impacts the magnetic interaction strengths. The only weakly reducing nature of EuII inhibits any radical character of the ligand scaffold, as confirmed by bond length analysis (Fig. 8), magnetometry, and X-ray absorption spectroscopy. The 4f7 electronic configuration of EuII leads to the largest possible spin magnetic moment and the absence of an orbital angular momentum results in a vanishingly small magnetic anisotropy. These qualities are tantalizing for the development of EuII-based coolant materials for adiabatic demagnetization refrigerators relying on the magnetocaloric effect (MCE).126,127 The MCE occurs due to the change in magnetic entropy of a material when a magnetic field is applied or removed. When the field is applied, the constituent spins of the magnet align, reducing magnetic entropy, and the system releases heat. When the field is removed, entropy increases as the spin moments randomize, absorbing heat and cooling the material. MCE materials operating in the sub-Kelvin regime are touted as alternatives to traditional 3He cooling used, e.g., for cryogenic applications needed in quantum technology and space exploration. In this temperature window, even weak antiferromagnetic interactions are undesired, as they minimize the ground state spin moment and thus the magnetic entropy. Nevertheless, complete eradication of magnetic interactions between electronic spins is only possible in magnetically highly dilute materials, which concurrently will have an only low cooling capacity. The creation of ideal triangular motifs bypasses this obstacle. An equilateral triangular net with pairwise antiferromagnetic interactions exhibits magnetic frustration, and thereby the absence of long-range magnetic order, and a retention of the magnetic entropy extending into the sub-Kelvin domain. Indeed, EuI2(pyz)5/2 fulfills these requirements by featuring triangular motifs with weak antiferromagnetic interactions, on the order of −0.3 cm−1, as suggested by DFT calculations. The specific heat shows a broad maximum in the zero-field cp between ca. 0.7 and 3 K, attributed to 2D intra-layer fluctuations. However, below 0.5 K, there is the onset of a magnetic phase transition, negating the observation of a magnetocaloric effect below this temperature. Despite this, a sizeable MCE is observed with a maximum isothermal magnetic entropy change of 24.3 J kg−1 K−1 for a 7 T field change and at a high temperature of 3 K. To further improve the MCE at sub-K temperature in EuII tessellations, we sought to realize strictly triangular motifs. This situation is fulfilled in the triangular tessellation, featuring 3- or 6-fold symmetric nodes, yet to be observed in a coordination solid. As the ionic radius of EuII is slightly too small to accommodate the required six ligands in plane, the larger BaII was used. This resultant compound, BaI2(pyz)3 (Fig. 7(f)), features sheets with a triangular tessellation, with crystallographic D3d local symmetry of the Ba site.119 Despite the nonexistence of the EuII analogue, performing the synthesis with a mixture of EuI2 and BaI2 leads to alloyed networks, Ba1−xEuxI2(pyz)3, with up to 92% of EuII (hereafter denoted as Ba0.1Eu0.9I2(pyz)3). In the absence (or less than 8%) of templating BaII, EuI2(pyz)3 is instead formed. Here, a different connectivity is observed, with the smaller EuII nodes enjoying a fivefold in-plane coordination (Fig. 7(e)). Structural analysis of Ba0.1Eu0.9I2(pyz)3 is consistent with an EuII oxidation state, and reveals the longest reported unsupported Eu–N bond lengths (2.90 Å). The resultant long Eu⋯Eu distances lead to exceedingly weak exchange interactions, which is reflected by the weak antiferromagnetic J = −0.006 cm−1 obtained from the concerted analysis of magnetometric and heat capacity data obtained down to 0.3 K (Fig. 10(a)). The ensuing magnetic frustration tied to the interactions and the high crystallographic symmetry leads to a magnetocaloric working temperature as low as T = 0.17 K (Fig. 10(b)). This value is remarkably low for a relatively magnetically dense refrigerant and illustrates well how coordination chemistry approaches towards novel MCE materials may challenge the much more established purely inorganic materials.
|
| Fig. 10 (a) The change in magnetic entropy, −ΔSm, at selected magnetic fields for Ba0.1Eu0.9I2(pyz)3, as obtained from the experimental cp and magnetization data, and calculated cp for J/kB = −0.008 K and D/kB = 0.068 K (lines). (b) Demagnetization (0.5 T → 0 T) processes at starting temperatures T0 = 0.58, 0.47 K, and 0.38 K. Inset: Time-dependence of the temperature at T0 = 0.38 K following a linear decrease in the magnetic field strength (grey trace). Reproduced from ref. 119. | |
4.5 Emergence of lanthanide valence swings
As for EuI2(pyz)3, the combination of LnI2 (Ln = Sm, Yb) with pyz leads to the formation of LnI2(pyz)3 (Fig. 7(e)), featuring pseudo-1D chains of {LnI2pyz2} with interdigitating pendant pyz ligands.118 All data for both EuI2(pyz)3 and YbI2(pyz)3 are consistent with the presence of a LnII ion. For SmII, despite the diamagnetic 7F0 ground state, thermal population of the 7F1 state gives rise to a sizable magnetic moment at room temperature (χT ≈ 1.4 cm3 mol−1 K−1), as found in, e.g., SmIII2(CH3CN)5.109 For SmI2(pyz)3, the room-temperature χT product is commensurate with an SmII oxidation state assignment. Upon descending temperature, an abrupt decrease in χT of ∼80% is observed at T ≈ 190 K. Sm–I bond length analysis of the single-crystal X-ray structures of the HT and LT phases suggests the presence of a SmII(pyz0)3 HT valence tautomer and a SmIII{(pyz0)2(pyz˙−)} LT valence tautomer (Fig. 8 and 11(a)). As expected for an SmII → SmIII oxidation, a 4% decrease in Ln–I bond lengths is observed and an overall 6% reduction in the unit cell volume. The structural analysis of the LT phase clearly shows the localization of the pyz˙− moieties in a zigzag chain between the SmIII ions (Fig. 11(a)). Additional evidence of a VT transition is provided by XANES of the HT and LT phases (Fig. 11(c)). The Sm L3 absorption edge resonance shows a shift of 8 eV, consistent with a quantitative and reversible SmII(pyz0)3 ⇄ SmIII{(pyz0)2(pyz˙−)} VT transition, and no indication of intermediate valency. The VT transition is cooperative and has an open thermal hysteresis of 14 K (sweep rate = 0.5 K min−1) and is reversible over at least 100 temperature cycles with only minor loss of conversion (Fig. 11(b)). The electronic ground state of SmIII, 6H5/2 with gJ = 2/7, results in an almost vanishing magnetic moment. As such, the magnetic moment of the LT phase is dominated by the magnetic moment of pyz˙−. The close proximity of the pyz˙− in the zigzag chain results in significant antiferromagnetic radical–radical exchange interactions (J = −76cm−1). The isomorphism of the LnI2(pyz)3 (Ln = Sm, Eu, Yb) series provides the opportunity to tune the VT transition by alloying, as often used in inorganic solids exhibiting VT transitions. Indeed, syntheses involving mixtures of SmI2 and YbI2, in any ratio, yield monophasic Sm1−xYbxI2(pyz)3 solid-solutions. The smaller ionic radius of YbII over SmII, results in a progressive reduction in the HT unit cell volume with increasing Yb content. Introducing only 10% YbII yields an alloy retaining a full VT transition, but at a lower critical temperature (Fig. 12). A further increase in Yb content leads to a further shifting of the VT critical temperature, a softening of the VT transition, and a narrowing of the thermal hysteresis, consistent with decreasing cooperativity. Notably, this is fully analogous to both the situation encountered in SCO coordination polymers, as well as the established Ln valence change transitions in purely inorganic materials. In stark contrast to the purely inorganic solids, however, the VT transition in SmI2(pyz)3 and Sm1−xYbxI2(pyz)3 appears to involve only distinct and integer oxidation states, and in this regard bear closer resemblance to the VT transitions observed in transition metal-based systems.
|
| Fig. 11 (a) HT and LT valence tautomers of the pseudo-1D chains of {Sm(pyz)2} in SmI2(pyz)3. Local coordination environment of the Sm ions shown on the right. (b) Durability of the VT transition as reflected in the χT data obtained by T cycling between 175 K and 225 K (sweep rate = 2 K min−1). (c) Sm L3 XANES spectra of polycrystalline SmI2(pyz)3 measured at 300K, at 44K and again at 300K. Reproduced from Springer Nature (2024).118 | |
|
| Fig. 12 Magnetic susceptibility-temperature product, χT (μ0H = 1.0T; temperature ramping rate, 0.25Kmin−1) of Sm1−xYbxI2(pyz)3 solid solutions, normalized to the samarium content, 1 − x. The grey surface represents the area spanned by the expected χT product for the LT phase and HT phase. The solid black line represents the 50% molar fraction of the LT ⇄ HT phase conversion. Inset: Temperature dependence of Tcversus x, with the best linear fit (grey line, Tc (K) = −3.55x + 202). Reproduced from Springer Nature (2024).118 | |
5. Conclusions and outlook
The self-assembly of trans-{LnI2}0/+ nodes and ditopic linkers, pyrazine and 4,4′-bipyridine, yields a variety of genuinely new chemical motifs. The robust structure-directing nature of the nodes provide rare opportunities for tailoring symmetries in f-element-based materials. The 2D tessellations escape the periodicity-defying pentagonal nodes by tiling Archimedean patterns, that may verge on even quasicrystallinity. Harnessing a novel templation effect, even hexagonal nodes can be realized. The structural symmetry has decisive implications for the electronic structure and ensuing physical properties. This is exemplified by the spin-frustrated Ba0.1Eu0.9I2(pyz)3 network, which escapes magnetic order, even when approaching absolute zero temperature. The wide span of LnIII/LnII reduction potentials and the chemical resemblance allows for the design of isostructural materials hosting either LnII–ligand(0) or LnIII–ligand(˙−) linkages. Despite being well-established in purely inorganic materials, stimuli-induced VT transitions linking the LnII–ligand(0) or LnIII–ligand(˙−) states are far from well-established. SmII and pyrazine are on the brink of electron transfer. The resultant compound, SmI2(pyz)3, is the first coordination solid of the f-block to exhibit an unambiguous and complete VT transition. The phenomenon is robust to structural distortions, although tunable, as illustrated by molecular alloying. The realization of a VT switch in a coordination network opens perspectives for further chemically encoding of cooperative VT events by crystal engineering, as successfully pursued in the field of SCO materials. However, the fundamental advantage to coordination networks featuring VT transitions is the concerted change to the extended electronic structure of the pillaring ligands. Such activation of the otherwise innocent scaffold is currently at the origin of new horizons of metal–organic framework chemistry, such as high electrical conductivity.128,129 We envision a plethora of magnetoelectrical properties, not currently found in molecular chemistry, could be found in the chemical space of valence-fluctuating lanthanide ions and redox-non-innocent scaffolds, and we do not see any reason why these materials would be limited to the metal ion and ligand combinations discussed in this account.
Author contributions
The manuscript was written by K. S. P. and M. A. D., and both authors have given their consent to the publication of the manuscript.
Data availability statement
No primary research results, software or code have been included and no new data were generated or analysed as part of this review.
Conflicts of interest
There are no conflicts to declare.
Acknowledgements
K. S. P. thanks the VILLUM Foundation for a VILLUM Young Investigator+ (42094) grant, the Independent Research Fund Denmark for a DFF-Sapere Aude Starting grant (no. 0165-00073B), and the Carlsberg Foundation for a research infrastructure grant (no. CF17-0637).
References
- O. Sato, Nat. Chem., 2016, 8, 644–656 CrossRef CAS.
- M. A. Halcrow, Chem. Soc. Rev., 2011, 40, 4119–4142 RSC.
- T. Tezgerevska, K. G. Alley and C. Boskovic, Coord. Chem. Rev., 2014, 268, 23–40 CrossRef CAS.
-
G. K. Gransbury and C. Boskovic, Encyclopedia of Inorganic and Bioinorganic Chemistry, John Wiley & Sons Ltd., New Jersey, USA, 2021, pp. 1–24 Search PubMed.
- A. S. Attia and C. G. Pierpont, Inorg. Chem., 1995, 34, 1172–1179 CrossRef CAS.
- A. A. Kamin, I. P. Moseley, J. Oh, E. J. Brannan, P. M. Gannon, W. Kaminsky, J. M. Zadrozny and D. J. Xiao, Chem. Sci., 2023, 14, 4083–4090 RSC.
- J. Chen, Y. Sekine, Y. Komatsumaru, S. Hayami and H. Miyasaka, Angew. Chem., Int. Ed., 2018, 57, 12043–12047 CrossRef CAS PubMed.
- L.-T. Yue, S. O. Shapovalova, J.-S. Hu, M. G. Chegerev, Y.-M. Zhao, C.-D. Liu, M. Yu, A. A. Starikova, A. A. Guda, Z.-S. Yao and J. Tao, Angew. Chem., Int. Ed., 2024, 63, e202401950 CrossRef CAS PubMed.
- Y.-M. Zhao, J. P. Wang, X. Y. Chen, M. Yu, A. A. Starikova and J. Tao, Inorg. Chem. Front., 2023, 10, 7251–7264 RSC.
- K. KC, T. Woods and L. Olshansky, Angew. Chem., Int. Ed., 2023, 62, e202311790 CrossRef CAS.
- S. Wiesner, A. Wagner, E. Kaifer and H.-J. Himmel, Chem. – Eur. J., 2016, 22, 10438–10445 CrossRef CAS.
- L. Greb, Eur. J. Inorg. Chem., 2022, e202100871 CrossRef CAS.
- J. Bendix and K. M. Clark, Angew. Chem., Int. Ed., 2016, 55, 2748–2752 CrossRef CAS PubMed.
- R. M. Buchanan and C. G. Pierpont, J. Am. Chem. Soc., 1980, 102(15), 4951–4957 CrossRef CAS.
- O. Sato, A. Cui, R. Matsuda, J. Tao and S. Hayami, Acc. Chem. Res., 2007, 40(5), 361–369 CrossRef CAS PubMed.
- L. Leroy, T. M. Francisco, H. J. Shepherd, M. R. Warren, L. K. Saunders, D. A. Shultz, P. R. Raithby and C. B. Pinheiro, Inorg. Chem., 2021, 60, 8665–8671 CrossRef CAS PubMed.
- W. H. Xu, Y. B. Huang, W.-W. Zheng, S. Q. Su, S. Kanegawa, S. Q. Wu and O. Sato, Dalton Trans., 2024, 53, 2512–2516 RSC.
- A. Summers, F. Z. M. Zahir, G. F. Turner, M. A. Hay, A. Riboldi-Tunnicliffe, R. Williamson, S. Bird, L. Goerigk, C. Boskovic and S. A. Moggach, Nat. Commun., 2024, 15, 8922 CrossRef CAS PubMed.
- C. Roux, D. M. Adams, J. P. Itié, A. Polian, D. N. Hendrickson and M. Verdaguer, Inorg. Chem., 1996, 35, 2846–2852 CrossRef CAS.
- G. Poneti, M. Mannini, L. Sorace, P. Sainctavit, M. A. Arrio, E. Otero, J. Criginski Cezar and A. Dei, Angew. Chem., Int. Ed., 2010, 122, 1998–2001 CrossRef.
- G. Poneti, L. Poggini, M. Mannini, B. Cortigiani, L. Sorace, E. Otero, P. Sainctavit, A. Magnani, R. Sessoli and A. Dei, Chem. Sci., 2015, 6, 2268–2274 RSC.
- S. Bin-Salamon, S. Brewer, S. Franzen, D. L. Feldheim, S. Lappi and D. A. Shultz, J. Am. Chem. Soc., 2005, 127, 5328–5329 CrossRef CAS PubMed.
- O. Kahn, J. Kröber and C. Jay, Adv. Mater., 1992, 4, 718–728 CrossRef CAS.
- J. Krober, E. Codjovi, O. Kahn, F. Groliere and C. Jay, J. Am. Chem. Soc., 1993, 115, 9810–9811 CrossRef CAS.
- J. Cruddas and B. J. Powell, Inorg. Chem. Front., 2020, 7, 4424–4437 RSC.
- M. Ahmed, K. A. Zenere, N. F. Sciortino, K. S. A. Arachchige, G. F. Turner, J. Cruddas, C. Hua, J. R. Price, J. K. Clegg, F. J. Valverde-Muñoz, J. A. Real, G. Chastanet, S. A. Moggach, C. J. Kepert, B. J. Powell and S. M. Neville, Inorg. Chem., 2022, 61, 6641–6649 CrossRef CAS.
- A. Günther, Y. Deja, M. Kilic, K. Tran, P. Kotra, F. Renz, W. Kowalsky and B. Roth, Sci. Rep., 2024, 14, 5897 CrossRef.
- C. H. Pham and F. Paesani, Inorg. Chem., 2018, 57, 9839–9843 CrossRef CAS.
- P. D. Southon, L. Liu, E. A. Fellows, D. J. Price, G. J. Halder, K. W. Chapman, B. Moubaraki, K. S. Murray, J.-F. Létard and C. J. Kepert, J. Am. Chem. Soc., 2009, 131, 10998–11009 CrossRef CAS PubMed.
- B. Li, L.-Q. Chen, R.-J. Wei, J. Tao, R.-B. Huang, L.-S. Zheng and Z. Zheng, Inorg. Chem., 2011, 50, 424–426 CrossRef CAS PubMed.
- O. Drath, R. W. Gable, B. Moubaraki, K. S. Murray, G. Poneti, L. Sorace and C. Boskovic, Inorg. Chem., 2016, 55, 4141–4151 CrossRef CAS PubMed.
- O. Drath and C. Boskovic, Coord. Chem. Rev., 2018, 375, 256–266 CrossRef CAS.
- B. Li, Y.-M. Zhao, A. Kirchon, J.-D. Pang, X.-Y. Yang, G.-L. Zhuang and H.-C. Zhou, J. Am. Chem. Soc., 2019, 141, 6822–6826 CrossRef CAS.
- A. Lannes, Y. Suffren, J. B. Tommasino, R. Chiriac, F. Toche, L. Khrouz, F. Molton, C. Duboc, I. Kieffer, J.-L. Hazemann, C. Reber, A. Hauser and D. Luneau, J. Am. Chem. Soc., 2016, 138, 16493–16501 CrossRef CAS PubMed.
- E. Evangelio and D. Ruiz-Molina, C. R. Chim., 2008, 11, 1137–1154 CrossRef CAS.
- G. K. Gransbury, M.-E. Boulon, S. Petrie, R. W. Gable, R. J. Mulder, L. Sorace, R. Stranger and C. Boskovic, Inorg. Chem., 2019, 58, 4230–4243 CrossRef CAS PubMed.
- F. Z. M. Zahir, M. A. Hay, J. T. Janetzki, R. W. Gable, L. Goerigk and C. Boskovic, Chem. Sci., 2024, 15, 5694–5710 RSC.
- D. J. R. Brook, J. DaRos, A. Ponnekanti, S. Agrestini and E. Pellegrin, Dalton Trans., 2024, 53, 7536–7545 RSC.
- J. M. D. Coey, Engineering, 2020, 6, 119–131 CrossRef CAS.
- C. A. P. Goodwin, F. Ortu, D. Reta, N. F. Chilton and D. P. Mills, Nature, 2017, 548, 439–442 CrossRef CAS PubMed.
- F.-S. Guo, B. M. Day, Y.-C. Chen, M.-L. Tong, A. Mansikkamäki and R. A. Layfield, Science, 2018, 362, 1400–1403 CrossRef CAS PubMed.
- C. A. Gould, K. R. McClain, D. Reta, J. G. C. Kragskow, D. A. Marchiori, E. Lachman, E.-S. Choi, J. G. Analytis, R. D. Britt, N. F. Chilton, B. G. Harvey and J. R. Long, Science, 2022, 375, 198–202 CrossRef CAS PubMed.
- Z. Zhu and J. Tang, Chem. Soc. Rev., 2022, 51, 9469–9481 RSC.
- A. Fatima, M. W. Ahmad, A. K. A. Al Saidi, A. Choudhury, Y. Chang and G. H. Lee, Nanomaterials, 2021, 11, 2449 CrossRef CAS PubMed.
- A. V. Pavlishchuk and V. V. Pavlishchuk, Theor. Exp. Chem., 2020, 56, 1–25 CrossRef CAS.
- J. C. G. Bünzli, Eur. J. Inorg. Chem., 2017, 5058–5063 CrossRef.
- M. E. Fieser, M. R. MacDonald, B. T. Krull, J. E. Bates, J. W. Ziller, F. Furche and W. J. Evans, J. Am. Chem. Soc., 2015, 137, 369–382 CrossRef CAS PubMed.
- M. R. MacDonald, J. E. Bates, J. W. Ziller, F. Furche and W. J. Evans, J. Am. Chem. Soc., 2013, 135, 9857–9868 CrossRef CAS PubMed.
- K. R. Meihaus, M. E. Fieser, J. F. Corbey, W. J. Evans and J. R. Long, J. Am. Chem. Soc., 2015, 137, 9855–9860 CrossRef CAS.
- M. R. MacDonald, J. E. Bates, M. E. Fieser, J. W. Ziller, F. Furche and W. J. Evans, J. Am. Chem. Soc., 2012, 134, 8420–8423 CrossRef CAS PubMed.
- N. T. Rice, I. A. Popov, D. R. Russo, T. P. Gompa, A. Ramanathan, J. Bacsa, E. R. Batista, P. Yang and H. S. La Pierre, Chem. Sci., 2020, 11, 6149–6159 RSC.
- H. Tateyama, A. C. Boggiano, C. Liao, K. S. Otte, X. Li and H. S. La Pierre, J. Am. Chem. Soc., 2024, 146, 10268–10273 CrossRef CAS PubMed.
- Y. Qiao, H. Yin, L. M. Moreau, R. Feng, R. F. Higgins, B. C. Manor, P. J. Carroll, C. H. Booth, J. Autschbach and E. J. Schelter, Chem. Sci., 2021, 12, 3558–3567 RSC.
- A. Ramanathan, J. Kaplan, D. C. Sergentu, J. A. Branson, M. Ozerov, A. I. Kolesnikov, S. G. Minasian, J. Autschbach, J. W. Freeland, Z. Jiang, M. Mourigal and H. S. La Pierre, Nat. Commun., 2023, 14, 3134 CrossRef CAS.
- A. R. Willauer, C. T. Palumbo, F. Fadaei-Tirani, I. Zivkovic, I. Douair, L. Maron and M. Mazzanti, J. Am. Chem. Soc., 2020, 142, 5538–5542 CrossRef CAS PubMed.
- N. T. Rice, I. A. Popov, D. R. Russo, J. Bacsa, E. R. Batista, P. Yang, J. Telser and H. S. La Pierre, J. Am. Chem. Soc., 2019, 141, 13222–13233 CrossRef CAS.
- A. R. Willauer, I. Douair, A. S. Chauvin, F. Fadaei-Tirani, J. C. G. Bünzli, L. Maron and M. Mazzanti, Chem. Sci., 2022, 13, 681–691 RSC.
- T. P. Gompa, A. Ramanathan, N. T. Rice and H. S. La Pierre, Dalton Trans., 2020, 49, 15945–15987 RSC.
- M. A. Hay and C. Boskovic, Chem. – Eur. J., 2021, 27, 3608–3637 CrossRef CAS PubMed.
- M. Suta and C. Wickleder, J. Lumin., 2019, 210, 210–238 CrossRef CAS.
- C. A. Gould, K. R. McClain, J. M. Yu, T. J. Groshens, F. Furche, B. G. Harvey and J. R. Long, J. Am. Chem. Soc., 2019, 141, 12967–12973 CrossRef CAS PubMed.
- M. Szostak and D. J. Procter, Angew. Chem., Int. Ed., 2012, 51, 9238–9256 CrossRef CAS.
- T. C. Jenks, M. D. Bailey, J. L. Hovey, S. Fernando, G. Basnayake, M. E. Cross, W. Lia and M. J. Allen, Chem. Sci., 2018, 9, 1273–1278 RSC.
- E. A. Boyd, C. Shin, D. J. Charboneau, J. C. Peters and S. E. Reisman, Science, 2024, 385, 847–853 CrossRef CAS PubMed.
- K. C. Nicolaou, S. P. Ellery and J. S. Chen, Angew. Chem., Int. Ed., 2009, 48, 7140–7165 CrossRef CAS.
- K. Gopalaiah and H. B. Kagan, Chem. Rec., 2013, 13, 187–208 CrossRef CAS.
- L. R. Morss, Chem. Rev., 1976, 76, 827–841 CrossRef CAS.
- S. Demir, I. R. Jeon, J. R. Long and T. D. Harris, Coord. Chem. Rev., 2015, 289, 149–176 CrossRef.
- N. Ishikawa, M. Sugita, T. Ishikawa, S.-Y. Koshihara and Y. Kaizu, J. Am. Chem. Soc., 2003, 125, 8694–8695 CrossRef CAS.
- J. D. Rinehart, M. Fang, W. J. Evans and J. R. Long, Nat. Chem., 2011, 3, 538–542 CrossRef CAS PubMed.
- J. D. Rinehart, M. Fang, W. J. Evans and J. R. Long, J. Am. Chem. Soc., 2011, 133, 14236–14239 CrossRef CAS PubMed.
- S. Demir, M. I. Gonzalez, L. E. Darago, W. J. Evans and J. R. Long, Nat. Commun., 2017, 8, 2144 CrossRef PubMed.
- N. Mavragani, D. Errulat, D. A. Gálico, A. A. Kitos, A. Mansikkamäki and M. Murugesu, Angew. Chem., Int. Ed., 2021, 60, 24206–24213 CrossRef CAS.
- N. Bajaj, N. Mavragani, A. A. Kitos, D. Chartrand, T. Maris, A. Mansikkamäki and M. Murugesu, Chem, 2024, 10, 2484–2499 CAS.
- M. Owen, Ann. Phys., 1912, 37, 657–699 CrossRef CAS.
- A. W. Lawson and T.-Y. Tang, Phys. Rev., 1949, 76, 301 CrossRef CAS.
- K. A. Gschneidner Jr. and R. Smoluchowski, J. Less-Common Met., 1963, 5, 374–385 CrossRef.
- A. Sousanis, P. F. Smet and D. Poelman, Materials, 2017, 10, 953 CrossRef.
- J. M. Lawrence and R. D. Parks, J. Phys. Colloq., 1976, 37, C4–249 CrossRef.
- A. Jayaraman, V. Narayanamurti, E. Bucher and R. G. Maines, Phys. Rev. Lett., 1970, 25, 1430 CrossRef CAS.
- M. B. Maple and D. Wohlleben, Phys. Rev. Lett., 1971, 27, 511 CrossRef CAS.
- Y. Haga, J. Derr, A. Barla, B. Salce, G. Lapertot, I. Sheikin, K. Matsubayashi, N. K. Sato and J. Flouquet, Phys. Rev. B, 2004, 70, 220406 CrossRef.
- A. Jayaraman, E. Bucher, P. D. Dernier and L. D. Longinotti, Phys. Rev. Lett., 1973, 31, 700–703 CrossRef CAS.
- S. Ariponnammal and S. Chandrasekaran, J. Nano-Electron. Phys., 2011, 3, 529–535 Search PubMed.
- A. P. Menushenkov, R. V. Chernikov, V. V. Sidorov, K. V. Klementiev, P. A. Alekseev and A. V. Rybina, JETP Lett., 2006, 84, 119–123 CrossRef CAS.
- K. Imura, M. Saito, M. Kaneko, T. Ito, T. Hajiri, M. Matsunami, S. Kimura, K. Deguchi, H. S. Suzuki and N. K. Sato, J. Phys.: Conf. Ser., 2015, 592, 012028 CrossRef.
- S. N. Gilbert Corder, X. Chen, S. Zhang, F. Hu, J. Zhang, Y. Luan, J. A. Logan, T. Ciavatti, H. A. Bechtel, M. C. Martin, M. Aronson, H. S. Suzuki, S.-I. Kimura, T. Iizuka, Z. Fei, K. Imura, N. K. Sato, T. H. Tao and M. Liu, Nat. Commun., 2017, 8, 2262 CrossRef PubMed.
- J. Arvanitidis, K. Papagelis, S. Margadonna, K. Prassides and A. N. Fitch, Nature, 2003, 425, 599–602 CrossRef CAS PubMed.
- S. Margadonna, J. Arvanitidis, K. Papagelis and K. Prassides, Chem. Mater., 2005, 17, 4474–4478 CrossRef CAS.
- J. Arvanitidis, K. Papagelis, S. Margadonna and K. Prassides, Dalton Trans., 2004, 3144–3146 RSC.
- Z. X. Yin, X. Du, W. Z. Cao, J. Jiang, C. Chen, S. R. Duan, J. S. Zhou, X. Gu, R. Z. Xu, Q. Q. Zhang, W. X. Zhao, Y. D. Li, Y.-F. Yang, H. F. Yang, A. J. Liang, Z. K. Liu, H. Yao, Y. P. Qi, Y. L. Chen and L. X. Yang, Phys. Rev. B, 2022, 105, 245106 CrossRef CAS.
- H. Yamaoka, A. Ohmura, N. Tsujii, H. Ishii, N. Hiraoka, H. Sato and M. Sawada, Phys. Rev. B, 2024, 109, 155147 CrossRef CAS.
- M. Hofmann, S. J. Campbell and A. V. J. Edge, J. Magn. Magn. Mater., 2004, 272, E489–E490 CrossRef.
- N. V. Musnikov, Low Temp. Phys., 2015, 41, 946–964 CrossRef.
- Z. E. Brubaker, R. L. Stillwell, P. Chow, Y. Xiao, C. Kenney-Benson, R. Ferry, D. Popov, S. B. Donald, P. Söderlind, D. J. Campbell, J. Paglione, K. Huang, R. E. Baumbach, R. J. Zieve and J. R. Jeffries, Phys. Rev. B, 2018, 98, 214115 CrossRef CAS.
- I. L. Fedushkin, O. V. Maslova, E. V. Baranov and A. S. Shavyrin, Inorg. Chem., 2009, 48, 2355–2357 CrossRef CAS.
- I. L. Fedushkin, O. V. Maslova, A. G. Morozov, S. Dechert, S. Demeshko and F. Meyer, Angew. Chem., Int. Ed., 2012, 51, 10584–10587 CrossRef CAS PubMed.
- D. A. Lukina, A. A. Skatova, R. V. Rumyantcev, S. V. Demeshko, F. Meyer and I. L. Fedushkin, Dalton Trans., 2024, 53, 8850–8856 RSC.
- S. L. Selektor, A. V. Shokurov, V. V. Arslanov, Y. G. Gorbunova, K. P. Birin, O. A. Raitman, F. Morote, T. Cohen-Bouhacina, C. Grauby-Heywang and A. Y. Tsivadze, J. Phys. Chem. C, 2014, 118, 4250–4258 CrossRef CAS.
- A. V. Shokurov, D. S. Kutsybala, A. G. Martynov, A. V. Bakirov, M. A. Shcherbina, S. N. Chvalun, Y. G. Gorbunova, A. Y. Tsidave, A. V. Zaytseva, D. Novikov, V. V. Arslanov and S. L. Selektor, Langmuir, 2020, 36, 1423–1429 CrossRef CAS.
- I. Saiful, M. I. Hossain, K. Katoh, M. Yamashita, R. Arafune, S. M. Fakruddin Shahed and T. Komeda, J. Phys. Chem. C, 2022, 126, 17152–17163 CrossRef CAS.
- M. Tricoire, N. Mahieu, T. Simler and G. Nocton, Chem. – Eur. J., 2021, 27, 6860–6879 CrossRef CAS PubMed.
- C. H. Booth, D. Kazhdan, E. L. Werkema, M. D. Walter, W. W. Lukens, E. D. Bauer, Y.-J. Hu, L. Maron, O. Eisenstein, M. Head-Gordon and R. A. Andersen, J. Am. Chem. Soc., 2010, 132, 17537–17549 CrossRef CAS PubMed.
- M. D. Walter, D. J. Berg and R. A. Andersen, Organometallics, 2006, 25, 3228 CrossRef CAS.
- K. S. Pedersen, P. Perlepe, M. L. Aubrey, D. N. Woodruff, S. E. Reyes-Lillo, A. Reinholdt, L. Voigt, Z. Li, K. Borup, M. Rouzières, D. Samohvalov, F. Wilhelm, A. Rogalev, J. B. Neaton, J. R. Long and R. Clérac, Nat. Chem., 2018, 10, 1056–1061 CrossRef CAS PubMed.
- P. Perlepe, I. Oyarzabal, K. S. Pedersen, P. Negrier, D. Mondieig, M. Rouzières, E. A. Hillard, F. Wilhelm, A. Rogalev, E. A. Suturina, C. Mathonière and R. Clérac, Polyhedron, 2018, 153, 248–253 CrossRef CAS.
- P. Perlepe, I. Oyarzabal, A. Mailman, M. Yquel, M. Platunov, I. Dovgaliuk, M. Rouzières, P. Négrier, D. Mondeig, E. S. Sututina, M.-A. Dourges, S. Bonhommeau, R. A. Musgrave, K. S. Pedersen, D. Chernyshov, F. Wilhelm, A. Rogalev, C. Mathonière and R. Clérac, Science, 2020, 370, 587–592 CrossRef CAS PubMed.
- P. Perlepe, I. Oyarzabal, L. Voigt, M. Kubus, D. N. Woodruff, S. E. Reyes-Lillo, M. L. Aubrey, P. Négrier, M. Rouzières, F. Wilhelm, A. Rogalev, J. B. Neaton, J. R. Long, C. Mathonière, B. Vignolle, K. S. Pedersen and R. Clérac, Nat. Commun., 2022, 13, 5766 CrossRef CAS.
- M. Kubus, L. Voigt and K. S. Pedersen, Inorg. Chem. Commun., 2020, 114, 107819 CrossRef CAS.
- M. N. Bochkarev, I. L. Fedushkin, S. Dechert, A. A. Fagin and H. Schumann, Angew. Chem., Int. Ed., 2001, 40, 3176–3178 CrossRef CAS.
- W. J. Evans, T. S. Gummersheimer and J. W. Ziller, J. Am. Chem. Soc., 1995, 117, 8999–9002 CrossRef CAS.
- G. Heckmann and M. Niemeyer, J. Am. Chem. Soc., 2000, 122, 4227–4228 CrossRef CAS.
- J. R. van den Hende, P. B. Hitchcock, S. A. Holmes, M. F. Lappert, W.-P. Leung, T. C. W. Mak and S. Prashar, J. Chem. Soc., Dalton Trans., 1995, 1427–1433 RSC.
- L. Voigt, M. Kubus and K. S. Pedersen, Nat. Commun., 2020, 11, 4705 CrossRef CAS.
- Z. Xie, K.-Y. Chiu, B. Wu and T. C. W. Mak, Inorg. Chem., 1996, 35, 5957–5958 CrossRef CAS.
- H. Chen, L. Voigt, M. Kubus, D. Mihrin, S. Mossin, R. W. Larsen, S. Kegnæs, S. Piligkos and K. S. Pedersen, J. Am. Chem. Soc., 2021, 143, 14041–14045 CrossRef CAS PubMed.
- H. Chen, A. S. Manvell, M. Kubus, M. A. Dunstan, G. Lorusso, D. Gracia, M. S. B. Jørgensen, S. Kegnæs, F. Wilhelm, A. Rogalev, M. Evangelisti and K. S. Pedersen, Chem. Commun., 2023, 59, 1609–1612 RSC.
- M. A. Dunstan, A. S. Manvell, N. J. Yutronkie, F. Aribot, J. Bendix, A. Rogalev and K. S. Pedersen, Nat. Chem., 2024, 16, 735–740 CrossRef CAS.
-
A. S. Manvell, M. A. Dunstan, D. Gracia, J. Hrubý, M. Kubus, J. McPherson, E. Palacios, H. Weihe, S. Hill, J. Schnack, M. Evangelisti and K. S. Pedersen, ChemRxiv, 2024, preprint, DOI: 10.26434/chemrxiv-2024-swrbm, This content is a preprint and has not been peer-reviewed.
- D. Écija, J. I. Urgel, A. C. Papgeorgiou, S. Joshi, W. Auwärter, A. P. Seitsonen, S. Klyatskaya, M. Ruben, S. Fischer, S. Vijayaraghavan, J. Reichert and J. V. Barth, Proc. Natl. Acad. Sci. U. S. A., 2013, 110, 6678–6681 CrossRef.
- J. J. Oppenheim, G. Skorupskii and M. Dincă, Chem. Sci., 2020, 11, 11094–11103 RSC.
- V. Smetana, S. P. Kelley, A. V. Mudring and R. D. Rogers, Sci. Adv., 2020, 6, eaay7685 CrossRef CAS PubMed.
- T. Gorai, W. Schmitt and T. Gunnlaugsson, Dalton Trans., 2021, 50, 770–784 RSC.
- Y. Zhang, S. Liu, Z.-S. Zhao, Z. Wang, R. Zhang, L. Liu and Z.-B. Han, Inorg. Chem. Front., 2021, 8, 590–619 RSC.
- S. Sun, Y. Zhao, J. Wang and R. Pei, J. Mater. Chem. B, 2022, 10, 9535–9564 RSC.
- F. Guillou, A. K. Pathak, D. Paudyal, Y. Mudryk, F. Wilhelm, A. Rogalev and V. K. Pecharsky, Nat. Commun., 2018, 9, 2925 CrossRef CAS PubMed.
- Z. Mo, Q. Liu, W. Hao, L. Li, H. Xie, Q. Fu, X. Gao and J. Shen, Mater. Today Phys., 2024, 41, 101351 CrossRef CAS.
- T. Wang, J. Lei, Y. Wang, L. Pang, F. Pan, K.-J. Chen and H. Wang, Small, 2022, 18, 2203307 CrossRef CAS.
- C. Li, L. Zhang, J. Chen, X. Li, J. Sun, J. Zhu, X. Wang and Y. Fu, Nanoscale, 2021, 13, 485–509 RSC.
|
This journal is © The Royal Society of Chemistry 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.