DOI:
10.1039/D5QI01477H
(Research Article)
Inorg. Chem. Front., 2025, Advance Article
Magnetic dynamics and exchange coupling in dinuclear lanthanide complexes bridged by naphthoquinone and anthraquinone radicals
Received
12th July 2025
, Accepted 27th August 2025
First published on 28th August 2025
Abstract
The use of the bis(bidentate) quinoid ligands 5,8-dihydroxy-1,4-naphthoquinone (dhnqH2) and 1,5-dihydroxyanthraquinone (dhaqH2) in Ln chemistry has afforded four new dinuclear DyIII complexes, viz., [Dy2(dhnq)(Tp)4] (1), [Dy2(dhaq)(Tp)4] (2), {K(18-crown-6)}[Dy2(dhnq)(Tp)4] (3) and {K(18-crown-6)}[Dy2(dhaq)(Tp)4] (4) (Tp = tris(pyrazolyl)borate). In compounds 1 and 2, the DyIII ions are bridged by the diamagnetic dianionic forms of dhnq2− and dhaq2, while complexes 3 and 4 are bridged by the one-electron reduced, radical forms of these ligands. The presence of the ligand-centered radical has been confirmed by X-ray crystallography and SQUID magnetometry. Alternating current (ac) magnetic susceptibility studies revealed that the relaxation dynamics of 1 and 2 are primarily governed by fast quantum tunneling of magnetization (QTM). Conversely, the radical-bridged complexes 3 and 4 behave as single molecule magnets (SMMs) with energy barriers for the magnetization reversal, Ueff, of 24.17 K and 16.70 K, respectively, in the absence of a direct current (dc) applied field. The strong ferromagnetic Dy–radical interactions, computed using ab initio POLY_ANISO calculations, led to coupling constants of J = +5.0 and +1.2 cm−1 for 3 and 4, respectively, which explains the SMM behavior in these complexes.
Introduction
Single Molecule Magnets (SMMs) are discrete species that exhibit magnetic hysteresis of the pure molecular origin and an effective energy barrier (Ueff) to reversal of the magnetization.1–4 Apart from their electronic structures, which have been extensively studied over the past several decades, the electron-transport properties of these materials are also of paramount interest for many potential technological applications, including high-density data storage,5 molecular electronic devices,6 and quantum computation.7 The syntheses of these materials often rely on self-assembly reactions between paramagnetic metal ions and organic bridging ligands. Lanthanide ions, in particular DyIII and TbIII,8–11 possess remarkably large single-ion anisotropies compared to other paramagnetic ions of the periodic table, attributes that have been highly successful for the design of new SMMs. Indeed, the first Ln SMM was reported in 200312 and since then, a variety of mononuclear13–20 and polynuclear11,21,22 4f metal SMMs have been characterized.
Despite the remarkable progress that has been made in the area of lanthanide metal SMMs, some challenges remain, particularly in the case of polynuclear complexes. Typically, the origin of the magnetic relaxation of these systems arises from the presence of dominant single ion effects rather than the result of a unified spin ground state. This situation is due to the poor radial extension of the 4f orbitals, which limits orbital overlap with bridging ligands, the result of which is weak magnetic exchange.4 In addition, the majority of bridging ligands that have been used are closed-shell ligands that promote very weak interactions between 4f metal ions and favour fast relaxation processes or quantum tunnelling of the magnetization (QTM). To overcome this limitation, the metal-radical approach23 has been pursued, which makes use of open-shell bridging ligands with diffuse spin orbitals that can penetrate the core electron density of the lanthanide ions and achieve strong direct coupling.24–27 Over the last 10 years, considerable effort has been directed at the synthesis of radical-bridged lanthanide complexes with a focus on dinuclear compounds. Indeed, a small but growing number of interesting dinuclear SMMs have been reported,24,27 including the remarkable compound [K(18-crown-6)(THF)2][{[(Me3Si)2N]2(THF)Ln}2(μ–η2:η2-N2)] which exhibits magnetic hysteresis up to 14 K.28
Following the landmark discovery that the N23˙− radical can strongly couple two LnIII centers, attention has shifted toward identifying organic radicals capable of achieving similarly efficient magnetic exchange. A key determinant in this context is the nature of the radical ligand, specifically, the number of atoms over which the unpaired electron is delocalized and the degree of metal-radical overlap. Numerous N-donor radical ligands have since been explored, including bipyrimidyl (bpym),29 tetrapyridylpyrazine (tppz),30 bisbenzimidazole (Bbim),31 hexaazatrinaphthylene (HAN),32 nitronyl nitroxides,27 indigo,33 triazinyl, and tetrazine.34,35 In this vein, our group has focused on developing lanthanide complexes featuring the radical form of the tetrazine-based ligand 3,6-bis(pyridyl)-1,2,4,5-tetrazine (bptz). Using this ligand in 4f metal chemistry afforded a dimetallic radical-bridged complex, [Cp2Co][[Dy(tmhd)3]2(bptz˙−)] (tmhd = 2,2,6,6-tetramethyl-3,5-heptanedionate),36 and a supramolecular metallacyclic triangle, [Dy3(hfac)6(bptz˙−)3] (hfac = 1,1,1,5,5,5-hexafluoro-2,4-pentanedionate).37
While N-donor radicals have dominated the field, O-donor radical ligands remain significantly underexplored. Bis(bidentate) benzoquinoids are a well-known category of organic molecules that undergo redox reactions to generate semiquinoid radicals, which have been successfully employed for the synthesis of many radical-bridged metal complexes. These ligands exhibit rich electronic versatility since they can accommodate a variety of donor atoms, including N, O, and S,38–40 and non-donor atoms or substituents such as Cl, Br, OMe, NO2, and SMe2;41–43 hence, a wide variety of magnetostructural correlations are rendered possible. Indeed, a number of 3d semiquinoid radical-bridged metal complexes have been reported, ranging from discrete dinuclear compounds44–49 to extended chains50 and multidimensional frameworks.51 In 4f chemistry, Ln2 radical-bridged compounds have been reported, bearing the dichloro- and dibromo-2,5-dihydroxy-1,4-benzoquinone radical derivatives.41,42
Motivated by these insights and building on our previous research in radical bridged complexes,36,37,52–55 we decided to extend this work to relatively unexplored bulkier bis(bidentate) quinoid molecules, namely 5,8-dihydroxy-1,4-naphthoquinone (dhnqH2) and 1,5-dihydroxyanthraquinone (dhaqH2) (Scheme 1), to ascertain if the addition of extra aromatic rings on the semiquinoid radical scaffold affects the strength of metal-radical coupling and the SMM behaviour of the resulting compounds. Although the bridging capabilities and redox activity of these ligands were explored in complexes with heavier d-block metal ions, there is currently no crystallographic evidence of the radical form of these organic molecules.56–59
 |
| Scheme 1 Redox states of the ligands used in this work. | |
Herein, we report the syntheses, crystal structures, magnetic properties, and theoretical studies of two dinuclear DyIII complexes [Dy2(dhnq)(Tp)4]·6CH2Cl2 (1·6CH2Cl2) and [Dy2(dhaq)(Tp)4]·6CH2Cl2 (2·6CH2Cl2), and their radical-bridged analogues {K(18-crown-6)}[Dy2(dhnq)(Tp)4]·2THF (3·2THF) and {K(18-crown-6)}[Dy2(dhaq)(Tp)4]·2THF (4·2THF). These results constitute rare examples of O-donor radical-bridged Dy2 complexes and the first examples of structurally characterized metal complexes bridged by a naphthoquinone or an anthraquinone radical.
Results and discussion
Synthetic comments and description of structures
Complexes 1 and 2 were synthesised by reacting DyCl3·6H2O, LH2 (L = dhnq2− and dhnq2−), KTp, and NaOH in a 2
:
1
:
4
:
2 molar ratio in EtOH. The solvent was removed in vacuo, the remaining solid was redissolved in CH2Cl2, and the resulting solution was left undisturbed for a day, which afforded crystallographically suitable crystals of [Dy2(L)(Tp)4] [L = dhnq2− (1) and dhaq2− (2)] in 75–85% yields.
Compounds 1 and 2 crystallize in the triclinic space group P
. Both structures (Fig. 1) feature two crystallographically equivalent [Dy(Tp)2]+ moieties linked together via a doubly deprotonated bridging dhnq2− or dhaq2− ligand for 1 and 2, respectively, with a site of inversion located at the center of the ligand. Each DyIII ion is 8-coordinate, exhibiting a {N6O2} coordination sphere; six coordination sites are occupied by the N atoms of two capping Tp groups, with the remaining two positions being occupied by the O atoms of the bridging ligand. The [Dy(Tp)2]+ moiety exhibits a bent sandwich-type architecture with a B–Dy–B angle of 125.0° in 1 and 128.3° in 2. The bridging dhnq2− and dhaq2− ligands adopt a bis-bidentate binding mode, forming two six-member chelating rings with the metal ions with bite angles of 72.9° and 73.0° and bite distances of the six-member chelate ring of 2.693 Å and 2.698 Å, for 1 and 2, respectively. Within the bridging ligands, the average C–O bond distances are 1.2853 Å and 1.272 Å in 1 and 2, respectively, consistent with reported values for the doubly deprotonated diamagnetic form of dhnq2− and dhaq2− in coordination compounds with 4d metal ions.56–59 The average Dy–O and Dy–N bond distances are similar in both compounds (2.267 Å and 2.514 Å in 1, and 2.267 Å and 2.508 Å in 2). The coordination geometry of the DyIII ions was estimated using the SHAPE60 program and found to be distorted square antiprismatic in both compounds (CShM = 0.99 and 0.76, respectively; Table S2). The metal ions are almost in the same plane with the bridging ligand (0.033 Å) in 1, while in 2 the DyIII ions are displaced by 0.123 Å above/below the mean plane of the planar dhaq2− ligand. The intramolecular Dy⋯Dy separation is 9.078 Å in 1 and 9.441 Å in 2. To the best of our knowledge, compounds 1 and 2 are the first examples of lanthanide complexes bearing the dhnq2− and dhaq2− ligands.
 |
| Fig. 1 Partially labeled representation of 1 (top) and 2 (bottom). Color Scheme: Dy, yellow; N, blue; C, black; B, magenta; O, red. Hydrogen atoms are omitted for the sake of clarity. | |
Chemical reduction of compounds 1 and 2 with one equivalent of KC8 in THF followed by the addition of 18-crown-6 yielded dark green solutions which were layered with Et2O to afford dark green crystals of the one electron reduced species 3 and 4, {K(18-crown-6)}[Dy2(L˙)(Tp)4] [L˙ = dhnq3˙− (3) and dhaq3˙− (4)], in 45–55% yields.
Complex 3 crystallizes in the triclinic space group P
, and complex 4 crystallizes in the monoclinic space group C2/c. In each case (Fig. 2), the asymmetric unit consists of a half-molecule of the anionic complex [Dy2(L˙)(Tp)4]− [L˙ = dhnq3˙− (3) and dhaq3˙− (4)], one half of the {K(18-crown-6)}+ counter cation, and a THF solvent molecule. The anions in 3 and 4 exhibit structures similar to those observed in 1 and 2. The DyIII ions in 3 and 4 are also 8-coordinate, possessing the same {N6O2} coordination environment. In the case of 3, the metal centers exhibit square antiprismatic geometry, while in 4, the coordination geometry of the DyIII ions is best described as distorted triangular dodecahedral (Table S2). The average Dy–O distances are 2.207 Å and 2.208 Å in 3 and 4, respectively. The average Dy–N bond distance is similar in both compounds (2.557 Å in 3, and 2.539 Å in 4). The [Dy(Tp)2]+ moieties exhibit B–Dy–B angles comparable to those observed for the unreduced analogues (122.3° in 3 and 127.0° in 4). The radical bridging ligands dhnq3˙− and dhaq3˙− adopt the same binding mode as their diamagnetic counterparts, with similar bite angles (74.9° and 75.5° for 3 and 4, respectively) and bite distances of the six-member chelate ring (2.683 Å and 2.703 Å, for 3 and 4, respectively). The metal ions are displaced by 0.115 Å or 0.153 Å above/below the mean plane of the bridging dhnq3˙− or dhaq3˙− ligand for 3 and 4, respectively. The intramolecular Dy⋯Dy separation is 9.008 Å in 3 and 9.316 Å in 4.
 |
| Fig. 2 Partially labeled representation of 3 (top) and 4 (bottom). Color Scheme: Dy, yellow; N, blue; C, black; B, magenta; O, red; K, purple. Hydrogen atoms are omitted for the sake of clarity. | |
While there are several structural similarities among complexes 1–4, a detailed comparison of the bond distances in 1–4 has revealed significant differences and provided crucial insight into the oxidation state of the dhnq and dhaq bridging ligands. A comparison of 1 with 3 and 2 with 4 reveals a net increase in the average C–O bond order by 1.7% and 2.3% (Table 1), respectively, indicative of the presence of an additional electron in the molecular orbitals of the ligands. Upon reduction, the average C–C bond distances remain unchanged within error, but it is worth noting that these values are influenced by the presence of crystallographic inversion symmetry, which complicates their interpretation.61 Furthermore, in both 3 and 4, there is a notable 2.6% decrease in the average Dy–O bond distance. This decrease signifies a stronger metal–ligand interaction, in line with the enhanced donating capability of the reduced dhnq3˙− or dhaq3˙− ligands compared to their dianionic counterparts. To the best of our knowledge, compounds 3 and 4 are the first examples of lanthanide radical-bridged complexes bearing the dhnq3˙− and dhaq3˙− ligands.
Table 1 Selected structural parameters: average distances (Å), angles (°), and coordination geometries for the 8-coordinate DyIII ions in compounds 1·6CH2Cl2, 2·6CH2Cl2, 3·2THF and 4·2THF
|
Diamagnetic ligand |
Radical ligand |
Complex |
l·6CH2Cl2 |
2·6CH2Cl2 |
3 BHF |
4 BHF |
B–Dy–B angle |
125.0 |
128.3 |
122.3 |
127.0 |
Dy–L bite angle |
72.9 |
73.0 |
74.9 |
75.5 |
Dy bite distance |
2.639 |
2.698 |
2.683 |
2.703 |
Dy–O distance |
2.267 |
2.267 |
2.207 |
2.208 |
Dy–N distance |
2.514 |
2.508 |
2.557 |
2.539 |
C–O distance |
1.284 |
1.272 |
1.306 |
1.302 |
Dy–L plane distance |
0.033 |
0.123 |
0.115 |
0.153 |
Dy–Dy intramolecular distance |
9.078 |
9.441 |
9.008 |
9.316 |
Dy coordination geometry |
Square antiprismatic |
Square antiprismatic |
Square antiprismatic |
Triangular dodecahedral |
Magnetic studies
Static properties
Direct current (dc) magnetic susceptibility measurements were conducted on powdered polycrystalline samples of 1–4 using a magnetic field of 0.1 T over a temperature range of 2–300 K. The data are displayed as χMT versus T plots in Fig. 3. For compounds 1 and 2, the experimental χMT values (28.20 and 28.03 cm3 K mol−1, for 1 and 2, respectively) at 300 K are in excellent agreement with the theoretical value (28.34 cm3 K mol−1) expected for two non-interacting DyIII (6H15/2, S = 5/2, L = 5, g = 4/3) ions.62 Upon cooling, the χMT product for complexes 1 and 2 decreases smoothly from 300 K to a value of 27.00 cm3 K mol−1 for 1 and 28.25 cm3 K mol−1 for 2 at 100 K. For 1, a more pronounced decrease in the χMT product is observed below 100 K, reaching a minimum value of 22.67 cm3 K mol−1 at 2 K. Compound 2 exhibits a similar trend, but only down to 4 K. Below this temperature, there is a slight increase in χMT for 2, reaching a value of 26.05 cm3 K mol−1 at 2 K. The decrease observed below 100 K, for 1 and 2, can be attributed to the depopulation of the excited Stark sublevels of the DyIII ions and/or weak antiferromagnetic interactions between the metal centers, while the slight low-temperature increase observed for 2 may indicate weak ferromagnetic coupling between intramolecular DyIII ions.63
 |
| Fig. 3 Plots of χMT versus T for 1–4. | |
A pronounced disparity in magnetic behavior is observed for the radical-bridged analogues 3 and 4. In both cases, the experimental χMT values (29.62 cm3 K mol−1 for 3, and 29.67 cm3 K mol−1 for 4) at 300 K are in agreement with the theoretical value (28.72 cm3 K mol−1) predicted for two non-interacting DyIII (6H15/2, S = 5/2, L = 5, g = 4/3) ions and an S = 1/2 organic radical. For 3, the χMT product exhibits a slight increase as the temperature is decreased, reaching a value of 32.31 cm3 K mol−1 at 28 K. Below this temperature, the χMT product increases more sharply, peaking at a value of 51.70 cm3 K mol−1 at 4 K. After reaching this maximum, the χMT product falls to a minimum value of 49.17 cm3 K mol−1 as the temperature is lowered to 2 K. The χMT product for 4 decreases slightly to a minimum value of 29.53 cm3 K mol−1 at 28 K before increasing to a maximum value of 47.28 cm3 K mol−1 at 2 K. The sudden increase of χMT product observed below 28 K, for 3 and 4, can be attributed to the spin alignment of the DyIII ions induced by the strong magnetic interactions between the dhnq3˙− or dhaq3˙− radicals and the metal centers. These results align with observations made for other radical-bridged lanthanide complexes. The decrease in χMT at lower temperatures for 3 can be ascribed to magnetic blocking.41,42
Field-dependent magnetization measurements were performed on 1–4 at 2, 5, and 7 K temperatures over the range of 0–7 T (Fig. S1–S4). For 1–4, magnetization shows a relatively rapid increase at low fields without reaching saturation at ∼7 T, which indicates significant magnetic anisotropy. Furthermore, the observed magnetization values for the reduced complexes, 3 and 4, at 2 K and 7 T are slightly lower compared to those of compounds 1 and 2. This difference is more likely a consequence of magnetic anisotropy effects, rather than of antiferromagnetic coupling between the metal ions and radical ligands in 3 and 4.
Dynamic properties
Alternating current (ac) magnetic susceptibility measurements were conducted using a 2 Oe ac field to investigate the magnetic dynamics of 1–4. Compounds 1 and 2, which feature the diamagnetic form of the dhnq2− and dhaq2− ligands, respectively, displayed both in-phase (χ′) and out-of-phase (χ′′) ac susceptibility signals that are frequency and temperature dependent in the absence of an applied dc field (Fig. S5 and S6).
However, within the frequency range of 1–1000 Hz and temperatures spanning from 2 to 25 K, no maxima were detected in the χ′′ signals, suggesting considerable quantum tunnelling of the magnetization (QTM). Additionally, ac susceptibility measurements were performed at various static fields (0–2000 Oe) (Fig. S7 and S8), but they did not improve the SMM properties of 1 and 2. Due to the pronounced QTM, Ueff could not be determined in either case.
In stark contrast, the reduced complexes, 3 and 4, exhibit out-of-phase (χ′′) ac susceptibility peaks that shift to lower frequencies as the temperature decreases, indicative of slow magnetic relaxation (Fig. 4a,b, and Fig. S9–S14). The experimental data between 1.8–4 K and 1.8–3.6 K, for 3 and 4, respectively, were used for the construction of the Cole–Cole plots for complexes 3 and 4 (Fig. S11 and S14). Fitting of the Cole–Cole plots for 3 and 4 using a generalized Debye model64,65 was performed, allowing for the extraction of the τ and α parameters. The α values are in the range of 0.19–0.08 for 3 and 0.17–0.03 for 4, in accordance with a single temperature-dependent relaxation process. The relaxation times (τ) obtained from the above fitting were plotted vs. 1/T (Fig. 4c) and were analyzed using the following eqn (1):
|
 | (1) |
where
τQTM−1,
CTn, and
τ0−1![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif)
exp(−
Ueff/
kBT) represent QTM, Raman, and Orbach relaxation processes, respectively. For
3, the best fitting gave
Ueff = 24.17 K,
τ0 = 3.00 × 10
−8 s,
C = 2.53 × 10
−2 s
−1 K
−6.3 and
n = 6.3, and for
4 Ueff = 16.70 K,
τ0 = 4.04 × 10
−7 s,
C = 1.27 × 10
−1 s
−1 K
−4.0 and
n = 4.0.
 |
| Fig. 4 Frequency dependence of the out-of-phase (χ′′) signals of (a) 3 and (b) under 0 Oe; solid lines are guides for the eye. (c) Comparison of the Arrhenius plots of 3 and 4. Solid lines highlight the fits to eqn (1) (see the text for the fit parameters). | |
These findings indicate that the presence of the radical on the ligand in 3 and 4 effectively suppresses the quantum tunnelling of magnetization prevalent in the unreduced compounds 1 and 2, in accord with other reports for radical-bridged lanthanide complexes. The obtained Ueff values for 3 and 4 are generally lower than those reported for most N-donor radical-bridged Dy2 complexes. However, as compared to O-donor radical-bridged Dy2 complexes, the Ueff values for 3 and 4 are slightly higher than the value reported for the bromoanilate radical (Ueff = 10.4 K, Hdc = 0 Oe)33 and comparable with those reported for the chloroanilate radical (Ueff = 31 K), which were measured in the presence of an external field (Hdc = 1000 Oe).32
Computational studies
To investigate the observed magnetic behavior and the nature of magnetic anisotropy of each DyIII center in 1–4, ab initio CASSCF/RASSI/SINGLE_ANISO calculations using the MOLCAS 8.066 (see computational details) were performed. The magnetic anisotropy direction of each DyIII ion in 1 and 2 is depicted in Fig. 5 (a similar anisotropy direction is observed in 3 and 4). The calculated gz values for both DyIII sites in 1–4 have similar values, signifying the same type of single-ion anisotropy behavior for both DyIII sites (Table 2 and Table S3). Both DyIII ions in 1–4 exhibit small transverse components (gx and gy) and a strong axial component (gz) in their ground Kramers doublet (KD) states. This suggests a low chance of mixing between the mJ states, which may suppress quantum tunneling of the magnetization (QTM), and supports the presence of optimal Ising-type anisotropy required for slow magnetization relaxation. Thus, at the single-ion level, their capacity to blockade the magnetization is unlikely to happen in the ground KDs, since the mJ = ±15/2 was found to be the ground state, but it could possibly occur in the excited KDs; this implies that the compounds are likely to show SMM behavior at zero dc field. Although reduction of 1 and 2 to the radical-bridged analogues 3 and 4 introduces modest changes in Dy–O distances (and coordination geometry in 4), CASSCF/RASSI/SINGLE_ANISO calculations show that the Dy single ion anisotropy remains essentially the same across the series, indicating that the enhanced SMM behavior of the radical-bridged complexes 3 and 4 arises primarily from Dy–radical exchange, with coordination changes acting only as a secondary tuner.
 |
| Fig. 5 Orientations of the local magnetic moments in the ground doublet of each DyIII center in (top) 1 and (bottom) 2. The blue arrows show the magnetic anisotropy directions of DyIII ions in the ground states. Similar magnetic anisotropy directions were observed for their respective radical-bridged complexes 3 and 4. | |
Table 2 Ab initio computed energies of the lowest Kramers Doublets (KDs) and ground-state g-tensors of each DyIII centers in all complexes
KDs |
1 (dhnq-neutral) |
2 (dhaq-neutral) |
3 (dhnq-radical) |
4 (dhaq-radical) |
Dy1 |
Dy2 |
Dy1 |
Dy2 |
Dy1 |
Dy2 |
Dy1 |
Dy2 |
1 |
0.0 |
0.0 |
0.0 |
0.0 |
0.0 |
0.0 |
0.0 |
0.0 |
2 |
148.1 |
148.1 |
149.7 |
149.7 |
154.4 |
154.4 |
157.8 |
157.8 |
3 |
195.4 |
195.4 |
261.9 |
261.8 |
241.4 |
241.4 |
248.6 |
248.6 |
4 |
226.1 |
226.1 |
313.6 |
313.6 |
318.9 |
318.9 |
326.2 |
326.3 |
5 |
255.7 |
255.7 |
329.7 |
329.7 |
338.8 |
338.8 |
389.9 |
390.0 |
6 |
286.7 |
286.7 |
346.4 |
346.4 |
383.3 |
383.3 |
428.3 |
428.4 |
7 |
330.5 |
330.5 |
378.7 |
378.7 |
464.8 |
464.8 |
493.1 |
493.2 |
8 |
641.6 |
641.6 |
697.3 |
697.3 |
740.2 |
740.2 |
740.7 |
740.7 |
gx |
0.0473 |
0.0471 |
0.0138 |
0.0138 |
0.0690 |
0.0688 |
0.0367 |
0.0367 |
gy |
0.0867 |
0.0868 |
0.0216 |
0.0217 |
0.0441 |
0.0437 |
0.0691 |
0.0691 |
gz |
19.7155 |
19.7162 |
19.7806 |
19.7894 |
19.4401 |
19.4396 |
19.5550 |
19.5653 |
In 1 and 2, the ground state gzz axis aligns with the O-atoms of the dhnq2− and dhaq2− ligands (see Fig. 5). This is likely owing to the shorter Dy–O (avg.) bond lengths (∼2.27 Å in 1 and ∼2.25 Å in 2) of dhnq2− and dhaq2− ligands compared to the longer Dy–N distances (∼2.5 Å) of Tp ligands. The oblate electron density of the DyIII ion induces the gzz axis to preferentially align in the same plane towards the shortest Dy–O bond (O atom of dhnq2− and dhaq2− ligands), avoiding significant electrostatic repulsion with the other coordinated atoms, resulting in a greater ground-first excited state gap in all complexes. Similar magnetic anisotropy directions have been observed for 3 and 4, as shown in Fig. 5.
The computed energies of the eight low-lying KDs (Table 2) for both DyIII sites have the same energy values, a reflection of the fact that they are symmetrically equivalent67 and are found to span an energy gap of 641.6, 697.3, 740.2, and 740.7 cm−1 for 1–4, respectively. To validate the calculated energy barriers, we developed relaxation mechanisms for magnetization blockade for both DyIII sites in 1–4 (Fig. 6 and Fig. S15). The ground state KDs of both DyIII sites in 1–4 have small QTM (0.02μB for 1, 0.006μB for 2, 0.03μB for 3, and 0.02μB for 4), hence, it allows the magnetization to relax via higher excited states. The Crystal-Field (CF, Bkq) parameters explain the QTM probability of ground-state KDs. The computed CF parameters for 1–4 are listed in Table S5. In both DyIII sites in 1–4, significant axial terms (in which q = 0 and k = 2, 4, 6) are computed compared to non-axial terms (in which q ≠ 0 and k = 2, 4, 6), especially the axial terms are dominating in k = 2 terms, suggesting negligible QTM effects.67–70 A significant TA-QTM (temperature assisted – QTM) in the first excited states was observed (0.83μB for 1, 0.1μB for 2, 0.34μB for 3, and 0.2μB for 4), which limits the relaxation of magnetization via these states with energy barriers of 148.1, 149.7, 154.2, and 157.8 cm−1 (Fig. 6) for 1–4, respectively. Single-ion calculation results indicate that the radical-bridged complexes 3 and 4 have relatively large energy barriers compared to their respective neutral counterparts (1 and 2). However, the conclusive relaxation mechanisms in 1–4 are best understood in terms of exchange-coupled systems.
 |
| Fig. 6 Ab initio computed magnetization blocking barrier for Dy1 site (same applicable for Dy2 site) in (top) 1, and (bottom) 2. The thick blue line indicates the Kramers doublets (KDs) as a function of computed magnetic moment. The red arrows represent the presence of QTM/TA-QTM between the connecting pairs. The green/purple arrows show the possible pathway via Orbach/Raman relaxation. The numbers provided at each arrow are the mean absolute value for the corresponding matrix element of the transition magnetic moment. | |
To understand the overall SMM behavior of 1–4 and the effect of Dy–radical coupling on magnetic relaxation, the exchange coupled states relaxation mechanism was developed by considering the magnetic coupling constants (Jexch + Jdip = Jtot) between DyIII–rad (only in 3 and 4) and DyIII–DyIII, as well as an intermolecular coupling constant zJ using the POLY_ANISO program71 within the Lines model.72 The Lines model in POLY_ANISO treats DyIII ions as isotropic spins with S = 5/2, which is a simplification used for exchange fitting. The full anisotropy is retained from the ab initio CASSCF/RASSI calculations. The computed magnetic coupling constants are tabulated in Table 3. Using the lowest energy states of individual DyIII ions and the possible exchange interactions, the ab initio calculated magnetic susceptibility data (Fig. 3; solid lines) were computed and are consistent with the experimental data, indicating that the retrieved J values and the Ucal parameters are reliable. These calculations reveal a set of exchange split states (Fig. 7), indicating the formation of coupled spin manifolds due to the bridging radicals in 3 and 4. These states are well isolated from higher excited states and show suppressed QTM in the ground levels for 3 and 4. The observed χMT product increase at low temperatures and magnetization values (relative to 1 and 2) is consistent with strong Dy–radical exchange interactions. Thus, the relaxation likely proceeds via thermally activated processes within the exchange-coupled ladder, distinct from the tunneling-dominated mechanisms operative in 1 and 2. The DyIII–DyIII interaction is found to be very weakly antiferromagnetic in 1 and 2, whereas it is found to be ferromagnetic in 3 and 4 and larger in magnitude than that of JDy–Dy. The presence of ferromagnetic DyIII–DyIII and DyIII–rad interactions in 3 and 4 is attributed to the presence of vacant antibonding orbitals of dhnq and dhaq ligands, which facilitate increased exchange; such an increase in magnitude was previously observed for other LnIII–rad complexes.24,47 The JDy–rad of 3 is large, with a value of 5.0 cm−1 is among the highest ferromagnetic coupling constants reported for a Ln–rad system.24,27 Ferromagnetic Ln–rad coupling is uncommon, and to our knowledge, it has been observed primarily in systems containing nitronyl nitroxide radicals.27 We found that the Jtot of the DyIII–rad interactions in 3 and 4 primarily arises from the Jexch (4.8 cm−1 for 3 and 1.0 cm−1 for 4) with a comparatively small Jdip. In general, the DyIII–DyIII exchange interaction is expected to be very weak in dinuclear or polynuclear complexes that significantly reduces the energy of low-lying doublet states of DyIII ions, resulting in weak or no SMM behavior in a few Dy-based complexes.11,73 This hypothesis is clearly operative in the cases of 1 and 2 where the presence of large tunneling gaps (see Tables S6, S7 and Fig. S16) between the ground exchange coupled states (non-Kramer doublet states) allow the magnetization to relax in the ground state itself which suggest no SMM behvaiour for 1 and 2. However, the presence of ferromagnetic DyIII–rad interactions in 3 and 4 produces several low-lying exchange-coupled states and quenches the QTM/TA-QTM in the ground and first excited exchange-coupled states (KDs) to some extent. The TA-QTM becomes larger in the third and fourth exchange-coupled states, which lead to SMM behaviour with an energy barrier of 19.1 cm−1 (27.5 K) and 10.2 cm−1 (14.7 K), respectively (see Tables S8, S9, and Fig. 7). These Ucal parameters are consistent with the experimentally determined barrier of 24.2 K and 16.7 K for 3 and 4, respectively.
 |
| Fig. 7 Low-lying exchange spectra in (top) 3, and (bottom) 4. The exchange-coupled states are placed on the diagram according to their magnetic moments (bold blue lines). The red arrows show the QTM/TA-QTM between the connecting states (KDs), while the green/purple arrows show the possible spin–phonon transitions. The numbers provided at each arrow are the mean absolute value for the corresponding matrix element of the transition magnetic moment. | |
Table 3 POLY_ANISO fitted magnetic exchange interactions (cm−1, Jexch + Jdip = Jtot) in 1–4
|
Jexch |
Jdip |
Jtot |
zJ |
JDy–Dy |
1 (dhnq-neutral) |
−0.01 |
−0.02 |
−0.03 |
−0.001 |
2 (dhaq-neutral) |
0.005 |
−0.015 |
−0.01 |
0.01 |
3 (dhnq-radical) |
0.58 |
−0.03 |
0.55 |
−0.002 |
4 (dhaq-radical) |
0.62 |
−0.02 |
0.6 |
0.008 |
JDy–rad |
3 (dhnq-radical) |
4.8 |
0.2 |
5.0 |
−0.002 |
4 (dhaq-radical) |
1.0 |
0.2 |
1.2 |
0.008 |
The majority of radical-bridged SMMs utilize anionic radical bridging ligands that align perpendicularly to the primary anisotropy axis of each metal ion.29,74 For example, in the [(Cp2Me4Tb)2(μ–η2: η2-N2)], it is evident that the presence of the radical spin perpendicular to the local magnetic anisotropy axes caused by the Cp* ligands is harmful to the magnetic anisotropy at the TbIII and DyIII sites.74 Based on these results, it was predicted that it would be advantageous for the spin of the radical bridge to be collinear with the local anisotropy axes, as this arrangement would facilitate ferromagnetic coupling between the metal ions through space and ensure that the anionic radical bridge enhances the existing magnetic anisotropy rather than diminishing it.4 This theory has been fulfilled in our study: the dianionic or trianionic napthaquinone and anthraquinone ligands, together with improved delocalization of the radical spin on the ligand surface, facilitate a co-parallel orientation between the radical spin and the gzz axis, which promotes the observed ferromagnetic exchange. A similar scenario was reported for the {CpiPr5LnI3LnCpiPr5} complex, where the unpaired electron residing in the σ-type Ln–Ln bond is co-parallel to the local magnetic anisotropy axes determined by the CpiPr5 ligands, increasing the magnetic anisotropy.20
Conclusions
The first dinuclear Dy2 compounds, 1 and 2, featuring dianions of the quinoid ligands 5,8-dihydroxy-1,4-naphthoquinone (dhnqH2) and 1,5-dihydroxyanthraquinone (dhaqH2) as bridges, are reported. These molecules were further reduced to produce the radical-bridged Dy2 species, 3 and 4, containing the dhnq3˙− and dhaq3˙− radicals, respectively. These are the first, crystallographically characterized, compounds that contain these radical ligands coordinated to a paramagnetic metal ion. Compounds 1–4 represent rare examples of pairs of Dy2 compounds with both neutral and radical forms of the same organic bridging ligands, allowing the investigation of the influence of the exchange coupling on the magnetic dynamics. Indeed, magnetic and theoretical studies revealed that in 1 and 2, the coupling is negligible, resulting in weak SMM behaviour (tails of χ′′ signals) and fast magnetic relaxation through QTM. In contrast, placing an additional electron on the bridging ligands in 3 and 4 led to a strong ferromagnetic exchange, which mitigates the occurrence of QTM observed in the unreduced species. Compounds 3 and 4 exhibit maxima of the χ′′ signals below 4 K under zero static field, indicative of SMM behaviour. Fitting of the data, considering all possible relaxation pathways, gave energy barriers of 24.17 K and 16.70 K for 3 and 4, respectively, for the thermally-activated relaxation. The presence of ferromagnetic coupling between the DyIII ions and the radical, with a coupling constant of +5.0 and +1.2 cm−1 for 3 and 4, was confirmed by the Lines model's POLY_ANISO calculations. The computed Ucal parameters are in excellent agreement with the experimentally determined barriers. Current efforts are underway to isolate other anisotropic LnIII analogues (i.e., LnIII = TbIII, ErIII, and HoIII) as well as to extend this synthetic strategy to the use of other supporting ligands, which are known to promote a strong axial crystal field for oblate LnIII ions.
Experimental
Synthetic procedures
For compounds 1 and 2, all manipulations were performed under aerobic conditions using reagent-grade materials and solvents as received. For 3 and 4, all manipulations were carried out under an inert atmosphere of N2 using standard Schlenk and glovebox techniques. Compounds 1 and 2 were dried under vacuum at 100 °C and stored in the glovebox prior to use. Anhydrous THF lacking butylated hydroxytoluene as an inhibitor was purchased from Sigma Aldrich and stored under an inert atmosphere.
[Dy2(dhnq)(Tp)4]·6CH2Cl2 (1·6CH2Cl2). To a stirred, beige/brown suspension of dhnqH2 (0.19 g, 1.0 mmol), NaOH (0.08 g, 2.0 mmol), and KTp (1.01 g, 4.0 mmol) in EtOH (80 mL) was added solid DyCl3·6H2O (0.75 g, 2.0 mmol). The resulting dark blue solution was stirred for a further 2 h, after which time the solvent was removed in vacuo, and the crude material was extracted using 30 mL of CH2Cl2. The extract was filtered, and the solution was layered with Et2O (30 mL). Slow diffusion after 1 day led to blue needles of 1; these were collected by filtration and washed with Et2O (3 × 5 mL); the yield was 85%. Upon dryness (under vacuum), the crystalline solid was analyzed as 1. Anal. calc. for C46H44N24B4O4Dy2 (1): C, 40.47; H, 3.25; N, 24.62%. Found: C, 40.43; H, 3.21; N, 24.67%. Selected ATR data (Nujol mull, cm−1): 1565 (m), 1500 (w), 1403 (w), 1300 (m), 1269 (m), 1213 (m), 1148 (w), 973 (m), 922 (w), 859 (w), 807 (w), 757 (m), 722 (s), 668 (m), 620 (m), 611 (w).
[Dy2(dhaq)(Tp)4]·6CH2Cl2 (2·6CH2Cl2). To a stirred, beige/brown suspension of dhaqH2 (0.24 g, 1.0 mmol), NaOH (0.08 g, 2.0 mmol), and KTp (1.01 g, 4.0 mmol) in EtOH (80 mL) was added solid DyCl3·6H2O (0.75 g, 2.0 mmol). The resulting dark red solution was stirred for a further 2 h, after which time the solvent was removed in vacuo, and the crude material was extracted using 30 mL of CH2Cl2. The extract was filtered, and the solution was layered with Et2O (30 mL). Slow diffusion after 1 day led to red needles of 2; these were collected by filtration and washed with Et2O (3 × 5 mL); the yield was 75%. Upon dryness (under vacuum), the crystalline solid was analyzed as 2. Anal. calc. for C50H46N24B4O4Dy2 (2): C, 42.43; H, 3.28; N, 23.75%. Found: C, 42.38; H, 3.31; N, 23.79%. Selected ATR data (Nujol mull, cm−1): 1733 (w), 1603 (s), 1584 (s), 1617 (s), 1502 (s), 1404 (m), 1300 (s), 1257 (s), 1212 (s), 1199 (m), 1167 (m), 1147 (w), 1120 (s), 1093 (w), 1048 (s), 976 (s), 922 (w), 887 (w), 864 (w), 864 (w), 781 (m), 750 (s), 722 (s), 671 (m), 636 (w), 621 (m), 572 (m), 474 (w).
{K(18-crown-6)}[Dy2(dhnq)(Tp)4]·2THF (3·2THF). In a nitrogen-filled glovebox, crystals of [Dy2(dhnq)(Tp)4] (1) (0.10 g, 0.07 mmol) were dissolved in THF (10 mL) to yield a dark blue solution. Solid KC8 (0.01 g, 0.07 mmol) was then added, and the color of the solution turned dark green. The mixture was stirred at room temperature for 1 h, after which time solid 18-crown-6 (0.04 g, 0.15 mmol) was added and the solution was filtered. The filtrate was layered with Et2O (10 mL). Slow mixing after 1 day produced dark green platelets of 3; these were collected by filtration and washed with Et2O (3 × 5 mL); the yield was 45%. Upon dryness (under vacuum), the crystalline solid was analyzed as 3. Anal. calc. for C58H68N24B4O10KDy2 (3): C, 41.75; H, 4.11; N, 20.15%. Found: C, 41.79; H, 3.98; N, 20.11%. Selected ATR data (Nujol mull, cm−1): 1542 (m), 1495 (s), 1398 (m), 1351 (w), 1300 (s), 1266 (s), 1215 (s), 1199 (m), 1114 (m), 1099 (s), 1043 (s), 972 (m), 953 (s), 920 (w), 894 (w), 832 (m), 806 (w), 774 (s), 752 (s), 722 (s), 667 (m), 623 (m), 533 (w), 499 (w).
{K(18-crown-6)}[Dy2(dhaq)(Tp)4]·2THF (4·2THF). In a nitrogen-filled glovebox, crystals of [Dy2(dhaq)(Tp)4] (2) (0.10 g, 0.07 mmol) were dissolved in THF (10 mL) to yield a dark red solution. Solid KC8 (0.01 g, 0.07 mmol) was then added, and the color of the solution turned dark green. The mixture was stirred at room temperature for 1 h, after which time solid 18-crown-6 (0.04 g, 0.15 mmol) was added, and the solution was filtered. The filtrate was layered with Et2O (10 mL). Slow diffusion after 1 day led to isolation of dark green plates of 4; these were collected by filtration and washed with Et2O (3 × 5 mL); the yield was 55%. Upon dryness (under vacuum), the crystalline solid was analyzed as 4. Anal. calc. for C70H86N24B4O10KDy2 (4): C, 43.33; H, 4.10; N, 19.56%. Found: C, 43.29; H, 3.06; N, 19.51%. Selected ATR data (Nujol mull, cm−1): 1563 (w), 1535 (w), 1349 (m), 1300 (s), 1251 (m), 1198 (m), 1187 (m), 1149 (w), 1108 (s), 1080 (m), 1042 (s), 972 (m), 960 (m), 922 (w), 855 (m), 796 (w), 780 (m), 737 (m), 723 (s), 704 (w), 670 (m), 622 (m), 585 (m), 486 (w).
Crystallography
Diffraction data for complexes 1·6CH2Cl2, 2·6CH2Cl2, 3·2THF, and 4·2THF were collected on a Bruker D8 diffractometer (Mo Kα graphite-monochromated radiation, λ = 0.71073 Å) with the acquisition controlled by the APEX2 software package.75 The data collection temperature, 150(2) K, was controlled by an Oxford Cryosystems Series 700. Images were processed with the software SAINT+,76 and absorption effects were corrected with the multi-scan method implemented in SADABS.77 The structure was solved using SHELXTL incorporated in the Bruker APEX-III software package and refined using the SHELXLE.78–80 All non-hydrogen atoms were successfully refined using anisotropic displacement parameters. H-atoms bound to carbon atoms were placed at geometrical positions using the suitable HFIX instructions in SHELXL and included in subsequent refinement cycles in riding-motion approximation with isotropic thermal displacement parameters (Uiso) fixed at the carbon atom to which they are attached. H-atoms associated with the coordinated and non-coordinated MeOH molecules were clearly visible in the difference Fourier maps and included in subsequent refinement stages with the O–H distances restrained to 0.85(2) Å and by using a riding-motion approximation with an isotropic thermal displacement parameter (Uiso) fixed at 1.5 × Ueq of the parent O-atom.
Figures of the structures were created using Mercury81 and Diamond82 software packages. The unit cell parameters, structure solution, and refinement details of 1·6CH2Cl2, 2·6CH2Cl2, 3·2THF, and 4·2THF are summarized in Table S1. Further crystallographic details can be found in the corresponding CIF files provided in the SI.
Physical studies
Infrared spectra were obtained using a Nicolet Nexus 470 FT-IR Spectrometer. Elemental analyses were performed by Atlantic Microlabs, Inc., Norcross, GA. Variable-temperature direct current (dc) and alternating current (ac) magnetic susceptibility data were collected on a Quantum Design MPMS-XL SQUID magnetometer equipped with a 7 T magnet and operating in the 2–300 K range. The diamagnetic contribution of the polypropylene bag used to hold the sample was subtracted from the raw data. Pascal's constants83 were used to estimate the diamagnetic corrections, which were subtracted from the experimental susceptibilities to give the molar paramagnetic susceptibilities (χM). Note: All physicochemical characterizations, except X-ray crystallography, were performed on dried samples. Lattice solvents were removed during this process, and elemental analysis confirmed the desolvated forms of the compounds 1–4.
Computational details
Using MOLCAS 8.0,84 ab initio calculations were performed on the DyIII ions using their crystal structures. The anisotropy of a single DyIII ion in all complexes was computed using the X-ray structure and by substituting a diamagnetic LuIII ion for the nearby DyIII ion. The presence of radical spins of dhnq3˙− and dhaq3˙− ligands in 3 and 4 was considered as point charges while performing the calculations. Relativistic effects were taken into account based on the Douglas–Kroll Hamiltonian.85 The spin-free eigenstates are achieved by the Complete Active Space Self-Consistent Field (CASSCF) method.86 The basis sets were taken from the ANO-RCC library for the calculations. We employed the [ANO-RCC⋯8s7p5d3f2g1h.] basis set87 for DyIII atoms, the [ANO-RCC⋯3s2p.] basis set for B and C atoms, the [ANO-RCC⋯2s.] basis set for H atoms, the [ANO-RCC⋯3s2p1d.] basis set for O and N atoms, and the [ANO-RCC⋯7s6p4d2f.] basis set for the Lu atom. In the first step, we ran a guessorb calculation using a Seward module to create the starting guess orbitals. Here, we included nine electrons across seven 4f orbitals of the DyIII ion. Then, using these guess orbitals, we chose the active space based on the number of active electrons in the number of active orbitals and carried out the SA-CASSCF calculations. Here, the Configuration Interaction (CI) procedure was computed for DyIII ions, and we considered twenty-one sextet excited states, two hundred and twenty-four quartet excited states, and four hundred and eighty doublet excited states in the calculations to compute the anisotropy. All the excited states corresponding to each multiplet of ions were computed in the CASSCF module. After computing these excited states, we mixed all the low-lying excited states (<50000 cm−1) using the RASSI-SO88 module to calculate the spin–orbit coupled states. Moreover, these computed SO states were considered in the SINGLE_ANISO89 program to compute the g-tensors. The g-tensors for the Kramers doublets of DyIII were computed based on the pseudospin S = ½ formalism.89 Crystal-field (CF) parameters have been extracted using the SINGLE_ANISO code, as implemented in MOLCAS 8.0. The CF parameters for all complexes were analyzed for deeper insight into the mechanism of magnetic relaxation. The corresponding crystal field Hamiltonian is given in the equation: |
 | (2) |
where Bkq is the crystal field parameter and Okq is Steven's operator.
The exchange/dipolar interactions for DyIII–radical and DyIII–DyIII were computed by fitting the experimental data using the POLY_ANISO program.71 The exchange Hamiltonian adapted for all complexes is shown below.
|
 | (3) |
(here
Ji =
Jidipolar +
Jiexch;
i.e.,
Ji is the total magnetic interaction in the combination of calculated
Jidipolar and fitted
Jiexch parameters; this describes the interaction between all the neighboring metal centers.)
The low-lying exchange spectra were derived for all complexes using these exchange interactions, and by considering six KDs of each Dy(III) ion, and an additional doublet state of radical spin was considered for complexes 3 and 4.
Conflicts of interest
There are no conflicts to declare.
Data availability
The data supporting this article have been included as part of the SI: crystallographic/refinement tables, dc/ac magnetic measurements and analyses, and ab initio computational figures and tables for complexes 1–4) have been included as part of the SI. See DOI: https://doi.org/10.1039/d5qi01477h.
CCDC 2471757–2471760 (1–4) contain the supplementary crystallographic data for this paper.90a–d
Acknowledgements
We gratefully acknowledge support for this work by the National Science Foundation (CHE-1808779) and the Robert A. Welch Foundation (Grant A-1449). The SQUID magnetometer used was purchased with funds provided by the Texas A&M University Vice President of Research. We are grateful to the HPRC at Texas A&M for the computing resources. D. I. A. acknowledges finance support from the programme “MEDICUS” of the University of Patras. L. C.-S. thanks financial support received from the PT national funds (FCT/MECI, Fundação para a Ciência e Tecnologia/Ministério da Educação, Ciência e Inovação) through the project UID/50006 – Laboratório Associado para a Química Verde – Tecnologias e Processos Limpos.
References
- R. Bagai and G. Christou, The Drosophila of single-molecule magnetism: [Mn12O12(O2CR)16(H2O)4], Chem. Soc. Rev., 2009, 38, 1011–1026 RSC.
- A. Zabala-Lekuona, J. M. Seco and E. Colacio, Single-Molecule Magnets: From Mn12-ac to dysprosium metallocenes, a travel in time, Coord. Chem. Rev., 2021, 441, 213984 CrossRef.
- E. Coronado, Molecular magnetism: from chemical design to spin control in molecules, materials and devices, Nat. Rev. Mater., 2020, 5, 87–104 CrossRef.
- N. F. Chilton, Molecular Magnetism, Annu. Rev. Mater. Res., 2022, 52, 79–101 CrossRef.
- R. Sessoli, D. Gatteschi, A. Caneschi and M. A. Novak, Magnetic bistability in a metal-ion cluster, Nature, 1993, 365, 141 CrossRef.
- L. Bogani and W. Wernsdorfer, Molecular spintronics using single-molecule magnets, Nat. Mater., 2008, 7, 179 CrossRef PubMed.
- E. Moreno-Pineda, C. Godfrin, F. Balestro, W. Wernsdorfer and M. Ruben, Molecular spin qudits for quantum algorithms, Chem. Soc. Rev., 2018, 47, 501–513 RSC.
- R. Sessoli and A. K. Powell, Strategies towards single molecule magnets based on lanthanide ions, Coord. Chem. Rev., 2009, 253, 2328–2341 CrossRef.
- L. Sorace, C. Benelli and D. Gatteschi, Lanthanides in molecular magnetism: old tools in a new field, Chem. Soc. Rev., 2011, 40, 3092–3104 RSC.
- J. D. Rinehart and J. R. Long, Exploiting single-ion anisotropy in the design of f-element single-molecule magnets, Chem. Sci., 2011, 2, 2078–2085 RSC.
- D. N. Woodruff, R. E. P. Winpenny and R. A. Layfield, Lanthanide Single-Molecule Magnets, Chem. Rev., 2013, 113, 5110–5148 CrossRef CAS PubMed.
- N. Ishikawa, M. Sugita, T. Ishikawa, S.-y. Koshihara and Y. Kaizu, Lanthanide Double-Decker Complexes Functioning as Magnets at the Single-Molecular Level, J. Am. Chem. Soc., 2003, 125, 8694–8695 CrossRef CAS PubMed.
- Y.-S. Ding, N. F. Chilton, R. E. P. Winpenny and Y.-Z. Zheng, On Approaching the Limit of Molecular Magnetic Anisotropy: A Near-Perfect Pentagonal Bipyramidal Dysprosium(III) Single-Molecule Magnet, Angew. Chem., Int. Ed., 2016, 55, 16071–16074 CrossRef CAS PubMed.
- A. K. Bar, P. Kalita, M. K. Singh, G. Rajaraman and V. Chandrasekhar, Low-coordinate mononuclear lanthanide complexes as molecular nanomagnets, Coord. Chem. Rev., 2018, 367, 163–216 CrossRef CAS.
- B. M. Day, F.-S. Guo and R. A. Layfield, Cyclopentadienyl Ligands in Lanthanide Single-Molecule Magnets: One Ring To Rule Them All?, Acc. Chem. Res., 2018, 51, 1880–1889 CrossRef CAS PubMed.
- T. G. Ashebr, H. Li, X. Ying, X.-L. Li, C. Zhao, S. Liu and J. Tang, Emerging Trends on Designing High-Performance Dysprosium(III) Single-Molecule Magnets, ACS Mater. Lett., 2022, 4, 307–319 CrossRef CAS.
- F.-S. Guo, B. M. Day, Y.-C. Chen, M.-L. Tong, A. Mansikkamäki and R. A. Layfield, A Dysprosium Metallocene Single-Molecule Magnet Functioning at the Axial Limit, Angew. Chem., Int. Ed., 2017, 56, 11445–11449 CrossRef CAS PubMed.
- C. A. P. Goodwin, F. Ortu, D. Reta, N. F. Chilton and D. P. Mills, Molecular magnetic hysteresis at 60 kelvin in dysprosocenium, 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, Magnetic hysteresis up to 80 kelvin in a dysprosium metallocene single-molecule magnet, Science, 2018, 362, 1400–1403 CrossRef CAS.
- 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 and N. F. Chilton, et al., Ultrahard magnetism from mixed-valence dilanthanide complexes with metal-metal bonding, Science, 2022, 375, 198–202 CrossRef CAS PubMed.
- P. Zhang, L. Zhang and J. Tang, Lanthanide single molecule magnets: progress and perspective, Dalton Trans., 2015, 44, 3923–3929 RSC.
- J. Wang, C.-Y. Sun, Q. Zheng, D.-Q. Wang, Y.-T. Chen, J.-F. Ju, T.-M. Sun, Y. Cui, Y. Ding and Y.-F. Tang, Lanthanide Single-molecule Magnets: Synthetic Strategy, Structures, Properties and Recent Advances, Chem. – Asian J., 2023, 18, e202201297 CrossRef CAS PubMed.
- A. Caneschi, D. Gatteschi, R. Sessoli and P. Rey, Toward molecular magnets: the metal-radical approach, Acc. Chem. Res., 1989, 22, 392–398 CrossRef CAS.
- S. Demir, I.-R. Jeon, J. R. Long and T. D. Harris, Radical ligand-containing single-molecule magnets, Coord. Chem. Rev., 2015, 289–290, 149–176 CrossRef CAS.
- Y.-C. Chen and M.-L. Tong, Single-molecule magnets beyond a single lanthanide ion: the art of coupling, Chem. Sci., 2022, 13, 8716–8726 RSC.
- A. Swain, T. Sharma and G. Rajaraman, Strategies to quench quantum tunneling of magnetization in lanthanide single molecule magnets, Chem. Commun., 2023, 59, 3206–3228 RSC.
- H.-D. Li, S.-G. Wu and M.-L. Tong, Lanthanide–radical single-molecule magnets: current status and future challenges, Chem. Commun., 2023, 59, 6159–6170 RSC.
- J. D. Rinehart, M. Fang, W. J. Evans and J. R. Long, A N23− Radical-Bridged Terbium Complex Exhibiting Magnetic Hysteresis at 14 K, J. Am. Chem. Soc., 2011, 133, 14236–14239 CrossRef PubMed.
- S. Demir, J. M. Zadrozny, M. Nippe and J. R. Long, Exchange coupling and magnetic blocking in bipyrimidyl radical-bridged dilanthanide complexes, J. Am. Chem. Soc., 2012, 134, 18546–18549 CrossRef PubMed.
- S. Demir, M. Nippe, M. I. Gonzalez and J. R. Long, Exchange coupling and magnetic blocking in dilanthanide complexes bridged by the multi-electron redox-active ligand 2,3,5,6-tetra(2-pyridyl)pyrazine, Chem. Sci., 2014, 5, 4701–4711 RSC.
- F. Benner, L. La Droitte, O. Cador, B. Le Guennic and S. Demir, Magnetic hysteresis and large coercivity in bisbenzimidazole radical-bridged dilanthanide complexes, Chem. Sci., 2023, 14, 5577–5592 RSC.
- C. A. Gould, L. E. Darago, M. I. Gonzalez, S. Demir and J. R. Long, A trinuclear radical-bridged lanthanide single-molecule magnet, Angew. Chem., Int. Ed., 2017, 56, 10103–10107 CrossRef PubMed.
- F.-S. Guo and R. A. Layfield, Strong direct exchange coupling and single-molecule magnetism in indigo-bridged lanthanide dimers, Chem. Commun., 2017, 53, 3130–3133 RSC.
- N. Mavragani, A. A. Kitos, J. L. Brusso and M. Murugesu, Enhancing magnetic communication between metal centres: the role of s-tetrazine based radicals as ligands, Chem. – Eur. J., 2021, 27, 5091–5106 CrossRef PubMed.
- N. Mavragani, D. Errulat, D. A. Gálico, A. A. Kitos, A. Mansikkamäki and M. Murugesu, Radical-bridged Ln4 metallocene complexes with strong magnetic coupling and a large coercive field, Angew. Chem., Int. Ed., 2021, 60, 24206–24213 CrossRef CAS PubMed.
- B. S. Dolinar, S. Gómez-Coca, D. I. Alexandropoulos and K. R. Dunbar, An air stable radical-bridged dysprosium single molecule magnet and its neutral counterpart: redox switching of magnetic relaxation dynamics, Chem. Commun., 2017, 53, 2283–2286 RSC.
- B. S. Dolinar, D. I. Alexandropoulos, K. R. Vignesh, T. A. James and K. R. Dunbar, Lanthanide triangles supported by radical bridging ligands, J. Am. Chem. Soc., 2018, 140, 908–911 CrossRef CAS PubMed.
- J. A. DeGayner, I.-R. Jeon and T. D. Harris, A series of tetraazalene radical-bridged M2 (M = CrIII, MnII, FeII, CoII) complexes with strong magnetic exchange coupling, Chem. Sci., 2015, 6, 6639–6648 RSC.
- K. S. Min, A. G. DiPasquale, J. A. Golen, A. L. Rheingold and J. S. Miller, Synthesis, Structure, and Magnetic Properties of Valence Ambiguous Dinuclear Antiferromagnetically Coupled Cobalt and Ferromagnetically Coupled Iron Complexes Containing the Chloranilate(2−) and the Significantly Stronger Coupling Chloranilate(˙3−) Radical Trianion, J. Am. Chem. Soc., 2007, 129, 2360–2368 CrossRef CAS PubMed.
- C. Hua, J. A. DeGayner and T. D. Harris, Thiosemiquinoid Radical-Bridged CrIII2 Complexes with Strong Magnetic Exchange Coupling, Inorg. Chem., 2019, 58, 7044–7053 CrossRef CAS PubMed.
- P. Zhang, M. Perfetti, M. Kern, P. P. Hallmen, L. Ungur, S. Lenz, M. R. Ringenberg, W. Frey, H. Stoll and G. Rauhut, et al., Exchange coupling and single molecule magnetism in redox-active tetraoxolene-bridged dilanthanide complexes, Chem. Sci., 2018, 9, 1221–1230 RSC.
- W. R. Reed, M. A. Dunstan, R. W. Gable, W. Phonsri, K. S. Murray, R. A. Mole and C. Boskovic, Tetraoxolene-bridged rare-earth complexes: a radical-bridged dinuclear Dy single-molecule magnet, Dalton Trans., 2019, 48, 15635–15645 RSC.
- A. E. Thorarinsdottir, R. Bjornsson and T. D. Harris, Insensitivity of Magnetic Coupling to Ligand Substitution in a Series of Tetraoxolene Radical-Bridged Fe2 Complexes, Inorg. Chem., 2020, 59, 4634–4649 CrossRef CAS PubMed.
- D. Guo and J. K. McCusker, Spin Exchange Effects on the Physicochemical Properties of Tetraoxolene-Bridged Bimetallic Complexes, Inorg. Chem., 2007, 46, 3257–3274 CrossRef CAS PubMed.
- H. S. Das, A. K. Das, R. Pattacini, R. Hübner, B. Sarkar and P. Braunstein, First structurally characterized mono- and dinuclear ruthenium complexes derived from zwitterionic quinonoid ligands, Chem. Commun., 2009, 4387–4389 RSC.
- P. B. Chatterjee, K. Bhattacharya, N. Kundu, K.-Y. Choi, R. Clérac and M. Chaudhury, Vanadium-Induced Nucleophilic IPSO Substitutions in a Coordinated Tetrachlorosemiquinone Ring: Formation of the Chloranilate Anion as a Bridging Ligand, Inorg. Chem., 2009, 48, 804–806 CrossRef CAS PubMed.
- D. Schweinfurth, M. M. Khusniyarov, D. Bubrin, S. Hohloch, C.-Y. Su and B. Sarkar, Tuning Spin–Spin Coupling in Quinonoid-Bridged Dicopper(II) Complexes through Rational Bridge Variation, Inorg. Chem., 2013, 52, 10332–10339 CrossRef CAS PubMed.
- I.-R. Jeon, J. G. Park, D. J. Xiao and T. D. Harris, An Azophenine Radical-Bridged Fe2 Single-Molecule Magnet with Record Magnetic Exchange Coupling, J. Am. Chem. Soc., 2013, 135, 16845–16848 CrossRef CAS PubMed.
- D. Schweinfurth, Y. Rechkemmer, S. Hohloch, N. Deibel, I. Peremykin, J. Fiedler, R. Marx, P. Neugebauer, J. van Slageren and B. Sarkar, Redox-Induced Spin-State Switching and Mixed Valency in Quinonoid-Bridged Dicobalt Complexes, Chem. – Eur. J., 2014, 20, 3475–3486 CrossRef CAS PubMed.
- J. A. DeGayner, K. Wang and T. D. Harris, A Ferric Semiquinoid Single-Chain Magnet via Thermally-Switchable Metal–Ligand Electron Transfer, J. Am. Chem. Soc., 2018, 140, 6550–6553 CrossRef CAS.
- A. E. Thorarinsdottir and T. D. Harris, Metal–Organic Framework Magnets, Chem. Rev., 2020, 120, 8716–8789 CrossRef CAS.
- T. J. Woods, M. F. Ballesteros-Rivas, S. M. Ostrovsky, A. V. Palii, O. S. Reu, S. I. Klokishner and K. R. Dunbar, Strong Direct Magnetic Coupling in a Dinuclear CoII Tetrazine Radical Single-Molecule Magnet, Chem. – Eur. J., 2015, 21, 10302–10305 CrossRef CAS PubMed.
- T. J. Woods, H. D. Stout, B. S. Dolinar, K. R. Vignesh, M. F. Ballesteros-Rivas, C. Achim and K.
R. Dunbar, Strong Ferromagnetic Exchange Coupling Mediated by a Bridging Tetrazine Radical in a Dinuclear Nickel Complex, Inorg. Chem., 2017, 56, 12094–12097 CrossRef CAS.
- D. I. Alexandropoulos, B. S. Dolinar, K. R. Vignesh and K. R. Dunbar, Putting a New Spin on Supramolecular Metallacycles: Co3 Triangle and Co4 Square Bearing Tetrazine-Based Radicals as Bridges, J. Am. Chem. Soc., 2017, 139, 11040–11043 CrossRef CAS PubMed.
- D. I. Alexandropoulos, K. R. Vignesh, H. Xie and K. R. Dunbar, Quinoxaline radical-bridged transition metal complexes with very strong antiferromagnetic coupling, Chem. Commun., 2020, 56, 9122–9125 RSC.
- S. Maji, B. Sarkar, S. M. Mobin, J. Fiedler, F. A. Urbanos, R. Jimenez-Aparicio, W. Kaim and G. K. Lahiri, Valence-state alternatives in diastereoisomeric complexes [(acac)2Ru(μ-QL)Ru(acac)2]n (QL2− = 1,4-dioxido-9,10-anthraquinone,n = +2, +1, 0, −1, −2), Inorg. Chem., 2008, 47, 5204–5211 CrossRef CAS.
- S. Ghumaan, B. Sarkar, S. Maji, V. G. Puranik, J. Fiedler, F. A. Urbanos, R. Jimenez-Aparicio, W. Kaim and G. K. Lahiri, Valence-state analysis through spectroelectrochemistry in a series of quinonoid-bridged diruthenium complexes [(acac)2Ru(μ-L)Ru(acac)2]n (n = +2, +1, 0, −1, −2), Chem. – Eur. J., 2008, 14, 10816–10828 CrossRef CAS PubMed.
- T. S. Kamatchi, S. Mondal, T. Scherer, M. Bubrin, K. Natarajan and W. Kaim, Near-infrared-absorbing organometallic diruthenium complex intermediates: evidence for bridging anthrasemiquinone formation and against mixed valency, Chem. – Eur. J., 2017, 23, 17810–17816 CrossRef CAS PubMed.
- K. Herasymchuk, M. Allain, G. A. MacNeil, V. Carré, F. Aubriet, D. B. Leznoff, M. Sallé and S. Goeb, Exciton Coupling in Redox-Active Salen based Self-Assembled Metallacycles, Chem. – Eur. J., 2021, 27, 16161–16172 CrossRef CAS PubMed.
- S. Alvarez, P. Alemany, D. Casanova, J. Cirera, M. Llunell and D. Avnir, Shape maps and polyhedral interconversion paths in transition metal chemistry, Coord. Chem. Rev., 2005, 249, 1693–1708 CrossRef CAS.
- K. Chakarawet, T. D. Harris and J. R. Long, Semiquinone radical-bridged M2 (M = Fe, Co, Ni) complexes with strong magnetic exchange giving rise to slow magnetic relaxation, Chem. Sci., 2020, 11, 8196–8203 RSC.
- C. Benelli and D. Gatteschi, Magnetism of Lanthanides in Molecular Materials with Transition-Metal Ions and Organic Radicals, Chem. Rev., 2002, 102, 2369–2388 CrossRef CAS PubMed.
- J. J. Le Roy, J. Cremers, I. A. Thomlinson, M. Slota, W. K. Myers, P. H. Horton, S. J. Coles, H. L. Anderson and L. Bogani, Tailored homo- and hetero-lanthanide porphyrin dimers: a synthetic strategy for integrating multiple spintronic functionalities into a single molecule, Chem. Sci., 2018, 9, 8474–8481 RSC.
- K. S. Cole and R. H. Cole, Dispersion and Absorption in Dielectrics I. Alternating Current Characteristics, J. Chem. Phys., 1941, 9, 341–351 CrossRef CAS.
- S. M. J. Aubin, Z. M. Sun, L. Pardi, J. Krzystek, K. Folting, L. C. Brunel, A. L. Rheingold, G. Christou and D. N. Hendrickson, Reduced Anionic Mn12 Molecules with Half-Integer Ground States as Single-Molecule Magnets, Inorg. Chem., 1999, 38, 5329–5340 CrossRef CAS.
- F. Aquilante, J. Autschbach, R. K. Carlson, L. F. Chibotaru, M. G. Delcey, L. De Vico, I. F. Galván, N. Ferré, L. M. Frutos and L. Gagliardi, et al., Molcas 8: New capabilities for multiconfigurational quantum chemical calculations across the periodic table, J. Comput. Chem., 2016, 37, 506–541 CrossRef CAS PubMed.
- K. R. Vignesh, S. K. Langley, K. S. Murray and G. Rajaraman, Quenching
the Quantum Tunneling of Magnetization in Heterometallic Octanuclear {TMIII4DyIII4} (TM = Co and Cr) Single-Molecule Magnets by Modification of the Bridging Ligands and Enhancing the Magnetic Exchange Coupling, Chem. – Eur. J., 2017, 23, 1654–1666 CrossRef CAS PubMed.
- K. R. Vignesh, S. K. Langley, B. Moubaraki, K. S. Murray and G. Rajaraman, Understanding the Mechanism of Magnetic Relaxation in Pentanuclear {MnIVMnIII2LnIII2} Single-Molecule Magnets, Inorg. Chem., 2018, 57, 1158–1170 CrossRef CAS PubMed.
- P. P. Hallmen, C. Köppl, G. Rauhut, H. Stoll and J. van Slageren, Fast and reliable ab initio calculation of crystal field splittings in lanthanide complexes, J. Chem. Phys., 2017, 147, 164101 CrossRef CAS PubMed.
- Y. Rechkemmer, J. E. Fischer, R. Marx, M. Dörfel, P. Neugebauer, S. Horvath, M. Gysler, T. Brock-Nannestad, W. Frey and M. F. Reid, et al., Comprehensive Spectroscopic Determination of the Crystal Field Splitting in an Erbium Single-Ion Magnet, J. Am. Chem. Soc., 2015, 137, 13114–13120 CrossRef CAS PubMed.
- L. F. Chibotaru and L. Ungur, POLY_ANISO program, University of Leuven, 2006 Search PubMed.
- M. E. Lines, Orbital Angular Momentum in the Theory of Paramagnetic Clusters, J. Chem. Phys., 1971, 55, 2977–2984 CrossRef CAS.
- L. Ungur, S. K. Langley, T. N. Hooper, B. Moubaraki, E. K. Brechin, K. S. Murray and L. F. Chibotaru, Net toroidal magnetic moment in the ground state of a {Dy6}-triethanolamine ring, J. Am. Chem. Soc., 2012, 134, 18554–18557 CrossRef CAS PubMed.
- G. T. Nguyen and L. Ungur, The Role of Radical Bridges in Polynuclear Single-Molecule Magnets, Chem. – Eur. J., 2022, 28, e202200227 CrossRef CAS.
- APEX2, Data Collection Software Version 2012.4, Bruker AXS, Delft, The Netherlands, 2012 Search PubMed.
- SAINT+, SAINT+, Data Integration Engine v. 7.23a ©, Bruker AXS, Madison, Wisconsin, USA, 1997–2005 Search PubMed.
- G. M. Sheldrick, SADABS v.2.01, Bruker/Siemens Area Detector Absorption Correction Program, Bruker AXS, Madison, Wisconsin, USA, 1998 Search PubMed.
- APEX-III, Bruker AXS Inc., Madison, Wisconsin, USA, 2016.
- T. Kottke and D. Stalke, Crystal handling at low temperatures, J. Appl. Crystallogr., 1993, 26, 615–619 CrossRef.
- C. B. Hubschle, G. M. Sheldrick and B. Dittrich, ShelXle: a Qt graphical user interface for SHELXL, J. Appl. Crystallogr., 2011, 44, 1281–1284 CrossRef.
- C. F. Macrae, P. R. Edgington, P. McCabe, E. Pidcock, G. P. Shields, R. Taylor, M. Towler and J. van de Streek, Mercury: visualization and analysis of crystal structures, J. Appl. Crystallogr., 2006, 39, 453–457 CrossRef CAS.
- K. Brandenburg and H. Putz, DIAMOND, Crystal Impact GbR, Bonn, Germany, 2006 Search PubMed.
- G. A. Bain and J. F. Berry, Diamagnetic Corrections and Pascal's Constants, J. Chem. Educ., 2008, 85, 532 CrossRef CAS.
- F. Aquilante, J. Autschbach, R. K. Carlson, L. F. Chibotaru, M. G. Delcey, L. De Vico, I. F. Galván, N. Ferré, L. M. Frutos and L. Gagliardi, et al., MOLCAS 8: New Capabilities for Multiconfigurational Quantum Chemical Calculations across the Periodic Table, J. Comput. Chem., 2016, 37, 506–541 CrossRef CAS.
- B. A. Hess, C. M. Marian, U. Wahlgren and O. Gropen, A mean-field spin-orbit method applicable to correlated wavefunctions, Chem. Phys. Lett., 1996, 251, 365–371 CrossRef CAS.
- B. O. Roos, P.-Å. Malmqvist, Relativistic quantum chemistry: the multiconfigurational approach, Phys. Chem. Chem. Phys., 2004, 6, 2919–2927 RSC.
- B. O. Roos, R. Lindh, P.-Å. Malmqvist, V. Veryazov, P.-O. Widmark and A. C. Borin, New Relativistic Atomic Natural Orbital Basis Sets
for Lanthanide Atoms with Applications to the Ce Diatom and LuF3, J. Phys. Chem. A, 2008, 112, 11431–11435 CrossRef CAS PubMed.
- P. Å. Malmqvist, B. O. Roos and B. Schimmelpfennig, The restricted active space (RAS) state interaction approach with spin-orbit coupling, Chem. Phys. Lett., 2002, 357, 230–240 CrossRef CAS.
- L. F. Chibotaru and L. Ungur, Ab initio calculation of anisotropic magnetic properties of complexes. I. Unique definition of pseudospin Hamiltonians and their derivation, J. Chem. Phys., 2012, 137, 064112 CrossRef CAS PubMed.
-
(a) D. I. Alexandropoulos, K. R. Vignesh, L. Cunha-Silva and K. R. Dunbar, CCDC 2471757: Experimental Crystal Structure Determination, 2025, DOI:10.5517/ccdc.csd.cc2nz23v;
(b) D. I. Alexandropoulos, K. R. Vignesh, L. Cunha-Silva and K. R. Dunbar, CCDC 2471758: Experimental Crystal Structure Determination, 2025, DOI:10.5517/ccdc.csd.cc2nz24w;
(c) D. I. Alexandropoulos, K. R. Vignesh, L. Cunha-Silva and K.R. Dunbar, CCDC 2471759: Experimental Crystal Structure Determination, 2025, DOI:10.5517/ccdc.csd.cc2nz25x;
(d) D. I. Alexandropoulos, K. R. Vignesh, L. Cunha-Silva and K.R. Dunbar, CCDC 2471760: Experimental Crystal Structure Determination, 2025, DOI:10.5517/ccdc.csd.cc2nz26y.
|
This journal is © the Partner Organisations 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.