Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Quenching quantum tunnelling of the magnetization utilizing 4d–4f exchange interactions in butterfly-shaped {RuIII2LnIII2} (Ln = Gd, Tb, Dy, Ho, and Er) single-molecule magnets

Imon J. Dutta a, Vignesh Kasi b, Deepanshu Chauhan b, Keith S. Murray c, Stuart K. Langley d, Maheswaran Shanmugam *b and Kuduva R. Vignesh *a
aDepartment of Chemical Sciences, Indian Institute of Science Education and Research Mohali, Sector-81, Knowledge City, S.A.S. Nagar, Mohali 140306, Punjab, India. E-mail: vigneshkuduvar@iisermohali.ac.in
bDepartment of Chemistry, Indian Institute of Technology Bombay, Powai, Mumbai 400076, Maharashtra, India. E-mail: eswar@chem.iitb.ac.in
cSchool of Chemistry, Monash University, Clayton, VIC, Australia 3800
dDepartment of Natural Sciences, Chemistry, Manchester Metropolitan University, Manchester, UK

Received 6th August 2025 , Accepted 30th September 2025

First published on 6th October 2025


Abstract

The performance of Single-Molecule Magnets (SMMs) can be enhanced by reducing quantum tunneling of magnetization (QTM), which can be achieved with the aid of strong magnetic exchange interactions and strong magnetic anisotropy. When 4d ions are mixed with 4f ions, the strong spin–orbit coupling and moderate ligand-field strength of 4d elements complement the highly anisotropic 4f ions, and the combined effect decreases the transverse anisotropy and stabilizes the magnetic easy axis, both of which are essential for lowering QTM. Despite their synthetic challenges, the unique electronic properties of 4d–4f systems make them promising candidates for designing advanced single-molecule magnets. Here, we report the synthesis, magnetic, and theoretical studies of five tetranuclear “butterfly-shaped” complexes of formulae [RuIII2LnIII2(OMe)2(o-tol)4(mdea)2(NO3)2] (Ln = Gd (1-Gd), Tb (2-Tb), Dy (3-Dy), Ho (4-Ho), mdeaH2 = N-methyldiethanolamine) and [RuIII2ErIII2(OMe)2(o-tol)6(mdea)2]·2MeOH (5-Er). Complexes 2-Tb, 4-Ho, and 5-Er are the first reported examples of Ru-4f complexes using these LnIII ions. DC magnetic susceptibility studies revealed an antiferromagnetic exchange interaction between the LnIII and RuIII metal ions in all complexes. Among the complexes reported, 3-Dy and 4-Ho show out-of-phase susceptibility signals in the absence of an external magnetic field, while the remaining complexes exhibit field-induced slow relaxation of magnetization. These observations were attributed to the presence of an ideal crystal field combined with the strong exchange interaction exerted between the RuIII and LnIII ions in the complexes. Importantly, the 4-Ho is the first non-dysprosium Ru–Ln polynuclear complex to show zero-field SMM behavior. Detailed ab initio CASSCF + RASSI-SO + SINGLE_ANISO + POLY_ANISO and Density Functional Theory (DFT) calculations reveal that the magnitude of exchange interaction increases in the following order: 1-Gd (−1.3 cm−1) < 2-Tb (−1.8 cm−1) < 3-Dy (−2.7 cm−1) < 4-Ho (−2.8 cm−1) < 5-Er (−4.0 cm−1). The larger single-ion anisotropy of the DyIII ions in 3-Dy leads to the largest barrier of 101.8 cm−1 (146 K), which increases further to 111 cm−1 (158.7 K) when considering the exchange-coupled model, with the strong Ru⋯Ln exchange further quenching QTM effects. These results highlight the importance of strong magnetic exchange between 4d–4f (RuIII⋯LnIII) ions in enhancing the SMM behavior of 4d–4f complexes.


Introduction

Single-Molecule Magnets have attracted considerable attention because of their potential use in digital data storage and processing,1,2 quantum computing,1,3–5 nanotechnology,6 and spintronic devices.7 The main challenges in designing high-performing SMMs are controlling the numerous magnetic relaxation pathways, especially the QTM.8–10 In order to improve the SMM properties for polynuclear systems, metal ions with a strong anisotropic nature and significant magnetic exchange interaction have become essential.11 Either a suitable combination of certain metal ions or the use of radical bridges can significantly increase the magnetic exchange coupling between the metal ions.12,13 To attain strong magnetic anisotropy, lanthanide (Ln) ions are chosen, which constitute strong unquenched orbital angular momentum compared to 3d transition metals. Some well-known Ln-based SMMs to date include (i) double-decker Tb(Pc)2 (Pc = phthalocyanine) with effective energy barriers (Ueff) of 330.9 K;14 (ii) air-sensitive dysprosocenium complexes show Ueff more than 2000 K and hysteresis up to 80 K;15,16 (iii) air-stable Dy(III) single-ion magnets (SIMs) with pseudo-D5h symmetry exhibit a magnetization blocking (TB) up to 12 K and Ueff above 735 K;17,18 (iv) hexagonal bipyramidal (HBP)-shaped Dy(III)-SIMs with pseudo-D6h symmetry exhibit Ueff = 779 K–1360 K;18–20 (v) mixed-valence dilanthanide complexes of formula (CpiPr5)2Ln2I3 exhibit TB in the range of 28 K–72 K and Ueff in the range of 795 K–2345 K.21,22 At the polynuclear level, the anisotropic properties of Ln ions have also been investigated by incorporating them with 3d metal ions. This strategy revolves around increasing the magnetic exchange interaction (J) between 3d–4f ions to greater magnitudes than possible between 4f–4f ions to reduce the onset of QTM, which is often the cause of fast magnetic relaxation times.23–26 Thus, larger magnetic exchange interactions are essential for polynuclear systems with long magnetic relaxation timescales. Some of us have reported several {CrIII2DyIII2} tetranuclear “butterfly” complexes and explored the incremental increase of the energy barrier and blocking temperature (Ueff = 31.6 K (22 cm−1) to 93.5 K (65 cm−1) and TB = 1.8 K to 4.7 K) by modulating the CrIII⋯LnIII magnetic exchange interaction (−11.2 and −8.3 cm−1 to −20.7 and −16.3 cm−1).27–29 Similarly, we reported an octanuclear {CrIII4DyIII4} SMM with an energy barrier of 79 K and TB = 3.5 K, with Cr⋯Dy magnetic exchange J = −1.8 cm−1.30 To demonstrate the importance of 3d–4f magnetic exchange interactions in observing SMM behavior, we replaced the CrIII with a diamagnetic CoIII ion in both tetranuclear and octanuclear cases, thus removing the exchange interaction. This resulted in a reduced anisotropy barrier for both the butterfly (38 K),31 and octanuclear complexes (56 K), with a loss of magnetic hysteresis above 1.8 K for all complexes.32 Many other 3d–4f complexes have been reported with other 3d ions, for example, a {FeIII4DyIII4} wheel exhibits Ueff of 30.5 K with ferromagnetic exchange interaction,33,34 a 20-nuclearity compound {Fe8Dy12} with an Ueff of 3.54 K,34 a {NiII2DyIII2} complex with an anisotropy barrier of 19 K and JNi–Dy = +0.94 cm−1, a pentanuclear {NiII2DyIII3} with Ueff of 85 K and TB = 3 K,35,36 a [MnIII4LnIII4] wheel (Ln = Sm, Gd, Tb, Dy, Ho, Er, and Y) and a {MnIII8LnIII8} wheel (Ln = Dy, Ho, Er, Y, and Yb) where MnIII–DyIII magnetic exchange is estimated to be −0.1 cm−1 and their DyIII analogs are SMMs.37,38 We believe that these complexes show good SMM behaviour mainly because of their 3d–4f exchange interactions; however, the J values were not determined for some of them either by analyzing DC magnetic data or through computational calculations.

Similar to Ln(III) ions, 4d and 5d metals exhibit stronger spin–orbit coupling (SOC) and larger magnetic anisotropies (highly anisotropic g-tensors with an unusually large ZFS value), leading to slower magnetic relaxation. Among 4d/5d elements, the reported mononuclear SMMs are mostly based on ReIV, NbIV, MoIII/V, RuIII, WIII/V, and OsIII metal ions, which exhibit very large magnetic anisotropy due to their sizeable spin–orbit coupling constant.39 Among a variety of examples, the complexes involving ReIV, RuIII, OsIII, MoIII, and NbIII/IV centers such as [ReX4(ox)]2− (X = Cl and Br),40 [ReCl4(CN)2]2–,41trans-[RuIII(salen)(CN)2], trans-[OsIII(salen)(CN)2], fac-MoIII(Me3tacn)(CN)3, and [NbIII/IV(CN)8]5−/4−, [ReIV(CN)7]3− are primarily fascinating because they possess large negative ZFS.42,43 Further to this, a hexanuclear {RuIII6} complex44 has been shown to exhibit robust magnetic exchange interactions, with J values exceeding −800 cm−1, hence suitable for combining with lanthanide ions to target stronger magnetic exchange interactions compared to 3d–4f complexes. Hence, 4d and 5d ions are preferred compared to their comparable 3d–4f and 4f–4f counterparts, since the more diffuse nature of the 4d/5d orbitals may encourage stronger magnetic exchange and help to reduce QTM and boost SMM performance.8,45 Thus, despite their synthetic hurdles, 4d/5d–4f systems’ unique electronic characteristics will make them interesting candidates for developing improved SMMs. However, there is a paucity of work in this area, with only a handful of polynuclear 4d–4f examples reported in the literature.

One such example is a trinuclear 4d–4f compound [Co(C5Me5)2][(C5Me5)2Ln(μ-S)2Mo(μ-S)2Ln(C5Me5)2], (Ln = Y, Gd, Tb, Dy) reported by Long and co-workers in which the 4d⋯4f (JGd–Mo) magnetic exchange interaction is found to be 16.1 cm−1, one of the largest magnetic exchange constants thus far reported with a 4f ion. The DyIII analogue shows slow magnetic relaxation with a Ueff = 68 cm−1 (98 K), which is the largest reported for a complex containing a paramagnetic 4d metal center.8 Dunbar group reported CN-bridged {ReII2LnIII3} (Ln = Dy and Tb) clusters; however, they are weak SMMs.46 The two above-mentioned examples are air-sensitive. Similarly, we have shown SMM behavior on a series of air-stable {RuIII2DyIII2} butterfly complexes, which are the first examples of 4d–4f complexes using the Ru(III) ion.45,47 The idea of observing strong exchange in 4d–4f complexes is reflected by the findings of this work, which revealed that the exchange interaction was significantly better than any 3d–4f example, except for a {CrIII2DyIII2} complex.27 It was shown that the dominant magnetic exchange interaction was observed between the LnIII⋯RuIII ions varying between −1.9 and −2.42 cm−1. We also showed that by altering the co-ligands, such as the amine polyalcohol and benzoic acid vs. acetylacetone, the anisotropy barrier changes, with values ranging from 15.4 K (ref. 45) to 45 K (Scheme 1).47


image file: d5qi01665g-s1.tif
Scheme 1 Schematic representation of a different type of butterfly-shaped 2p/3d/4d–4f SMMs and new 4d–4f SMMs reported in this work.

The understanding of what controls the sign and magnitude of the RuIII–LnIII exchange interactions and SMM behavior in these {RuIII2LnIII2} butterfly complexes needs to be further verified with the development of magneto-structural correlations. Hence, to develop structure–property correlations in these {RuIII2LnIII2} butterfly complexes, we have expanded the family of complexes by varying substitution in the benzoic acid as well as replacing the anisotropic DyIII ion with other anisotropic LnIII ions such as TbIII, HoIII, and ErIII. In this context, we have isolated five {RuIII2LnIII2} butterfly complexes with the molecular formulae [RuIII2LnIII2(OMe)2(o-tol)4(mdea)2(NO3)2] (Ln = 1-Gd, 2-Tb, 3-Dy, 4-Ho) and [RuIII2ErIII2(OMe)2(o-tol)6(mdea)2]·2MeOH (5-Er) (o-tol = o-toluate, (mdea)2− = doubly deprotonated N-methyldiethanolamine). To the best of our knowledge, complexes 2-Tb, 4-Ho, and 5-Er are the first mixed RuIII-4f SMMs containing anisotropic non-dysprosium ions. Herein, we discuss the synthesis, crystal structure, magnetism, and computational analysis of the {RuIII2LnIII2} butterfly complexes and explore how the magnetic behaviour is influenced by changing the carboxylate and lanthanide metal ion.

Results and discussion

Syntheses and crystal structures

The reaction of RuCl3·xH2O, Ln(NO3)3·xH2O [Ln = Gd, Tb, Dy, Ho, and Er; x = 5 or 6], ortho-toluic acid, N-methyldiethanolamine (mdeaH2), and triethylamine in acetonitrile solvent generates a dark brown solution.

Subsequent removal of solvent and redissolution of the residue in MeOH gives brown crystals of [RuIII2LnIII2(OMe)2(o-tol)4(mdea)2(NO3)2] [Ln = Gd (1-Gd), Tb (2-Tb), Dy (3-Dy), and Ho (4-Ho)] and [RuIII2ErIII2(OMe)2(o-tol)6(mdea)2] (5-Er) within 2 weeks.

Complexes 3-Dy, 4-Ho, and 5-Er crystallize in the triclinic P[1 with combining macron], space group. For 3-Dy and 4-Ho, the asymmetric unit contains half of the complex, i.e., one RuIII and one DyIII metal ion, both of which are positioned on an inversion center. Complex 5-Er contains two half molecules in the asymmetric unit, which lie upon an inversion centre (Table S1). The tetranuclear complexes [RuIII2LnIII2] (Ln = Dy, Ho, and Er) exhibit a butterfly-shaped core structure, where the two lanthanide ions occupy the central body positions and the two RuIII ions are located at the peripheral wing positions (Fig. 1). Each LnIII ion is eight-coordinated, surrounded by a combination of oxygen and nitrogen donor atoms from bridging and chelating ligands, resulting in a distorted coordination geometry. The RuIII centers adopt distorted octahedral environments, coordinated by both N and O donor atoms. Two μ3-methoxide (μ3-OMe) ligands are present, bridging two LnIII ions to a RuIII ion. Furthermore, the complex consists of two mdea2− ligands, which coordinate to one RuIII ion via the N-atom, which then bridges from the RuIII to two LnIII ions via the two μ2-O-atoms. Four o-toluate ligands are found in complexes 3-Dy and 4-Ho, each bridging between a RuIII and a LnIII metal ion. For 5-Er, six o-toluate ligands are present, four again bridge a RuIII and ErIII ion, while the final two each chelate to a single ErIII ion. For 3-Dy and 4-Ho, the final coordination sites are completed by two nitrate (NO3) ions, which chelate to each LnIII ion. There is no difference in the synthesis method used for 5-Er compared to those used for the other complexes; the difference in structure may be a consequence of the size of the ion (the smallest of the 4f ions used). The average Ru–O/N bond lengths for 3-Dy, 4-Ho, and 5-Er are found to be 2.036, 2.046, and 2.048 Å, respectively. The average Ln–O bond lengths for 3-Dy, 4-Ho, and 5-Er are found to be 2.369, 2.351, and 2.346 Å, respectively. The RuIII ions of 3-Dy, 4-Ho, and 5-Er possess a six-coordinate octahedral geometry, which, when analyzed by SHAPE analysis, show a CShM of 0.411, 0.400, and 0.399 (Table S2). Similarly, DyIII, HoIII, and ErIII metal ions of 3-Dy, 4-Ho, and 5-Er complexes exhibit square antiprismatic geometry with a CShM of 1.747, 1.719, and 1.671 (Table S2).


image file: d5qi01665g-f1.tif
Fig. 1 Molecular structures of complexes 3-Dy and 5-Er. The solvent and H atoms are omitted for clarity. Color scheme: RuIII, pink; DyIII, green; ErIII, lavender; O, red; N, blue; C, light grey. Complexes 1-Gd, 2-Tb, and 4-Ho are isostructural with 3-Dy, with the GdIII, TbIII, and HoIII ions replacing the DyIII sites.

Single crystals suitable for X-ray diffraction were not obtained for complexes 1-Gd and 2-Tb; thus, the solid-state structures of 1-Gd and 2-Tb were elucidated through their powder X-ray diffraction (PXRD) patterns (Fig. S1). The experimental PXRD peaks of 1-Gd and 2-Tb align closely with the simulated pattern, with all major reflections accurately matched in both position and intensity, showing that these samples possess the same crystal structure as 3-Dy obtained from SCXRD. Minor discrepancies in relative intensities at lower 2θ, attributable to preferred orientation effects within the powder sample. For 3-Dy, 4-Ho, and 5-Er complexes, sample phase purity is checked by comparing the powder XRD pattern with the simulated single-crystal XRD data (Fig. S1). The core is stabilized by multiple μ-bridging alkoxo or carboxylate ligands, creating a compact and rigid framework that enables short Ru⋯Ln distances, crucial for effective 4d–4f magnetic exchange interactions. While the overall topology remains consistent across the series, subtle variations in coordination geometry arise from the differing ionic radii and anisotropies of the DyIII and ErIII centers, which in turn influence their magnetic behavior and relaxation dynamics.

Magnetic properties

Direct current magnetic susceptibility measurements. The direct current (DC) magnetic susceptibility measurements on complexes 1-Gd, 2-Tb, 3-Dy, 4-Ho, and 5-Er have been conducted across the temperature range of 1.8–300 K in a DC magnetic field of 0.1 T (Fig. 2). The room temperature χT values of 1-Gd, 2-Tb, 3-Dy, 4-Ho, and 5-Er are 16.76, 23.90, 29.20, 27.43, and 24.13 cm3 K mol−1, respectively. These are in good agreement with the calculated values of non-interacting two GdIII (S = 7/2, L = 0, 8S7/2, g = 2, C = 7.88 cm3 K mol−1), two TbIII (S = 3, L = 3, 7F6, g = 3/2, C = 11.82 cm3 K mol−1), two DyIII (S = 5/2, L = 5, 6H15/2, g = 4/3, C = 14.17 cm3 K mol−1), two HoIII (S = 2, L = 6, 5I8, g = 5/4, C = 14.08 cm3 K mol−1) ions, two ErIII (S = 3/2, L = 6, 4I15/2, g = 6/5, C = 11.48 cm3 K mol−1) ions, and two RuIII (S = 1/2, g = 2, C = 0.37 cm3 K mol−1) ions of 16.50, 24.39, 29.08, 28.89, and 23.71 cm3 K mol−1, respectively (Fig. 2).
image file: d5qi01665g-f2.tif
Fig. 2 Experimental χT vs. T for complexes 1-Gd, 2-Tb, and 4-Ho (top), 3-Dy, and 5-Er (bottom) in an applied magnetic field of 0.1 T. The fitted lines are from the POLY_ANISO48–50/PHI program51 (see below for more details).

As the temperature is decreased, the χT product decreases down to 25 K, below which an upturn is observed for all complexes 1-Gd, 2-Tb, 3-Dy, 4-Ho, and 5-Er. The high temperature decrease (300–25 K) is due to the depopulation of the excited mJ states of the LnIII ions, while the increase at lower temperatures suggests non-negligible exchange interactions between the RuIII and the LnIII ions (see ab initio and DFT analysis). Below 2 K, a sudden drop is observed for all five complexes, due to antiferromagnetic intermolecular interactions (zJ) or dipolar interaction, or the magnetic anisotropy. For 1-Gd, 2-Tb, 3-Dy, 4-Ho, and 5-Er, the isothermal magnetization (M) measurements are plotted as a function of the magnetic field (H) and shown in Fig. S2. All the complexes exhibit a sharp rise in magnetization below 2 T, followed by a more steady, linear-like increase that doesn't reach saturation, indicating the presence of low-lying excited states and/or a considerable magnetic anisotropy (Fig. S2). Using the PHI program,51 the exchange interactions (see Fig. S10) were derived by fitting the dc susceptibility data, which yielded the magnetic exchange constants as follows: J1(Ru–Gd) = −1.0 cm−1, J2(Gd–Gd) = 0.03 cm−1, and J3(Ru–Ru) = −0.07 cm−1. These experimental fitting data suggest that the combination of antiferromagnetic and ferromagnetic interactions within the complex leads to an overall ferrimagnetic behavior.

Alternating current magnetic susceptibility measurements. To study the magnetization dynamics of slow magnetic relaxation, variable temperature and variable frequency AC magnetic susceptibility measurements were carried out. Using a 5 Oe oscillating AC field (0.1–750 Hz) and a zero DC field, complexes 3-Dy and 4-Ho showed SMM characteristics. These two complexes show clear frequency-dependent maxima in the out-of-phase (χ′′M) susceptibility component. Over the whole frequency range, peak maxima are found for χ′′M between 1.8–4.5 K and 1.8–3 K for 3-Dy and 4-Ho, respectively (Fig. 3). Upon application of an optimum external magnetic field of 1600 Oe, 4-Ho shows clear frequency-dependent maxima in the out-of-phase (χ′′M) signals between 1.8–4 K (Fig. S4) with two relaxations; the maxima at higher frequency are due to QTM, and the maxima at lower frequency were analyzed further. Complexes 2-Tb and 5-Er show no out-of-phase (χ′′M) susceptibility peaks in a zero static DC field, but maxima are observed under an applied DC field of 2000 and 2400 Oe, respectively, signifying field-induced SMMs (Fig. 3). This behavior suggests QTM is more significant for 2-Tb and 5-Er, which is quenched to some extent under the application of the static DC field. For 1-Gd, AC measurements performed under a DC field of 3600 Oe revealed maxima in χ′′Mvs. frequency plot (Fig. 3). Since the GdIII ion is isotropic, the origin of the slow magnetic relaxation for 1-Gd is attributed to the presence of the anisotropic RuIII ions and the exchange interaction between the RuIII and GdIII ions.
image file: d5qi01665g-f3.tif
Fig. 3 χ′′Mvs. frequency plots for 3-Dy and 4-Ho (middle-left and middle-right) at Hdc = 0 Oe and 1-Gd, 2-Tb, and 5-Er (top-left and top-right and bottom), at Hdc = 3600, 2000, and 2400 Oe, respectively. The solid lines are fitted values obtained from the CC-fit52 program using the parameters given in the text.

Fig. 3, S5, S6, and S7 are comprised of χ′′Mvs. frequency, χMvs. frequency, and χ′′Mvs. χM for all five complexes. The Cole-Cole plots of χ′′Mvs. χM of 1-Gd and 3-Dy showed a single semicircular pattern, confirming a single relaxation process, whereas complexes 2-Tb, 4-Ho (0 and 1600 Oe), and 5-Er showed two semicircular patterns, referring to more than one relaxation process (Fig. S6 and S7). The second relaxation process for 2-Tb, 4-Ho, and 5-Er is observed to be largely temperature-independent, suggesting relaxation via QTM.32,62

Using CC-FIT,52 the temperature-dependent relaxation data for 1-Gd, 2-Tb, 3-Dy, 4-Ho (at 0 and 1600 Oe), and 5-Er were used to extract the relaxation times. The relaxation times (τ) from the CC-FIT,52 were plotted against 1/T, resulting in a non-linear and Arrhenius plot (Fig. 4, S8, and S9). The best-fit magnetic relaxation parameters are tabulated in Table 1. Fitting the frequency-dependent AC data (solid lines in Fig. 3, S5, S6, and S7) using CC-FIT52 yields the relaxation times and the associated α parameters (Tables S4–S9). For 3-Dy and 4-Ho, the α parameter remains above 0.2 up to 4 K. These higher α values hint at wider distributions that are observed at these temperatures, which are probably a consequence of the interaction of the Raman, QTM, and Orbach relaxation processes. The threshold for 1-Gd, 2-Tb, and 5-Er is 2 or 3.5 K, signifying narrow distributions of relaxation times at higher temperatures, indicating that only the Orbach process is present above 2.5 K and QTM is dominant below that. Eqn (1) was used to fit the data for 3-Dy and 4-Ho (0 Oe) by considering all relaxation processes.

 
1/τ = 1/τQTM + CTn + τ0−1[thin space (1/6-em)]exp(Ueff/kBT),(1)
where the 1st, 2nd, and 3rd terms correspond to quantum tunneling, Raman relaxation, and Orbach relaxation pathway, respectively. Only fitting the relaxation data to the Arrhenius equation (eqn (2)) was used for 1-Gd, 2-Tb, 4-Ho (1600 Oe), and 5-Er.
 
τ = τ0[thin space (1/6-em)]exp(Ueff/kBT)(2)


image file: d5qi01665g-f4.tif
Fig. 4 Magnetization relaxation time (τ), plotted as ln(τ) versus T−1 for complexes 3-Dy and 4-Ho.
Table 1 Fitting parameters of the relaxation processes for 1-Gd, 2-Tb, 3-Dy, 4-Ho (at 0 and 1600 Oe), and 5-Er in the presence and absence of an external DC field
Complex 1-Gd 2-Tb 3-Dy 4-Ho 4-Ho 5-Er
H dc/Oe 3600 2000 0 0 1600 2400
τ QTM/s 1.77 0.00003 0.003 0.008 0.00002 0.4
n 5.64 9
C/s−1 Kn 5 0.4
τ 0/s 7 × 10−2 1 × 10−11 1 × 10−6 1.5 × 10−11 9.5 × 10−11 2.6 × 10−9
U eff/K 5.5 30.5 60.0 20.0 22.2 38.9


The non-linear equation fit (eqn (1)) for the ln(τ) vs. T−1 data yields an energy barrier of 60 K for 3-Dy and 20 K for 4-Ho with small QTM values, which supports the observation of zero-field SMM behavior. The QTM is quenched for 4-Ho upon application of a dc field, which slightly increases the barrier to 22.2 K. Similarly, the QTM is quenched upon application of dc fields for 2-Tb and 5-Er, which yield an anisotropy barrier of 30.5 K and 38.9 K, respectively. However, the application of the DC field does not reduce QTM for 1-Gd, yielding a negligible barrier of 5.5 K. The origin of such a small barrier could be attributed to the anisotropic RuIII ions, considering the isotropic nature of the GdIII ion.

To infer the magnetization blockade, the M vs. H hysteresis loops under optimal conditions for all complexes on their polycrystalline samples were measured at 2 K with an average sweep rate of 150 Oe s−1. Unfortunately, no coercivity was observed in any of the reported complexes, at the temperature and sweep rate measured (Fig. S3 in the SI).

Computational studies

Ab initio and DFT calculations were performed to investigate the observed magnetic properties using MOLCAS 8.2 and Gaussian 16 software, respectively (see Computational details). To investigate the anisotropic nature of individual LnIII and RuIII ions in 2-Tb, 3-Dy, 4-Ho, and 5-Er, CASSCF/SO-RASSI/SINGLE_ANISO calculations were conducted, followed by the overall magnetic behavior of these complexes, which was rationalized using POLY_ANISO. In all SINGLE_ANISO calculations, the neighboring LnIII and RuIII ions were substituted with diamagnetic LuIII and YIII ions in each of these computations. In POLY_ANISO calculations, the RuIII–LnIII, LnIII–LnIII, and RuIII–RuIII magnetic exchange interactions were computed in order to develop an exchange-coupled states relaxation mechanism for 1-Gd, 2-Tb, 3-Dy, 4-Ho, and 5-Er.
Single-ion relaxation mechanism. Tables S10–S12 comprise the energies of low-lying Kramers doublets (KDs)/Ising doublets (IDs) of the individual LnIII and RuIII ions and their respective g-tensor values for 2-Tb, 3-Dy, 4-Ho, and 5-Er. These computed values suggest that both LnIII centers and both RuIII centers are identical to each other in their respective complexes. The ground-state KDs of DyIII ions in 3-Dy, and IDs of TbIII and HoIII ions in 2-Tb and 4-Ho are highly axial, revealing a large gz value (∼19.8 for Dy, ∼17.2 for Tb, and ∼19.1 for Ho) and negligible transverse components (gx, gy), suggesting these LnIII ions are likely to show magnetic relaxation in the excited states. Whereas in 5-Er, the local g-tensors of each ErIII ion in the ground KD possess a large transverse component (gx and gy), suggesting a high probability of fast QTM.

Fig. 5 displays the orientations of the main anisotropy axes (gz) in the ground doublets for 2-Tb, 3-Dy, 4-Ho, and 5-Er. The anisotropic axes of LnIII centers in these complexes are nearly parallel to each other, and the same scenario was observed between the RuIII centers. The local magnetic moments in the ground exchange doublet state of RuIII centers are aligned opposite to the main anisotropy axis of LnIII sites. The LnIII anisotropic axes deviate from the μ3-O-atoms of methoxides since they possess large LoProp charges of −0.77 to −0.79, which helps them to align in the direction of the Omeda–Ln–Omeda (O–Ln–O) bond. The parallel alignments between the DyIII ions deviate slightly from the O–Ln–O bond vectors due to the large and differing LoProp charges on the μ2-O-atoms of mdea2− ligands. In 2-Tb, 3-Dy, and 4-Ho, the anisotropy directions of the Ln⋯Ln bond vectors lie between the O–Ln–O bond of the mdea2− ligands and the O–Ln–O bond of the o-toluate ligands, which can be attributed to the presence of anisotropy directions of the Ru⋯Ru bond vectors.


image file: d5qi01665g-f5.tif
Fig. 5 Magnetic anisotropy directions of complexes 2-Tb (top, left), 3-Dy (top, right), 4-Ho (bottom, left), and 5-Er (bottom, right), along with the computed LoProp charges on metal and donor atoms. The arrows show the orientation of local magnetic moments in the ground exchange doublet state.

The Ln⋯Ln vectors are tilted by an angle of 30.8°/50.4°, 18.4°/22.3°, and 20.9°/30.7° from the linear O–Ln–O bond vector in 2-Tb, 3-Dy, and 4-Ho, respectively.

Such a small deviation helps to maintain the stronger magnetic anisotropy, which is reflected in their gz values. Whereas in 5-Er, the Er⋯Er bond vectors also lie between the O–Ln–O bond of the mdea2− ligands and the O–Ln–O bond of the o-toluate ligands, but closer towards the Oo-tol–ErIII bonds, and they are tilted by an angle of 89.8°/91.8° from the O–Ln–O bond of the mdea2− ligands, which hints at the presence of weaker anisotropy for the ErIII ions.53,54

The pronounced tilt of the anisotropy axes in the Er complex originates from the prolate nature of the ErIII f-electron density, which favours anisotropy axes oriented perpendicular to the equatorial ligand field. As a result, in 5-Er the Er⋯Er vectors are nearly orthogonal (≈90°) to the O–Er–O bond directions of the mdea2− ligands, leading to much weaker axial anisotropy. In contrast, TbIII, DyIII, and HoIII possess oblate electron densities, for which the anisotropy axes preferentially align along the axial O–Ln–O bond vectors, which preserves strong axial anisotropy, consistent with the larger gz values observed.

To verify the presence of stronger magnetic anisotropies of TbIII, DyIII, and HoIII ions in 2-Tb, 3-Dy, and 4-Ho and the weaker magnetic anisotropy of ErIII ions in 5-Er in their magnetic relaxation pathways, the energy values of eight low-lying Kramers doublets for 3-Dy and 5-Er, and the low-lying Ising doublets for 2-Tb and 4-Ho were calculated (Table S10). These low-lying doublet states span up to 478.2 cm−1, 551.1 cm−1, 263.2 cm−1, and 338.1 cm−1 in 2-Tb, 3-Dy, 4-Ho, and 5-Er, respectively. The energy gap between the ground and first excited KDs/IDs is found to be 118.3 cm−1, 101.8 cm−1, 70.1 cm−1, and 36.8 cm−1 in 2-Tb, 3-Dy, 4-Ho, and 5-Er, respectively. These low-lying states and the larger gap between each doublet state suggest that 3-Dy can exhibit the best SMM behavior with a maximum barrier of 551.1 cm−1 if QTM is quenched in all the states compared to other analogs.

To corroborate these assumptions, the magnetic relaxation mechanism for each Ln center was developed, and a qualitative explanation of the magnetic relaxation mechanism for each TbIII, DyIII, HoIII, and ErIII ion in 2-Tb, 3-Dy, 4-Ho, and 5-Er, respectively, is illustrated in Fig. 6 and S11. For both the Dy ions (Dy1 and Dy2) in 3-Dy, QTM is very small in the ground state (0.003 μB), which is mainly due to the presence of a highly pure magnetic ground state of mJ = ±15/2, and allows the magnetic relaxation to occur in the excited states. The magnetic relaxation occurs between the first excited state (mJ = ±13/2) as it is highly mixed with other magnetic states mJ = ±11/2, etc., leading to significant thermally assisted-QTM (TA-QTM) values (0.24 μB) and yielding a barrier of 101.8 cm−1 (146 K). This is in reasonable agreement with the experimental Ueff value of 60 K. The two Er ions (Er1 and Er2) in 5-Er possess a significant QTM in the ground state (0.72 μB), which is caused by the mixing of the mJ = ±15/2 states with other states such as ±13/2, ±11/2, ±7/2, and ±1/2, etc., suggesting magnetic relaxation between the ground KDs itself. The magnetic relaxation is likely to occur in the first excited state if QTM is quenched upon applying a dc field and yields energy barrier of Ucal = 36.8 cm−1 (52.6 K), which slightly overestimates the experimental Ueff value of 38.9 K under a dc field of 2400 Oe.


image file: d5qi01665g-f6.tif
Fig. 6 The ab initio computed magnetization blocking barrier for Tb1, Dy1, Ho1, and Er1 ions in 2-Tb, 3-Dy, 4-Ho, and 5-Er. The thick black line indicates the KDs/IDs as a function of computed magnetic moment. The green/blue arrows show the possible pathway through Orbach/Raman relaxation. The dotted red lines represent the presence of QTM/TA-QTM between the connecting pairs. The numbers provided at each arrow are the mean absolute value for the corresponding matrix element of the transition magnetic moment.

In the magnetic relaxation diagram of the TbIII ions of 2-Tb, the ground pseudo-doublet state is a mixture of mJ = |±6〉 and mJ = |±4〉 states. Due to the highly mixed mJ levels, this results in a non-negligible tunnel splitting of 0.73 cm−1. Hence, strong mixing causes tunneling and quenches the out-of-phase signals at zero dc field. In the case of 4-Ho, the ground pseudo-doublet state possesses a highly axial gzz value of 19.08, which is closer to the expected value of 20,55 indicating the strong anisotropic nature of HoIII ions. However, the mixing of the ground mJ = |±8〉 state with the mJ = |±4〉 state causes a large tunnelling gap of 1.93 cm−1, and induces QTM. In contrast, slow magnetic relaxation for 4-Ho occurs in the absence of an applied magnetic field. Therefore, the SMM behaviour of 4-Ho cannot be accounted for by only considering the single-ion behavior of the HoIII ions. The tunneling gap/QTM could be significantly reduced by applying a DC field, as shown in Fig. S4, or by quenching it with a strong magnetic exchange (Fig. 7).56,57


image file: d5qi01665g-f7.tif
Fig. 7 Low-lying exchange spectra in 1-Gd, 2-Tb, 3-Dy, and 4-Ho. The exchange states are placed on the diagram according to their magnetic moments (bold black lines). The red arrows show the tunneling transitions (energy splitting) within each doublet state, while the green/blue arrows show the possible pathway through Orbach/Raman relaxation. The numbers on the paths are averaged transition moments in μB, connecting the corresponding states.

The role of the anisotropic nature of RuIII ions needs to be considered in these complexes, especially for 1-Gd, where the origin of the SMM behaviour under an applied DC field can be accounted for based on its anisotropic behaviour. The RuIII ions’ (t2g5) unquenched orbital angular momentum is the result of the 4d orbitals’ large crystal field splitting and anisotropy. This splitting is a consequence of the strong interaction between the 4d electrons of the RuIII ions and the surrounding ligand field, which significantly alters the energy levels of the d-orbitals, stabilizing the 2T2g configuration as the ground state.58 The ground-state KDs of the RuIII sites exhibit a rhombic set of g-tensors for all four complexes (Table S12); this is anticipated to impact the SMM features, and dominantly in the case of 1-Gd. The overall magnetic dynamics of these complexes are anticipated to originate from the exchange-coupled state of the entire molecule. According to the single-ion viewpoint, complex 3-Dy should have the best SMM parameters among the five complexes. However, the overall magnetic relaxation will be determined by a combination of the exchange-coupled state and the single-ion behaviour of both LnIII and RuIII ions.

Exchange-coupled states relaxation mechanism

The exchange-coupled states relaxation mechanisms for 1-Gd, 2-Tb, 3-Dy, 4-Ho, and 5-Er were developed considering all the possible exchange interactions using the POLY_ANISO program within the Lines model59 (see Computational details) and by fitting the experimental magnetic data. There are three types of magnetic exchange interactions for each complex: (i) four J1(Ru⋯Ln) exchange interactions (Ln1–Ru1, Ln1–Ru2, Ln2–Ru1, and Ln2–Ru2), (ii) one J2(Ln⋯Ln) exchange interaction, and (iii) one J3(Ru⋯Ru) exchange coupling, as shown in Fig. S10. The calculated J values are tabulated in Table 2. All five complexes possess moderate antiferromagnetic J1(Ru–Ln) interactions and nearly identical and weak ferromagnetic J2(Ln–Ln) exchange interactions. Antiferromagnetic exchange couplings exist between Ru⋯Ru metal ions for all complexes, where 1-Gd, 2-Tb, 3-Dy and 4-Ho have similar values (∼−0.05 cm−1); however, 5-Er has the largest (−0.16 cm−1). The intermolecular exchange interaction is found to be antiferromagnetic for all five complexes, which accounts for the decrease in χMT at the lowest temperatures.
Table 2 Exchange coupling constants (J; in cm−1) were estimated using POLY_ANISO and DFT calculations for 1-Gd, 2-Tb, 3-Dy, 4-Ho, and 5-Er
  DFT Lines model
J (cm−1) 1-Gd 1-Gd 2-Tb 3-Dy 4-Ho 5-Er
J 1 (Ru–Ln) −1.32 −0.4 −1.8 −2.7 −2.8 −4.0
J 2 (Ln–Ln) 0.01 0.01 0.03 0.05 0.01 0.01
J 3 (Ru–Ru) −0.24 −0.05 −0.05 −0.05 −0.05 −0.16
zJ −0.001 −0.015 −0.014 −0.015 −0.034


To verify the POLY_ANISO computed exchange interactions between the metal centers, the DFT calculations were performed by utilizing Noodleman's Broken Symmetry approach60 in the Gaussian 16 software61 (See Computational details). Different spin configurations used to calculate the exchange interactions are shown in Table S15. DFT calculations yield antiferromagnetic exchange (JAF) coupling values for J1(Gd–Ru) = −1.32 cm−1, ferromagnetic exchange (JF) value for J2(Gd–Gd) = ∼0.01 cm−1, and antiferromagnetic exchange value of J3(Ru–Ru) = −0.24 cm−1, leading to an S = 6 ground state. These DFT-computed J values are in good agreement with PHI-fitted J values (see Table 2) and also align well with the POLY_ANISO computed J values for other complexes, with slight variation in their magnitudes (see more discussion in SI). Among all the complexes, 5-Er possesses the largest J1(Ru⋯Ln) exchange interaction, which is a common observation for Er-based complexes, and this can be attributed to the prolate nature of erbium's electron density.62,63 For 2-Tb, 3-Dy, and 4-Ho, which are oblate-Ln ions, the Ru–O–Ln angle is found to be 99.1°, 99.1°, and 98.9°, respectively, which is in accordance with their oblate nature (Tb ∼ Dy > Ho), and their J1(Ru⋯Ln) values are also in the reverse order of oblate nature. Whereas, the Gd ion is spherical in shape, giving the smallest J1(Ru⋯Gd) value among all. The overall J1(Ru–Ln) exchange interaction is dominant among all three exchange interactions and follows the trend in the order of complexes 5-Er > 4-Ho > 3-Dy > 2-Tb > 1-Gd. This trend, as elucidated above, can therefore be attributed to the gradual transition in the nature of the 4f electron density distribution of the lanthanide ions, which increases when they possess prolate Ln ions and decreases for spherical Gd ions. Furthermore, this can be linked to the combination of JF and JAF, leading to the overall J value. The number of orbital overlaps is greatest for RuIII–GdIII (one 4d orbital and seven 4f orbitals), with some of these contributing to JF, reducing the magnitude of the antiferromagnetic J1(Ru–Gd) exchange.64–67 As we move across the series, the number of magnetic exchange overlaps is reduced (e.g., only three overlaps in Ru–Er, i.e., one 4d orbital and three 4f orbitals, which lowers the JF part), resulting in the increment of the magnitude of antiferromagnetic J1(Ru–Ln) exchange (5-Er > 4-Ho > 3-Dy > 2-Tb > 1-Gd).

Fig. 7 and S15 show the POLY_ANISO computed exchange-coupled states’ mechanism of magnetic relaxation of 1-Gd, 2-Tb, 3-Dy, 4-Ho, and 5-Er that was developed using all three exchange coupling constants. The two lanthanide metal ions in each complex, possessing ground Kramers/Ising doublet states, coupled with two RuIII ions, yielded a non-Kramers exchange spectrum for all complexes. The lowest exchange doublets from the spectrum with their tunnel splitting, and the gz value of each doublet are given in Tables S16–20. In the case of 3-Dy, a very small tunneling gap (∼10−6 to 10−4 cm−1) was observed at the ground state as well as in lower excited states. This gap is enlarging at the 12th excited state, which allows the magnetic relaxation to happen via 12th excited exchange states and yields an energy barrier (Ucal) of 111 cm−1 (158.7 K). This corroborates the observation of SMM behaviour for complex 3-Dy in the absence of a DC field, but this Ucal value is overestimated, and this can be attributed to the exclusion of hyperfine interactions in our calculations.68,69 Similarly, for complexes 2-Tb and 4-Ho, the tunneling gap is small in the ground state, but it becomes larger in the 1st excited state, which allows the magnetic relaxation with a barrier of 3.9 cm−1 (5.6 K), and 6.9 cm−1 (10 K), respectively. These values suggest that 2-Tb and 4-Ho can show SMM behaviour with a small dc field or without a dc field, which is in accordance with the experimental observation, but slightly underestimated. Though comparatively a large tunnelling gap between the ground doublet is predicted at each Ho(III) single-ion for 4-Ho, resulting in efficient QTM, the Ru–Ho exchange interaction raised the degeneracy and produced a ladder of exchange-coupled states, which in turn helped to reduce the tunnelling gap. Slow magnetic relaxation takes over as a result of the suppression of the tunnelling gap at the ground exchange-doublet states, which supports the observation of the maxima in the AC magnetic susceptibility data at lower temperatures from 1.8–3.5 K in the absence of a dc magnetic field. Thus, the interplay between the large intrinsic tunnelling gap and the exchange-induced level splitting explains why Ho exhibits observable zero-field dynamics despite its unfavorable single-ion characteristics. For 1-Gd, the ground state, as well as several low-lying excited states, possess a very small tunneling gap (∼10−8 to 10−3 cm−1), but these states are stacked very closely, leading to a negligible energy barrier of 0.3 cm−1 (0.4 K). Whereas for 5-Er, the tunneling gap is large even at the ground state, allowing for relaxation with a negligible energy barrier (Ucal). These negligible Ucal values of 1-Gd and 5-Er support the field-induced SMM behaviour observed from the experiment.

A correlation has been developed to compare the energy barriers from single-ion relaxation of lanthanide ions, the exchange-coupled states mechanism, and experimental data with the RuIII–LnIIIJ values (Fig. 8). At the single-ion level, the first excited KDs/IDs values are taken considering a low QTM and tunneling gap that allows relaxation in those states, or it could be quenched upon exchange or application of a dc field. The correlation suggests that 3-Dy has a better Ucal/Ueff than other complexes, which could be attributed to the combination of oblate electron density that favors axial anisotropy and a moderate J (Ru–Ln) value. This balance between electron density distribution and exchange strength stabilizes the relaxation barrier and enhances SMM behaviour. For the 2-Tb and 4-Ho, the Ueff falls between the Ucal from single-ion and Ucal from coupled-states, which hints that the tunneling gap between the ground IDs of oblate TbIII and HoIII ions is quenched with the combined effect of moderate J1(Ru–Ln) value and an applied dc magnetic field. Though 5-Er has a greater J value among all, the prolate electron density disfavors the exchange-coupled relaxation and leads to weak SMM behaviour even upon the application of a dc field. Hence, in the Er analogue, the exchange interactions are insufficient to effectively separate the excited states, which therefore remain closer in energy to the single-ion levels, further limiting its performance. Overall, due to strong Ru⋯Ln exchange interactions, the Ucal value derived from the POLY_ANISO computations of all the complexes was improved compared to their single-ion computations. For example: (i) for 3-Dy the Ucal of the individual DyIII ion is 101.8 cm−1, which has increased to 111 cm−1 when they are coupled with the RuIII ion; (ii) for 2-Tb and 4-Ho at the single-ion level, both the TbIII and HoIII ions, showed a large tunneling gap in their GS itself, which was quenched when they coupled with RuIII ions and yielded reasonable energy barrier of 3.9 cm−1 and 6.9 cm−1, respectively, importantly we observed zero-field SMM behaviour for 4-Ho, highlighting the importance of the exchange interaction; (iii) the larger QTM observed between the ground KDs of ErIII ion was also quenched to some extent in the exchange-coupled mechanism; (iv) for 1-Gd the isotropic GdIII ions coupled with the anisotropic RuIII ions with a moderate Ru–Gd exchange interaction resulted in a ground state value of S = 6, and showed SMM behavior, becoming a rare example of Gd-based SMMs.70,71


image file: d5qi01665g-f8.tif
Fig. 8 Comparison of exchange interaction (J, cm−1) with experimental energy barrier (Ueff, K) and computed energy barrier from single-ion as well as exchange-coupled states (Ucal, K) of 1-Gd, 2-Tb, 3-Dy, 4-Ho, and 5-Er.

The observed higher centrosymmetric space group P[1 with combining macron], in these complexes, seemingly, brought small geometric changes with different lanthanide ions that have significant effects on the balance between anisotropy and magnetic exchange interactions, even though the overall crystallographic symmetry is invariant. Such an effect of symmetry and space groups of crystal structures usually has an important influence on magnetism, which was proven by single-crystal measurements for a few 3d–4f complexes;62 however, our detailed ab initio calculations provide an adequate explanation here. For example, a stronger transverse anisotropy component is introduced by the 5-Er, which significantly increases the QTM. Conversely, the 3-Dy and 4-Ho that contain Dy and Ho exhibit a more axial form of anisotropy, favoring slower, thermally induced relaxation pathways. Additionally, determining the magnetic exchange between the metal centres is largely dependent on the spatial arrangement and relative orientation of the Ln⋯Ln vectors as well as the Ru–Ln exchange pathways integrated into the inversion-symmetric P[1 with combining macron] space group. The intricate interactions among global symmetry, exchange topology, and local coordination geometry in controlling the single-ion and collective magnetic characteristics of this Ru–Ln compound family are emphasised in this study.

As stated earlier, some of us reported {RuIII2DyIII2} butterfly complexes bearing energy barriers of 15.4 K (ref. 45) and 45 K.47 In the present work, the effective energy barrier is enhanced to 60 K in 3-Dy due to the improvement of the magnetic exchange interaction between RuIII⋯DyIII from −2.4 cm−1 to −2.7 cm−1, which can be attributed to the presence of more acidic o-toluate ligands, which helped to provide better alignment of anisotropic directions compared with the previously reported benzoate {RuIII2DyIII2} analog.45,47 In the previously reported cases, the anisotropic directions of Dy⋯Dy ions are aligned parallel with the Ru⋯Ru anisotropic direction. However, in 3-Dy, these anisotropic directions are not completely parallel to each other, which eventually helps in the enhancement of the 4d–4f exchange interaction.

Compared to 3d–4f magnetic exchange interaction as explored in {CrIII2DyIII2} complexes, the RuIII⋯LnIII exchange interaction is found to be smaller, irrespective of the ligand and LnIII metal ions used. For example, the [CrIII2DyIII2(OMe)2(O2CPh)4(mdea)2(NO3)2] complex27 shows CrIII–DyIII magnetic exchange interaction (J) of −20.7 to −16.3 cm−1 with Ueff = 54 cm−1 and TB = 3.7 K, and [CrIII2DyIII2(OMe)2(acac)4(mdea)2(NO3)2], shows J (Cr–Dy) of −11.2 to −8.3 cm−1 with Ueff = 24 cm−1 and TB = 1.8 K.28 The J values are better compared to the reported Ru–Ln complexes; however, the Ueff values are similar to the RuIII–DyIII complexes. This is because the RuIII ion not only participates in the exchange interaction but also lifts the degeneracy of the crystal-field splitting due to its anisotropic nature, improving the magnetization reversal barrier. Whereas the CrIII ion is isotropic, which helps only in the exchange part.

Conclusions

We demonstrated the utilization of the magnetic exchange interactions and magnetic anisotropies in quenching QTM in five {RuIII2LnIII2} butterfly complexes (1-Gd, 2-Tb, 3-Dy, 4-Ho, and 5-Er) by detailed experimental and computational studies. Among them, 2-Tb, 4-Ho, and 5-Er are the first set of non-DyIII-based Ru-4f complexes. AC magnetic studies suggested that 3-Dy and 4-Ho show slow magnetic relaxation in the absence of an external magnetic field with Ueff of 60 K and 20 K, respectively. Whereas 1-Gd, 2-Tb, and 5-Er show field-induced SMM behavior with a Ueff of 5.5 K, 30.5 K, and 38.9 K, under an applied dc field of 3600 Oe, 2000 Oe, and 2400 Oe, respectively. An anisotropy barrier (Ueff) of 60 K was determined for 3-Dy, representing the highest value reported to date for 4d–4f butterfly-type complexes. For 4-Ho, zero-field SMM behavior is recognized, which is a rare occurrence for complexes containing this ion due to the usual large QTM, making it a unique addition of a zero-field HoIII SMM and the first {Ru–Ln} SMM not containing a DyIII ion. Additionally, the field-induced SMM for 1-Gd is a rare Gd-based SMM and expands the diversity of heterometallic complexes in the field of molecular magnetism. The CASSCF/SO-RASSI/SINGLE_ANISO/POLY_ANISO/DFT calculations support the experimental observation of SMM behavior for all five complexes. These calculations predict that the Ru⋯Ln exchange interaction plays a major role in enhancing the barrier height and quenching QTM in the case of 1-Gd, 2-Tb, 3-Dy, and 4-Ho, demonstrating the potential of using RuIII ions in SMM design.

Overall, the present work reveals that a strong magnetic exchange interaction can be found between RuIII and LnIII ions (−0.4 to −4 cm−1), and highlights the importance of the RuIII⋯LnIII magnetic exchange interaction for observing SMM behaviour with LnIII ions such as DyIII and HoIII. The dominant Ru–Ln exchange interaction progresses from 5-Er to 4-Ho, 3-Dy, 2-Tb, and 1-Gd, and this pattern is correlated to the number of orbital overlaps that determines the magnitude of the Ru-Ln exchange interaction. Furthermore, we showed improvements of the effective anisotropy barrier compared to previously reported {RuIII2DyIII2} butterfly complexes by increasing the exchange interaction via ligand manipulation.45,47

Experimental and computational methodologies

Synthesis procedures

Materials and methods. All reactions were conducted under aerobic conditions. Reagents and solvents were sourced commercially and utilized without additional purification.
Synthesis of [RuIII2GdIII2(OMe)2(o-tol)4(mdea)2(NO3)2] (1-Gd). A similar method to that used to make [RuIII2DyIII2(OMe)2(O2CPh)4(mdea)2(NO3)2] was employed.45,47 RuCl3·xH2O (0.25 mmol) and Gd(NO3)3·6H2O (0.25 mmol) were dissolved in 15 mL of acetonitrile (MeCN). ortho-Toluic acid (0.5 mmol), N-methyldiethanolamine (0.25 mmol), and triethylamine (1.0 mmol) were then added, resulting in a dark brown solution while being heated to 60 °C. The mixture was stirred at this temperature for 6 hours, after which the solvent was evaporated, leaving behind a brown oil. This brown oil was re-dissolved in methanol (MeOH) and layered with diethyl ether (Et2O). Within 2 weeks, brown crystals were isolated, which are suitable for single-crystal X-ray diffraction. Brown-coloured crystals of 1-Gd appeared within 2 weeks, with an approximate yield (based on Gd(NO3)3 of 1-Gd) of 47% (crystalline product). Anal. Calculated (found) for 1-Gd: Ru2Gd2C44H56O20N4: C, 35.77 (35.55); H, 3.82 (3.54); N, 3.79 (3.51). IR (KBr, cm−1): 3432 (m), 3063 (w), 2959 (w), 2928 (w), 2863 (w), 2826 (w), 1606 (m), 1583 (m), 1553 (s), 1463 (s), 1441 (w), 1383 (s), 1300 (m), 1157 (w), 1061 (m), 1010 (m), 900 (w), 857 (w), 814 (w), 788 (w), 741 (m), 669 (w), 612 (w).
Synthesis of [RuIII2TbIII2(OMe)2(o-tol)4(mdea)2(NO3)2] (2-Tb). The synthetic method for 1-Gd was carried out, but Tb(NO3)3·6H2O (0.25 mmol) was used instead of Gd(NO3)3·6H2O. Brown-coloured crystals of 2-Tb appeared within 2 weeks, with an approximate yield (based on Tb(NO3)3 of 2-Tb) of 45% (crystalline product). Anal. Calculated (found) for 2-Tb: Ru2Tb2C44H56O20N4: C, 35.69 (35.60); H, 3.81 (3.76); N, 3.78 (3.89). IR (KBr, cm−1): 3432 (m), 3063 (w), 2959 (w), 2926 (w), 2863 (w), 2824 (w), 1606 (m), 1583 (m), 1555 (s), 1461 (s), 1439 (w), 1381 (s), 1302 (m), 1155 (w), 1061 (m), 1010 (m), 900 (w), 857 (w), 788 (w), 741 (m), 669 (w), 610 (w), 514 (w).
Synthesis of [RuIII2DyIII2(OMe)2(o-tol)4(mdea)2(NO3)2] (3-Dy). The synthetic method for 1-Gd was carried out, but Dy(NO3)3·6H2O (0.25 mmol) was used instead of Gd(NO3)3·6H2O. Within 2 weeks, brown crystals were isolated, which are suitable for single-crystal X-ray diffraction. Yield (based on Dy(NO3)3) of 3-Dy = 42% (crystalline product). Elemental Analysis: calculated (found) for 3-Dy: Ru2Dy2C44H56O20N4: C, 35.51 (35.60); H, 3.79 (3.51); N, 3.77 (4.01). IR (KBr, cm−1): 3430 (m), 3061 (w), 2963 (w), 2930 (w), 2863 (w), 2828 (w), 1606 (m), 1583 (m), 1553 (s), 1465 (s), 1441 (w), 1381 (s), 1302 (m), 1157 (w), 1061 (m), 1010 (m), 902 (w), 855 (w), 816 (w), 790 (w), 741 (m), 669 (w), 610 (w).
Synthesis of [RuIII2HoIII2(OMe)2(o-tol)4(mdea)2(NO3)2] (4-Ho). The synthetic method for 1-Gd was carried out, but Ho(NO3)3·5H2O (0.25 mmol) was used instead of Gd(NO3)3·6H2O. Brown-coloured crystals of 4-Ho appeared within 2 weeks, with an approximate yield (based on Ho(NO3)3 of 4-Ho) of 47% (crystalline product). Anal. Calculated (found) for 4-Ho: Ru2Ho2C44H56O20N4: C, 35.40 (35.73); H, 3.78 (3.31); N, 3.75 (4.21). IR (KBr, cm−1): 3432 (m), 3063 (w), 2961 (w), 2930 (w), 2863 (w), 2828 (w), 1606 (m), 1583 (m), 1553 (s), 1467 (s), 1441 (w), 1381 (s), 1304 (m), 1157 (w), 1059 (m), 1010 (m), 902 (w), 857 (w), 816 (w), 790 (w), 741 (m), 669 (w), 610 (w).
Synthesis of [RuIII2ErIII2(OMe)2(o-tol)6(mdea)2]·2MeOH (5-Er). The synthetic method for 1-Gd was carried out, but Er(NO3)3·5H2O (0.25 mmol) was used instead of Gd(NO3)3·6H2O. Brown-coloured crystals of 5-Er appeared within 2 weeks, with an approximate yield (based on Er(NO3)3 of 5-Er) of 45% (crystalline product). Anal. Calculated (found) for 5-Er: Ru2Er2C61H74O19N2: C, 43.60 (43.23); H, 4.60 (4.53); N, 1.64 (1.91). IR (KBr, cm−1): 3432 (m), 3063 (w), 2957 (w), 2930 (w), 2863 (w), 2828 (w), 1606 (m), 1583 (m), 1553 (s), 1467 (s), 1439 (w), 1381 (s), 1306 (m), 1157 (w), 1061 (m), 1008 (m), 902 (w), 857 (w), 816 (w), 786 (w), 743 (m), 669 (w), 614 (w), 514 (w).

X-ray crystallography

Single-crystal X-ray diffraction data for 3-Dy, 4-Ho, and 5-Er were collected using a Bruker diffractometer with Mo Kα radiation (λ = 0.71073 Å). Crystals were selected under mineral oil, mounted on fiber loops, and cooled using an Oxford Cryo-system. A complete hemispherical dataset was acquired using ω and φ scan techniques (1° per frame, 2 seconds per frame). Cell refinement and data reduction were carried out using CrysAlisPro 1.171.38.43. The structures were solved by direct methods and refined by full-matrix least-squares techniques on F2 using SHELXL as implemented in Olex2-1.5. All non-hydrogen atoms were refined anisotropically, while hydrogen atoms were placed in calculated positions and refined using a riding model. The CCDC numbers are 2456392–2456394. Crystallographic data and refinement parameters are summarized in Table S1 in the SI.

Powder X-ray diffraction (PXRD) data for 1-Gd and 2-Tb, as well as for other samples, were collected on a Rigaku Ultima IV fully automatic high-resolution X-ray diffractometer system with a Theta–Theta (θθ) Goniometer, which was utilized for routine characterization of powder samples. Scans were performed at room temperature in the 2θ range 5–50° and compared with predicted patterns based on low-temperature single-crystal X-ray data.

Magnetic measurements

Magnetic susceptibility measurements were carried out on a Quantum Design SQUID magnetometer MPMS-3 and PPMS, operated between 1.8 and 300 K for DC-applied fields ranging from 0 to 70 kOe and 0 to 90 kOe, respectively. To prevent the crystallites from torquing, microcrystalline samples were distributed in Vaseline. A calibrated gelatin capsule containing the sample mulls was fixed at the end of the sample rod and held in the middle of a drinking straw. Alternating current (AC) magnetic susceptibility measurements were carried out under an oscillating AC field with frequencies ranging from 0.1 Hz to 1 kHz (MPMS-3) and 10 Hz to 10 kHz (PPMS).

Computational details

Ab initio calculations. The ab initio calculations were performed using MOLCAS 8.272 on the anisotropic metal ions TbIII, DyIII, HoIII, ErIII, and RuIII in complexes 2-Tb, 3-Dy, 4-Ho, and 5-Er. Calculations for 3-Dy, 4-Ho, and 5-Er were performed using their X-ray structures, whereas for 2-Tb, the X-ray structure of 3-Dy was utilized by replacing DyIII with TbIII. The Douglas-Kroll Hamiltonian is employed to incorporate relativistic effects.73 The Complete Active Space Self-Consistent Field (CASSCF) method is employed to compute the spin-free eigenstates.74 The following basis sets are used for the respective metal ions such as [ANO-RCC…8s7p5d3f2g1h.] basis set for paramagnetic Ln(III) ions, [ANO-RCC…7s6p4d3f1g.] for Lu(III), [ANO-RCC…6s5p3d1f.] for Ru(III) ion, [ANO-RCC…6s5p3d.] for Y(III) ion, [ANO-RCC…3s2p.] for C atoms, [ANO-RCC…3s2p.] for N atoms, [ANO-RCC…4s3p2d.] for O atoms and [ANO-RCC…2s.] for H atoms.75 The neighboring LnIII (TbIII, DyIII, HoIII, and ErIII) and RuIII ions were replaced by diamagnetic LuIII and YIII, respectively, while performing the single-ion calculations. In the CASSCF calculations, the active spaces (8,7), (9,7), (10,7), and (11,7) were considered for the TbIII, DyIII, HoIII, and ErIII ions, respectively, and (5,5) in the case of RuIII ion. To compute the SOC, 7 septet, 140 quintet, and 195 triplet states for TbIII; 21 sextet, 224 quartet, and 490 doublet states for DyIII; 35 quintet, 210 triplet, and 195 singlet states for HoIII; and 35 quartet and 112 doublet states for ErIII; and 1 sextet, 24 quartet, and 75 doublet states for RuIII were considered. After calculating the excited spin–orbit states, we mixed the roots that are below 50000 cm−1 using RASSI-SO.51 Moreover, the mixed spin–orbit (SO) states were used in the SINGLE_ANISO program to calculate the g-tensors.76 We have also determined the crystal-field parameters using the SINGLE_ANISO module in MOLCAS 8.2 by following the equation below.
 
image file: d5qi01665g-t1.tif(3)
POLY_ANISO calculations. The observed magnetic data for each complex were analyzed using the POLY_ANISO software48–50 within the Lines model to determine the exchange interaction between the metal ions (eqn (4)).
 
image file: d5qi01665g-t2.tif(4)
(here Ji = Jdipolari + Jexchi; i.e., Ji is the total magnetic interaction of the fitted Jdipolari and Jexchi parameters).
Density functional theory (DFT) calculations. To compute the exchange interactions in the Gd analog (1-Gd), DFT calculations were performed using the B3LYP functional77 within the Gaussian 16 software package.61 BS-DFT calculations were performed by utilizing the X-ray structure of 3-Dy and replacing DyIII with GdIII, while the positions of the other atoms were kept consistent with those determined by X-ray crystallography. We employed the Cundari–Stevens (CS) relativistic effective core potential (ECP) on Gd,78 def2-TZVP basis set for Ru, 6-31G** basis set for C, N, O, H. The magnetic exchange (J) values were calculated using the Broken Symmetry (BS) method60 in association with DFT computations.

The exchange interaction (J) is calculated by using the following Hamiltonian:

 
image file: d5qi01665g-t3.tif(5)

Conflicts of interest

There are no conflicts to declare.

Data availability

All data supporting the findings of this study are included in the supplementary information (SI). Supplementary information: crystallographic data, PXRD, magnetic plots and tables, and computational analysis plots and tables. See DOI: https://doi.org/10.1039/d5qi01665g.

CCDC 2456392–2456394 contain the supplementary crystallographic data for this paper.79a–c

Acknowledgements

I. J. D. thanks the DST-INSPIRE (IF220060) for providing a PhD fellowship. V. K. is thankful to the Prime Minister's Research Fellowship for financial assistance. M. S. thanks the funding agencies SERB (CRG/2023/002178), BRNS (58/14/07/2023-BRNS/37029), CSIR (01/3092/23/EMR-II), IISc-STARs (MoE-STARS/STARS-2/2023-0158), and IIT Bombay for the financial support. K. R. V. thanks the IISER Mohali-Central PXRD facility, IISER Mohali-High-Performance Computing Facility, and Prof. Gopalan Rajaraman, IIT Bombay, for providing computing resources. K. R. V. also thanks SERB for providing start-up research grant funding (SRG/2023/000286).

References

  1. M. N. Leuenberger and D. Loss, Quantum computing in molecular magnets, Nature, 2001, 410, 789–793 CrossRef CAS PubMed.
  2. A. Ardavan, O. Rival, J. J. L. Morton, S. J. Blundell, A. M. Tyryshkin, G. A. Timco and R. E. P. Winpenny, Will spin-relaxation times in molecular magnets permit quantum information processing?, Phys. Rev. Lett., 2007, 98, 057201 CrossRef.
  3. S. Thiele, F. Balestro, R. Ballou, S. Klyatskaya, M. Ruben and W. Wernsdorfer, Electrically driven nuclear spin resonance in single-molecule magnets, Science, 2014, 344, 1135–1138 CrossRef CAS.
  4. J. M. Zadrozny, J. Niklas, O. G. Poluektov and D. E. Freedman, Millisecond coherence time in a tunable molecular electronic spin qubit, ACS Cent. Sci., 2015, 1, 488–492 CrossRef CAS.
  5. M. Shiddiq, D. Komijani, Y. Duan, A. Gaita-Ariño, E. Coronado and S. Hill, Enhancing coherence in molecular spin qubits via atomic clock transitions, Nature, 2016, 531, 348–351 CrossRef CAS.
  6. M. F. Crommie, Manipulating magnetism in a single molecule, Science, 2005, 309, 1501–1502 CrossRef CAS PubMed.
  7. M. Urdampilleta, N.-V. Nguyen, J.-P. Cleuziou, S. Klyatskaya, M. Ruben and W. Wernsdorfer, Molecular quantum spintronics: Supramolecular spin valves based on single-molecule magnets and carbon nanotubes, Int. J. Mol. Sci., 2011, 12, 6656–6667 CrossRef CAS PubMed.
  8. L. E. Darago, M. D. Boshart, B. D. Nguyen, E. Perlt, J. W. Ziller, W. W. Lukens, F. Furche, W. J. Evans and J. R. Long, Strong ferromagnetic exchange coupling and single-molecule magnetism in MoS43−-bridged dilanthanide complexes, J. Am. Chem. Soc., 2021, 143, 8465–8475 CrossRef CAS PubMed.
  9. 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.
  10. R. Sessoli, D. Gatteschi, A. Caneschi and M. A. Novak, Magnetic bistability in a metal-ion cluster, Nature, 1993, 365, 141–143 CrossRef CAS.
  11. R. Tiron, W. Wernsdorfer, N. Aliaga-Alcalde and G. Christou, Quantum tunneling in a three-dimensional network of exchange-coupled single-molecule magnets, Phys. Rev. B: Condens. Matter Mater. Phys., 2003, 68, 140407 CrossRef.
  12. J. D. Rinehart, M. Fang, W. J. Evans and J. R. Long, Strong exchange and magnetic blocking in N23−-radical-bridged lanthanide complexes, Nat. Chem., 2011, 3, 538–542 CrossRef CAS PubMed.
  13. Y. Peng and A. K. Powell, What do 3d–4f butterflies tell us?, Coord. Chem. Rev., 2021, 426, 213490 CrossRef CAS.
  14. N. Ishikawa, M. Sugita, T. Ishikawa, S.-y. Koshihara and Y. Kaizu, Mononuclear lanthanide complexes with a long magnetization relaxation time at high temperatures: A new category of magnets at the single-molecular level, J. Phys. Chem. B, 2004, 108, 11265–11271 CrossRef CAS.
  15. 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.
  16. A. H. Vincent, Y. L. Whyatt, N. F. Chilton and J. R. Long, Strong axiality in a dysprosium(III) bis (borolide) complex leads to magnetic blocking at 65 K, J. Am. Chem. Soc., 2023, 145, 1572–1579 CrossRef CAS.
  17. S. K. Gupta, T. Rajeshkumar, G. Rajaraman and R. Murugavel, An air-stable Dy(III) single-ion magnet with high anisotropy barrier and blocking temperature, Chem. Sci., 2016, 7, 5181–5191 RSC.
  18. A. B. Canaj, S. Dey, E. R. Martí, C. Wilson, G. Rajaraman and M. Murrie, Insight into D6h Symmetry: Targeting Strong Axiality in Stable Dysprosium(III) Hexagonal Bipyramidal Single–Ion Magnets, Angew. Chem., Int. Ed., 2019, 58, 14146–14151 CrossRef CAS.
  19. A. S. Armenis, A. Mondal, S. R. Giblin, C. P. Raptopoulou, V. Psycharis, D. I. Alexandropoulos, J. Tang, R. A. Layfield and T. C. Stamatatos, Unveiling new [1 + 1] Schiff-base macrocycles towards high energy-barrier hexagonal bipyramidal Dy(III) single-molecule magnets, Chem. Commun., 2024, 60, 12730–12733 RSC.
  20. A. S. Armenis, A. Mondal, S. R. Giblin, D. I. Alexandropoulos, J. Tang, R. A. Layfield and T. C. Stamatatos, ‘Kick-in the head’: high-performance and air-stable mononuclear DyIII single-molecule magnets with pseudo-D6h symmetry from a [1 + 1] Schiff-base macrocycle approach, Inorg. Chem. Front., 2025, 12, 1214–1224 RSC.
  21. H. Kwon, K. R. McClain, J. G. C. Kragskow, J. K. Staab, M. Ozerov, K. R. Meihaus, B. G. Harvey, E. S. Choi, N. F. Chilton and J. R. Long, Coercive fields exceeding 30 T in the mixed-valence single-molecule magnet (CpiPr5)2Ho2I3, J. Am. Chem. Soc., 2024, 146, 18714–18721 CrossRef CAS.
  22. 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, Ultrahard magnetism from mixed-valence dilanthanide complexes with metal-metal bonding, Science, 2022, 375, 198–202 CrossRef CAS PubMed.
  23. D. N. Woodruff, R. E. P. Winpenny and R. A. Layfield, Lanthanide single-molecule magnets, Chem. Rev., 2013, 113, 5110–5148 CrossRef CAS.
  24. L. R. Piquer and E. C. Sañudo, Heterometallic 3d–4f single-molecule magnets, Dalton Trans., 2015, 44, 8771–8780 RSC.
  25. K. Liu, W. Shi and P. Cheng, Toward heterometallic single-molecule magnets: Synthetic strategy, structures and properties of 3d–4f discrete complexes, Coord. Chem. Rev., 2015, 289, 74–122 CrossRef.
  26. J. L. Liu, J. Y. Wu, Y. C. Chen, V. Mereacre, A. K. Powell, L. Ungur, L. F. Chibotaru, X. M. Chen and M. L. Tong, A heterometallic FeII–DyIII single–molecule magnet with a record anisotropy barrier, Angew. Chem., Int. Ed., 2014, 53, 12966–12970 CrossRef CAS PubMed.
  27. S. K. Langley, D. P. Wielechowski, V. Vieru, N. F. Chilton, B. Moubaraki, B. F. Abrahams, L. F. Chibotaru and K. S. Murray, A {CrIII2DyIII2} single-molecule magnet: enhancing the blocking temperature through 3d magnetic exchange, Angew. Chem., Int. Ed., 2013, 52, 12014–12019 CrossRef CAS PubMed.
  28. S. K. Langley, D. P. Wielechowski, V. Vieru, N. F. Chilton, B. Moubaraki, L. F. Chibotaru and K. S. Murray, Modulation of slow magnetic relaxation by tuning magnetic exchange in {Cr2Dy2} single molecule magnets, Chem. Sci., 2014, 5, 3246–3256 RSC.
  29. S. K. Langley, D. P. Wielechowski, B. Moubaraki and K. S. Murray, Enhancing the magnetic blocking temperature and magnetic coercivity of {CrIII2LnIII2} single-molecule magnets via bridging ligand modification, Chem. Commun., 2016, 52, 10976–10979 RSC.
  30. S. K. Langley, C. M. Forsyth, B. Moubaraki and K. S. Murray, A fluoride bridged {CrIII4DyIII4} single molecule magnet, Dalton Trans., 2015, 44, 912–915 RSC.
  31. S. K. Langley, N. F. Chilton, B. Moubaraki and K. S. Murray, Single-molecule magnetism in three related {CoIII2DyIII2}-acetylacetonate complexes with multiple relaxation mechanisms, Inorg. Chem., 2013, 52, 7183–7192 CrossRef CAS PubMed.
  32. 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.
  33. D. Schray, G. Abbas, Y. Lan, V. Mereacre, A. Sundt, J. Dreiser, O. Waldmann, G. E. Kostakis, C. E. Anson and A. K. Powell, Combined magnetic susceptibility measurements and 57Fe mössbauer spectroscopy on a ferromagnetic {FeIII4Dy4} ring, Angew. Chem., Int. Ed., 2010, 49, 5185–5188 CrossRef CAS PubMed.
  34. Y. Peng, J. Braun, M. Schulze, H. Kaemmerer, Y. F. Schneider, C. E. Anson, W. Wernsdorfer and A. K. Powell, A nested spin structure and single molecule magnet behaviour in an Fe8Dy12 heterometallic cyclic coordination cluster, Dalton Trans., 2024, 53, 894–897 RSC.
  35. N. Ahmed, C. Das, S. Vaidya, S. K. Langley, K. S. Murray and M. Shanmugam, Nickel(II)–Lanthanide (III) Magnetic Exchange Coupling Influencing Single–Molecule Magnetic Features in {Ni2Ln2} Complexes, Chem. – Eur. J., 2014, 20, 14235–14239 CrossRef CAS PubMed.
  36. V. Chandrasekhar, P. Bag, W. Kroener, K. Gieb and P. Müller, Pentanuclear heterometallic {Ni2Ln3}(Ln = Gd, Dy, Tb, Ho) assemblies. Single-molecule magnet behavior and multistep relaxation in the dysprosium derivative, Inorg. Chem., 2013, 52, 13078–13086 CrossRef CAS.
  37. M. Li, Y. Lan, A. M. Ako, W. Wernsdorfer, C. E. Anson, G. Buth, A. K. Powell, Z. Wang and S. Gao, A family of 3d–4f octa-nuclear [MnIII4LnIII4] wheels (Ln = Sm, Gd, Tb, Dy, Ho, Er, and Y): synthesis, structure, and magnetism, Inorg. Chem., 2010, 49, 11587–11594 CrossRef CAS PubMed.
  38. K. R. Vignesh, S. K. Langley, B. Moubaraki, K. S. Murray and G. Rajaraman, Large hexadecametallic {MnIII–LnIII} wheels: Synthesis, structural, magnetic, and theoretical characterization, Chem. – Eur. J., 2015, 21, 16364–16369 CrossRef CAS.
  39. C. Benelli and D. Gatteschi, Introduction to molecular magnetism: From transition metals to lanthanides, John Wiley & Sons, 2015 Search PubMed.
  40. C. Kremer, Substitution on Re(IV) complexes: a tool for the synthesis of novel mono-and polynuclear compounds, Commun. Inorg. Synth., 2013, 1, 8–15 Search PubMed.
  41. T. D. Harris, M. V. Bennett, R. Clerac and J. R. Long, [ReCl4(CN)2]2−: A high magnetic anisotropy building unit giving rise to the single-chain magnets (DMF)4MReCl4(CN)2 (M = Mn, Fe, Co, Ni), J. Am. Chem. Soc., 2010, 132, 3980–3988 CrossRef CAS PubMed.
  42. X.-Y. Wang, C. Avendaño and K. R. Dunbar, Molecular magnetic materials based on 4d and 5d transition metals, Chem. Soc. Rev., 2011, 40, 3213–3238 RSC.
  43. S. K. Singh and G. Rajaraman, Deciphering the origin of giant magnetic anisotropy and fast quantum tunnelling in Rhenium(IV) single-molecule magnets, Nat. Commun., 2016, 7, 10669 CrossRef CAS.
  44. A. Upadhyay, J. Rajpurohit, M. Kumar Singh, R. Dubey, A. Kumar Srivastava, A. Kumar, G. Rajaraman and M. Shanmugam, Hydroxo–Bridged Dimers of Oxo–Centered Ruthenium(III) Triangle: Synthesis and Spectroscopic and Theoretical Investigations, Chem. – Eur. J., 2014, 20, 6061–6070 CrossRef CAS.
  45. S. K. Langley, D. P. Wielechowski, V. Vieru, N. F. Chilton, B. Moubaraki, L. F. Chibotaru and K. S. Murray, The first 4d/4f single-molecule magnet containing a {RuIII2DyIII2} core, Chem. Commun., 2015, 51, 2044–2047 RSC.
  46. M. R. Saber and K. R. Dunbar, Trigonal bipyramidal 5d–4f molecules with SMM behavior, Chem. Commun., 2014, 50, 2177–2179 RSC.
  47. A. Swain, R. Martin, K. R. Vignesh, G. Rajaraman, K. S. Murray and S. K. Langley, Enhancing the barrier height for magnetization reversal in 4d/4f RuIII2LnIII2 “butterfly” single molecule magnets (Ln = Gd, Dy) via targeted structural alterations, Dalton Trans., 2021, 50, 12265–12274 RSC.
  48. L. Chibotaru, L. Ungur and A. Soncini, The origin of nonmagnetic Kramers doublets in the ground state of dysprosium triangles: evidence for a toroidal magnetic moment, Angew. Chem., Int. Ed., 2008, 47, 4126–4129 CrossRef CAS.
  49. L. Ungur, W. Van den Heuvel and L. F. Chibotaru, Ab initio investigation of the non-collinear magnetic structure and the lowest magnetic excitations in dysprosium triangles, New J. Chem., 2009, 33, 1224–1230 RSC.
  50. L. F. Chibotaru, L. Ungur, C. Aronica, H. Elmoll, G. Pilet and D. Luneau, Structure, magnetism, and theoretical study of a mixed-valence CoII3CoIII4 heptanuclear wheel: lack of SMM behavior despite negative magnetic anisotropy, J. Am. Chem. Soc., 2008, 130, 12445–12455 CrossRef CAS.
  51. N. F. Chilton, R. P. Anderson, L. D. Turner, A. Soncini and K. S. Murray, PHI: A powerful new program for the analysis of anisotropic monomeric and exchange–coupled polynuclear d-and f-block complexes, J. Comput. Chem., 2013, 34, 1164–1175 CrossRef CAS PubMed.
  52. D. Reta and N. F. Chilton, Uncertainty estimates for magnetic relaxation times and magnetic relaxation parameters, Phys. Chem. Chem. Phys., 2019, 21, 23567–23575 RSC.
  53. B. Dutta, T. Guizouarn, A. Gharu, F. Pointillart, K. R. Vignesh and D. Ray, New Family of Dinuclear Ni–Ln Complexes from Ternary Coordination by β-Diketonates around the Lanthanoids: Synthesis, Structures, and Magnetic Behavior with Theoretical Supports, Cryst. Growth Des., 2024, 25, 143–157 CrossRef.
  54. N. Barman, P. Halder, S. Mukhopadhyay, B. Schwarz, E. Colacio, R. Rana, G. Rajaraman and J. Goura, Synthesis, structure, and magnetic properties of diamagnetic Co (iii) ion-based heterometallic CoIII–LnIII (Ln = Dy, Tb, Ho, Er) complexes, New J. Chem., 2024, 48, 15735–15746 RSC.
  55. S. K. Gupta, T. Rajeshkumar, G. Rajaraman and R. Murugavel, Is a strong axial crystal-field the only essential condition for a large magnetic anisotropy barrier? The case of non-Kramers Ho(III) versus Tb(III), Dalton Trans., 2018, 47, 357–366 RSC.
  56. S. K. Singh, M. F. Beg and G. Rajaraman, Role of Magnetic Exchange Interactions in the Magnetization Relaxation of {3d–4f} Single–Molecule Magnets: A Theoretical Perspective, Chem. – Eur. J., 2016, 22, 672–680 CrossRef CAS.
  57. K. S. Pedersen, G. Lorusso, J. J. Morales, T. Weyhermüller, S. Piligkos, S. K. Singh, D. Larsen, M. Schau-Magnussen, G. Rajaraman and M. Evangelisti, Fluoride–Bridged {GdIII3MIII2} (M = Cr, Fe, Ga) Molecular Magnetic Refrigerants, Angew. Chem., Int. Ed., 2014, 53, 2394–2397 CrossRef CAS.
  58. K. Wade, Ligand field theory and its applications, ed. B. N. Figgis and M. A. Hitchman, Wiley-VCH, New York, 2000, xviii+354, £ 51.95, ISBN 0.471-31776-4, 2000, pp. 449–450.
  59. M. E. Lines, Orbital angular momentum in the theory of paramagnetic clusters, J. Chem. Phys., 1971, 55, 2977–2984 CrossRef CAS.
  60. L. Noodleman, Valence bond description of antiferromagnetic coupling in transition metal dimers, J. Chem. Phys., 1981, 74, 5737–5743 CrossRef CAS.
  61. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, G. A. Petersson and H. Nakatsuji, Gaussian 16 Revision C. 01, 2016, Gaussian Inc., Wallingford, CT, 2016, vol. 1, pp. 572 Search PubMed.
  62. K. R. Vignesh, S. K. Langley, A. Swain, B. Moubaraki, M. Damjanović, W. Wernsdorfer, G. Rajaraman and K. S. Murray, Slow Magnetic Relaxation and Single–Molecule Toroidal Behaviour in a Family of Heptanuclear {CrIIILnIII6} (Ln = Tb, Ho, Er) Complexes, Angew. Chem., Int. Ed., 2018, 57, 779–784 CrossRef CAS.
  63. J. J. Le Roy, L. Ungur, I. Korobkov, L. F. Chibotaru and M. Murugesu, Coupling strategies to enhance single-molecule magnet properties of erbium–cyclooctatetraenyl complexes, J. Am. Chem. Soc., 2014, 136, 8003–8010 CrossRef CAS.
  64. O. Kahn, Molecular magnetism, Courier Dover Publications, 2021 Search PubMed.
  65. G. Rajaraman, F. Totti, A. Bencini, A. Caneschi, R. Sessoli and D. Gatteschi, Density functional studies on the exchange interaction of a dinuclear Gd(III)–Cu(II) complex: method assessment, magnetic coupling mechanism and magneto-structural correlations, Dalton Trans., 2009, 3153–3161 RSC.
  66. S. K. Singh, N. K. Tibrewal and G. Rajaraman, Density functional studies on dinuclear {NiIIGdIII} and trinuclear {NiIIGdIIINiII} complexes: magnetic exchange and magneto-structural maps, Dalton Trans., 2011, 40, 10897–10906 RSC.
  67. K. R. Vignesh, S. K. Langley, C. J. Gartshore, I. Borilović, C. M. Forsyth, G. Rajaraman and K. S. Murray, Rationalizing the sign and magnitude of the magnetic coupling and anisotropy in dinuclear manganese(III) complexes, Dalton Trans., 2018, 47, 11820–11833 RSC.
  68. K. R. Vignesh, S. K. Langley, K. S. Murray and G. Rajaraman, Exploring the influence of diamagnetic ions on the mechanism of magnetization relaxation in {CoIII2LnIII2} (Ln = Dy, Tb, Ho) “butterfly” complexes, Inorg. Chem., 2017, 56, 2518–2532 CrossRef CAS PubMed.
  69. A. Gharu and K. R. Vignesh, Theoretical exploration of single-molecule magnetic and single-molecule toroic behaviors in peroxide-bridged double-triangular {MII3LnIII3} (M = Ni, Cu and Zn; Ln = Gd, Tb and Dy) complexes, Dalton Trans., 2024, 53, 13394–13408 RSC.
  70. D. Schray, D. Westerbeck, J. Braun, Y. Lan, S. Gomez-Coca, W. Wernsdorfer, E. Ruiz, C. E. Anson, J. r. Schnack and A. K. Powell, Fe–Gd Ferromagnetic Cyclic Coordination Cluster [FeIII4GdIII4(teaH)8(N3)8(H2O)] with Magnetic Anisotropy – Theory and Experiment, Inorg. Chem., 2023, 62, 6642–6648 CrossRef CAS PubMed.
  71. N. C. Jana, Z. Jagličić, R. Herchel, P. Brandão, R. Nandy and A. Panja, Insight into the Magnetic Exchange Interactions and Anisotropy in Heterobimetallic CuII–LnIII Complexes: A Rare Example of Cu–Gd Single-Molecule Magnets, Cryst. Growth Des., 2024, 24, 7537–7548 CrossRef CAS.
  72. F. Aquilante, J. Autschbach, R. K. Carlson, L. F. Chibotaru, M. G. Delcey, L. De Vico, N. Ferré, L. M. Frutos, L. Gagliardi and M. Garavelli, Molcas 8: New capabilities for multiconfigurational quantum chemical calculations across the periodic table, J. Comput. Chem., 2016, 37, 506–541 CrossRef CAS.
  73. B. O. Roos and P.-Å. Malmqvist, Relativistic quantum chemistry: the multiconfigurational approach, Phys. Chem. Chem. Phys., 2004, 6, 2919–2927 RSC.
  74. 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.
  75. 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.
  76. 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.
  77. A. D. Becke, Density–functional thermochemistry. III. The role of exact exchange, J. Chem. Phys., 1993, 98, 5648–5652 CrossRef CAS.
  78. T. R. Cundari and W. J. Stevens, Effective core potential methods for the lanthanides, J. Chem. Phys., 1993, 98, 5555–5565 CrossRef CAS.
  79. (a) CCDC 2456392: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ng2gp; (b) CCDC 2456393: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ng2hq; (c) CCDC 2456394: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ng2jr.

This journal is © the Partner Organisations 2026
Click here to see how this site uses Cookies. View our privacy policy here.