DOI:
10.1039/D3DT01448G
(Paper)
Dalton Trans., 2023,
52, 8956-8963
A bis-calix[4]arene-supported [CuII16] cage†
Received
15th May 2023
, Accepted 8th June 2023
First published on 9th June 2023
Abstract
Reaction of 2,2′-bis-p-tBu-calix[4]arene (H8L) with Cu(NO3)2·3H2O and N-methyldiethanolamine (Me-deaH2) in a basic dmf/MeOH mixture affords [CuII16(L)2(Me-dea)4(μ4-NO3)2(μ-OH)4(dmf)3.5(MeOH)0.5(H2O)2](H6L)·16dmf·4H2O (4), following slow evaporation of the mother liquor. The central core of the metallic skeleton describes a tetracapped square prism, [Cu12], in which the four capping metal ions are the CuII ions housed in the calix[4]arene polyphenolic pockets. The [CuII8] square prism is held together “internally” by a combination of hydroxide and nitrate anions, with the N-methyldiethanolamine co-ligands forming dimeric [CuII2] units which edge-cap above and below the upper and lower square faces of the prism. Charge balance is maintained through the presence of one doubly deprotonated H6L2− ligand per [Cu16] cluster. Magnetic susceptibility measurements reveal the predominance of strong antiferromagnetic exchange interactions and an S = 1 ground state, while EPR is consistent with a large zero-field splitting.
Introduction
Interest in the magnetic behaviour of polymetallic complexes of CuII can be traced back to studies of copper acetate1 in the 1950s and the subsequent magneto-structural correlations developed for [Cu2] dimers bridged by hydroxides, halides and azides.2–6 Later studies examined larger, more complex structure types such as [Cu3] triangles,7 [Cu4] cubes,8 [Cu]n chains9 and heterometallic species containing transition metal10 or lanthanide metal centres.11 Many of these studies revealed fascinating magnetic behaviours, such as the spin frustration prevalent in equilateral triangles12 which represent the simplest molecular analogues of the kagome lattice,13 a topology also of relevance in multicopper oxidases.14
A search of the Cambridge Structural Database (CSD) reveals there are several hundred homometallic O/N-bridged CuII clusters deposited that possess a nuclearity of two or more, with the largest being [Cu44], reported in 2004.15 However, as nuclearity increases this number decreases significantly – for example there are less than twenty with a nuclearity of twenty or more. A seemingly ideal ligand framework for housing the Jahn–Teller (JT) distorted octahedral CuII ion is p-tBu-calix[4]arene (H4TBC[4]; Fig. 1A), its tetraphenolic pocket being perfectly suited to accommodating four short equatorial bonds. Despite this, and the success of H4TBC[4] in the coordination chemistry of the JT distorted octahedral MnIII ion,16 there are just three homometallic CuII compounds known. The first two, [CuII9(OH)3(TBC[4])3Cl2(dmso)5.5(EtOH)0.5][CuICl2] (1) and [CuII9(OH)3(TBC[4])3(NO3)2(dmso)6](NO3) (2) (Fig. 2A), are isostructural and were both reported in the same paper.17 Their metallic skeletons describe tricapped trigonal prisms in which [Cu-TBC[4]]2− moieties act as metalloligands that encapsulate an hydroxide-bridged [Cu6] trigonal prism.
 |
| Fig. 1 Single crystal X-ray structures of H4TBC[4] (A) and H8L (B). Colour code C = grey, O = red, H = white. | |
 |
| Fig. 2 Partial single crystal X-ray structures of 1 (A) and 3 (B). The disordered central Cu ion in 3 has been omitted. Colour code: C = grey, O = red, N = dark blue, S = yellow, CuII = light blue. H atoms and co-crystallised solvent/anions omitted for clarity. | |
Extension of this chemistry to examine the coordination behaviour of the related ligand 2,2′-bis-p-tBu-calix[4]arene (H8L),18 in which two H4TBC[4]s are linked via a methylene bridge (Fig. 1B), affords the larger, but related [CuII13(L)2(NO3)(μ-OH)8(dmf)7](OH) (3) whose metallic skeleton describes a centred, tetracapped square prism (Fig. 2B).19 The expansion from [Cu9] to [Cu13] is a result of the conformational flexibility (ring inversion) of H8L which, upon metal coordination, provides an organic skeleton with eight proximal phenolic O-atoms. In order to explore this reaction space further we have adapted the synthesis of complex 3 to include a flexible bridging co-ligand, N-methyldiethanolamine (Me-deaH2), which enables the formation of the larger complex [CuII16(L)2(Me-dea)4(μ4-NO3)2(μ-OH)4(dmf)3.5(MeOH)0.5(H2O)2](H6L)·16dmf·4H2O (4) whose metallic skeleton is also based on a multiply-capped square prism.
Experimental
Synthesis
Cu(NO3)2·3H2O was purchased from commercial suppliers and used without further purification. 2,2′-Bis-p-tBu-calix[4]arene (H8L) was prepared as previously described.18
[CuII16(L)2(Me-dea)4(μ4-NO3)2(μ-OH)4(dmf)3.5(MeOH)0.5(H2O)2](H6L)·16dmf·4H2O (4).
Cu(NO3)2·3H2O (1 mmol, 250 mg), H8L (0.154 mmol, 200 mg) and Me-deaH2 (0.174 mmol, 20 μL) were dissolved in a 1
:
1 dmf/MeOH mixture (24 mL) and stirred for 10 minutes. An excess of NEt3 (2.87 mmol, 400 μL) was added and the resulting brown solution was stirred for 2 hours and then filtered. Slow evaporation of the mother liquor afforded dark brown crystals suitable for X-ray diffraction. Elemental analysis calculated for (4), C343H510.5Cu16N25.5O68: C, 58.06%; H, 7.25%: N, 5.03%. Found C, 57.82%; H, 6.93; N, 4.69%.
X-ray crystallography
Diffraction data for 4 were collected on a Bruker D8 diffractometer equipped with a PHOTON 100 detector and operating with synchrotron radiation (0.7749 Å). Crystal data for 4 (CCDC 2101166†): C343H510.5Cu16N25.5O68, M = 7095.88 g mol−1, triclinic, space group P
(no. 2), a = 21.4369(9) Å, b = 21.9078(9) Å, c = 22.0660(9) Å, α = 105.596(3)°, β = 111.468(2)°, γ = 95.527(3)°, V = 9067.1(7) Å3, Z = 1, T = 100(2) K, μ(synchrotron) = 1.251 mm−1, Dcalc. = 1.300 g cm−3, 56
519 reflections measured (4.312° ≤ 2Θ ≤ 47.784°), 21
580 unique (Rint = 0.0705, Rsigma = 0.0826) which were used in all calculations. The final R1 was 0.0566 (I > 2σ(I)) and wR2 was 0.1551 (all data).
Magnetic data
Magnetic susceptibility and magnetisation data were collected on a powdered microcrystalline sample of 4 using a Quantum Design PPMS Dynacool. Susceptibility data were collected in the T = 2–300 K range under and applied magnetic field, B = 0.1 T. Magnetisation data were collected in the T = 2–10 K and B = 0.5–9.0 T ranges. A unit cell check was performed prior to measurement.
EPR spectroscopy
EPR spectra were measured on powder and solution (CH2Cl2/toluene) samples of 4, at 10 and 20 K and over the 0–1.8 T field range, on a Bruker EMXPlus spectrometer.
Computational details
Density functional theory in conjunction with the broken symmetry approach20 has been used to estimate the magnetic exchange interactions in complex 4 employing Gaussian 09.21 The hybrid B3LYP functional22 together with Ahlrichs TZV basis set for was used for the Cu atoms, the SVP basis set for O and N atoms, and the SV basis set for C and H atoms.23 This methodology has been shown to reproduce experimental magnetic exchange interactions accurately.24
Results and discussion
Reaction of H8L with Cu(NO3)2·3H2O and Me-deaH2 in a basic dmf/MeOH mixture affords single crystals of formula [CuII16(L)2(Me-dea)4(μ4-NO3)2(μ-OH)4(dmf)3.5(MeOH)0.5(H2O)2](H6L)·16dmf·4H2O (4, Fig. 3 and 4), following slow evaporation of the mother liquor. The crystals were found to be in a triclinic cell and structure solution was carried out in the space group P
. The asymmetric unit (ASU) comprises half of the formula.
 |
| Fig. 3 Orthogonal views of the molecular structure of the cation of 4 shown ‘face-on’ (A) and ‘side-on’ (B). Colour code: C = grey, O = red, N = dark blue, CuII = light blue. H atoms and co-crystallised solvent/anions omitted for clarity. | |
 |
| Fig. 4 Cluster core (A) and metallic skeleton (B) of 4. Colour code: O = red, N = dark blue, CuII = light blue. The eight Cu ions in the square prism are Cu3–6 and symmetry equivalents. The four face-capping Cu ions are Cu1–2. The four edge-capping Cu ions are Cu7–8. | |
The metallic skeleton describes a tetracapped (Cu1–2 and symmetry equivalent) square prism (Cu3–6) in which the ‘upper’ and ‘lower’ square faces (as drawn in Fig. 4B) are edge-capped by a {Cu2} (Cu7–8) moiety. The Cu ions in the upper and lower faces of the square prism are connected to each other by four μ-OH ions to form a [Cu8(OH)4]12+ central unit (Fig. S1†). The four face-capping Cu ions are housed in the polyphenolic pockets of the two fully deprotonated bis-calixarene ligands whose O-atoms further bridge to the Cu ions in the square prism (Cu3–Cu6; Fig. S1 and S2†). The L8− ligands thus wrap around the four square faces of the prism in the ‘equatorial’ plane (as drawn in Fig. 1A), completely encapsulating the [Cu8(OH)4]12+ central unit. The ‘upper’ and ‘lower’ faces of the prism are connected to the {Cu2} edge-cap by a μ4-bridging NO3− ion which connects Cu3–4 to Cu7–8. The latter are further bridged to each other and the Cu ions on the vertices of the square prism by two Me-dea2− ligands that bond in two different ways. The first μ3-bridges, using one O-arm (O16) to link between Cu7–Cu8 and the other (O17) to link between Cu8 and Cu6 in the square prism. The second chelates to Cu7 with one O-arm terminally coordinated (O15) and one arm (O14) bridging to Cu4 in the square prism.
The Cu ions (Cu1–2) that sit in the tetraphenolic pocket of L8− are square planar {CuO4} with a fifth, longer contact to a disordered molecule of dmf that fills the calixarene cavity. The Cu ions in the square prism (Cu3–6) are all five coordinate and in distorted square pyramidal geometry, the remaining sites on Cu3 and Cu5 being occupied by dmf/H2O molecules. The Cu ions in the edge-capping {Cu2} moiety (Cu7, Cu8) bonded to the Me-dea2− ligands are five-coordinate, square pyramidal {CuO4N}, and four-coordinate, square planar {CuO3N}, respectively.
The complex displays numerous intra- and inter-molecular H-bonding interactions (Fig. S3–S6†). The water molecule bonded to Cu5 is H-bonded to two O-atoms of the nitrate anion (O19⋯O12/O11, ∼2.83–3.01 Å). O12 is also H-bonded to a μ-OH ion (O12⋯O9, ∼2.95 Å), and the third O-atom in the nitrate anion (O13) is H-bonded to a dmf molecule bonded to Cu3 (O13⋯O18, ∼3.00 Å). There is also a complex H-bonding network in the region of one of the two Me-dea2− co-ligands. The O-atom (O15) belonging to the Me-dea2− ligand supporting Cu7 is H-bonded to a H2O molecule of crystallisation (O15⋯O21, ∼2.62 Å). The latter is also interacting with the lower-rim oxygen atom of the co-crystallised H6L2− ligand (O21⋯O26, ∼2.72 Å). Upon symmetry expansion, it is possible to appreciate the importance of this network of H-bonds as it dictates the way compound 4 and the co-crystallised molecules pack, through the formation of H-bonded chains (Fig. S3–S6†). The [Cu16] cluster is surrounded at its four ‘corners’ by co-crystallised H6L2−. This doubly deprotonated molecule affords overall charge balance, and although there are waters of crystallisation present (and therefore potentially OH− counter anions), we favour the presence of this dianion for two reasons. (1) The inter-molecular interactions discussed above; (2) refinement suggests each TBC[4] moiety in H6L2− is singly deprotonated, consistent with the highly acidic nature of H4TBC[4] hydroxyl groups reported in literature.25 The clusters are well isolated from each other, with the closest Cu⋯Cu distance being >12 Å between the Cu8 ions of distinct molecules. A search of the CSD shows that there are a total of thirteen [CuII16] clusters known, none of which have a topology similar to 4.26
A comparison of the metallic skeletons of [Cu9], [Cu13] and [Cu16] (Fig. 5) shows some striking similarities. [Cu9] is a tricapped trigonal prism in which the square faces are all capped by a Cu ion housed in the pocket of a TBC[4] moiety with halide/nitrate anions capping the upper/lower triangular faces. [Cu13] is a centred, tetracapped square prism in which the ‘equatorial’ square faces are all capped by a Cu ion housed in the pocket of a TBC[4] moiety with nitrate anions capping the upper/lower triangular faces. [Cu16] has a structure analogous to [Cu13] but without the central Cu ion, and with two additional {Cu2} edge-capping moieties introduced through the presence of a co-ligand. These structural similarities reflect the dominant, structure-directing influence of the calix[4]arene ligands.
 |
| Fig. 5 A comparison of the metallic skeletons of [Cu9] (1–2, A and B), [Cu13] (3, C and D) and [Cu16] (4, E and F). The central, distorted Cu ion in [Cu13] has been removed to aid comparison. [Cu9] describes a tricapped trigonal prism, [Cu13] a (centred) tetracapped square prism, and [Cu16] a tetracapped square prism containing two additional {Cu2} edge-caps. The Cu ions in the prisms are in light blue, those housed in the tetraphenolic calixarene pocket in green and the edge-caps in pink. | |
Magnetic measurements
Direct current magnetic susceptibility (χ) studies were performed on a polycrystalline sample of 4 over the temperature range T = 2–300 K, in an applied magnetic field B = 0.1 T (Fig. 6), where χ = M/B and M is the magnetisation. At 300 K, the χT value of 5.9 cm3 mol−1 K is below the expected value for spin-only contributions to the susceptibility for a [CuII16] unit (7.26 cm3 mol−1 K, g = 2.2; 6.00 cm3 mol−1 K, g = 2.0). Upon cooling, the χT product decreases rapidly until approximately T = 10 K where it reaches 1.1 cm3 mol−1 K, wherefrom it increases upon further cooling reaching a value of 1.5 cm3 mol−1 K at 2 K. The data are therefore suggestive of the presence of competing exchange interactions, in agreement with the range of Cu–O–Cu angles present, and previous magneto-structural correlations developed for smaller O-bridged CuII clusters.1–6
 |
| Fig. 6 Experimental χT versus T data for 4 measured in the T = 2–300 K temperature range in an applied field, B = 0.1 T. | |
The quantitative interpretation of the temperature dependence of the χT product of 4 was performed by numerically fitting the experimental data to the full matrix representation of spin-Hamiltonian (1), of dimension 65
536 by 65
356, by use of the Levenberg–Marquardt algorithm.27
|  | (1) |
where the indices refer to the constituent Cu
II ions,
μB is the Bohr magneton,
g is the
g-factor,
Ŝ is a spin operator and
Jij is the bilinear pairwise isotropic exchange interaction parameter. Based on the structure of
4, four different isotropic exchange parameters could be considered to fit the experimental temperature dependence of the
χT product (
Fig. 6, insets).
We employed irreducible tensor operator algebra to block-diagonalise the matrix representation of spin-Hamiltonian (1),28 fixing the g factor for all CuII centres to g = 2, affording only the four different isotropic exchange parameters as free fit parameters. This resulted in very good agreement with the experimental data with the best-fit parameters J1, J2 and J3 being strongly antiferromagnetic (AF), ca. −70 to −100 cm−1, and J4 being ferromagnetic (F), ca. +20 cm−1. However, all these exchange parameters are strongly correlated, with only the correlation coefficients between J4 and the others being less than 95%. Use of a single exchange parameter fails to satisfactorily reproduce the experimental data at all temperature regimes. Use of a model containing two different exchange parameters (J1–3 = J, J4 = J′; Fig. 6, insets) results in a very good agreement with the experimental data (Fig. 6), with J = −122 ± 12 cm−1 and J′ = 22 ± 8 cm−1, with a correlation coefficient of 87%. With these best-fit parameters the ground spin-state of 4 is an S = 1 spin-state, with excited S = 0, 0 and 2 states lying 23, 35 and 46 cm−1 above the ground state, respectively, followed by a quasi-continuum of states (Fig. 7). We note that the exchange constant in the [Cu12] prism of 4 is similar to the values found in [Cu13] (3), which are all AF and in the range −66 ≤ J ≤ −84 cm−1.19 Employing the same isotropic model to fit the VTVB magnetisation data of 4 was not possible. This is because the VTVB magnetisation of 4 presents significant nesting (Fig. S7†) when plotted against B/T, which indicates anisotropy splittings of the same order of magnitude as the experimental conditions (T = 2 to 10 K and B = 1 to 9 T). Given that individual CuII centres are devoid of anisotropy (bar the g-factor), the observed anisotropy splittings can only originate (within the spin-Hamiltonian formalism) from magnetic exchange.29 The experimental data do not allow for the determination of these exchange contributions to anisotropy.
 |
| Fig. 7 Energy spectrum of 4 calculated from the fit of the susceptibility data, as described in the text. The ground spin-state of 4 is an S = 1 spin-state, with excited S = 0, 0 and 2 states lying 23, 35 and 46 cm−1 above the ground state. | |
Given the relatively well-isolated S = 1 ground state evidenced by the magnetic data and modelling, we attempted to probe this state by EPR spectroscopy at low temperatures. Spectra measured at 10 K and over the 0–1.8 T magnetic field range, at both X- and Q-band frequencies, only gave weak signals that could be attributed to small amounts of monomeric CuII species. This indicates that the zero-field splitting (ZFS) of the S = 1 state is significantly larger than the microwave frequencies applied. In order to estimate the ZFS, attempts were made to fit the VTVB magnetisation data to an isolated S = 1 model: while these fits were not entirely satisfactory (Fig. S7†) they indicate an axial ZFS parameters (D) of several cm−1. This would be consistent with 4 being EPR silent under the measurement conditions. Given the lack of single-ion ZFS for CuII, the origin of such a ZFS must lie in anisotropic components of the exchange interaction.29 Although a ZFS in the region of several cm−1 would be unusual for a Cu cluster, the project coefficients of anisotropic exchange are larger for lower total spins,28 which may provide an explanation given that the ground state of 4 is the lowest possible total spin. It is more usual for Cu clusters to either give rise to a diamagnetic ground state or a high spin state, so there are few data to compare this against.
Theoretical studies
In order to analyse the sign and magnitude of the exchange interactions in more detail we have constructed three pentametallic models (4A, 4B, 4C) and one bimetallic model (4D) based on the X-ray structure of 4 to determine twelve exchange interactions (Fig. S8†). These are given in Table S2† and consist of ten antiferromagnetic interactions in the range −3.9 < J < −55.7 cm−1, and two weakly ferromagnetic interactions in the range +2.6 < J < +6.9 cm−1. The strongest antiferromagnetic exchange occurs between Cu2 (face-cap) and Cu4 (vertex) mediated by a single μ-OPh bridge with a Cu–O–Cu angle of ∼131°. The second strongest antiferromagnetic exchange interaction is between Cu3 (vertex) and Cu4 (vertex) mediated via a single μ-OH bridge with a Cu–O–Cu angle of ∼126°. The two ferromagnetic exchange interactions occur between Cu1 (face cap) and Cu3 (vertex) mediated by μ-OH/μ-OPh bridges with an average Cu–O–Co angle of ∼96° and a Cu–O–Cu–O dihedral angle of ∼24°, and between Cu6 (vertex) and Cu8 (edge-cap) mediated by μ-O(NO3)/OR groups with an average Cu–O–Co angle of ∼98° and a Cu–O–Cu–O dihedral angle of ∼17°. A detailed examination of the structure reveals that the sign and magnitude of J strongly depends on both the Cu–O–Cu angle and the Cu–O–Cu–O dihedral angle.30 Large Cu–O–Cu angles lead to strong antiferromagnetic exchange whose magnitude decreases with decreasing angle (Fig. S9†). The switch to ferromagnetic occurs at a Cu–O–Cu angle of ∼97–99°.1–6 A large Cu–O–Cu–O dihedral angle in conjunction with a small Cu–O–Cu angle results in accidental orbital orthogonality (a counter complementarity effect arising from the presence of two different bridging ligands) which leads to a weak ferromagnetic interaction. DFT calculated spin density analysis (Fig. S10–S13†) suggests strong spin delocalisation with spin densities on the CuII ions of between 0.391–0.673. The bridging hydroxides/phenolic O-atoms have the largest spin density among the coordinating atoms, consistent with the strongest antiferromagnetic exchange through these moieties. The pattern of calculated exchange interactions (Table S2†) and the spin density analysis therefore strongly suggests the presence of spin frustration between the face-capping Cu ions housed in the calixarene polyphenolic pockets and the Cu ions at the vertices of the square prism. The small spin density of the N-atom of the nitrate points to a small, near-negligible, exchange through the Cu–O–N–O–Cu pathway (Fig. S14†).
Conclusions
The reaction between Cu(NO3)2·3H2O and 2,2′-bis-p-tBu-calix[4]arene (H8L) in a basic dmf/MeOH mixture and in the presence of the co-ligand Me-deaH2 results in the formation of the complex [CuII16(L)2(Me-dea)4(NO3)2(OH)4(dmf)3(MeOH)(H2O)2](H6L)·16dmf·4H2O. Its metallic skeleton describes a tetra-capped square prism, [Cu12], with two additional edge-capping [Cu2] units introduced by the co-ligand sitting above and below the “upper” and “lower” square faces. The cage structure is held together internally via a combination of bridging hydroxides and nitrates, and externally by the two bis-p-tBu-calix[4]arenes. The structural similarity of [Cu16] to [Cu13], a centred, tetra-capped square prism also built with H8L (in the absence of co-ligands), and [Cu9], a tri-capped trigonal prism built with H4TBC[4], reflects the dominant structure-directing role played by the calix[n]arene ligands. Magnetic susceptibility measurements reveal strong antiferromagnetic exchange interactions between neighbouring CuII ions within the central [Cu12] tetra-capped square prism and between the CuII ions in the prism and the caps (J = −122 ± 12 cm−1), and strong ferromagnetic exchange between the CuII ions in the caps (J′ = +22 ± 8 cm−1), leading to a ground state of S = 1. DFT calculations reveal a strong dependence of J on both the Cu–O–Cu and O–Cu–O–Cu angles, and suggest significant spin frustration in the central [Cu12] tetra-capped square prism. Magnetisation and EPR data suggest the presence of significant ZFS in the S = 1 ground state, originating from anisotropic components of the exchange interaction. It will be interesting to examine the effect of extending the bridge length between the H4TBC[4] moieties in bis-calix[4]arene ligands and/or of increasing the size of the calix[n]arene macrocycle, e.g. to p-tBu-calix[8]arene, on cage nuclearity and topology. Expansion into Cu-4f chemistry also promises the discovery of some interesting cages. This work is currently in progress.
Author contributions
LRBW, MC performed the synthetic chemistry and measured the magnetic data. LRBW and SP fitted the magnetic data. SJT/SJD collected and solved the XRD data. AB, MS and EJLM collected and analysed the EPR data. MKS performed theoretical calculations. SJD and EKB conceived the idea. All authors wrote/edited the manuscript.
Conflicts of interest
There are no conflicts to declare.
Acknowledgements
The authors thank the EPSRC for funding a studentship (LRBW) and grant EP/N01331X/1 (MC), and for funding of the EPSRC EPR National Research Facility (EP/W014521/1 and EP/V035231/1). This research used resources of the Advanced Light Source, which is a DOE Office of Science User Facility under contract no. DE-AC02-05CH11231. For the purpose of open access, the author has applied a Creative Commons Attribution (CC BY) license to any Author Accepted Manuscript version arising from this submission.
Notes and references
- B. Bleaney and K. D. Bowers, Proc. R. Soc. London, Ser. A, 1952, A214, 451 Search PubMed.
- W. H. Crawford, H. W. Richardson, J. R. Wasson, D. J. Hodgson and W. E. Hatfield, Inorg. Chem., 1976, 15, 2107 CrossRef.
- W. E. Hatfield, Comments Inorg. Chem., 1981, 1, 105 CrossRef CAS.
- W. E. Marsh, K. C. Patel, W. E. Hatfield and D. J. Hodgson, Inorg. Chem., 1983, 22, 511 CrossRef CAS.
- C. P. Landee and R. E. Greeney, Inorg. Chem., 1986, 25, 3371 CrossRef.
- S. S. Tandon, L. K. Thompson, M. E. Manuel and J. N. Bridson, Inorg. Chem., 1994, 33, 5555 CrossRef CAS.
- R. Boča, L. Dlhan, G. Mezei, T. Ortiz-Perez, R. G. Raptis and J. Telser, Inorg. Chem., 2003, 42, 5801 CrossRef PubMed.
- J. Tercero, E. Ruiz, S. Alvarez, A. Rodriguez-Fortea and P. Alemany, J. Mater. Chem., 2006, 16, 2729 RSC.
- E. Colacio, J. M. Dominguez-Vera, J.-P. Costes, R. Kivekas, J.-P. Laurent, J. Ruiz and M. Sundberg, Inorg. Chem., 1992, 31, 774 CrossRef CAS.
- O. Kahn, J. Galy, Y. Journaux, J. Jaud and I. Morgenstern-Badarau, J. Am. Chem. Soc., 1982, 104, 2165 CrossRef CAS.
- G. Rajaraman, F. Totti, A. Bencini, A. Caneschi, R. Sessoli and D. Gatteschi, Dalton Trans., 2009, 3153 RSC.
- A. Escuer, G. Vlahopoulou, F. Lloret and F. A. Mautner, Eur. J. Inorg. Chem., 2014, 83 CrossRef CAS.
- M. Kohno, O. A. Starykh and L. Balents, Nat. Phys., 2007, 3, 790 Search PubMed.
- I. Bento, L. O. Martins, G. G. Lopes, M. A. Carrondo and P. F. Lindley, Dalton Trans., 2005, 3507 RSC.
- M. Murugesu, R. Clérac, C. E. Anson and A. K. Powell, Inorg. Chem., 2004, 43, 7269 CrossRef CAS PubMed.
-
(a) S. M. Taylor, R. D. McIntosh, S. Piligkos, S. J. Dalgarno and E. K. Brechin, Chem. Commun., 2012, 48, 11190 RSC;
(b) S. M. Taylor, R. D. McIntosh, C. M. Beavers, S. J. Teat, S. Piligkos, S. J. Dalgarno and E. K. Brechin, Chem. Commun., 2011, 47, 1440 RSC;
(c) M. Coletta, S. Sanz, E. K. Brechin and S. J. Dalgarno, Dalton Trans., 2020, 49, 9882 RSC;
(d) M. Coletta, S. Sanz, L. J. McCormick, S. J. Teat, E. K. Brechin and S. J. Dalgarno, Dalton Trans., 2017, 46, 16807 RSC;
(e) G. Karotsis, S. J. Teat, W. Wernsdorfer, S. Piligkos, S. J. Dalgarno and E. K. Brechin, Angew. Chem., Int. Ed., 2009, 48, 8285 CrossRef CAS PubMed;
(f) S. M. Taylor, G. Karotsis, R. D. McIntosh, S. Kennedy, S. J. Teat, C. M. Beavers, W. Wernsdorfer, S. Piligkos, S. J. Dalgarno and E. K. Brechin, Chem. – Eur. J., 2011, 17, 7521 CrossRef CAS PubMed;
(g) M. A. Palacios, R. McLellan, C. M. Beavers, S. J. Teat, H. Weihe, S. Piligkos, S. J. Dalgarno and E. K. Brechin, Chem. – Eur. J., 2015, 21, 11212 CrossRef CAS PubMed;
(h) L. R. B. Wilson, M. Coletta, M. Evangelisiti, S. Piligkos, S. J. Dalgarno and E. K. Brechin, Dalton Trans., 2022, 51, 4213 RSC.
- G. Karotsis, S. Kennedy, S. J. Dalgarno and E. K. Brechin, Chem. Commun., 2010, 46, 3884 RSC.
- L. T. Carroll, P. Aru Hill, C. Q. Ngo, K. P. Klatt and J. L. Fantini, Tetrahedron, 2013, 69, 5002 CrossRef CAS.
- R. McLellan, M. A. Palacios, C. M. Beavers, S. J. Teat, S. Piligkos, E. K. Brechin and S. J. Dalgarno, Chem. – Eur. J., 2015, 21, 2804 CrossRef CAS PubMed.
-
(a) L. Noodleman and D. A. Case, Adv. Inorg. Chem., 1992, 38, 423 CrossRef CAS;
(b) L. Noodleman and E. R. Davidson, Chem. Phys., 1986, 109, 131 CrossRef;
(c) L. Noodleman, J. Chem. Phys., 1981, 74, 5737 CrossRef CAS;
(d) L. Noodleman and J. G. Norman, J. Chem. Phys., 1979, 70, 4903 CrossRef CAS.
-
M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A. Montgomery, J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. Austin, J. R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, Ö. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski and D. J. Fox, Gaussian 09, Revision E.01, Wallingford CT, 2013 Search PubMed.
-
(a) C. Lee, W. Yang and R. G. Parr, Phys. Rev. B: Condens. Matter, 1988, 37, 785 CrossRef CAS PubMed;
(b) A. D. Becke, J. Chem. Phys., 1993, 98, 5648 CrossRef CAS;
(c) A. D. Becke, J. Chem. Phys., 1993, 98, 1372 CrossRef CAS;
(d) P. J. Stephens, F. J. Devlin, C. F. Chabalowski and M. J. Frisch, J. Phys. Chem., 1994, 98, 11623 CrossRef CAS.
-
(a) A. Schäfer, H. Horn and R. Ahlrichs, J. Chem. Phys., 1992, 97, 2571 CrossRef;
(b) A. Schäfer, C. Huber and R. Ahlrichs, J. Chem. Phys., 1994, 100, 5829 CrossRef;
(c) G. E. Scuseria and H. F. Schaefer, III, J. Chem. Phys., 1989, 90, 3700 CrossRef CAS.
-
(a) E. K. Brechin, D. J. Cutler, M. Coletta, M. Singh, A. B. Tsanai, L. J. McCormick McPherson, S. Coles and J. Schnack, Dalton Trans., 2022, 51, 8945 RSC;
(b) M. Coletta, T. G. Tziotzi, M. Gray, G. S. Nichol, M. K. Singh, C. J. Milios and E. K. Brechin, Chem. Commun., 2021, 57, 4122 RSC;
(c) D. J. Cutler, M. K. Singh, G. S. Nichol, M. Evangelisti, J. Schnack, L. Cronin and E. K. Brechin, Chem. Commun., 2021, 57, 8925 RSC;
(d) S. S. Woodhouse, T. N. Dais, E. H. Payne, M. K. Singh, E. K. Brechin and P. G. Plieger, Dalton Trans., 2021, 50, 5318 RSC;
(e) M. K. Singh and G. Rajaraman, Inorg. Chem., 2019, 58, 3175 CrossRef CAS PubMed;
(f) C. McDonald, S. Sanz, E. K. Brechin, M. K. Singh, G. Rajaraman, D. Gaynor and L. F. Jones, RSC Adv., 2014, 4, 38182 RSC;
(g) S. Hazra, S. Bhattacharya, M. K. Singh, L. Carrella, E. Rentschler, T. Weyhermueller, G. Rajaraman and S. Mohanta, Inorg. Chem., 2013, 52, 12881 CrossRef CAS PubMed;
(h) A. E. Dearle, D. J. Cutler, H. W. L. Fraser, S. Sanz, E. Lee, S. Dey, I. F. Diaz-Ortega, G. S. Nichol, H. Nojiri, M. Evangelisti, G. Rajaraman, J. Schnack, L. Cronin and E. K. Brechin, Angew. Chem., Int. Ed., 2019, 58, 16903 CrossRef CAS PubMed;
(i) K. R. Vignesh, S. K. Langley, K. S. Murray and G. Rajaraman, Chem. – Eur. J., 2015, 21, 2881 CrossRef CAS PubMed;
(j) P. A. Tsami, T. G. Tziotzi, A. B. Canaj, M. K. Singh, S. J. Dalgarno, E. K. Brechin and C. J. Milios, Dalton Trans., 2022, 51, 15128 RSC.
- S. Shinkai, K. Araki, P. D. J. Grootenhuis and D. N. Reinhoudt, J. Chem. Soc., Perkin Trans., 1991, 2, 1883 RSC.
- L. N. Dawe and L. K. Thompson, Angew. Chem., Int. Ed., 2007, 46, 7440 CrossRef CAS PubMed; Y. Li, D. Li, J. Xu, H. Hao, S. Wang, J. Dou, T. Hu and X. Bu, Dalton Trans., 2014, 43, 15708 RSC; A. Adhikary, S. Goswami, J. A. Sheikh and S. Konar, Eur. J. Inorg. Chem., 2014, 963 CrossRef; T. Liu, T. C. Stamatatos, K. A. Abboud and G. Christou, Dalton Trans., 2010, 39, 3554 RSC; V. Chandrasekhar and L. Nagarajan, Dalton Trans., 2009, 34, 6712 RSC; G. A. Craig, M. Schutze, D. Aguila, O. Roubeau, J. Ribas-Arino, S. Vela, S. J. Teat and G. Aromí, Inorg. Chem., 2014, 53(7), 3290 CrossRef PubMed; A. K. Kostopoulos, A. D. Katsenis, J. M. Frost, V. G. Kessler, E. K. Brechin and G. S. Papaefstathiou, Chem. Commun., 2014, 50, 15002 RSC; T. Zhao, X. Jing, J. Wang, D. Wang, G. Li, Q. Huo and Y. Liu, Cryst. Growth Des., 2012, 12, 5456 CrossRef; Y. Bai, V. Tangoulis, R. Huang, L. Zheng and J. Tao, Chem. – Eur. J., 2009, 15, 2377 CrossRef PubMed; X. Zhang, T. Han, Q. Zou, R. Zhu, J. Wang and Z. Li, Z. Anorg. Allg. Chem., 2016, 642, 4 CrossRef; P. Klüfers and J. Schumacher, Angew. Chem., Int. Ed. Engl., 1995, 34, 2119 CrossRef.
-
W. H. Press, S. A. Teukolsky, W. T. Vetterling and B. P. Flannery, Numerical recipes in C: The Art of Scientific Computing, Cambridge University Press, Cambridge, 2nd edn, 1992 Search PubMed.
-
A. Bencini and D. Gatteschi, Electron Paramagnetic Resonance of Exchange Coupled Systems, Springer, Heidelberg, 1990 Search PubMed.
-
(a) I. Dzyaloshinskii, J. Phys. Chem. Solids, 1958, 4, 241 CrossRef;
(b) T. Moriya, Phys. Rev., 1960, 120, 91 CrossRef CAS;
(c) S. Ferrer, F. Lloret, E. Pardo, J. M. Clemente-Juan, M. Liu-González and S. García-Granda, Inorg. Chem., 2012, 51, 985 CrossRef CAS PubMed;
(d) M.-A. Bouammalim, N. Suaud, N. Guihéry and R. Maurice, Inorg. Chem., 2022, 61, 12138 CrossRef PubMed;
(e) J. Yoon and E. Solomon, Coord. Chem. Rev., 2007, 251, 379 CrossRef CAS;
(f) E. J. L. McInnes and D. Collison, eMagRes, 2016, 5, 1445 CAS.
- E. Ruiz, P. Alemany, S. Alvarez and J. Cano, J. Am. Chem. Soc., 1997, 119, 1297 CrossRef CAS.
Footnote |
† Electronic supplementary information (ESI) available: Additional figures and further computational details. CCDC 2101166. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d3dt01448g |
|
This journal is © The Royal Society of Chemistry 2023 |
Click here to see how this site uses Cookies. View our privacy policy here.