Next Article in Journal
Effect of an External Magnetic Field on the Hydrogen Reduction of Magnetite Nanoparticles in a Polymer Matrix
Next Article in Special Issue
A Chain of Vertex-Sharing {CoIII2CoII2}n Squares with Single-Ion Magnet Behavior
Previous Article in Journal
Unusual Compositions of Fe-Nb Alloy Precipitates in Iron-Implanted LiNbO3
Previous Article in Special Issue
Diverse Magnetic Properties of Two New Binuclear Complexes Affected by [FeN6] Octahedral Distortion: Two-Step Spin Crossover versus Antiferromagnetic Interactions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Cu12 Metallacycle Assembled from Four C3-Symmetric Spin Frustrated Triangular Units

1
State Key Laboratory of Rare Earth Resource Utilization, Changchun Institute of Applied Chemistry, Chinese Academy of Sciences, Changchun 130022, China
2
School of Applied Chemistry and Engineering, University of Science and Technology of China, Hefei 230026, China
3
Univ Rennes, CNRS, ISCR (Institut des Sciences Chimiques de Rennes)—UMR 6226, F-35000 Rennes, France
4
Key Laboratory of Advanced Energy Materials Chemistry (Ministry of Education), College of Chemistry, Nankai University, Tianjin 300071, China
5
Institut für Anorganische und Analytische Chemie, Friedrich-Schiller-Universität Jena, D-07743 Jena, Germany
*
Authors to whom correspondence should be addressed.
Magnetochemistry 2023, 9(5), 122; https://doi.org/10.3390/magnetochemistry9050122
Submission received: 10 April 2023 / Revised: 26 April 2023 / Accepted: 3 May 2023 / Published: 6 May 2023
(This article belongs to the Special Issue Advances in Molecular Magnetism)

Abstract

:
Assembling metallacycles with interesting topological arrangements is a critical task for chemists. We report here a novel dodecanuclear CuII compound, [{Cu3L(µ-N3)}4(Py)14]·2Py (Cu12) (where Py = pyridine and [H6L]Cl = tris(2-hydroxybenzylidine)triaminoguanidinium chloride, respectively), with the topology of a cycle accomplished by four two-connecting approximately flat C3-symmetric guanidine-based ligands. Each ligand affords three tridentate metal-binding cavities and the four node-to-node connections through single azido bridges are provided by pairs of metal centers. A theoretical investigation using CASSCF in addition to DFT calculations showed strong antiferromagnetic coupling within the Cu3-triangles, resulting in spin-frustrated systems. However, these calculations were not able to properly reproduce the very weak antiferromagnetic couplings between the triangle units, highlighting the challenge of describing the magnetic behavior of this compound.

Graphical Abstract

1. Introduction

Coordination chemistry provides an opportunity for construction by self-assembling an apparently limitless variety of structures [1,2]. Specifically, the construction of the supramolecular coordination chemistry of transition metal complexes has received much attention over the last couple of decades, due to their vital roles in biology [3,4,5], catalysis [6,7], and more importantly, in magneto-chemistry [8,9,10,11,12,13,14]. In this context, polytopic threefold symmetric ligands based on triaminoguanidine are of particular interest in crystal engineering and supramolecular chemistry [15,16,17,18] due to their structural diversity and the chemical versatility of the resulting systems [19,20,21,22,23,24]. Ligands possessing triaminoguanidine core have been successfully employed to develop high-nuclear supramolecular cages, but with diamagnetic metal ions [2,25,26,27,28,29]. Only a couple of open-shell metal ion systems [30,31], copper(II) coordination polymers [32,33], and a series of trinuclear cobalt(II) [34], nickel(II) [35,36], Iron (III) [37], and copper(II) [38,39] triangles have been studied magnetically, with the latter two showing the phenomenon of spin frustration, having significant applications in molecular spintronics [40].
Triaminoguanidine-based ligands with three equivalent coordinating sites could coordinate with three metal ions, leading to an almost planar structure style and rather small metal∙∙∙metal distances bridged through the N−N diazine channels of the central triaminoguanidine moiety [30]. Consequently, the established complexes display strong antiferromagnetic exchange interactions through the σ bond of the N−N diazine arms of the ligand among the metal centers [34,35,36]. In this regard, Plass and co-workers [38,39] developed two trinuclear copper(II) triangles, [Cu3L(py)6]ClO4 (Cu3py) and [Cu3L(bpy)3]ClO4·3DMF (Cu3bpy) (py and bpy are pyridine and bipyridine, respectively). Both complexes exhibit strong antiferromagnetic coupling between the copper(II) ions, leading to a spin-frustrated system. Very recently, by fluctuating the synthetic approach while keeping the same metal–ligand combination, we synthesized two new antiferromagnetic hexa-nuclear CuII complexes, [Cu6L2Cl(µ-OAc)(DMF)3]·DMF (Cu6) and [Cu6L2(µ-Cl)2(DMF)4] (Cu6Cl), in which two planar triangles with strict C3-symmetry were deliberately bridged through acetate and chloride anions in cis and trans fashions, respectively [41]. It was assumed that both Cu6 and Cu6Cl would also present the phenomenon of spin frustration based on their similar structural and experimental magnetic properties to the reference compounds, Cu3py and Cu3bpy.
Coordination supramolecular cages with tetrahedral topologies have been constructed using triaminoguanidine-based ligands, tris (2-hydroxybenzylidine) triaminoguanidinium [H6L]+, and tris (5-bromo-2-hydroxybenzylidine) triaminoguanidinium [H6L1]+, in which the triangular units were connected through (−CdO−)2 four-membered rings (O from the phenolate of the ligand itself), resulting in [{(CdCl)3L}]8− and [{(CdCl)3L1}]8−, respectively [28,42]. The authors concluded that the construction of a comparable discrete single-walled supramolecular tetrahedron with Zn2+ ions was impossible due to the much smaller size of the Zn2+ ion relative to the Cd2+ ion (Zn2+, 0.74 Å; Cd2+, and 0.97 Å) [29]. In the case of a Zn2+ single-walled tetrahedron, the triangular faces would result in closer contact in view of its smaller size, which is sterically unfavorable. Hence, based on this assumption, they developed a double-walled tetrahedron using two different triaminoguanidine-based mixed ligands [29]. Keeping this task in mind, this work planned to use an azide anion as a bridging co-ligand, owing to its linear structure and least steric hindrance. Surprisingly, we synthesized a dodecanuclear metallacycle based on a Cu2+ ion (0.73 Å), [{Cu3L(µ-N3)}4(Py)14]·2Py (Cu12), assembled from four C3-symmetric isosceles triangular faces in close contact. Furthermore, the experimental and theoretical magnetic calculations revealed very strong intra-triangular antiferromagnetic exchange interactions between the copper ions through the σ bond pathway of the N−N diazine channels of the ligand, very similar to that of the reference compounds, Cu3py and Cu3bpy. To our knowledge, this is the first example of a supramolecular Cu12 metallacycle based on triaminoguanidine ligands with four isosceles triangular sub-units that has been synthesized and studied magnetically.

2. Materials and Methods

2.1. Materials and General Information

Tris(2-hydroxybenzylidine)triaminoguanidinium chloride [H6L]Cl (Scheme S1) was synthesized according to the methods in the literature [28,29]. Sodium azide was purchased and used without further purification. The elemental analyses of C, H, and N were completed on a PerkinElmer 2400 analyzer. The powder X-ray diffraction analysis was achieved on a Burker-D8 advance diffractometer using Cu-Kα (λ = 1.5418 Å) radiation at room temperature.

Synthesis of [{Cu3L(µ-N3)}4(Py)14]·2Py·4[CH3OH]·0.8[H2O] (Cu12)

Two freshly prepared solutions, copper(II) salt Cu(OAc)2·6H2O (43.8 mg, 0.15 mmol) in 10 mL of methanol and ligand [H6L]Cl (22.6 mg, 0.05 mmol) in 5 mL of pyridine, were mixed slowly with continued stirring. Then, the sodium azide (0.013 mg, 0.2 mmol) dissolved in 1 mL of water was added dropwise. After 3 h of further stirring, the final solution was filtered out and put aside for slow evaporation. After three days, well-shaped dark green cube-like crystals of Cu12 in a quantitative yield suitable for single crystal measurements were obtained. The elemental analysis (%) calcd for C168H140Cu12N52O12: C: 52.52, H: 3.67, and N: 18.95; and found C: 52.54, H: 3.65, and N: 18.93.

2.2. Crystallography

The details of the crystal data and structure refinement parameters of the Cu12 are summarized in Table S1, while the obtained bond lengths and angles are given in Table S2. To measure the single crystals of the titled complex, they were mounted on glass fiber under a microscope and the diffractional data were collected at 120 (2) K using a Bruker AXS D8 Venture single-crystal diffractometer equipped with graphite-monochromatized Mo Kα (λ = 0.71073 Å) (Table S1). The drawing of the molecular structure and mean plane analysis were obtained using the DIAMOND (version 3.1) software. The structure was solved using direct methods and SHELXT and refined using full-matrix least-squares methods based on F2 (SHELXL) in the Olex2 package [43,44].

2.3. Magnetic Measurements

The magnetic susceptibility measurements were performed on a polycrystalline sample using a Quantum Design MPMS-XL7 SQUID magnetometer in the 2−300 K temperature range, equipped with a 7 T magnet. The diamagnetic corrections of the constituent atoms were determined from Pascal’s constants [45].

2.4. Theoretical Calculations

Based on the wavefunction theory (WFT), using the OpenMolcas software package with the complete active space (CAS) self-consistent field (SCF) approach [46], the electronic structures and magnetic properties of each Cu center were first calculated [47]. This was achieved by replacing the eleven other CuII ions with diamagnetic ZnII centers using the X-ray structures. In combination with the all-electron atomic natural orbital relativistically contracted basis set (ANO-RCC) [48,49], the calculations were first carried out at the scalar (SR) level using the second-order Douglas–Kroll–Hess scalar relativistic Hamiltonian [50,51,52]. The basis sets were contracted to the triple-ζ plus polarization (TZP) quality for the Cu (21s15p10d6f4g2h/6s5p3d2f1g) and Cl (17s12p5d4f2g/5s4p2d1f) atoms, as well as for the N and O atoms coordinated to the paramagnetic center (N, O = 14s9p5d3f2g/4s3p2d1f). The Zn atoms were treated with a double-ζ plus polarization (DZP) basis set (21s15p10d6f4g2h/5s4p2d1f), whereas the rest of the C, N, O, and H atoms were treated with a double-ζ basis set (C, N, O = 14s9p5d3f2g/3s2p; H = 8s4p3d1f/2s). By considering the five lowest spin-doublet states and an active space formed by 11 electrons spanning 11 orbitals, the state-average calculations were performed accordingly. These 11 orbitals contained five 3d orbitals of the CuII ion, five 3d’ orbitals by considering the double-shell effect, and one ligand-centered orbital that could form a bonding σ interaction with the metal-centered orbitals. By using the restricted active space state interaction (RASSI) approach [53], spin–orbit coupling (SOC) was introduced afterwards via a state interaction within the basis of the spin–orbit free states. The EPR g-factors were calculated based on the Reference [54], as implemented in the RASSI module of OpenMolcas, whereas the magnetic susceptibility and magnetization were calculated using the Single-Aniso and Poly-Aniso modules of OpenMolcas, as detailed in Reference [55].
Using the Amsterdam Density Functional (ADF) software package [56,57,58], Kohn–Sham density functional theory (DFT) calculations were carried out to evaluate the magnetic coupling between the CuII centers The calculations of the magnetic coupling constants were performed based on the X-ray structures, in which ten of the paramagnetic CuII atoms were replaced by diamagnetic ZnII centers. Such a replacement produced a model system of two CuII centers with a spin of 1/2, leading to a triplet or a singlet spin state.
These calculations utilized the scalar all-electron zeroth-order regular approximation (ZORA) [59] in combination with the spin-unrestricted formalism. The hybrid functionals, PBE0 [60,61] (Perdew–Burke–Ernzerhof) with 25% of exact Hartree–Fock exchange (HFX), B3LYP [62,63,64,65] (Becke, three-parameter Lee–Yang–Parr) with 20% of HFX, and B3LYP* [66] with 15% of HFX, were employed along with the triple-ζ polarized Slater-type orbital (STO) all-electron basis set, with one set of polarization function for all the atoms (TZP) [67].
The magnetic exchange coupling J was determined from broken symmetry (BS) calculations first proposed by Noodleman [68], in addition to the spin decontamination scheme from Yamaguchi [69,70],
J = 2 E BS E T S ^ 2 T S ^ 2 B S
where E BS and E T are the energy of the BS configuration and the triplet state, respectively, and S ^ 2 BS and S ^ 2 T are the expectation values of the S ^ 2 operator.
The magnetic susceptibility of Cu12 was then calculated using the PolyAniso module, as implemented in the OpenMolcas package, with the following spin Hamiltonian,
H ^ = μ B i g i · S ^ i · B i , j J i j · S ^ i · S ^ j
where the magnetic moments and magnetic coupling constants of each CuII center were obtained from the WFT and DFT calculations, respectively. It is worth noting that the validity of such a computational strategy has been successfully confirmed on a related trinuclear copper(II) complex [39] and closely related Cu6 and CuCl6 complexes [41].

3. Results and Discussions

3.1. Crystal Structure of Cu12

The structural data obtained for [{Cu3L(µ-N3)}4(Py)14]·2Py·4[CH3OH]·0.8[H2O] (Cu12) revealed that the complex crystalizes in the triclinic space group P 1 . The unit cell contains one crystallographically independent dodeca-nuclear neutral complex [{Cu3L(µ-N3)}4(Py)14] and two co-crystallized pyridine molecules (Figure 1, left and Figure S1). The supporting information provides the selected crystallographic data, bond lengths, and angles for Cu12 in Tables S1 and S2.
The asymmetric unit of Cu12 contains half of the molecule [{Cu3L(µ-N3)}2(Py)7]·Py (Figure S2). Each asymmetric unit contains two almost-planar C3-symmetric triangular fragments, which are bridged via Cu3 and Cu6 (6.11 Å) through azide anions (Figure 1, right and Figure S2). Finally, both asymmetric units are again node-to-node connected via Cu2 and Cu4 (5.95 Å) through azide anions to construct a dodeca-nuclear Cu12 metallacycle. In total, four two-connecting node-to-node connections via single azide bridges are provided by pairs of metal centers (Figure 1, right). Within each triangle, the tri-topic ligand leads to an approximately planar symmetrical triangular arrangement through the N−N diazine bridges among the copper centers. In Cu1 to Cu5, all five copper centers are five-coordinated, in which three equatorial coordination sites (NNO) are occupied by the tridentate pockets of the ligand, along with two trapped pyridine molecules for both the Cu1 and Cu5 centers to provide trigonal bipyramid (D3h) arrangements, while in the case of the Cu2 to Cu4 centers, the remaining two coordination sites are filled by one pyridine and one bridging azide anion to accomplish spherical square pyramid (C4v) geometries, respectively (Figure S3 and Table S3). Cu6 is the only metal ion that is four-coordinated and whose fourth position is occupied by the bridging azide anion in cooperating with the planar tridentate pocket (NNO) of the ligand, displaying a square planar (D4h) geometry (Figure S3). The distances within intra-triangular copper ions range between 4.70 Å and 4.90 Å (Figure 1 right). The Cu−N−N−Cu dihedral angles are almost linear, ranging from 147.32° to 177.97°. The calculated bond lengths and dihedral angles of Cu12 are very consistent with the reported spin-frustrated triangles, Cu3py and Cu3bpy [38,39].
In addition, at first glance, an offset face-to-face π···π interaction may exist between the two perfectly parallel benzene rings of the ligands related by an inversion center (Figure 2, left), proved by the short distance of 4.06 (1) Å between the centroids of the benzene rings. Although the vertical distance of 3.36 (4) Å between them corresponds to the formation of a weak π···π interaction between the benzene rings, the planes of the two adjacent benzene rings hardly overlap with each other at all, seeing that the centroid shift distance of 2.26 (9) Å between the two benzene rings is almost equal to the centroid distance of 2.42 (1) Å between the two adjacent side-by-side benzene rings (Figure 2, right). Thus, there was nearly no π···π interaction in the Cu12 and a further negligible exchange interaction attached to the centers of Cu6 and Cu6a can be transmitted along this path.

3.2. Magnetic Properties

Temperature-dependent direct current (dc) molar susceptibility (χMT) measurements for Cu12 were performed in the temperature range of 2 to 300 K (Figure 3). The room temperature χMT value of 3.30 cm3 K mol−1 is significantly lower than the spin-only value expected for the twelve independent CuII ions (χMT = 0.375 cm3 K mol−1 for S = 1/2 and g = 2.0). The χMT products decreased gradually with lowering the temperature, reaching a plateau value of about 1.8 cm3 K mol−1 at around 40 K. This behavior indicates strong antiferromagnetic interactions within the intra-triangular CuII ions, with a ground spin state of S = 1/2. Below 15 K, the χMT value dropped down sharply to reach a value of 1.51 cm3 K mol−1 at 2 K, indicating the existence of weak inter-triangular interactions between the four Cu3 triangles in the Cu12 [34,35,36,38,39]. The field dependence of the magnetization curve was also measured for Cu12 under 2.0 K between 0 and 70 kOe (Figure S4). The obtained magnetization plateau value was ca. 3.96 μB at 70 kOe, which is consistent with the S = 2 spin ground state for the four triangular units (each with S = 1/2) and fully in line with the recently reported spin-frustrated triangles, Cu3py and Cu3bpy, based on the same ligand.

3.3. Theoretical Calculations

The calculated magnetic susceptibility obtained with the different density functionals and experimental data are reported in Figure 3. Between them, a relatively good qualitative agreement was reached for the high-temperature range. This resulted from a relevant determination of the strong antiferromagnetic intra-couplings (Jintra) between the CuII ions of the same triangular unit (Table S4). However, the couplings determined from the different functionals used in this work were not able to properly reproduce the trend of (χMT) in the low-temperature range. The magnetic behavior at low temperatures mainly resulted from the interactions between the triangular units through the cis or trans azido bridges. These (Jazido) couplings were determined as very weakly ferromagnetic, with values in the order of magnitude of the cm−1 (Table S4). This order of magnitude is basically the expected accuracy of the DFT BS evaluation. Hence, one may have some doubts about the ability of this level of theory to properly evaluate these couplings. In addition, the magnetic couplings between copper ions mediated by such end-to-end azido bridges have been deeply investigated in the literature and are expected to be weakly antiferromagnetic [71,72,73,74]. However, using higher levels of theory such as WFT-based methods is untractable regarding the size of the system considered here.
In order to confirm the possible antiferromagnetic behavior of Jazido, we simulated the χMT curve using an average value of Jintra = −250 cm−1 and different values of Jazido. Presented in Figure 4, these results show that using a weak antiferromagnetic Jazido qualitatively allows the reproduction of the magnetic behavior of Cu12, even at the low-temperature range. Using these different antiferromagnetic Jazido results in a non-magnetic ground state. The χMT product increased rapidly with respect to the temperature as the low-lying excited states were populated. For Jazido = −2, −5, and −15 cm−1, the χMT curves reached a plateau at about 5 K, 15 K, and 40 K, respectively, corresponding to populating the triply degenerated excited states at 0.96, 2.37, and 6.88 cm−1, respectively.
The determination of Jazido highlights the difficulty in computing such subtle interactions within complex systems. To explain this difficulty, one may wonder about the relevance of reducing the interactions between the triangular units with the interactions between the closest CuII ions of each unit. However, a more sophisticated treatment appears difficult in the BS-DFT framework, while WFT-based approaches would be untractable.

4. Conclusions

Through synthetic engineering, for the first time, based on a triaminoguanidine ligand, we deliberately constructed a supramolecular [{Cu3L(µ-N3)}4(Py)14]·2Py (Cu12) metallacycle with four triangular sub-units. Each ligand traps three CuII ions in its tridentate-binding cavities (NNO), forming rigid planar triangular structures. This is also the first example of a higher nuclear compound in this metal–ligand combination that has been studied magnetically. While comparing Cu12 with reported spin-frustrated Cu3py and Cu3bpy systems, similar strong antiferromagnetic interactions within the intra-triangular fragments can be observed and we can also expect spin frustration in Cu12. This type of phenomenon in molecular systems is expected to give rise to spin–electric coupling, which allows for a manipulation of the molecular spin states. Hence, Cu12 can be seen as a potential candidate in the future for molecular spintronics.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/magnetochemistry9050122/s1, Figure S1: Molecular structure of Cu12; Figure S2: Asymmetric unit of Cu12 in giving direction; Figure S3: Representation of copper environment geometries for Cu12; Figure S4: Magnetization curve for Cu12 at 2 K; Table S1: Crystal data and structure refinement parameters for Cu12; Table S2: Selected bond lengths (Å) and angles (°) for Cu12; Table S3: Calculated geometries of Cu12 [75]; Table S4: Principal magnetic coupling constants (J in cm−1) for Cu12 calculated at the DFT level with the B3LYP, B3LYP* and PBE0 functionals; Table S5: Calculated energy gap ∆EES1−GS (cm−1) between the first excited state (ES1) and the ground state (GS), and the electronic g-factors for the ground state of the different metal centers in Cu12; Table S6: Relative energy (in cm−1) of the 20 lowest states of the spin Hamiltonian.

Author Contributions

Research design, data manipulation and writing original draft, B.A.; writing—review and editing, X.-L.L. and J.T.; theoretical calculations, G.D., F.G., O.C., W.P. and B.L.G.; supervising the project, J.T. project administration and funding acquisition, J.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China; Chinese Academy of Sciences; Natural Science Foundation of Jilin Province of China; China Postdoctoral Science Foundation; Stratégie d’Attactivité Durable.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data are available by corresponding authors.

Acknowledgments

We thank the National Natural Science Foundation of China (Grant 92261103), the Natural Science Foundation of Jilin Province of China (20200201244JC), the China Postdoctoral Science Foundation (2021M693397) and the Key Research Program of Frontier Sciences, CAS (Grant ZDBS-LY-SLH023) for financial support. X.-L. L. is thankful to the Chinese Academy of Sciences for a Special Research Assistant Grant (2021000162). B. A is grateful for the support through CAS-TWAS President’s Fellowship. G. D., F. G. and B. L. G. thank the French GENCI/IDRIS-CINES centers for high-performance computing resources and acknowledge the Stratégie d’Attactivité Durable (SAD18006—LnCPLSMM) for financial support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Leininger, S.; Olenyuk, B.; Stang, P.J. Self-Assembly of Discrete Cyclic Nanostructures Mediated by Transition Metals. Chem. Rev. 2000, 100, 853–908. [Google Scholar] [CrossRef]
  2. Müller, I.M.; Robson, R. A New Class of Easily Obtained Carbonate-Related μ3-Ligands and a Protein-Sized Doughnut-Shaped Coordination Oligomer. Angew. Chem. Int. Ed. 2000, 39, 4357–4359. [Google Scholar] [CrossRef]
  3. Buccella, D.; Lim, M.H.; Morrow, J.R. Metals in Biology: From Metallomics to Trafficking. Inorg. Chem. 2019, 58, 13505–13508. [Google Scholar] [CrossRef]
  4. Holm, R.H.; Kennepohl, P.; Solomon, E.I. Structural and functional aspects of metal sites in biology. Chem. Rev. 1996, 96, 2239–2314. [Google Scholar] [CrossRef] [PubMed]
  5. Brown, A.C.; Suess, D.L.M. An Open-Cuboidal [Fe3S4] Cluster Characterized in Both Biologically Relevant Redox States. J. Am. Chem. Soc. 2023, 145, 2075–2080. [Google Scholar] [CrossRef] [PubMed]
  6. Manicke, N.; Hoof, S.; Keck, M.; Braun-Cula, B.; Feist, M.; Limberg, C. A Hexanuclear Iron(II) Layer with Two Square-Planar FeO4 Units Spanned by Tetrasiloxide Ligands: Mimicking of Minerals and Catalysts. Inorg. Chem. 2017, 56, 8554–8561. [Google Scholar] [CrossRef]
  7. Nesterov, D.S.; Nesterova, O.V.; Pombeiro, A.J.L. Homo- and heterometallic polynuclear transition metal catalysts for alkane C-H bonds oxidative functionalization: Recent advances. Coord. Chem. Rev. 2018, 355, 199–222. [Google Scholar] [CrossRef]
  8. Doroshenko, I.; Babiak, M.; Buchholz, A.; Tucek, J.; Plass, W.; Pinkas, J. Hexanuclear iron(III) alpha-aminophosphonate: Synthesis, structure, and magnetic properties of a molecular wheel. New J. Chem. 2018, 42, 1931–1938. [Google Scholar] [CrossRef]
  9. Doroshenko, I.; Babiak, M.; Buchholz, A.; Görls, H.; Plass, W.; Pinkas, J. New molecular heptanuclear cobalt phosphonates: Synthesis, structures and magnetic properties. New J. Chem. 2018, 42, 9568–9577. [Google Scholar] [CrossRef]
  10. Kostakis, G.E.; Perlepes, S.P.; Blatov, V.A.; Proserpio, D.M.; Powell, A.K. High-nuclearity cobalt coordination clusters: Synthetic, topological and magnetic aspects. Coord. Chem. Rev. 2012, 256, 1246–1278. [Google Scholar] [CrossRef]
  11. Frost, J.M.; Harriman, K.L.M.; Murugesu, M. The rise of 3-d single-ion magnets in molecular magnetism: Towards materials from molecules? Chem. Sci. 2016, 7, 2470–2491. [Google Scholar] [CrossRef] [PubMed]
  12. Liu, J.; Mrozek, J.; Ullah, A.; Duan, Y.; Baldoví, J.J.; Coronado, E.; Gaita-Ariño, A.; Ardavan, A. Quantum coherent spin–electric control in a molecular nanomagnet at clock transitions. NatPh 2021, 17, 1205–1209. [Google Scholar] [CrossRef]
  13. Boudalis, A.K. Half-Integer Spin Triangles: Old Dogs, New Tricks. Chem. Eur. J. 2021, 27, 7022–7042. [Google Scholar] [CrossRef] [PubMed]
  14. Cañón-Mancisidor, W.; Hermosilla-Ibáñez, P.; Spodine, E.; Paredes-García, V.; Gómez-García, C.J.; Espallargas, G.M.; Venegas-Yazigi, D. Slow Relaxation of the Magnetization on Frustrated Triangular FeIII Units with S = 1/2 Ground State: The Effect of the Highly Ordered Crystal Lattice and the Counteranions. Cryst. Growth Des. 2021, 21, 6213–6222. [Google Scholar] [CrossRef]
  15. Gangu, K.K.; Maddila, S.; Mukkamala, S.B.; Jonnalagadda, S.B. A review on contemporary Metal-Organic Framework materials. Inorg. Chim. Acta 2016, 446, 61–74. [Google Scholar] [CrossRef]
  16. Chen, T.-H.; Popov, I.; Kaveevivitchai, W.; Miljanic, O.S. Metal-Organic Frameworks: Rise of the Ligands. Chem. Mater. 2014, 26, 4322–4325. [Google Scholar] [CrossRef]
  17. Moulton, B.; Zaworotko, M.J. From molecules to crystal engineering: Supramolecular isomerism and polymorphism in network solids. Chem. Rev. 2001, 101, 1629–1658. [Google Scholar] [CrossRef]
  18. Kitagawa, S.; Kitaura, R.; Noro, S.-i. Functional Porous Coordination Polymers. Angew. Chem. Int. Ed. 2004, 43, 2334–2375. [Google Scholar] [CrossRef]
  19. Akintola, O.; Hornig, D.; Buchholz, A.; Görls, H.; Plass, W. Solvent-dependent selective cation exchange in anionic frameworks based on cobalt(II) and triphenylamine linkers: Reactor-dependent synthesis and sorption properties. Dalton Trans. 2017, 46, 8037–8050. [Google Scholar] [CrossRef]
  20. Akintola, O.; Ziegenbalg, S.; Buchholz, A.; Görls, H.; Plass, W. A robust anionic pillared-layer framework with triphenylamine-based linkers: Ion exchange and counterion-dependent sorption properties. CrystEngComm 2017, 19, 2723–2732. [Google Scholar] [CrossRef]
  21. Conerney, B.; Jensen, P.; Kruger, P.E.; MacGloinn, C. The ‘Trinity’ helix: Synthesis and structural characterisation of a C3-symmetric tris-bidentate ligand and its coordination to Ag(I). Chem. Commun. 2003, 11, 1274–1275. [Google Scholar] [CrossRef] [PubMed]
  22. Kisets, I.; Gelman, D. Carbometalated Complexes Possessing Tripodal Pseudo-C3-Symmetric Triptycene-Based Ligands. Organometallics 2018, 37, 526–529. [Google Scholar] [CrossRef]
  23. Lusby, P.J.; Müller, P.; Pike, S.J.; Slawin, A.M.Z. Stimuli-Responsive Reversible Assembly of 2D and 3D Metallosupramolecular Architectures. J. Am. Chem. Soc. 2009, 131, 16398–16400. [Google Scholar] [CrossRef]
  24. Liu, Y.; Xuan, W.; Zhang, H.; Cui, Y. Chirality- and Threefold-Symmetry-Directed Assembly of Homochiral Octupolar Metal−Organoboron Frameworks. Inorg. Chem. 2009, 48, 10018–10023. [Google Scholar] [CrossRef] [PubMed]
  25. Müller, I.M.; Möller, D. A New Ligand for the Formation of Triangular Building Blocks in Supramolecular Chemistry. Eur. J. Inorg. Chem. 2005, 2005, 257–263. [Google Scholar] [CrossRef]
  26. Müller, I.M.; Spillmann, S.; Franck, H.; Pietschnig, R. Rational Design of the First Closed Coordination Capsule with Octahedral Outer Shape. Chem. Eur. J. 2004, 10, 2207–2213. [Google Scholar] [CrossRef]
  27. Müller, I.M.; Möller, D.; Föcker, K. From a Monomer to a Protein-Sized, Doughnut-Shaped Coordination Oligomer—The Influence of Side Chains of C3-Symmetric Ligands in Supramolecular Chemistry. Chem. Eur. J. 2005, 11, 3318–3324. [Google Scholar] [CrossRef]
  28. Müller, I.M.; Robson, R.; Separovic, F. A metallosupramolecular capsule with the topology of the tetrahedron, 33, assembled from four guanidine-based ligands and twelve cadmium centers. Angew. Chem. Int. Ed. 2001, 40, 4385–4386. [Google Scholar] [CrossRef]
  29. Oppel, I.M.; Föcker, K. Rational design of a double-walled tetrahedron containing two different C3-symmetric ligands. Angew. Chem. Int. Ed. 2008, 47, 402–405. [Google Scholar] [CrossRef]
  30. Plass, W. Structural variety and magnetic properties of polynuclear assemblies based on 2-aminoglucose and tritopic triaminoguanidine ligands. Coord. Chem. Rev. 2009, 253, 2286–2295. [Google Scholar] [CrossRef]
  31. Böhme, M.; Ion, A.E.; Kintzel, B.; Buchholz, A.; Görls, H.; Plass, W. Pentanucletear Nickel(II) Complex with two Vertex-Shared Triaminoguanidine Fragments and Symmetric Capping Ligand. Z. Anorg. Allg. Chem. 2020, 646, 999–1009. [Google Scholar] [CrossRef]
  32. Zharkouskaya, A.; Görls, H.; Vaughan, G.; Plass, W. A new three-dimensional Cu(II) coordination polymer with triaminoguanidin building blocks. Inorg. Chem. Commun. 2005, 8, 1145–1148. [Google Scholar] [CrossRef]
  33. Zharkouskaya, A.; Buchholz, A.; Plass, W. A new coordination polymer architecture with (10,3)-a network containing chiral hydrophilic 3-D channels. Eur. J. Inorg. Chem. 2005, 2005, 4875–4879. [Google Scholar] [CrossRef]
  34. Plaul, D.; Böhme, M.; Ostrovsky, S.; Tomkowicz, Z.; Görls, H.; Haase, W.; Plass, W. Modeling Spin Interactions in a Triangular Cobalt(II) Complex with Triaminoguanidine Ligand Framework: Synthesis, Structure, and Magnetic Properties. Inorg. Chem. 2018, 57, 106–119. [Google Scholar] [CrossRef]
  35. Böhme, M.; Schuch, D.; Buchholz, A.; Görls, H.; Plass, W. Spin Interactions and Magnetic Anisotropy in a Triangular Nickel(II) Complex with Triaminoguanidine Ligand Framework. Z. Anorg. Allg. Chem. 2020, 646, 166–174. [Google Scholar] [CrossRef]
  36. Hamblin, J.; Tuna, F.; Bunce, S.; Childs, L.J.; Jackson, A.; Errington, W.; Alcock, N.W.; Nierengarten, H.; Van Dorsselaer, A.; Leize-Wagner, E.; et al. Supramolecular Circular Helicates Formed by Destabilisation of Supramolecular Dimers. Chem. Eur. J. 2007, 13, 9286–9296. [Google Scholar] [CrossRef]
  37. Kintzel, B.; Böhme, M.; Plaul, D.; Görls, H.; Yeche, N.; Seewald, F.; Klauss, H.-H.; Zvyagin, A.A.; Kampert, E.; Herrmannsdörfer, T.; et al. A Trinuclear High-Spin Iron(III) Complex with a Geometrically Frustrated Spin Ground State Featuring Negligible Magnetic Anisotropy and Antisymmetric Exchange. Inorg. Chem. 2023, 62, 3420–3430. [Google Scholar] [CrossRef]
  38. Spielberg, E.T.; Gilb, A.; Plaul, D.; Geibig, D.; Hornig, D.; Schuch, D.; Buchholz, A.; Ardavan, A.; Plass, W. A Spin-Frustrated Trinuclear Copper Complex Based on Triaminoguanidine with an Energetically Well-Separated Degenerate Ground State. Inorg. Chem. 2015, 54, 3432–3438. [Google Scholar] [CrossRef]
  39. Kintzel, B.; Böhme, M.; Liu, J.; Burkhardt, A.; Mrozek, J.; Buchholz, A.; Ardavan, A.; Plass, W. Molecular electronic spin qubits from a spin-frustrated trinuclear copper complex. Chem. Commun. 2018, 54, 12934–12937. [Google Scholar] [CrossRef]
  40. Liu, J.; Mrozek, J.; Myers, W.K.; Timco, G.A.; Winpenny, R.E.P.; Kintzel, B.; Plass, W.; Ardavan, A. Electric Field Control of Spins in Molecular Magnets. Phys. Rev. Lett. 2019, 122, 037202. [Google Scholar] [CrossRef]
  41. Ali, B.; Gendron, F.; Li, X.-L.; Le Guennic, B.; Tang, J. Cis and Trans Linkage of Spin Frustrated Copper Triangles Creating Cu6 Clusters. Chem. Sq. 2020, 4, 4. [Google Scholar] [CrossRef]
  42. Müller, I.M.; Möller, D.; Schalley, C.A. Rational design of tightly closed coordination tetrahedra that are stable in the solid state, in solution, and in the gas phase. Angew. Chem. Int. Ed. 2005, 44, 480–484. [Google Scholar] [CrossRef]
  43. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. Sect. C Struct. Chem. 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  44. Sheldrick, G.M. SHELXT—Integrated space-group and crystal-structure determination. Acta Crystallogr. Sect. A Found. Adv. 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  45. Bain, G.A.; Berry, J.F. Diamagnetic Corrections and Pascal’s Constants. J. Chem. Educ. 2008, 85, 532. [Google Scholar] [CrossRef]
  46. Roos, B.O.; Taylor, P.R.; Sigbahn, P.E.M. A complete active space SCF method (CASSCF) using a density matrix formulated super-CI approach. Chem. Phys. 1980, 48, 157–173. [Google Scholar] [CrossRef]
  47. Fdez. Galván, I.; Vacher, M.; Alavi, A.; Angeli, C.; Aquilante, F.; Autschbach, J.; Bao, J.J.; Bokarev, S.I.; Bogdanov, N.A.; Carlson, R.K.; et al. OpenMolcas: From Source Code to Insight. J. Chem. Theory Comput. 2019, 15, 5925–5964. [Google Scholar] [CrossRef]
  48. Roos, B.O.; Lindh, R.; Malmqvist, P.-Å.; Veryazov, V.; Widmark, P.-O. Main Group Atoms and Dimers Studied with a New Relativistic ANO Basis Set. J. Phys. Chem. A 2004, 108, 2851–2858. [Google Scholar] [CrossRef]
  49. Widmark, P.-O.; Malmqvist, P.-Å.; Roos, B.O. Density matrix averaged atomic natural orbital (ANO) basis sets for correlated molecular wave functions. Theor. Chim. Acta 1990, 77, 291–306. [Google Scholar] [CrossRef]
  50. Douglas, M.; Kroll, N.M. Quantum electrodynamical corrections to the fine structure of helium. Ann. Phys.-N. Y. 1974, 82, 89–155. [Google Scholar] [CrossRef]
  51. Hess, B.A. Applicability of the no-pair equation with free-particle projection operators to atomic and molecular structure calculations. Phys. Rev. A 1985, 32, 756–763. [Google Scholar] [CrossRef] [PubMed]
  52. Hess, B.A. Relativistic electronic-structure calculations employing a two-component no-pair formalism with external-field projection operators. Phys. Rev. A 1986, 33, 3742–3748. [Google Scholar] [CrossRef] [PubMed]
  53. Malmqvist, P.Å.; Roos, B.O.; Schimmelpfennig, B. The restricted active space (RAS) state interaction approach with spin–orbit coupling. Chem. Phys. Lett. 2002, 357, 230–240. [Google Scholar] [CrossRef]
  54. Bolvin, H. An Alternative Approach to the g-Matrix: Theory and Applications. Chemphyschem 2006, 7, 1575–1589. [Google Scholar] [CrossRef] [PubMed]
  55. Chibotaru, L.F.; Ungur, L. Ab initio calculation of anisotropic magnetic properties of complexes. I. Unique definition of pseudospin Hamiltonians and their derivation. J. Chem. Phys. 2012, 137, 064112. [Google Scholar] [CrossRef] [PubMed]
  56. Fonseca Guerra, C.; Snijders, J.G.; te Velde, G.; Baerends, E.J. Towards an order-N DFT method. Theor. Chem. Acc. 1998, 99, 391–403. [Google Scholar] [CrossRef]
  57. te Velde, G.; Bickelhaupt, F.M.; Baerends, E.J.; Fonseca Guerra, C.; van Gisbergen, S.J.A.; Snijders, J.G.; Ziegler, T. Chemistry with ADF. J. Comput. Chem. 2001, 22, 931–967. [Google Scholar] [CrossRef]
  58. Baerends, A.J.A.; Atkins, E.J.Z.T.; Bashford, J.; Baseggio, D.; Bérces, O.; Bo, A.B.F.M.; Boerritger, C.; Cavallo, P.M.; Daul, L.; Chong, C.; et al. ADF2017, SCM, Theoretical Chemistry; Vrije Universiteit: Amsterdam, The Netherlands, 2017; Available online: https://www.scm.com (accessed on 1 January 2022).
  59. van Lenthe, E.; Baerends, E.J.; Snijders, J.G. Relativistic regular two-component Hamiltonians. J. Chem. Phys. 1993, 99, 4597–4610. [Google Scholar] [CrossRef]
  60. Ernzerhof, M.; Scuseria, G.E. Assessment of the Perdew–Burke–Ernzerhof exchange-correlation functional. J. Chem. Phys. 1999, 110, 5029–5036. [Google Scholar] [CrossRef]
  61. Adamo, C.; Barone, V. Toward reliable density functional methods without adjustable parameters: The PBE0 model. J. Chem. Phys. 1999, 110, 6158–6170. [Google Scholar] [CrossRef]
  62. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  63. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef]
  64. Vosko, S.H.; Wilk, L.; Nusair, M. Accurate spin-dependent electron liquid correlation energies for local spin density calculations: A critical analysis. Can. J. Phys. 1980, 58, 1200–1211. [Google Scholar] [CrossRef]
  65. Stephens, P.J.; Devlin, F.J.; Chabalowski, C.F.; Frisch, M.J. Ab Initio Calculation of Vibrational Absorption and Circular Dichroism Spectra Using Density Functional Force Fields. J. Phys. Chem. 1994, 98, 11623–11627. [Google Scholar] [CrossRef]
  66. Salomon, O.; Reiher, M.; Hess, B.A. Assertion and validation of the performance of the B3LYP functional for the first transition metal row and the G2 test set. J. Chem. Phys. 2002, 117, 4729–4737. [Google Scholar] [CrossRef]
  67. Van Lenthe, E.; Baerends, E.J. Optimized Slater-type basis sets for the elements 1–118. J. Comput. Chem. 2003, 24, 1142–1156. [Google Scholar] [CrossRef] [PubMed]
  68. Noodleman, L. Valence bond description of antiferromagnetic coupling in transition metal dimers. J. Chem. Phys. 1981, 74, 5737–5743. [Google Scholar] [CrossRef]
  69. Yamaguchi, K.; Fukui, H.; Fueno, T. Molecular Orbital (Mo) Theory for Magnetically Interacting Organic Compounds. Ab-Initio Mo Calculatins of The Effective Exchange Integrals for Cyclophane-Type Carbene Dimers. Chem. Lett. 1986, 15, 625–628. [Google Scholar] [CrossRef]
  70. Yamaguchi, K.; Takahara, Y.; Fueno, T.; Houk, K.N. Extended Hartree-Fock (EHF) theory of chemical reactions. Theor. Chim. Acta 1988, 73, 337–364. [Google Scholar] [CrossRef]
  71. Fabrizi de Biani, F.; Ruiz, E.; Cano, J.; Novoa, J.J.; Alvarez, S. Magnetic Coupling in End-to-End Azido-Bridged Copper and Nickel Binuclear Complexes:  A Theoretical Study. Inorg. Chem. 2000, 39, 3221–3229. [Google Scholar] [CrossRef]
  72. Cabrero, J.; de Graaf, C.; Bordas, E.; Caballol, R.; Malrieu, J.-P. Role of the Coordination of the Azido Bridge in the Magnetic Coupling of Copper(II) Binuclear Complexes. Chem. Eur. J. 2003, 9, 2307–2315. [Google Scholar] [CrossRef] [PubMed]
  73. Chastanet, G.; Guennic, B.L.; Aronica, C.; Pilet, G.; Luneau, D.; Bonnet, M.-L.; Robert, V. Tuning magnetic exchange using the versatile azide ligand. Inorg. Chim. Acta 2008, 361, 3847–3855. [Google Scholar] [CrossRef]
  74. Le Guennic, B.; Robert, V. From magnetic molecules to magnetic solids: An ab initio expertise. C R Chim. 2008, 11, 650–664. [Google Scholar] [CrossRef]
  75. Casanova, D.; Llunell, M.; Alemany, P.; Alvarez, S. The Rich Stereochemistry of Eight-Vertex Polyhedra: A Continuous Shape Measures Study. Chem. Eur. J. 2005, 11, 1479–1494. [Google Scholar] [CrossRef]
Figure 1. (left) Molecular structure of Cu12, coordinated and solvent pyridine molecules have been omitted for clarity; and (right) triangular arrangements with calculated distances between the copper centers along with single-bridged mode (through azide anion) of four planar trinuclear units. Color code: orange, Fe; red, O; blue, N; grey, C; and light grey, H.
Figure 1. (left) Molecular structure of Cu12, coordinated and solvent pyridine molecules have been omitted for clarity; and (right) triangular arrangements with calculated distances between the copper centers along with single-bridged mode (through azide anion) of four planar trinuclear units. Color code: orange, Fe; red, O; blue, N; grey, C; and light grey, H.
Magnetochemistry 09 00122 g001
Figure 2. The view of the π···π stacking interaction within the Cu12 attached to the centers of Cu6 and Cu6a; the coordinated and solvent pyridine molecules have been omitted for clarity; the relative benzene rings were highlighted with purple and blue colors, respectively. Color code: orange, Fe; red, O; blue, N; grey, C; and light grey, H.
Figure 2. The view of the π···π stacking interaction within the Cu12 attached to the centers of Cu6 and Cu6a; the coordinated and solvent pyridine molecules have been omitted for clarity; the relative benzene rings were highlighted with purple and blue colors, respectively. Color code: orange, Fe; red, O; blue, N; grey, C; and light grey, H.
Magnetochemistry 09 00122 g002
Figure 3. Calculated (solid lines) and experimental (empty circles) magnetic susceptibilities (χMT) as a function of the temperature. Theoretical results are shown for magnetic couplings computed from DFT broken symmetry calculations using B3LYP (red), B3LYP* (green), and PBE0 (blue) functions. Magnetic moments of CuII are obtained with the CASSCF method. Details of calculations are presented in Tables S4–S6.
Figure 3. Calculated (solid lines) and experimental (empty circles) magnetic susceptibilities (χMT) as a function of the temperature. Theoretical results are shown for magnetic couplings computed from DFT broken symmetry calculations using B3LYP (red), B3LYP* (green), and PBE0 (blue) functions. Magnetic moments of CuII are obtained with the CASSCF method. Details of calculations are presented in Tables S4–S6.
Magnetochemistry 09 00122 g003
Figure 4. Calculated (solid lines) and experimental (filled circles) magnetic susceptibilities (χMT) as a function of the temperature. Theoretical results are shown for an average Jintra of −250 cm−1 and different values for Jazido. Magnetic moments of CuII are obtained with the CASSCF method. Details of calculations are presented in Table S6.
Figure 4. Calculated (solid lines) and experimental (filled circles) magnetic susceptibilities (χMT) as a function of the temperature. Theoretical results are shown for an average Jintra of −250 cm−1 and different values for Jazido. Magnetic moments of CuII are obtained with the CASSCF method. Details of calculations are presented in Table S6.
Magnetochemistry 09 00122 g004
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ali, B.; David, G.; Gendron, F.; Li, X.-L.; Cador, O.; Plass, W.; Le Guennic, B.; Tang, J. A Cu12 Metallacycle Assembled from Four C3-Symmetric Spin Frustrated Triangular Units. Magnetochemistry 2023, 9, 122. https://doi.org/10.3390/magnetochemistry9050122

AMA Style

Ali B, David G, Gendron F, Li X-L, Cador O, Plass W, Le Guennic B, Tang J. A Cu12 Metallacycle Assembled from Four C3-Symmetric Spin Frustrated Triangular Units. Magnetochemistry. 2023; 9(5):122. https://doi.org/10.3390/magnetochemistry9050122

Chicago/Turabian Style

Ali, Basharat, Grégoire David, Frédéric Gendron, Xiao-Lei Li, Olivier Cador, Winfried Plass, Boris Le Guennic, and Jinkui Tang. 2023. "A Cu12 Metallacycle Assembled from Four C3-Symmetric Spin Frustrated Triangular Units" Magnetochemistry 9, no. 5: 122. https://doi.org/10.3390/magnetochemistry9050122

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

Article Metrics

Back to TopTop