Next Article in Journal
Superparamagnetic-like Micrometric Single Crystalline Magnetite for Biomedical Application Synthesis and Characterization
Previous Article in Journal
Iron Oxides Nanoparticles as Components of Ferroptosis-Inducing Systems: Screening of Potential Candidates
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Assembly of Grid-type Lanthanide Cluster

1
School of Chemistry and Chemical Engineering, Northwestern Polytechnical University, Xi’an 710072, China
2
Institute of Flexible Electronics (IFE), Northwestern Polytechnical University, Xi’an 710072, China
*
Authors to whom correspondence should be addressed.
Magnetochemistry 2023, 9(1), 4; https://doi.org/10.3390/magnetochemistry9010004
Submission received: 20 November 2022 / Revised: 20 December 2022 / Accepted: 21 December 2022 / Published: 24 December 2022

Abstract

:
A dicompartmental Schiff base ligand was synthesized and used for the assembly of a lanthanide grid-like complex. Dinuclear Dy2 and tetranuclear Dy4 complexes were isolated from the reaction of the ligand with different dysprosium salt. Single crystal X-ray diffractions show that the two DyIII ions in Dy2 are adopted in the N3O coordination pockets of the ligand and further coordinated by water molecules, whereas, for Dy4, the four DyIII ions are clamped by four ligands through their terminal N3O coordination pockets, forming a grid-type assembly. Magnetic studies reveal that complex Dy2 shows field-induced single-molecule magnetic behavior under 1000 Oe dc field, complex Dy4 shows fast relaxation under zero field and field-induced single-molecule magnet (SMM) behavior under 500 Oe. The difference in the magnetic relaxation is related to the various deprotonation of the ligand and distinct topology of the assemblies.

Graphical Abstract

1. Introduction

Lanthanides single-molecule magnets (Ln-SMMs) have attracted much attention since the first discovery of (Bu4N)[Tb(Pc)2] [1], which shows record energy barriers and different relaxation mechanisms from that of the transition-metal-cluster SMMs. Since then, lanthanide has become the best candidate for the design and synthesis of high-performance SMMs [2,3,4,5] due to its potential applications in high-density information storage, quantum information processing [6], and molecular spintronic materials [7,8,9,10,11,12]. During the last two decades, thousands of Ln-SMM have been reported, and the record energy barrier (Ueff) and blocking temperature (TB) were refreshed up to 1541 cm−1 and 80 K,[13] respectively, pushing the potential application of SMMs into reality. However, due to the presence of fast quantum tunneling of magnetization (QTM), the relaxations of magnetizations in most of the reported SMMs are short cut, resulting in a relatively low blocking temperature [14,15,16].
One efficient way for suppressing QTM in Ln-SMM is restricting the lanthanide ions in an axial ligand field, such as C∞v [17,18], D4d [19,20,21,22], D5h [23,24,25], and D6d coordination geometries for DyIII ions. Due to the inherent high coordination number and flexible coordination nature of lanthanide ions, it is very difficult and challenging to fix the lanthanide ions in a perfect due coordination geometry. Alternatively, introducing magnetic interaction between lanthanide ions can suppress the QTM efficiently [26], especially in polynuclear SMMs. For instance, the reported single-electron Ln–Ln coupling in dimetallofullerenes shows very strong exchange interactions between 4f moments that result in a gigantic coercivity of 8.2 teslas at 5 K and blocking temperature up to 25.2 K [27,28,29]. Long and co-workers reported a mixed-valence dilanthanide complexes (CpiPr5)2Dy2I3 that gives rise to an enormous coercive magnetic field with a lower bound of 14 tesla up to 60 K [30]. Additionally, high-performance polynuclear SMMs are also found in radical [31,32,33] and μ2-OH bridged [34,35,36] dinuclear lanthanide complexes, in which the spin centers are in axial ligand field and strongly coupled, suppressing the QTM efficiently. For Ln-SMMs with higher nuclearity, the poor axial ligand field, as well as the complicated topology of the molecule, make it difficult to manipulate the relaxation processes and suppress the QTM [37,38,39].
Our research interests are focused on the unique magnetic behavior [40] in grid-type lanthanide assemblies, in which the QTM could be suppressed by the magnetic interaction between lanthanide ions [41]. For this, we have designed several types of ditopic and tritopic hydrazone ligands with suitable coordination pockets to accommodate lanthanide ions and arrange them in a grid [42,43,44] or hexagonal [45,46,47] topology. Herein, the linkers in the ligands are very important for the transmission of intermolecular magnetic interactions. As was reported in the literature, pyrimidine is a good linker to arrange transition metal ions into grid-type topologies [48,49,50,51,52]. However, the assembly of lanthanide grids based on pyrimidine ligands is rare [53]. Herein, we designed a pyrimidine-based dicompartmental Schiff base ligand with a pyrimidine linker, by using which a dinuclear Dy2 was isolated (Scheme 1). After optimizing the reaction conditions, tetranuclear Dy4 featuring a [2 × 2] grid topology was also obtained. The magnetic study revealed that complex Dy2 shows field-induced single-molecule magnetic behavior under a 1000 Oe dc field. Dy4 shows fast relaxation under zero field, which was slowed down by the application of 500 Oe, resulting in the field-induced single-molecule magnet.

2. Materials and Methods

2.1. Synthesis of Ligand H2L

A 100-mL flask was charged with a magnetic stirrer bar and 4,6-dihydrazineylpyrimidine [54] (1.4 g, 10 mmol, 1 eq) in 50 mL methanol. 6-(hydroxymethyl)-2-carbaldehyde-pyridine [55] (2.74 g, 20 mmol, 2 eq) was then added. The reaction was stirred and refluxed overnight. Upon cooling, the precipitate was filtered, washed with ether, and dried under vacuum, giving the pure product H2L (((pyrimidine-4,6-diylbis(hydrazin-2-yl-1-ylidene))bis(methaneylylidene))bis(pyridine-6,2- diyl))dimethanol) as a pale yellow powder. Yield: 3.1 g (82%). Elemental analysis (%) calcd for C18H18N8O2, MW = 378.16: C, 57.14, H, 4.79, N, 29.61; found C, 57.28, H, 4.81, N, 29.65. 1H NMR (500 MHz, DMSO-d6) δ = 11.43 (2 H, s), 8.24 (1 H, d, J 0.9), 8.14 (2 H, s), 7.93–7.85 (4 H, m), 7.46 (2 H, d, J 7.3), 6.94 (1 H, s), 5.47 (2 H, d, J 6.0), 4.62–4.58 (4 H, m). IR (KBr disks) [cm−1] = 3394 (w), 3190 (m), 3047 (w), 1608 (s), 1585 (s), 1562 (s), 1504 (m), 1462 (m), 1421 (m), 1375 (m), 1323 (m), 1255 (m), 1198 (m), 1134 (s), 1084 (m), 993 (m), 916 (w), 833 (s), 746 (s), 719 (m), 651 (m), 624 (w), 561 (w), 501 (w) 463 (w).

2.2. Synthesis of Dy2

DyBr3·9H2O (0.6 mmol) and ligand H2L (0.2 mmol) in a mixture of 10 mL methanol and 5 ml DCM was stirred at room temperature for 10 min. The reaction was stirred at room temperature overnight. The solution was filtered and exposed to air to allow the slow evaporation of the solvent. Yellow crystals of Dy2 suitable for X-ray diffraction were obtained after 4 days. Yield: 90 mg, (32%, based on metal salt). Elemental analysis (%) calcd for [Dy2(H2L)(H2O)10]·Br6·3H2O (C18H44Br6Dy2N8O15, MW = 1417.07): C, 15.26, H, 3.14, N, 7.91; found C, 15.28, H, 3.11, N, 7.85. IR (KBr disks) [cm−1] = 3336 (br, w), 3207 (br, w), 3060 (br, w), 1635 (s), 1610 (s), 1568 (m), 1531 (w), 1514 (w), 1487 (m), 1456 (m), 1410 (w), 1290 (m), 1221 (w), 1188 (s), 1146 (m), 1101 (w), 1049 (w), 1012 (w), 827 (w), 740 (w), 696 (w), 655 (w), 605 (w), 567 (w), 544 (m), 501 (w), 476 (m), 459 (w), 447 (w).

2.3. Synthesis of Dy4

A mixture of Dy(ClO4)3·6H2O (0.3 mmol) and ligand H2L (0.2 mmol) in a mixture of 10 mL methanol and 5 ml DCM was stirred at room temperature for 10 min, triethylamine (0.2 mmol) was then added dropwise. The reaction was stirred at room temperature overnight. The solution was filtered and exposed to air to allow the slow evaporation of the solvent. Yellow crystals of Dy4 suitable for X-ray diffraction were obtained after 5 days. Yield: 97 mg (65%, based on metal salt). Elemental analysis (%) calcd for [Dy4L4]·(ClO4)4·12MeOH·4H2O (C84H120Cl4Dy4N32O40, MW = 3009.91): C, 33.52, H, 4.02, N, 14.89; found C, 33.49, H, 4.03, N, 14.78. IR (KBr disks) [cm−1] = 3456 (br, w), 3343 (br, w), 3016 (br, w), 2981 (br, w), 1635 (s), 1606 (s), 1558 (m), 1533 (m), 1477 (s), 1417 (m), 1375 (w), 1297 (m), 1223 (w), 1193 (s), 1145 (w), 1101 (s), 1036 (w), 1006 (w), 898 (w), 771 (w), 654 (w), 625 (m), 597 (w), 578 (w), 525 (w), 501 (w), 469 (w).

2.4. Crystallography

Single-crystal X-ray diffraction data were collected by the Bruker Apex II CCD diffractometer (Bruker AXS GMBH, Germany), using graphite-monochromatized Mo-Kα radiation (λ = 0.71073 Å). In the Olex2 package [56], the structures were solved by using SHELXT [57] (direct methods), and all non-hydrogen atoms were refined by using SHELXL [58] (full-matrix least-squares techniques) on F2 with anisotropic thermal parameters). All hydrogen atoms were introduced in calculated positions and refined with fixed geometry relative to their carrier atoms. The solvent voids (164.1 Å3 and 14.1 e peer cell) in Dy4 were refined using a solvent masking routine in the Olex2 package. Due to the poor diffraction of Dy4, it is difficult to introduce hydrogen atoms on some of the solvent molecules in the lattice. Crystallographic data of Dy2 and Dy4 are listed in Table S1. CCDC 2220302 and 2220303 contain supplementary crystallographic data for this paper.

2.5. Magnetic Measurements

Magnetic measurements were performed by using a Quantum Design MPMS-XL-7 SQUID magnetometer (Quantum Design, United States) equipped with a 7 T magnet. Susceptibility measurements were carried out on the polycrystalline sample of the two complexes. In the temperature range of 2–300 K, the direct-current (dc) susceptibility measurements were obtained under an applied field of 1000 Oe. Diamagnetic corrections were made with Pascal’s constants [59] for all the constituent atoms and the contributions of the sample holder. The field-dependent magnetizations were obtained in the field range of 0−7 T. In the frequency range of 1–1488 Hz, the alternating-current (ac) susceptibility measurements were obtained in a 3 Oe ac oscillating field under various dc fields.

3. Results and Discussions

3.1. Structures of Dy2 and Dy4

The ligand H2L ((((pyrimidine-4,6-diylbis(hydrazin-2-yl-1-ylidene)) bis(methaneylylidene))bis(pyridine-6,2-diyl))dimethanol) was synthesized from the condensation of 4,6-dihydrazineylpyrimidine and 6-(hydroxymethyl)-2- carbaldehyde-pyridine in good yield (Figure S1). As shown in Scheme 1, two four-membered O-N-N-N pockets reside at both sides of the ligand, which can capture two DyIII ions in a planar arrangement. Indeed, a dinuclear Dy2, with formula of [Dy2(H2L)(H2O)10]·Br6·3H2O, was obtained from the reaction of DyBr3·9H2O (0.6 mmol) and ligand H2L. Yellow crystals of Dy2 suitable for X-ray diffraction were obtained after the workup. Single-crystal X-ray diffraction revealed that complex Dy2 crystallizes in the triclinic space group P-1 with Z = 2 (Table S1). The asymmetric unit contains one ligand, two DyIII ions, ten coordinated water, six free Br-, and three water molecules in the lattice. The ligand is not deprotonated and coordinates with the DyIII ions by its coordination pockets on both sides (Figure 1a). Dy1 and Dy2 reside in the four-membered O-N-N-N pocket, with each further coordinated with five H2O molecules, forming a nine-coordination environment (O6N3). The Dy-O bonds are in the range of 2.37–2.46 Å, which is much shorter than the Dy-N bond (2.50–2.57 Å, Table S1), probably because of the hard-base behavior of the coordinated O atoms. The DyIII ions in the molecule are separated by the pyrimidine group with a distance of 6.67 Å. The Br- anions reside in the lattice to balance the positive charge of the molecule and form hydrogen bondings with the coordinated and free H2O, bridging the molecules into three-dimensional frameworks (Figure S6).
After adjusting the ligand–metal ratio and reaction conditions, fortunately, a tetranuclear grid-like complex Dy4 ([Dy4L4]·(ClO4)4·12MeOH·4H2O) was isolated. Single crystal X-ray diffraction shows that complex Dy4 crystallizes in the tetragonal space group P42/nmc with Z = 2. The molecule contains four double deprotonated ligands, four DyIII ions, four perchloride, and some solvent molecules (MeOH and H2O) in the lattice. The molecule features a [2 × 2] grid topology with the four ligands as the edges and four DyIII ions bending in the corners (Figure 1b). Due to the high symmetry of the molecule, all the DyIII ions and ligands are identical, with each DyIII ion clamped by two N3O coordination pockets from two crossed ligands, forming an eight-coordination environment. The ligands are deprotonated on both hydroxymethyl group sites, resulting in relatively short Dy-O bonds (2.27 Å), which is also much shorter than that of in complex Dy2 (Dy-O > 2.39 Å). The shortest intramolecular Dy···Dy distance is 6.9 Å longer than that of Dy2 (6.67 Å). This can be ascribed to the deprotonation of the hydroxymethyl groups that drag the DyIII ions close to the terminal sites of the ligand. The long Dy···Dy distance suggests the weak or negligible magnetic exchange interaction within the complex. The molecules pack along the crystallographic C axis and are separated by the ClO4- and solvent molecules in the lattice.
To gain more insight into the coordination difference of DyIII ions in the two complexes, the coordination geometries of the DyIII centers were quantified by the SHAPE analyses [60,61]. As shown in Figure S8, the DyIII ions in Dy2 are both nine coordinated, and the coordination geometry of Dy1 and Dy2 in Dy2 are close to Muffin (Cs) with CShM value of 1.369 and capped square antiprism (C4v) with CShM value of 0.743 (Table S4), respectively. In Dy4, the four DyIII ions are identical, and eight are coordinated in an N6O2 coordination environment, the coordination geometry for Dy1 in a large distorted augmented trigonal prism (C2v) with a CShM value of 3.284. The difference in the coordinate geometry of the DyIII ions in Dy2 and Dy4 will influence the magnetic relaxation significantly (see below).

3.2. Magnetic Properties of Dy2 and Dy4

3.2.1. Static Magnetic Properties of Dy2 and Dy4

Direct current (dc) magnetic susceptibility measurements were carried out on polycrystalline samples under an applied field of 1000 Oe in the temperature range of 2–300 K. As shown in Figure 2, the temperature-dependent χMT product (χM = molar magnetic susceptibility) of the two complexes show similar profiles. The χMT at 300 K are 27.44 and 53.39 cm3·K·mol−1 for Dy2 and Dy4, respectively, which are close to the expected value for two (28.34 cm3·K·mol−1) and four (56.68 cm3·K·mol−1) free-ion approximation of DyIII ions (S = 5/2, L = 5, 6H15/2, g = 4/3). Upon cooling, the χMT products decrease gradually before a quick drop below 10 K and reach the value of 17.78 cm3·K·mol−1 (Dy2) and 40.41 cm3·K·mol−1 (Dy4) at 2 K. The quick drop of the χMT products at low temperatures can be ascribed to the thermal depopulation of the Stark sublevels as well as the possibility of dominated weak antiferromagnetic interaction between the DyIII ions.
The field-dependent molar magnetizations (M) vs. H plot for Dy2 and Dy4 at 1.9 K (Figures S10 and S11) shows a sharp increase at low fields and then linear increases at high fields, reaching the values of 10.68 μB (Dy2) and 19.85 μB (Dy4) at 7 T, which are close to the expected value for two (10 μB) and four DyIII ions (20 μB) in a pure mJ = |±15/2 > ground state. The M versus H curves at 3 and 5 K also show similar increasing tendencies but do not overlap on the curve at 1.9 k, which can be attributed to significant magnetic anisotropy and/or the existence of low-lying excited states.

3.2.2. Dynamic Magnetic Properties of Dy2 and Dy4

To investigate the dynamics of magnetization, alternating current (AC) susceptibility measurements were carried out with an oscillating field of 3.0 Oe. However, complex Dy2 does not show any relaxation under zero dc field. It seems the magnetic interactions between the two DyIII ions are very weak and cannot suppress the fast relaxation of the complex. Field-dependent ac susceptibility measurements were carried out at 1.9 K with a frequency of 1000 Hz. As shown in Figure S12, the out-of-phase ac susceptibility (χ″) peak is detected around 1000 Oe, which should be the optimized dc field that can slow down the magnetic relaxation. Thus, we performed the ac susceptibility measurements under 1000 Oe dc field, and the temperature-dependent χ″ showed frequency-dependent peaks at low temperatures (Figure S13), indicating the slow relaxation of the magnetization. The frequency-dependent χ″ peaks shift to a higher frequency when raising the temperature, suggesting the thermal relaxation behaviors. For complex Dy4, the χ″ signals are observed under zero dc field but without peaks (Figure S14). This is probably related to the fast relaxation of QTM that shortcuts the relaxation of the magnetization.[62,63] This kind of relaxation can usually be suppressed by the application of dc field.[64,65] The optimized dc field was determined through the field-dependent ac susceptibility measurements (Figure S15), giving the optimized dc field of 500 Oe. After applying this dc field, frequency- and temperature-dependent χ″ peaks were observed in the ac measurements (Figure 3 and Figure S16), indicative of the field-induced single-molecule magnetic property.
To explore the magnetic relaxation mechanics, Cole–Cole plots were represented as χ″ versus χ′ (Figure 4). For complexes Dy2 and Dy4, the Cole–Cole plots show semi-circular profiles (Figure 4), suggesting the presence of a homogeneous relaxation process. Fitting the Cole–Cole plots with the CC-FIT2 program [66] using the Debye model [67] gives the temperature-dependent relaxation time. The relevant parameters of the fitting results are listed in Tables S5 and S6. Thereafter, we fit temperature-dependent relaxation by the following equation [68] to investigate the mechanics of the relaxation process:
1 τ obs = 1 τ Q T M + A H 4 T + C T n + τ 0 1 exp ( U eff / T )
where 1/τQTM, AH4T, CTn, and τ0−1·exp(−Ueff/T) represent quantum tunneling, direct, Raman, and Orbach relaxation processes, respectively. The relevant parameters obtained from the best fittings are listed in Table S7. Both complexes show Raman and Orbach relaxation processes with comparative Raman relaxation parameters (Table S7). The energy barriers of the two complexes are also comparative with Ueff = 17(4) K, τ0 = 4.3(3) × 10−7 s, C = 144(5) s−1·Kn, and n = 3(0.3) for Dy2, and Ueff = 16(2) K, τ0 = 1.0(2) × 10−3 s, C = 98(9) s−1·K−n, and n = 3(0.8) for Dy4.
The low energy barriers for both complexes are probably related to the poor axial ligand field of the DyIII ions. In complex Dy2, the DyIII ions reside in nine coordination environments with Dy-O bond length in the range of 2.37–2.46 Å and Dy-N bond of 2.50–2.57 Å, which indicate a relatively weak ligand field. For Dy4, the deprotonated hydroxymethyl groups give two shortest Dy-O bonds of 2.27 Å, which probably provides an axial ligand field that dominates the anisotropy of the complex. In order to investigate the anisotropies differences in the two complexes, we use the Magellan program [69] to calculate the anisotropy axes of each DyIII ion. The calculation is based on the X-ray structures and electrostatic model of the complexes, with the DyIII ions containing three positive charges and the coordinated O and free Br- possessing one negative charge. For complex Dy2, the anisotropy axes of Dy1 and Dy2 are close to the cape-point of the capped square antiprism coordination polyhedron (Figure 5). For complex Dy4, the anisotropy axis of the Dy1 is parallel to the O1···O1’ direction (Figure 5), which is beneficial from the shortest Dy-O bonds (2.27 Å) that provides a relatively stronger axial ligand field. Therefore, complex Dy4 shows magnetic relaxation under zero field and a relatively slower relaxation time at low temperatures than that of Dy2.

4. Conclusions

In conclusion, we synthesized a novel dicompartmental Schiff base ligand based on pyrimidine, by using which a dinuclear Dy2 and tetranuclear Dy4 grid were constructed. Due to the distinct coordination environments and topology in the two complexes, the magnetic relaxation of these complexes is different. In Dy2, the DyIII ions are in a relatively weak ligand field with relatively long coordination bonds; therefore, the complexes only show field-induced single-molecule magnet behavior. For complexes Dy4, the deprotonated ligand gives a pair of relatively short Dy-O bonds, resulting in the axial ligand field that slows down the relaxation. Therefore, complex Dy4 shows relaxation under zero dc field and field-induced single-molecule magnetic behavior under 500 Oe dc field. Although the magnetic interactions in both complexes are weak and have no positive effect on the magnetic relaxation, the assembly of grid-type lanthanide clusters in the work provides an efficient strategy to regulate the topology of lanthanide ions. The next step will be focused on the enhancement of the magnetic coupling between lanthanide ions.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/magnetochemistry9010004/s1, Figure S1: Schematic drawings of the synthetic route of ligand H2L, Figure S2: 1H-NMR spectrum of ligand H2L in DMSO-d6, Figures S3–S5: IR spectra of ligand H2L, complexes Dy2 and Dy4, Figures S6 and S7: Packing models of complexes Dy2 and Dy4, Figures S8 and S9: Coordination polyhedrons of DyIII ions in complexes Dy2 and Dy4, Figures S10 and S11: Field-dependent molar magnetization of Dy2 and Dy4, Figures S12–S16: Field- and temperature-dependent ac susceptibility of Dy2 and Dy4, Figures S17 and S18: Plots of τ vs. T−1 for Dy2 and Dy4, Figures S19 and S20: Orientations of the main magnetic axes of the ground state of Dy2 and Dy4, Table S1: Crystallographic data of Dy2 and Dy4, Tables S2 and S3: Selected bond distances (Å) and angles (º) of Dy2 and Dy4, Table S4: The CShM values calculated by SHAPE 2.1 of DyIII ions in Dy2 and Dy4, Tables S5 and S6: CC-Fit results for frequency-dependent ac susceptibility of Dy2 and Dy4, Table S7: Parameters obtained from fitting the plots of the relaxation time τ vs. 1/T for Dy2 and Dy4.

Author Contributions

Conceptualization, J.L.; Methodology, F.Z.; Software, D.L.; Formal analysis, F.Z. and X.G.; Investigation, J.L. and F.Z.; Data curation, X.G. and D.L.; Writing—original draft, J.L.; Writing—review & editing, D.L. and J.W.; Visualization, J.W.; Supervision, J.W.; Project administration, J.W.; Funding acquisition, J.W. 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, 22101229” and “The Open Funds of the State Key Laboratory of Rare Earth Resource Utilization, RERU2022021”.

Data Availability Statement

The data are available by corresponding authors.

Acknowledgments

We thank the National Natural Science Foundation of China (22101229) and the Open Funds of the State Key Laboratory of Rare Earth Resource Utilization (RERU2022021) for financial support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ishikawa, N.; Sugita, M.; Ishikawa, T.; Koshihara, S.-Y.; Kaizu, Y. Lanthanide Double-Decker Complexes Functioning as Magnets at the Single-Molecular Level. J. Am. Chem. Soc. 2003, 125, 8694–8695. [Google Scholar] [CrossRef] [PubMed]
  2. Feng, M.; Tong, M.L. Single Ion Magnets from 3d to 5f: Developments and Strategies. Chem. Eur. J. 2018, 24, 7574–7594. [Google Scholar] [CrossRef] [PubMed]
  3. Meng, Y.-S.; Jiang, S.-D.; Wang, B.-W.; Gao, S. Understanding the Magnetic Anisotropy toward Single-Ion Magnets. Acc. Chem. Res. 2016, 49, 2381–2389. [Google Scholar] [CrossRef] [PubMed]
  4. Sessoli, R.; Powell, A.K. Strategies towards single molecule magnets based on lanthanide ions. Coord. Chem. Rev. 2009, 253, 2328–2341. [Google Scholar] [CrossRef]
  5. Clemente-Juan, J.M.; Coronado, E.; Gaita-Arino, A. Magnetic polyoxometalates: From molecular magnetism to molecular spintronics and quantum computing. Chem. Soc. Rev. 2012, 41, 7464–7478. [Google Scholar] [CrossRef]
  6. Akiyoshi, R.; Zenno, H.; Sekine, Y.; Nakaya, M.; Akita, M.; Kosumi, D.; Lindoy, L.F.; Hayami, S. A Ferroelectric Metallomesogen Exhibiting Field-Induced Slow Magnetic Relaxation. Chem. Eur. J. 2022, 28, e202103367. [Google Scholar] [CrossRef]
  7. Sessoli, R.; Tsai, H.L.; Schake, A.R.; Wang, S.; Vincent, J.B.; Folting, K.; Gatteschi, D.; Christou, G.; Hendrickson, D.N. High-spin molecules: [Mn12O12(O2CR)16(H2O)4]. J. Am. Chem. Soc. 1993, 115, 1804–1816. [Google Scholar] [CrossRef]
  8. Gatteschi, D.; Caneschi, A.; Pardi, L.; Sessoli, R. Large Clusters of Metal Ions: The Transition from Molecular to Bulk Magnets. Science 1994, 265, 1054–1058. [Google Scholar] [CrossRef]
  9. Cosquer, G.; Shen, Y.; Almeida, M.; Yamashita, M. Conducting single-molecule magnet materials. Dalton Trans. 2018, 47, 7616–7627. [Google Scholar] [CrossRef]
  10. Dey, A.; Kalita, P.; Chandrasekhar, V. Lanthanide(III)-Based Single-Ion Magnets. ACS Omega 2018, 3, 9462–9475. [Google Scholar] [CrossRef]
  11. Guo, F.-S.; Bar, A.K.; Layfield, R.A. Main Group Chemistry at the Interface with Molecular Magnetism. Chem. Rev. 2019, 119, 8479–8505. [Google Scholar] [CrossRef] [PubMed]
  12. Zhu, Z.; Li, X.-L.; Liu, S.; Tang, J. External stimuli modulate the magnetic relaxation of lanthanide single-molecule magnets. Inorganic Chemistry Frontiers 2020, 7, 3315–3326. [Google Scholar] [CrossRef]
  13. Guo, F.-S.; Day, B.M.; Chen, Y.-C.; Tong, M.-L.; Mansikkamäki, A.; Layfield, R.A. Magnetic hysteresis up to 80 kelvin in a dysprosium metallocene single-molecule magnet. Science 2018, 362, 1400–1403. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Ishikawa, N.; Sugita, M.; Wernsdorfer, W. Quantum Tunneling of Magnetization in Lanthanide Single-Molecule Magnets: Bis(phthalocyaninato)terbium and Bis(phthalocyaninato)dysprosium Anions. Angew. Chem. Int. Ed. 2005, 44, 2931–2935. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Wernsdorfer, W.; Sessoli, R. Quantum Phase Interference and Parity Effects in Magnetic Molecular Clusters. Science 1999, 284, 133–135. [Google Scholar] [CrossRef] [Green Version]
  16. Gatteschi, D.; Sessoli, R. Quantum Tunneling of Magnetization and Related Phenomena in Molecular Materials. Angew. Chem. Int. Ed. 2003, 42, 268–297. [Google Scholar] [CrossRef]
  17. Guo, F.S.; Day, B.M.; Chen, Y.C.; Tong, M.L.; Mansikkamaki, A.; Layfield, R.A. A Dysprosium Metallocene Single-Molecule Magnet Functioning at the Axial Limit. Angew. Chem. Int. Ed. 2017, 56, 11445–11449. [Google Scholar] [CrossRef] [Green Version]
  18. Goodwin, C.A.P.; Ortu, F.; Reta, D.; Chilton, N.F.; Mills, D.P. Molecular magnetic hysteresis at 60 kelvin in dysprosocenium. Nature 2017, 548, 439–442. [Google Scholar] [CrossRef] [Green Version]
  19. Martynov, A.G.; Horii, Y.; Katoh, K.; Bian, Y.; Jiang, J.; Yamashita, M.; Gorbunova, Y.G. Rare-earth based tetrapyrrolic sandwiches: Chemistry, materials and applications. Chem. Soc. Rev. 2022, 51, 9262–9339. [Google Scholar] [CrossRef]
  20. Jiang, S.-D.; Wang, B.-W.; Su, G.; Wang, Z.-M.; Gao, S. A Mononuclear Dysprosium Complex Featuring Single-Molecule-Magnet Behavior. Angew. Chem. Int. Ed. 2010, 49, 7448–7451. [Google Scholar] [CrossRef]
  21. Wu, J.; Jung, J.; Zhang, P.; Zhang, H.; Tang, J.; Le Guennic, B. Cis-Trans Isomerism Modulates the Magnetic Relaxation of Dysprosium Single-Molecule Magnets. Chem. Sci. 2016, 7, 3632–3639. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Wu, J.; Cador, O.; Li, X.-L.; Zhao, L.; Le Guennic, B.; Tang, J. Axial Ligand Field in D4d Coordination Symmetry: Magnetic Relaxation of Dy SMMs Perturbed by Counteranions. Inorg. Chem. 2017, 56, 11211–11219. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Chen, Y.-C.; Liu, J.-L.; Ungur, L.; Liu, J.; Li, Q.-W.; Wang, L.-F.; Ni, Z.-P.; Chibotaru, L.F.; Chen, X.-M.; Tong, M.-L. Symmetry-Supported Magnetic Blocking at 20 K in Pentagonal Bipyramidal Dy(III) Single-Ion Magnets. J. Am. Chem. Soc. 2016, 138, 2829–2837. [Google Scholar] [CrossRef] [PubMed]
  24. Liu, J.; Chen, Y.-C.; Liu, J.-L.; Vieru, V.; Ungur, L.; Jia, J.-H.; Chibotaru, L.F.; Lan, Y.; Wernsdorfer, W.; Gao, S.; et al. A Stable Pentagonal Bipyramidal Dy(III) Single-Ion Magnet with a Record Magnetization Reversal Barrier over 1000 K. J. Am. Chem. Soc. 2016, 138, 5441–5450. [Google Scholar] [CrossRef] [PubMed]
  25. Yu, K.-X.; Kragskow, J.G.C.; Ding, Y.-S.; Zhai, Y.-Q.; Reta, D.; Chilton, N.F.; Zheng, Y.-Z. Enhancing Magnetic Hysteresis in Single-Molecule Magnets by Ligand Functionalization. Chem 2020, 6, 1777–1793. [Google Scholar] [CrossRef]
  26. Chen, Y.-C.; Tong, M.-L. Single-Molecule Magnets beyond a Single Lanthanide Ion: The Art of Coupling. Chem. Sci. 2022, 13, 8716–8726. [Google Scholar] [CrossRef] [PubMed]
  27. Liu, F.; Krylov, D.S.; Spree, L.; Avdoshenko, S.M.; Samoylova, N.A.; Rosenkranz, M.; Kostanyan, A.; Greber, T.; Wolter, A.U.B.; Büchner, B.; et al. Single molecule magnet with an unpaired electron trapped between two lanthanide ions inside a fullerene. Nat. Commun. 2017, 8, 16098. [Google Scholar] [CrossRef] [Green Version]
  28. Liu, F.; Velkos, G.; Krylov, D.S.; Spree, L.; Zalibera, M.; Ray, R.; Samoylova, N.A.; Chen, C.-H.; Rosenkranz, M.; Schiemenz, S.; et al. Air-stable redox-active nanomagnets with lanthanide spins radical-bridged by a metal–metal bond. Nat. Commun. 2019, 10, 571. [Google Scholar] [CrossRef]
  29. Paschke, F.; Birk, T.; Enenkel, V.; Liu, F.; Romankov, V.; Dreiser, J.; Popov, A.A.; Fonin, M. Exceptionally High Blocking Temperature of 17 K in a Surface-Supported Molecular Magnet. Adv Mater 2021, 33, 2102844. [Google Scholar] [CrossRef]
  30. Gould, C.A.; McClain, K.R.; Reta, D.; Kragskow, J.G.C.; Marchiori, D.A.; Lachman, E.; Choi, E.-S.; Analytis, J.G.; Britt, R.D.; Chilton, N.F.; et al. Ultrahard magnetism from mixed-valence dilanthanide complexes with metal-metal bonding. Science 2022, 375, 198–202. [Google Scholar] [CrossRef]
  31. Rinehart, J.D.; Fang, M.; Evans, W.J.; Long, J.R. Strong exchange and magnetic blocking in N2 32-radical-bridged lanthanide complexes. Nat. Chem. 2011, 3, 538–542. [Google Scholar] [CrossRef] [PubMed]
  32. Demir, S.; Jeon, I.-R.; Long, J.R.; Harris, T.D. Radical ligand-containing single-molecule magnets. Coord. Chem. Rev. 2015, 289-290, 149–176. [Google Scholar] [CrossRef] [Green Version]
  33. Demir, S.; Gonzalez, M.I.; Darago, L.E.; Evans, W.J.; Long, J.R. Giant coercivity and high magnetic blocking temperatures for N23− radical-bridged dilanthanide complexes upon ligand dissociation. Nat. Commun. 2017, 8, 2144. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Wang, J.; Li, Q.-W.; Wu, S.-G.; Chen, Y.-C.; Wan, R.-C.; Huang, G.-Z.; Liu, Y.; Liu, J.-L.; Reta, D.; Giansiracusa, M.J.; et al. Opening magnetic hysteresis by axial ferromagnetic coupling: From mono-decker to double-decker metallacrown. Angew. Chem. Int. Ed. 2021, 60, 5299–5306. [Google Scholar] [CrossRef] [PubMed]
  35. Xiong, J.; Ding, H.-Y.; Meng, Y.-S.; Gao, C.; Zhang, X.-J.; Meng, Z.-S.; Zhang, Y.-Q.; Shi, W.; Wang, B.-W.; Gao, S. Hydroxide-bridged five-coordinate DyIII single-molecule magnet exhibiting the record thermal relaxation barrier of magnetization among lanthanide-only dimers. Chem. Sci. 2017, 8, 1288–1294. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Wu, J.; Demeshko, S.; Dechert, S.; Meyer, F. Macrocycle Based Dinuclear Dysprosium(III) Single Molecule Magnets with Local D5h Coordination Geometry. Dalton Trans. 2021, 50, 17573–17582. [Google Scholar] [CrossRef]
  37. Lu, J.; Li, X.-L.; Jin, C.; Yu, Y.; Tang, J. Dysprosium-based linear helicate clusters: Syntheses, structures, and magnetism. New J. Chem. 2020, 44, 994–1000. [Google Scholar] [CrossRef]
  38. Zhang, Y.; Ali, B.; Wu, J.; Guo, M.; Yu, Y.; Liu, Z.; Tang, J. Construction of Metallosupramolecular Coordination Complexes: From Lanthanide Helicates to Octahedral Cages Showing Single-Molecule Magnet Behavior. Inorg. Chem. 2019, 58, 3167–3174. [Google Scholar] [CrossRef]
  39. Wu, J.; Li, X.-L.; Zhao, L.; Guo, M.; Tang, J. Enhancement of Magnetocaloric Effect through Fixation of Carbon Dioxide: Molecular Assembly from Ln4 to Ln4 Cluster Pairs. Inorg. Chem. 2017, 56, 4104–4111. [Google Scholar] [CrossRef]
  40. Amlani, I.; Orlov, A.O.; Toth, G.; Bernstein, G.H.; Lent, C.S.; Snider, G.L. Digital Logic Gate Using Quantum-Dot Cellular Automata. Science, 1999; 284, 289–291. [Google Scholar] [CrossRef]
  41. Yang, Q.; Tang, J. Heterometallic grids: Synthetic strategies and recent advances. Dalton Trans. 2019, 48, 769–778. [Google Scholar] [CrossRef]
  42. Wu, J.; Guo, M.; Li, X.-L.; Zhao, L.; Sun, Q.-F.; Layfield, R.; Tang, J. From Double-Shelled Grids to Supramolecular Frameworks. Chem. Commun. 2018, 54, 12097–12100. [Google Scholar] [CrossRef] [PubMed]
  43. Wu, J.; Zhao, L.; Guo, M.; Tang, J. Constructing supramolecular grids: From 4f square to 3d-4f grid. Chem. Commun. 2015, 51, 17317–17320. [Google Scholar] [CrossRef] [PubMed]
  44. Wu, J.; Zhao, L.; Zhang, L.; Li, X.-L.; Guo, M.; Tang, J. Metallosupramolecular Coordination Complexes: The Design of Heterometallic 3d–4f Gridlike Structures. Inorg. Chem. 2016, 55, 5514–5519. [Google Scholar] [CrossRef] [PubMed]
  45. Wu, J.; Zhao, L.; Zhang, L.; Li, X.-L.; Guo, M.; Powell, A.K.; Tang, J. Macroscopic Hexagonal Tubes of 3 d–4 f Metallocycles. Angew. Chem. Int. Ed. 2016, 55, 15574–15578. [Google Scholar] [CrossRef] [PubMed]
  46. Wu, J.; Li, X.-L.; Guo, M.; Zhao, L.; Zhang, Y.; Tang, J. Realization of toroidal magnetic moments in heterometallic 3d-4f metallocycles. Chem. Commun. 2018, 54, 1065–1068. [Google Scholar] [CrossRef] [PubMed]
  47. Wu, J.; Liu, D.; Yang, Q.; Ge, Y.; Tang, J.; Qi, Z. Magnetic investigation in di- and tetranuclear lanthanide complexes. New J. Chem. 2021, 45, 2200–2207. [Google Scholar] [CrossRef]
  48. Alam, M.S.; Strömsdörfer, S.; Dremov, V.; Müller, P.; Kortus, J.; Ruben, M.; Lehn, J.-M. Addressing the Metal Centers of [2×2] CoII4 Grid-Type Complexes by STM/STS. Angew. Chem. Int. Ed. 2005, 44, 7896–7900. [Google Scholar] [CrossRef]
  49. Barboiu, M.; Stadler, A.-M.; Lehn, J.-M. Controlled Folding, Motional, and Constitutional Dynamic Processes of Polyheterocyclic Molecular Strands. Angew. Chem. Int. Ed. 2016, 55, 4130–4154. [Google Scholar] [CrossRef] [Green Version]
  50. Bassani, D.M.; Lehn, J.-M.; Fromm, K.; Fenske, D. Toposelective and Chiroselective Self-Assembly of [2 × 2] Grid-Type Inorganic Arrays Containing Different Octahedral Metallic Centers. Angew. Chem. Int. Ed. 1998, 37, 2364–2367. [Google Scholar] [CrossRef]
  51. Wang, Y.-T.; Li, S.-T.; Wu, S.-Q.; Cui, A.-L.; Shen, D.-Z.; Kou, H.-Z. Spin Transitions in Fe(II) Metallogrids Modulated by Substituents, Counteranions, and Solvents. J. Am. Chem. Soc. 2013, 135, 5942–5945. [Google Scholar] [CrossRef]
  52. Dhers, S.; Mondal, A.; Aguilà, D.; Ramírez, J.; Vela, S.; Dechambenoit, P.; Rouzières, M.; Nitschke, J.R.; Clérac, R.; Lehn, J.-M. Spin State Chemistry: Modulation of Ligand pKa by Spin State Switching in a [2 × 2] Iron(II) Grid-Type Complex. J. Am. Chem. Soc. 2018, 140, 8218–8227. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Li, X.-L.; Wu, J.; Zhao, L.; Shi, W.; Cheng, P.; Tang, J. End-to-end azido-pinned interlocking lanthanide squares. Chem. Commun. 2017, 53, 3026–3029. [Google Scholar] [CrossRef] [PubMed]
  54. Uppadine, L.H.; Gisselbrecht, J.-P.; Kyritsakas, N.; Nättinen, K.; Rissanen, K.; Lehn, J.-M. Mixed-Valence, Mixed-Spin-State, and Heterometallic [2 × 2] Grid-type Arrays Based on Heteroditopic Hydrazone Ligands: Synthesis and Electrochemical Features. Chem. Eur. J. 2005, 11, 2549–2565. [Google Scholar] [CrossRef] [PubMed]
  55. Join, B.; Möller, K.; Ziebart, C.; Schröder, K.; Gördes, D.; Thurow, K.; Spannenberg, A.; Junge, K.; Beller, M. Selective Iron-Catalyzed Oxidation of Benzylic and Allylic Alcohols. Adv. Synth. Catal. 2011, 353, 3023–3030. [Google Scholar] [CrossRef]
  56. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.K.; Puschmann, H. OLEX2: A complete structure solution, refinement and analysis program. J. Appl. Crystallogr. 2009, 42, 339–341. [Google Scholar] [CrossRef]
  57. Sheldrick, G. SHELXT - Integrated space-group and crystal-structure determination. Acta Crystallographica Section A 2015, 71, 3–8. [Google Scholar] [CrossRef] [Green Version]
  58. Sheldrick, G. Crystal structure refinement with SHELXL. Acta Crystallographica Section C 2015, 71, 3–8. [Google Scholar] [CrossRef] [Green Version]
  59. Boudreaux, E.A.; Mulay, L.N. Theory and Applications of Molecular Paramagnetism; John Wiley & Sons: New York, NY, USA, 1976. [Google Scholar]
  60. Casanova, D.; Alemany, P.; Bofill, J.M.; Alvarez, S. Shape and Symmetry of Heptacoordinate Transition-Metal Complexes: Structural Trends. Chem. Eur. J. 2003, 9, 1281–1295. [Google Scholar] [CrossRef]
  61. 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]
  62. Acharya, J.; Ahmed, N.; Flores-Gonzalez, J.; Kumar, P.; Pointillart, F.; Cador, O.; Singh, S.K.; Chandrasekhar, V. Slow Magnetic Relaxation in a Homo Dinuclear Dy(III) Complex in a Pentagonal Bipyamidal Geometry. Dalton Trans. 2020, 49, 13110–13122. [Google Scholar] [CrossRef]
  63. Yang, H.; Liu, S.-S.; Meng, Y.-S.; Zhang, Y.-Q.; Pu, L.; Wang, X.; Lin, S. Four mononuclear dysprosium complexes with neutral Schiff-base ligands: Syntheses, crystal structures and slow magnetic relaxation behavior. Dalton Trans. 2022, 51, 1415–1422. [Google Scholar] [CrossRef] [PubMed]
  64. Yang, Q.; Wang, G.-L.; Zhang, Y.-Q.; Tang, J. Self-assembly of fish-bone and grid-like CoII-based single-molecule magnets using dihydrazone ligands with NNN and NNO pockets. Dalton Trans. 2022, 51, 13928–13937. [Google Scholar] [CrossRef] [PubMed]
  65. Ashebr, T.G.; Li, X.-L.; Zhao, C.; Yang, Q.; Tang, J. Bis-pyrazolone-based dysprosium(iii) complexes: Zero-field single-molecule magnet behavior in the [2 × 2] grid DyIII4 cluster. CrystEngComm 2022, 24, 6688–6695. [Google Scholar] [CrossRef]
  66. Reta, D.; Chilton, N.F. Uncertainty estimates for magnetic relaxation times and magnetic relaxation parameters. PCCP 2019, 21, 23567–23575. [Google Scholar] [CrossRef] [PubMed]
  67. Debye, P. Einige Bemerkungen zur Magnetisierung bei tiefer Temperatur. Ann. Phys. 1926, 386, 1154–1160. [Google Scholar] [CrossRef]
  68. Demir, S.; Zadrozny, J.M.; Long, J.R. Large Spin-Relaxation Barriers for the Low-Symmetry Organolanthanide Complexes [Cp*2Ln(BPh4)] (Cp*=pentamethylcyclopentadienyl; Ln=Tb, Dy). Chem. Eur. J. 2014, 20, 9524–9529. [Google Scholar] [CrossRef]
  69. Chilton, N.F.; Collison, D.; McInnes, E.J.L.; Winpenny, R.E.P.; Soncini, A. An electrostatic model for the determination of magnetic anisotropy in dysprosium complexes. Nat. Commun. 2013, 4, 2551. [Google Scholar] [CrossRef] [Green Version]
Scheme 1. Schematic drawing of the assemblies of complexes Dy2 and Dy4 from the dicompartmental Schiff base ligand.
Scheme 1. Schematic drawing of the assemblies of complexes Dy2 and Dy4 from the dicompartmental Schiff base ligand.
Magnetochemistry 09 00004 sch001
Figure 1. Structures of complexes Dy2 (a) and Dy4 (b). The cations and solvents have been omitted for clarity.
Figure 1. Structures of complexes Dy2 (a) and Dy4 (b). The cations and solvents have been omitted for clarity.
Magnetochemistry 09 00004 g001
Figure 2. Temperature-dependent χMT products at 1 kOe between 2 and 300 K for Dy2 and Dy4.
Figure 2. Temperature-dependent χMT products at 1 kOe between 2 and 300 K for Dy2 and Dy4.
Magnetochemistry 09 00004 g002
Figure 3. Frequency-dependent ac susceptibility of Dy2 (a) and Dy4 (b) under indicated dc field and temperature. The dots and solid lines are the data from ac measurements.
Figure 3. Frequency-dependent ac susceptibility of Dy2 (a) and Dy4 (b) under indicated dc field and temperature. The dots and solid lines are the data from ac measurements.
Magnetochemistry 09 00004 g003
Figure 4. Cole-Cole plots for Dy2 under 1000 Oe (a) and Dy4 under 500 Oe (b) at the indicated temperature range. The solid lines represent the best fit.
Figure 4. Cole-Cole plots for Dy2 under 1000 Oe (a) and Dy4 under 500 Oe (b) at the indicated temperature range. The solid lines represent the best fit.
Magnetochemistry 09 00004 g004
Figure 5. Orientations of the main magnetic axes of the ground state of Dy2 (a) and Dy4 (b) calculated based on the molecule structure.
Figure 5. Orientations of the main magnetic axes of the ground state of Dy2 (a) and Dy4 (b) calculated based on the molecule structure.
Magnetochemistry 09 00004 g005
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

Li, J.; Zhang, F.; Guo, X.; Liu, D.; Wu, J. The Assembly of Grid-type Lanthanide Cluster. Magnetochemistry 2023, 9, 4. https://doi.org/10.3390/magnetochemistry9010004

AMA Style

Li J, Zhang F, Guo X, Liu D, Wu J. The Assembly of Grid-type Lanthanide Cluster. Magnetochemistry. 2023; 9(1):4. https://doi.org/10.3390/magnetochemistry9010004

Chicago/Turabian Style

Li, Jinsong, Fan Zhang, Xuefeng Guo, Dan Liu, and Jianfeng Wu. 2023. "The Assembly of Grid-type Lanthanide Cluster" Magnetochemistry 9, no. 1: 4. https://doi.org/10.3390/magnetochemistry9010004

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