Next Article in Journal
Solid-State Heating Using the Multicaloric Effect in Multiferroics
Previous Article in Journal
Perspectives on Molecular Materials—A Tribute to Professor Peter Day
Previous Article in Special Issue
Synthesis, Crystal Structure and Magnetic Properties of 1D Chain Complexes Based on Azo Carboxylate Oxime Ligand
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Syntheses, Structures and Magnetic Properties of M2 (M = Fe, Co) Complexes with N6 Coordination Environment: Field-Induced Slow Magnetic Relaxation in Co2

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
*
Author to whom correspondence should be addressed.
Magnetochemistry 2021, 7(12), 153; https://doi.org/10.3390/magnetochemistry7120153
Submission received: 29 October 2021 / Revised: 12 November 2021 / Accepted: 19 November 2021 / Published: 23 November 2021
(This article belongs to the Special Issue Magnetic Properties of Metal Complexes)

Abstract

:
Two dinuclear complexes [M2(H2L)2](ClO4)4·2MeCN (M = Co for Co2 and Fe for Fe2) were synthesized using a symmetric hydrazone ligand with the metal ions in an N6 coordination environment. The crystal structures and magnetic properties were determined by single-crystal X-ray diffraction and magnetic susceptibility measurements. The crystal structure study revealed that the spin centers were all in the high-spin state with a distorted octahedron (Oh) geometry. Dynamic magnetic properties measurements revealed that complex Co2 exhibited field-induced single-molecule magnet properties with two-step relaxation in which the fast relaxation path was from QTM and the slow relaxation path from the thermal relaxation under an applied field.

1. Introduction

Since the discovery of the first single-molecule magnet (SMM) Mn12 cluster [1], research on SMMs has aroused interest among the scientific community due to the fact of their potential applications such as high-density information storage, quantum computing, and spintronics [2,3,4,5,6,7,8,9,10,11,12]. Subsequently, a large number of SMMs based on transition mental [13,14,15,16,17], lanthanide [18,19,20,21,22,23,24,25], and mixed-metal ions [26,27,28,29,30,31] have been designed and reported. For SMMs, the combination of a large negative zero-field splitting parameter (D < 0) or strong uniaxial magnetic anisotropy (gz) and large spin ground state (S) could result in an SMM with a large effective energy barrier (Ueff) [32]. The energy barrier can be calculated by Ueff = |D|S2 or |D|(S2 − 1/4) for integer and half-integer spin, respectively [33].
For transition metal complexes, crystal field splitting is much stronger than spin-orbit coupling; therefore, the orbital angular momentum is almost completely quenched [34]. However, the zero-field splitting parameter is usually only several dozens of wavenumbers for transition metals, and most of the complexes based on 3d ions are field-induced single-molecule magnets, especially in the presence of quantum tunneling of magnetization (QTM) [35,36,37]. The divalent cobalt ion is in the d7 configuration, and there is an important orbital contribution to the magnetization. The coordination number of CoII ions ranges from two coordination to eight coordination with different coordination geometries [38,39,40,41,42,43,44]. As a result, the CoII ion is a good candidate for the construction of SMMs due to the fact of its magnetic anisotropy. Up to now, numerous SMMs based on CoII ions have been reported [45,46,47,48].
By reducing the coordination number [49] or regulating the coordinated atoms [50], it is possible to obtain a relatively weak ligand field and increase the orbital contribution. However, low-coordination molecular magnets are usually unstable in air. Therefore, the design of high-coordination and stable single-molecule magnets has attracted much attention [41,48]. Herein, we designed and synthesized a symmetrical hydrazone ligand 2,6-bis((E)-(2-(6-(1H-pyrazol-1-yl)pyridin-2-yl)hydrazineylidene)methyl)pyridine (H2L) in which only nitrogen acts as donor atoms to coordinate with metal ions. Two new dinuclear complexes [M2(H2L)2](ClO4)4·2MeCN (M = Co for Co2 and Fe for Fe2) were obtained from the ligand. The crystal structures and magnetic properties were determined by single-crystal X-ray diffraction and magnetic susceptibility measurements. The spin centers were all in the high-spin state with a distorted octahedron (Oh) geometry. Importantly, dynamic magnetic properties measurements revealed that complex Co2 exhibited field-induced single-molecule magnet properties with two-step relaxation in which the fast relaxation (FR) path was from the QTM and the slow relaxation (SR) path from the thermal relaxation.

2. Materials and Methods

2.1. Syntheses of Ligand and Complexes

All chemicals and solvents were obtained commercially without further purification. 2-Hydrazineyl-6-(1H-pyrazol-1-yl)pyridine and pyridine-2,6-dicarbaldehyde were synthesized according to the previously reported literature [51,52]. As a caution, the salts Co(ClO4)2·6H2O and Fe(ClO4)2·6H2O should be used carefully, as they are potentially explosive.

2.1.1. Synthesis of Ligand H2L

2-Hydrazineyl-6-(1H-pyrazol-1-yl)pyridine (1.75 g, 10 mmol) was added into a solution of pyridine-2,6-dicarbaldehyde (0.67 g, 5 mmol) in ethanol (50 mL). Then, the reaction mixture was allowed to reflux overnight giving a dark orange solid. After filtering, the product was washed with ice ethanol and dried in a vacuum, yielding the ligand 2,6-bis((E)-(2-(6-(1H-pyrazol-1-yl)pyridin-2-yl)hydrazineylidene)methyl)pyridine (H2L) (1.78 g, 79.4%). 1H NMR (400 MHz, DMSO-d6): δ = 11.41 (s, 2H), 8.51–8.49 (m, 2H), 8.14 (s, 2H), 7.96–7.93 (m, 2H), 7.89–7.84 (m, 3H), 7.81 (d, J = 1.0, 2H), 7.35 (d, J = 7.5, 2H), 7.26 (d, J = 8.1, 2H), and 6.58 (dd, J = 2.5, 1.7, 2H). Selected IR (solid, ATR) (cm−1): 543 (m), 566 (w), 578 (w), 588 (w), 605 (m), 644 (m), 667 (w), 688 (w), 715 (w), 728 (m), 738 (m), 769 (s), 790 (s), 811 (w), 863 (w), 883 (w), 902 (m), 916 (m), 964 (m), 989 (w), 1047 (m), 1072 (w), 1087 (w), 1130 (m), 1141 (m), 1176 (m), 1253 (m), 1270 (w), 1284 (w), 1321 (m), 1336 (w), 1346 (w), 1392 (m), 1427 (m), 1452 (s), 1508 (s), 1575 (s), 1606 (s), 2954 (w), 3089 (w), and 3317 (w).

2.1.2. Synthesis of Co2

A mixture of H2L (0.1 mmol) and Co(ClO4)2·6H2O (0.1 mmol) in acetonitrile (15 mL) was stirred for one hour, yielding a dark red clear solution after filtering. Subsequently, the filtrate was allowed to stand and evaporate for four days to obtain red crystals suitable for single-crystal X-ray diffraction. Yield: 15.99 mg, (21.36%, based on metal salts). Elemental analysis calculated for C50H44Cl4Co2N24O16 (%): C, 40.12; H, 2.96; N, 22.46. Found (%): C, 40.08; H, 2.91; N, 22.42.

2.1.3. Synthesis of Fe2

A mixture of H2L (0.1 mmol) and Fe(ClO4)2·6H2O (0.1 mmol) in acetonitrile (15 mL) was stirred for one hour, yielding a dark red clear solution after filtering. Subsequently, the filtrate was allowed to stand and evaporate for four days to obtain red crystals suitable for single-crystal X-ray diffraction. Yield: 13.82 mg, (18.54%, based on metal salts). Elemental analysis calculated for C50H44Cl4Fe2N24O16 (%): C, 40.29; H, 2.98; N, 22.55. Found (%): C, 40.23; H, 2.91; N, 22.53.

2.2. Physical Measurements

The 1H NMR spectrum of H2L was recorded on a Bruker Avance 400 MHz spectrometer (Bruker, Switzerland, Figure S1). Elemental analyses (i.e., C, H, and N) were measured on a PerkinElmer 2400 analyzer (PerkinElmer, United States). Fourier transform infrared spectrometer (FTIR) spectra were obtained using a Nicolet 6700 Flex FTIR spectrometer (Thermo Fisher, United States) equipped with a smart iTR attenuated total reflectance (ATR) sampling accessory (Figures S2 and S3). Powder X-ray diffraction (XRD) measurements were carried out using a Bruker D8 advance X-ray diffractometer (Bruker AXS GMBH, Germany) with Cu-Kα radiation.

2.2.1. Crystallography

Single-crystal X-ray diffraction data were collected by the Bruker D8 venture CCD diffractometer (Bruker AXS GMBH, Germany) using graphite-monochromatized Mo-Kα radiation (λ = 0.71073 Å). In the Olex2 package, the structures were solved using SHELXT [53] (direct methods), and all non-hydrogen atoms were refined using SHELXL [54] (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 crystallographic data for Co2 and Fe2 are listed in Table S1. CCDC 2116748 and 2116749 contain the supplementary crystallographic data for this paper.

2.2.2. Magnetic Measurements

Magnetic measurements were measured 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 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 [55] for all 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 0 and 3500 Oe dc fields.

3. Results and Discussions

3.1. Structures of Co2 and Fe2

The crystal structures of Co2 and Fe2 were determined by single-crystal X-ray diffraction at 173 and 180 K, respectively. The two complexes were isostructural; therefore, the structure of Co2 is described here only. Co2 crystallized in the triclinic space group P 1 ¯ with the crystallographic data and refinement details shown in Table S1. The asymmetric unit of the complex consisted of one neutral ligand H2L, one crystallographically independent CoII center, two ClO4 anions, and one MeCN solvent molecule in the lattice. The spin centers were in the N6 coordination pocket from the two H2L ligands, forming a Co2 core (Figure 1 and Figure S5). Four perchloride anions crystallized in the crystal lattice to balance the positive charges.
The Co–N bond distances were in the range of 2.05–2.25 Å, which indicates that the CoII ions were in the high spin (HS) state (Table S2). The coordination geometry of the CoII center was evaluated by the SHAPE software [56,57] (Figure S4 and Table S3). The coordination geometry of the CoII ion was closest to an octahedron (Oh) with a CShM value of 5.37. Such a large value suggests a large distortion of the coordination geometry. For Fe2, the Fe–N bond distances were in the range of 2.11–2.27 Å, which also indicates that the FeII ions were in the HS state. The coordination geometry of the FeII ion was closest to an octahedron (Oh) with a CShM value of 6.34, suggesting a larger distortion of the coordination geometry than the CoII ion in Co2 (Figure S4 and Table S3). Due to the long coordination bonds, complex Fe2 did not exhibit spin-crossover properties [58]. As depicted in Figure S5 and S6, the intramolecular distance and the shortest intermolecular distance between two CoII ions were 7.45 and 9.01 Å, respectively. The intermolecular interaction could be ignored because of the relatively long intermolecular distance (9.01 Å) [59], while dinuclear cobalt complexes with shorter Co···Co distances (3.10–3.11 Å) usually show strong ferromagnetic interaction with the coupling parameter regulated by changing the ligand field of one CoII center as well as the effects on the dc magnetic susceptibility [60]. The intramolecular CoII···CoII ions distance of 7.45 Å probably induced very weak intramolecular interaction and, thus, played a role in the magnetic properties. The intramolecular and the shortest intermolecular Fe···Fe distances were 7.47 and 9.05 Å for Fe2, respectively. As shown in Figure S7, the phase purity of the bulk samples of the two complexes was confirmed by powder XRD analyses.

3.2. Magnetic Properties of Co2 and Fe2

3.2.1. Static Magnetic Properties of Co2 and Fe2

Direct current magnetic susceptibility measurements were measured on polycrystalline samples with a 1000 Oe field in the temperature range of 2–300 K (Figure 2). At room temperature, the χMT values (χM is molar magnetic susceptibility) were 4.99 and 7.23 cm3Kmol1 for Co2 and Fe2, respectively. These values were higher than the expected values for the two spin-only HS ions (CoII, S = 3/2, χMT = 1.875 cm3Kmol−1; FeII, S = 2, χMT = 3 cm3Kmol−1), which may be due to the existence of orbital contribution. In low-temperature regions, the χMT values decreased gradually for Co2 and sharply for Fe2, which probably resulted from the magnetic interaction or zero-field splitting of the spin center. The temperature-dependent magnetic susceptibility plots were approximately simulated by using a spin Hamiltonian (Equation (1)) [61,62]:
H ^ = 2 J S ^ 1 S ^ 2 + { D i [ S Z , i 2 S i ( S i + 1 ) / 3 ] + E [ S x , i 2 S y , i 2 ] } + g μ B B S i
where J, D, E, g, μB, and B correspond to magnetic exchange, the axial and rhombic zero-field splitting parameter, Landé factor, Bohr magneton, and magnetic field vector, respectively. The best fit provided large |D| values of 30 cm−1 and rhombic a parameter E (E/D = 0.15) for Co2, indicating the presence of magnetic anisotropy, which probably relates to the large distortion of the coordination geometry. The g values were 2.32, revealing the presence of orbital contribution. It is worth noting that the sign of D cannot be determined by the simulation of temperature-dependent magnetic susceptibility plots but can be further determined by theoretical calculation or EPR measurements [63,64,65]. In addition, not only CoII complexes with negative D values can exhibit single-molecule magnet properties [50,62,66,67], but some cobalt complexes with positive D values can also act as SMMs [68,69,70,71]. The two CoII ions were ferromagnetically coupled, with J = 0.08 cm−1. In contrast, the D values of Fe2 were smaller (|D| = 1.0 cm−1, E/D = 0.2, g = 2.2), suggesting the possible absence of SMM properties, which is probably because of the difference in the electron structure. Moreover, the magnetic interaction parameter (J = −0.25 cm−1) was different from Co2, indicating the presence of weak antiferromagnetic interaction.
The field-dependent magnetization for Co2 was performed in the range of field 0−7 T at 1.9, 3.0, and 5.0 K. The magnetization value did not reach saturation value (6.96 μB) at 7 T. The plots of M vs. H/T at various temperatures were non-superimposable, suggesting the presence of magnetic anisotropy (Figure 3).

3.2.2. Dynamic Magnetic Properties of Co2

To investigate the dynamic magnetic properties, ac susceptibility measurements were performed on Co2 under various applied dc fields. No out-of-phase (χ″) susceptibilities signals were observed under a zero dc field (Figure S8), possibly due to the presence of QTM. We then measured field-dependent ac susceptibility at 1.9 K to determine the optimal dc field. The peak of out-of-phase (χ″) susceptibilities appealed at 3500 Oe (Figure S9); thus, temperature-dependent and frequency-dependent ac susceptibility were then carried out under this dc field (Figure 4). In the temperature-dependent out-of-phase (χ″) plots, the maximum appeared up to 6.0 K at a frequency of 1488 Hz, suggesting the slow relaxation of magnetization. In the low-temperature region, an upturning appeared, which is probably ascribed to QTM. The peaks of frequency-dependent out-of-phase (χ″) susceptibilities shifted to high frequency when increasing the temperature, suggesting the typical field-induced single-molecule magnet properties.
From the frequency-dependent ac susceptibility data in the temperature region of 1.9–8.0 K, Cole–Cole plots were represented as χ″ vs. χ′ and fitted with the double-relaxation Debye model [72] and CC-FIT program [73] (Figure 5). The Cole–Cole plots show two semi-circular profiles, suggesting the presence of two-step relaxation. Extracting the relaxation time from the frequency-dependent susceptibility provided two relaxation regimes (Table S4). The fast relaxation (FR) path showed typical thermal relaxation, while the slow relaxation (SR) path was almost temperature-independent, which resulted from QTM (Figure 6). To analyze the relaxation procedure, the plots of τ vs. T−1 were fitted using the following equation [74,75,76]:
1 τ = 1 τ Q T M + A T + C T n + τ 0 1 exp ( U e f f / T )
where 1/τQTM, AT, CTn, and τ0−1exp(–Ueff/T) correspond to quantum tunneling, direct, Raman, and Orbach relaxation processes [77,78], respectively. The best fit of the FR path included direct, Raman, and Orbach relaxation processes with Ueff = 43 K, τ0 = 1.07 × 10−7 s, A = 432, C = 1.72 × 10−12 s−1·K−n, n = 3. The quantum tunneling relaxation time of the SR path was linearly fitted for the τ versus 1/T plot, giving τQTM = 0.14 s.
The effective barrier obtained from the fitting above (approximately 43 K) for the FR path was much lower than the energy level 2|D| = 60 cm−1. Therefore, we ignored the Orbach relaxation process and fitted the FR path with Equation (3) (where 1/τQTM, AT, and CTn correspond to quantum tunneling, direct, and Raman relaxation processes) for the whole temperature region (Figure 7). The best fit gives τQTM = 1.48 × 10−3 s, A = 124, C = 0.106 s−1·K−n, and n = 6.3.
1 τ = 1 τ Q T M + A T + C T n

4. Conclusions

In conclusion, we successfully designed and synthesized two dinuclear complexes, Co2 and Fe2, using a symmetric hydrazone ligand with the metal ions in an N6 coordination environment. The crystal structures of the two complexes revealed that the spin centers were in the N6 coordinated environment with a distorted octahedron (Oh) geometry. The analysis of the crystal structures and the dc susceptibility measurements indicated that the spin centers were in the high-spin state. Dynamic magnetic properties measurements revealed that complex Co2 exhibited field-induced single-molecule magnet properties due to the magnetic anisotropy of CoII ions. The complex exhibited two-step relaxation with the FR and SR paths resulting from QTM and thermal relaxation, respectively. This work may open a new opportunity for the design of dinuclear 3d-SMMs based on octahedron coordination geometry.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/magnetochemistry7120153/s1, Figure S1: 1H-NMR spectrum of H2L in DMSO-d6 at room temperature; Figure S2: IR spectrum of H2L; Figure S3: IR spectra of the complexes Co2 (blue curve) and Fe2 (red curve); Figure S4: Coordination polyhedrons of CoII (left) and FeII (right) in the complexes Co2 and Fe2; Figure S5: Packing model along the a- and c-axes of complex Co2; Figure S6: Packing model along the a- and c-axes of complex Fe2; Figure S7: Powder XRD analyses of the complexes Co2 and Fe2; Figure S8: Temperature-dependent ac susceptibility of Co2 under a 0 Oe dc field; Figure S9: Field-dependent ac susceptibility of Co2 at 1.9 K with an ac frequency of 997 Hz; Table S1: Crystallographic data for Co2 and Fe2; Table S2: Selected bond distances (Å) and angles (º) of Co2 and Fe2; Table S3: Short contacts and interatomic distances (Å) for 1·ClO4 and 1·BF4; Table S3: The CShM values calculated by SHAPE 2.1 of CoII and FeII ions in Co2 and Fe2; Table S4: Parameters for the best fit of frequency-dependent ac susceptibility of Co2 under 3500 Oe dc field.

Author Contributions

Methodology, software, data curation and writing—original draft preparation, Q.Y.; writing—review and editing, X.-L.L. and J.T.; project administration and funding acquisition, J.T. All authors have read and agreed to the published version of the manuscript.

Funding

National Natural Science Foundation of China.

Data Availability Statement

The data are available by corresponding authors.

Acknowledgments

We thank the National Natural Science Foundation of China (21871247), the Key Research Program of Frontier Sciences, and CAS (ZDBS-LY-SLH023) for the financial support. J.T. gratefully acknowledges support from the Royal Society’s Newton Advanced Fellowship (NA160075).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sessoli, R.; Gatteschi, D.; Caneschi, A.; Novak, M.A. Magnetic bistability in a metal-ion cluster. Nature 1993, 365, 141–143. [Google Scholar] [CrossRef]
  2. Gatteschi, D. Molecular Magnetism: A Basis for New Materials. Adv. Mater. 1994, 6, 635–645. [Google Scholar] [CrossRef]
  3. Bogani, L.; Wernsdorfer, W. Molecular spintronics using single-molecule magnets. Nat. Mater. 2008, 7, 179–186. [Google Scholar] [CrossRef] [PubMed]
  4. Dei, A.; Gatteschi, D. Molecular (Nano) Magnets as Test Grounds of Quantum Mechanics. Angew. Chem. Int. Ed. 2011, 50, 11852–11858. [Google Scholar] [CrossRef]
  5. Woodruff, D.N.; Winpenny, R.E.P.; Layfield, R.A. Lanthanide Single-Molecule Magnets. Chem. Rev. 2013, 113, 5110–5148. [Google Scholar] [CrossRef] [PubMed]
  6. Zhang, P.; Guo, Y.-N.; Tang, J. Recent advances in dysprosium-based single molecule magnets: Structural overview and synthetic strategies. Coord. Chem. Rev. 2013, 257, 1728–1763. [Google Scholar] [CrossRef]
  7. Moreno-Pineda, E.; Godfrin, C.; Balestro, F.; Wernsdorfer, W.; Ruben, M. Molecular spin qudits for quantum algorithms. Chem. Soc. Rev. 2018, 47, 501–513. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Zhu, Z.; Guo, M.; Li, X.-L.; Tang, J. Molecular magnetism of lanthanide: Advances and perspectives. Coord. Chem. Rev. 2019, 378, 350–364. [Google Scholar] [CrossRef]
  9. Mannini, M.; Pineider, F.; Sainctavit, P.; Danieli, C.; Otero, E.; Sciancalepore, C.; Talarico, A.M.; Arrio, M.A.; Cornia, A.; Gatteschi, D.; et al. Magnetic memory of a single-molecule quantum magnet wired to a gold surface. Nat. Mater. 2009, 8, 194–197. [Google Scholar] [CrossRef]
  10. Atzori, M.; Sessoli, R. The Second Quantum Revolution: Role and Challenges of Molecular Chemistry. J. Am. Chem. Soc. 2019, 141, 11339–11352. [Google Scholar] [CrossRef]
  11. Fittipaldi, M.; Cini, A.; Annino, G.; Vindigni, A.; Caneschi, A.; Sessoli, R. Electric field modulation of magnetic exchange in molecular helices. Nat. Mater. 2019, 18, 329–334. [Google Scholar] [CrossRef] [PubMed]
  12. Serrano, G.; Poggini, L.; Briganti, M.; Sorrentino, A.L.; Cucinotta, G.; Malavolti, L.; Cortigiani, B.; Otero, E.; Sainctavit, P.; Loth, S.; et al. Quantum dynamics of a single molecule magnet on superconducting Pb(111). Nat. Mater. 2020, 19, 546–551. [Google Scholar] [CrossRef] [PubMed]
  13. 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] [PubMed]
  14. Barra, A.-L.; Debrunner, P.; Gatteschi, D.; Schulz, C.E.; Sessoli, R. Superparamagnetic-like behavior in an octanuclear iron cluster. Europhys. Lett. 1996, 35, 133–138. [Google Scholar] [CrossRef]
  15. Castro, S.L.; Sun, Z.; Grant, C.M.; Bollinger, J.C.; Hendrickson, D.N.; Christou, G. Single-Molecule Magnets:  Tetranuclear Vanadium(III) Complexes with a Butterfly Structure and an S = 3 Ground State. J. Am. Chem. Soc. 1998, 120, 2365–2375. [Google Scholar] [CrossRef]
  16. Caneschi, A.; Gatteschi, D.; Lalioti, N.; Sangregorio, C.; Sessoli, R.; Venturi, G.; Vindigni, A.; Rettori, A.; Pini, M.G.; Novak, M.A. Cobalt(II)-Nitronyl Nitroxide Chains as Molecular Magnetic Nanowires. Angew. Chem. Int. Ed. 2001, 40, 1760–1763. [Google Scholar] [CrossRef]
  17. Chakarawet, K.; Harris, T.D.; Long, J.R. Semiquinone radical-bridged M2 (M = Fe, Co, Ni) complexes with strong magnetic exchange giving rise to slow magnetic relaxation. Chem. Sci. 2020, 11, 8196–8203. [Google Scholar] [CrossRef]
  18. 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]
  19. Tang, J.; Hewitt, I.; Madhu, N.T.; Chastanet, G.; Wernsdorfer, W.; Anson, C.E.; Benelli, C.; Sessoli, R.; Powell, A.K. Dysprosium Triangles Showing Single-Molecule Magnet Behavior of Thermally Excited Spin States. Angew. Chem. Int. Ed. 2006, 45, 1729–1733. [Google Scholar] [CrossRef]
  20. AlDamen, M.A.; Clemente-Juan, J.M.; Coronado, E.; Martí-Gastaldo, C.; Gaita-Ariño, A. Mononuclear Lanthanide Single-Molecule Magnets Based on Polyoxometalates. J. Am. Chem. Soc. 2008, 130, 8874–8875. [Google Scholar] [CrossRef] [PubMed]
  21. Rinehart, J.D.; Fang, M.; Evans, W.J.; Long, J.R. Strong exchange and magnetic blocking in N23- radical-bridged lanthanide complexes. Nat. Chem. 2011, 3, 538–542. [Google Scholar] [CrossRef]
  22. Guo, Y.-N.; Xu, G.-F.; Wernsdorfer, W.; Ungur, L.; Guo, Y.; Tang, J.; Zhang, H.-J.; Chibotaru, L.F.; Powell, A.K. Strong Axiality and Ising Exchange Interaction Suppress Zero-Field Tunneling of Magnetization of an Asymmetric Dy2 Single-Molecule Magnet. J. Am. Chem. Soc. 2011, 133, 11948–11951. [Google Scholar] [CrossRef] [PubMed]
  23. Wang, Y.-X.; Ma, Y.; Chai, Y.; Shi, W.; Sun, Y.; Cheng, P. Observation of Magnetodielectric Effect in a Dysprosium-Based Single-Molecule Magnet. J. Am. Chem. Soc. 2018, 140, 7795–7798. [Google Scholar] [CrossRef] [PubMed]
  24. 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] [Green Version]
  25. Briganti, M.; Garcia, G.F.; Jung, J.; Sessoli, R.; Le Guennic, B.; Totti, F. Covalency and magnetic anisotropy in lanthanide single molecule magnets: The DyDOTA archetype. Chem. Sci. 2019, 10, 7233–7245. [Google Scholar] [CrossRef]
  26. Osa, S.; Kido, T.; Matsumoto, N.; Re, N.; Pochaba, A.; Mrozinski, J. A Tetranuclear 3d−4f Single Molecule Magnet:  [CuIILTbIII(hfac)2]2. J. Am. Chem. Soc. 2004, 126, 420–421. [Google Scholar] [CrossRef]
  27. Zaleski, C.M.; Depperman, E.C.; Kampf, J.W.; Kirk, M.L.; Pecoraro, V.L. Synthesis, Structure, and Magnetic Properties of a Large Lanthanide–Transition-Metal Single-Molecule Magnet. Angew. Chem. Int. Ed. 2004, 43, 3912–3914. [Google Scholar] [CrossRef]
  28. Mereacre, V.M.; Ako, A.M.; Clérac, R.; Wernsdorfer, W.; Filoti, G.; Bartolomé, J.; Anson, C.E.; Powell, A.K. A Bell-Shaped Mn11Gd2 Single-Molecule Magnet. J. Am. Chem. Soc. 2007, 129, 9248–9249. [Google Scholar] [CrossRef]
  29. Kong, X.-J.; Ren, Y.-P.; Chen, W.-X.; Long, L.-S.; Zheng, Z.; Huang, R.-B.; Zheng, L.-S. A Four-Shell, Nesting Doll-like 3d–4f Cluster Containing 108 Metal Ions. Angew. Chem. Int. Ed. 2008, 47, 2398–2401. [Google Scholar] [CrossRef]
  30. Pugh, T.; Chilton, N.F.; Layfield, R.A. A Low-Symmetry Dysprosium Metallocene Single-Molecule Magnet with a High Anisotropy Barrier. Angew. Chem. Int. Ed. 2016, 55, 11082–11085. [Google Scholar] [CrossRef] [PubMed]
  31. 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]
  32. 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]
  33. Gatteschi, D.; Sessoli, R.; Villain, J. Molecular Nanomagnets; Oxford University Press: New York, NY, USA, 2006. [Google Scholar]
  34. Benelli, C.; Gatteschi, D. Magnetism of Lanthanides in Molecular Materials with Transition-Metal Ions and Organic Radicals. Chem. Rev. 2002, 102, 2369–2388. [Google Scholar] [CrossRef] [PubMed]
  35. Habib, F.; Luca, O.R.; Vieru, V.; Shiddiq, M.; Korobkov, I.; Gorelsky, S.I.; Takase, M.K.; Chibotaru, L.F.; Hill, S.; Crabtree, R.H.; et al. Influence of the Ligand Field on Slow Magnetization Relaxation versus Spin Crossover in Mononuclear Cobalt Complexes. Angew. Chem. Int. Ed. 2013, 52, 11290–11293. [Google Scholar] [CrossRef]
  36. Fortier, S.; Le Roy, J.J.; Chen, C.-H.; Vieru, V.; Murugesu, M.; Chibotaru, L.F.; Mindiola, D.J.; Caulton, K.G. A Dinuclear Cobalt Complex Featuring Unprecedented Anodic and Cathodic Redox Switches for Single-Molecule Magnet Activity. J. Am. Chem. Soc. 2013, 135, 14670–14678. [Google Scholar] [CrossRef] [PubMed]
  37. Wernsdorfer, W.; Sessoli, R. Quantum Phase Interference and Parity Effects in Magnetic Molecular Clusters. Science 1999, 284, 133–135. [Google Scholar] [CrossRef] [Green Version]
  38. Murrie, M. Cobalt(ii) single-molecule magnets. Chem. Soc. Rev. 2010, 39, 1986–1995. [Google Scholar] [CrossRef]
  39. Saber, M.R.; Dunbar, K.R. Ligands effects on the magnetic anisotropy of tetrahedral cobalt complexes. Chem. Commun. 2014, 50, 12266–12269. [Google Scholar] [CrossRef]
  40. Ruamps, R.; Batchelor, L.J.; Guillot, R.; Zakhia, G.; Barra, A.-L.; Wernsdorfer, W.; Guihéry, N.; Mallah, T. Ising-type magnetic anisotropy and single molecule magnet behaviour in mononuclear trigonal bipyramidal Co(II) complexes. Chem. Sci. 2014, 5, 3418–3424. [Google Scholar] [CrossRef]
  41. Huang, X.-C.; Zhou, C.; Shao, D.; Wang, X.-Y. Field-Induced Slow Magnetic Relaxation in Cobalt(II) Compounds with Pentagonal Bipyramid Geometry. Inorg. Chem. 2014, 53, 12671–12673. [Google Scholar] [CrossRef]
  42. Meng, Y.-S.; Mo, Z.; Wang, B.-W.; Zhang, Y.-Q.; Deng, L.; Gao, S. Observation of the single-ion magnet behavior of d8 ions on two-coordinate Co(I)-NHC complexes. Chem. Sci. 2015, 6, 7156–7162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Deng, Y.-F.; Han, T.; Yin, B.; Zheng, Y.-Z. On Balancing the QTM and the Direct Relaxation Processes in Single-Ion Magnets – the importance of Symmetry Control. Inorg. Chem. Front. 2017, 4, 1141–1148. [Google Scholar] [CrossRef]
  44. Paul, A.; Viciano-Chumillas, M.; Puschmann, H.; Cano, J.; Manna, S.C. Field-induced slow magnetic relaxation in mixed valence di- and tri-nuclear Co(II)-Co(III) complexes. Dalton Trans. 2020, 49, 9516–9528. [Google Scholar] [CrossRef] [PubMed]
  45. Shiga, T.; Matsumoto, T.; Noguchi, M.; Onuki, T.; Hoshino, N.; Newton, G.N.; Nakano, M.; Oshio, H. Cobalt Antiferromagnetic Ring and Grid Single-Molecule Magnet. Chem. Asian J. 2009, 4, 1660–1663. [Google Scholar] [CrossRef] [PubMed]
  46. Yang, F.; Zhou, Q.; Zhang, Y.; Zeng, G.; Li, G.; Shi, Z.; Wang, B.; Feng, S. Inspiration from old molecules: Field-induced slow magnetic relaxation in three air-stable tetrahedral cobalt(II) compounds. Chem. Commun. 2013, 49, 5289–5291. [Google Scholar] [CrossRef]
  47. Moilanen, J.O.; Chilton, N.F.; Day, B.M.; Pugh, T.; Layfield, R.A. Strong Exchange Coupling in a Trimetallic Radical-Bridged Cobalt(II)-Hexaazatrinaphthalene Complex. Angew. Chem. Int. Ed. 2016, 55, 5521–5525. [Google Scholar] [CrossRef] [Green Version]
  48. Diego, R.; Pavlov, A.; Darawsheh, M.; Aleshin, D.; Nehrkorn, J.; Nelyubina, Y.; Roubeau, O.; Novikov, V.; Aromí, G. Coordination [CoII2] and [CoIIZnII] Helicates Showing Slow Magnetic Relaxation. Inorg. Chem. 2019, 58, 9562–9566. [Google Scholar] [CrossRef]
  49. Bunting, P.C.; Atanasov, M.; Damgaard-Møller, E.; Perfetti, M.; Crassee, I.; Orlita, M.; Overgaard, J.; van Slageren, J.; Neese, F.; Long, J.R. A linear cobalt(II) complex with maximal orbital angular momentum from a non-Aufbau ground state. Science 2018, 362, eaat7319. [Google Scholar] [CrossRef]
  50. Novikov, V.V.; Pavlov, A.A.; Nelyubina, Y.V.; Boulon, M.-E.; Varzatskii, O.A.; Voloshin, Y.Z.; Winpenny, R.E.P. A Trigonal Prismatic Mononuclear Cobalt(II) Complex Showing Single-Molecule Magnet Behavior. J. Am. Chem. Soc. 2015, 137, 9792–9795. [Google Scholar] [CrossRef]
  51. Du, W.; Wang, Q.; Wang, L.; Yu, Z. Ruthenium Complex Catalysts Supported by a Bis(trifluoromethyl)pyrazolyl–Pyridyl-Based NNN Ligand for Transfer Hydrogenation of Ketones. Organometallics 2014, 33, 974–982. [Google Scholar] [CrossRef]
  52. Abebayehu, A.; Dutta, R.; Lee, C.-H. Synthesis, Characterization and Properties of Expanded Pyriporphyrins: A New Family of Alkylidenyl Porphyrin Homologues Bearing meso-Exocyclic Double Bonds. Chem. Eur. J. 2016, 22, 13850–13856. [Google Scholar] [CrossRef]
  53. Sheldrick, G. SHELXT – Integrated space-group and crystal-structure determination. Acta Crystallogr. Sect. A 2015, 71, 3–8. [Google Scholar] [CrossRef] [Green Version]
  54. Sheldrick, G. Crystal structure refinement with SHELXL. Acta Crystallographica Section C 2015, 71, 3–8. [Google Scholar]
  55. Boudreaux, E.A.; Mulay, L.N. Theory and Applications of Molecular Paramagnetism; John Wiley & Sons: New York, NY, USA, 1976. [Google Scholar]
  56. Alvarez, S.; Llunell, M. Continuous symmetry measures of penta-coordinate molecules: Berry and non-Berry distortions of the trigonal bipyramid. J. Chem. Soc. Dalton Trans. 2000, 3288–3303. [Google Scholar] [CrossRef]
  57. 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]
  58. Alvarez, S. Distortion Pathways of Transition Metal Coordination Polyhedra Induced by Chelating Topology. Chem. Rev. 2015, 115, 13447–13483. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Ishikawa, R.; Horii, Y.; Nakanishi, R.; Ueno, S.; Breedlove, B.K.; Yamashita, M.; Kawata, S. Field-Induced Single-Ion Magnetism Based on Spin-Phonon Relaxation in a Distorted Octahedral High-Spin Cobalt(II) Complex. Eur. J. Inorg. Chem. 2016, 3233–3239. [Google Scholar] [CrossRef]
  60. Kharwar, A.K.; Mondal, A.; Sarkar, A.; Rajaraman, G.; Konar, S. Modulation of Magnetic Anisotropy and Exchange Interaction in Phenoxide-Bridged Dinuclear Co(II) Complexes. Inorg. Chem. 2021, 60, 11948–11956. [Google Scholar] [CrossRef]
  61. Bill, E. julX: A Program for the Simulation and Analysis of Magnetic Susceptibility Data. V. 1.4; Max Planck Institute for Chemical Energy Conversion: Mülheim/Ruhr, Germany, 2008. [Google Scholar]
  62. Tong, J.; Demeshko, S.; John, M.; Dechert, S.; Meyer, F. Redox-Induced Single-Molecule Magnetism in Mixed-Valent [2 × 2] Co4 Grid Complexes. Inorg. Chem. 2016, 55, 4362–4372. [Google Scholar] [CrossRef]
  63. Schweinfurth, D.; Sommer, M.G.; Atanasov, M.; Demeshko, S.; Hohloch, S.; Meyer, F.; Neese, F.; Sarkar, B. The Ligand Field of the Azido Ligand: Insights into Bonding Parameters and Magnetic Anisotropy in a Co(II)–Azido Complex. J. Am. Chem. Soc. 2015, 137, 1993–2005. [Google Scholar] [CrossRef]
  64. Gomez-Coca, S.; Cremades, E.; Aliaga-Alcalde, N.; Ruiz, E. Mononuclear single-molecule magnets: Tailoring the magnetic anisotropy of first-row transition-metal complexes. J. Am. Chem. Soc. 2013, 135, 7010–7018. [Google Scholar] [CrossRef]
  65. Craig, G.A.; Murrie, M. 3d single-ion magnets. Chem. Soc. Rev. 2015, 44, 2135–2147. [Google Scholar] [CrossRef] [Green Version]
  66. Zadrozny, J.M.; Long, J.R. Slow Magnetic Relaxation at Zero Field in the Tetrahedral Complex [Co(SPh)4]2–. J. Am. Chem. Soc. 2011, 133, 20732–20734. [Google Scholar] [CrossRef] [PubMed]
  67. Chen, L.; Wang, J.; Wei, J.-M.; Wernsdorfer, W.; Chen, X.-T.; Zhang, Y.-Q.; Song, Y.; Xue, Z.-L. Slow Magnetic Relaxation in a Mononuclear Eight-Coordinate Cobalt(II) Complex. J. Am. Chem. Soc. 2014, 136, 12213–12216. [Google Scholar] [CrossRef] [PubMed]
  68. Gomez-Coca, S.; Urtizberea, A.; Cremades, E.; Alonso, P.J.; Camon, A.; Ruiz, E.; Luis, F. Origin of slow magnetic relaxation in Kramers ions with non-uniaxial anisotropy. Nat. Commun. 2014, 5, 4300. [Google Scholar] [CrossRef] [Green Version]
  69. Roy, S.; Oyarzabal, I.; Vallejo, J.; Cano, J.; Colacio, E.; Bauza, A.; Frontera, A.; Kirillov, A.M.; Drew, M.G.B.; Das, S. Two Polymorphic Forms of a Six-Coordinate Mononuclear Cobalt(II) Complex with Easy-Plane Anisotropy: Structural Features, Theoretical Calculations, and Field-Induced Slow Relaxation of the Magnetization. Inorg. Chem. 2016, 55, 8502–8513. [Google Scholar] [CrossRef]
  70. Vallejo, J.; Castro, I.; Ruiz-García, R.; Cano, J.; Julve, M.; Lloret, F.; De Munno, G.; Wernsdorfer, W.; Pardo, E. Field-Induced Slow Magnetic Relaxation in a Six-Coordinate Mononuclear Cobalt(II) Complex with a Positive Anisotropy. J. Am. Chem. Soc. 2012, 134, 15704–15707. [Google Scholar] [CrossRef] [PubMed]
  71. Zadrozny, J.M.; Liu, J.; Piro, N.A.; Chang, C.J.; Hill, S.; Long, J.R. Slow magnetic relaxation in a pseudotetrahedral cobalt(II) complex with easy-plane anisotropy. Chem. Commun. 2012, 48, 3927–3929. [Google Scholar] [CrossRef]
  72. Guo, Y.-N.; Xu, G.-F.; Gamez, P.; Zhao, L.; Lin, S.-Y.; Deng, R.; Tang, J.; Zhang, H.-J. Two-Step Relaxation in a Linear Tetranuclear Dysprosium(III) Aggregate Showing Single-Molecule Magnet Behavior. J. Am. Chem. Soc. 2010, 132, 8538–8539. [Google Scholar] [CrossRef]
  73. Reta, D.; Chilton, N.F. Uncertainty estimates for magnetic relaxation times and magnetic relaxation parameters. PCCP 2019, 21, 23567–23575. [Google Scholar] [CrossRef]
  74. Lucaccini, E.; Sorace, L.; Perfetti, M.; Costes, J.-P.; Sessoli, R. Beyond the anisotropy barrier: Slow relaxation of the magnetization in both easy-axis and easy-plane Ln(trensal) complexes. Chem. Commun. 2014, 50, 1648–1651. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Zadrozny, J.M.; Atanasov, M.; Bryan, A.M.; Lin, C.-Y.; Rekken, B.D.; Power, P.P.; Neese, F.; Long, J.R. Slow magnetization dynamics in a series of two-coordinate iron(II) complexes. Chem. Sci. 2013, 4, 125–138. [Google Scholar] [CrossRef]
  76. Wu, J.; Demeshko, S.; Dechert, S.; Meyer, F. Hexanuclear [Cp*Dy]6 single-molecule magnet. Chem. Commun. 2020, 56, 3887–3890. [Google Scholar] [CrossRef] [PubMed]
  77. Orbach, R. Spin-Lattice Relaxation in Rare-Earth Salts. Proc. R. Soc. Lond. A 1961, 264, 458–484. [Google Scholar]
  78. Orbach, R. Spin-Lattice Relaxation in Rare-Earth Salts: Field Dependence of the Two-Phonon Process. Proc. R. Soc. Lond. A 1961, 264, 485–495. [Google Scholar]
Figure 1. Crystal structure of the complex Co2. Color code: CoII, orange-red; N, blue. The hydrogen atoms, counter-ions, and solvent molecules have been omitted for clarity.
Figure 1. Crystal structure of the complex Co2. Color code: CoII, orange-red; N, blue. The hydrogen atoms, counter-ions, and solvent molecules have been omitted for clarity.
Magnetochemistry 07 00153 g001
Figure 2. The plots of χMT vs. T of the complexes Co2 (a) and Fe2 (b) (black circles, experimental data; red lines, fits to the data) between 2 and 300 K at 1000 Oe.
Figure 2. The plots of χMT vs. T of the complexes Co2 (a) and Fe2 (b) (black circles, experimental data; red lines, fits to the data) between 2 and 300 K at 1000 Oe.
Magnetochemistry 07 00153 g002
Figure 3. Field-dependent molar magnetization measurements for Co2 at 1.9, 3.0, and 5.0 K (inset: the M vs. H/T plots).
Figure 3. Field-dependent molar magnetization measurements for Co2 at 1.9, 3.0, and 5.0 K (inset: the M vs. H/T plots).
Magnetochemistry 07 00153 g003
Figure 4. (a) Temperature-dependent and (b) frequency-dependent ac susceptibility of Co2 under 3500 Oe dc field.
Figure 4. (a) Temperature-dependent and (b) frequency-dependent ac susceptibility of Co2 under 3500 Oe dc field.
Magnetochemistry 07 00153 g004
Figure 5. Cole–Cole plots for Co2 under a 3500 Oe dc field in the temperature range of 1.9–8.0 K. The red lines represent the best fits.
Figure 5. Cole–Cole plots for Co2 under a 3500 Oe dc field in the temperature range of 1.9–8.0 K. The red lines represent the best fits.
Magnetochemistry 07 00153 g005
Figure 6. Plots of τ vs. T−1 for Co2 obtained under 3500 Oe dc fields. The red lines represent the best fits using Equation (2).
Figure 6. Plots of τ vs. T−1 for Co2 obtained under 3500 Oe dc fields. The red lines represent the best fits using Equation (2).
Magnetochemistry 07 00153 g006
Figure 7. Plots of τ vs. T–1 for Co2 obtained under 3500 Oe dc fields. The red lines represent the best fits using Equation (3).
Figure 7. Plots of τ vs. T–1 for Co2 obtained under 3500 Oe dc fields. The red lines represent the best fits using Equation (3).
Magnetochemistry 07 00153 g007
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yang, Q.; Li, X.-L.; Tang, J. Syntheses, Structures and Magnetic Properties of M2 (M = Fe, Co) Complexes with N6 Coordination Environment: Field-Induced Slow Magnetic Relaxation in Co2. Magnetochemistry 2021, 7, 153. https://doi.org/10.3390/magnetochemistry7120153

AMA Style

Yang Q, Li X-L, Tang J. Syntheses, Structures and Magnetic Properties of M2 (M = Fe, Co) Complexes with N6 Coordination Environment: Field-Induced Slow Magnetic Relaxation in Co2. Magnetochemistry. 2021; 7(12):153. https://doi.org/10.3390/magnetochemistry7120153

Chicago/Turabian Style

Yang, Qianqian, Xiao-Lei Li, and Jinkui Tang. 2021. "Syntheses, Structures and Magnetic Properties of M2 (M = Fe, Co) Complexes with N6 Coordination Environment: Field-Induced Slow Magnetic Relaxation in Co2" Magnetochemistry 7, no. 12: 153. https://doi.org/10.3390/magnetochemistry7120153

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