Next Article in Journal
TiO2-La2O3 as Photocatalysts in the Degradation of Naproxen
Next Article in Special Issue
Structural and Photoluminescent Properties of a Novel Terbium Bis(thiocyanato)aurate, Tb[Au(SCN)2]3·6H2O
Previous Article in Journal
Spontaneous Adsorption and Efficient Photodegradation of Indigo Carmine under Visible Light by Bismuth Oxyiodide Nanoparticles Fabricated Entirely at Room Temperature
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Tetranuclear Dysprosium Schiff Base Complex Showing Slow Relaxation of Magnetization

1
Department of Industrial Chemistry, College of Applied Sciences and Nanotechnology Excellence Center, Addis Ababa Science and Technology University, Addis Ababa P.O. Box 16417, Ethiopia
2
Department of Chemistry, Faculty of Natural Sciences, University of SS Cyril and Methodius, 91701 Trnava, Slovakia
3
Institute of Nanotechnology, Karlsruhe Institute of Technology, Hermann-von-Helmholtz-Platz 1, 6344 Karlsruhe, Germany
4
Institute of Quantum Materials and Technologies (IQMT), Karlsruhe Institute of Technology, Hermann-von-Helmholtz-Platz 1, 76344 Karlsruhe, Germany
5
Centre Européen de Science Quantique (CESQ), Institut de Science et d’Ingénierie Supramoléculaires (ISIS, UMR 7006), CNRS-Université de Strasbourg, 8 Allée Gaspard Monge, BP 70028, CEDEX, 67083 Strasbourg, France
*
Authors to whom correspondence should be addressed.
Submission received: 26 April 2022 / Revised: 17 May 2022 / Accepted: 19 May 2022 / Published: 21 May 2022
(This article belongs to the Special Issue Synthesis, Structure and Properties of f-Block Complexes)

Abstract

:
A tetranuclear dysprosium Schiff base complex was isolated by reacting dysprosium chloride with 2-hydroxy-3-methoxybenzaldehyde and 2-(aminomethyl)pyridine in-situ under basic conditions. The isolated Dy(III) complex was characterized by elemental analyses, single crystal X-ray diffraction and molecular spectroscopy. The complex crystallizes in the triclinic space group P-1 with unit cell parameters of a = 10.2003 (4), b = 13.8602 (5), c = 14.9542 (6), α = 94.523 (3), β = 109.362 (4), and γ = 99.861 (3). The magnetic properties of 1 have been investigated by DC and AC susceptibility measurements. The DC measurements reveal weak exchange coupling of antiferromagnetic nature. In the AC measurement, the complex shows a slow relaxation of magnetization in the absence of an external magnetic field.

1. Introduction

Single molecule magnets (SMMs) are individual molecular compounds that exhibit slow relaxation of magnetization at zero field [1] below a certain temperature, which is called the blocking temperature [2]. Such slow relaxation is achieved by magnetic anisotropy causing an effective energy barrier Ueff to spin reversal. Mechanisms to overcome the barrier include spin-lattice processes (Ram, Orb, Direct) or quantum tunneling of magnetization (QTM) [3].
After the first SMM, [Mn12O12(O2CMe)16(H2O)4] [4], was discovered, high spin, strongly coupled 3d transition metal complexes were initially considered to have the most potential for the development of SMMs. However, it has been confirmed over recent years that the f-block elements can also be used, which is based on the interaction between the electron density of the 4f ions and the crystal field environments in which it is placed. Lanthanides-based SMMs entered the field in 2003 with the report on the slow relaxation of magnetization in the LnPc2 double-decker complexes [5], and hundreds of lanthanide SMMs were reported then after [6,7,8,9,10,11,12,13,14,15,16,17,18].
Due to their molecular properties, such as solubility, crystallinity, and so forth, SMMs have been proposed as promising candidates for various modern technological advancements, such as information and data storage as well as molecular spin qudits for quantum algorithms [19] The coordination chemistry of Ln(III) ions has attracted the interest of many researchers for their ability to form clusters (polynuclear) with unprecedented and nanosized structures [20]. Lanthanide-based SMMs received special attention in molecular magnetism on account of their large spin state and high magnetic anisotropy [21].
Schiff bases are ideal candidates because of their fine tunability for the ligand field by varying substituents of both aldehyde and amine precursors [22,23]. Schiff base ligands containing two coordination pockets are believed to be effective for the preparation of heterometallic transition–lanthanide(3d-4f) single molecule magnets by forming individual pockets for 3d and 4f metal ions. These kinds of single molecule magnets are obtained by the exchange coupling of both 3d and 4f ions through oxide, hydroxide, or alkoxide bridges in the ligands. In the present study, we aimed at occupying both coordination pockets with dysprosium to try to understand their exchange coupling, which indeed led to the generation of a single molecule magnet.
Previous studies showed that the Schiff base ligand (HL) in the present study has been used to generate mixed metal isostructural hexanuclear Zn2Ln4 [24], dinuclear dysprosium [25], tetranuclear dysprosium cluster [26], copper [27], and palladium [28] complexes.
Here, we are reporting a novel pure 4f lanthanide compound with the formula of C56H54Cl6Dy4N8O10 (1) which synthesized by the in situ condensation of o-vanillin with 2-(aminomethyl)pyridine in the presence of triethylamine. Single-crystal XRD and magnetic studies reveal that it is a tetranuclear dysprosium Schiff base complex with weak magnetic exchange. Frequency-dependent AC susceptibility measurements suggest SMM behavior having a slow relaxation of magnetization at lower and higher frequencies.

2. Results and Discussion

2.1. Structural Description of Tetranuclear Dysprosium Complex 1

Single crystal XRD studies confirm that complex 1 crystallizes in triclinic space group P-1 with half the molecule in the asymmetric unit (Figure 1a). The aggregate forms a dicationic planar tetranuclear dysprosium complex [(Dy4(L)42-OH)2Cl4]2+ (Figure 1 and Scheme 1). In total, the molecule consists of four Dy(III) ions (Dy1, Dy2 and their symmetry equivalents (Dy1* (1-x, 1-y, 1-z), Dy2*(1-x, 1-y, 1-z)), four tetradentate Schiff base ligands (L), two bridging hydroxo, two bridging and two terminating chlorido ligands. The four Dy(III) ions are held together by four deprotonated Schiff base ligands (L), two (μ2-OH) and two (μ2-chloro). The other coordination sphere is completed by one chlorido ligand for Dy2 and Dy2*, respectively. The charge on the Schiff base complex is balanced by two chloride anions located outside of the sphere. The Schiff base ligand (HL) (Scheme 2a) adopts different bridging behavior of µ-1, µ-2, µ-1, µ-1 respectively after deprotonating (L) for Dy1 and Dy2, as shown in Scheme 2b,c, respectively. In a nutshell, the Schiff base ligand is acting in both chelating and bridging mode in the molecular structure of complex 1 (Figure 1b).
The four Dy(III) ions of the core almost lie on one plane, and the compound exhibits a planar butterfly-type structural topology. The metal center Dy1 and its symmetry equivalent Dy1* define the body of the butterfly, and Dy2 and its symmetry equivalent Dy2* define the wing tips of the butterfly motif (Figure 1b–d). The body and wing tips of the butterfly core are connected through the two μ3-OH bridges. These bridging ligands are located above and below the plane. The peripheral part of the metal hydroxo cores are being bridged by four Schiff base ligands and two chloro ligands. Dy2 is connected to Dy1 via one phenoxido oxygen (O4) of one independent Schiff base ligand. Dy2 and Dy2* are further coordinated by one chlorido ligand (Cl2, Cl2*), respectively.
The bond distances of Dy2-O1 and Dy1–O1 are 2.379 (4) Å and 2. 295 (4) Å, respectively. Taking one of the half asymmetric molecular units, the bond angles of Dy1*--Cl1--Dy2, Dy1*--O2--Dy2, Dy1*--O1--Dy2, Dy2--O1--Dy1, Dy2--O4--Dy1 and Dy1*--O1--Dy1 are 80.12 (3)°, 99.6 (1)°, 100.0 (1)°, 107.3 (1)°, 112.6 (1)° and 108.9 (1)°, respectively (Table S1). The central dysprosium atoms Dy1 and Dy1* and the outer Dy2 and Dy2* are eight coordinate: with four oxygen, two chlorine and two nitrogen atoms for Dy2 and Dy2* and with five oxygen, one chlorine and two nitrogen atoms for Dy1 and Dy1*. Calculations utilizing the shape v2.1 program [29] show that the coordination polyhedra of the two Dy(III)-ion central atoms in complex 1 belong to biaugmented trigonal prism (BTP) geometry (8-BTP, C2v, 2.303) (Figure 2d, Table S3). Therefore, each eight-coordinate Dy(III) ion possesses a distorted biaugmented trigonal prism (BTP) geometry (Figure 2d) with a N2O5Cl coordination environment for Dy1 and Dy1* and N2O4Cl2 coordination environment for Dy2 and Dy2*. The two square bases of the biaugmented trigonal prism for Dy2 comprise 01, O2, O3, O4, N3 and N4, respectively. However, for Dy1, the two square bases are well-defined by the atoms of O1, O1*, O2, O4, O5, N1 and N2 respectively (Figure 2). The complex under investigation is somewhat close to previously reported planar tetranuclear lanthanide hydroxo aggregates in terms of their bond distances and bond angles [30,31,32,33,34,35,36]. Wang et al. [26] have reported a planar tetranuclear dysprosium mixed ligand Schiff base aggregate, generating somewhat the same topology close to our report. The main difference between the two structures are on the coordination environment around the dysprosium metal ions. However, in both the aggregates, each Dy(III) ion possesses an eight-coordinated biaugmented trigonal prism (BTP) geometry as a common and additionally a triangular dodecahedron geometry in the reported structure of Wang et al. In addition to the preparative similarity, the similarities in the Dy–O, Dy–N, and Dy–Dy bond distances are also noticeable.
From the unit cell packing, it has been revealed that there is one molecule exiting per unit cell (Figure 2a–c) without hydrogen bonding. However, there are several intermolecular connections happening in the molecule in the unit cell of half of the asymmetric unit. The bond distances observed in the intermolecular contacts/bonding in H26–Cl3, H23A–Cl3, H15–Cl1, H15A–Cl2, C15–Cl1 and H8–Cl3 are in the range of 2.760–3.255 Å. The molecular packing of the extended network in different directions is shown in Figure S1.

2.2. DC Susceptibility

The temperature dependence of the magnetic susceptibility for 1 was taken between T = 2 and 300 K at the magnetic field BDC = 0.1 T. Raw data were corrected for the estimated underlying diamagnetism and transformed to the dimensionless product function χT/C0 ( C 0 = N A μ 0 μ B 2 / k B is the reduced Curie constant containing only the fundamental physical constants in their usual meaning) that is displayed in Figure 3.
The value of the product function with the coupling switched off (as assumed at room temperature) for four Dy(III) centers is χT/C0 = 4 × 37.78 = 151.1 (χT = 52.1 cm3 K mol−1 in the cgs and emu units). The observed r.t. value is χT/C0 = 118.5 due to the effect that not Jmax = 4 × (15/2) = 60/2 is the ground molecular state. On cooling, the product function decreases and falls down to the value of 51.9 at T = 2.0 K.
The ground electronic state (multiplet) of a Dy(III) center is 6H15/2 with gJ = 4/3. Four Dy(III) centers produce N = 164 = 65,536 magnetic states and M = 2736 zero field states with equal energy. When only the isotropic exchange is considered, the total molecular angular momentum (J) is a good quantum number, and the large-dimensional interaction matrix can be factored to low-dimensional blocks. The molecular values vary between Jmin = 0 and Jmax = 4·(15/2) = 60/2. Using the technique of irreducible tensor operators [37] the dimensions of these blocks are 16, 45, 71, 94, 114, 131, 145, 156, 164, 169, 171, 170, 166, 159, 149, 136, 120, 105, 91, 78, 66, 55, 45, 36, 28, 21, 15, 10, 6, 3, and 1; these are irrespective of the exchange coupling scheme. The largest dimension MJ = 171 occurs for J = 20/2. After the diagonalization of each block, all eigenvalues are collected and inserted to the partition function. A Zeeman term is added to each zero-field energy level for three working fields (this stays diagonal in the basis set of J-manifold assuming a uniform g-factor for each of the Dy centers). Finally, the magnetization and the magnetic susceptibility are calculated by means of statistical thermodynamics as the first and the second field-derivatives of the partition function, respectively.
A reasonable exchange coupling model will require three coupling constants as shown in Table 1. However, in order to avoid an overparameterization and the mutual dependence of the coupling constants, in the simplest model, all of them were set equal. Such a simple exchange coupling model gave the following set of magnetic parameters: Jex/hc = –0.036 (14) cm−1, g = 1.19 (1) and the temperature-independent magnetism χTIM = 4.9 × 10−6 m3 mol−1; the discrepancy factor of the fit R = 0.077). A release of the constraint for the J-constants gave J1/hc = –0.0078 cm−1, J2/hc = –0.094 cm−1, J3/hc = –0.036 cm−1, g = 1.187 (2) and χTIM = 4.8 × 10−6 m3 mol−1 (R = 0.011). The calculated susceptibility is drawn in Figure 3 as a solid line.
The zero-field energy levels (M = 2736) are displayed in Figure 4 showing that the ground molecular state is J = 0. This explains an observed value of the product function. A further improvement of the model would be based on the zero-field splitting and other crystal-field effects. In such a case, however, there would be the off-diagonal matrix elements mixing the blocks of the different angular momentum so that the blocking of the interaction matrix is not possible anymore.

2.3. AC Susceptibility

In order to test a possible slow magnetic relaxation in 1, AC susceptibility measurements were performed using the small amplitude of the oscillating field BAC = 0.3 mT, at T = 2 K for a variable set of frequencies f = 1 − 1488 Hz, and applied external field BDC = 0 to 0.5 T; the results are presented in Figure 5. The out-of-phase susceptibility is non-zero even in the absence of the external field. This means that 1 behaves as a true single molecule magnet. With increasing external field, the profile of χ” alters.
Both AC susceptibility components were fitted simultaneously by employing the two-set Debye model. This model contains seven free parameters: a pair of isothermal susceptibilities χT1 and χT2, two distribution parameters α1 and α2, two relaxation times τ1 (low-frequency, LF) and τ2 (high-frequency, HF) and the common adiabatic susceptibility χS. At T = 2.0 K and BDC = 0, the relaxation times are τLF = 29 ms and τHF = 75 μs, and the mole fraction xLF = 0.24 (xHF = 1 − xLF). With BDC = 0.5 T, these parameters alter to τLF = 9 ms and τHF = 66 μs, and xLF = 0.52. To this end, 1 is the single-molecule magnet even in the absence of the external magnetic field.
The temperature dependence of the AC susceptibility for various frequencies of the oscillating field at BDC = 0 is shown in Figure 6. It is seen that the out-of-phase susceptibility survives until T > 10 K.
The same dataset has been rearranged as a frequency dependence of the AC susceptibility for a set of temperatures as shown in Figure 7. This function can be fitted to the two-set Debye model, since two relaxation channels are evident: the low-frequency (LF) and the high-frequency (HF) one.
Temperature evolution of the (fitted) relaxation time, individually for the LF and HF relaxation channels, is presented in Figure 8. It can be seen that the HF relaxation time below 8 K is almost temperature independent, which indicates a quantum tunnelling relaxation process.

3. Materials and Methods

3.1. General Procedures

All the starting materials, o-vanillin, 2-(aminomethyl)pyridine, triethylamine, dysprosium chloride hexahydrate and solvents were of analytical reagent grade and were used without any further purification.
Elemental analysis for C, H, and N was carried out on a Flash 2000 CHNSO apparatus (Thermo Scientific). FTIR spectra were measured by the ATR method in the region of 400–4000 cm−1 (Shimadzu IR Affinity−1, Quest ATR holder).
Magnetic susceptibility data were collected at temperatures between 2 and 300 K using a Quantum Design MPMS-XL SQUID magnetometer equipped with a 1 T magnet at an external field of 0.1 T. The samples were grounded and fixed in a gelatine capsule using small amounts of eicosane to avoid any movement of the sample. The data obtained were corrected for diamagnetic contributions of the sample, the eicosane, the gelatine capsule and the sample holder.

3.2. Synthesis of Tetranuclear Dysprosium Complex (1)

The tetranuclear complex (C56H54Cl6Dy4N8O10) (1) (Scheme 1) was synthesized by slowly adding a methanol solution of DyCl3·6H2O (1.0 mmol, 0.376 g) dissolved in 15 mL into a stirring solution of o-vanillin (1 mmol, 0.152 g) and 2-(aminomethyl)pyridine (1 mmol, 0.108 mL) in the presence of triethylamine (1 mmol, 0.101 mL) in methanol (50 mL). The mixture was refluxed for 4 h in an oil bath. The yellow solution obtained on reflux was cooled to room temperature and filtered. Vapor diffusion of diethyl ether to the filtered yellow solution yielded X-ray quality yellow block crystals. The complex was collected by filtration and washed with cold MeOH and dried in air and vacuum. Yield, 200 mg, 52.25%. Anal. Calc. for 1, 38.97 (C), 7.15 (N), 4.95 (H); Found: 38.84 (C), 7.02 (N), 4.42 (H). IR (KBr disc)/cm−1: 2981.25 (w), 2943.44 (w), 2609.96 (m), 2496.53 (m), 2358.15 (m), 2333.20 (w), 1634.48 (s), 1558.86 (m), 1508.95 (m), 1464.34 (s), 1301 (m), 1212.53 (s), 1168.67 (m), 1036.34 (s), 961.48 (w), 854.10 (w), 740.67 (w), 627.24 (m), 552.38 (m), 413.24 (s).

3.3. X-ray Crystallography

Single crystal X-ray diffraction measurements were performed on a STOE StadiVadi 25 diffractometer using a GeniX 3D HF micro focus with MoKα-radiation (λ = 0.71073 Å) and a CCD image plate detector. The crystals were mounted using crystallographic oil and placed in a cold nitrogen stream. All the data were corrected for absorption using CrysAlisPro [38]. The structures were solved by direct methods and refined against F2 using the SHELXL-97 package [39] in Olex2.25. All non-hydrogen atoms were refined anisotropically, and hydrogens were placed based on a riding model approach. Full crystallographic details can be found in CIF format: see the Cambridge Crystallographic Data Centre database 2161049 for complex 1. Crystal parameters and refinement results for the complex are collated in Table S2, and the selected bond lengths and bond angles of complex 1 are presented in Table S1.

4. Conclusions

In summing up, a tetranuclear dysprosium Schiff base complex was synthesized and characterized by single crystal XRD, elemental analysis and molecular spectroscopy. DC susceptibility measurements of the aggregate reveal an antiferromagnetic behavior of the complex. AC susceptibility proves that the complex is exhibiting single molecule magnetic behavior, showing slow magnetic relaxation at zero field. As the coordination environments clearly change the anisotropic nature of the four dysprosium ions, the present case opens ample opportunities to play around the anisotropy by changing the coordination environment, introducing new ligand functionalities as bridging or by coordination capacities. In general, we are expecting that the present work with modified ligand architecture will be promising to design and synthesize dysprosium-based single molecule magnets with superior characteristics.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/inorganics10050066/s1, Figure S1: View on intermolecular contacts formed in complex 1 through the building of extended networks in the three crystallographic directions (along a, along b and along c); Table S1: Selected bond lengths (Å) and bond angles (°) in 1; Table S2: Crystal data and structure refinement for 1.

Author Contributions

The major work for this article, designing, execution and writing of the original draft, was completed by the first author: M.G., which is part of his Ph.D. program. The second author: S.S. participated in characterizations (solving the structure and SQUID measurement). The remaining authors: C.R., A.S., M.R., M.T. and R.B. were responsible for supervision, editing, and reviewing of this article. In addition, the authors: C.R., M.R. and R.B. have completed the DC and AC data analysis of the compound. All authors have read and agreed to the published version of the manuscript.

Funding

This research project was supported by the National Scholarship Programme of the Slovak Republic 2021 and Addis Ababa Science and Technology University, Ethiopia for (M.G). Financial support of Slovak grants agencies (APVV 18-0016, APVV 19-0087, VEGA 1/0191/22 and VEGA 1/0086/21) are thankfully acknowledged.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We are thankful to National Scholarship Programme of the Slovak Republic 2021 for a research stay and Addis Ababa Science and Technology University, Ethiopia for a Ph.D. studentship for M.G. We acknowledge our thankfulness to R. Mičová UCM, Trnava, Slovakia, for the spectroscopic and microanalysis. We are also thankful to S. Klyatskaya from KIT, Germany for her support during this work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gatteschi, D.; Sessoli, R.; Villain, J. Molecular Nanomagnets; Oxford University Press: Oxford, UK, 2006; ISBN 9780191718298. [Google Scholar]
  2. Ritter, S.K. Single-molecule magnets evolve. Chem. Eng. News 2004, 82, 29–32. [Google Scholar] [CrossRef]
  3. Layfield, R.A.; Murugesu, M. Lanthanides and Actinides in Molecular Magnetism; John Wiley & Sons: Hoboken, NJ, USA, 2015; ISBN 9781119950837. [Google Scholar]
  4. Sessoli, R.; Gatteschi, D.; Caneschi, A.; Novak, M.A. Magnetic bistability in a metal-ion cluster. Nature 1993, 365, 141–143. [Google Scholar] [CrossRef]
  5. Ishikawa, N.; Sugita, M.; Ishikawa, T.; Koshihara, S.; 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]
  6. Zhang, P.; Zhang, L.; Tang, J. Lanthanide single molecule magnets: Progress and perspective. Dalt. Trans. 2015, 44, 3923–3929. [Google Scholar] [CrossRef] [PubMed]
  7. Woodruff, D.N.; Winpenny, R.E.P.; Layfield, R.A. Lanthanide Single-Molecule Magnets. Chem. Rev. 2013, 113, 5110–5148. [Google Scholar] [CrossRef]
  8. Boča, R.; Stolárová, M.; Falvello, L.R.; Tomás, M.; Titiš, J.; Černák, J. Slow magnetic relaxations in a ladder-type Dy(III) complex and its dinuclear analogue. Dalt. Trans. 2017, 46, 5344–5351. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Dolai, M.; Ali, M.; Rajnák, C.; Titiš, J.; Boča, R. Slow magnetic relaxation in Cu(ii)-Eu(iii) and Cu(ii)-La(iii) complexes. New J. Chem. 2019, 43, 12698–12701. [Google Scholar] [CrossRef]
  10. Hazra, S.; Titiš, J.; Valigura, D.; Boča, R.; Mohanta, S. Bis-phenoxido and bis-acetato bridged heteronuclear {CoIIIDyIII} single molecule magnets with two slow relaxation branches. Dalt. Trans. 2016, 45, 7510–7520. [Google Scholar] [CrossRef]
  11. Luzon, J.; Sessoli, R. Lanthanides in molecular magnetism: So fascinating, so challenging. Dalt. Trans. 2012, 41, 13556–13567. [Google Scholar] [CrossRef]
  12. Sorace, L.; Benelli, C.; Gatteschi, D. Lanthanides in molecular magnetism: Old tools in a new field. Chem. Soc. Rev. 2011, 40, 3092–3104. [Google Scholar] [CrossRef]
  13. Liddle, S.T.; Slageren, J. van Improving f-element single molecule magnets. Chem. Soc. Rev. 2015, 44, 6655–6669. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Habib, F.; Murugesu, M. Lessons learned from dinuclear lanthanide nano-magnets. Chem. Soc. Rev. 2013, 42, 3278–3288. [Google Scholar] [CrossRef] [Green Version]
  15. Pinkowicz, D.; Southerland, H.I.; Avendan, C.; Prosvirin, A.; Sanders, C.; Wernsdorfer, W.; Pedersen, K.S.; Dreiser, J.; Cle, R.; Nehrkorn, J.; et al. Cyanide Single-Molecule Magnets Exhibiting Solvent Dependent Reversible “On” and “Off ” Exchange Bias Behavior. J. Am. Chem. Soc. 2015, 45, 14406–14422. [Google Scholar] [CrossRef] [PubMed]
  16. Dreiser, J. Molecular lanthanide single-ion magnets: From bulk to submonolayers. J. Phys. Condens. Matter 2015, 27, 183203. [Google Scholar] [CrossRef] [PubMed]
  17. Lin, S.Y.; Tang, J. Versatile tetranuclear dysprosium single-molecule magnets. Polyhedron 2014, 83, 185–196. [Google Scholar] [CrossRef]
  18. Vráblová, A.; Tomás, M.; Falvello, L.R.; Dlháň, Ľ.; Titiš, J.; Černák, J.; Boča, R. Slow magnetic relaxation in Ni-Ln (Ln = Ce, Gd, Dy) dinuclear complexes. Dalt. Trans. 2019, 48, 13943–13952. [Google Scholar] [CrossRef] [Green Version]
  19. 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] [Green Version]
  20. Peng, J.B.; Kong, X.J.; Zhang, Q.C.; Orendáč, M.; Prokleška, J.; Ren, Y.P.; Long, L.S.; Zheng, Z.; Zheng, L.S. Beauty, symmetry, and magnetocaloric effect-four-shell keplerates with 104 lanthanide atoms. J. Am. Chem. Soc. 2014, 136, 17938–17941. [Google Scholar] [CrossRef]
  21. Vincent, R.; Klyatskaya, S.; Ruben, M.; Wernsdorfer, W.; Balestro, F. Electronic read-out of a single nuclear spin using a molecular spin transistor. Nature 2012, 488, 357–360. [Google Scholar] [CrossRef]
  22. Gebrezgiabher, M.; Bayeh, Y.; Gebretsadik, T.; Gebreslassie, G.; Elemo, F.; Thomas, M.; Linert, W. Lanthanide-based single-molecule magnets derived from schiff base ligands of salicylaldehyde derivatives. Inorganics 2020, 8, 66. [Google Scholar] [CrossRef]
  23. Senthil Kumar, K.; Bayeh, Y.; Gebretsadik, T.; Elemo, F.; Gebrezgiabher, M.; Thomas, M.; Ruben, M. Spin-crossover in iron(ii)-Schiff base complexes. Dalt. Trans. 2019, 48, 15321–15337. [Google Scholar] [CrossRef] [PubMed]
  24. Khan, A.; Akhtar, M.N.; Lan, Y.; Anson, C.E.; Powell, A.K. Linear shaped hetero-metallic [Zn2Ln4] clusters with Schiff base ligand: Synthesis, characterization and magnetic properties. Inorg. Chim. Acta 2021, 524, 5–10. [Google Scholar] [CrossRef]
  25. Khan, A.; Fuhr, O.; Akhtar, M.N.; Lan, Y.; Thomas, M.; Powell, A.K. Synthesis, characterization and magnetic studies of dinuclear lanthanide complexes constructed with a Schiff base ligand. J. Coord. Chem. 2020, 73, 1045–1054. [Google Scholar] [CrossRef]
  26. Wang, H.L.; Peng, J.M.; Zhu, Z.H.; Mo, K.Q.; Ma, X.F.; Li, B.; Zou, H.H.; Liang, F.P. Step-by-Step and Competitive Assembly of Two Dy(III) Single-Molecule Magnets with Their Performance Tuned by Schiff Base Ligands. Cryst. Growth Des. 2019, 19, 5369–5375. [Google Scholar] [CrossRef]
  27. Maxim, C.; Pasatoiu, T.D.; Kravtsov, V.C.; Shova, S.; Muryn, C.A.; Winpenny, R.E.P.; Tuna, F.; Andruh, M. Copper(II) and zinc(II) complexes with Schiff-base ligands derived from salicylaldehyde and 3-methoxysalicylaldehyde: Synthesis, crystal structures, magnetic and luminescence properties. Inorg. Chim. Acta 2008, 361, 3903–3911. [Google Scholar] [CrossRef]
  28. Patil, S.A.; Weng, C.M.; Huang, P.C.; Hong, F.E. Convenient and efficient Suzuki-Miyaura cross-coupling reactions catalyzed by palladium complexes containing N,N,O-tridentate ligands. Tetrahedron 2009, 65, 2889–2897. [Google Scholar] [CrossRef]
  29. Llunell, S.A.M.; Casanova, D.; Cirera, J.; Alemany, P.; Alvarez, S. SHAPE-Program for the Stereochemical Analysis of Molecular Fragments by Means of Continuous Shape Measures and Associated Tools Version 2.1; Universitat de Barcelona: Barcelona, Spain, 2013. [Google Scholar]
  30. Zheng, Y.Z.; Lan, Y.; Anson, C.E.; Powell, A.K. Anion-perturbed magnetic slow relaxation in planar {Dy4} clusters. Inorg. Chem. 2008, 47, 10813–10815. [Google Scholar] [CrossRef]
  31. Langley, S.K.; Chilton, N.F.; Gass, I.A.; Moubaraki, B.; Murray, K.S. Planar tetranuclear lanthanide clusters with the Dy4 analogue displaying slow magnetic relaxation. Dalt. Trans. 2011, 40, 12656–12659. [Google Scholar] [CrossRef]
  32. Abbas, G.; Lan, Y.; Kostakis, G.E.; Wernsdorfer, W.; Anson, C.E.; Powell, A.K. Series of isostructural planar lanthanide complexes [LnIII4(μ3-OH)2(mdeaH)2(piv) 8] with single molecule magnet behavior for the Dy4 analogue. Inorg. Chem. 2010, 49, 8067–8072. [Google Scholar] [CrossRef]
  33. Jami, A.K.; Ali, J.; Mondal, S.; Homs-Esquius, J.; Sañudo, E.C.; Baskar, V. Dy2 and Dy4 hydroxo clusters assembled using o-vanillin based Schiff bases as ligands and β-diketone co-ligands: Dy4 cluster exhibits slow magnetic relaxation. Polyhedron 2018, 151, 90–99. [Google Scholar] [CrossRef]
  34. Lin, P.H.; Burchell, T.J.; Ungur, L.; Chibotaru, L.F.; Wernsdorfer, W.; Murugesu, M. A polinuclear lanthanide single-molecule magnet with a record anisotropic barrier. Angew. Chemie-Int. Ed. 2009, 48, 9489–9492. [Google Scholar] [CrossRef] [PubMed]
  35. Kuang, W.W.; Zhu, L.L.; Xu, Y.; Yang, P.P. A tetranuclear holmium compound exhibiting single molecule magnet behavior. Inorg. Chem. Commun. 2015, 61, 169–172. [Google Scholar] [CrossRef]
  36. Yan, P.F.; Lin, P.H.; Habib, F.; Aharen, T.; Murugesu, M.; Deng, Z.P.; Li, G.M.; Sun, W. Bin Planar tetranuclear dy(III) single-molecule magnet and its Sm(III), Gd(III), and Tb(III) analogues encapsulated by salen-type and β-diketonate ligands. Inorg. Chem. 2011, 50, 7059–7065. [Google Scholar] [CrossRef] [PubMed]
  37. Boča, R. A Handbook of Magnetochemical Formulae; Elsevier: Amsterdam, The Netherlands, 2012; ISBN 978-0-12-416014-9. [Google Scholar]
  38. Agilent, CrysAlisPro Data Collection and Processing Software for Agilent X-ray Diffractometers; Technol. UK Ltd.: Oxford, UK, 2014; Volume 44, pp. 1–53.
  39. Sheldrick, G.M. A short history of SHELX. Acta Crystallogr. Sect. A Found. Crystallogr. 2008, 64, 112–122. [Google Scholar] [CrossRef] [Green Version]
Scheme 1. Schematic representation of the reaction procedure for complex 1.
Scheme 1. Schematic representation of the reaction procedure for complex 1.
Inorganics 10 00066 sch001
Figure 1. (a) Asymmetric unit of complex 1 (b) Full molecular structure of complex 1; (c) The dysprosium core structure with coordinating atoms from each ligand for complex 1 (d) The dysprosium central core of complex 1: Carbon atoms in (c,d) and hydrogen atoms in all the figures are omitted for clarity. Symmetry code (1-x, 1-y, -z). Color codes: cyan Dy(III), red O, black C, blue N, and green Cl and * represents symmetry generated atom.
Figure 1. (a) Asymmetric unit of complex 1 (b) Full molecular structure of complex 1; (c) The dysprosium core structure with coordinating atoms from each ligand for complex 1 (d) The dysprosium central core of complex 1: Carbon atoms in (c,d) and hydrogen atoms in all the figures are omitted for clarity. Symmetry code (1-x, 1-y, -z). Color codes: cyan Dy(III), red O, black C, blue N, and green Cl and * represents symmetry generated atom.
Inorganics 10 00066 g001
Scheme 2. This Structure of the Schiff base ligand HL (a) and its binding modes after deprotonation (L) (b,c) with Dy1 and Dy2, Dy1*, and Dy2* is in coordination mode of µ-1, µ-2, µ-1, µ-1. All hydrogen atoms in (b,c) are removed for clarity. Color codes: cyan Dy(III), red O, black and * represents symmetry generated atom.
Scheme 2. This Structure of the Schiff base ligand HL (a) and its binding modes after deprotonation (L) (b,c) with Dy1 and Dy2, Dy1*, and Dy2* is in coordination mode of µ-1, µ-2, µ-1, µ-1. All hydrogen atoms in (b,c) are removed for clarity. Color codes: cyan Dy(III), red O, black and * represents symmetry generated atom.
Inorganics 10 00066 sch002
Figure 2. Molecular packing arrangements along the crystallographic (a) a-axis (b) b-axis and (c) c-axis (d) Coordination polyhedra of distorted biaugmented trigonal prism (BTP) geometry observed for Dy(III) ion in complex 1. All hydrogen atoms are omitted for clarity. Color codes: cyan Dy(III), red O, black, C and green Cl.
Figure 2. Molecular packing arrangements along the crystallographic (a) a-axis (b) b-axis and (c) c-axis (d) Coordination polyhedra of distorted biaugmented trigonal prism (BTP) geometry observed for Dy(III) ion in complex 1. All hydrogen atoms are omitted for clarity. Color codes: cyan Dy(III), red O, black, C and green Cl.
Inorganics 10 00066 g002
Figure 3. Plot of the χT vs. T for 1 in an applied magnetic field of BDC = 0.1 T. Solid lines—fitted (see main text). Right—field dependence of the magnetization per formula unit.
Figure 3. Plot of the χT vs. T for 1 in an applied magnetic field of BDC = 0.1 T. Solid lines—fitted (see main text). Right—field dependence of the magnetization per formula unit.
Inorganics 10 00066 g003
Figure 4. Calculated energy spectrum of 1.
Figure 4. Calculated energy spectrum of 1.
Inorganics 10 00066 g004
Figure 5. Field dependence (top) and frequency dependence (bottom) of the AC susceptibility components for 1 at T = 2.0 K. Dashed—guide for eyes; solid lines—fitted.
Figure 5. Field dependence (top) and frequency dependence (bottom) of the AC susceptibility components for 1 at T = 2.0 K. Dashed—guide for eyes; solid lines—fitted.
Inorganics 10 00066 g005
Figure 6. Temperature dependence of the AC susceptibility for 1.
Figure 6. Temperature dependence of the AC susceptibility for 1.
Inorganics 10 00066 g006
Figure 7. The frequency dependence of the AC susceptibility for 1. Solid lines—fitted.
Figure 7. The frequency dependence of the AC susceptibility for 1. Solid lines—fitted.
Inorganics 10 00066 g007
Figure 8. Arrhenius like plot of lnτ vs. T−1 for 1.
Figure 8. Arrhenius like plot of lnτ vs. T−1 for 1.
Inorganics 10 00066 g008
Table 1. Topological matrix for the exchange coupling with the bond angles Dy-O-Dy* in deg a.
Table 1. Topological matrix for the exchange coupling with the bond angles Dy-O-Dy* in deg a.
Dy/Dy*Dy1Dy2Dy1*Dy2*
Dy1-J1, 107, 112J2, 109J3, 100, 99, 80(Cl)
Dy2 -J3, 100, 99, 80(Cl)0
Dy1* -J1, 107, 112
Dy2* -
a According to the core depicted in Figure 1d.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Gebrezgiabher, M.; Schlittenhardt, S.; Rajnák, C.; Sergawie, A.; Ruben, M.; Thomas, M.; Boča, R. A Tetranuclear Dysprosium Schiff Base Complex Showing Slow Relaxation of Magnetization. Inorganics 2022, 10, 66. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics10050066

AMA Style

Gebrezgiabher M, Schlittenhardt S, Rajnák C, Sergawie A, Ruben M, Thomas M, Boča R. A Tetranuclear Dysprosium Schiff Base Complex Showing Slow Relaxation of Magnetization. Inorganics. 2022; 10(5):66. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics10050066

Chicago/Turabian Style

Gebrezgiabher, Mamo, Sören Schlittenhardt, Cyril Rajnák, Assefa Sergawie, Mario Ruben, Madhu Thomas, and Roman Boča. 2022. "A Tetranuclear Dysprosium Schiff Base Complex Showing Slow Relaxation of Magnetization" Inorganics 10, no. 5: 66. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics10050066

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