Next Article in Journal / Special Issue
Metallated [3]Ferrocenophanes Containing P3M Bridges (M = Li, Na, K) §
Previous Article in Journal
Group 4 Metallocene Polymers—Selected Properties and Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Synthesis, Structural, and Magnetic Characterization of a Mixed 3d/4f 12-Metallacrown-4 Family of Complexes

by
Angeliki A. Athanasopoulou
1,2,
Luca M. Carrella
1 and
Eva Rentschler
1,*
1
Institute of Inorganic and Analytical Chemistry, Johannes Gutenberg University Mainz, Duesbergweg 10-14, D-55128 Mainz, Germany
2
Graduate School Materials Science in Mainz, Staudinger Weg 9, D-55128 Mainz, Germany
*
Author to whom correspondence should be addressed.
Inorganics 2018, 6(3), 66; https://doi.org/10.3390/inorganics6030066
Submission received: 21 June 2018 / Revised: 2 July 2018 / Accepted: 5 July 2018 / Published: 7 July 2018
(This article belongs to the Special Issue Organometallic Macrocycles and Their Applications)

Abstract

:
A new family of complexes (tBu4N){[LnIII(O2CBut)4][12-MC-Mn(III)N(shi)-4]}·5CH2Cl2 (Ln = Gd (1) and Tb (2)), (tBu4N)2{[YIII(O2CBut)4][12-MC-Mn(III)N(shi)-4]}(ClO4) (3); where shiH3 = salicylhydroxamic acid; ButCO2 = pivalate ions; tBu4N = tetrabutylammonium and ClO4 = perchlorate ions, has been isolated. The reaction of salicylhydroxamic acid with Mn(O2CBut)2·2H2O, Ln(NO3)3·xH2O, tBu4NClO4 in the presence of morpholine (C4H9NO) led to the isolation of compounds 13. The complexes belong to the 12-MC-4 family of Metallacrowns (MCs) possessing a central {Mn4IIILnIII(µ-NO)4}11+ core with the four MnIII atoms occupying the periphery positions, while the formed [Mn–N–O] repeating unit, assists in the accommodation of the LnIII atom in the center of the ring. Peripheral ligation is provided by four η11:μ pivalate ions. Direct current magnetic susceptibility (dc) measurements revealed the presence of predominant antiferromagnetic exchange interactions within the metal centers. A 1-J fitting model was used in order to quantify the MnIII–MnIII interactions and fitting of the data, for the diamagnetic YIII analogue, gave J = −3.74 cm−1 and gMn(III) = 2.07. Fitting of the {Mn4Gd} compound using a 2-J model, counting additionally for the MnIII–GdIII interactions, revealed values of J1 = −3.52 cm−1, J2 = −0.45 cm−1, and gMn(III) = 1.99.

Graphical Abstract

1. Introduction

Heterometallic 3d/4f complexes continue to attract the interest of the scientific community as they have been proven to be good candidates for possible applications in various fields such as optics [1,2], catalysis [3], and molecular magnetism [4]. In the field of molecular magnetism, the use of paramagnetic 3d metal ions in combination with highly anisotropic lanthanides such as DyIII or TbIII with large and unquenched orbital angular momenta [5] can lead to single-molecule magnetism (SMM) behavior with large anisotropy barriers (or energy barriers) for the magnetization reversal. In a common understanding, an SMM is able to retain its magnetization only as long as it is kept below a characteristic blocking temperature, TΒ, in the absence of an applied magnetic field [6]. The magnitude of the energy barrier to spin reversal (Ueff) in 3d SMMs is equal to S2|D| for integer and (S2 − 1/4)|D| for half-integer spin systems, where D is the zero-field splitting parameter. Thus, the total spin of the molecule, S, and the Ising-type magnetic anisotropy, are the two factors that block the magnetization reversal. In transition metal complexes, the ground state bistability arises from the total spin S with the ensuing [2S + 1] ms microstates, while in lanthanides, the spin-orbit-coupled ground term 2S+1LJ splits into [2J + 1] mJ microstates that are responsible for the magnetic bistability of those complexes [5]. Experimentally, we can detect the slow magnetic relaxation of SMMs by performing alternating-current (ac) susceptibility measurements and, most importantly, by the observation of hysteresis loops, which is the ultimate diagnostic property of bulk classical magnets [7]. Recently, quantum tunneling of magnetization (QTM) [8,9] and quantum interference promote the discussion of SMMs as being ideal candidates for even more advanced applications such as spintronics and quantum computing [9,10,11].
Metallacrowns (MCs) is a class of compounds that since their discovery has attracted the immense attention of the scientific community [12,13,14]. Most of these complexes have repeatedly demonstrated their ability to encapsulate a central metal ion in their MC cavity, similar to crown ethers, and till now a wide range of MC sizes has been reported [15]. The first example of a 12-MC-4 was reported in 1989 by Pecoraro and Lah, and it was a Mn(OAc)2[12-MC-MnIII(N)shi-4] complex where OAc is acetate ions and shi3− is salicylhydroximate ions [13]. Pecoraro and coworkers have exceedingly demonstrated that salicylhydroxamic acid (shaH2, Scheme 1) can possibly be subjected to a metal-assisted 2-amide-iminol tautomerism, which leads to salicylhydroxime (shiH3, Scheme 1); the latter being an excellent chelating-bridging ligand which has also been shown to possess the appropriate geometry to afford the 12-MC-4 motif [1,15]. The formed repeating unit of [MnIII–N–O], along with the triply deprotonated salicylhydroximate, the central MnII ion and the two acetate bridges were responsible for the aforementioned configuration. Usually, in the central MC cavity sits a transition metal ion, even though there are also reports where alkali and alkaline earth metals occupy the cavity [15,16]. Lately, the focus has been turned into the incorporation of lanthanide ions in the center, since these compounds have been proposed as excellent candidates for molecular recognition [17,18], molecular magnetism [19] and luminescent [1,19,20,21,22] technologies.
Although there have been numerous 12-MC-4 complexes reported to date, only a few comprise the 3d/4f motif with salicylhydroxamic acid (shiH3) [23,24,25]. If we further restrict the above requirements, by exclusively using MnIII as the periphery ring metal ion, only a very few papers have been published featuring the above qualifications. Pecoraro and Zaleski reported a complex with the general formula being LnIIIMI(OAc)4[12-MC-MnIII(N)shi-4](H2O)4·6DMF where MI = NaI and KI, and Ln = various lanthanides [26,27,28]. This family of compounds has been extensively structurally studied but the authors did focus their investigations mainly towards the effect that the NaI or KI ions had on magnetic measurements and not on the pure {Mn4IIILnIII(µ-NO)4}11+ magnetic core, without emphasizing on the evaluation of exchange interactions by fitting of the data. Our group has a great interest in the synthesis and magnetic characterization of MCs and so far has dealt with the isolation of homometallic and heterometallic ones based on transition metal ions [29,30,31,32]. Herein, we report the synthesis, crystal structures and magnetic studies of a rare family of isostructural (tBu4N){[LnIII(O2CBut)4][12-MC-Mn(III)N(shi)-4]}·5CH2Cl2 (Ln = Gd (1) and Tb (2)) and a (tBu4N)2{[YIII(O2CBut)4][12-MC-Mn(III)N(shi)-4]}·(ClO4) (3) compound. This is an unprecedented example of Ln(III)[12-MC-Mn(III)N(shi)-4] without the presence of any alkali or alkaline earth metals.

2. Results and Discussion

2.1. Crystal Structures of Compounds 13

The general reaction of Mn(O2CBut)2·2H2O, M(NO3)3·xH2O (M = GdIII, TbIII, YIII), shaH2, tBu4NClO4 and morpholine, in a 4:1:4:1:4 molar ratio, in CH2Cl2 gave dark brown solutions which were layered with hexanes to give dark brown crystals of (tBu4N){[LnIII(O2CBut)4][12-MC-Mn(III)N(shi)-4]}·5CH2Cl2 (or (tBu4N)[MnIII4LnIII(O2CBut)4(shi)4]·5CH2Cl2) for 1 and 2 and (tBu4N)2{[YIII(O2CBut)4][12-MC-Mn(III)N(shi)-4]}·(ClO4)·((tBu4N)2[MnIII4LnIII(O2CBut)4(shi)4]·(ClO4)) for 3 in high yields (>59%). The chemical and structural identities of the compounds were confirmed by single-crystal X-ray crystallography, elemental analyses (C, H, N) and IR spectral data (Supplementary Materials).
Single-crystal diffraction studies revealed that compounds 1 and 2 are isostructural and crystallize in the P4/n tetragonal space group, while complex 3 crystallizes in the P4cc tetragonal space group (Table S1). Although complex 3 is not isostructural with complexes 1 and 2, still it does possess the same core with them and thus, only complex 2 will be thoroughly described for simplicity reasons. The structure of 2 consists of a [TbIII(O2CBut)4][12-MC-Mn(III)N(shi)-4]} anion (Figure 1), one tBu4N+ cation and, five non-coordinated CH2Cl2 molecules. Its asymmetric unit features one-quarter of the [TbIII(O2CBut)4][12-MC-Mn(III)N(shi)-4]} anion, with the C4 axis passing through the central TbIII ion. There are two isomers in the structure and only the main part will be discussed. Selected interatomic distances and angles for all complexes are listed in Tables S2 and S3. The core of 2 comprises four MnIII and one TbIII atoms arranged in a square pyramidal-like topology with the Tb atom occupying the apical position of the pyramid and the Mn atoms completing the base (Figure S1).
The basal MnIII atoms are bridged by diatomic oximate bridges provided by the shi3− ligand, giving a Mn···Mn separation of 4.642(9) Å. Note that the coordinated anion of shi3− was generated in situ from the metal ion-assisted transformation of shaH2 under basic conditions. The large Mn–N–O–Mn torsion angle (173.3(4)°), which is very close to the ideal linearity of 180°, is responsible for the approximately ideal planarity of the Mn4 assembly, whilst the TbIII atom lies 1.789(8) Å out of the Mn4 plane. The connection between the basal MnIII and the TbIII atoms is provided by the oximate O atoms of the shi3− ligand resulting in a Mn…Tb separation of 3.739(2) Å. Further ligation is provided by four η11:μ bridging pivalate ions (Figure 1). The coordination sphere around the Mn atoms is completed by the alkoxido and phenoxido O atoms provided by the organic moiety which possesses a η11123 coordination mode leading to an overall inorganic core of {Mn4IIILnIII(µ-NO)4}11+ (Figure S1).
All MnIII ions are five-coordinate with almost perfect square pyramidal geometry. This has been confirmed by the analysis of the shape-determining bonds and angles using the Reedjik and Addison et al. method [33], which gives us a trigonality index, τ, of 0.11 for the four MnIII ions. The τ = 0.1 value is consistent with a square pyramidal geometry, as for a perfect square pyramidal geometry a τ value of 0 is expected, while a τ value of 1 is consistent for a trigonal bipyramidal geometry. The oxidation states of Mn atoms were established by charge balance considerations, metric parameters and bond-valence sum (BVS) calculations [34], with the last providing us with a value of 3.03 for Mn1. Note that the oxidation of MnII to MnIII occurs undoubtedly by the atmospheric O2 under the prevailing basic conditions [35,36]. The central lanthanide ion is eight-coordinate possessing a slightly distorted square antiprismatic geometry with a continuous shape measurement factor of CShM = 0.71 (Figure S2) [37]. The closest this number is to zero, the closest is the geometry of the lanthanide to the ideal one. Finally, the crystal packing of complex 2 (Figure S2) has revealed that the ionic compounds do not communicate which each other (well separated) in the crystal by any means such as hydrogen bonding or π–π stacking interactions but are solely surrounded by the tetrabutylammonium cations and the solvate molecules in the lattice.
Several pivotal geometrical parameters were obtained for complex 2 in order to gain a better understanding of the inner coordination sphere around the lanthanide ion (Figure 2). To be more descriptive, the angle between the four-fold axis and the Ln–O bond direction, θ, corresponds to compression or elongation along the tetragonal axis, depending on its value. The magic value for perfect square-antiprismatic (SAP) geometry is θ = 54.74°, while smaller angles correspond to elongation and wider ones lead to compression [38,39,40]. In complex 2, the average value of θ was found to be 56.15°, indicating axial compression. The distance between the upper and lower O4-planes, interplanar distance (dpp), was found to be 2.639(0) Å, while the distances d1 and d2 were found to be 1.103(4) Å and 1.535(6) Å, counting for the plane spanned by the carboxylate oxygen atom (O5) and the one by the oximate oxygen (O3), respectively. The symmetry of lanthanide’s coordination geometry can be further described by another important parameter which is called skew or twist angle. This is the φ angle which basically defines the angle between the diagonals of the two O4-planes. This is a vital parameter for the determination of point group symmetry at the lanthanide site, which consequently leads to assisting on the description of the crystal field substrate composition of lanthanide complexes. When φ = 0 an ideal square prismatic geometry is expected, while when φ = 45° an ideal square antiprismatic geometry is observed. In complex 2 an average φ value of 43.28(4)° was calculated, further supporting the square antiprismatic geometry of the TbIII ion.

2.2. Magnetic Studies of Complexes 13

Solid state, direct-current (dc) magnetic susceptibility (χM) measurements were collected in the temperature range of 2.0–300 K for freshly prepared crystalline samples of 1, 2, and 3, under an applied field of 0.1 T. The obtained data are presented as a χMT vs. T plot in Figure 3. The experimental values at 300 K for all complexes (18.1 cm3·mol−1·K for 1, 21.1 cm3·mol−1·K for 2 and 11.4 cm3·mol−1·K for 3) are lower than the theoretical ones (19.88 cm·mol−1 K for 1, 23.82 cm3·mol−1·K for 2 and 12.0 cm3·mol−1·K for 3) expected for four non-interacting MnIII ions (S = 2, g = 2) and one GdIII (8S7/2, S = 7/2, L = 0, g = 2), one TbIII (7F6, S = 3, L = 3, g = 3/2), or one diamagnetic YIII ion [41]. All complexes possess a similar magnetic behavior, with the χMT steadily decreasing with decreasing temperature from 300 K till 2 K, where it reaches values of 6.95 cm3·K·mol−1 for 1, 3.30 cm3·K·mol−1 for 2 and 0.2 cm3·K·mol−1 for 3, respectively. The shape of the χMT vs. T plots for all complexes indicates the presence of predominant antiferromagnetic exchange interactions within the metal centers. This is further supported by the fact that the χMT values at 300 K for all complexes are lower than the expected theoretical ones.
In order to gain better insight in the strength of the intramolecular MnIII–MnIII magnetic exchange interactions, the magnetic susceptibility data of complex 3 were fit using PHI [42] program. The magnetic susceptibility data of complex 3, which comprises the diamagnetic YIII ion in the central cavity, were fit using a 1-J model according to the spin Hamiltonian:
H ^ =   2 J   ( S ^ Mn 1 · S ^ Mn 2   +   S ^ Mn 2 · S ^ Mn 3   +   S ^ Mn 3 · S ^ Mn 4   +   S ^ Mn 4 · S ^ Mn 1 )
An excellent simulation of the data (solid green line, Figure 3) was achieved with a J = −3.74 cm−1 and g = 2.07. The antiferromagnetic exchange interactions are anticipated for a system that is exclusively coupled via oximate bridges that give very large Mn–N–O–Mn torsion angles, which are known to promote antiferromagnetic exchange interactions [43].
In order to also quantify the nature of the MnIII–LnIII exchange interactions, complex 1 was also fit, using a 2-J model, which counts for the outer MnIII–MnIII (J1) interactions as well as for inner the MnIII–GdIII (J2) interaction, shown in Figure 3b. The data were fit according to the spin Hamiltonian shown below:
H ^ =   2 J 1   ( S ^ Mn 1 · S ^ Mn 2   +   S ^ Mn 2 · S ^ Mn 3   +   S ^ Mn 3 · S ^ Mn 4   +   S ^ Mn 4 · S ^ Mn 1 )     2 J 2 ( S ^ Mn 1 · S ^ Gd   +   S ^ Mn 2 · S ^ Gd   +   S ^ Mn 3 · S ^ Gd   +   S ^ Mn 4 · S ^ Gd )
An excellent fit of the data (solid green line, Figure 3) could be obtained with J1 = −3.35 cm−1, J2 = −0.45 cm−1 and gMn(III) = 1.99. The values of both models are in an excellent agreement. Note that antiferromagnetic exchange interactions, in MnIII–GdIII and other 3d–GdIII species, possessing the 3dx2−y2 orbital unoccupied, are quite often observed [44,45,46,47]. Here is the first time that fitting of the magnetic susceptibility data for 3d/GdIII interactions has been reported, within the MC family of complexes.
Field-dependent magnetization measurements were also performed for complexes 1 and 2 at temperatures between 2 K and 10 K over the range of 0–7 T (Figure 4 and Figure S8). The magnetization of 1 and 2 shows a rapid increase below 1 T followed by a slow, nearly linear increase without reaching saturation. The lack of saturation in magnetization of 1 and 2 indicates the presence of magnetic anisotropy and/or population of the LnIII low-lying excited states, as well as the effect from some weak antiferromagnetic components between the metal centers.

3. Materials and Methods

3.1. General Information

All chemicals and solvents used for synthesis were of reagent grade and used as purchased without further purification. The starting material Mn(O2CBut)2·2H2O was synthesized using literature procedures [48]. C, H and N elemental analyses were carried out on a Foss Heraeus Vario EL (Elementar Analysensysteme GmbH, Langenselbold, Germany) at the Institute of Organic Chemistry at the Johannes Gutenberg University Mainz. Infrared absorption spectra were recorded at room temperature in a range of 3000–400 cm−1 on a Thermo Fischer NICOLET Nexus FT/IR-5700 spectrometer (Thermo Fischer Scientific, Waltham, MA, USA) equipped with Smart Orbit ATR Diamond cell (Thermo Fischer Scientific, Waltham, MA, USA). UV–Vis absorption measurements were performed between for complexes 1, 2 and 3 in MeCN between 200 mm and 1000 nm on a JASCO V-570 UV/Vis/NIR spectrophotometer (JASCO Inc., Easton, MD, USA) (Figure S7 in Supplementary Materials).
A similar procedure has been used to isolate compounds 13.
(tBu4N){[GdIII(O2CBut)4][12-MC-Mn(III)N(shi)-4]}·5CH2Cl2 (1·5CH2Cl2): To an almost colorless solution of shiH3 (30.50 mg, 0.2 mmol) and morpholine (18 µL, 0.2 mmol) in CH2Cl2 was added Mn(O2CBut)2·2H2O (55.00 mg, 0.2 mmol) followed by stirring for 5 min. To the resulting dark brown almost clear solution Gd(NO3)3·H2O (6.00 mg, 0.025 mmol) was added along with tBu4NClO4 (26.00 mg, 0.075 mmol) and left for stirring for another 40 min. The solution was subsequently filtered and left for crystallization. Layering hexane gave diffraction quality crystals of 1·5CH2Cl2 after 5 days which were collected by filtration, washed with hexanes (3 × 5 mL) and dried in air. Yield: 0.032 g (59%) based on the GdIII ion. The air-dried solid was analyzed as (1): C, 47.32; H, 5.46; N, 4.3. Found: C, 47.44; H, 5.49; N, 4.38. Selected ATR data (cm−1): 2961 (w), 2292 (w), 2875 (w), 1598 (w), 1569 (s), 1538 (w), 1421 (w), 1096 (m), 867 (w), 768 (w), 683 (s), 649 (w), 617 (m), 482 (m).
(tBu4N){[TbIII(O2CBut)4][12-MC-Mn(III)N(shi)-4]}·5CH2Cl2 (2·5CH2Cl2): Compound 2·5CH2Cl2 was synthesized with the similar procedure as compound 1·5CH2Cl2, except that Tb(NO3)3·H2O (8.50 mg, 0.025 mmol) was used instead of Gd(NO3)3·H2O. Yield: 0.039 g (76%) based on the TbIII ion. The air-dried solid was analyzed as (2): C, 47.27; H, 5.46; N, 4.31. Found: C, 47.38; H, 5.52; N, 4.34. Selected ATR data (cm−1): 2961 (w), 2929 (w), 2874 (w), 1597 (w), 1568 (s), 1537 (w), 1421 (w), 1099 (m), 865 (w), 771 (w), 721 (s), 649 (w), 683 (s), 600 (m), 482 (m).
(tBu4N)2{[YIII(O2CBut)4][12-MC-Mn(III)N(shi)-4]}·(ClO4) (3): Compound 3 was synthesized with the similar procedure as compounds 1 and 2, except that this time Y(NO3)3·H2O (0.01 mg, 0.025 mmol) was used. Yield: 0.028 g (59%) based on the YIII ion. The air-dried solid was analyzed as (3): C, 50.63; H, 6.59; N, 4.43. Found: C, 50.74; H, 6.71; N, 4.51. Selected ATR data (cm−1): 2961 (w), 2929 (w), 2874 (w), 1596 (w), 1569 (m), 1424 (w), 1099 (m), 865 (w), 771 (w), 684 (s), 649 (w), 600 (m), 482 (m).

3.2. X-ray Crystallography

X-ray diffraction data for the structure analysis were collected from suitable single crystals on a Bruker SMART with an APEX II CCD detector (1, 2) (Bruker AXS GmbH, Karlsruhe, Germany) and on a STOE IPDS 2 T (3) (STOE & Cie GmbH, Darmstadt, Germany) equipped with an Oxford cooling system (Oxford Cryosystems Ltd., Oxford, UK) operating at 173(2) K (1, 2) and at 120(2)K (3), respectively. Graphite-monochromated Mo-Kα radiation (λ = 0.71073 Å) from long-fine focus sealed X-ray tube was used throughout. Data reduction and absorption correction were done with Bruker Apex v3.0 [49,50] and SADABS[×11] (1, 2) or with STOE X-RED [51] (3). Structures were solved with SHELXT [52] and refined by full-matrix least-squares on F-squared using SHELXL [53], interfaced through Olex2 [54]. All non-hydrogen atoms were refined with anisotropic displacement parameters, while all hydrogen atoms have been placed on idealized positions using a riding model. In complexes 1, 2 and 3 the anionic (metallacrown) part show isomerism and are disordered over two positions. The metallacrowns can be arranged clockwise [M–NO–M] or anticlockwise [M–ON–M], with slightly different position for the transition metal ions, while the central lanthanide ion remains on its position in both isomers. The site occupation factor of the isomers were refined free to 0.82/0.18, 0.80/0.20 and 0.91/0.09 for 1, 2 and 3, respectively. While in 1 and 2 the whole anionic part was refined over two positions, in 3 only the manganese ions were refined over two positions, due to the low occupancy of the second isomer. The cationic counter ions Bu4N+ were refined over two positions with a fixed ratio of 0.6/0.4 in 1 and 2. CCDC 1849727–1849729 (13) contains the supplementary crystallographic data for the structure reported in this paper.

3.3. Magnetic Measurements

Variable-temperature direct current (dc) magnetic susceptibility measurements were performed on polycrystalline samples with the use of Quantum Design SQUID magnetometer MPMS-7 equipped with a 7 T magnet. The samples were embedded in eicosane to avoid orientation of the crystallites under the applied field. Experimental susceptibility data were corrected for the underlying diamagnetism using Pascal’s constants [55]. The temperature dependent magnetic contribution of the holder and of the embedding matrix eicosane were experimentally determined and subtracted from the measured susceptibility data. Variable temperature susceptibility data were collected in a temperature range of 2–300 K under an applied field of 0.1 Tesla, while magnetization data were collected between 2 K and 10 K and magnetic fields up to 7 Tesla. Alternating-current (ac) measurements were performed with an oscillating magnetic field of 3 Oe at frequencies ranging from 1 Hz to 1400 Hz.

4. Conclusions

In summary, we reported a new family of MnIII/LnIII 12-MC-4 complexes, derived from the reaction of Mn(O2CBut)2·2H2O with various nitrate salts of lanthanides in the presence of salicylhydroxamic acid. Direct-current (dc) magnetic susceptibility studies revealed the presence of antiferromagnetic exchange interactions between the metal centers, while fitting of the data using the {Mn4Y} complex allowed us to quantify the strength of the interactions within the outer MnIII ions, which was found to be J = −3.74 cm−1 with g = 2.07. Moreover, fitting of the {Mn4Gd} (1) data gave an extra insight into the strength of the magnetic exchange interactions, especially for the MnIII–GdIII intramolecular interaction, revealing values of JMn–Mn = −3.35 cm−1, JMn–Gd = −0.45 cm−1 and gMn(III) = 1.99. Note that this is the first example within the family of metallacrowns (MCs), where simulation of the magnetic data has been reported. In-phase and out-of-phase (ac) magnetic susceptibility measurements as a function of temperature did not reveal any slow relaxation in fields of Hdc = 0–3000 Oe. In order to further improve the magnetic properties of such compounds, the chemistry will be broadened to the use of other magnetic or diamagnetic 3d and 4f metal ions, by means of modifying the structural and/or physical properties of the resulting molecular compounds.

Supplementary Materials

Supplementary materials are available online at https://www.mdpi.com/2304-6740/6/3/66/s1, CIF and checkCIF files of compounds 1, 2 and 3, Figure S1: Coordination modes of ligands in complex 2, Figure S2: Labeled schematic representation of the core {Mn4IIILnIII(µ-NO)4}11+ of complex 2. Color scheme: Tb, yellow; MnIII, blue; N, green; O, red, Figure S3: Crystal packing representation in complex 2, Figure S4: IR spectrum for complex 1, Figure S5: IR spectrum for complex 2, Figure S6: IR spectrum for complex 3, Figure S7: UV–Vis spectra of 1 (black), 2 (red), 3 (blue), and shiH3 (green) in MeCN, Figure S8: M vs. H plots for complex 1 in various temperatures as indicated. Solid lines are guidelines for the eyes. Table S1: Crystallographic date for complexes 13, Table S2: Selected bond Lengths for complex 2, Table S3: Selected Bond Angles for 2, Table S4: Shape measurements of the 8-coordinate lanthanide coordination polyhedra. The bold numbers indicate the closest polyhedron according to SHAPE calculations.

Author Contributions

A.A.A. designed and performed the experiments. She analyzed most of the data and wrote the paper. L.M.C. refined the structures and helped a lot with his knowledgeable input throughout the paper-writing process. E.R. was involved during the writing of the paper to every step of the process giving valuable feedback.

Funding

This research received no external funding.

Acknowledgments

We are very greateful to Regine Jung-Pothmann and Dieter Schollmeyer for the collection of the X-ray diffraction data of all our complexes. Angeliki A. Athanasopoulou is a member of (SFB/TRR) 173 “Spin+X–Spin its collective environment” and a fellow of the Excellence Initiative by the Graduate School Materials Science in Mainz, Germany (DFG/GSC 266), both initiated by the Deutsche Forschungsgemeinschaft (DFG, GermanResearchFoundation).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chow, C.Y.; Trivedi, E.R.; Pecoraro, V.; Zaleski, C.M. Heterometallic Mixed 3d-4f Metallacrowns: Structural Versatility, Luminescence, and Molecular Magnetism. Comments Inorg. Chem. 2015, 35, 214–253. [Google Scholar] [CrossRef]
  2. Jia, R.; Li, H.-F.; Chen, P.; Gao, T.; Sun, W.-B.; Li, G.-M.; Yan, P.-F. Synthesis, Structure, and Tunable White Light Emission of Heteronuclear Zn2 Ln2 Arrays Using a Zinc Complex as Ligand. CrystEngComm 2016, 18, 917–923. [Google Scholar] [CrossRef]
  3. Evangelisti, F.; Moré, R.; Hodel, F.; Luber, S.; Patzke, G.R. 3d4f {CoII3Ln(OR)4} Cubanes as Bio-Inspired Water Oxidation Catalysts. J. Am. Chem. Soc. 2015, 137, 11076–11084. [Google Scholar] [CrossRef] [PubMed]
  4. Rosado Piquer, L.; Sañudo, E.C. Heterometallic 3d4f Single-Molecule Magnets. Dalton Trans. 2015, 44, 8771–8780. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  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. Christou, G.; Gatteschi, D.; Hendrickson, D.N.; Sessoli, R. Single-Molecule Magnets. MRS Bull. 2000, 25, 66–71. [Google Scholar] [CrossRef]
  7. Bagai, R.; Christou, G. The Drosophila of Single-Molecule Magnetism: [Mn12O12(O2CR)16(H2O)4]. Chem. Soc. Rev. 2009, 38, 1011. [Google Scholar] [CrossRef] [PubMed]
  8. Friedman, J.R.; Sarachik, M.P.; Tejada, J.; Ziolo, R. Macroscopic Measurement of Resonant Magnetization Tunneling in High-Spin Molecules. Phys. Rev. Lett. 1996, 76, 3830–3833. [Google Scholar] [CrossRef] [PubMed]
  9. Thomas, L.; Lionti, F.; Ballou, R.; Gatteschi, D.; Sessoli, R.; Barbara, B. Macroscopic Quantum Tunnelling of Magnetization in a Single Crystal of Nanomagnets. Nature 1996, 383, 145–147. [Google Scholar] [CrossRef]
  10. Bogani, L.; Wernsdorfer, W. Molecular Spintronics Using Single-Molecule Magnets. Nat. Mater. 2008, 7, 179–186. [Google Scholar] [CrossRef] [PubMed]
  11. Urdampilleta, M.; Klyatskaya, S.; Cleuziou, J.-P.; Ruben, M.; Wernsdorfer, W. Supramolecular Spin Valves. Nat. Mater. 2011, 10, 502–506. [Google Scholar] [CrossRef] [PubMed]
  12. Lah, M.S.; Kirk, M.L.; Hatfield, W.; Pecoraro, V.L. The Tetranuclear Cluster FeIII[FeIII (Salicylhydroximato)(MeOH)(Acetate)]3 Is an Analogue of M3+(9-Crown-3). J. Chem. Soc. Chem. Commun. 1989, 1606–1608. [Google Scholar] [CrossRef]
  13. Lah, M.S.; Pecoraro, V.L. Isolation and Characterization of {MnII[MnIII(Salicylhydroximate)]4(Acetate)2(DMF)6}·2DMF: An Inorganic Analog of M2+(12-Crown-4). J. Am. Chem. Soc. 1989, 111, 7258–7259. [Google Scholar] [CrossRef]
  14. Dutton, J.C.; Murray, K.S.; Tiekink, E.R.T. Magnetism of Oxovanadium(IV) Complexes of Binucleating Ligands. Oxidation to and Structure of a Mononuclear Oxovanadium(V) Complex of N,N′-(Pentan-3-Ol) Bis(Salicylaldimine). Inorg. Chim. Acta 1989, 166, 5–8. [Google Scholar] [CrossRef]
  15. Mezei, G.; Zaleski, C.M.; Pecoraro, V.L. Structural and Functional Evolution of Metallacrowns. Chem. Rev. 2007, 107, 4933–5003. [Google Scholar] [CrossRef] [PubMed]
  16. Tegoni, M.; Furlotti, M.; Tropiano, M.; Lim, C.S.; Pecoraro, V.L. Thermodynamics of Core Metal Replacement and Self-Assembly of Ca2+ 15-Metallacrown-5. Inorg. Chem. 2010, 49, 5190–5201. [Google Scholar] [CrossRef] [PubMed]
  17. Jones, L.F.; Kilner, C.A.; Halcrow, M.A. A Cobalt Metallacrown Anion Host with Guest-Dependent Redox Activity. Chem. Eur. J. 2009, 15, 4667–4675. [Google Scholar] [CrossRef] [PubMed]
  18. Jones, L.F.; Barrett, S.A.; Kilner, C.A.; Halcrow, M.A. Ammonium, Alkylammonium, and Amino Acid Complexes of a Hexacopper Fluoro-Metallacrown Cavitand. Chem. Eur. J. 2008, 14, 223–233. [Google Scholar] [CrossRef] [PubMed]
  19. Chow, C.Y.; Eliseeva, S.V.; Trivedi, E.R.; Nguyen, T.N.; Kampf, J.W.; Petoud, S.; Pecoraro, V.L. Ga3+/Ln3+ Metallacrowns: A Promising Family of Highly Luminescent Lanthanide Complexes That Covers Visible and Near-Infrared Domains. J. Am. Chem. Soc. 2016, 138, 5100–5109. [Google Scholar] [CrossRef] [PubMed]
  20. Trivedi, E.R.; Eliseeva, S.V.; Jankolovits, J.; Olmstead, M.M.; Petoud, S.; Pecoraro, V.L. Highly Emitting Near-Infrared Lanthanide “Encapsulated Sandwich” Metallacrown Complexes with Excitation Shifted Toward Lower Energy. J. Am. Chem. Soc. 2014, 136, 1526–1534. [Google Scholar] [CrossRef] [PubMed]
  21. Ostrowska, M.; Fritsky, I.O.; Gumienna-Kontecka, E.; Pavlishchuk, A.V. Metallacrown-Based Compounds: Applications in Catalysis, Luminescence, Molecular Magnetism, and Adsorption. Coord. Chem. Rev. 2016, 327–328, 304–332. [Google Scholar] [CrossRef]
  22. Athanasopoulou, A.A.; Gamer, C.; Völker, L.; Rentschler, E. Novel Magnetic Nanostructures: Unique Properties and Applications; ELSEVIER: New York, NY, USA, 2018. [Google Scholar]
  23. Qin, Y.; Gao, Q.; Chen, Y.; Liu, W.; Lin, F.; Zhang, X.; Dong, Y.; Li, Y. Four Mixed 3d-4f 12-Metallacrown-4 Complexes: Syntheses, Structures and Magnetic Properties. J. Clust. Sci. 2017, 28, 891–903. [Google Scholar] [CrossRef]
  24. Cao, F.; Wang, S.; Li, D.; Zeng, S.; Niu, M.; Song, Y.; Dou, J. Family of Mixed 3d–4f Dimeric 14-Metallacrown-5 Compounds: Syntheses, Structures, and Magnetic Properties. Inorg. Chem. 2013, 52, 10747–10755. [Google Scholar] [CrossRef] [PubMed]
  25. Lou, T.; Yang, H.; Zeng, S.; Li, D.; Dou, J. A New Family of Heterometallic LnIII[12-MCFeIIIN(Shi)-4] Complexes: Syntheses, Structures and Magnetic Properties. Crystals 2018, 8, 229. [Google Scholar] [CrossRef]
  26. Azar, M.R.; Boron, T.T.; Lutter, J.C.; Daly, C.I.; Zegalia, K.A.; Nimthong, R.; Ferrence, G.M.; Zeller, M.; Kampf, J.W.; Pecoraro, V.L.; et al. Controllable Formation of Heterotrimetallic Coordination Compounds: Systematically Incorporating Lanthanide and Alkali Metal Ions into the Manganese 12-Metallacrown-4 Framework. Inorg. Chem. 2014, 53, 1729–1742. [Google Scholar] [CrossRef] [PubMed]
  27. Travis, J.R.; Zeller, M.; Zaleski, C.M. Facile Carboxylate Ligand Variation of Heterotrimetallic 12-Metallacrown-4 Complexes. Polyhedron 2016, 114, 29–36. [Google Scholar] [CrossRef]
  28. Boron, T.T.; Lutter, J.C.; Daly, C.I.; Chow, C.Y.; Davis, A.H.; Nimthong-Roldán, A.; Zeller, M.; Kampf, J.W.; Zaleski, C.M.; Pecoraro, V.L. The Nature of the Bridging Anion Controls the Single-Molecule Magnetic Properties of DyX4M 12-Metallacrown-4 Complexes. Inorg. Chem. 2016, 55, 10597–10607. [Google Scholar] [CrossRef] [PubMed]
  29. Happ, P.; Rentschler, E. Enforcement of a High-Spin Ground State for the First 3d Heterometallic 12-Metallacrown-4 Complex. Dalton Trans. 2014, 43, 15308–15312. [Google Scholar] [CrossRef] [PubMed]
  30. Happ, P.; Plenk, C.; Rentschler, E. 12-MC-4 Metallacrowns as Versatile Tools for SMM Research. Coord. Chem. Rev. 2015, 289–290, 238–260. [Google Scholar] [CrossRef]
  31. Plenk, C.; Krause, J.; Beck, M.; Rentschler, E. Rational Linkage of Magnetic Molecules Using Click Chemistry. Chem. Commun. 2015, 51, 6524–6527. [Google Scholar] [CrossRef] [PubMed]
  32. Happ, P.; Sapozhnik, A.; Klanke, J.; Czaja, P.; Chernenkaya, A.; Medjanik, K.; Schuppler, S.; Nagel, P.; Merz, M.; Rentschler, E.; et al. Analyzing the Enforcement of a High-Spin Ground State for a Metallacrown Single-Molecule Magnet. Phys. Rev. B 2016, 93, 174404. [Google Scholar] [CrossRef]
  33. Addison, A.W.; Rao, T.N.; Reedijk, J.; van Rijn, J.; Verschoor, G.C. Synthesis, Structure, and Spectroscopic Properties of Copper(II) Compounds Containing Nitrogen–sulphur Donor Ligands; the Crystal and Molecular Structure of Aqua[1,7-Bis(N-Methylbenzimidazol-2′-Yl)-2,6-Dithiaheptane]Copper(II) Perchlorate. J. Chem. Soc. Dalton Trans. 1984, 1349–1356. [Google Scholar] [CrossRef]
  34. Liu, W.; Thorp, H.H. Bond Valence Sum Analysis of Metal-Ligand Bond Lengths in Metalloenzymes and Model Complexes. 2. Refined Distances and Other Enzymes. Inorg. Chem. 1993, 32, 4102–4105. [Google Scholar] [CrossRef]
  35. Alaimo, A.A.; Koumousi, E.S.; Cunha-Silva, L.; McCormick, L.J.; Teat, S.J.; Psycharis, V.; Raptopoulou, C.P.; Mukherjee, S.; Li, C.; Gupta, S.D.; et al. Structural Diversities in Heterometallic Mn–Ca Cluster Chemistry from the Use of Salicylhydroxamic Acid: {MnIII4Ca2}, {MnII/III6Ca2}, {MnIII/IV8Ca}, and {MnIII8Ca2} Complexes with Relevance to Both High- and Low-Valent States of the Oxygen-Evolving Complex. Inorg. Chem. 2017, 56, 10760–10774. [Google Scholar] [CrossRef] [PubMed]
  36. Stoumpos, C.C.; Gass, I.A.; Milios, C.J.; Lalioti, N.; Terzis, A.; Aromí, G.; Teat, S.J.; Brechin, E.K.; Perlepes, S.P. A MnII4 Cubane and a Novel MnII10MnIII4 Cluster from the Use of Di-2-Pyridyl Ketone in Manganese Acetate Chemistry. Dalton Trans. 2009, 307–317. [Google Scholar] [CrossRef] [PubMed]
  37. Alvarez, S.; Alemany, P.; Casanova, D.; Cirera, J.; Llunell, M.; Avnir, D. Shape Maps and Polyhedral Interconversion Paths in Transition Metal Chemistry. Coord. Chem. Rev. 2005, 249, 1693–1708. [Google Scholar] [CrossRef]
  38. Sorace, L.; Benelli, C.; Gatteschi, D. Lanthanides in Molecular Magnetism: Old Tools in a New Field. Chem. Soc. Rev. 2011, 40, 3092. [Google Scholar] [CrossRef] [PubMed]
  39. Baldoví, J.J.; Cardona-Serra, S.; Clemente-Juan, J.M.; Coronado, E.; Gaita-Ariño, A.; Palii, A. Rational Design of Single-Ion Magnets and Spin Qubits Based on Mononuclear Lanthanoid Complexes. Inorg. Chem. 2012, 51, 12565–12574. [Google Scholar] [CrossRef] [PubMed]
  40. Sørensen, M.A.; Weihe, H.; Vinum, M.G.; Mortensen, J.S.; Doerrer, L.H.; Bendix, J. Imposing High-Symmetry and Tuneable Geometry on Lanthanide Centres with Chelating Pt and Pd Metalloligands. Chem. Sci. 2017, 8, 3566–3575. [Google Scholar] [CrossRef]
  41. 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]
  42. Chilton, N.F.; Anderson, R.P.; Turner, L.D.; Soncini, A.; Murray, K.S. PHI: A Powerful New Program for the Analysis of Anisotropic Monomeric and Exchange-Coupled Polynuclear d- and f-Block Complexes. J. Comput. Chem. 2013, 34, 1164–1175. [Google Scholar] [CrossRef] [PubMed]
  43. Milios, C.J.; Piligkos, S.; Brechin, E.K. Ground State Spin-Switching via Targeted Structural Distortion: Twisted Single-Molecule Magnets from Derivatised Salicylaldoximes. Dalton Trans. 2008, 1809–1817. [Google Scholar] [CrossRef] [PubMed]
  44. Papatriantafyllopoulou, C.; Abboud, K.A.; Christou, G. Carboxylate-Free MnIII2LnIII2 (Ln = Lanthanide) and MnIII2YIII2 Complexes from the Use of (2-Hydroxymethyl)pyridine: Analysis of Spin Frustration Effects. Inorg. Chem. 2011, 50, 8959–8966. [Google Scholar] [CrossRef] [PubMed]
  45. Savva, M.; Skordi, K.; Fournet, A.D.; Thuijs, A.E.; Christou, G.; Perlepes, S.P.; Papatriantafyllopoulou, C.; Tasiopoulos, A.J. Heterometallic MnIII4Ln2 (Ln = Dy, Gd, Tb) Cross-Shaped Clusters and Their Homometallic MnIII4MnII2 Analogues. Inorg. Chem. 2017, 56, 5657–5668. [Google Scholar] [CrossRef] [PubMed]
  46. Liu, S.-J.; Xie, X.-R.; Zheng, T.-F.; Bao, J.; Liao, J.-S.; Chen, J.-L.; Wen, H.-R. Three-Dimensional Two-Fold Interpenetrated CrIII–GdIII Heterometallic Framework as an Attractive Cryogenic Magnetorefrigerant. CrystEngComm 2015, 17, 7270–7275. [Google Scholar] [CrossRef]
  47. Gómez, V.; Vendier, L.; Corbella, M.; Costes, J.-P. Tetranuclear [Co–Gd]2 Complexes: Aiming at a Better Understanding of the 3d-Gd Magnetic Interaction. Inorg. Chem. 2012, 51, 6396–6404. [Google Scholar] [CrossRef] [PubMed]
  48. Yu, S.B.; Lippard, S.J.; Shweky, I.; Bino, A. Dinuclear Manganese(II) Complexes with Water and Carboxylate Bridges. Inorg. Chem. 1992, 31, 3502–3504. [Google Scholar] [CrossRef]
  49. Bruker Apex 3; Bruker AXS Inc.: Madison, WI, USA, 2016.
  50. Sheldrick, G.M. SADABS-2016/2; Bruker AXS Inc.: Madison, WI, USA, 2016. [Google Scholar]
  51. Stoe & Cie X-RED; Stoe & Cie: Darmstadt, Germany, 2002.
  52. Sheldrick, G.M. SHELXT—Integrated Space-Group and Crystal-Structure Determination. Acta Crystallogr. Sect. Found. Adv. 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  53. Sheldrick, G.M. Crystal Structure Refinement with SHELXL. Acta Crystallogr. Sect. C Struct. Chem. 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  54. 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]
  55. Bain, G.A.; Berry, J.F. Diamagnetic Corrections and Pascal’s Constants. J. Chem. Educ. 2008, 85, 532. [Google Scholar] [CrossRef]
Scheme 1. Illustrative representation and abbreviation of organic molecules discussed in the text.
Scheme 1. Illustrative representation and abbreviation of organic molecules discussed in the text.
Inorganics 06 00066 sch001
Figure 1. Schematic representation of molecular structure and labeled schematic representation of the core {Mn4IIILnIII(µ-NO)4}11+ of complex 2. Color scheme: Tb, yellow; MnIII, blue; N, green; O, red. H atoms are omitted for clarity.
Figure 1. Schematic representation of molecular structure and labeled schematic representation of the core {Mn4IIILnIII(µ-NO)4}11+ of complex 2. Color scheme: Tb, yellow; MnIII, blue; N, green; O, red. H atoms are omitted for clarity.
Inorganics 06 00066 g001
Figure 2. Details of structural parameters discussed in the text for complex 2. Yellow ball: terbium, red ball: oxygen.
Figure 2. Details of structural parameters discussed in the text for complex 2. Yellow ball: terbium, red ball: oxygen.
Inorganics 06 00066 g002
Figure 3. (a) Temperature dependence of magnetic susceptibility for complexes 1, 2, and 3. Green solid line represents simulation of the data in complexes 1 and 3; see text for details and fitting parameters. (b) Fitting model for complex 1.
Figure 3. (a) Temperature dependence of magnetic susceptibility for complexes 1, 2, and 3. Green solid line represents simulation of the data in complexes 1 and 3; see text for details and fitting parameters. (b) Fitting model for complex 1.
Inorganics 06 00066 g003
Figure 4. M vs. H plots for complex 2 in various temperatures as indicated. Solid lines are guidelines for the eyes.
Figure 4. M vs. H plots for complex 2 in various temperatures as indicated. Solid lines are guidelines for the eyes.
Inorganics 06 00066 g004

Share and Cite

MDPI and ACS Style

Athanasopoulou, A.A.; Carrella, L.M.; Rentschler, E. Synthesis, Structural, and Magnetic Characterization of a Mixed 3d/4f 12-Metallacrown-4 Family of Complexes. Inorganics 2018, 6, 66. https://doi.org/10.3390/inorganics6030066

AMA Style

Athanasopoulou AA, Carrella LM, Rentschler E. Synthesis, Structural, and Magnetic Characterization of a Mixed 3d/4f 12-Metallacrown-4 Family of Complexes. Inorganics. 2018; 6(3):66. https://doi.org/10.3390/inorganics6030066

Chicago/Turabian Style

Athanasopoulou, Angeliki A., Luca M. Carrella, and Eva Rentschler. 2018. "Synthesis, Structural, and Magnetic Characterization of a Mixed 3d/4f 12-Metallacrown-4 Family of Complexes" Inorganics 6, no. 3: 66. https://doi.org/10.3390/inorganics6030066

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