Next Article in Journal
Recent Advances of Cellulase Immobilization onto Magnetic Nanoparticles: An Update Review
Next Article in Special Issue
Diversity of Coordination Modes in a Flexible Ditopic Ligand Containing 2-Pyridyl, Carbonyl and Hydrazone Functionalities: Mononuclear and Dinuclear Cobalt(III) Complexes, and Tetranuclear Copper(II) and Nickel(II) Clusters
Previous Article in Journal / Special Issue
Chloranilato-Based Layered Ferrimagnets with Solvent-Dependent Ordering Temperatures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Novel Family of Triangular CoII2LnIII and CoII2YIII Clusters by the Employment of Di-2-Pyridyl Ketone †

by
Constantinos G. Efthymiou
1,
Áine Ní Fhuaráin
1,
Júlia Mayans
2,
Anastasios Tasiopoulos
3,
Spyros P. Perlepes
4 and
Constantina Papatriantafyllopoulou
1,*
1
SSPC, Synthesis and Solid State Pharmaceutical Centre, School of Chemistry, National University of Ireland Galway, University Road, H91 TK33 Galway, Ireland
2
Departament de Quimica Inorgànica i Orgànica, Secció Inorgànica and Institut de Nanociencia I, Nanotecnologia (IN2UB), Universitat de Barcelona, Marti i Franquès 1-11, 08028 Barcelona, Spain
3
Department of Chemistry, University of Cyprus, 1678 Nicosia, Cyprus
4
Department of Chemistry, University of Patras, 26504 Patras, Greece
*
Author to whom correspondence should be addressed.
This paper is dedicated to Professor Masahiro Yamashita, a leading and inspiring Scientist on Molecular Magnetism, on the occasion of his 65th birthday.
Magnetochemistry 2019, 5(2), 35; https://doi.org/10.3390/magnetochemistry5020035
Submission received: 31 March 2019 / Revised: 22 May 2019 / Accepted: 24 May 2019 / Published: 4 June 2019

Abstract

:
The synthesis, structural characterization and magnetic study of novel CoII/4f and CoII/YIII clusters are described. In particular, the initial employment of di-2-pyridyl ketone, (py)2CO, in mixed metal Co/4f chemistry, provided access to four triangular clusters, [CoII2MIII{(py)2C(OEt)(O)}4(NO3)(H2O)]2[M(NO3)5](ClO4)2 (M = Gd, 1; Dy, 2; Tb, 3; Y, 4), where (py)2C(OEt)(O) is the monoanion of the hemiketal form of (py)2CO. Clusters 14 are the first reported Co/4f (13) and Co/Y (4) species bearing (py)2CO or its derivatives, despite the fact that over 200 metal clusters bearing this ligand have been reported so far. Variable-temperature, solid-state dc and ac magnetic susceptibility studies were carried out on 14 and revealed the presence of weak ferromagnetic exchange interactions between the metal ions (JCo-Co = +1.3 and +0.40 cm−1 in 1 and 4, respectively; JCo-Gd = +0.09 cm−1 in 1). The ac susceptibility studies on 2 revealed nonzero, weak out-of-phase (χ’’M) signals below ~5 K.

Graphical Abstract

1. Introduction

The synthesis and characterization of new mixed-metal 3d/4f clusters has attracted immense interest over the last few decades, due to their fascinating structural features (high nuclearities, unprecedented metal topologies, aesthetically pleasing architectures, etc.), as well as due to their interesting magnetic properties [1,2,3]. In particular, 4f ions often favor the formation of heterometallic compounds that possess exceptionally high nuclearities, with representative examples being clusters of Ni64Gd96 [4], Ni76La60 [5], Ni54Gd54 [6], Cu36Dy24 [7], Ni10Gd42 [8], Ni30La20 [9,10], etc. This intriguing ability of 4f ions possibly stems from their strong oxophilicity, which, in combination with their high coordination numbers, results in the formation of hydroxo/oxo species that readily promote the aggregation process. Concerning the magnetic properties of the 3d/4f compounds, the 4f ions bring several advantages, such as their considerable number of unpaired electrons (e.g., Gd3+ has seven unpaired e−) and their large single ion anisotropy (e.g., Tb3+, Dy3+, Ho3+, etc.) as a result of their orbital angular momentum. The above properties make them ideal candidates for the synthesis of heterometallic clusters with single-molecule magnetism behavior (SMMs) [11,12], fulfilling the desirable features for a compound to behave as an SMM, namely (i) high spin ground state (S) and (ii) negative axial zero field splitting parameter (D). SMMs are discrete metal compounds that exhibit superparamagnetic behavior below a blocking temperature TB and have been proposed for several technological applications including high-density information storage, molecular spintronics and qubits for quantum computation [11,12,13,14,15].
The study of mixed-metal 3d/4f reaction systems, as a means for the isolation of new SMMs with a high energy barrier for the magnetization reversal, has led to a large variety of such species that now include Mn/4f [16,17,18,19,20,21], Fe/4f [22,23,24,25], Ni/4f [8,26], Cu/4f [27,28,29,30] and Co/4f [8,31,32,33] compounds [1,2]. It is noteworthy that the majority of 3d/4f SMMs are Mn/4f clusters containing some MnIII centers with an S = 2 spin state and a significant uniaxial anisotropy. Some remarkable examples, e.g., a Mn6Tb2 [18] and a Mn21Dy [17] cluster, display high energy barriers for the magnetization reversal (Ueff = 103 K and 74 K, respectively), which are of comparable magnitude to the family of the most thoroughly studied homometallic carboxylato Mn12 SMMs [11]. On the other hand, the reported Co/4f SMMs are significantly less, despite the fact that the combination of the anisotropic 3d7 CoII with the 4f ions has a great potential to yield SMMs with high Ueff and distinctively different properties from other heterometallic species. A possible explanation for this could be related to synthetic challenges such as the oxidation of the Co2+ to the diamagnetic and low spin Co3+, which occurs easily in the presence of a base under ambient conditions.
Many carboxylate and O or N,O-ligands have been used for the synthesis of 3d/4f metal clusters [1,2,3]; amongst them, di-2-pyridyl ketone ((py)2CO, Scheme 1) is very attractive as its carbonyl group can easily undergo nucleophilic attack, providing a wide range of hemiaketal and gem-diol derivatives that are able to link many metal ions, favoring the formation of high nuclearity metal clusters with interesting structural features and magnetic properties [34,35]. Over 200 homo- and heterometallic compounds have now been reported, containing (py)2CO and its derivatives, thus the absence of such Co/4f clusters is noticeable considering the great development of this research field.
In 2010, our groups employed (py)2CO in Ni/4f cluster chemistry and reported a new family of triangular Ni2Ln compounds with interesting magnetic properties [36,37]. Expanding this research, we herein report the synthesis and characterization of four isostructural triangular Co2M clusters (M=Tb, Dy, Gd, Y); these compounds are the first examples of Co/4f or Y species bearing (py)2CO and its derivatives, with the Co2Dy analogue displaying out-of-phase ac magnetic susceptibility signals at low temperatures, indicative of the slow relaxation of the magnetization.

2. Results and Discussion

2.1. Synthetic Comments

We have developed an intense interest in the synthesis of 3d/4f metal clusters by the employment of various pyridyl oximate- and alkoxide-containing ligands; these research efforts have yielded a variety of new mixed-metal species with interesting structural features and magnetic properties, including Ni8Dy8 [38,39], Ni2Ln2 [40], Ni3Ln [26], Ni2Ln [36,37,40], Mn4Ln2 [41], etc. Restricting further discussion to the use of (py)2CO in this field, we recently reported the first Mn/4f compounds, which belong to a family of cross-shaped Mn4Ln2 clusters, where some of them exhibit slow relaxation of magnetization; whereas, in the past, we reported the first Ni/4f compounds with the monoanionic form of (py)2CO. Wishing to expand this work, we recently decided to investigate the previously unexplored reaction system of Co2+/Ln3+/(py)2CO.
The reaction of Co(ClO4)2·6H2O, Ln(NO3)3·6H2O (Ln=Gd, 1; Dy, 2; Tb, 3) or Y(NO3)3·6H2O (4), (py)2CO and CH3CO2Na·3H2O in EtOH afforded a red solution from which well-shaped red crystals of compounds 14 with the general formula [Co2M{(py)2C(OEt)(O)}4(NO3)(H2O)]2[M(NO3)5](ClO4)2 were subsequently isolated. The formation of 14 is summarized in Equation (1).
EtOH            4   Co ( ClO 4 ) 2 · 6 H 2 O   +   3 M ( NO 3 ) 3 · 6 H 2 O   +   8 ( py ) 2 CO   +   8 NaO 2 CMe · 3 H 2 O   +   8 EtOH      [ C o 2 M { ( py ) 2 C ( OEt ) ( O ) } 4 ( N O 3 ) ( H 2 O ) ] 2 [ M ( N O 3 ) 5 ] ( Cl O 4 ) 2 + 8 MeC O 2 H +   2 NaN O 3 +   6 NaCl O 4 +   66 H 2 O           Gd , 1 ;   Dy , 2 ;   Tb , 3 ;   Y , 4
The nature of the base and the crystallization method are not crucial for the identity of the products and affect only their crystallinity and the reaction yield; we were able to isolate 14 (IR evidence) by using other bases, such as NaOMe, NaOEt, LiOH·H2O, etc. On the other hand, the ratio of the reactants and the nature of solvent affect the product identity, as by further increasing the excess of (py)2CO, mononuclear CoII compounds are isolated. EtOH is the only solvent that favors the formation of 14, whereas the use of different solvents yields amorphous products that could not be further characterized.

2.2. Description of Structures

A representation of the cationic [Co2Gd{(py)2C(OEt)(O)}4(NO3)(H2O)]2+ that is present in the molecular structure of 1 is shown in Figure 1. A representation of the elipsoid plot for 1 is shown in Figure S1 in the supplementary material. Selected interatomic distances and angles for 1 are listed in Table 1.
Complex 1 crystallizes in the monoclinic space group C 2/c. Its structure consists of two isostructural triangular cationic clusters [Co2Gd{(py)2C(OEt)(O)}4(NO3)(H2O)]2+, which are symmetrically related with a 2-fold crystallographic axis. The positive charge of the cation is balanced by one [Gd(NO3)5]2− and two NO3 counterions. The cationic cluster is comprised of two Co2+ and one Gd3+ ions, which are held together by four (py)2C(OEt)(O) ligands. The {Co2GdO4}3+ core of this complex displays an oxo-centered triangular arrangement, in which one μ3-alkoxo group coming from one (py)2C(OEt)(O) ligand bridges the three metal centres; in addition, three μ2-O2− ions, from three different (py)2C(OEt)(O) ligands, are located peripherally, bridging the two metal ions in each edge of the triangle. Alternatively, the structural core in 1 can be described as a defective cubane, in which one vertex and three edges are missing. The central μ3-O2− ion deviates 1.12(2) Å from the plane formed by the metal ions. The intermetallic distances in 1 are Co1…Gd = 3.471 Å, Co2…Gd = 3.546 Å and Co1…Co2 = 3.192 Å.
The monoanionic (py)2C(OEt)(O) ligands are derived from the nucleophilic attack of one EtOH molecule on the central C atom of the carbonyl group of (py2)CO. The three (py)2C(OEt)(O) ligands adopt a η1212 coordination mode, with the fourth one being coordinated to the metals in a η1313 fashion (Scheme 2). The two CoII ions are six-coordinate with their coordination spheres ({O1, O5, O3, N1, N3, N5} for Co1; {O3, O5, O7, N4, N6, N7} for Co2) displaying distorted octahedral geometries. The three O and the three N donor atoms around each CoII ion adopt the facial, fac- topological arrangement; each CoII ion is surrounded by three five-membered chelate rings, formed by three different (py)2C(OEt)(O) ligands. The Co oxidation state was assigned by charge considerations and bond-valence sum (BVS) calculations [42].
Gd1 is eight-coordinate and its {O1, O3, O7, O9, O10, O12, N2, N7} coordination sphere is rich in O donor atoms as a consequence of the oxophilic character of the lanthanides. Its coordination environment is formed by two five-membered chelate rings, the central μ3-O2− ion, one bidentate chelate NO3 ion and one terminal H2O molecule. Gd2 in the [Gd(NO3)5]2− ion is 10-coordinated, surrounded by five bidentate chelating nitrate groups. Gd2 lies on a crystallographic 2-fold axis, which passes through the N11 atom of a NO3 group.
To deduce the coordination polyhedra defined by the donor atoms around Gd1, a comparison of the experimental structural data with the theoretical data for the most common polyhedral structures with eight vertices was performed by means of the program SHAPE [43,44]; a reliable, high-quality fit was not achieved.
Closer inspection of the crystal structure of 1 reveals the absence of strong H-bonding interactions. This might be a result of the very well-separated neighboring Co2Gd units. The shortest metal∙∙∙metal distance between neighboring trinuclear clusters is 10.564 Å (Gd1…Gd1), while the shortest metal∙∙∙metal distance between a trinuclear cluster with a neighboring [Gd(NO3)5]2− anion is 7.491 Å (Gd1…Gd2).
Compounds 24 are isostructural with 1, as confirmed by a comparison of their unit cell dimensions. The identity, purity and stability of these compounds was also studied by powder X-ray diffraction (pxrd) studies (Figure S2 in the Supplementary Material).
Compound 1 and its structural analogues (24) are the first Co/Ln or Y clusters bearing (py)2CO and/or its transformed gem-diol or hemiketal derivatives. They also join the very small family of heterometallic 3d/4f/(py)2CO clusters [26,36,37,41,45,46]; thus, they provide insight into the coordination chemistry of this versatile ligand and unlock the chemical and structural features, which can further lead to the isolation of higher nuclearity heterometallic species.

2.3. Magnetism Studies

Solid-state, variable-temperature dc magnetic susceptibility (χM) data were collected on vacuum-dried microcrystalline samples of complexes 14 in the 2.0–300 K range, and they are shown in Figure 2, top, as χMT vs. T plots. The experimental values for 14 at 300 K are 16.04, 26.86, 23.09 and 5.5 cm3·K·mol−1, respectively, being close to the expected ones for one and a half non-interacting LnIII cations (1, Gd, S = 7/2, L = 0, 8S7/2, g = 2; 2, Dy, S = 5/2, L = 5, 6H15/2, g = 4/3; 3, Tb, S = 3, L = 3, 7F6, g = 3/2; 4, Y, S = 0) and two non-interacting high spin CoII cations (S = 3/2, g = 2) of 15.05, 26.1, 21.5 and 3.8 cm3·K·mol−1, respectively.
The study of the static magnetic properties of highly anisotropic LnIII cations with high-spin CoII ions (S = 3/2) within the same molecule is challenging because both types of paramagnetic centers present spin-orbit contribution due to the strong orbital contribution to the magnetic moment; this yields high anisotropies, which prevent the use of spin-only Hamiltonians for the mathematical interpretation and fitting of the experimental curves [47,48]. Although L is not fully quenched, spin-only Hamiltonians are used to fit the curves for practical reasons, where feasible, in the reported compounds.
For complex 4, the χMT vs. T curve remains almost constant with the decreasing temperature from 300 to 50 K and then drops to 3.7 cm3·K·mol−1. This complex contains a diamagnetic YIII, which allows the study of the interaction between the Co(II) ions using the spin Hamiltonian H = −2J (ŜCo1·ŜCo2) + DŜz2 + Σ i μ Β g e f f H S ^ i in the full range of temperature; the exchange interactions between the CoII ions are weak ferromagnetic with J = +0.40 cm−1, D = 9.5 cm−1 and g = 2.35.
The χMT vs. T curve for 1 remains almost constant until 20 K and then starts to increase, reaching the value of 20.01 cm3·K·mol−1 at 2 K, which shows an extremely weak ferromagnetic coupling between the metal ions. The fitting of the experimental data to the Hamiltonian equation H = −2J(ŜCo1 ŜGd + ŜCo2 ŜGd)−2J’(ŜCo1 ŜCo2) + DŜz2 + Σ i μ Β g e f f H S ^ i , in the whole temperature range, provided the coupling values between CoII-CoII ions (J = +1.3 cm−1) and CoII-GdIII ions (J = +0.09 cm−1), respectively, with a mean g value of 2.35. This magnetic coupling is in agreement with previous studies in CoII-GdIII complexes, which always present a ferromagnetic coupling when the CoII is a high spin cation [49,50].
Complexes 2 and 3 exhibit a similar magnetic behavior to that of complex 1, with a very smooth drop while cooling down due to the depopulation of the Stark sublevels, reaching a minimum (21.88 cm3·K·mol−1 for 2; 18.99 cm3·K·mol−1for 3) at 12 K.
The field dependence of the magnetization at 2 K for complexes 14 is shown in Figure 2, bottom. For complexes 13, the magnetization increases rapidly below 1 T. For 4, magnetization presents a value of 3.8 cm3·K·mol−1, which corresponds to the value of two ferromagnetically coupled CoII cations at 2 K (S = 1/2 for each one). For 13, the values of the magnetization at 5 T are 13.5, 10.9 and 10.9 μβ, respectively.
The study of the dynamic magnetic properties was also performed for all compounds under a zero magnetic field, revealing a clear dependency of the χM’’ on temperature and frequency for complex 2 (Figure S3, Supplementary Material), indicating that 2 might be an extremely weak SMM.

3. Materials and Methods

3.1. Materials, Physical and Spectroscopic Measurements

All manipulations were performed under aerobic conditions using materials (reagent grade) and solvents as received. Elemental analyses (C, H, N) were performed by the University of Patras microanalysis service. IR spectra (4000–400 cm−1) were recorded using a Perkin Elmer 16PC FT-IR spectrometer with samples prepared as KBr pellets. Powder X-ray diffraction data (pxrd) were collected using an Inex Equinox 6000 diffractometer. Solid-state, variable-temperature and variable-field magnetic data were collected on powdered samples using an MPMS5 Quantum Design magnetometer operating at 0.03 T in the 300–2.0 K range for the magnetic susceptibility and at 2.0 K in the 0–5 T range for the magnetization measurements. Diamagnetic corrections were applied to the observed susceptibilities using Pascal’s constants. Alternating current (ac) magnetic susceptibility experiments were carried out at 1000 Hz.

3.2. Synthesis of [Co2Gd{(py)2C(OEt)(O)}4(NO3)(H2O)]2[Gd(NO3)5](ClO4)2 (1)

Solid (py)2CO (0.111 g, 0.60 mmol) and NaO2CMe·3H2O (0.041 g, 0.30 mmol) were added to a pink solution of Co(ClO4)2·6H2O (0.110 g, 0.30 mmol) in EtOH (15 mL) under stirring, yielding a red solution. Gd(NO3)3·6H2O (0.046 g, 0.10 mmol) was then added and the resulting solution was stirred for 30 min. The red solution was allowed to stand undisturbed in a closed flask. Red prismatic crystals appeared after 2 days, which were collected by filtration, washed with EtOH (2 × 5 mL) and Et2O (2 × 5 mL) and dried in air. Yield: ~65%. Anal. Calcd (Found) for 1: C, 38.91 (38.80); H, 3.39 (3.72); N, 10.03 (9.73) %. Selected IR data (KBr, cm−1): 3390 (s,b), 2972 (m), 2928 (w), 2897 (w), 1602 (m), 1568 (w), 1470 (s), 1441 (m), 1384 (s), 1317 (s), 1222 (m), 1090 (s), 1053 (s), 903 (w), 777 (m), 686 (m), 635 (m), 624 (m), 541 (w), 474 (m).

3.3. Synthesis of [Co2Dy{(py)2C(OEt)(O)}4(NO3)(H2O)]2[Dy(NO3)5](ClO4)2 (2)

This was prepared in the same manner as complex 1 but using Dy(NO3)3·6H2O (0.046 g, 0.10 mmol) in place of Gd(NO3)3·6H2O. After 2 days, red prismatic crystals of 2 appeared, which were collected by filtration, washed with EtOH (2 × 5 mL) and Et2O (2 × 5 mL) and dried in air. Yield: ~60%. Anal. Calcd (Found) for 2: C, 38.72 (38.91); H, 3.37 (3.75); N, 9.99 (10.08) %. Selected IR data (KBr, cm−1): 3394 (s,b), 2974 (m), 2930 (w), 2897 (w), 1604 (m), 1570 (w), 1472 (s), 1443 (m), 1384 (s), 1315 (s), 1225 (m), 1090 (s), 1054 (s), 904 (w), 780 (m), 686 (m), 635 (m), 625 (m), 542 (w), 472 (m).

3.4. Synthesis of [Co2Tb{(py)2C(OEt)(O)}4(NO3)(H2O)]2[Tb(NO3)5](ClO4)2 (3)

This was prepared in the same manner as complex 1 but using Tb(NO3)3·6H2O (0.046 g, 0.10 mmol) in place of Gd(NO3)3·6H2O. After 2 days, red prismatic crystals of 3 appeared, which were collected by filtration, washed with EtOH (2 × 5 mL) and Et2O (2 × 5 mL) and dried in air. Yield: ~65%. Anal. Calcd (Found) for 3: C, 38.85 (38.73); H, 3.39 (2.99); N, 10.02 (9.84) %. Selected IR data (KBr, cm−1): v = 3394 (s,b), 2974 (m), 2930 (w), 2896 (w), 1604 (m), 1570 (w), 1472 (s), 1442 (m), 1384 (s), 1316 (s), 1224 (m), 1089 (s), 1054 (s), 904 (w), 780 (m), 686 (m), 636 (m), 626 (m), 542 (w), 474 (m).

3.5. Synthesis of [Co2Y{(py)2C(OEt)(O)}4(NO3)(H2O)]2[Y(NO3)5](ClO4)2 (4)

This was prepared in the same manner as complex 1 but using Y(NO3)3·6H2O (0.038 g, 0.10 mmol) in place of Gd(NO3)3·6H2O. After 2 days, red prismatic crystals of 4 appeared, which were collected by filtration, washed with EtOH (2 × 5 mL) and Et2O (2 × 5 mL) and dried in air. Yield: ~65%. Anal. Calcd (Found) for 4: C, 41.56 (41.47); H, 3.62 (3.53); N, 10.72 (11.09) %. Selected IR data (KBr, cm−1): 3390 (s,b), 2972 (m), 2928 (w), 2897 (w), 1602 (m), 1568 (w), 1470 (s), 1441 (m), 1384 (s), 1317 (s), 1222 (m), 1090 (s), 1053 (s), 903 (w), 777 (m), 686 (m), 635 (m), 624 (m), 541 (w), 474 (m).
Caution! Although no such behavior was observed during the present work, perchlorate and nitrate salts are potentially explosive; such compounds should be synthesized and used in small quantities, and treated with utmost care at all times.

3.6. Single-Crystal X-ray Crystallography

Data were collected at the University of Cyprus on an Oxford-Diffraction SuperNova diffractometer, equipped with a CCD area detector and a graphite monochromator utilizing Mo-Kα radiation (λ = 0.71073 Å). Suitable crystals were attached to glass fiber using paratone-N oil and transferred to a goniostat, where they were cooled to 100 K for data collection. Empirical absorption corrections (multi-scan based on symmetry-related measurements) were applied using CrysAlis RED software [51]. The structure was solved by direct methods using SIR92 [52] and refined on F2 vai the full-matrix least squares method using SHELXL97 [53] and SHELXL-2014/7 [54]. Software packages used are listed as follows: CrysAlisCC for data collection, CrysAlisRED for cell refinement and data reduction [51], WINGX for geometric calculations [55], DIAMOND [56] and MERCURY [57] for molecular graphics. The program SQUEEZE [58], a part of the PLATON package of crystallographic software, was used to remove the contribution of highly disordered solvent molecules. The non-H atoms were treated anisotropically, whereas the H atoms were placed in calculated, ideal positions and refined as riding on their respective C atoms. Unit cell parameters and structure solution and refinement data for 1 are listed in Table S1. An initial search of reciprocal space for 24 revealed monoclinic cells with dimensions similar to those of 1; thus, full data collection of their structures was not pursued.
Several crystals of compound 1, from different preparations and at different periods of time, were carefully tested on the X-rays (using CuKa and MoK radiation) at ambient and low (100 K) temperatures. The diffraction quality of the crystals proved to be moderate and structure determination was eventually carried out by means of the best data set collected. It is important to mention that compound 1 has a unit cell and structure similar to a Ni2Gd analogous compound, as previously reported by us [36,37], though the latter differs mainly in the nature of the 3d metal ion, i.e., it contains NiII instead of CoII; thus, although the crystallographic data are not of the best quality, the information they provide about the structure is absolutely reliable.
The X-ray crystallographic data for 1 have been deposited with a CCDC reference number CCDC 1906734. They can be obtained free of charge at http://www.ccdc.cam.ac.uk/conts/retrieving.html or from the Cambridge Crystallographic Data Center, 12 Union Road, Cambridge, CB2 1EZ, UK: Fax: +44-1223-336033; or e-mail: [email protected].

4. Conclusions

Four new mixed-metal CoII2Ln (Ln = Gd, 1; Dy, 2; Tb, 3) and CoII2Y (4) clusters are described, bearing the anionic hemiaketalic form of di-2-pyridyl ketone as an organic ligand. Compounds 14 display a triangular metal topology and were synthesized by the reaction of Co(ClO4)2·6H2O, M(NO3)3·6H2O, (py)2CO and CH3CO2Na·3H2O in EtOH. They are the first heterometallic Co/4f or Y clusters containing (py)2CO or its derivatives, and join a very small family of such compounds with this ligand. dc and ac magnetic susceptibility studies revealed the presence of weak ferromagnetic exchange interactions between the metal ions, with 2 exhibiting nonzero, weak out-of-phase (χ’’M) signals at temperatures below ~5 K.
(py)2CO remains a rich wellspring of new metal clusters with interesting structural features and magnetic properties, after many years of intense research efforts that have yielded a massive number of compounds. Further studies on the use of this ligand for the synthesis of new 3d/4f metal clusters are in progress and will be reported in due course.

Supplementary Materials

The following are available online at https://www.mdpi.com/2312-7481/5/2/35/s1: Figure S1. Representation of the elipsoid plot for 1, Figure S2. Theoretical and experimental pxrd patterns for 14, Figure S3: Representation of χ (black line) and χ’’ (red line) for 2, Figure S4: Linear fit of the ac magnetic suscetibility data for 2 at the frequency of 1000 Hz using the generalized Debye model to extract the slow relaxation parameters, Table S1: Crystallographic data for complex 1.

Author Contributions

C.G.E. contributed to the synthesis, crystallization and preliminary characterization of all the compounds, and he wrote the relevant draft of the paper. Á.N.F. contributed to the synthesis of 14. J.M. performed the magnetic measurements, interpreted the results and wrote the relevant part of the paper. A.T. contributed to the structural characterization of 14. S.P.P. contributed to the coordination of the research and interpretation of the results. C.P. contributed to the coordination of the research, collected single crystal X-ray crystallographic data and solved the structure of 1, and performed refinement in the structure. She also wrote the paper based on the reports of her collaborators.

Funding

J.M. thanks the Ministerio de Economía y Competitividad, Project CTQ2015-63614-P for funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rosado Piquer, L.; Sañudo, E.C. Heterometallic 3d–4f single-molecule magnets. Dalton Trans. 2015, 44, 8771–8780. [Google Scholar] [CrossRef] [PubMed]
  2. Liu, K.; Shia, W.; Cheng, P. Toward heterometallic single-molecule magnets: Synthetic strategy, structures and properties of 3d–4f discrete complexes. Coord. Chem. Rev. 2015, 289–290, 74–122. [Google Scholar] [CrossRef]
  3. Sharples, J.W.; Collison, D. The coordination chemistry and magnetism of some 3d–4f and 4f amino-polyalcohol compounds. Coord. Chem. Rev. 2014, 260, 1–20. [Google Scholar] [CrossRef] [PubMed]
  4. Chen, W.P.; Liao, P.Q.; Yu, Y.; Zheng, Z.; Chen, X.-M.; Zheng, Y.Z. A Mixed-Ligand Approach for a Gigantic and Hollow Heterometallic Cage {Ni64RE96} for Gas Separation and Magnetic Cooling Applications. Angew. Chem. Int. Ed. 2016, 55, 9375–9379. [Google Scholar] [CrossRef] [PubMed]
  5. Kong, X.J.; Long, L.-S.; Huang, R.B.; Zheng, L.-S.; Harris, T.D.; Zheng, Z. A four-shell, 136-metal 3d-4f heterometallic cluster approximating a rectangular parallelepiped. Chem. Commun. 2009, 4354–4356. [Google Scholar] [CrossRef]
  6. 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]
  7. Leng, J.-D.; Liu, J.-L.; Tong, M.-L. Unique nanoscale {CuII36LnIII24} (Ln = Dy and Gd) metallo-rings. Chem. Commun. 2012, 48, 5286–5288. [Google Scholar] [CrossRef]
  8. Peng, J.-B.; Zhang, Q.-C.; Kong, X.-L.; Zheng, Y.-Z.; Ren, Y.-P.; Long, L.-S.; Huang, R.-B.; Zheng, L.-S.; Zheng, Z. High-Nuclearity 3d−4f Clusters as Enhanced Magnetic Coolers and Molecular Magnets. J. Am. Chem. Soc. 2012, 134, 3314–3317. [Google Scholar] [CrossRef]
  9. Kong, X.J.; Ren, Y.P.; Long, L.-S.; Zheng, Z.; Nichol, G.; Huang, R.-B.; Zheng, L.-S. Dual Shell-like Magnetic Clusters Containing NiII and LnIII (Ln = La, Pr, and Nd) Ions. Inorg. Chem. 2008, 47, 2728–2739. [Google Scholar] [CrossRef]
  10. Kong, X.J.; Ren, Y.P.; Long, L.-S.; Zheng, Z.; Huang, R.-B.; Zheng, L.-S. A Keplerate Magnetic Cluster Featuring an Icosidodecahedron of Ni(II) Ions Encapsulating a Dodecahedron of La(III) Ions. J. Am. Chem. Soc. 2007, 129, 7016–7017. [Google Scholar] [CrossRef]
  11. Bagai, R.; Christou, G. The Drosophila of single-molecule magnetism: [Mn12O12(O2CR)16(H2O)4]. Chem. Soc. Rev. 2009, 38, 1011–1026. [Google Scholar] [CrossRef]
  12. Christou, G.; Gatteschi, D.; Hendrickson, D.N.; Sessoli, R. Single-Molecule Magnets. MRS Bull. 2000, 25, 66–71. [Google Scholar] [CrossRef]
  13. Wernsdorfer, W.; Bogani, L. Molecular spintronics using single-molecule magnets. Nat. Mater. 2008, 7, 179–186. [Google Scholar]
  14. Hill, S.; Edwards, R.S.; Aliaga-Alcalde, N.; Christou, G. Quantum Coherence in an Exchange-Coupled Dimer of Single-Molecule Magnets. Science 2003, 302, 1015–1018. [Google Scholar] [CrossRef] [Green Version]
  15. Tiron, R.; Wernsdorfer, W.; Aliaga-Alcalde, N.; Christou, G. Quantum tunneling in a three-dimensional network of exchange-coupled single-molecule magnets. Phys. Rev. B 2003, 68, 140407. [Google Scholar] [CrossRef] [Green Version]
  16. Akhtar, M.N.; Lan, Y.; AlDamen, M.A.; Zheng, Y.-Z.; Anson, C.E.; Powell, A.K. Effect of ligand substitution on the SMM properties of three isostructural families of double-cubane Mn4Ln2 coordination clusters. Dalton Trans. 2018, 47, 3485–3495. [Google Scholar] [CrossRef]
  17. Papatriantafyllopoulou, C.; Wernsdorfer, W.; Abboud, K.A.; Christou, G. Mn21Dy Cluster with a Record Magnetization Reversal Barrier for a Mixed 3d/4f Single-Molecule Magnet. Inorg. Chem. 2011, 50, 421–423. [Google Scholar] [CrossRef]
  18. Holynska, M.; Premuzic, D.; Jeon, I.-R.; Wernsdorfer, W.; Clérac, R.; Dehnen, S. [MnIII6O3Ln2] Single-Molecule Magnets: Increasing the Energy Barrier Above 100 K. Chem. Eur. J. 2011, 17, 9605–9610. [Google Scholar] [CrossRef]
  19. Stamatatos, T.C.; Teat, S.J.; Wernsdorfer, W.; Christou, G. Enhancing the Quantum Properties of Manganese-Lanthanide Single-Molecule Magnets: Observation of Quantum Tunneling Steps in the Hysteresis Loops of a {Mn12Gd} Cluster. Angew. Chem. Int. Ed. 2009, 48, 521–524. [Google Scholar] [CrossRef]
  20. Mereacre, V.; Ako, A.M.; Clerac, R.; Wernsdorfer, W.; Filoti, G.; Bartolome, 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]
  21. Mereacre, V.; Ako, A.M.; Clerac, R.; Wernsdorfer, W.; Hewitt, I.J.; Anson, C.E.; Powell, A.K. Heterometallic [Mn5-Ln4] Single-Molecule Magnets with High Anisotropy Barriers. Chem. Eur. J. 2008, 14, 3577–3584. [Google Scholar] [CrossRef]
  22. Li, H.; Meng, X.; Wang, M.; Wang, Y.-X.; Shi, W.-S.; Cheng, P. A {Tb2Fe3} Pyramid Single-Molecule Magnet with Ferromagnetic Tb-Fe Interaction. Chin. J. Chem. 2019, 37, 373–377. [Google Scholar] [CrossRef]
  23. Baniodeh, A.; Liang, Y.; Anson, C.E.; Magnani, N.; Powell, A.K.; Unterreiner, A.N.; Seyfferle, S.; Slota, M.; Dressel, M.; Bogani, L.; et al. Unraveling the Influence of Lanthanide Ions on Intra- and Inter-Molecular Electronic Processes in Fe10Ln10 Nano-Toruses. Adv. Funct. Mater. 2014, 24, 6280–6290. [Google Scholar] [CrossRef]
  24. Badia-Romano, L.; Bartolomé, F.; Bartolomé, J.; Luzón, J.; Prodius, D.; Turta, C.; Mereacre, V.; Wilhelm, F.; Rogalev, A. Field-induced internal Fe and Ln spin reorientation in butterfly {Fe3LnO2} (Ln = Dy and Gd) single-molecule magnets. Phys. Rev. B 2013, 87, 184403:1–184403:11. [Google Scholar] [CrossRef]
  25. Schmidt, S.; Prodius, D.; Mereacre, V.; Kostakis, G.E.; Powell, A.K. Unprecedented chemical transformation: Crystallographic evidence for 1,1,2,2-tetrahydroxyethane captured within an Fe6Dy3 single molecule magnet. Chem. Commun. 2013, 49, 1696–1698. [Google Scholar] [CrossRef]
  26. Efthymiou, C.G.; Stamatatos, T.C.; Papatriantafyllopoulou, C.; Tasiopoulos, A.J.; Wernsdorfer, W.; Perlepes, S.P.; Christou, G. Nickel/Lanthanide Single-Molecule Magnets: {Ni3Ln} “Stars” with a Ligand Derived from the Metal-Promoted Reduction of Di-2-pyridyl Ketone under Solvothermal Conditions. Inorg. Chem. 2010, 49, 9737–9739. [Google Scholar] [CrossRef]
  27. Zhu, Q.; Xiang, S.; Sheng, T.; Yuan, D.; Shen, C.; Tan, C.; Hua, S.; Wu, X. A series of goblet-like heterometallic pentanuclear [LnIIICuII4] clusters featuring ferromagnetic coupling and single-molecule magnet behavior. Chem. Commun. 2012, 48, 10736–10738. [Google Scholar] [CrossRef]
  28. Ghosh, S.; Ida, Y.; Ishida, T.; Ghosh, A. Linker Stoichiometry-Controlled Stepwise Supramolecular Growth of a Flexible Cu2Tb Single Molecule Magnet from Monomer to Dimer to One-Dimensional Chain. Cryst. Growth Des. 2014, 14, 2588–2598. [Google Scholar] [CrossRef]
  29. Jose Heras Ojea, M.; Milway, V.A.; Velmurugan, G.A.; Thomas, L.H.; Coles, S.J.; Wilson, C.; Wernsdorfer, W.; Rajaraman, G.; Murrie, M. Enhancement of TbIII–CuII Single-Molecule Magnet Performance through Structural Modification. Chem. Eur. J. 2016, 22, 12839–12848. [Google Scholar] [CrossRef]
  30. Baskar, V.; Gopal, K.; Helliwell, M.; Tuna, F.; Wernsdorfer, W.; Winpenny, R.E.P. 3d–4f Clusters with large spin ground states and SMM behavior. Dalton Trans. 2010, 39, 4747–4750. [Google Scholar] [CrossRef]
  31. Li, X.-L.; Min, F.-Y.; Wang, C.; Lin, S.-Y.; Liu, Z.; Yang, J. Utilizing 3d-4f magnetic interaction to slow the magnetic relaxation of heterometallic complexes. Inorg. Chem. 2015, 54, 4337–4344. [Google Scholar] [CrossRef]
  32. Chandrasekhar, V.; Pandian, B.M.; Azhakar, R.; Vittal, J.J.; Clérac, R. Linear Trinuclear Mixed-Metal CoII−GdIII−CoII Single-Molecule Magnet: [L2Co2Gd][NO3]2CHCl3(LH3 = (S)P[N(Me)N=CH−C6H3-2-OH-3-OMe]3). Inorg. Chem. 2007, 46, 5140–5142. [Google Scholar] [CrossRef]
  33. Funes, A.V.; Carrella, L.; Rentschler, E.; Alborés, P. {CoIII2DyIII2} single molecule magnet with two resolved thermal activated magnetization relaxation pathways at zero field. Dalton Trans. 2014, 43, 2361–2364. [Google Scholar] [CrossRef]
  34. Stamatatos, T.C.; Efthymiou, C.G.; Stoumpos, C.C.; Perlepes, S.P. Adventures in the Coordination Chemistry of Di-2-pyridyl Ketone and Related Ligands: From High-Spin Molecules and Single-Molecule Magnets to Coordination Polymers, and from Structural Aesthetics to an Exciting New Reactivity Chemistry of Coordinated Ligands. Eur. J. Inorg. Chem. 2009, 2009, 3361–3391. [Google Scholar]
  35. Papaefstathiou, G.S.; Perlepes, S.P. Families of Polynuclear Manganese, Cobalt, Nickel and Copper Complexes Stabilized by Various Forms of Di-2-pyridyl Ketone. Comments Inorg. Chem. 2002, 23, 249–274. [Google Scholar] [CrossRef]
  36. Efthymiou, C.G.; Georgopoulou, A.N.; Papatriantafyllopoulou, C.; Terzis, A.; Raptopoulou, C.P.; Escuer, A.; Perlepes, S.P. Initial employment of di-2-pyridyl ketone as a route to nickel(II)/lanthanide(III) clusters: Triangular Ni2Ln complexes. Dalton Trans. 2010, 39, 8603–8605. [Google Scholar] [CrossRef]
  37. Georgopoulou, A.N.; Efthymiou, C.G.; Papatriantafyllopoulou, C.; Psycharis, V.; Raptopoulou, C.P.; Manos, M.; Tasiopoulos, A.; Escuer, A.; Perlepes, S.P. Triangular NiII2LnIII and NiII2YIII complexes derived from di-2-pyridyl ketone: Synthesis, structures and magnetic properties. Polyhedron 2011, 30, 2978–2986. [Google Scholar] [CrossRef]
  38. Papatriantafyllopoulou, C.; Stamatatos, T.C.; Efthymiou, C.G.; Cunha-Silva, L.; Almeida Paz, F.; Perlepes, S.P.; Christou, G. A High-Nuclearity 3d/4f Metal Oxime Cluster: An Unusual Ni8Dy8 “Core-Shell” Complex from the Use of 2-Pyridinealdoxime. Inorg. Chem. 2010, 49, 9743–9745. [Google Scholar] [CrossRef]
  39. Polyzou, C.D.; Efthymiou, C.G.; Escuer, A.; Cunha-Silva, L.; Papatriantafyllopoulou, C.; Perlepes, S.P. In search of 3d/4f-metal single-molecule magnets: Nickel(II)/lanthanide(III) coordination clusters. Pure Appl. Chem. 2013, 85, 315–327. [Google Scholar] [CrossRef]
  40. Papatriantafyllopoulou, C.; Estrader, M.; Efthymiou, C.G.; Dermitzaki, D.; Gkotsis, K.; Terzis, A.; Diaz, C.; Perlepes, S.P. In search for mixed transition metal/lanthanide single-molecule magnets: Synthetic routes to NiII/TbIII and NiII/DyIII clusters featuring a 2-pyridyl oximate ligand. Polyhedron 2009, 28, 1652–1655. [Google Scholar] [CrossRef]
  41. 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]
  42. 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]
  43. Ruiz-Martinez, A.; Casanova, D.; Alvarez, S. Polyhedral Structures with an Odd Number of Vertices: Nine-Coordinate Metal Compounds. Chem. Eur. J. 2008, 14, 1291–1303. [Google Scholar] [CrossRef]
  44. Llunell, M.; Casanova, D.; Girera, J.; Alemany, P.; Alvarez, S. SHAPE, version 2.0; Universitat de Barcelona: Barcelona, Spain, 2010. [Google Scholar]
  45. Georgopoulou, A.N.; Adam, R.; Raptopoulou, C.P.; Psycharis, V.; Ballesteros, R.; Abarca, B.; Boudalis, A.K. Expanding the 3d-4f heterometallic chemistry of the (py)2CO and pyCOpyCOpy ligands: Structural, magnetic and Mössbauer spectroscopic studies of two FeII–GdIII complexes. Dalton Trans. 2011, 40, 8199–8205. [Google Scholar] [CrossRef]
  46. Alexandropoulos, D.I.; Cunha-Silva, L.; Tang, J.; Stamatatos, T.C. Heterometallic Cu/Ln cluster chemistry: Ferromagnetically-coupled {Cu4Ln2} complexes exhibiting single-molecule magnetism and magnetocaloric properties. Dalton Trans. 2018, 47, 11934–11941. [Google Scholar] [CrossRef]
  47. Tang, J.; Zhang, P. Lanthanide Single Molecule Magnets; Springer: Berlin/Heidelberg, Germany, 2015. [Google Scholar]
  48. Lloret, F.; Julve, M.; Cano, J.; Ruiz-García, R.; Pardo, E. Magnetic properties of six-coordinated high-spin cobalt(II) complexes: Theoretical background and its application. Inorg. Chim. Acta 2008, 361, 3432–3445. [Google Scholar] [CrossRef]
  49. Xu, Y.; Luo, F.; Zheng, J.-M. Syntheses, Structures, and Magnetic Properties of a Series of Heterotri-, Tetra- and Pentanuclear LnIII–CoII Compounds. Polymers 2019, 11, 196. [Google Scholar] [CrossRef]
  50. Bartolomé, J.; Filoti, G.; Kuncser, V.; Schinteie, G.; Mereacre, V.; Anson, C.E.; Powell, A.K.; Prodius, D.; Turta, C. Magnetostructural correlations in the tetranuclear series of {Fe3LnO2} butterfly core clusters:Magnetic and Mössbauer spectroscopic study. Phys. Rev. B. 2009, 80, 014430:1–014430:16. [Google Scholar] [CrossRef]
  51. Oxford Diffraction. CrysAlis CCD and CrysAlis RED, version 1.171.32.15; Oxford Diffraction Ltd.: Oxford, UK, 2008. [Google Scholar]
  52. Altomare, A.; Cascarano, G.; Giaconazzo, C.; Guagliardi, A.; Burla, M.C.; Polidori, G.; Camalli, M. SIR92: A program for automatic solution of crystal structures by direct methods. J. Appl. Crystallogr. 1994, 27, 435–436. [Google Scholar] [CrossRef]
  53. Sheldrick, G.M. SHELXL97; University of Göttingen: Göttingen, Germany, 1997. [Google Scholar]
  54. Sheldrick, G.M. SHELXL-2014/7; Program for Refinement of Crystal Structures, University of Göttingen: Göttingen, Germany, 2014. [Google Scholar]
  55. Farrugia, L.J. WinGX suite for small-molecule single-crystal crystallography. J. Appl. Crystallogr. 1999, 32, 837–838. [Google Scholar] [CrossRef]
  56. Brandenburg, K. DIAMOND; Version 3.1d; Crystal Impact GbR: Bonn, Germany, 2006. [Google Scholar]
  57. Macrae, C.F.; Edgington, P.R.; McCabe, P.; Pidcock, E.; Shields, G.P.; Taylor, R.; Towler, M.; Van de Streek, J. Mercury: Visualization and analysis of crystal structures. J. Appl. Crystallogr. 2006, 39, 453–457. [Google Scholar] [CrossRef]
  58. Van der Sluis, P.; Spek, A.L. BYPASS: An Effective Method for the Refinement of Crystal Structures Containing Disordered Solvent Regions. Acta Crystallogr. Sect. A Found Crystallogr. 1990, A46, 194–201. [Google Scholar] [CrossRef]
Scheme 1. A schematic representation of (py)2CO (top) and its transformation to (py)2C(OEt)(OH) (bottom).
Scheme 1. A schematic representation of (py)2CO (top) and its transformation to (py)2C(OEt)(OH) (bottom).
Magnetochemistry 05 00035 sch001
Figure 1. Structure of the cationic cluster 1. The hydrogen atoms and the counteranions were omitted for clarity.
Figure 1. Structure of the cationic cluster 1. The hydrogen atoms and the counteranions were omitted for clarity.
Magnetochemistry 05 00035 g001
Scheme 2. A schematic representation of the coordination modes of (py)2C(OEt)(O) in 1.
Scheme 2. A schematic representation of the coordination modes of (py)2C(OEt)(O) in 1.
Magnetochemistry 05 00035 sch002
Figure 2. χMT vs. T plots (top) and field dependence of the magnetization at 2 K (bottom) for 14. Solid line represents the best fit for 1 and 4.
Figure 2. χMT vs. T plots (top) and field dependence of the magnetization at 2 K (bottom) for 14. Solid line represents the best fit for 1 and 4.
Magnetochemistry 05 00035 g002
Table 1. Selected interatomic distances (Å) and angles (degrees) for 1.
Table 1. Selected interatomic distances (Å) and angles (degrees) for 1.
Gd(1)-O(1)2.252(15)Co(1)-N(5)2.220(20)
Gd(1)-O(3)2.387(15)Co(2)-O(3)2.142(13)
Gd(1)-O(7)2.304(13)Co(2)-O(5)2.048(14)
Gd(1)-N(2)2.582(17)Co(2)-O(7)2.046(14)
Gd(1)-N(8)2.578(19)Co(2)-N(4)2.070(16)
Co(1)-O(1)2.021(15)Co(2)-N(6)2.08(2)
Co(1)-O(3)2.265(15)Co(2)-N(7)2.176(19)
Co(1)-O(5)1.984(16)Gd(1)-Co(1)3.471(3)
Co(1)-N(1)2.083(18)Gd(1)-Co(2)3.546(3)
Co(1)-N(3)2.093(19)Co(1)-Co(2)3.192(4)
Co(1)-O(1)-Gd(1)108.5(6)Co(2)-O(7)-Gd(1)109.1(5)
Co(1)-O(3)-Gd(1)96.5(5)Co(1)-O(3)-Co(2)92.8(5)
Co(2)-O(3)-Gd(1)102.9(6)Co(1)-O(5)-Co(2)104.6(7)

Share and Cite

MDPI and ACS Style

Efthymiou, C.G.; Ní Fhuaráin, Á.; Mayans, J.; Tasiopoulos, A.; Perlepes, S.P.; Papatriantafyllopoulou, C. A Novel Family of Triangular CoII2LnIII and CoII2YIII Clusters by the Employment of Di-2-Pyridyl Ketone. Magnetochemistry 2019, 5, 35. https://doi.org/10.3390/magnetochemistry5020035

AMA Style

Efthymiou CG, Ní Fhuaráin Á, Mayans J, Tasiopoulos A, Perlepes SP, Papatriantafyllopoulou C. A Novel Family of Triangular CoII2LnIII and CoII2YIII Clusters by the Employment of Di-2-Pyridyl Ketone. Magnetochemistry. 2019; 5(2):35. https://doi.org/10.3390/magnetochemistry5020035

Chicago/Turabian Style

Efthymiou, Constantinos G., Áine Ní Fhuaráin, Júlia Mayans, Anastasios Tasiopoulos, Spyros P. Perlepes, and Constantina Papatriantafyllopoulou. 2019. "A Novel Family of Triangular CoII2LnIII and CoII2YIII Clusters by the Employment of Di-2-Pyridyl Ketone" Magnetochemistry 5, no. 2: 35. https://doi.org/10.3390/magnetochemistry5020035

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