Next Article in Journal
Preparation of CD3 Antibody-Conjugated, Graphene Oxide Coated Iron Nitride Magnetic Beads and Its Preliminary Application in T Cell Separation
Next Article in Special Issue
Magnetic and Structural Properties of Organic Radicals Based on Thienyl- and Furyl-Substituted Nitronyl Nitroxide
Previous Article in Journal
Pressureless Sintering of YIG Ceramics from Coprecipitated Nanopowders
Previous Article in Special Issue
New Radical Cation Salts Based on BDH-TTP Donor: Two Stable Molecular Metals with a Magnetic [ReF6]2− Anion and a Semiconductor with a [ReO4] Anion
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

New Cyanido-Bridged Heterometallic 3d-4f 1D Coordination Polymers: Synthesis, Crystal Structures and Magnetic Properties

by
Diana Dragancea
1,2,*,
Ghenadie Novitchi
3,*,
Augustin M. Mădălan
1 and
Marius Andruh
1,*
1
Inorganic Chemistry Laboratory, Faculty of Chemistry, University of Bucharest, Str. Dumbrava Rosie nr. 23, 020464 Bucharest, Romania
2
Institute of Chemistry, Str. Academiei nr. 3, MD 2028 Chișinău, Moldova
3
CNRS, University Grenoble Alpes, LNCMI, F-38000 Grenoble, France
*
Authors to whom correspondence should be addressed.
Submission received: 8 April 2021 / Revised: 21 April 2021 / Accepted: 26 April 2021 / Published: 28 April 2021

Abstract

:
Three new 1D cyanido-bridged 3d-4f coordination polymers, {[Gd(L)(H2O)2Fe(CN)6]·H2O}n (1GdFe), {[Dy(L)(H2O)2Fe(CN)6]·3H2O}n (2DyFe), and {[Dy(L)(H2O)2Co(CN)6]·H2O}n (3DyCo), were assembled following the building-block approach (L = pentadentate bis-semicarbazone ligand resulting from the condensation reaction between 2,6-diacetyl-pyridine and semicarbazide). The crystal structures consist of crenel-like LnIII-MIII alternate chains, with the LnIII ions connected by the hexacyanido metalloligands through two cis cyanido groups. The magnetic properties of the three complexes have been investigated. Field-induced slow relaxation of the magnetization was observed for compounds 2DyFe and 3DyCo. Compound 3DyCo is a new example of chain of Single Ion Magnets.

1. Introduction

The discovery of slow relaxation of the magnetization phenomena for discrete metal complexes (Single Molecule Magnets, SMMs) and 1D coordination polymers (Single Chain Magnets, SCMs) has stimulated the development of an intensive interdisciplinary research field. Beyond their relevance in fundamental Physics and Chemistry, spectacular applications in quantum computing and high-density information storage from these molecules are expected [1]. Although the field of SMMs was initially dominated by transition metal-based-systems, the focus of research shifted to lanthanides, which increase the energy barriers of SMMs [2,3]. The lanthanide ions (especially TbIII, DyIII, and HoIII), bring large magnetic moments and high uniaxial magnetic anisotropy, which are essential prerequisites for the observation of slow relaxation of the magnetization. While SMMs can be mono- and oligonuclear (homo- and heteronuclear) complexes, most of the SCMs are constructed from two different spin carriers (e.g., 3d-3d′, 3d-4f, 2p-3d, and 2p-4f) [4]. Homospin SCMs are rare and rather serendipitously obtained [5,6]. Examples of lanthanide-based Single-Chain Magnets are also limited, and most of them result from the association of lanthanide ions with nitronyl-nitroxide (paramagnetic) ligands [7,8]. Lanthanide ions can be linked to other paramagnetic metal ions through small bridging ligands, which facilitate the exchange interactions. The building-block approach, relying on the employment of metalloligands, represents an excellent strategy to generate heterometallic coordination compounds [9]. Anionic cyanido complexes are very popular in this respect. The self-assembly processes between [M(CN)6]3− metalloligands and cationic LnIII complexes led to a rich variety of structural architectures. The assembling complex cations are either solvated LnIII species or heteroleptic complexes, containing chelating ligands and weakly coordinated anions or solvent molecules, which can be easily replaced by the cyanido bridge. The dimensionality of the resulting coordination polymers is dependent on the number of accessible positions at the lanthanide ions. For example, nitrogen donor blocking ligands attached to LnIII ions, such as 1,10-phenanthroline, 2,2′-bipyridine, 2,2′:6′,2″—terpyridine, 2,4,6-tri(2-pyridyl)-1,3,5-triazine, favor the aggregation of 1D coordination polymers, employing [M(CN)6]3− as metalloligands [10,11,12,13,14,15,16,17,18]. When the reactions between the lanthanide salts and the hexacyanido building block occur in dimethylformamide (DMF) or dimethyl sulfoxide (DMSO), depending on the experimental conditions, discrete species, 1D, 2D, or even 3D coordination polymers have been obtained [19,20,21,22,23,24]. By decreasing the number of cyanido groups within the metalloligand, the formation of low-dimensionality coordination polymers is favored [25]. Most of these 3d-4f cyanido-bridged complexes show interesting physical properties, mainly magnetic [26,27,28] and optical [29,30,31], and in some cases combined slow magnetic relaxation and light emission were revealed [32,33].
In this paper, we report on a new family of 1D coordination polymers, which are assembled from [LnL]3+ and [M(CN)6]3− ions (L = bis-semicarbazone ligand, M = Fe, Co). The pentadentate bis-semicarbazone ligand, L (Scheme 1), was previously used for the synthesis of both 3d-3d/3d-4d cyanido-bridged discrete [34] and polymeric structures [35,36,37,38,39,40,41].
The new compounds have been characterized by single-crystal X-ray diffraction, and their magnetic properties have been investigated.

2. Experimental Section

2.1. Materials and Physical Measurements

All reagents and solvents for synthesis were commercially purchased and used without any further purification. The bis-semicarbazone ligand L was synthesized according to the method reported in the literature [42].
IR spectra were recorded on a FTIR Bruker Tensor V-37 spectrophotometer (KBr pellets) in the range of 4000–400 cm−1. Elemental analysis was performed on a EuroEA Elemental Analyzer.
Magnetic Studies: DC magnetic susceptibility data (2–300 K) were collected on powdered samples using a SQUID magnetometer (Quantum Design MPMS-XL), applying a magnetic field of 0.1 T. All data were corrected for the contribution of the sample holder and the diamagnetism of the samples estimated from Pascal’s constants [43,44]. The field dependence of the magnetization (up to 5 T) was measured between 2.0 and 5.0 K. AC magnetic susceptibility was measured between 2 and 7 K with an oscillating field magnitude of Hac = 3.0 Oe and frequency ranging between 1 and 1488 Hz in presence of a dc field up to Hdc = 4000 Oe. Fitting of the variable parameters and estimation of errors was performed with lsqcurvefit solver in MATLAB, and jacobian matrix was used to generate 95% confidence intervals on the fitted parameters. Typical examples of this analysis are presented in Figures S3–S6.
X-ray powder diffraction data were measured on a Proto AXRD benchtop using Cu-Kα radiation with a wavelength of 1.54059 Å in the range of 5–35º (2θ).

2.2. Single Crystal X-ray Crystallography

X-ray diffraction data were collected at 293 K on a Rigaku XtaLAB Synergy-S diffractometer operating with Mo-Kα (λ = 0.71073 Å) micro-focus sealed X-ray tube. The structures were solved by direct methods and refined by full-matrix least squares techniques based on F2. The non-H atoms were refined with anisotropic displacement parameters. Calculations were performed using SHELX-2014 or SHELX-2018 crystallographic software package [45,46]. Supplementary X-ray crystallographic data in CIF format have been deposited with the CCDC with the following reference numbers: 2069217 (1GdFe), 2069215 (2DyFe), and 2069216 (3DyCo). A summary of the crystallographic data and the structure refinement for crystals 13 are given in Table S1.

2.3. Synthesis of Complexes

The three compounds are synthesized following the same general procedure: LnCl3·6H2O (0.06 mmol) and L (0.06 mmol) in 10 mL H2O were stirred at 80 °C for 30 min. The above-cooled solution was filtered and transferred to a 30 mL test tube. Additional 5 mL of water was layered over the aqueous solution of the mononuclear complexes, and finally, a 10 mL H2O solution containing 0.06 mmol K3[Fe(CN)6] or K3[Co(CN)6] was then slowly layered on top. The whole set up was kept undisturbed and slow diffusion of these two solutions led, after 2 weeks, to single crystals. The reaction mixture was mechanically stirred and was filtered off through frit followed by drying under vacuum to obtain a polycrystalline solid. Single crystals required for the X-ray data collections were picked up from the crystalline mixtures prior to mechanical stirring.
{[Gd(L)(H2O)2Fe(CN)6]·H2O}n, 1GdFe: Orange crystalline solid, mass (yield): 22 mg (51 %). Selected IR data (KBr, cm−1): 3457 (m), 3349 (m), 3189 (m), 2147 (mw), 2121 (vs), 1677 (vs), 1629 (m), 1614 (m), 1542 (s), 1461 (mw), 1367 (mw), 1307 (mw), 1266 (mw), 1196 (s), 1174 (mw), 1138 (mw), 1106 (mw), 1004 (w), 815 (mw), 769 (mw), 705 (mw), 656 (mw), 563 (mw), 485 (mw), 417 (mw). Elemental analysis. Calcd. for C17H21N13O5FeGd: C, 29.15; H, 3.02; N, 25.99%; found C, 29.09; H, 2.96; N, 26.01%.
{[Dy(L)(H2O)2Fe(CN)6]·3H2O}n, 2DyFe: Orange crystalline solid, mass (yield): 27 mg (60 %). Selected IR data (KBr, cm−1): 3457 (m), 3348 (w), 3236 (mw), 1676 (m), 1631 (m), 1609 (m), 1542 (m), 1461 (m), 1371 (m), 1309 (s), 1267 (s), 1196 (s), 1136 (vs), 1105 (m), 816 (m), 769 (mw), 704 (m), 654 (mw), 570 (mw), 507 (mw), 487 (mw). Elemental analysis. Calcd. for C17H25N13O7DyFe: C, 27.53; H, 3.40; N, 24.55%; found C, 27.25; H, 3.37; N, 24.69%.
{[Dy(L)(H2O)2Co(CN)6]·H2O}n, 3DyCo: White crystalline solid, mass (yield): 22 mg (51 %). Selected IR data (KBr, cm−1): 3458 (s), 3349 (s), 3238 (s), 3189 (mw), 2917 (mw), 2849 (w), 2362 (mw), 2165 (mw), 2148 (mw), 2132 (vs), 2091 (w), 1677 (vs), 1631 (m), 1614 (m), 1542 (vs), 1464 (m), 1368 (m), 1309 (mw), 1268 (mw), 1199 (ms), 1174 (mw), 1138 (m), 1107 (m), 1005 (w), 816 (mw), 770 (m), 708 (mw), 656 (mw), 556 (mw), 495 (m), 478 (m), 459 (mw), 422 (m). Elemental analysis. Calcd. for C17H21N13O5CoDy: C, 28.80; H, 2.99; N, 25.69%; found C, 28.64; H, 2.98; N, 25.45%.

3. Results and Discussion

3.1. Synthesis and Structures of the Complexes

The three compounds, {[Gd(L)(H2O)2Fe(CN)6]·H2O}n, 1GdFe, {[Dy(L)(H2O)2Fe(CN)6]·3H2O}n, 2DyFe, and {[Dy(L)(H2O)2Co(CN)6]·H2O}n, 3DyCo, have been obtained via slow diffusion of water solutions containing [Ln(L)(H2O)4]Cl3-K3[M(CN)6] in 1:1 molar ratio. All complexes show C=O and C=N IR absorption peaks in the range of 1654–1657 cm−1, which indicate the presence of the semicarbazone ligand. The split bands at 2200–2100 cm−1 are assigned to both the monodentate and bridging cyanido groups [47]. The crystalline phase purity of the samples was confirmed by the good agreement between the PXRD patterns and the ones simulated using single-crystal data (Figure S1). The FTIR spectra are displayed in Figure S2.
Complexes 1GdFe and 3DyCo are isostructural, and they crystallize in the orthorhombic space group Pbca with one crystallization water molecule/formula unit. Complex 2DyFe crystallizes in the monoclinic system, space group P21/c, with three lattice water molecules/formula unit. In all complexes, the metal ions have similar coordination environments, and the topology of the heterometallic chains is identical.
Compounds 1GdFe and 3DyCo consist of heterometallic chains with alternating distributions of the 3d and 4f metal ions. Since the two compounds are isostructural, we will describe only the crystal structure of the compound 1GdFe. The general appearance of the chains is crenel-like, due to the fact that the [Fe(CN)6]3− metalloligand acts as a bridge trough two cis cyanido groups and the two neighboring connecting [Fe(CN)6]3− moieties are placed on the same side of the organic ligand coordinated to the lanthanide ion (Figure 1). One of the two water molecules coordinated to the lanthanide ion is involved in intra-chain hydrogen interaction with a cyanido group from a [Fe(CN)6]3− metalloligand coordinated to a neighboring lanthanide ion. The chains are running along the crystallographic a axis.
The lanthanide ions are nine-coordinated by the pentadentate organic ligand (O1, O2, N3, N4, and N5), two nitrogen atoms arising from the cyanido bridges (N8, N13′; symmetry code: ′ = −0.5 + x, 0.5 − y, 1 − z), and two aqua ligands (O3, O4). The Ln-N bond lengths with the organic ligand are in the range of 2.584(6) − 2.598(6) Å for 1GdFe and 2.562(2) − 2.574(2) for 3DyCo, respectively, while for the cyanido groups are Gd1-N8 = 2.534(6), Gd1-N13′ = 2.566(6), respectively, Dy1-N8 = 2.519(3) and Dy1-N13′ = 2.544(3). The Ln-O bond lengths are slightly longer with the organic ligand than aqua ligands: Gd1-O1 = 2.423(5), Gd1-O2 = 2.406(5), Gd1-O3 = 2.330(5), Gd1-O4 = 2.375(5), Dy1-O1 = 2.398(2), Dy1-O2 = 2.382(2), Dy1-O3 = 2.308(2), Dy1-O4 = 2.3518(19) Å. The coordination geometry of the gadolinium ion can be described as spherical capped square antiprism, according to the calculations made with SHAPE software (Table S2) [48].
The O3 aqua ligands are further connected by hydrogen bonding to neighboring chains generating a 2D supramolecular architecture in the crystallographic ac plane (Figure 2). The distances for the hydrogen interactions are: (O3)H1O···N9″ = 1.85 and (O3)H2O···N11″′ = 1.93 Å, while the corresponding angles are: O3-H1O···N9″ = 173.4 and O3-H2O···N11″′ = 170.6º (symmetry codes: ″ = 0.5 + x, 0.5 − y, 1 − z; ″′ = x, 0.5 − y, −0.5 + z).
The extension of the supramolecular architecture to 3D is also mediated by hydrogen bond interactions involving the second aqua ligand, the crystallization water molecules and cyanido groups of the anionic metalloligand (Figure 3). Each O4 coordinated water molecule is involved as donor in hydrogen interactions with two crystallization water molecules. Each crystallization water molecule is acceptor for two hydrogen interactions with two coordinated water molecules from neighboring layers and donor for two cyan groups also from the two neighboring layers. The distances for these hydrogen interactions are: (O4)H4O···O5 = 1.95, (O4)H3O···O5i = 1.92, (O5)H5O···N10 = 2.18 and (O5)H6O···N12ii = 2.14 Å, while the corresponding angles are: O4-H4O···O5 = 166.7, O4-H3O···O5i = 161.8, O5-H5O···N10 = 158.7 and O5-H6O···N12ii = 169.7º (symmetry codes: i = 1 − x, −y, 1 − z; ii = 1.5 − x, −0.5 + y, z).
Compound 2DyFe consists also of heterometallic chains running in this case along the crystallographic b axis, and crystallization water molecules. The 1D chains are formed in a similar manner by connecting [Dy(L)(H2O)2]3+ complex cations by the [Fe(CN)6]3− metalloligands, which employ two cis cyanido groups for bridging (Figure 4). The DyIII ion is nine-coordinated by the pentadentate ligand (O1, O2, N3, N4, and N5), two nitrogen atoms arising from the cyanido bridges (N8, N13), and two aqua ligands (O3, O4). The coordination geometry of the dysprosium ion is also spherical capped square antiprism (Table S2). The Dy-N and Dy-O bond lengths are in the range of 2.503(2)–2.553(2) and 2.3298(18)–2.3783(13) Å, respectively. The two Dy-N bond distances (nitrogen atoms arising from the bridging cyanido groups) are Dy1-N8 = 2.556(2) and Dy1-N13′ = 2.533(2) Å (symmetry code: ′ = 1 − x, −0.5 + y, 1.5 − z). The FeIII ions show a slightly distorted octahedral geometry with Fe1-C bond lengths ranging from 1.929(3) to 1.957(3) Å. Each {Dy(L)(H2O)2} module links two {Fe(CN)6} fragments in cis positions (the Dy···Fe···Dy angle is 95.78°), and each {Fe(CN)6} metalloligand connects two DyIII ions, resulting in a crenel-like chain structure.
The O3 aqua ligand is also involved in intra- and interchain hydrogen bonding generating an analogous 2D supramolecular architecture in the ab crystallographic plane (Figure 5). The distances for the hydrogen interactions are: (O3)H1O···N12″ = 1.88 and (O3)H2O···N10″′ = 2.05 Å, while the corresponding angles are: O3-H1O···N12″ = 172.4 and O3-H2O···N10″′ = 167.9º (symmetry codes: ″ = x, −1 + y, z; ″′ = 2 − x, −0.5 + y, 1.5 − z).
The main differences between crystals 1GdFe and 2DyFe appear in hydrogen interactions established between the supramolecular layers. Compound 2DyFe has two more crystallization water molecules per unit comparing with the crystals 1GdFe and 3DyCo. The O4 coordinated water molecule is also involved as donor in hydrogen interactions with two crystallization water molecules, O5 and O5i, (Figure 6). Each of these crystallization water molecules is acceptor for two hydrogen interactions with two coordinated water molecules from neighboring layers and acts as donor for only one cyanido group (O5 is donor for N11i atom). The other two crystallization water molecules are involved in hydrogen bonding with one NH2 group and one cyanido group (O6), respectively, and two cyanido groups (O7). The distances for the hydrogen interactions are: (O4)H4O···O5 = 1.91, (O4)H3O···O5i = 2.02, (O5)H5O···N11i = 2.28, (N7)H5N···O6 = 2.16, (O6)H8O···N12iii = 2.28, (O7)H9O···N9ii = 2.15, and (O7)H10O···N10i = 2.20 Å, while the corresponding angles are: O4-H4O···O5 = 174.4, O4-H3O···O5i = 171.4, O5-H5O···N11i = 146.7, N7-H5N···O6 = 162.3, O6-H8O···N12iii = 166.1, O7-H9O···N9ii = 153.9, and O7-H10O···N10i = 148.8 º (symmetry codes: i = 1 − x, 1 − y, 1 − z; ii = 1 − x, −0.5 + y, 1.5 − z; iii = x, 1.5 − y, −0.5 + z).
The shortest intramolecular Fe···Gd distances in 1GdFe are 5.513 and 5.557 Å, while the intramolecular Dy···Fe distances in 2DyFe are 5.542 and 5.494 Å. Selected bond distances and angles for compounds 13 are listed in Table S3.

3.2. Magnetic Properties of the Complexes

Static magnetic characterizations. The magnetic susceptibility data for compounds 13 were measured on polycrystalline samples in the temperature range of 2–300 K as shown in Figure 7, in the form of χMT vs. T curves. The observed χMT values at 300 K for 1GdFe, 3DyCo, and 2DyFe are of 8.265, 15.454, and 16.305 cm3 mol−1K, which are slightly higher than the expected values for a non-interacting spin system of one GdIII (7.88 cm3 mol−1 K, S = 7/2, 8S7/5, g = 2.00), DyIII (14.17 cm3 mol−1 K, S = 5/2, 6H15/2, g = 4/3) [49], and one low-spin S = ½ FeIII ion or one diamagnetic CoIII ion [44]. Upon cooling, the χMT values stay almost constant in the high temperature region, while at low temperatures, the χMT values show a rapid decrease and reach the values of 6.833, 12.722, and 11.210 cm3 mol−1 K, respectively, at 2.0 K. In the case of 1GdFe, the decrease in χMT with temperature may be associated with FeIII-GdIII antiferromagnetic interactions along the heterometallic alternating chain. The possible presence of intermolecular interactions can also contribute to this decrease. The expected ferrimagnetic behavior (i.e., the characteristic minimum on the χMT vs. T curve) is not observed, probably due to the small magnitude of the exchange interactions along the chain. The evolution of the temperature dependence of the magnetic susceptibility for 3DyCo is exclusively defined by the presence of strongly anisotropic DyIII ions, which are isolated by diamagnetic low spin CoIII ions. The decrease in χMT with the temperature is due to the depopulation of MJ (Stark) sublevels of the DyIII centers in 3DyCo [50]. This effect is certainly present in the case of compound 2DyFe. Additionally, a FeIII-DyIII magnetic coupling along the chain can be expected. In the case of 2DyFe, the evolution of χMT shows a more important slope compared to compound 3DyCo (Figure 7), with a lower value of susceptibility at 2.0 K (11.210 cm3 mol−1 K). This indicates the presence of some antiferromagnetic impact, which contributes to the observed decreasing χMT values. For 2DyFe, the existence of magnetic interactions similar to those in 1GdFe also suggests the formation of a ferrimagnetic chain, which, associated with strong magnetic anisotropy, could lead to a Single Chain Magnet. Unfortunately, as in the case of 1GdFe, the increase in χMT at low temperatures and the characteristic minimum were not detected for 2DyFe (Figure 7). This behavior is probably due to the very small antiferromagnetic interactions along the chain. The magnetization measurements (Figure S3) support the presence of an important magnetic anisotropy in 2DyFe and 3DyCo.
Dynamic magnetic characterizations. Dynamic magnetic properties of the compounds 1GdFe, 2DyFe, and 3DyCo were studied by measuring the temperature and field dependence ac (alternative current) magnetic susceptibility. Compound 1GdFe does not have any manifestation of the out-of-phase component (χ″ac) of the ac magnetic susceptibility at 2 K and zero dc (direct current) field. After applying small dc field (2000 Oe) no modification was observed in χ″ac component of ac susceptibility.
For 3DyCo, no signal was observed under zero dc field at 2.0 K, in χ″ac component of ac susceptibility. After applying of small dc fields (up to 4000 Oe), a frequency dependent out-of-phase signal appears (Figure 8b) and has a rich evolution in function of the field. Such behavior is consistent with presence of strongly anisotropic paramagnetic centers DyIII and indicates the presence of field-induced slow magnetic relaxation. The intensity of the out-of-phase signals gradually increases till about 2000 Oe, and then, it slightly decreases. To investigate the nature of slow magnetic relaxation, additional ac susceptibility data were collected under fixed dc field (2000 Oe) and stable temperatures between 2.0 and 5.0 K (with a 0.2 K increment)—Figure 8d–f. The temperature sweeping of the ac susceptibility shows the important evolution of the χ″ac component and supports the presence of field-induced slow magnetic relaxation in 3DyCo. Since the CoIII ion is diamagnetic, compound 3DyCo can be described as being a chain of Single Ion Magnets.
A similar strategy of measurements was used in the dynamic analysis of 2DyFe. As in the case of 3DyCo, at T = 2.0 K and zero dc field, no signal was detected in the out-of-phase component of the ac susceptibility. The signals appear when a small magnetic dc field was applied and has similar evolutions as in the case of 3DyCo (Figure S8a–c). A dc field of 3000 Oe was used to perform the temperature sweeping measurements of the ac susceptibility in the case of 2DyFe (Figure S8d–f).
For both compounds (2DyFe and 3DyCo), the visual analysis of the out-of-phase susceptibilities, as well of the χ″ac vs. χ′ac plots (Cole–Cole plots), suggests the presence of at least two distinct relaxation processes. In consequence, the ac susceptibility data (field sweeping and temperature sweeping measurements) for 3DyCo and 2DyFe were evaluated with generalized (extended) Debye equations combining two-relaxation processes [51,52,53]. The two relaxation times (τ1, and τ2,) and two distribution parameters (α1, and α2) occur along with two isothermal susceptibilities (χT1 and χT2) and one common adiabatic susceptibility (χs) (see Equations (S1) and (S2)). The deconvolution of two relaxations process is presented in Figures S4–S7. Variable parameters derived from the best fits of the ac susceptibility are shown in Figures S9–S12.
In both compounds, the first relaxation process LF (Low Frequency) is well defined, while the second HF (High Frequency) process has large errors on the variable parameters. The distributions of relaxation times for the LF process are rather broad (α1 = 0.3 ÷ 0.5). The extracted temperature and field dependence of relaxation times for 2DyFe and 3DyCo can be modulated based on four relaxation mechanisms according to the following equation [54,55,56,57,58]:
τ T / H 1 ( T , H ) = Q 1 1 + Q 2 H 2 + τ 0 1 exp ( U e f f k T ) + A H 4 T + C T n .
The first term represents the Quantum Tunneling of Magnetization (QTM), the second term corresponds to Orbach, the third to Direct, and the last one to Raman process; moreover, H = applied dc magnetic field and T = temperature. In order to constrain the variable parameters and avoid the overparameterization problem, temperature and field dependence of relaxation times were fitted simultaneously [56] (vector of data: τ−1 in s−1, T in Kelvin, and H in kOe). Only LF signals will be discussed below, as the second process (HF) is poorly defined. Different combinations of the four mechanisms of relaxation have been used in order to simulate the evolution of the relaxation times. For the LF signals, with both compounds, the contribution of Quantum Tunneling of Magnetization (QTM) is indispensable to simulate the relaxation data. The continuous decreasing trend of τ−1 vs. applied field (H) excludes the presence of significant contribution of Direct relaxation mechanism. The other contributions in relaxation times can be Raman and/or Orbach, which have the same increasing evolutions with temperature variation [57]. In this restricted range of temperature, it is difficult to separate these two components. In order to have some information regarding the manifestation of these mechanisms, the comparative fits on the temperature dependence of relaxation time for 2DyFe and 3DyCo have been done (Figures S13–S15). Both mechanisms can reproduce the time of evolution. The quality of the Orbach mechanism is slightly better. It should be mention here that the distribution parameters (α) have important impact on uncertainties relaxation time [59] and can also be an argument in favor of one or another mechanism. The analysis presented in Figures S13–S15 shows the similarity in uncertainties of relaxation times for both mechanisms. Based on this argument and a low temperature range (2–4 K) of relaxation data for 3DyCo and 2DyFe, as well the traditional representation of relaxation phenomena in SMM, our analysis of LF relaxation process is limited to two contributions: QTM and Orbach. The best fit for LF relaxation processes based on the two mechanisms has been obtained for the following sets of parameters:
3DyCo: Ueff/k = 7.1 K; τ0 = 7.5 × 10−5 s; Q1 = 121 s−1; Q2 = 1.05 kOe−2
2DyFe: Ueff/k = 10.8 K; τ0 = 5.9 × 10−4 s; Q1 = 200 s−1; Q2 = 0.05 kOe−2.
The obtained relaxation parameters are similar for compounds 2DyFe and 3DyCo. Due to the diamagnetic CoIII ions in 3DyCo, the slow magnetic relaxation is solely associated with the anisotropic DyIII ions. The existence of intrachain magnetic interaction in 2DyFe does not change significantly the energy barrier of slow relaxation of the magnetization (see the temperature dependence in Figure 9), but affects more the field dependence, which becomes much more redistributed. This probably can be associated to redistribution/mixing the different energy levels in the 2DyFe as a result of the small antiferromagnetic interaction along the chain and of intermolecular (interchain) interaction. The splitting of MJ (Stark) sublevels of the DyIII centers under variation of the magnetic field also contributes to this redistribution. As in the case of static magnetic measurements, these competitive interactions cannot be quantified at the reported range of temperatures.
In a recent paper, Ma et al. report on a family of discrete, tetranuclear 3d-4f complexes assembled from a cationic lanthanide complexes and [M(CN)6]3− metalloligands (M = Fe, Co), the ligand attached to the lanthanide(III) ions (Tb, Dy, and Ho) being also pentadentate [60]. The field-induced slow relaxation of the magnetization, with a low energy barrier (11.17 K), was observed only with the [Dy2Co2] derivative. The presence of the paramagnetic FeIII ion does not improve the SMM behavior: for the [Dy2Fe2] derivative, the slow relaxation is not observed even by applying dc fields.

4. Conclusions

In this paper, we have shown that the pentadentate bis-semicarbazone ligand, L, generates robust cationic LnIII complexes, which are useful modules for constructing heterometallic coordination polymers. The metalloligands, [M(CN)6]3−, employ two cis cyanido groups as bridges against the LnIII ions, resulting in a wave-like chain topology for the three compounds. The investigation of the magnetic properties reveals that the two DyIII-containing coordination polymers exhibit slow relaxation of the magnetization, with rather low energy barriers. From the magnetic point of view, compound 3DyCo behaves like a chain of Single Ion Magnets. These results open interesting perspectives for the synthesis of new cyanido-bridged 3d-4f complexes, using not only homoleptic but also heteroleptic cyanido tectons, as well as other types of metalloligands. Further work is in progress in our laboratory.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/magnetochemistry7050057/s1, Crystallographic data (single crystal and PXRD), bond distances and angles, infrared spectra; magnetic data; treatment of the ac magnetic data.

Author Contributions

Conceptualization, D.D. and M.A.; methodology, M.A., D.D. and G.N.; formal analysis, G.N. and A.M.M.; investigation, D.D., A.M.M. and G.N.; writing—original draft preparation, D.D. and M.A.; writing—review and editing, M.A. and G.N. All authors have read and agreed to the published version of the manuscript.

Funding

MAGMOLMET, grant ID: 867445.

Acknowledgments

D.D. is grateful to the European Union’s Horizon 2020 research and innovation programme, for financial support (MAGMOLMET, grant ID: 867445).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Woodruff, D.N.; Winpenny, R.E.P.; Layfield, R.A. Lanthanide Single-Molecule Magnets. Chem. Rev. 2013, 113, 5110–5148. [Google Scholar] [CrossRef]
  2. Goodwin, C.A.P.; Ortu, F.; Reta, D.; Chilton, N.F.; Mills, D.P. Molecular magnetic hysteresis at 60 Kelvin in dysprosocenium. Nature 2017, 548, 439–442. [Google Scholar] [CrossRef]
  3. Guo, F.S.; Day, B.M.; Chen, Y.C.; Tong, M.L.; Mansikkamaki, A.; Layfield, R.A. Magnetic hysteresis up to 80 kelvin in a dysprosium metallocene single-molecule magnet. Science 2018, 362, 1400–1403. [Google Scholar] [CrossRef] [Green Version]
  4. Dhers, S.; Feltham, H.L.C.; Brooke, S. A toolbox of building blocks, linkers and crystallisation methods used to generate single-chain magnets. Coord. Chem. Rev. 2015, 296, 24–44. [Google Scholar] [CrossRef]
  5. Zheng, Y.-Z.; Lan, Y.; Wernsdorfer, W.; Anson, C.E.; Powell, A.K. Polymerisation of the Dysprosium Acetate Dimer Switches on Single-Chain Magnetism. Chem. Eur. J. 2009, 15, 12566–12570. [Google Scholar] [CrossRef] [PubMed]
  6. Liu, Z.-Y.; Xia, Y.-F.; Jiao, J.; Yang, E.-C.; Zhao, X.-J. Two water-bridged cobalt(ii) chains with isomeric naphthoate spacers: From metamagnetic to single-chain magnetic behaviour. Dalton Trans. 2015, 44, 19927–19934. [Google Scholar] [CrossRef] [PubMed]
  7. Bogani, L.; Sangregorio, C.; Sessoli, R.; Gatteschi, D. Molecular engineering for single-chain-magnet behavior in a one-dimensional dysprosium–nitronyl nitroxide compound. Angew. Chem. Int. Ed. 2005, 44, 5817–5821. [Google Scholar] [CrossRef] [PubMed]
  8. Bernot, K.; Bogani, L.; Caneschi, A.; Gatteschi, D.; Sessoli, R. A family of rare-earth-based single chain magnets: Playing with anisotropy. J. Am. Chem. Soc. 2006, 128, 7947–7956. [Google Scholar] [CrossRef]
  9. Pedersen, K.S.; Bendix, J.; Clérac, R. Single-molecule magnet engineering: Building-block approaches. Chem. Commun. 2014, 50, 4396–4415. [Google Scholar] [CrossRef] [Green Version]
  10. Pal, S.; Dey, K.; Benmansour, S.; Gómez-García, C.J.; Nayek, H.P. Syntheses, structures and magnetic properties of cyano-bridged one-dimensional Ln3+–Fe3+ (Ln = La, Dy, Ho and Yb) coordination polymers. New J. Chem. 2019, 43, 6228–6233. [Google Scholar] [CrossRef]
  11. Yu, D.-Y.; Li, L.; Zhou, H.; Yuan, A.-H.; Li, Y.-Z. Cyano-Bridged 4f-3d Assemblies with Achiral Helical Chains: Syntheses, Structures, and Magnetic Properties. Eur. J. Inorg. Chem. 2012, 3394–3397. [Google Scholar] [CrossRef]
  12. Estrader, M.; Ribas, J.; Tangoulis, V.; Solans, X.; Font-Bardía, M.; Maestro, M.; Diaz, C. Synthesis, Crystal Structure, and Magnetic Studies of One-Dimensional Cyano-Bridged Ln3+−Cr3+ Complexes with bpy as a Blocking Ligand. Inorg. Chem. 2006, 45, 8239–8250. [Google Scholar] [CrossRef] [PubMed]
  13. Figuerola, A.; Ribas, J.; Casanova, D.; Maestro, M.; Alvarez, S.; Diaz, C. Magnetism of Cyano-Bridged Ln3+−M3+ Complexes. Part II:  One-Dimensional Complexes (Ln3+ = Eu, Tb, Dy, Ho, Er, Tm; M3+ = Fe or Co) with bpy as Blocking Ligand. Inorg. Chem. 2005, 44, 6949–6958. [Google Scholar] [CrossRef] [PubMed]
  14. Figuerola, A.; Diaz, C.; Ribas, J.; Tangoulis, V.; Sangregorio, C.; Gatteschi, D.; Maestro, M.; Mahía, J. Magnetism of Cyano-Bridged Hetero-One-Dimensional Ln3+−M3+ Complexes (Ln3+ = Sm, Gd, Yb; M3+ = FeLS, Co). Inorg. Chem. 2003, 42, 5274–5281. [Google Scholar] [CrossRef] [PubMed]
  15. Petrosyants, S.P.; Ilyukhin, A.B.; Efimov, N.N.; Gavrikov, A.V.; Novotortsev, V.M. Self-assembly and SMM properties of lanthanide cyanocobaltate chain complexes with terpyridine as blocking ligand. Inorg. Chim. Acta 2018, 482, 813–820. [Google Scholar] [CrossRef]
  16. Muddassir, M.; Song, X.-J.; Chen, Y.; Cao, F.; Weia, R.-M.; Song, Y. Ion-induced diversity in structure and magnetic properties of hexacyanometalate–lanthanide bimetallic assemblies. CrystEngComm. 2013, 15, 10541–10549. [Google Scholar] [CrossRef]
  17. Figuerola, A.; Ribas, J.; Solans, X.; Font-Bardía, M.; Maestro, M.; Diaz, C. One Dimensional 3d–4f Heterometallic Compounds: Synthesis, Structure and Magnetic Properties. Eur. J. Inorg. Chem. 2006, 1846–1852. [Google Scholar] [CrossRef]
  18. Zhao, H.; Lopez, N.; Prosvirin, A.; Chifotidesa, H.T.; Dunbar, K.R. Lanthanide–3d cyanometalate chains Ln(III)–M(III) (Ln = Pr, Nd, Sm, Eu, Gd, Tb; M = Fe) with the tridentate ligand2,4,6-tri(2-pyridyl)-1,3,5-triazine (tptz): Evidence of ferromagnetic interactions for the Sm(III)–M(III) compounds (M = Fe, Cr). Dalton Trans. 2007, 878–888. [Google Scholar] [CrossRef]
  19. Liu, J.; Knoeppel, D.W.; Liu, S.; Meyers, E.A.; Shore, S.G. Cyanide-Bridged Lanthanide(III)−Transition Metal Extended Arrays:  Interconversion of One-Dimensional Arrays from Single-Strand (Type A) to Double-Strand (Type B) Structures. Complexes of a New Type of Single-Strand Array (Type C). Inorg. Chem. 2001, 40, 2842–2850. [Google Scholar] [CrossRef]
  20. Kou, H.-Z.; Gao, S.; Sun, B.-W.; Zhang, J. Metamagnetism of the First Cyano-Bridged Two-Dimensional Brick-Wall-like 4f−3d Array. Chem. Mater. 2001, 13, 1431–1433. [Google Scholar] [CrossRef]
  21. Wilson, D.C.; Liu, S.; Chen, X.; Meyers, E.A.; Bao, X.; Prosvirin, A.V.; Dunbar, K.R.; Hadad, C.M.; Shore, S.G. Water-Free Rare Earth-Prussian Blue Type Analogues: Synthesis, Structure, Computational Analysis, and Magnetic Data of {LnIII(DMF)6FeIII(CN)6}∞ (Ln = Rare Earths Excluding Pm). Inorg. Chem. 2009, 48, 5725–5735. [Google Scholar] [CrossRef]
  22. Chen, W.-T.; Guo, G.-C.; Wang, M.-S.; Xu, G.; Cai, L.-Z.; Akitsu, T.; Akita-Tanaka, M.; Matsushita, A.; Huang, J.-S. Self-Assembly and Characterization of Cyano-Bridged Bimetallic [Ln−Fe] and [Ln−Co] Complexes (Ln = La, Pr, Nd and Sm). Nature of the Magnetic Interactions between the Ln3+ and Fe3+ Ions. Inorg. Chem. 2007, 46, 2105–2114. [Google Scholar] [CrossRef]
  23. Chen, W.-T.; Wu, A.-Q.; Guo, G.-C.; Wang, M.-S.; Ca, L.-Z.; Huang, J.-S. Cyano-Bridged 2D Bimetallic 4f–3d Arrays with Monolayered Stair-Like, Brick-Wall-Like, or Bilayered Topologies–Rational Syntheses and Crystal Structures. Eur. J. Inorg. Chem. 2010, 2826–2835. [Google Scholar] [CrossRef]
  24. Xin, Y.; Wang, J.; Zychowicz, M.; Zakrzewski, J.J.; Nakabayashi, K.; Sieklucka, B.; Chorazy, S.; Ohkosh, S. Dehydration−Hydration Switching of Single-Molecule Magnet Behavior and Visible Photoluminescence in a Cyanido-Bridged DyIIICoIII Framework. J. Am. Chem. Soc. 2019, 141, 18211–18220. [Google Scholar] [CrossRef]
  25. Visinescu, D.; Toma, L.M.; Fabelo, O.; Ruiz-Pérez, C.; Lloret, F.; Julve, M. Low-Dimensional 3d–4f Complexes Assembled by Low-Spin [FeIII(phen)(CN)4]− Anions. Inorg. Chem. 2013, 52, 1525–1537. [Google Scholar] [CrossRef] [PubMed]
  26. Chorazy, S.; Rams, M.; Wyczesany, M.; Nakabayashi, K.; Ohkoshi, S.; Sieklucka, B. Antiferromagnetic exchange and long-range magnetic ordering in supramolecular networks constructed of hexacyanido-bridged LnIII(3-pyridone)–CrIII (Ln = Gd, Tb) chains. CrystEngComm. 2018, 20, 1271–1281. [Google Scholar] [CrossRef]
  27. Xue, A.-Q.; Liu, Y.-Y.; Li, J.-X.; Zhang, Y.; Meng, Y.-S.; Zhu, W.-H.; Zhang, Y.-Q.; Sun, H.-L.; Wang, F.; Qiu, G.-X.; et al. The differential magnetic relaxation behaviours of slightly distorted triangular dodecahedral dysprosium analogues in a type of cyano-bridged 3d–4f zig-zag chain compounds. Dalton Trans. 2020, 49, 6867–6875. [Google Scholar] [CrossRef] [PubMed]
  28. Petrosyants, S.P.; Ilyukhin, A.B.; Efimov, N.N.; Novotortsev, V.M. Cyano-Bridged d–f Ensembles of the Dysprosium Tetrapyridine Complexes with the Hexacyanoferrate Anion. Russ. J. Coord. Chem. 2018, 44, 660–666. [Google Scholar] [CrossRef]
  29. Chorazy, S.; Zakrzewski, J.J.; Wang, J.; Ohkoshi, S.; Sieklucka, B. Incorporation of hexacyanidoferrate(III) ion in photoluminescent trimetallic Eu(3-pyridone)[Co1−xFex(CN)6] chains exhibiting tunable visible light absorption and emission properties. CrystEngComm 2018, 20, 5695–5706. [Google Scholar] [CrossRef]
  30. Chorazy, S.; Kumar, K.; Nakabayashi, K.; Sieklucka, B.; Ohkoshi, S. Fine Tuning of Multicolored Photoluminescence in Crystalline Magnetic Materials Constructed of Trimetallic EuxTb1−x[Co(CN)6] Cyanido-Bridged Chains. Inorg. Chem. 2017, 56, 5239–5252. [Google Scholar] [CrossRef]
  31. Chorazy, S.; Wyczesany, M.; Sieklucka, B. Lanthanide Photoluminescence in Heterometallic Polycyanidometallate-Based Coordination Networks. Molecules 2017, 22, 1902. [Google Scholar] [CrossRef] [Green Version]
  32. Chorazy, S.; Rams, M.; Nakabayashi, K.; Sieklucka, B.; Ohkoshi, S. White Light Emissive DyIII Single-Molecule Magnets Sensitized by Diamagnetic [CoIII(CN)6]3– Linkers. Chem. Eur. J. 2016, 22, 7371–7375. [Google Scholar] [CrossRef]
  33. Chorazy, S.; Rams, M.; Wang, J.; Sieklucka, B.; Ohkoshi, S. Octahedral Yb(III) complexes embedded in [CoIII(CN)6]-bridged coordination chains: Combining sensitized near-infrared fluorescence with slow magnetic relaxation. Dalton Trans. 2017, 46, 13668–13672. [Google Scholar] [CrossRef]
  34. Qian, K.; Huang, X.-C.; Zhou, C.; You, X.-Z.; Wang, X.-Y.; Dunbar, K.R. A Single-Molecule Magnet Based on Heptacyanomolybdate with the Highest Energy Barrier for a Cyanide Compound. J. Am. Chem. Soc. 2013, 135, 13302–13305. [Google Scholar] [CrossRef]
  35. Sasnovskaya, V.D.; Kopotkov, V.A.; Talantsev, A.D.; Morgunov, R.B.; Yagubskii, E.B.; Simonov, S.V.; Zorina, L.V.; Mironov, V.S. Synthesis, Structure, and Magnetic Properties of 1D {[MnIII(CN)6][MnII(dapsc)]}n Coordination Polymers: Origin of Unconventional Single-Chain Magnet Behavior. Inorg. Chem. 2017, 56, 8926–8943. [Google Scholar] [CrossRef]
  36. Zorina, L.V.; Simonov, S.V.; Sasnovskaya, V.D.; Talantsev, A.D.; Morgunov, R.B.; Mironov, V.S.; Yagubskii, E.B. Slow Magnetic Relaxation, Antiferromagnetic Ordering, and Metamagnetism in MnII(H2dapsc)-FeIII(CN)6 Chain Complex with Highly Anisotropic Fe-CN-Mn Spin Coupling. Chem. Eur. J. 2019, 25, 14583–14597. [Google Scholar] [CrossRef]
  37. Bar, A.K.; Gogoi, N.; Pichon, C.; Goli, V.M.L.D.P.; Thlijeni, M.; Duhayon, C.; Suaud, N.; Guihéry, N.; Barra, A.-L.; Ramasesha, S.; et al. Pentagonal Bipyramid FeII Complexes: Robust Ising-spin Units Towards Heteropolynuclear Nano-Magnets. Chem. Eur. J. 2017, 23, 4380–4396. [Google Scholar] [CrossRef] [PubMed]
  38. Pichon, C.; Suaud, N.; Duhayon, C.; Guihéry, N.; Sutter, J.-P. Cyano-Bridged Fe(II)−Cr(III) Single-Chain Magnet Based on Pentagonal Bipyramid Units: On the Added Value of Aligned Axial Anisotropy. J. Am. Chem. Soc. 2018, 140, 7698–7704. [Google Scholar] [CrossRef]
  39. Sommerer, S.O.; Westcott, B.L.; Cundari, T.R.; Krause, J.A. A structural and computational study of tetraaqua[2,6-diacetylpyridinebis-(semicarbazone)]-gadolinium(III) trinitrate. Inorg. Chim. Acta 1993, 209, 101–104. [Google Scholar] [CrossRef]
  40. Gioia, M.; Crundwell, G.; Westcott, B.L. Tetraaqua[2,6-diacetylpyridine bis(semicarbazone)]samarium(III) trinitrate. IUCrData 2018, 3, x181454. [Google Scholar] [CrossRef]
  41. Sasnovskaya, V.D.; Kopotkov, V.A.; Kazakova, A.V.; Talantsev, A.D.; Morgunov, R.B.; Simonov, S.V.; Zorina, L.V.; Mironov, V.S.; Yagubskii, E.B. Slow magnetic relaxation in mononuclear complexes of Tb, Dy, Ho and Er with the pentadentate (N3O2) Schiff-base dapsc ligand. New J. Chem. 2018, 42, 14883–14893. [Google Scholar] [CrossRef]
  42. Palenik, G.J.; Wester, D.W.; Rychlewska, U.; Palenik, R.C. Pentagonal-Bipyramidal Complexes. Synthesis and Crystal Structures of Diaqua [2,6-diacetylpyridine bis(semicarbazone)]chromium(III) Hydroxide Dinitrate Hydrate and Dichloro[2,6-diacetylpyridine bis(semicarbazone)] iron(III) Chloride Dihydrate. Inorg. Chem. 1976, 15, 1814–1819. [Google Scholar] [CrossRef]
  43. Pascal, P. Magnochemical studies. Ann. Chim. Phys. 1910, 19, 5–70. [Google Scholar]
  44. Kahn, O. Molecular Magnetism; VCH Publishers: New York, NY, USA, 1993. [Google Scholar]
  45. Sheldrick, G.M. SHELXT—Integrated space-group and crystal-structure determination. Acta Cryst. 2015, A71, 3–8. [Google Scholar] [CrossRef] [Green Version]
  46. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Cryst. 2015, C71, 3–8. [Google Scholar] [CrossRef]
  47. Nakamoto, K. Infrared and Raman Spectra of Inorganic and Coordination Compounds, 4th ed.; John Wiley & Sons: Hoboken, NJ, USA, 1986; p. 245. [Google Scholar]
  48. Llunell, 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; University of Barcelona: Barcelona, Spain, 2013. [Google Scholar]
  49. 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]
  50. Kahn, M.L.; Ballou, R.; Porcher, P.; Kahn, O.; Sutter, J.-P. Analytical Determination of the {Ln–Aminoxyl Radical} Exchange Interaction Taking into Account Both the Ligand-Field Effect and the Spin–Orbit Coupling of the Lanthanide Ion (Ln = DyIII and HoIII). Chem. A Eur. J. 2002, 8, 525–531. [Google Scholar] [CrossRef]
  51. Guo, Y.-N.; Xu, G.-F.; Gamez, P.; Zhao, L.; Lin, S.-Y.; Deng, R.; Tang, J.; Zhang, H.-J. Two-Step Relaxation in a Linear Tetranuclear Dysprosium(III) Aggregate Showing Single-Molecule Magnet Behavior. J. Am. Chem. Soc. 2010, 132, 8538–8539. [Google Scholar] [CrossRef]
  52. Grahl, M.; Kötzler, J.; Seßler, I. Correlation between Domain-Wall Dynamics and Spin-Spin Relaxation in Uniaxial Ferromagnets. J. Magn. Magn. Mater. 1990, 90–91, 187–188. [Google Scholar] [CrossRef]
  53. Dolai, M.; Ali, M.; Titiš, J.; Boča, R. Cu(II)–Dy(III) and Co(III)–Dy(III) Based Single Molecule Magnets with Multiple Slow Magnetic Relaxation Processes in the Cu(II)–Dy(III) Complex. Dalton Trans. 2015, 44, 13242–13249. [Google Scholar] [CrossRef] [PubMed]
  54. Lucaccini, E.; Sorace, L.; Perfetti, M.; Costes, J.-P.; Sessoli, R. Beyond the Anisotropy Barrier: Slow Relaxation of the Magnetization in Both Easy-Axis and Easy-Plane Ln(Trensal) Complexes. Chem. Commun. 2014, 50, 1648–1651. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Zadrozny, J.M.; Atanasov, M.; Bryan, A.M.; Lin, C.Y.; Rekken, B.D.; Power, P.P.; Neese, F.; Long, J.R. Slow Magnetization Dynamics in a Series of Two-Coordinate Iron(II) Complexes. Chem. Sci. 2013, 4, 125–138. [Google Scholar] [CrossRef]
  56. Feng, X.; Liu, J.L.; Pedersen, K.S.; Nehrkorn, J.; Schnegg, A.; Holldack, K.; Bendix, J.; Sigrist, M.; Mutka, H.; Samohvalov, D.; et al. Multifaceted Magnetization Dynamics in the Mononuclear Complex [ReIVCl4(CN)2]2–. Chem. Commun. 2016, 52, 12905–12908. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Ding, Y.S.; Yu, K.X.; Reta, D.; Ortu, F.; Winpenny, R.E.P.; Zheng, Y.Z.; Chilton, N.F. Field- and Temperature-Dependent Quantum Tunnelling of the Magnetisation in a Large Barrier Single-Molecule Magnet. Nat. Commun. 2018, 9, 1–10. [Google Scholar] [CrossRef] [PubMed]
  58. Liu, X.; Feng, X.; Meihaus, K.R.; Meng, X.; Zhang, Y.; Li, L.; Liu, J.; Pedersen, K.S.; Keller, L.; Shi, W.; et al. Coercive Fields Above 6 T in Two Cobalt(II)–Radical Chain Compounds. Angew. Chem. Int. 2020, 59, 10610–10618. [Google Scholar] [CrossRef] [PubMed]
  59. Reta, D.; Chilton, N.F. Uncertainty estimates for magnetic relaxation times and magnetic relaxation parameters. Phys. Chem. Chem. Phys. 2019, 21, 23567–23575. [Google Scholar] [CrossRef]
  60. Wang, R.; Wang, H.; Wang, J.; Bai, F.; Ma, Y.; Li, L.; Wang, Q.; Zhao, B.; Cheng, P. The different magnetic relaxation behaviors in [Fe(CN)6]3− or [Co(CN)6]3− bridged 3d–4f heterometallic compounds. CrystEngComm 2020, 22, 2998–3004. [Google Scholar] [CrossRef]
Scheme 1. The structure of bis-semicarbazone ligand, L.
Scheme 1. The structure of bis-semicarbazone ligand, L.
Magnetochemistry 07 00057 sch001
Figure 1. Perspective view of the 1D coordination polymer in compound 1GdFe (symmetry codes: ′ = −0.5 + x, 0.5 − y, 1 − z; ″ = 0.5 + x, 0.5 − y, 1 − z).
Figure 1. Perspective view of the 1D coordination polymer in compound 1GdFe (symmetry codes: ′ = −0.5 + x, 0.5 − y, 1 − z; ″ = 0.5 + x, 0.5 − y, 1 − z).
Magnetochemistry 07 00057 g001
Figure 2. View along the crystallographic b axis of the packing diagram of compound 1GdFe showing a supramolecular layer (symmetry codes: ′ = −0.5 + x, 0.5 − y, 1 − z; ″ = 0.5 + x, 0.5 − y, 1 − z; ″′ = x, 0.5 − y, −0.5 + z).
Figure 2. View along the crystallographic b axis of the packing diagram of compound 1GdFe showing a supramolecular layer (symmetry codes: ′ = −0.5 + x, 0.5 − y, 1 − z; ″ = 0.5 + x, 0.5 − y, 1 − z; ″′ = x, 0.5 − y, −0.5 + z).
Magnetochemistry 07 00057 g002
Figure 3. Detail of the packing diagram with the hydrogen interactions established between chains from neighboring supramolecular layers and crystallization water molecules in crystal 1GdFe (symmetry codes: i = 1 − x, −y, 1 − z; ii = 1.5 − x, −0.5 + y, z).
Figure 3. Detail of the packing diagram with the hydrogen interactions established between chains from neighboring supramolecular layers and crystallization water molecules in crystal 1GdFe (symmetry codes: i = 1 − x, −y, 1 − z; ii = 1.5 − x, −0.5 + y, z).
Magnetochemistry 07 00057 g003
Figure 4. Perspective view of the 1D coordination polymer in compound 2DyFe. The inset shows a detail of the coordination environment of the DyIII ion (symmetry codes: ′ = 1 − x, −0.5 + y, 1.5 − z; ″ = x, −1 + y, z).
Figure 4. Perspective view of the 1D coordination polymer in compound 2DyFe. The inset shows a detail of the coordination environment of the DyIII ion (symmetry codes: ′ = 1 − x, −0.5 + y, 1.5 − z; ″ = x, −1 + y, z).
Magnetochemistry 07 00057 g004
Figure 5. View along the crystallographic c axis of the packing diagram of compound 2DyFe showing a supramolecular layer (symmetry codes: ″ = x, −1 + y, z; ″′ = 2 − x, −0.5 + y, 1.5 − z).
Figure 5. View along the crystallographic c axis of the packing diagram of compound 2DyFe showing a supramolecular layer (symmetry codes: ″ = x, −1 + y, z; ″′ = 2 − x, −0.5 + y, 1.5 − z).
Magnetochemistry 07 00057 g005
Figure 6. Detail of the packing diagram with the hydrogen interactions established between chains from neighboring supramolecular layers and crystallization water molecules in crystal 2DyFe (symmetry codes: i = 1 − x, 1 − y, 1 − z; ii = 1 − x, −0.5 + y, 1.5 − z; iii = x, 1.5 − y, −0.5 + z).
Figure 6. Detail of the packing diagram with the hydrogen interactions established between chains from neighboring supramolecular layers and crystallization water molecules in crystal 2DyFe (symmetry codes: i = 1 − x, 1 − y, 1 − z; ii = 1 − x, −0.5 + y, 1.5 − z; iii = x, 1.5 − y, −0.5 + z).
Magnetochemistry 07 00057 g006
Figure 7. Temperature dependence of χMT vs. T for complexes 1GdFe, 2DyFe, and 3DyCo.
Figure 7. Temperature dependence of χMT vs. T for complexes 1GdFe, 2DyFe, and 3DyCo.
Magnetochemistry 07 00057 g007
Figure 8. Field dependence (left, a,b) and temperature dependence (right, d,e) of ac susceptibility (Hac = 3.0 Oe) and Cole–Cole plots, (c,f), for 3DyCo at the indicated temperatures and fields. The solid lines represent the best fits according to the generalized Debye model for two relaxations processes (Equations (S1) and (S2)).
Figure 8. Field dependence (left, a,b) and temperature dependence (right, d,e) of ac susceptibility (Hac = 3.0 Oe) and Cole–Cole plots, (c,f), for 3DyCo at the indicated temperatures and fields. The solid lines represent the best fits according to the generalized Debye model for two relaxations processes (Equations (S1) and (S2)).
Magnetochemistry 07 00057 g008
Figure 9. Field (up) and temperature (down) dependence of relaxation times for 3DyCo and 2DyFe. The solid lines correspond to the fit using QTM and Orbach mechanisms for the magnetic relaxation.
Figure 9. Field (up) and temperature (down) dependence of relaxation times for 3DyCo and 2DyFe. The solid lines correspond to the fit using QTM and Orbach mechanisms for the magnetic relaxation.
Magnetochemistry 07 00057 g009
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Dragancea, D.; Novitchi, G.; Mădălan, A.M.; Andruh, M. New Cyanido-Bridged Heterometallic 3d-4f 1D Coordination Polymers: Synthesis, Crystal Structures and Magnetic Properties. Magnetochemistry 2021, 7, 57. https://0-doi-org.brum.beds.ac.uk/10.3390/magnetochemistry7050057

AMA Style

Dragancea D, Novitchi G, Mădălan AM, Andruh M. New Cyanido-Bridged Heterometallic 3d-4f 1D Coordination Polymers: Synthesis, Crystal Structures and Magnetic Properties. Magnetochemistry. 2021; 7(5):57. https://0-doi-org.brum.beds.ac.uk/10.3390/magnetochemistry7050057

Chicago/Turabian Style

Dragancea, Diana, Ghenadie Novitchi, Augustin M. Mădălan, and Marius Andruh. 2021. "New Cyanido-Bridged Heterometallic 3d-4f 1D Coordination Polymers: Synthesis, Crystal Structures and Magnetic Properties" Magnetochemistry 7, no. 5: 57. https://0-doi-org.brum.beds.ac.uk/10.3390/magnetochemistry7050057

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