Investigation of slow magnetic relaxation in a series of 1D polymeric cyclobutane-1,1-dicarboxylates based on LnIIIVIV2 units (LnIII = Tb, Dy, Ho, Er, Tm, Yb): rare examples of VIV-4f single-molecule magnets

Evgeniya S. Bazhina *a, Maxim A. Shmelev a, Natalia V. Gogoleva a, Konstantin A. Babeshkin a, Ivan V. Kurganskii b, Nikolay N. Efimov a, Matvey V. Fedin b, Mikhail A. Kiskin a and Igor L. Eremenko a
aN.S. Kurnakov Institute of General and Inorganic Chemistry, Russian Academy of Sciences, Leninsky prosp. 31, Moscow 119991, Russian Federation. E-mail: bazhina@igic.ras.ru
bInternational Tomography Center, Siberian Branch of Russian Academy of Sciences, Institutskaya St. 3a, Novosibirsk 630090, Russian Federation

Received 18th June 2024 , Accepted 25th September 2024

First published on 24th October 2024


Abstract

The reactions of VOSO4·3H2O with Na2(cbdc) (cbdc2− – dianion of cyclobutane-1,1-dicarboxylic acid) and lanthanide(III) nitrates taken in a molar ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]2[thin space (1/6-em)]:[thin space (1/6-em)]1 were found to yield a series of isostructural heterometallic compounds [NaLn(VO)2(cbdc)4(H2O)10]n (1Ln, Ln = Tb, Dy, Ho, Er, Tm, Yb). These compounds are constructed from trinuclear anionic units [Ln(VO)2(cbdc)4(H2O)8] ({LnV2}) linked by Na+ ions into 1D polymeric chains. The crystal structures of 1Dy and 1Er were determined by single-crystal X-ray diffraction (XRD), and their isostructurality with 1Tb, 1Ho, 1Tm, and 1Yb was proved by powder X-ray diffraction (PXRD). According to alternating current (ac) magnetic susceptibility measurements, 1Dy, 1Er, and 1Yb exhibited field-induced slow relaxation of magnetization. Compound 1Er is the first representative of ErIII–VIV single-molecule magnets. Measuring the temperature dependences of the phase memory time (Tm) for 1Dy and 1Yb using pulsed EPR spectroscopy allowed us to observe the phenomenon of phase relaxation enhancement (PRE) at temperatures below 30 K. In future, this phenomenon may contribute to the evaluation of relaxation times of the lanthanide ions.


Introduction

Effective ways for the synthesis of heterometallic 3d–4f coordination compounds exhibiting properties of a bulk magnet at the molecular level are currently being developed.1 While constructing such complex molecules, paramagnetic 3d-metal ions are used, which are often involved in spin–spin exchange interactions via a super-exchange mechanism through a diamagnetic bridging ligand.2 Thus, the value of the exchange parameter affects the relaxation characteristics of a single-molecule magnet (SMM): strong exchange interactions enable reduction of the contribution of quantum tunneling of magnetization (QTM) to the relaxation of magnetization, increasing the value of the energy barrier for magnetization reversal, while the presence of weak and dipole–dipole interactions in the compound reduces the value of this parameter.3

Among all known SMMs based on heterometallic 3d–4f systems, the magnetic properties of CuII–LnIII complexes are the most well-studied; the influence of the geometric characteristics of the molecule on the parameter of exchange interactions was shown for these compounds.4 Similar to CuII, the VIV ion also has S = 1/2, however, the electronic structure of these ions is different: a single unpaired electron of VIV is located on the dxy orbital, but not on the dx2y2 orbital, as in the case of CuII (Fig. S1).5 In this regard, CuII–LnIII and VIV–LnIII compounds cannot be expected to exhibit identical magnetic properties. In addition, the geometric features of the VIVO2+ ion, i.e. the presence of an oxo group, exclude the formation of VIV–LnIII compounds similar in structure to their analogues containing the ions of CuII and other divalent 3d-metals. Despite ongoing studies of VIV–LnIII complexes, very few such compounds exhibiting SMM properties have been obtained so far.6 Therefore, the synthesis and detailed study of the magnetic properties of VIV–LnIII systems is an urgent task of modern coordination chemistry. The presence of a single unpaired d-electron also gives rise to the interest in VIV compounds as potential candidates for molecular-based spin qubits7 and makes them convenient objects for the study by EPR spectroscopy.

On the other hand, it is of interest to study 3d–4f SMMs containing lanthanide ions rarely used for these purposes, for example, ErIII and YbIII (see ref. 8) or non-Kramers lanthanide ions, HoIII and TmIII (see ref. 9 and 10), for which the coordination environment is of great importance to control the magnetic anisotropy. To date, 3d–4f SMMs with HoIII, ErIII, TmIII, and YbIII ions still remain much less studied than their TbIII and DyIII-containing counterparts.

Polydentate ligands that combine chelation with a variety of bridging coordination modes enable the constructing stable anionic blocks with atoms of 3d-elements and binding them with 4f-metal ions in a polynuclear molecule or coordination polymer without the use of additional ligands. An additional factor in the design of 3d–4f compounds can be the alkali metal ions introduced at the synthesis stage, which are diamagnetic, but often play a structure-directing role in the formation of the structure in the crystal, and therefore can influence the molecular geometry of the complex and, in particular, the coordination environment of paramagnetic metal centers. The structure-directing role of alkali metal ions has been studied for coordination polymers of s-elements,11 heteronuclear compounds of s-3d (see ref. 12) and s-4f metals.13 Very few such studies are known for heterometallic 3d–4f systems.14

In our previous studies, we investigated the influence of the ionic radii of MI (M = Na, K, Rb, Cs) and LnIII on the composition, structure, and magnetic properties of heterometallic compounds formed in the MI–LnIII–VIV systems with anions of cyclobutane-1,1-dicarboxylic acid (H2cbdc).6a,c,15 According to X-ray diffraction studies, alkali metal ions in this series of compounds affect not only the crystal packing of the compound, but also the geometric characteristics of the LnIII–VIV molecular fragments that form it, and, as a consequence, the exhibited magnetic properties. For NaI, the formation of 1D polymeric structures [NaLn(VO)2(cbdc)4(H2O)10]n was found in the systems with diamagnetic rare-earth metal ions YIII and LuIII (see ref. 16), as well as with GdIII (see ref. 15).

The present work is the logical continuation of our previous research, so it sets out to synthesize heterometallic NaI–LnIII–VIV compounds with paramagnetic lanthanide ions from TbIII to YbIII having high magnetic anisotropy and to study the slow magnetic relaxation phenomenon in the resulting complexes.

Results and discussion

Synthesis

The reactions of aqueous oxovanadium(IV) nitrate (prepared via metathesis between VOSO4·3H2O and Ba(NO3)2 in water), Na2(cbdc) and lanthanide(III) nitrates (TbIII, DyIII, HoIII, ErIII, TmIII, YbIII) in a molar ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]2[thin space (1/6-em)]:[thin space (1/6-em)]1 yielded blue crystals of a series of heterometallic compounds [NaLn(VO)2(cbdc)4(H2O)10]n (1Ln). Barium nitrate was added to the reaction mixture to remove sulfate anions from the solution, because the presence of these anions causes the crystallization of LnIII sulfates.

The description of crystal structures

According to PXRD data, all 1Ln compounds are isostructural (Fig. S2 in ESI). The crystal structures of 1Dy and 1Er were determined by single-crystal XRD. Compounds 1Dy and 1Er crystallize in monoclinic space group C2/c. The asymmetric units of 1Dy and 1Er contain one vanadium atom (V1), one lanthanide atom (Dy1 or Er1), and one sodium atom (Na1). In addition to the vanadyl oxo group, V1 atom coordinates two chelating cbdc2− anions and one water molecule, thus forming mononuclear bis-chelate {VO(cbdc)2(H2O)}2− moiety.

The geometry of vanadium coordination polyhedron is a distorted octahedron (VO6) whose equatorial plane is formed by four carboxylate O atoms. The O atoms of the oxo group and water molecule occupy the axial positions and form the shortened (∼1.60 Å) and the elongated (∼2.30 Å) bonds with a metal center, respectively (Fig. 1 and Table 1). The V1 atom deviates from the equatorial plane by 0.371 Å (in 1Dy) and 0.374 Å (in 1Er) towards oxo group, giving rise to the increase of V[double bond, length as m-dash]O/V–O(cbdc) bond angles and the decrease of V–O(cbdc)/V–O(H2O) ones (Table 2).


image file: d4dt01779j-f1.tif
Fig. 1 The fragment of the 1D polymeric chain of 1Dy (cyclobutane moieties are omitted for clarity) [symmetry codes: (a) 1 − x, y, 0.5 − z; (b) −x, y, 0.5 − z; (c) −1 + x, y, z; (d) 1 + x, y, z].
Table 1 Selected bond lengths and the shortest interatomic distances (d, Å) in structures 1Dy and 1Er
Compound 1Dy (Ln = Dy) 1Er (Ln = Er)
Bond/distance d
V[double bond, length as m-dash]O 1.600(2) 1.602(4)
V–O(cbdc) 1.965(2)–2.024(2) 1.965(4)–2.024(4)
V–O(H2O) 2.299(2) 2.310(4)
Ln–O(cbdc) 2.361(2) 2.335(3)
Ln–O(H2O) 2.349(2)–2.384(2) 2.317(4)–2.372(4)
Na–O(cbdc) 2.536(2), 2.637(2) 2.530(4), 2.638(4)
Na–O(H2O) 2.411(2), 2.447(2) 2.413(4), 2.449(4)
Ln⋯V 5.719(1) 5.699(2)
V⋯V 6.245(1) 6.249(3)


Table 2 Selected bond angles (ω, °) characterizing the coordination polyhedra of vanadium in complexes 1Dy and 1Er
Compound 1Dy 1Er
Angle ω
V[double bond, length as m-dash]O/V–O(cbdc) 100.00(10)–101.90(11) 99.98(18)–102.08(18)
V[double bond, length as m-dash]O/V–O(H2O) 177.65(10) 177.91(18)
V–O(cbdc)/V–O(cbdc) (acute) 85.62(9)–90.12(8) 85.39(15)–90.24(15)
V–O(cbdc)/V–O(H2O) (acute) 76.84(8)–81.88(8) 76.76(14)–81.73(14)


Each lanthanide atom binds to the two {VO(cbdc)2(H2O)}2− moieties via the coordination of two unchelated carboxylate O atoms. In the resulting {LnV2} trinuclear unit, the central lanthanide atom additionally coordinates six water molecules completing its polyhedron having the geometry of a triangular dodecahedron (TDD-8, the deviations from the ideal figure, CShM values are 0.630 and 0.640 for 1Dy and 1Er respectively) (see Table S1 in the ESI).17 Selected bond angles in the coordination polyhedra of lanthanides in structures 1Dy and 1Er are given in Tables S2 and S3 in the ESI.

The neighboring [Ln(VO)2(cbdc)4(H2O)8] units are linked into a 1D polymeric structure due to the coordination of Na atoms to the carboxylate O atoms involved in the chelation of vanadium (Fig. 1 and Table 1). Each Na atom coordinates two monodentate water molecules. The crystal structures of 1Dy and 1Er are additionally stabilized by the network of hydrogen bonds, whose formation involves all the coordinated water molecules, the carboxylate O atoms, and vanadyl oxo group (Tables S4 and S5 in the ESI).

Magnetic properties

The magnetic properties of the series of isostructural compounds 1Ln (Ln = Tb, Dy, Ho, Er, Tm, Yb) were studied by measuring the temperature dependences of molar magnetic susceptibility (χ) in the 2–300 K temperature range under 5000 Oe dc-magnetic field. For all compounds, the experimental χT values at 300 K are in satisfactory agreement with theoretical ones for two magnetically isolated VIV ions and one LnIII ion (Table 3). For TbIII, DyIII, HoIII, ErIII, and TmIII-containing compounds, the χT values at 300 K slightly exceed the theoretical ones, but generally fit into the range of acceptable deviations (about 10% from the corresponding theoretical χT value).
Table 3 The χT values for 1Ln under 5000 Oe field
Compound χT (theor.), cm3 K mol−1 χT (300 K), cm3 K mol−1 χT (2 K), cm3 K mol−1
1Tb (TbIIIVIV2) 12.58 13.74 6.76
1Dy (DyIIIVIV2) 14.93 15.06 8.49
1Ho (HoIIIVIV2) 14.83 15.15 4.68
1Er (ErIIIVIV2) 12.24 12.53 6.58
1Tm (TmIIIVIV2) 7.91 8.60 6.69
1Yb (YbIIIVIV2) 3.33 3.37 2.24


For 1Dy, 1Ho, 1Er (Fig. 2), the χT values monotonously decrease in the range from 300 to 100 K and then gradually decrease with decreasing temperature. On cooling below 10 K, the χT values sharply drop and reach a minimum at 2 K.


image file: d4dt01779j-f2.tif
Fig. 2 The experimental χT vs. T plots for compounds 1Ln in the range of 2–300 K under 5000 Oe field.

For 1Tb, a monotonous increase in the χT is observed in the range from 300 to 30 K, probably indicating the presence of weak ferromagnetic interactions. It should be noted that the field-induced orientation of polycrystals was excluded by using the mineral oil during the sample preparation (see Experimental part). With a further decrease in temperature, a sharp decrease in the χT value occurs, reaching a minimum at 2 K. For 1Tm, the χT value remains virtually constant up to 16 K and then sharply decrease with a further decrease in temperature down to 2 K.

For 1Yb, a monotonous decrease in the χT value is observed in the range from 300 to 2 K. Such a behavior of the compounds under study can be due to the possible presence of spin–spin antiferromagnetic interactions and/or the depopulation of the excited Stark sublevels.18 The M(H) and M(H/T) dependences for all obtained complexes were also measured at 2, 4, and 6 K (Fig. S3–S6 in the ESI).

In order to study magnetization relaxation of the compounds, ac-magnetic susceptibility measurements were carried out. In the absence of a dc-magnetic field, the values of the out-of-phase component of dynamic magnetic susceptibility (χ″) were close to zero for all the compounds, which may be due to a strong contribution from quantum tunneling to the relaxation of magnetization. Application of an external dc-field enabled to significantly reduce this effect and observe the χ″ non-zero values for 1Dy, 1Er, 1Tm, and 1Yb (Fig. S7–S12 in the ESI).

The highest relaxation times were achieved on applying the optimal fields of 1000 Oe for 1Dy, 1Er, 1Tm, and 2500 Oe for 1Yb (Fig. S13–S15). To produce the τ vs. 1/T plots, the χ″(ν) isotherms were approximated by the generalized Debye model (Fig. S16–S19 in the ESI). The plots of τ vs. 1/T thus obtained were approximated by the equations corresponding to different relaxation mechanisms and their combinations. In the high-temperature range, all the τ vs. 1/T dependences were approximated using only the Orbach relaxation mechanism (τ−1 = τ0−1·exp{−Δeff/kBT}) to estimate the value of the effective energy barrier (Fig. 3, 4 and 5).


image file: d4dt01779j-f3.tif
Fig. 3 The τ vs. 1/T plots for 1Dy under 1000 Oe field. Blue dashed lines represent the fittings of high-temperature ranges by the Orbach mechanism. Red solid lines represent the fittings in the whole temperature range by the sum of the Orbach and QTM relaxation mechanisms.

image file: d4dt01779j-f4.tif
Fig. 4 The τ vs. 1/T plots for 1Er under 1000 Oe field. Blue dotted line represents the fitting of a high-temperature range by the Orbach mechanism. Red solid line represents the fitting by the Raman relaxation mechanism.

image file: d4dt01779j-f5.tif
Fig. 5 The τ vs. 1/T plots for 1Yb under 2500 Oe field. Blue dotted line represents the fitting of a high-temperature range by the Orbach mechanism. Red solid line represents the fitting by the sum of the Raman and direct relaxation mechanisms.

According to the approximation of χvs. ν dependencies by the generalized Debye model (Fig. S16), there are at least two or even more relaxation processes for complex 1Dy also confirmed by Cole–Cole plots (Fig. S20 in the ESI). This may be due to the independent relaxation of DyIII and VIV ions19 and/or possible disorder of water molecules coordinated to DyIII (see ref. 14c and 20). Unfortunately, we failed to obtain isostructural analogue of complex 1Dy with diamagnetic d-metal ions (ZnII, CdII), so it was impossible to evaluate the contribution of DyIII ions to the magnetic relaxation dynamics of complex 1Dy. Previously, the magnetic properties of isostructural complexes [NaLn(VO)2(cbdc)4(H2O)10]n with diamagnetic rare-earth ions (Ln = YIII, LuIII) were studied.16 In both complexes, the presence of field-induced slow relaxation of magnetization was shown by ac-susceptibility measurements. This suggests the contribution of VIV ions to the magnetic relaxation dynamics in the case of complex 1Dy. Therefore, τ vs. 1/T plots for 1Dy were built using both low-frequency (LF) and high-frequency (HF) maxima of χvs. ν dependencies (Fig. 3). The good agreement between the experimental τ vs. 1/T plots and approximation equation can be achieved using parameters for the sum of the Orbach and QTM relaxation mechanisms (τ−1 = τ−10·exp{−Δeff/kBT} + B) both for LF and HF (Table 4).

Table 4 The best-fit parameters of magnetization relaxation for 1Dy
  Orbach Orbach + QTM
Δ eff/kB, K τ 0, s Δ eff/kB, K τ 0, s B, s−1
LF 50.4 ± 0.2 2.70 × 10−8 ± 8 × 10−10 52 (fixed) 2.1 × 10−8 ± 2 × 10−9 560 ± 11
HF 26 ± 2 4 × 10−7 ± 1 × 10−7 39 ± 3 6 × 10−8 ± 3 × 10−8 9066 ± 294


For 1Yb, the best-fit of the experimental τ vs. 1/T dependence in the whole temperature range was achieved by the sum of the Raman and direct relaxation mechanisms according to the equation τ−1 = CRamanTn_Raman + AdirectTH4 (Fig. 5). For 1Er, the corresponding fit was achieved with the use of only the Raman relaxation mechanism (τ−1 = CRamanTn_Raman) (Fig. 4).

For 1Tm, the χ″ values are less than the χ′ ones by more than 10 times (but χ′/χ″ ratio is close to 10), thus the presence of slow relaxation of magnetization is questionable in this case. All obtained magnetic relaxation data for 1Tm is presented in the ESI (Fig. S21 and Table S6).

The best-fit parameters for the approximations of τ vs. 1/T plots obtained for 1Er and 1Yb are given in Table 5.

Table 5 The best-fit parameters of magnetization relaxation for 1Er and 1Yb
Compound Orbach Raman + direct Raman
Δ eff/kB, K τ 0, s A direct, K−1 Oe−4 s−1 C Raman, s−1 Kn_Raman n_ Raman C Raman, c−1 Kn_Raman n_ Raman
1Er 19.2 ± 0.2 1.8 × 10−8 ± 1 × 10−9 41.0 ± 0.5 7 (fixed)
1Yb 44.8 ± 0.5 5.3 × 10−8 ± 4 × 10−9 1.20 × 10−11 ± 2 × 10−13 3.5 × 10−2 ± 4 × 10−3 7 (fixed)


It is worth pointing out that for 1Er, the value of nRaman = 7 is lower than the expected value for the Kramers systems (n = 9), indicating the presence of a Raman process through spin-phonon relaxation.21

In addition, for all compounds, the calculations of alternative magnetic relaxation parameters were performed using MagSuite v.3.2 software.22 The results obtained are presented in Fig. S22–S24 and Tables S7–S9.

The first VIV–DyIII SMM was described by K. Kotrle et al. (see ref. 6b), but the authors failed to determine possible relaxation mechanisms and estimate the effective energy barrier for this compound. The literature review showed that the value of Δeff/kB calculated for 1Dy is higher than those for the most known 3d-DyIII SMMs with paramagnetic 3d-metal ions and a similar triangular dodecahedral DyO8 coordination environment (Table 6).

Table 6 Parameters of slow magnetic relaxation of the reported 3d-DyIII SMMs with paramagnetic 3d-metal ions and triangular dodecahedral DyO8 coordination environment
Compound DyIIIO8 coordination polyhedron Δ eff/kB, K (Hdc, Oe) τ 0, s Ref.
[1]bpy = 2,2′-bipyridine; [2]NO2-benz = 3-nitrobenzoate; [3]hfac = 1,1,1,5,5,5-hexafluoroacetylacetonate; [4]bpca = bis(2-pyridylcarbonyl)amine; [5]H2vanox = o-vanillinoxime; [6]Br-benz = 4-bromobenzoate; [7]HL = 6-chloro-2-pyridinol; [8]piv = trimethylacetate; [9]H2L = N,N′-dimethyl-N,N′-bis(2-hydroxy-3,5-dimethylbenzyl)ethylenediamine; [10]PhCO2 = phenylacetate; [11]H2L = 6,6′-{(2-(dimethylamino)ethylazanediyl)-bis(methylene)}bis(2-methoxy-4-methylphenol); [12]H3L = N,N′-bis(3-methoxysalicylidene)-1,3-diamino-2-propanol; [13]OAc = acetate; [14]benz = benzoate; [15]H2L = (2-((2-hydroxy-3-methoxybenzylidene)amino)benzoic acid); [16]H2L = 2-{[(2-hydroxy-3-methoxybenzyl)imino]methyl}phenol; [17]H2L = 2-(benzothiazol-2-ylhydrazonomethyl)-6-methoxyphenol; [18]H3L = ligand formed from the in situ condensation reaction of 3-amino-1,2-propanediol with 2-hydroxy-1-naphthaldehyde; [19]phen = 1,10-phenanthroline; [20]mbenz = 3-methylbenzoate; [21]F-benz = 4-fluorobenzoate; [22]L = 1,3,5-Tris(2-di(2′-pyridyl)hydroxymethylphenyl)benzene; [23]OTf = trifluoromethanesulfonate; [24]4-tBubenz = 4-tert-butylbenzoate; [25]tBudeaH2 = N-tert-butyldiethanolamine; [26]Cl-benz = 4-chlorobenzoate; [27]ipO = 2-hydroxyisophthalate; [28]HCO2 = formate; [29]FcCO2 = ferrocenecarboxylate; [30]teaH3 = triethanolamine; [31]MePh = toluene; [32]HL = 8-hydroxyquinoline. a TDD-8 = triangular dodecahedron.b BTPR-8 = bicapped trigonal prism.c SAPR-8 = square antiprism.d Rough estimation using ln(χ″/χ′) = ln(2πντ0) + Δeff/kBT equation.e Recalculated from cm−1.
[DyIII2NiII2(bpy[1])2(NO2-benz[2])10] TDD-8a 2.8 (0) 5.47 × 10−6 23a
[{DyIII(hfac[3])3}2{NiII(bpca[4])2}]·CHCl3 TDD-8 4.9 (1000) 1.3 × 10−6 23b
[FeIII6DyIII3(OMe)9(vanox[5])6(Br-benz[6])6] TDD-8 4.9 (1000) 5.2 × 10−5 23c
[DyIII2CoII6(OH)4(L[7])6(piv[8])8(MeCN)2]·0.5CH2Cl2 TDD-8 7.7d (1000) 5.7 × 10−8 23d
[CoII2(L[9])2(PhCO2[10])2DyIII2(hfac)4] TDD-8 8.8d (0) 2.0 × 10−7 23e
7.8d (1000) 3.9 × 10−7
[{DyIII(hfac)3}2{FeII(bpca)2}]·CHCl3 TDD-8 9.7 (1000) 8.7 × 10−8 23b
[NiII3DyIII3(O)(OH)3(L[11])3(piv)3](ClO4)·8MeCN·3CH2Cl2·5.5H2O TDD-8 ∼10 (3000) ∼10−6 23f
[NiII2DyIII2(CO3)2(HL[12])(EtOH)(OAc[13])]·2EtOH TDD-8 11.52 (1200) 5.01 × 10−6 23g
[FeIII6DyIII3(OMe)9(vanox)6(benz[14])6] TDD-8 12.4 (2000) 8.0 × 10−5 23h
[CoII4DyIII4(L[15])4(piv)8(OH)4(MeOH)2] H2O·3MeOH TDD-8 12.5 (0) 1.51 × 10−6 23i
[DyIII2NiII2MnIII2(L[16])4(OAc)2(OH)4(MeOH)2](NO3)2·2MeOH TDD-8 13.0 (0) 2.8 × 10−7 23j
[DyIII2NiII4(L[16])4(OAc)2(OH)4(MeOH)2]·4MeOH TDD-8 13.4 (0) 3.4 × 10−7 23j
[DyIII2NiII2(OH)3(OAc)4(HL[17])2(MeOH)3](ClO4)3 3MeOH TDD-8 ↔ BTPR-8b 7.6 (1200) 7.5 × 10−6 23k
[DyIII2CoII8(OMe)2(L[18])4(HL[18])2(OAc)2(NO3)2(MeCN)2]·MeCN·H2O TDD-8 14.89 (0) 1.68 × 10−7 23l
[DyIIIFeII(H2O)(phen[19])(mbenz[20])5] TDD-8 17 (3000) 2.6 × 10−9 23m
[DyIIINiII(H2O)(phen)(mbenz)5] TDD-8 20 (5000) 1.38 × 10−8 23m
[FeIII6DyIII3(OMe)9(vanox)6(F-benz[21])6] TDD-8, TDD-8 ↔ SAPR-8c 21.3 (1500) 4.1 × 10−7 23c
[DyIII2NiII2(bipy)2(mbenz)10] TDD-8 25.9 (0) 1.16 × 10−6 23a
[(L[22])DyIIIMnIV3O4(OAc)3(DMF)2](OTf[23]) TDD-8 27 (0) 2.13 × 10−8 23n
[CrIII2DyIII2(OMe)(OH)(4-tBubenz[24])4(tBudea[25])2(NO3)2]·MeOH·2Et2O TDD-8 ↔ SAPR-8 31.3e (0) 7.7 × 10−8 23o
[FeIII6DyIII3(OMe)9(vanox)6(Cl-benz[26])6] TDD-8, TDD-8 ↔ SAPR-8 36.1 (2000) 3.4 × 10−7 23c
[DyIII2NiII2(bpy)2(benz)10] TDD-8 39.9 (0) 1.80 × 10−8 23a
[NaDy III (V IV O) 2 (cbdc) 4 (H 2 O) 10 ] n TDD-8 LF: 50.4 (1000) 2.70 × 10 −8 This work
HF: 26 (1000) 4 × 10 −7
[DyIII2CuII6(ipO[27])6(H2O)12]n TDD-8 63.68 (2000) 3.77 × 10−8 23p
[MnIV3MnIII18DyIIIO20(OH)2(piv)20(HCO2[28])4(NO3)3(H2O)7]·5MeNO2 H2O TDD-8 74 (0) 2.0 × 10−12 23q
[DyIII2CrIII2(OH)2(FcCO2[29])4(NO3)2(Htea[30])2]·2MePh[31]·2THF TDD-8 75 (0) 26 × 10−9 23r
[NiII6DyIII(L[32])8(OAc)2(NO3)(OH)2(OMe)2] TDD-8 122.73 (0) 7.64 × 10−13 23s


Considering the previously obtained magnetic data for potassium-containing analogues of 1Dy and 1Yb (see ref. 6a), it can be concluded that in VIV–LnIII systems with cbdc2−, the substitution of potassium by sodium ions giving rise to a significant change in the crystal structure and coordination environment of the lanthanide ion has a positive influence on their SMM behavior. For 1Dy, the appearance of slow magnetic relaxation is observed compared to KI–VIV–DyIII compound. One of the possible explanations for such differences in the magnetic behavior of two Dy-containing compounds may be the difference in DyIII coordination polyhedra, which is a biaugmented trigonal prism (C2v symmetry) in KI–DyIII–VIV and a triangular dodecahedron (D2d symmetry) in 1Dy. Another factor influencing SMM behavior is supposed to be the longer Dy⋯V distances in 1Dy (5.719 Å) compared to KI–VIV–DyIII (4.627 Å), that allow weakening of dipole–dipole interactions between DyIII and VIV ions.

For 1Yb, the increase in the Δeff/kB value to 44.8 K occurs compared to the KI–YbIII–VIV compound (Δeff/kB = 26 K), although, in these compounds, the YbIII coordination polyhedra have similar geometry (triangular dodecahedron) and the shortest Yb⋯V distances are also similar (5.684 Å in KI–YbIII–VIV and ∼5.7 Å in 1Yb). Thus, the possible influence of crystal packing and intermolecular interactions on SMM behavior can be assumed in this case.

The literature review showed that 3d-ErIII and 3d-YbIII SMMs containing paramagnetic 3d-metal ions are quite rare (Tables 7 and 8). To date, compound 1Er is the first representative of heterometallic VIV–ErIII single-molecule magnets. The value of Δeff/kB calculated for 1Er is comparable with those for compounds with triangular dodecahedral ErO8 coordination environment (Table 7). Among all reported heterometallic compounds of such type, 1Yb displays the record value of Δeff/kB (Table 8).

Table 7 The parameters of slow magnetic relaxation of the reported 3d-ErIII SMMs with paramagnetic 3d-metal ions
Compound ErIII environment, coordination polyhedron Δ eff/kB, K (Hdc, Oe) τ 0, s Ref.
[1]2-PNO = 2-picoline-N-oxide; [2]piv = trimethylacetate; [3]2-ma = 2-methylalanine; [4]H2pmide = N-(2-pyridylmethyl)iminodiethanol; [5]p-Me-benz = 4-methylbenzoate; [6]o-tol = o-toluate; [7]H3L = (E)-2-(hydroxymethyl)-6-(((2-hydroxyphenyl)imino)methyl)-4-methylphenol; [8]H2L = N,N′-dimethyl-N,N′-bis(2-hydroxy-3-formyl-5-bromo-benzyl)ethylenediamine; [9]OAc = acetate; [10]H6L = 2,2′-(propane-1,3-diyldiimino)bis[2-(hydroxylmethyl)propane-1,3-diol]; [11]{HB(pz)3} = hydrotris(pyrazolyl)borate; [12]pyim = 2-(1H-imidazol-2-yl)pyridine; [13]Ph3PO = triphenylphosphineoxide; [14]HL = 3-methoxy-N-[2-(methylsulfanyl)phenyl]salicylaldimine. a PBPY-7 = pentagonal bipyramid.b SAPR-8 = square antiprism.c TDD-8 = triangular dodecahedron.d CSAPR-9 = capped square antiprism.e MFF-9 = muffin.f JBCSAPR-10 = bicapped square antiprism.g Recalculated from cm−1.
{[FeIIIErIII(CN)6(2-PNO[1])5]·4H2O}n ErN2O5, PBPY-7a 43.55 (1000) 2.10 × 10−9 24a
(Et3NH)2[NiII2ErIII2(OH)2(piv[2])10] ErO8, SAPR-8b 18 (1000) 3.9 × 10−6 3
[CuII8ErIII(OH)8(2-ma[3])8Cl2](ClO4)·21H2O ErO8, SAPR-8 22.9 (0) 4.74 × 10−7 24b
33 (1000) 9.48 × 10−6
[FeIII2ErIII2(OH)2(pmide[4])2(p-Me-benz[5])6]·2MeCN ErN2O6, SAPR-8 16.51 (1000) 2.03 × 10−7 24c
[CrIIIErIII6(OH)8(o-tol[6])12(NO3)(MeOH)5]·2MeOH ErO8, TDD-8c 4.5 (3000) 9.1 × 10−8 24d
[NaEr III (V IV O) 2 (cbdc) 4 (H 2 O) 10 ] n ErO 8 , TDD-8 19.2 (1000) 1.8 × 10 −8 This work
[NiII4ErIII(L[7])2(HL[7])2(MeCN)3Cl]·2H2O·2MeCN ErO8, TDD-8 31.87g (4000) 7.94 × 10−11 24e
[NiIIErIII(L[8])(OAc[9])(NO3)2(MeCN)]·MeCN ErO9, CSAPR-9d 11.91 (1000) 5.12 × 10−8 24f
(NMe4)2[CuII3ErIII2(H3L[10])2(NO3)7(MeOH)2](NO3) ErO9, CSAPR-9 14.8 (0) 1.2 × 10−7 24g
[FeIIIErIII{HB(pz)3}[11](CN)3(NO3)2(pyim[12])(Ph3PO[13])]2·2MeCN ErN4O5, MFF-9e 57.6g (2500) 24h
[NiIIErIII(L[14])2(NO3)3]·0.5H2O ErO10, JBCSAPR-10f 12.1 (1000) 3.49 × 10−7 24i


Table 8 The parameters of slow magnetic relaxation of the reported 3d-YbIII SMMs with paramagnetic 3d-metal ions
Compound YbIII environment, coordination polyhedron Δ eff/kB, K (Hdc, Oe) τ 0, s Ref.
[1]phen = 1,10-phenanthroline; [2]OTf = trifluoromethanesulfonate; [3]AcrCN = acrylonitrile; [4]PrCN = propionitrile; [5]MalCN = malononitrile; [6]piv = trimethylacetate; [7]H2butyrat = 3-aminobutyric hydroxamic acid; [8]2-ma = 2-methylalanine. a OC-6 = octahedron.b SAPR-8 = square antiprism.c TDD-8 = triangular dodecahedron.
{YbIII(4-pyridone)4[FeII(phen[1])2(CN)2]2}(OTf[2])3·2MeCN YbN2O4, OC-6a 12.5/800 7.28 × 10−6 25a
{YbIII(4-pyridone)4[FeII(phen)2(CN)2]2}(OTf)3·2AcrCN[3] YbN2O4, OC-6 7.86/800 2.51 × 10−5 25a
{YbIII(4-pyridone)4[FeII(phen)2(CN)2]2}(OTf)3·2PrCN[4] YbN2O4, OC-6 10.28/800 1.46 × 10−5 25a
{YbIII(4-pyridone)4[FeII(phen)2(CN)2]2}(OTf)3·2MalCN[5]·MeOH YbN2O4, OC-6 4.83/800 5.82 × 10−5 25a
[Na2YbIII2CuII2(OH)2(piv[6])10(EtOH)2]·EtOH YbO8, SAPR-8b 8.5/1000 2.1 × 10−6 14c
[YbIII{CuII4(butyrat[7])4}2]Cl3·MeOH·26H2O YbO8, SAPR-8 6.84/1000 1.04 × 10−5 25b
[YbIIICuII8(OH)8(2-ma[8])8(Cl)2](ClO4)·21H2O YbO8, SAPR-8 22.5/700 1.48 × 10−8 24b
{[KYb(VO)2(cbdc)4(H2O)11]·2H2O}2 YbO8, TDD-8c 23/2000 5.6 × 10−7 6a
[NaYb(VO) 2 (cbdc) 4 (H 2 O) 10 ] n YbO 8 , TDD-8 44.8/2500 5.3 × 10 −8 This work


EPR spectroscopy of 1Yb and 1Dy

Continuous wave (CW) EPR spectra of 1Yb and 1Dy at room temperature are typical for oxovanadium(IV) complexes and display hyperfine structure of VIV ion (Fig. 6). No signatures of YbIII and DyIII ions are observed at room temperature, but the spectrum is dominated by VIV EPR signal. This signal can be simulated using typical26g- and A-tensors (the latter refers to the hyperfine interaction tensors): g = [1.975 1.975 1.938], A = [185 185 520] MHz for 1Dy, and g = [1.974 1.974 1.941], A = [164 164 521] MHz for 1Yb. Reference compound 1Y with diamagnetic rare-earth metal ion (YIII) shows almost the same EPR signal of VIV at room temperature and can be simulated using the very similar set of parameters g = [1.974 1.974 1.938], A = [184 184 518] MHz, which agrees well with previous data.16
image file: d4dt01779j-f6.tif
Fig. 6 CW EPR spectra of 1Dy, 1Yb and 1Y at 293 K and 10 K. Simulations are shown in red.

However, as the temperature lowers, EPR spectra of both 1Yb and 1Dy become broader, resulting in one line with unresolved structure at 10 K (Fig. 6). This trend is unusual as most relaxation processes become slower at low temperatures, leading to the narrowing of EPR lines. Remarkably, the EPR spectrum of reference compound 1Y still shows a resolved hyperfine structure at 10 K; therefore, drastic broadening of EPR spectra in cases of 1Yb and 1Dy at 10 K should be assigned to the interactions between VIV and YbIII/DyIII ions. Note that, in addition to VIV signal, the 10 K spectrum of 1Yb shows a small feature at ∼100 mT, that is tentatively assigned to the contribution of YbIII.27

These interactions are more clearly evident in pulse EPR. In order to complement ac-magnetic susceptibility data and shed light on faster processes on micro- and submicroseconds timescales, we performed measurements of phase memory time (Tm) for 1Yb and 1Dy at 10–60 K. Two-pulse (Hahn) echo was monitored as a function of interpulse delay, and stretched exponential analysis (image file: d4dt01779j-t1.tif, β = 2 ± 0.5) was then employed to obtain corresponding Tm values.

Fig. 7 shows the obtained Tm(T) dependences for 1Yb, 1Dy and the reference compound of molecular structure [KY(VO)2(cbdc)4(H2O)11]·2H2O (MY)6a,16 with diamagnetic rare-earth metal ion (see the ESI for details and choice of MY). The Tm(T) dependence is observed to have a non-monotonous behavior for 1Yb and 1Dy: the relaxation accelerates leading to a decrease of Tm values, reaching minima at T ∼ 10–12 K. Moreover, reference compound MY with diamagnetic YIII ion shows perfectly monotonous dependence without such peculiarity. Again, this means that the observed behavior for 1Yb and 1Dy owes to the interactions between VIV and YbIII/DyIII. This also confirms that the VIV–YbIII and VIV–DyIII units are present when 1Yb and 1Dy are dissolved in water/glycerol, since otherwise their Tm(T) dependences would be similar to that of MY.


image file: d4dt01779j-f7.tif
Fig. 7 T m(T) dependences for 1Yb, 1Dy and the reference compound MY. Solid line guides the eye.

In fact, such phenomenon is generally known in literature and is called phase relaxation enhancement (PRE), i.e. an increase of the relaxation rate (decrease of Tm) induced by a partner spin coupled with observer spin by dipolar interaction.28 In compounds 1Yb and 1Dy, we deal with the spins of two types – slow-relaxing S = 1/2 spins of VIV, and much faster relaxing spins of YbIII or DyIII. If the spin of lanthanide ion relaxes (flips) much faster than that of vanadium, the dipolar interaction will be averaged and no effect on vanadium EPR should be observed.

This situation corresponds to CW EPR spectra of 1Yb and 1Dy at room temperature and to the Tm(T) dependences at T > 30 K. In another limit, if (hypothetically) lanthanide spin relaxed too slowly, there should be no PRE of the vanadium spin as well (this situation is not reached experimentally). However, at intermediate relaxation rate of the lanthanide spins the influence of such fluctuations on the Tm value of vanadium ion is anticipated, due to the dipolar coupling between these ions. At the same time, CW EPR spectrum should broaden due to the contribution of lanthanide.

Previous theoretical consideration of similar phenomena derived general expression for the electron spin echo decay of slow-relaxing spin in the presence of dipolarly-coupled fast-relaxing spin (see ref. 28a):

 
image file: d4dt01779j-t2.tif(1)
where R2 = W2A2(r)/4, image file: d4dt01779j-t3.tif and image file: d4dt01779j-t4.tif with θ12 being the angle between image file: d4dt01779j-t5.tif (image file: d4dt01779j-t6.tif for the target compounds) and image file: d4dt01779j-t7.tif. In particular, for R = 0, which corresponds to the maximum PRE effect (minimum at Tm(T) dependence), the phase relaxation is enhanced up to:
 
image file: d4dt01779j-t8.tif(2)

The eqn (2) qualitatively explains the behavior observed for 1Yb and 1Dy in Fig. 7. The analysis of experimental data allows one to potentially obtain the unique information on spin relaxation times of the lanthanide ions, which are hardly available otherwise being often too short to measure by EPR. However, quantification of this approach requires more work. For instance, eqn (2) should result in a decrease of Teffm down to ≈TLn1, which can be estimated as ∼10 ns at PRE maximum (image file: d4dt01779j-t9.tif and A(r) ∼ 200 MHz in point-dipole approximation based on the crystal structures). This short Tm values are not observed experimentally, meaning that more experimental factors should be theoretically taken into account to describe PRE in LnIII–VIV complexes. First, when 1Yb and 1Dy are dissolved for pulse EPR measurements, one should ensure that there is only one type of spin pairs (or spin triads) present in frozen solution, because if a part of the compound is fully dissolved and separate vanadium, and rare-earth blocks are present, the apparent Tm(T) will have two contributions which should be treated properly. Second, a distribution over parameters TLn1 and gz of the pairs (eqn (1)) should be significantly broad29 and be treated accordingly. The other theoretical challenges are the proper introduction of an atom with strong spin–orbit coupling (relevant for all lanthanides) into the framework of the current PRE-theory and high sensitivity of Tm to the minor changes in the environment.29 The optimization of the theory might be the topic of our future study. At the same time, the development of clear manifestations of PRE in 1Yb and 1Dy complexes potentially outlooks the use of such phenomena in complex characterization of relaxation times in molecular magnet candidates.

Conclusions

In the NaI–LnIII–VIV system (LnIII = Tb, Dy, Ho, Er, Tm, Yb) with cyclobutane-1,1-dicarboxylate anions (cbdc2−), the lanthanide ionic radius was found to have no impact on the structure of the resulting heterometallic compound. All six new LnIII–VIV compounds obtained have the same 1D polymeric structure formed by trinuclear anionic units [Ln(VO)2(cbdc)4(H2O)8] linked by Na+ ions.

According to ac-magnetic susceptibility measurements, the DyIII-, ErIII-, and YbIII-containing compounds showed field-induced slow relaxation of magnetization. Slow magnetic relaxation observed can be best described by the sum of the Orbach and Raman relaxation mechanisms for DyIII–VIV complex, the sum of Raman and direct relaxation mechanisms for YbIII–VIV complex, and only the Raman relaxation mechanism for ErIII–VIV one.

For DyIII–VIV, two relaxation processes were suggested, which may result from the independent relaxation of DyIII and VIV centers and/or possible disorder of water molecules coordinated to DyIII. For complexes with TbIII, HoIII, and TmIII slow magnetic relaxation was not observed due to the possible appearance of weak intramolecular and/or dipole–dipole exchange interactions.

The ErIII-containing complex is the first representative of heterometallic ErIII–VIV compounds exhibiting slow magnetic relaxation.

For DyIII–VIV and YbIII–VIV studied by EPR spectroscopy, the phenomenon of phase relaxation enhancement (PRE) was observed, which can be used for complex characterization of relaxation times in molecular magnet candidates.

Experimental

Materials and methods

New compounds were synthesized in air, using distilled water as the solvent. Starting reagents included VOSO4·3H2O (>99%), Ba(NO3)2 (>98%), cyclobutane-1,1-dicarboxylic acid (H2cbdc, 99%, Acros Organics), NaOH (>99%), Tb(NO3)3·6H2O (99.9%, Lanhit), Dy(NO3)3·5H2O (99.9%, Lanhit), Ho(NO3)3·5H2O (99.9%, Lanhit), Er(NO3)3·5H2O (99.9%, Lanhit), Tm(NO3)3·5H2O (99.9%, Lanhit), Yb(NO3)3·5H2O (99.9%, Lanhit).

The infrared spectra of complexes 1Ln were recorded in the frequency range of 4000–400 cm−1 on a PerkinElmer Spectrum 65 Fourier transform infrared spectrometer equipped with a Quest ATR Accessory (Specac). Elemental analysis of the compounds synthesized was carried out on a EuroEA 3000 CHNS analyzer (EuroVector, S.p.A.).

The purity of compound samples was approved by powder X-ray diffraction. The patterns were measured on a Bruker D8 Advance diffractometer with a LynxEye detector in the Bragg–Brentano geometry, with the samples dispersed thinly on a zero-background Si sample holder, λ(CuKα) = 1.54060 Å, θ/θ scan with variable slits (beam length is 20 mm) in the 2θ-angle range from 5° to 50°, with a step size of 0.020°.

The magnetic properties of compounds 1Ln were studied in the dc- and ac-modes on a Quantum Design PPMS-9 magnetometer in the temperature range of 2–300 K. Dc-magnetic fields with an intensity of 0–5000 Oe and ac-magnetic fields with intensity of 5 Oe, 3 Oe and 1 Oe within frequency ranges 10–100, 100–1000 and 1000–10[thin space (1/6-em)]000 Hz, respectively, were applied using standard procedure.30 All magnetic behavior studies were performed using ground polycrystalline samples, sealed in polyethylene bags and frozen in mineral oil to prevent the orientation of crystallites in a magnetic field. The paramagnetic component of the magnetic susceptibility (χ) was determined taking into account the diamagnetic contribution of the sample, evaluated from Pascal's constant, and the diamagnetic contributions of the mineral oil and the sample holder.

All EPR data were collected using Bruker Elexsys E580 spectrometer at X-band (9 GHz) at the Center of Collective Use “Mass spectrometric investigations” SB RAS. The spectrometer was equipped with helium flow cryostat and temperature control system (4–300 K). Continuous wave EPR spectra were obtained on polycrystalline powder samples under conditions avoiding microwave saturation and modulation broadening. Phase memory time was measured using two-pulse Hahn electron spin echo sequence for glassy water/glycerol (C ∼ 0.2 mM) solutions of target compounds. In all cases samples were placed into quartz sample tubes and studied. Simulations were performed using EasySpin.31

General synthesis procedure for [NaLn(VO)2(cbdc)4(H2O)10]n (1Ln, Ln = Tb, Dy, Ho, Er, Tm, Yb). A weighed sample of VOSO4·3H2O (0.100 g, 0.46 mmol) was dissolved in H2O (15 mL), then Ba(NO3)2 (0.120 g, 0.46 mmol) was added, and the reaction mixture was stirred for 20 min at 40 °C. The solution of Na2(cbdc) prepared by neutralization of H2cbdc (0.133 g, 0.92 mmol) with NaOH (0.074 g, 1.84 mmol) in H2O (10 mL) was added to the reaction mixture, and the stirring was continued. After 10 min Ln(NO3)3·xH2O (m g, 0.46 mmol) was added. The reaction mixture was stirred for additional 10 min and allowed to stand for 1 hour, then BaSO4 precipitate was removed by filtration. The resulting blue solution (25 mL) was allowed to evaporate slowly in air at 22 °C. X-ray quality blue crystals were obtained within 2 months. The crystals were separated from the mother liquor by filtration, washed with cold H2O (t = 3 °C) and dried in air at 22 °C.

For 1Tb: x = 6, m = 0.208. The yield was 0.113 g (46.3% based on VOSO4·3H2O). Anal. Calc for C24H44NaO28TbV2: C, 27.08; H, 4.17. Found: C, 27.01; H, 4.26%. IR (ATR), ν/cm−1: 3642 w, 3348 br. m [ν(O–H)], 3234 m [ν(O–H)], 3000 w [ν(C–H)], 2957 w [ν(C–H)], 1634 m, 1582 s [νas(COO)], 1555 s [νas(COO)], 1443 m, 1431 m, 1391 s [νs(COO)], 1349 s, 1254 m, 1242 m, 1229 m [ν(C–C)cycle], 1195 w, 1162 w, 1122 m [γ(C(–C)2)], 1061 w, 1012 w, 1000 w, 968 s [ν(V[double bond, length as m-dash]O)], 952 s, 924 s, 875 w, 843 w, 807 w, 773 m, 762 m, 725 s [δ(COO)], 647 s, 560 s, 533 s, 471 s, 444 s, 418 s.

For 1Dy: x = 5, y = 0.202. The yield was 0.118 g (47.8% based on VOSO4·3H2O). Anal. Calc for C24H44DyNaO28V2: C, 26.99; H, 4.15. Found: C, 27.08; H, 4.20%. IR (ATR), ν/cm−1: 3641 vw, 3351 br. m [ν(O–H)], 3229 m [ν(O–H)], 2999 w [ν(C–H)], 2956 w [ν(C–H)], 1631 m, 1581 vs [νas(COO)], 1554 vs [νas(COO)], 1443 m, 1431 m, 1390 s [νs(COO)], 1348 s, 1254 m, 1242 m, 1229 m [ν(C–C)cycle], 1196 w, 1161 w, 1122 m [γ(C(–C)2)], 1061 w, 1012 w, 1000 w, 968 s [ν(V[double bond, length as m-dash]O)], 952 s, 924 s, 874 w, 843 w, 807 w, 773 m, 762 m, 725 s [δ(COO)], 649 vs, 604 s, 561 vs, 533 vs, 468 s, 450 vs, 440 vs, 415 s, 403 vs.

For 1Ho: x = 5, y = 0.203. The yield was 0.126 g (51.2% based on VOSO4·3H2O). Anal. Calc for C24H44HoNaO28V2: C, 26.93; H, 4.14. Found: C, 26.90; H, 4.19%. IR (ATR), ν/cm−1: 3641 vw, 3358 br. m [ν(O–H)], 3234 m [ν(O–H)], 3000 w [ν(C–H)], 2957 w [ν(C–H)], 1634 m, 1580 s [νas(COO)], 1557 s [νas(COO)], 1443 m, 1431 m, 1391 s [νs(COO)], 1349 s, 1254 w, 1242 w, 1230 m [ν(C–C)cycle], 1193 w, 1162 w, 1123 m [γ(C(–C)2)], 1061 v.w, 1012 w, 1000 w, 968 s [ν(V[double bond, length as m-dash]O)], 952 s, 924 m, 875 w, 843 w, 807 w, 773 m, 762 m, 725 s [δ(COO)], 648 s, 561 s, 533 s, 470 s, 448 s, 438 s, 415 s, 403 vs.

For 1Er: x = 5, y = 0.204. The yield was 0.113 g (45.8% based on VOSO4·3H2O). Anal. Calc for C24H44ErNaO28V2: C, 26.87; H, 4.13. Found: C, 26.79; H, 4.19%. IR (ATR), ν/cm−1: 3641 vw, 3356 br. m [ν(O–H)], 3234 m [ν(O–H)], 3000 w [ν(C–H)], 2957 w [ν(C–H)], 1634 m, 1580 s [νas(COO)], 1555 s [νas(COO)], 1443 m, 1432 m, 1390 s [νs(COO)], 1348 s, 1254 m, 1242 m, 1229 m [ν(C–C)cycle], 1193 w, 1162 w, 1122 m [γ(C(–C)2)], 1063 w, 1012 w, 1000 w, 968 s [ν(V[double bond, length as m-dash]O)], 953 m, 924 m, 875 w, 843 w, 807 w, 773 m, 762 m, 725 m [δ(COO)], 650 s, 595 s, 561 s, 532 s, 467 s, 446 s, 437 s, 420 s, 407 vs.

For 1Tm: x = 5, y = 0.205. The yield was 0.077 g (31.2% based on VOSO4·3H2O). Anal. Calc for C24H44NaO28TmV2: C, 26.83; H, 4.13. Found: C, 26.94; H, 4.14%. IR (ATR), ν/cm−1: 3639 vw, 3352 br. m [ν(O–H)], 3238 m [ν(O–H)], 3000 w [ν(C–H)], 2956 w [ν(C–H)], 1634 m, 1583 s [νas(COO)], 1557 s [νas(COO)], 1443 m, 1431 m, 1391 s [νs(COO)], 1348 s, 1254 m, 1242 m, 1229 m [ν(C–C)cycle], 1196 w, 1163 w, 1123 m [γ(C(–C)2)], 1063 w, 1012 w, 1000 w, 968 s [ν(V[double bond, length as m-dash]O)], 953 s, 924 m, 874 w, 843 w, 807 w, 773 m, 765 m, 725 s [δ(COO)], 653 s, 595 s, 561 s, 533 s, 471 s, 448 s, 423 s.

For 1Yb: x = 5, y = 0.207. The yield was 0.120 g (48.4% based on VOSO4·3H2O). Anal. Calc for C24H44NaO28V2Yb: C, 26.73; H, 4.11. Found: C, 26.67; H, 4.08%. IR (ATR), ν/cm−1: 3639 vw, 3354 br. m [ν(O–H)], 3229 m [ν(O–H)], 3000 w [ν(C–H)], 2956 w [ν(C–H)], 1634 m, 1582 s [νas(COO)], 1554 s [νas(COO)], 1443 m, 1431 m, 1391 s [νs(COO)], 1349 s, 1254 m, 1242 m, 1229 m [ν(C–C)cycle], 1196 w, 1162 w, 1123 m [γ(C(–C)2)], 1063 w, 1012 w, 1000 w, 968 s [ν(V[double bond, length as m-dash]O)], 953 s, 924 m, 875 w, 843 w, 807 w, 773 m, 762 m, 726 s [δ(COO)], 654 s, 561 s, 533 s, 466 s, 448 s, 440 s, 425 s, 416 s.

X-ray crystallography

The X-ray diffraction data sets for compounds 1Dy and 1Er were collected on a Bruker SMART APEX II diffractometer equipped with a CCD detector (Mo-Kα, λ = 0.71073 Å, graphite monochromator).32 A semiempirical absorption correction was applied using SADABS program.33 The structures were solved by direct methods and refined by the full-matrix least squares with anisotropic displacement parameters for non-hydrogen atoms. The hydrogen atoms of the OH groups were determined from the difference Fourier maps; with other hydrogen atoms calculated geometrically and refined using a riding model. The calculations were performed with the SHELX-2014 program package34via OLEX2 1.3 graphical user interface.35 The crystallographic data for 1Dy, 1Er, and the structure refinement statistics are given in Table 9.
Table 9 Crystallographic parameters and structure refinement statistics for compounds 1Dy and 1Er
Parameter 1Dy 1Er
a R 1 = ∑||Fσ| − |Fc||/∑|Fo|. b wR2 = [∑w(Fo2Fc2)2/∑w(Fo2)2]1/2.
Empirical formula C24H44DyNaO28V2 C24H44ErNaO28V2
Formula weight (g mol−1) 1067.96 1072.72
T (K) 150
Crystal system Monoclinic
Space group C2/c
a (Å) 9.097(2) 9.088(4)
b (Å) 24.739(5) 24.681(11)
c (Å) 17.116(3) 17.098(8)
β (°) 104.682(7) 104.589(8)
V3) 3726.2(14) 3711(3)
Z 4 4
D calc (g cm−3) 1.904 1.920
θ min –θ max (°) 2.97–33.14 2.46–31.83
μ (mm−1) 2.59 2.85
No. of measured, independent and observed [I > 2σ(I)] reflections 7547, 3639, 3336 6910, 3159, 2796
R int 0.026 0.040
GOF 1.045 1.036
R 1[thin space (1/6-em)]a, wR2[thin space (1/6-em)]b (I > 2σ(I)) 0.0259, 0.0548 0.0405, 0.0991
R 1[thin space (1/6-em)]a, wR2[thin space (1/6-em)]b (all data) 0.0298, 0.0567 0.0479, 0.1035
T min, Tmax 0.626, 0.747 0.456, 0.745
Δρmax, Δρmin (e Å−3) 0.99, −1.01 1.98, −1.34


CCDC 2266768 and 2266772 contain the supplementary crystallographic data for 1Dy and 1Er.

Author contributions

Investigation, E. S. B., M. A. S., N. V. G., M. A. K., K. A. B., I. V. K.; formal analysis, E. S. B., M. A. S., N. V. G., M. A. K., K. A. B., I. V. K., M. V. F.; methodology, E. S. B., M. A. S., N. V. G., M. A. K., K. A. B., I. V. K., M. V. F., N. N. E.; visualization, E. S. B., K. A. B., I. V. K., M. V. F.; writing – original draft, E. S. B., M. A. S., K. A. B., N. N. E., M. A. K., I. V. K., M. V. F.; writing – review and editing, E. S. B., N. N. E., M. A. K., M. V. F., I. L. E.; supervision, conceptualization: E. S. B., M. A. K., M. V. F., I. L. E. All authors have read and agreed to the published version of the manuscript.

Data availability

The data supporting this article have been included as part of the ESI.

Crystallographic data for 1Dy and 1Er have been deposited at the Cambridge Crystallographic Data Centre under CCDC 2266768 and 2266772 numbers and can be obtained from https://www.ccdc.cam.ac.uk/conts/retrieving.html.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work was supported by the Ministry of Science and Higher Education of the Russian Federation as part of the State Assignment of the Kurnakov Institute of General and Inorganic Chemistry of the Russian Academy of Sciences.

Single-crystal and powder X-ray diffraction analyses, IR spectroscopy, CHN elemental analysis, and magnetic measurements were performed using the equipment of the JRC PMR IGIC RAS.

I. V. K. and M. V. F. thank Ministry of Science and Higher Education of the Russian Federation for access to EPR equipment. We thank Dr Maxim Yulikov (ETH-Zurich) for fruitful discussions on PRE.

References

  1. A. Dey, J. Acharya and V. Chandrasekhar, Heterometallic 3d–4f Complexes as Single-Molecule Magnets, Chem. – Asian J., 2019, 14, 4433 CrossRef CAS PubMed.
  2. S. K. Langley, D. P. Wielechowski, V. Vieru, N. F. Chilton, B. Moubaraki, B. F. Abrahams, L. F. Chibotaru and K. S. Murray, A {CrIII2DyIII2} Single-Molecule Magnet: Enhancing the Blocking Temperature through 3d Magnetic Exchange, Angew. Chem., Int. Ed., 2013, 52, 12014 CrossRef CAS PubMed.
  3. E. Moreno Pineda, N. F. Chilton, F. Tuna, R. E. P. Winpenny and E. J. L. McInnes, Systematic Study of a Family of Butterfly-Like {M2Ln2} Molecular Magnets (M = MgII, MnIII, CoII, NiII, and CuII; Ln = YIII, GdIII, TbIII, DyIII, HoIII, and ErIII), Inorg. Chem., 2015, 54, 5930 CrossRef CAS PubMed.
  4. A. Dey, P. Bag, P. Kalita and V. Chandrasekhar, Heterometallic CuII–LnIII complexes: Single molecule magnets and magnetic refrigerants, Coord. Chem. Rev., 2021, 432, 213707 CrossRef CAS.
  5. (a) N. Subbulakshmi, M. S. Kumar, K. J. Sheela, S. R. Krishnan, V. M. Shanmugam and P. Subramanian, EPR and optical absorption studies of paramagnetic molecular ion (VO2+) in Lithium Sodium Acid Phthalate single crystal, Physica B, 2017, 526, 110 CrossRef CAS; (b) Q. H. Le, C. Friebe, W. C. Wang and L. Wondraczek, Spectroscopic properties of Cu2+ in alkaline earth metaphosphate, fluoride-phosphate and fluoride-phosphate-sulfate glasses, J. Non-Cryst. Solids, 2019, 4, 100037 CAS.
  6. (a) E. S. Bazhina, G. G. Aleksandrov, M. A. Kiskin, A. A. Korlyukov, N. N. Efimov, A. S. Bogomyakov, A. A. Starikova, V. S. Mironov, E. A. Ugolkova, V. V. Minin, A. A. Sidorov and I. L. Eremenko, The First Series of Heterometallic LnIII-VIV Complexes Based on Substituted Malonic Acid Anions: Synthesis, Structure and Magnetic Properties, Eur. J. Inorg. Chem., 2018, 2018, 5075 CrossRef CAS; (b) K. Kotrle, I. Nemec, J. Moncol, E. Čižmár and R. Herchel, 3d–4f magnetic exchange interactions and anisotropy in a series of heterobimetallic vanadium(IV)–lanthanide(III) Schiff base complexes, Dalton Trans., 2021, 50, 13883 RSC; (c) E. S. Bazhina, M. A. Shmelev, K. A. Babeshkin, N. N. Efimov, M. A. Kiskin and I. L. Eremenko, Two families of Ln(III)-V(IV) compounds (Ln(III) = Eu, Tb, Dy, Ho) of different structural types mediated by Rb+ and Cs+ cations: Slow magnetic relaxation of Eu(III)- and Ho(III)-containing members, Polyhedron, 2023, 236, 116385 CrossRef CAS.
  7. Y.-S. Ding, Y.-F. Deng and Y.-Z. Zheng, The Rise of Single-Ion Magnets as Spin Qubits, Magnetochemistry, 2016, 2, 40 CrossRef.
  8. (a) S. P. Petrosyants, K. A. Babeshkin, A. V. Gavrikov, A. B. Ilyukhin, E. V. Belova and N. N. Efimov, Towards comparative investigation of Er- and Yb-based SMMs: the effect of the coordination environment configuration on the magnetic relaxation in the series of heteroleptic thiocyanate complexes, Dalton Trans., 2019, 48, 12644 RSC; (b) K. A. Babeshkin, A. V. Gavrikov, S. P. Petrosyants, A. B. Ilyukhin, E. V. Belova and N. N. Efimov, Unexpected Supremacy of Non-Dysprosium Single-Ion Magnets within a Series of Isomorphic Lanthanide Cyanocobaltate(III) Complexes, Eur. J. Inorg. Chem., 2020, 2020, 4380 CrossRef CAS; (c) J.-Y. Ge, Z. Chen, H.-Y. Wang, H. Wang, P. Wang, X. Duan and D. Huo, Thiacalix[4]arene-supported mononuclear lanthanide compounds: slow magnetic relaxation in dysprosium and erbium analogues, New J. Chem., 2018, 42, 17968 RSC; (d) P. Antal, B. Drahoš, R. Herchel and Z. Trávníček, Muffin-like lanthanide complexes with an N5O2-donor macrocyclic ligand showing field-induced single-molecule magnet behaviour, Dalton Trans., 2016, 45, 15114 RSC.
  9. S. K. Langley, D. P. Wielechowski, N. F. Chilton, B. Moubaraki and S. Keith, Murray, A Family of {CrIII2LnIII2} Butterfly Complexes: Effect of the Lanthanide Ion on the Single-Molecule Magnet Properties, Inorg. Chem., 2015, 54, 10497 CrossRef CAS PubMed.
  10. (a) S.-D. Jiang, S.-S. Liu, L.-N. Zhou, B.-W. Wang, Z.-M. Wang and S. Gao, Series of Lanthanide Organometallic Single-Ion Magnets, Inorg. Chem., 2012, 51, 3079 CrossRef CAS PubMed; (b) Y.-S. Meng, Y.-S. Qiao, Y.-Q. Zhang, S.-D. Jiang, Z.-S. Meng, B.-W. Wang, Z.-M. Wang and S. Gao, Can Non-Kramers TmIII Mononuclear Molecules be Single-Molecule Magnets (SMMs)?, Chem. – Eur. J., 2016, 22, 4704 CrossRef CAS PubMed; (c) K. L. M. Harriman, I. Korobkov and M. Murugesu, From a Piano Stool to a Sandwich: A Stepwise Route for Improving the Slow Magnetic Relaxation Properties of Thulium, Organometallics, 2017, 36, 4515 CrossRef CAS; (d) O. Yu. Mariichak, S. Kaabel, Y. A. Karpichev, G. M. Rozantsev, G. M. Rozantsev, S. V. Radio, C. Pichon, H. Bolvin and J.-P. Sutter, Crystal Structure and Magnetic Properties of Peacock–Weakley Type Polyoxometalates Na9[Ln(W5O18)] (Ln = Tm, Yb): Rare Example of Tm(III) SMM, Magnetochemistry, 2020, 6, 53 CrossRef CAS.
  11. D. Banerjee and J. B. Parise, Recent Advances in s-Block Metal Carboxylate Networks, Cryst. Growth Des., 2011, 11, 4704 CrossRef CAS.
  12. (a) J. C. Goodwin, D. J. Price, A. K. Powell and S. L. Heath, Alkali Metal Templated Assembly of an Iron Trigonal Prism, Eur. J. Inorg. Chem., 2000, 2000, 1407 CrossRef; (b) W. L. Leong and J. J. Vittal, Alkali metal ion directed self-assembled Ni(II) molecular clusters, New J. Chem., 2010, 34, 2145 RSC; (c) S. Gupta, M. V. Kirillova, M. F. C. Guedes da Silva, A. J. L. Pombeiro and A. M. Kirillov, Alkali Metal Directed Assembly of Heterometallic Vv/M (M = Na, K, Cs) Coordination Polymers: Structures, Topological Analysis, and Oxidation Catalytic Properties, Inorg. Chem., 2013, 52, 8601 CrossRef CAS PubMed; (d) G. González-Riopedre, M. R. Bermejo, M. I. Fernández-García, A. M. González-Noya, R. Pedrido, M. J. Rodríguez-Doutón and M. Maneiro, Alkali-Metal-Ion-Directed Self-Assembly of Redox-Active Manganese(III) Supramolecular Boxes, Inorg. Chem., 2015, 54, 2512 CrossRef PubMed; (e) D. O. Blinou, E. N. Zorina-Tikhonova, J. K. Voronina, M. A. Shmelev, A. K. Matiukhina, P. N. Vasilyev, N. N. Efimov, E. V. Alexandrov, M. A. Kiskin and I. L. Eremenko, Impacts of Alkali Metals on the Structures and Properties of Fe(III) Heterometallic Cyclobutane-1,1-dicarboxylate Complexes, Cryst. Growth Des., 2023, 23, 5571 CrossRef CAS.
  13. Y.-L. Li, H.-L. Wang, Z.-H. Zhu, X.-L. Lu, F.-P. Liang and H.-H. Zou, Alkali metal-linked triangular building blocks assemble a high-nucleation lanthanoid cluster based on single-molecule magnets, iScience, 2022, 25, 105285 CrossRef CAS PubMed.
  14. (a) Y. Maruno, K. Yabe, H. Hagiwara, T. Fujinami, N. Matsumoto, N. Re and J. Mrozinski, yclic and Linear Structures Constructed by Ionic Bonds between Alkali Ion and Pinwheel Pentanuclear [GdIII(CuIIL)4] Core of M[GdIII(CuIIL)4] (M+ = Na+, K+, and Cs+; H3L = N-(4-Methyl-6-oxo-3-azahept-4-enyl)oxamic Acid), Bull. Chem. Soc. Jpn., 2010, 83, 1511 CrossRef CAS; (b) M. R. Azar, T. T. Boron, J. C. Lutter, C. I. Daly, K. A. Zegalia, R. Nimthong, G. M. Ferrence, M. Zeller, J. W. Kampf, V. L. Pecoraro and C. M. Zaleski, Controllable Formation of Heterotrimetallic Coordination Compounds: Systematically Incorporating Lanthanide and Alkali Metal Ions into the Manganese 12-Metallacrown-4 Framework, Inorg. Chem., 2014, 53, 1729 CrossRef CAS PubMed; (c) A. A. Bovkunova, E. S. Bazhina, I. S. Evstifeev, Yu. V. Nelyubina, M. A. Shmelev, K. A. Babeshkin, N. N. Efimov, M. A. Kiskin and I. L. Eremenko, Two types of Ln2Cu2 hydroxo-trimethylacetate complexes with 0D and 1D motifs: synthetic features, structural differences, and slow magnetic relaxation, Dalton Trans., 2021, 50, 12275 RSC.
  15. E. S. Bazhina, A. A. Korlyukov, J. K. Voronina, K. A. Babeshkin, E. A. Ugolkova, N. N. Efimov, M. V. Fedin, M. A. Kiskin and I. L. Eremenko, Effect of the Alkaline Metal Ion on the Crystal Structure and Magnetic Properties of Heterometallic GdIII-VIV Complexes Based on Cyclobutane-1,1-Dicarboxylate Anions, Magnetochemistry, 2021, 7, 82 CrossRef CAS.
  16. I. V. Kurganskii, E. S. Bazhina, A. A. Korlyukov, K. A. Babeshkin, N. N. Efimov, M. A. Kiskin, S. L. Veber, A. A. Sidorov, I. L. Eremenko and M. V. Fedin, Mapping Magnetic Properties and Relaxation in Vanadium(IV) Complexes with Lanthanides by Electron Paramagnetic Resonance, Molecules, 2019, 24, 4582 CrossRef CAS PubMed.
  17. M. Llunell, D. Casanova, J. Cirera, P. Alemany and S. Alvarez, SHAPE v.2.1, Program for the Stereochemical Analysis of Molecular Fragments by Means of Continuous Shape Measures and Associated Tools, Barselona, Spain, 2013 Search PubMed.
  18. (a) H.-R. Wen, X.-R. Xie, S.-J. Liu, J. Bao, F.-F. Wang, C.-M. Liu and J.-S. Liao, Homochiral luminescent lanthanide dinuclear complexes derived from a chiral carboxylate, RSC Adv., 2015, 5, 98097 RSC; (b) M. Gebrezgiabher, Y. Bayeh, T. Gebretsadik, G. Gebreslassie, F. Elemo, M. Thomas and W. Linert, Lanthanide-Based Single-Molecule Magnets Derived from Schiff Base Ligands of Salicylaldehyde Derivatives, Inorganics, 2020, 8, 66 CrossRef CAS.
  19. P. Singh, S. Schlittenhardt, D. Thakre, S. K. Kushvaha, S. Kumar, H. S. Karnamkkott, M. Ruben, M. Ibrahim, A. Banerjee and K. C. Mondal, Exploration of Vanadium(IV)-Based Single-Ion Magnet Properties in Diphosphonate-Supported Mixed-Valent Polyoxovanadates, Cryst. Growth Des., 2022, 22, 5666 CrossRef CAS.
  20. (a) A. V. Gavrikov, P. S. Koroteev, N. N. Efimov, Z. V. Dobrokhotova, A. B. Ilyukhin, A. K. Kostopoulos, A.-M. Ariciu and V. M. Novotortsev, Novel mononuclear and 1D-polymeric derivatives of lanthanides and (η6-benzoic acid)tricarbonylchromium: synthesis, structure and magnetism, Dalton Trans., 2017, 46, 3369 RSC; (b) A. V. Gavrikov, N. N. Efimov, Z. V. Dobrokhotova, A. B. Ilyukhin, P. N. Vasilyev and V. M. Novotortsev, Novel mononuclear Ln complexes with pyrazine-2-carboxylate and acetylacetonate co-ligands: remarkable single molecule magnet behavior of a Yb derivative, Dalton Trans., 2017, 46, 11806 RSC; (c) S. P. Petrosyants, K. A. Babeshkin, A. B. Ilyukhin and N. N. Efimov, Molecular Magnets Based on Mononuclear Aqua and Aqua-Chloro Lanthanide (Tb, Dy, Er, Yb) Complexes with Bipyridine, Russ. J. Coord. Chem., 2021, 47, 165 CrossRef CAS.
  21. (a) M. Briganti, F. Santanni, L. Tesi, F. Totti, R. Sessoli and A. Lunghi, J. Am. Chem. Soc., 2021, 143, 13633 CrossRef CAS PubMed; (b) L. Gu and R. Wu, Origin of the anomalously low Raman exponents in single molecule magnets, Phys. Rev. B, 2021, 103, 014401 CrossRef CAS.
  22. MagSuite v.3.2, M. Rouzières, Zenodo, 2023;  DOI:10.5281/zenodo.4030310.
  23. (a) Y. Li, Q. Shang, Y.-Q. Zhang, E.-C. Yang and X.-J. Zhao, Fine Tuning of the Anisotropy Barrier by Ligand Substitution Observed in Linear {Dy2Ni2} Clusters, Chem. – Eur. J., 2016, 22, 18840 CrossRef CAS PubMed; (b) F. Pointillart, K. Bernot, R. Sessoli and D. Gatteschi, Effects of 3d–4f Magnetic Exchange Interactions on the Dynamics of the Magnetization of DyIII-MII-DyIII Trinuclear Clusters, Chem. – Eur. J., 2007, 13, 1602 CrossRef PubMed; (c) I. A. Kühne, C. E. Anson and A. K. Powell, The Influence of Halide Substituents on the Structural and Magnetic Properties of Fe6Dy3 Rings, Front. Chem., 2020, 8, 701 CrossRef PubMed; (d) X.-Q. Zhao, J. Wang, D.-X. Bao, S. Xiang, Y.-J. Liu and Y.-C. Li, The ferromagnetic [Ln2Co6] heterometallic complexes, Dalton Trans., 2017, 46, 2196 RSC; (e) S. M. T. Abtab, M. C. Majee, M. Maity, J. Titiš, R. Boča and M. Chaudhury, Tetranuclear Hetero-Metal [CoII2LnIII2] (Ln = Gd, Tb, Dy, Ho, La) Complexes Involving Carboxylato Bridges in a Rare μ4–η22 Mode: Synthesis, Crystal Structures, and Magnetic Properties, Inorg. Chem., 2014, 53, 1295 CrossRef CAS PubMed; (f) J. Goura, R. Guillaume, E. Rivière and V. Chandrasekhar, Hexanuclear, Heterometallic, Ni3Ln3 Complexes Possessing O-Capped Homo- and Heterometallic Structural Subunits: SMM Behavior of the Dysprosium Analogue, Inorg. Chem., 2014, 53, 7815 CrossRef CAS PubMed; (g) L. Jiang, Y. Liu, X. Liu, J. Tian and S. Yan, Three series of heterometallic NiII–LnIII Schiff base complexes: synthesis, crystal structures and magnetic characterization, Dalton Trans., 2017, 46, 12558 RSC; (h) I. A. Kühne, V. Mereacre, C. E. Anson and A. K. Powell, Nine members of a family of nine-membered cyclic coordination clusters; Fe6Ln3 wheels (Ln = Gd to Lu and Y), Chem. Commun., 2016, 52, 1021 RSC; (i) M. Biswas, E. C. Sañudo and D. Ray, Carboxylate-Decorated Aggregation of Octanuclear Co4Ln4 (Ln = Dy, Ho, Yb) Complexes from Ligand-Controlled Hydrolysis: Synthesis, Structures, and Magnetic Properties, Inorg. Chem., 2021, 60, 11129 CrossRef CAS PubMed; (j) A. Bhanja, M. Schulze, R. Herchel, E. Moreno-Pineda, W. Wernsdorfer and D. Ray, Selective Coordination of Self-Assembled Hexanuclear [Ni4Ln2] and [Ni2Mn2Ln2] (Ln = DyIII, TbIII, and HoIII) Complexes: Stepwise Synthesis, Structures, and Magnetic Properties, Inorg. Chem., 2020, 59, 17929 CrossRef CAS PubMed; (k) Y. Gao, L. Zhao, X. Xu, G.-F. Xu, Y.-N. Guo, J. Tang and Z. Liu, Heterometallic Cubanes: Syntheses, Structures, and Magnetic Properties of Lanthanide(III)−Nickel(II) Architectures, Inorg. Chem., 2011, 50, 1304 CrossRef CAS PubMed; (l) S. Yu, H.-L. Wang, Z. Chen, H.-H. Zou, H. Hu, Z.-H. Zhu, D. Liu, Y. Liang and F.-P. Liang, Two Decanuclear DyIIIxCoII10−x (x = 2, 4) Nanoclusters: Structure, Assembly Mechanism, and Magnetic Properties, Inorg. Chem., 2021, 60, 4904 CrossRef CAS PubMed; (m) Y. Li, C. Zhang, J.-W. Yu, E.-C. Yang and X.-J. Zhao, Transition metal ion-directed magnetic behaviors observed in an isostructural heterobinuclear system, Inorg. Chim. Acta, 2016, 445, 110 CrossRef CAS; (n) P.-H. Lin, E. Y. Tsui, F. Habib, M. Murugesu and T. Agapie, Effect of the Mn Oxidation State on Single-Molecule-Magnet Properties: MnIII vs MnIV in Biologically Inspired DyMn3O4 Cubanes, Inorg. Chem., 2016, 55, 6095 CrossRef CAS PubMed; (o) S. K. Langley, D. P. Wielechowski, B. Moubaraki and K. S. Murray, Enhancing the magnetic blocking temperature and magnetic coercivity of {CrIII2LnIII2} single-molecule magnets via bridging ligand modification, Chem. Commun., 2016, 52, 10976 RSC; (p) K. Wang, Z.-L. Chen, H.-H. Zou, Z. Zhang, W.-Y. Sun and F.-P. Liang, Two Types of Cu-Ln Heterometallic Coordination Polymers with 2-Hydroxyisophthalate: Syntheses, Structures, and Magnetic Properties, Cryst. Growth Des., 2015, 15, 2883 CrossRef CAS; (q) C. Papatriantafyllopoulou, W. Wernsdorfer, K. A. Abboud and G. Christou, Mn21Dy Cluster with a Record Magnetization Reversal Barrier for a Mixed 3d/4f Single-Molecule Magnet, Inorg. Chem., 2011, 50, 421 CrossRef CAS PubMed; (r) P. S. Koroteev, Z. V. Dobrokhotova, A. B. Ilyukhin, E. V. Belova, A. D. Yapryntsev, M. Rouzières, R. Clérac and N. N. Efimov, Tetranuclear Cr–Ln ferrocenecarboxylate complexes with a defect-dicubane structure: synthesis, magnetism, and thermolysis, Dalton Trans., 2021, 50, 16990 RSC; (s) T. Zhang, L.-L. Zhang, C.-X. Ji, S. Ma, Y.-X. Sun, J.-P. Zhao and F.-C. Liu, Construction of Designated Heptanuclear Metal 8-hydroxyquinolates with Different Ions and Auxiliary Coligands, Cryst. Growth Des., 2019, 19, 3372 CrossRef CAS.
  24. (a) F. Wang, H.-W. Gong, Y. Zhang, A.-Q. Xue, W.-H. Zhu, Y.-Q. Zhang, Z.-N. Huang, H.-L. Sun, B. Liu, Y.-Y. Fang and S. Gao, The comparative studies on the magnetic relaxation behaviour of the axially-elongated pentagonal-bipyramidal dysprosium and erbium ions in similar one-dimensional chain structures, Dalton Trans., 2021, 50, 8736 RSC; (b) G.-J. Zhou, T. Han, Y.-S. Ding, N. F. Chilton and Y.-Z. Zheng, Metallacrowns as Templates for Diabolo-like {LnCu8} Complexes with Nearly Perfect Square Antiprismatic Geometry, Chem. – Eur. J., 2017, 23, 15617 CrossRef CAS PubMed; (c) Y. Peng, V. Mereacre, C. E. Anson and A. K. Powell, Butterfly MIII2Er2 (MIII = Fe and Al) SMMs: Synthesis, Characterization, and Magnetic Properties, ACS Omega, 2018, 3, 6360 CrossRef CAS PubMed; (d) K. R. Vignesh, S. K. Langley, A. Swain, B. Moubaraki, M. Damjanović, W. Wernsdorfer, G. Rajaraman and K. S. Murray, Slow Magnetic Relaxation and Single-Molecule Toroidal Behaviour in a Family of Heptanuclear {CrIIILnIII6} (Ln=Tb, Ho, Er) Complexes, Angew. Chem., Int. Ed., 2018, 57, 779 CrossRef CAS PubMed; (e) P. Shukla, S. Roy, D. Dolui, W. Cañón-Mancisidor and S. Das, Syntheses, Structures, and Magnetic Properties of Pentanuclear Spirocyclic Ni4Ln Derivative: Field Induced Slow Magnetic Relaxation by Dysprosium and Erbium Analogue, Eur. J. Inorg. Chem., 2020, 2020, 823 CrossRef CAS; (f) I. Oyarzabal, E. Echenique-Errandonea, E. San Sebastián, A. Rodríguez-Diéguez, J. M. Seco and E. Colacio, Synthesis, Structural Features and Physical Properties of a Family of Triply Bridged Dinuclear 3d-4f Complexes, Magnetochemistry, 2021, 7, 22 CrossRef CAS; (g) M. J. Heras Ojea, V. A. Milway, G. Velmurugan, L. H. Thomas, S. J. Coles, C. Wilson, W. Wernsdorfer, G. Rajaraman and M. Murrie, Enhancement of TbIII–CuII Single-Molecule Magnet Performance through Structural Modification, Chem. – Eur. J., 2016, 22, 12839 CrossRef CAS PubMed; (h) M.-G. Alexandru, D. Visinescu, B. Cula, S. Shova, R. Rabelo, N. Moliner, F. Lloret, J. Cano and M. Julve, A rare isostructural series of 3d–4f cyanidobridged heterometallic squares obtained by assembling [FeIII{HB(pz)3}(CN)3] and LnIII ions: synthesis, X-ray structure and cryomagnetic study, Dalton Trans., 2021, 50, 14640 RSC; (i) B. Dutta, T. Guizouarn, F. Pointillart, K. Kotrle, R. Herchel and D. Ray, Lanthanoid coordination prompts unusually distorted pseudo-octahedral NiII coordination in heterodinuclear Ni–Ln complexes: synthesis, structure and understanding of magnetic behaviour through experiment and computation, Dalton Trans., 2023, 52, 10402 RSC.
  25. (a) M. Liberka, K. Boidachenko, J. J. Zakrzewski, M. Zychowicz, J. Wang, S. Ohkoshi and S. Chorazy, Near-Infrared Emissive Cyanido-Bridged {YbFe2} Molecular Nanomagnets Sensitive to the Nitrile Solvents of Crystallization, Magnetochemisty, 2021, 7, 79 CrossRef CAS; (b) W.-W. Chang, H. Yang, H.-Q. Tian, D.-C. Li and J.-M. Dou, 3d–4f Metallacrown complexes with a new sandwich core: synthesis, structures and single molecule magnet behavior, New J. Chem., 2020, 44, 14145 RSC.
  26. (a) M. Atzori, E. Morra, L. Tesi, A. Albino, M. Chiesa, L. Sorace and R. Sessoli, Quantum Coherence Times Enhancement in Vanadium(IV)-based Potential Molecular Qubits: the Key Role of the Vanadyl Moiety, J. Am. Chem. Soc., 2016, 138, 11234 CrossRef CAS PubMed; (b) V. Lagostina, E. Romeo, A. M. Ferrari, V. Maurino and M. Chiesa, Monomeric (VO3+) and dimeric mixed valence (V2O33+) vanadium species at the surface of shape controlled TiO2 anatase nano crystals, J. Catal., 2022, 406, 28 CrossRef CAS.
  27. C. Bonardi, C. J. Magon, E. A. Vidoto, M. C. Terrile, L. E. Bausá, E. Montoya, D. Bravo, A. Martín and F. J. López, EPR spectroscopy of Yb3+ in LiNbO3 and Mg:LiNbO3, J. Alloys Compd., 2001, 323–324, 340 CrossRef CAS.
  28. (a) K. M. Salikhov, S. A. Dzuba and A. M. Raitsimring, The theory of electron spin-echo signal decay resulting from dipole-dipole interactions between paramagnetic centers in solids, J. Magn. Reson., 1981, 42, 255 CAS; (b) W. Hilczer, J. Goslar, M. Gramza, S. K. Hoffmann, W. Blicharski, A. Osyczka, B. Turyna and W. Froncisz, A resonance enhancement of the phase relaxation in the electron spin echo of nitroxide covalently attached to cytochrome c, Chem. Phys. Lett., 1995, 247, 601 CrossRef CAS; (c) P. R. Vennam, N. Fisher, M. D. Krzyaniak, D. M. Kramer and M. K. Bowman, A Caged, Destabilized, Free Radical Intermediate in the Q-Cycle, ChemBioChem, 2013, 14, 1745 CrossRef CAS PubMed.
  29. P. Lueders, S. Razzaghi, H. Jäger, R. Tschaggelar, M. A. Hemminga, M. Yulikov and G. Jeschke, Distance determination from dysprosium induced relaxation enhancement: a case study on membrane-inserted WALP23 polypeptides, Mol. Phys., 2013, 111, 2824–2833 CrossRef CAS.
  30. N. N. Efimov, K. A. Babeshkin and A. V. Rotov, Method of dynamic magnetic susceptibility in the study of coordination compounds, Russ. J. Coord. Chem., 2024, 50, 363 CrossRef CAS.
  31. S. Stoll and A. Schweiger, EasySpin, a comprehensive software package for spectral simulation and analysis in EPR, J. Magn. Reson., 2006, 178, 42 CrossRef CAS PubMed.
  32. SMART (control) and SAINT (integration) Software. Version 5.0, Bruker AXS, Inc., Madison (WI, USA), 1997 Search PubMed.
  33. G. M. Sheldrick, SADABS, Program for Scaling and Correction of Area Detector Data, Göttingen University, Göttingen, Germany, 1997 Search PubMed.
  34. G. M. Sheldrick, Crystal structure refinement with SHELXL, Acta Crystallogr., Sect. A: Found. Crystallogr., 2015, 71, 3 CrossRef PubMed.
  35. O. V. Dolomanov, L. J. Bourhis, R. J. Gildea, J. A. K. Howard and H. Puschmann, OLEX2: a complete structure solution, refinement and analysis program, J. Appl. Crystallogr., 2009, 42, 339 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available: Orbital diagrams for CuII and VIV ions; PXRD patterns for 1Ln series; continuous shape measures (CShM) for LnO8 coordination polyhedra in 1Dy, 1Er; tables of selected bond angles in structures 1Dy, 1Er; tables of hydrogen bond parameters in 1Dy, 1Er; the magnetization M(T) and M(H/T) dependences for 1Ln, frequency dependences of the in-phase and out-of-phase parts of dynamic magnetic susceptibility for 1Ln at T = 2 K under various dc-magnetic fields; frequency dependences of the in-phase and out-of-phase parts of dynamic magnetic susceptibility for 1Dy, 1Er, 1Tm (under 1000 Oe dc-field), and 1Yb (under 2500 Oe dc-field) at different temperatures; Cole–Cole plots for 1Dy measured at 2–7.5 K; the τ vs. H plots for 1Dy, 1Er, 1Yb at 2 K; the τ vs. 1/T plots and the best-fit parameters of magnetization relaxation for 1Tm; the τ vs. 1/T plots for 1Dy, 1Er, 1Yb processed using the MagSuite v.3.2 software and the best-fit parameters of magnetization relaxation calculated for 1Dy, 1Er, 1Yb; CW EPR spectra of MY and 1Y at 293 K and 10 K; IR spectra for 1Dy in aqueous solution and crystalline state. CCDC 2266768 and 2266772. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d4dt01779j

This journal is © The Royal Society of Chemistry 2024
Click here to see how this site uses Cookies. View our privacy policy here.