Abstract
A series of DyIII–AlIII metallacrowns (MC) with the ligand salicylhydroxamic acid (H3shi) were investigated for magneto-structural interactions. The four MCs demonstrate that slight variations in reaction conditions and molecular components can lead to different classes of MCs. The DyIII–AlIII MCs include an archetype Dy III Al III 4 [12-MC-4], the dimer Dy III 2 Al III 8 [12-MC-4] 2, a [3.3.1] metallacryptand (MCr) Dy III Al III 6 [3.3.1], and Dy III 2 Al III 6 [18-MC-6], representing a new type of MC. Dy III 2 Al III 6 [18-MC-6] consists of a MC ring with six ring AlIII ions that have an octahedral propeller configuration with a stereoisomer pattern ΔΛΛΛΔΔ about the MC ring. The 18-MC-6 captures a [DyIII 2(μ3-OH)2]4+ core, where each DyIII ion is nine-coordinate with muffin geometry (C s). Both the static and dynamic magnetic properties of the complexes were investigated. Models of the static magnetic data reveal that Dy III Al III 4 [12-MC-4] and Dy III 2 Al III 8 [12-MC-4] 2 do not have a significant axial ligand field, as the DyIII ions have a square antiprism geometry (D 4d). However, a similar analysis demonstrates that Dy III Al III 6 [3.3.1] and Dy III 2 Al III 6 [18-MC-6] do have a significant axial component to the ligand field, as the DyIII ions in these complexes have spherical capped square antiprism (C 4v) or muffin geometry, respectively. The DyIII geometry differences lead to differing dynamic magnetic susceptibility behavior. Dy III Al III 4 [12-MC-4] and Dy III 2 Al III 8 [12-MC-4] 2 do not display a frequency-dependent out-of-phase magnetic susceptibility signal in the absence of a static magnetic field. A frequency-dependent signal is observed only with the application of an 800 Oe magnetic field. However, Dy III Al III 6 [3.3.1] and Dy III 2 Al III 6 [18-MC-6] do exhibit SMM behavior in the absence and presence of a static magnetic field. For Dy III Al III 6 [3.3.1] and Dy III 2 Al III 6 [18-MC-6], the effective energy barrier to magnetization relaxation (U eff) with an 800 Oe magnetic field is 59 ± 2 and 35 ± 2 cm–1 with τ0 = 2.5 ± 0.7 × 10–8 and 3.5 ± 1.1 × 10–8 s, respectively.


1. Introduction
Single-molecule magnets (SMMs) continue to attract interest since the recognition of the first SMM, a Mn12O12(acetate)16(H2O)4 complex, by Gatteschi, Christou, Hendrickson, and co-workers. − Initially, research focused on 3d-transition metal only molecules; however, the field of SMMs soon expanded to mixed 3d-4f and 4f-only structures. − While most SMMs have multiple metal centers, some only contain one paramagnetic center and are subsequently named single-ion magnets (SIMs), though the principles behind SMMs and SIMs are the same. , SMMs have the potential to store information at cryogenic temperatures, have applications in spintronic devices, and act as qubits for quantum computing and quantum simulation. − The properties of SMMs rely on the slow reversal of magnetization due to an effective energy barrier (U eff) to this reversal. The slow relaxation dynamics of the magnetization in these SMMs is the result of an interplay between different temperature-dependent mechanisms that arise from the molecular structure and the vibrational spectrum. −
One of the key aspects that defines SMMs is the height (i.e., size) of the energy barrier, which is directly related to the overall molecular spin state of the molecule (ST) and the molecular magnetoanisotropy (D). In efforts to increase the operational temperatures of SMMs, scientists have focused on increasing the ST and D components through a bottom-up approach to molecular design.
Metallacrowns (MC), the metallamacrocycle analogues of organic macrocyclic crown ethers, offer a path to SMMs. MCs are a class of supramolecular species based on a ring structure with a metal–nitrogen-oxygen repeat unit. This cyclic structure generates a central cavity, and like their organic crown ether counterparts, MCs can bind various metal ions in the central cavity. Through careful choice of ring and central metal ions and surrounding ligands that form the framework of the macrocycle, MCs can be built in various sizes with 9-MC-3, 12-MC-4, and 15-MC-5 being the most common, where the first number represents the total number of atoms in the MC ring and the second number represents the number of oxygen atoms in the MC ring. Numerous MCs have been recognized as SMMs, − including one of the first 3d-4f SMMs, a DyIII 6MnIII 4MnIV 2 28-MC-10. Since that first report, lanthanide ions have been combined with an assortment of 3d transition metal ions to design MC-SMMs. The combination of LnIII ions with manganese, , iron, , nickel, , and copper , has been a common choice and has led to SMMs from a variety of different classes of MCs, with 12-MC-4 and 15-MC-5 being the most common.
One MC doctrine is the modular approach to the design of the molecules. Individual components can be substituted while retaining the overall MC structure. This can lead to the fine-tuning of the MC molecule to a particular function, such as white-light emission, , near-infrared luminescence, , or magnetism. For example, the first 12-MC-4 complex identified as a SMM was a MnIIMnIII 4 system with a central MnII ion surrounded by four ring MnIII ions. The overall shape of the molecule was a square due to the choice of the MC framework salicylhydroxamic acid (H3shi), which places the ring MnIII ions 90° relative to each other. Subsequently, we showed that two of the ring MnIII ions can be replaced with two CuII ions to generate a MnIIMnIII 2CuII 2 12-MC-4 that retains an overall square shape, and the MC maintained its SMM behavior. Moreover, we demonstrated that the central MnII ion of a 12-MC-4 with four ring MnIII ions can be replaced with a LnIII ion, while the overall square shape for the 12-MC-4 was maintained even though the much larger LnIII ion was captured in the center of the MC cavity. We then determined that some of the DyIIIMnIII 4 12-MC-4 complexes displayed SMM behavior, which depended on the identity of ancillary carboxylate anions. Furthermore, if the ring MnIII ions were replaced with GaIII or AlIII ions, the resulting LnIIIGaIII 4 or LnIIIAl4 12-MC-4 complexes were highly luminescent species. − Thus, the series of 12-MC-4 complexes was systematically altered to fulfill a particular application while retaining the MC framework. In addition, we showed that planar DyIIICuII 5 and HoIIICuII 5 15-MC-5 complexes with a central LnIII ion in an eight-coordinate triangular dodecahedron (D 2d) or biaugmented trigonal prism (C 2v) geometry behave as weak SMMs. Recent work has focused on planar 15-MC-5 complexes based on LnIII ions confined in a pentagonal bipyramidal geometry (D 5h) with a strong axial component of the ligand field. By judicious choice of the axial ligands for the LnIII ions, DyIIICuII 5, HoIIINiII 5, and DyIIINiII 5 15-MC-5 complexes have produced the highest U eff values for MCs thus far, with values ranging from 423 to 625 cm–1. ,, These examples highlight a hallmark of MC chemistry – the modular design of molecules through careful choice of reaction conditions and of components to tailor molecules to particular properties.
Herein, we report a series of DyIII–AlIII MCs with the MC framework ligand salicylhydroxamic acid that display SMM behavior: an archetype Dy III Al III 4 [12-MC-4], a Dy III 2 Al III 8 [12-MC-4] 2 dimer, a Dy III Al III 6 [3.3.1] MCr (metallacryptand), and a Dy III 2 Al III 6 [18-MC-6]. The four molecules represent how slight variations in stoichiometric ratios between reaction components and astute ligand choices can lead to different classes of MCs. The different MC structures and coordination geometries of the central DyIII ions lead to different magnetic properties; thus, providing an opportunity to investigate magneto-structural interactions. The new DyIII–AlIII systems are also compared to similar DyIII–GaIII MCs previously investigated (see Section ), highlighting the magneto-structural differences between the two families of compounds.
2. Experimental Section
2.1. Synthetic Materials
Dysprosium(III) nitrate pentahydrate (99.99%) was purchased from Alfa Aesar. Salicylhydroxamic acid (>98%) was purchased from TCI America. Aluminum(III) nitrate nonahydrate (ACS grade) was purchased from Fisher Scientific. Methanol (ACS grade) and pyridine (ACS grade) were purchased from VWR Chemicals BDH. All reagents were used as received and without further purification.
2.2. Syntheses
DyIIINa(ben)4[12-MCAl(III)N(shi)-4](H2O)4, Dy III Al III 4 [12-MC-4], where ben– is benzoate. The synthesis and X-ray crystal structure of the compound have been previously reported.
{DyIIINa[12-MCAl(III)N(shi)-4]}2(iph)4(DMF)2(H2O)8, Dy III 2 Al III 8 [12-MC-4] 2 , where iph2– is isophthalate and DMF is N,N-dimethylformamide. The synthesis and X-ray crystal structure of the compound have been previously reported.
[Hpy]2[DyIIIAlIII 6(H2shi)2(shi)7(py)1.891(H2O)2], Dy III Al III 6 [3.3.1] MCr (Metallacryptand), where py is pyridine. The synthesis and X-ray crystal structure of the compound have been previously reported.
DyIII 2AlIII 6(H2shi)4(shi)6(OH)2(H2O)2.24(CH3OH)1.759(py)2· 6.935py·5.631CH3OH·1.2H2O, Dy III 2 Al III 6 [18-MC-6]. Dysprosium(III) nitrate pentahydrate (0.25 mmol), aluminum(III) nitrate nonahydrate (0.5 mmol), and H3shi (1 mmol) were dissolved in 30 mL of methanol, resulting in a clear, light pink solution. Then 13 mL of pyridine were added, resulting in a clear, yellow solution. The solution was stirred for 1 min and gravity filtered. No precipitate was recovered, and the filtrate remained a clear, yellow color. X-ray quality pink, block crystals were grown in 27 days by slow evaporation of the solvent. The percent yield was 2.1% based on dysprosium(III) nitrate pentahydrate. Elemental based on loss of interstitial methanol molecules, partial loss of interstitial pyridine molecules and an additional waters of hydration: Dy2Al6(H2shi)4(shi)6(OH)2(H2O)2.24(CH3OH)1.759(py)2·2py·8H2O, C91.759H97.516Al6Dy2N14O43.999 [FW = 2587.32 g/mol] found % (calculated): C = 42.41 (42.60), H = 4.07 (3.80), N = 7.81 (7.58). FT-IR bands (ATR, cm–1): 1602, 1579, 1538, 1521, 1479, 1452, 1405, 1318, 1272, 1252, 1218, 1151, 1101, 1066, 1035, 1007, 960, 934, 865, 825, 753, 697, 680 638, 605, 564.
2.3. Magnetic Measurements
Magnetization and susceptibility measurements were carried out on powder samples of Dy III Al III 4 [12-MC-4], Dy III 2 Al III 8 [12-MC-4] 2 , Dy III Al III 6 [3.3.1] MCr, and Dy III 2 Al III 6 [18-MC-6] using a Quantum Design MPMS-XL5 SQUID magnetometer. The powders were placed in gelatin capsules and manually compressed with the top part of each capsule inverted to minimize grain movement. Each capsule was positioned at the center of a plastic straw that was mounted onto the magnetometer sample rod. The low thermal mass of the sample holder ensures efficient heat exchange between the sample and the magnetometer sample space. Field- and temperature-dependent magnetization measurements were performed using the DC mode of the SQUID magnetometer, employing a 4 cm scan length and 3 scans per point. Magnetization as a function of magnetic field was measured from 0 to 5 T at a fixed temperature of 2 K. Temperature-dependent magnetization was recorded under a static field of 0.1 T, with a cooling sweep from 300 to 2 K at a rate of 2 K/min.
AC susceptibility measurements were conducted both in zero and under an applied static field of 0.08 T, using the AC option of the SQUID magnetometer with an AC drive field of 4 Oe. Frequency-dependent susceptibility data were collected in the 1 Hz to 1 kHz range at various temperatures between 10 and 2 K, with 0.5 K steps, using the “settle mode” and a temperature sweep rate of 0.5 K/min. The static field of 0.08 T was applied with the sample held at 2 K.
Precise temperature control and measurement in the 1.9–400 K range (with an accuracy of 0.01 K) is ensured by the magnetometer advanced Temperature Control System. This system employs two factory-calibrated negative-temperature coefficient Cernox thermometers to account for potential thermal gradients within the sample space. Below 14 K, only the field-shielded thermometer located beneath the sample tube is used in order to eliminate magnetic field interference with temperature readings.
2.4. X-ray Crystallography
Crystals used for single-crystal X-ray diffraction were taken from the mother liquor and were not dried. A mineral oil-coated crystal of Dy III 2 Al III 6 [18-MC-6] was mounted on a MicroMesh MiTeGen micromount and transferred to the diffractometer. The data were collected on a Bruker AXS D8 Quest diffractometer equipped with a solid-state CMOS area detector and a fine focus sealed tube X-ray source using Mo Kα radiation (λ = 0.71073 Å) monochromated with a Triumph curved graphite crystal. All data were collected at 150 K, and data collection and cell refinement were performed using APEX3 (version 2018.1-0) and SAINT+ embedded in APEX3, respectively. The data were scaled and corrected for absorption with SADABS as built into APEX3. Space groups were assigned using the SHELXTL suite of programs. The structures were solved using direct methods with SHELXS-97 and refined using least-squares refinements based on F2 with SHELXL-2018/3 and the graphical interface SHELXLE. , , Additional crystallographic data and experimental parameters are provided in Table S1 and the individual CIF of the compound. Important bond distances are provided in Table S2.
3. Results and Discussion
A. Synthesis and Characterization
The synthesis and single-crystal X-ray structures of Dy III Al III 4 [12-MC-4], , Dy III 2 Al III 8 [12-MC-4] 2 , , and Dy III Al III 6 [3.3.1] MCr (Figure ) have been previously reported in extensive detail. Thus, only brief descriptions will be provided. Dy III Al III 4 [12-MC-4] and Dy III 2 Al III 8 [12-MC-4] 2 are both based on a 12-MC-4 scaffold using the common MC ligand salicylhydroxamic acid (Figure a). For Dy III Al 4 [12-MC-4], a slightly domed 12-MC-4 framework consists of 4 ring AlIII ions and 4 salicylhydroximate (shi3–) ligands that bind a central DyIII ion on the convex side of the dome and a central Na+ ion on the concave side. The molecule possesses the typical 12-MC-4 framework with an [Al–N–O] repeat unit that recurs four times to generate a square-shaped metallamacrocycle with a central cavity lined with the oxime oxygen atoms of the shi3– ligands. The central DyIII ion binds to the four oxime oxygen atoms of the MC cavity, and four benzoate anions further tether the DyIII ion to the MC as the carboxylate groups of the benzoate anions form bridges between each ring AlIII ion and the central DyIII ion. A SHAPE analysis − revealed that the eight-coordinate DyIII ion has a square antiprism geometry (D 4d). If the benzoate bridges are replaced by the dicarboxylate anion isophthalate, two DyIIIAlIII 4 12-MC-4 units are joined to form the dimer found in Dy III 2 Al III 8 [12-MC-4] 2 (Figure b). Each central DyIII ion maintains a coordination number of eight with a square antiprism geometry. The DyIII–DyIII distance in the dimer is 7.12 Å. The AlIII ions of both Dy III Al III 4 [12-MC-4] and Dy III 2 Al III 8 [12-MC-4] 2 are six-coordinate with octahedral geometry (O h). While still utilizing H3shi, Dy III Al III 6 [3.3.1] MCr represents a three-dimensional MC also known as a metallacryptand, MCr. − A three-dimensional cavity is built around the nine-coordinate central DyIII ion (Figure c). The MCr shell is composed of seven shi3– ligands, which provide the metallacryptand linkages. The [3.3.1] nomenclature is analogous to that used for organic cryptands and refers to the number of oxygen atoms in each Al–N–O chain that surrounds the central DyIII ion. In the case of Dy III Al III 6 [3.3.1] MCr, there are three chains with two long O–N–Al–O–N–Al–O–N linkages and one short N–O linkage. Two singly deprotonated H2shi– ligands bind to the central DyIII ion but do not contribute to the MCr linkages. As determined by a SHAPE analysis, the nine-coordinate DyIII ion has a spherical capped square antiprism geometry (C 4v). , Each AlIII ion is six-coordinate with an octahedral geometry. Two of the AlIII ions have a trans-configuration of two different shi3– ligands. The other four AlIII ions have a cis-arrangement of three different shi3– ligands, which provide a propeller configuration. Thus, there are four stereocenters in the MCr with two Λ and two Δ AlIII stereoconfigurations.
1.

X-ray structures of (a) Dy III Al III 4 [12-MC-4], (b) Dy III 2 Al III 8 [12-MC-4] 2 , and (c) Dy III Al III 6 [3.3.1] MCr with only the metallacryptate core shown (all other atoms were omitted for clarity). The ellipsoid plots are at the 50% level. Hydrogen atoms, disordered atoms, and solvent molecules have been omitted for clarity. Color scheme: DyIIIaqua, AlIIIgreen, Na+yellow, oxygenred, nitrogenblue, and carbongray.
The synthesis and characterization of Dy III 2 Al III 6 [18-MC-6] have not been previously reported. The synthesis is similar to that of the other three compounds, but some key differences lead to the generation of an 18-MC-6. For both Dy III Al III 4 [12-MC-4] and Dy III 2 Al III 8 [12-MC-4] 2 , a strict ratio of 1:4:4 between Dy(NO3)3:Al(NO3)3:H3shi with different carboxylate anions leads to the formation of a 12-MC-4 framework that maintains the 1:4:4 ratio between the central LnIII ion, the ring AlIII ions, and the shi3– ligands. For Dy III Al III 6 [3.3.1] MCr, the Dy(NO3)3:Al(NO3)3:H3shi ratio is changed to 1:6:9 with the addition of the weak bases triethylamine and pyridine. This leads to an MCr that maintains a 1:6:9 ratio among the central DyIII ion, the ring AlIII ions, and the shi3–/H2shi– ligands (7 shi3– and 2 H2shi–). For Dy III 2 Al III 6 [18-MC-6], the Dy(NO3)3:Al(NO3)3:H3shi ratio is altered to 1:2:4 with the addition of pyridine as the only base. This leads to an MC with a 1:3:5 ratio among the central DyIII ions, the ring AlIII ions, and the shi3–/H2shi– ligands (6 shi3– and 4 H2shi–). In this case, the ratio in the resulting molecule is different than that of the starting materials. At this time, we do not have a conclusive explanation for this discrepancy.
The Dy III 2 Al III 6 [18-MC-6] (Figure ) consists of two DyIII and six AlIII ions (total 24+ charge) that are counterbalanced by four singly deprotonated salicylhydroximate anions (H2shi–), six triply deprotonated salicylhydroximate anions (shi3–), and two μ3-hydroxide anions (total 24- charge). Solvent water, pyridine, and methanol molecules complete the coordination of some of the metal ions. In addition, there are disordered interstitial water, methanol, and pyridine solvent molecules. The CIF contains a detailed explanation of the treatment of the disordered solvate molecules.
2.

(a) X-ray structure of Dy III 2 Al III 6 [18-MC-6]. Hydrogen atoms, disorder, and solvent molecules have been omitted for clarity. (b) Highlight of the 18-MC-6 ring with the captured [DyIII 2(μ3-OH)2]4+ core. Hydrogen atoms were placed on the μ3-OH anions. The ellipsoid plots are at the 50% level. Color scheme: DyIIIyellow, AlIIIgreen, oxygenred, nitrogenblue, and carbongray [symmetry code: (i) −x + 1, −y + 1, −z].
The main structure of the Dy III 2 Al III 6 [18-MC-6] is positioned about an inversion center located between the μ3-OH anions that bridge between two AlIII and one DyIII ions. The overall connectivity of the molecule is that of an 18-MC-6, as the AlIII ions form a ring about the outer perimeter of the structure, and N–O bridges exist between each AlIII ion (Figure b). However, the connectivity does not strictly match that of archetypal MCs. In standard MC systems, the N–O linkage does not reverse about the ring to O–N. In the Dy III 2 Al III 6 [18-MC-6], the connectivity does not strictly remain −[Al–N–O]–. Instead, the connectivity in the Dy III 2 Al III 6 [18-MC-6] follows a pattern of −[Al1–O–N–Al2–N–O–Al3–O–N]– that recurs twice to generate the 18-MC-6. In this pattern, the middle AlIII ion does not bind to an oxime atom of shi3–; thus, the standard MC model is not produced. In the structure, only the six shi3– ligands provide nitrogen–oxygen connectivity between the AlIII centers of the 18-MC-6 ring. The four H2shi– ligands form bridges between the DyIII and AlIII ions, but the H2shi– ligands do not participate in the MC connectivity. In essence, the H2shi– ligand helps tether the DyIII ions to the MC cavity similar to carboxylate anions such as acetate do in archetypal LnIII[12-MCM(III)N(shi)-4] structures, where MIII = Mn, Al, or Ga. ,, Moreover, archetypal MCs capture a single metal ion in the central cavity. However, for the Dy III 2 Al III 6 [18-MC-6], a [DyIII 2(μ3-OH)2]4+ core comprises the center of the molecule. The capture of a bimetallic core is reminiscent of other nonarchetypal MCs based on lanthanide and manganese ions. A DyIII 4MnIII 4 16-MC-6 structure with a heterobimetallic MC ring and a [DyIII-O–MnIII–N–O–MnIII–N–O] connectivity also captures a [DyIII 2(μ3-OH)2]4+ core. Furthermore, a HoIII 4MnIII 6 22-MC-8 structure with a heterobimetallic MC ring and a [HoIII-O–N–MnIII–O–MnIII–N–O–MnIII–N–O] connectivity captures a [HoIII 2(μ3-OH)2]4+ core.
The captured DyIII ions of the [DyIII 2(μ3-OH)2]4+ core are nine-coordinate. A SHAPE analysis (SHAPE 2.1; Table S3) of the geometry reveals that the lowest Continuous Shapes Measures (CShM) value is for a muffin configuration (C s; CShM = 1.385; Figure S1), though the CShM value for a spherical capped square antiprism (C 4v; 1.405) is comparable. A muffin geometry is best described as a shape with a trigonal base, a pentagonal equatorial plane, and a single-point vertex. , The DyIII–DyIII distance across the core is 6.29 Å. Each DyIII ion binds to three shi3– ligands, two H2shi– ligands, a μ3-OH anion, an oxygen atom of a nondisordered water molecule, and an oxygen atom of a disordered solvent molecule (water or methanol). The three shi3– ligands and one of the H2shi- ligands donate an oxime oxygen atom in a monodentate fashion to the DyIII ion. The other H2shi– ligand binds in a bidentate fashion by using the carbonyl and oxime oxygen atoms of the ligand to form a five-membered chelate ring. The oxime oxygen atoms of the shi3– and H2shi– ligands also bridge to the ring AlIII ions (Al1–Al3), helping to secure the DyIII ions to the MC cavity. In addition, each DyIII ion binds to a μ3-OH anion that bridges to Al3 and the other Al3 related by the inversion center of the molecule. Lastly, the coordination sphere of each DyIII ion is completed by two solvent oxygen atoms, an oxygen atom (O17) of a nondisordered water molecule, and an oxygen atom of a disordered solvent molecule, with all solvent molecules associated with O19. The disordered solvent molecule is either a water molecule or a methanol molecule with its methyl group disordered over two positions. The occupancy of the water molecule was refined to 0.120(3), while the two positions of the methyl groups of the methanol molecule were refined to 0.6223(17) and 0.257(3).
The ring AlIII ions are six-coordinate with an octahedral geometry (Table S4 and Figure S2). The three unique AlIII ions of the ring each have slightly different coordination spheres. For Al1, the coordination consists of three bidentate ligands in a propeller configuration with Δ stereoisomerism. For the aluminum ion Al1i, which is related by the inversion center [symmetry code: (i) −x + 1, −y + 1, −z], the stereoisomerism is Λ. The propeller is constructed by one bidentate H2shi– ligand that binds with the oxime and carbonyl oxygen atoms to form a five-membered chelate ring, a shi3– ligand that forms a similar five-membered chelate ring, and a shi3– ligand that binds with the oxime nitrogen atom and phenolate oxygen atom to form a six-membered chelate ring. For Al2, the coordination consists of two cis bidentate shi3– ligands, an oxime oxygen atom of a H2shi–, and a nitrogen atom of a pyridine molecule that is disordered over two positions with occupancy rates of 0.538(16)–0.462(16). Both shi3– ligands form six-membered chelate rings by binding with the oxime nitrogen atom and phenolate oxygen atom of the ligands. The cis configuration of the ligands about Al2 imparts a Λ-like stereoisomerism to the metal center and a Δ-like stereoisomerism for the symmetry equivalent Al2i. For Al3, the coordination is composed of two cis bidentate shi3– ligands and two oxygen atoms of cis μ3-OH anions. Both shi3– ligands form five-membered chelate rings by binding with the oxime and carbonyl oxygen atoms of the ligands. The cis configuration of the ligands imparts a Λ-like stereoisomerism to the metal center and a Δ-like stereoisomerism for the symmetry equivalent Al3i. Thus, the stereoisomerimetric pattern of the AlIII ions [Al1–Al2–Al3–Al1i–Al2i–Al3i] about the 18-MC-6 ring is ΔΛΛΛΔΔ.
B. Magnetometry and Model Hamiltonian
The static and dynamic magnetic properties of Dy III Al III 4 [12-MC-4], Dy III 2 Al III 8 [12-MC-4] 2 , Dy III Al III 6 [3.3.1] MCr, and Dy III 2 Al III 6 [18-MC-6] were investigated, and the experimental results were modeled with a spin Hamiltonian framework, as detailed below.
In the complexes investigated in this work, the central magnetic ion hosted in the MCs/MCr cavity is a DyIII ([Xe] 6s 2 4f 9) with a ground configuration 6 H 15/2 for the free ion (S = 5/2, L = 5 and J = 15/2). In the weak field approximation (dominant spin–orbit coupling), J is a good quantum number, and the crystal field Hamiltonian acts on the |SLJM⟩ multiplet. The spin Hamiltonian of such a system is defined as follows:
| 1 |
The first term, , represents the Zeeman coupling, which accounts for the interaction with the applied static magnetic field that splits the ground-state Kramers degeneracy. The second term, , is the crystal field contribution, describing the effect of the ligand field environment on the energy levels of the |SLJM⟩ multiplets. Here, the crystal field potential is expressed using Steven’s equivalent operators Ô k i , derived from the ligand field symmetry of the DyIII ions, with B k i crystal field parameters and θ k operator equivalent factors (tabulated for lanthanides in |J,m J ⟩ basis). The third term in eq describes the interaction between the magnetic ions. Although this term can reasonably be assumed to be negligible for intermolecular interactions, it cannot be ignored for dimers, where intramolecular interactions occur over much shorter distances. In this case, through-space dipolar coupling is the most appropriate mechanism to describe the interaction, as the significant spatial separation between ions (even within the same molecule) strongly suppresses exchange-like coupling mechanisms. The dipolar coupling is characterized by an interaction coefficient J 12 , between the two magnetic centers, that in the point dipole approximation is written as
| 2 |
where R is the vector that connects the two magnetic ions.
For each complex, magnetometry and susceptibility measurements were collected on powder samples with a MPMSXL5 SQUID magnetometer (Quantum Design). The magnetization was measured as a function of temperature, M(T), under a direct current (DC) applied magnetic field of 10 kOe from 300 to 2 K. Field-dependent magnetization measurements, M(H), were collected at 2 K by changing the applied magnetic field from 0 to 5 T (Figure ). The temperature and field dependences of the system magnetization were simultaneously fitted for all compounds using the model spin Hamiltonian of eq (Figure ). The fitting procedure was performed using the PHI program.
3.

(a, b, e, f) Field dependence of the magnetization M(H), measured at 2 K, and (c, d, g, h) temperature dependence of the magnetic susceptibility product with temperature χT measured in an applied static field of 10 kOe. Inset of (a, b, e, f): a sketch of the corresponding molecular structure of Dy III Al III 4 [12-MC-4], Dy III 2 Al III 8 [12-MC-4] 2 , Dy III Al III 6 [3.3.1] MCr, and Dy III 2 Al III 6 [18-MC-6]. All data are corrected for the contribution of the sample diamagnetism, estimated from Pascal’s constants. The red lines represent the best-fit result obtained with Hamiltonian (1). Shaded error bars are estimated within a reasonable approximation of the cumulative experimental uncertainty to 5% of the measured values.
Finally, alternating current (AC) magnetic susceptibility measurements were conducted to assess the magnetization dynamics of these complexes (Figure ). The AC susceptibility was measured by varying the frequency of the AC field in the 2–10 K temperature range under zero and 800 Oe static magnetic field. Comparison of the results obtained reveals the main differences in the dynamic magnetic behavior between the structures with different coordination geometries about the DyIII ions.
4.

Frequency dependence of the out-of-phase susceptibilities for Dy III Al III 4 [12-MC-4], Dy III 2 Al III 8 [12-MC-4] 2 , Dy III Al III 6 [3.3.1] MCr, and Dy III 2 Al III 6 [18-MC-6] (specified in inset), measured at different temperatures in zero (a, c, e, g) and an applied (800 Oe,[b, d, f, h]) static field.
3.2.1. Static and Dynamic Magnetic Behavior
In the case of Dy III Al III 4 [12-MC-4], the geometry about the eight-coordinate DyIII ion is square antiprism (SAP), which is indicative of a D 4d coordination polyhedron with tetragonal symmetry. For exact D 4d coordination, the Steven’s operator set necessary to describe the crystal field simplifies from the broader category of tetragonal symmetries to only include the axial terms:
(Table S5). Further details on the molecular structure, obtained from X-ray diffraction, reveal that the characteristic cubic angle of SAP θc = 54.74°, for which the B 0 term vanishes, is here slightly distorted and even between the upper and lower pyramids of the SAP geometry. Therefore, in this regime in which the cubic angle approaches the condition in which B 0 vanishes and given its nonunivocal prolate or oblate deformation, a reasonable approximation that also reduces the number of free parameters is to model the CF assuming an effective exact SAP with only the B 0 and B 0 operators. The other characteristic angle of the SAP geometry, that is, the twist angle between the two base planes ϕ D4d = 45° is here only slightly reduced to ϕ = 41.8° and supports the assumption of an effective ideal D 4d to approximate the system CF. Moreover, in the point charge electrostatic model (PCEM) approximation, we assume a priori the signs of the crystal field parameters, depending on the values observed for the typical angles (θ c ,ϕ) of the SAP ligand field.
Assuming for the Zeeman term of eq the isotropic Landé factor of the DyIII free ion g j = 4/3, and neglecting the intermolecular dipolar coupling, the CF parameters defined above are the only fitting parameters of the model. In Figure a,c, we observe that the experimental magnetization and susceptibility data are well reproduced by the model Hamiltonian. The low-temperature magnetization shows a sharp increase for a small static field and a close to linear increment above 1 T. Not reaching magnetization saturation even at high field is symptomatic of the presence of low-lying excited states, contributed through Van Vleck mechanisms. The low temperature χT instead shows a sharp decrease, characteristic of the depopulation of the Zeeman-split crystal field levels, and a close-to-flat response above 50 K, reaching at room temperature a value compatible with that expected for a noninteracting DyIII ion in the 6 H 15/2, S = 5/2, L = 5, J = 15/2, and g = 4/3 configuration (14.17 cm3 K/mol).
According to the spin Hamiltonian that reproduces the data, the system eigenstates (Table S6) are structured with a |13/2⟩ ground-state doublet, unmixed with other CF states, because of the absence of off-diagonal matrix elements in the CF energy matrix (generated by q ≠ 0 parameters). The gap with the first excited doublet |11/2⟩ is 18.3 cm–1. This arrangement of states is in agreement with other examples of Dy complexes in D 4d symmetry with moderate CF, , while in contrast to the configuration of strong axial CF, such as the one found (in a markedly different DyIII coordination geometry) for the dysprosocenium complex, where the large B 0 CF term stabilizes the |15/2⟩ ground state. , Additionally, we found for Dy III Al III 4 [12-MC-4] a modest overall CF splitting of the |Jm J ⟩ multiplet (∼64 cm–1), compared to compounds with a more pronounced axial CF.
When the benzoate anions of Dy III Al III 4 [12-MC-4] are substituted with isophthalate anions, a dimeric 12-MC-4 complex is generated, Dy III 2 Al III 8 [12-MC-4] 2 , and the DyIII ions of the dimer maintain an SAP coordination geometry (D 4d). Therefore, for the fitting of the magnetometry data, we reasonably expect similar CF parameters obtained from the fitting of the Dy III Al III 4 [12-MC-4] data, which were thus chosen as a starting point for the fit of the dimer. Furthermore, the contribution arising from the dipolar coupling constant was calculated a priori from the distance between the Dy center of 7.12 Å, determined by X-ray diffraction. Once more, a satisfactory concordance is achieved between the experimental data for Dy III 2 Al III 8 [12-MC-4] 2 and the theoretical curves derived from the model spin Hamiltonian, with similar crystal field parameters (Table S5), yielding an eigenstate spectrum comparable to those of the monomer. Indeed, the profile of M(H) and χT is very close to what is measured for its monomer, with the exception of a factor of 2 in the magnetization values across the whole field range and a saturation value for the χT product at room temperature, which matches the value expected for 2 isolated DyIII ions (Figure b,d). Therefore, we conclude that because of the significant space separation between the DyIII ions, the dipolar contribution here plays a minor contribution. To validate the modification in the modeled CF, we simulated the system eigenstates. The total CF multiplet splitting of Dy III 2 Al III 8 [12-MC-4] 2 is slightly more than doubled with respect to Dy III Al III 4 [12-MC-4], while the eigenstate spectra are mainly unchanged, with a ground state of the |13/2,13/2⟩ quartet, in which the four states |±13/2,∓13/2⟩ and |±13/2,±13/2⟩ are quasi-degenerate due to the small dipolar interaction. The gap with the first excited |13/2,11/2⟩, |11/2,13/2⟩ octet is about 14.1 cm–1. The eigenstates of the system as a function of the applied field are reported in the Supporting Information for both Dy III Al III 4 [12-MC-4] and Dy III 2 Al III 8 [12-MC-4] 2 (Figure S3 and Table S6).
The small differences between the low-energy spectra of Dy III Al III 4 [12-MC-4] and Dy III 2 Al III 8 [12-MC-4] 2 deduced from DC magnetometry reflect on the analogies observed in AC susceptibility measurements. No out-of-phase susceptibility χ″ is detected for either Dy III Al III 4 [12-MC-4] or Dy III 2 Al III 8 [12-MC-4] 2 at temperatures above 2 K (Figure a,c); thus, they do not display SMM behavior in the experimentally accessible regime. However, a minor χ″ component is recovered with the application of a small static field (800 Oe). Similar behavior was noted for two YbIIIZnII 4 [12-MC-4] complexes and a DyIIICdII 16 [12-MC-4]2[24-MC-4] system, where the DyIII ion is sandwiched between two 12-MC-4 units. In these other systems, the YbIII and DyIII ions also have a square antiprism geometry (D 4d) with little axial CF contribution as in the Dy-Al 12-MC-4 complexes. For the DyIII and YbIII 12-MC-4 systems, the presence of a χ″ component with the application of a small state field could be associated with the suppression of fast relaxation dynamics induced by the quantum tunneling of magnetization (QTM). However, even in an applied static field, neither complex shows any χ″ peak maxima above 2 K. This indicates that the relaxation rate exceeds the maximum detectable limit (1 kHz) within this temperature range (Figure b,d). In this context, the dimer Dy III 2 Al III 8 [12-MC-4] 2 demonstrates a slightly more stable ground state, as indicated by its out-of-phase susceptibility at 2 K, which nearly reaches its maximum at the highest AC frequency measured. Conversely, the monomer Dy III Al III 4 [12-MC-4] does not exhibit a comparable approach to saturation within the same frequency range, implying faster relaxation processes, unresolved in the accessible frequency range.
For Dy III Al III 6 [3.3.1] MCr, the nine-coordinate DyIII ion has a distorted spherical capped square antiprism geometry (C 4v). The degree of deviation from the ideal geometry results in a continuous shape measure approaching 1. Here, the crystal field Hamiltonian is modeled with five parameters:
(Table S5). However, because of the significant distortions of the C 4v symmetry group and the already large number of CF parameters, the fitting of the magnetometry data could suffer from overparameterization and the resulting correlations in CF parameters. Therefore, because of the stronger CF axiality (compared to monomer Dy III Al III 4 [12-MC-4]) promoted by the monocapping of the tetragonal symmetry, for the interpretation of the experimental measurements, we opted for an effective purely axial CF. This effective CF is sufficient to model adequately the magnetometry data (Figure e,g). Introducing off-axial components does not significantly improve the model fitting and affects its unambiguity. Moreover, due to the significantly distorted coordination, we did not constrain the sign of the CF parameters, depending on the expectation from the PCEM approximation for the exact C 4v group, which are strongly influenced by the characteristic angle of the ligand coordination. Again, for this monomer system, we neglected the intermolecular dipolar coupling. Therefore, the magnetometry data are fitted (Figure e,g) only with the effective axial CF defined above. In contrast to the results from Dy III Al III 4 [12-MC-4] and Dy III 2 Al III 8 [12-MC-4] 2 , the M(H) of this metallacryptate system shows a less sloped profile above 1 T, almost approaching saturation (Figure e). This is associated with a mitigated contribution of the low-lying excited states through Van Vleck mechanisms. A similar trend is instead found for the χT product, where at low temperature a sharp decrease indicates the depopulation of the Zeeman split crystal field levels, and a plateau is observed above 50 K to the expected room temperature value for the noninteracting DyIII ion in the 6 H 15/2, S = 5/2, L = 5, J = 15/2 and g = 4/3 configuration (Figure g). The computed CF from fitting the magnetometry data reveals a significantly more pronounced B 0 term than what is observed in Dy III Al III 4 [12-MC-4] and Dy III 2 Al III 8 [12-MC-4] 2 , contributing to the stabilization of a |15/2⟩ ground state, characteristic of 4f systems under strong axial CF. ,,− Nonetheless, the CF in Dy III Al III 6 [3.3.1] MCr is insufficient for significantly isolating the ground state doublet from the excited states. Indeed, the first excited state is the |13/2⟩ multiplet, which is separated by a modest gap of ≈8 cm–1, while the total extent of the CF multiplet splitting is of ≈ 150 cm–1 (Figure S3 and Table S6).
For Dy III 2 Al III 6 [18-MC-6], a [DyIII 2(μ3-OH)2]4+ core is captured in the MC ring where both DyIII ions are nine-coordinate. Compared to Dy III Al III 6 [3.3.1] MCr, the symmetry of the DyIII ions in Dy III 2 Al III 6 [18-MC-6] is lowered to C s as the geometry is best described as muffin. Unfortunately, the number of CF parameters in Steven’s notation derived from this very low symmetry coordination is incompatible with a reliable fitting. Thus, the most convenient approach for interpreting the magnetometry data of Dy III 2 Al III 6 [18-MC-6] is to define an effective CF model, starting from the three axial CF parameters obtained from Dy III Al III 6 [3.3.1] MCr and refining the fit without the introduction of additional off-axial parameters. From the comparison (Figure f,h), we conclude that this effective axial CF is sufficient to appropriately model the magnetometry data of Dy III 2 Al III 6 [18-MC-6]. Similarly to Dy III 2 Al III 8 [12-MC-4] 2 , because of the space separation between DyIII-ions in Dy III 2 Al III 6 [18-MC-6] (6.29 Å), the contribution to the spin Hamiltonian 1 of the DyIII–DyIII interaction is assumed to be purely dipolar and the coefficient is calculated from eq . The best-fit curves reproduce both M(H) and χT(T) (Figure f,h). The resulting effective CF maintains more robust axial components, with a dominant B 0 term, as also observed in Dy III Al III 6 [3.3.1] MCr, compared to what is found for the 12-MC-4 complexes. This contributes to the stabilization in the easy-axis direction of a ground state |15/2,15/2⟩ quartet, in which the |±15/2,∓15/2⟩ and |±15/2,±15/2⟩ states are quasi-degenerate because of the small dipolar interaction. Similarly to Dy III Al III 6 [3.3.1], a modest gap of 8.5 cm–1 is found with the first excited octet (|15/2,13/2⟩,|13/2,15/2⟩). Additionally, we found an enhanced CF multiplet total splitting of ≈510 cm–1 (Figure S3 and Table S6). The calculated system eigenstates as a function of the applied field are reported in the Supporting Information for both Dy III Al III 6 [3.3.1] and Dy III 2 Al III 6 [18-MC-6] (Table S6).
AC susceptibility measurements reveal that both Dy III Al III 6 [3.3.1] and Dy III 2 Al III 6 [18-MC-6] exhibit SMM behavior in zero applied field. However, at 2 K, the particularly fast relaxation rates for both compounds exceed the upper detection limit (1 kHz), as evidenced by the absence of a maximum in the measured out-of-phase susceptibility peak within the accessible frequency range. Also, for these complexes, the dinuclear compound exhibits a slower magnetization relaxation time in a zero static field, with a χ″ that approaches its maximum close to 1 kHz for Dy III 2 Al III 6 [18-MC-6] at 2 K. Upon application of a small static field (800 Oe), the relaxation rates of both systems fall within the accessible frequency window for most of the temperatures investigated (below ∼8 K). Under these conditions, the χ″ peak becomes observable and shifts to higher frequencies as the temperature increases. This trend indicates that thermally activated relaxation mechanisms involving coupling to molecular vibrations play a dominant role in accelerating the magnetization dynamics at higher temperatures. Notably, the slowest relaxation rate is observed for Dy III Al III 6 [3.3.1] at 2 K.
The main structural difference between the compounds, besides the class of MC, is the coordination geometry of the DyIII ions. For Dy III Al III 4 [12-MC-4] and Dy III 2 Al III 8 [12-MC-4] 2 , the eight-coordinate DyIII ions reside in a square antiprism geometry (D 4d) with little axial contribution. However, for Dy III Al III 6 [3.3.1] MCr and Dy III 2 Al III 6 [18-MC-6], the nine-coordinate DyIII ions have a spherical capped square antiprism (C 4v) and muffin (C s) geometry, respectively, with a more pronounced axial contribution to the crystal field. As extensively reported in the literature, a strong axial ligand field is a key ingredient enhancing the magnetic anisotropy, stabilizing the magnetic ground state, and effectively suppressing QTM. Consequently, it contributes to the SMM behavior of oblate ions such as DyIII. ,, In fact, the strong axial ligand field has been identified as a crucial factor for the exceptionally high blocking temperatures observed in dysprosocenium complexes, which possess some of the largest U eff values reported to date. ,, Additionally, these systems benefit from a reduced coupling between the spin and molecular vibrations, further limiting relaxation dynamics. , For Dy III Al III 4 [12-MC-4] and Dy III 2 Al III 8 [12-MC-4] 2 , the lack of a significant axial ligand field results in the nonexistence of SMM behavior in the absence of a static magnetic field, and a frequency-dependent out-of-phase magnetic susceptibility signal is only observed when a 800 Oe static magnetic field is applied.
For the investigated MCs, the axial ligands of Dy III Al III 6 [3.3.1] MCr and Dy III 2 Al III 6 [18-MC-6] allow the observation of SMM behavior under an applied static magnetic field of 800 Oe and the determination of the U eff and the pre-exponential frequency factor (τ0) values. The AC data were analyzed with a Cole–Cole plot (Figure a,b) and fitted to the generalized Debye model to extract the relaxation times τ at different temperatures and their distribution α. , Here, the parameters of the model were refined by fitting simultaneously both the generalized definitions for χ′(ω) and χ″(ω):
| 3 |
where α > 0 indicates a wide distribution of relaxation times τ and χS, χT represents the adiabatic and isothermal limits of the susceptibility, respectively. Equation correctly reproduced the experimental AC data. The resulting values for α vary in the range 0.36–0.42 and 0.52–0.68 for Dy III Al III 6 [3.3.1] MCr and Dy III 2 Al III 6 [18-MC-6], respectively, indicating both a uniformly distributed relaxation process and a broader distribution of rates in the Dy III dinuclear system (fitting parameters in Table S7). This broadening can be attributed to a more pronounced structural disorder in the dinuclear complex, which is reflected in the magnetic environment responsible for the relaxation dynamics of the systems.
5.

Temperature dependence of the Cole–Cole plots for (a) Dy III Al III 6 [3.3.1] MCr and (b) Dy III 2 Al III 6 [18-MC-6] in an applied static field of 800 Oe (dots). The solid line represents the fit with a generalized Debye model, as specified in the text. Relaxation rates (dots) extracted from the fitting as a function of temperature for (c) Dy III Al III 6 [3.3.1] MCr and (d) Dy III 2 Al III 6 [18-MC-6], respectively, are modeled (solid lines) with a combination of Raman and Orbach relaxation mechanisms, as detailed in the main text.
The relaxation rates were extracted by fitting the susceptibility as a function of the temperature (Figure c,d). These are expected to be governed by a single rate Arrhenius-like behavior τOrbach = τ0 e –U eff/K B T in the higher temperature regime, where thermally activated mechanisms play a key role. Conversely, in the intermediate temperature regime, the relaxation rate is dominated by nonresonant Raman processes, which can be modeled with an effective power-law temperature dependence τRaman = CT n . − The relaxation rate is then fitted, in the corresponding temperature range, as the sum of these relaxation models. We extracted an effective energy barrier U eff of 59 ± 2 cm–1 with τ0 = 2.5 ± 0.7 × 10–8 s and a Raman exponent n = 4.38 ± 0.04 with C Raman = 1.3 ± 0.1 K–n s–1 for complex Dy III Al III 6 [3.3.1] MCr, and U eff = 38 ± 4 cm–1 with τ0 = 2.4 ± 1.1 × 10–8 s, and n = 2 ± 0.4 with C Raman = 110 ± 60 K–n s–1 for Dy III 2 Al III 6 [18-MC-6].
In the Orbach regime, the differences between the complexes are less pronounced. The effective energy barrier is nearly halved in the dinuclear complex Dy III 2 Al III 6 [18-MC-6], where this relaxation pathway becomes active at lower temperatures (see Figure c,d). For the mononuclear complex Dy III Al III 6 [3.3.1] MCr, the extracted barrier aligns well with the eigenstate spectrum obtained from fitting the DC susceptibility data. In contrast, for Dy III 2 Al III 6 [18-MC-6], the observed discrepancy can be attributed to the invasive assumption of neglecting nonaxial contributions to the CF (Table S6). In the Raman regime, although it is characterized here by a limited number of data points, a tentative comparison can still be made. The markedly different Raman prefactor C Raman between the two complexes may suggest a richer low-energy phonon spectrum for Dy III 2 Al III 6 [18-MC-6], leading to a steeper phonon density of states (pDOS). , Additionally, the significantly reduced Raman exponent in Dy III 2 Al III 6 [18-MC-6] (approaching the high-temperature limit ,, where n ≈ 2) points to a higher-energy Debye cutoff, with low-lying optical modes remaining dispersive up to higher energies. ,
3.2.2. Comparison between DyIII–AlIII and DyIII–GaIII MCs
While the magnetic properties of LnIII-AlIII MCs have not been reported before, the magnetic properties of several similar DyIII–GaIII MCs have been previously investigated (Table ). In 2019, Jiang, Shao, and co-workers reported a DyIIIGaIII 4 [12-MC-4] that displays slow magnetization relaxation under an applied DC magnetic field. However, unlike the slightly domed Dy III Al III 4 [12-MC-4], the DyIIIGaIII 4 [12-MC-4] is significantly nonplanar as it is made with MC framework ligand 3-hydroxy-2-naphthanoic hydroxamic acid (H3napt). Four napt3– comprise the MC unit, while two singly deprotonated H2napt– ligands bridge between the ring GaIII ions and the central DyIII ion. The use of H2napt– instead of benzoate likely leads to the nonplanarity of the 12-MC-4. In addition, the four-ring GaIII ions do not lie in the same plane as in the Dy III Al III 4 [12-MC-4]. Furthermore, the central DyIII of DyIIIGaIII 4 [12-MC-4] is nine-coordinate with a spherical tricapped trigonal prism geometry (D 3h).
1. Comparison of SMM Parameters of DyIII–AlIII and DyIII–GaIII Ms.
| MC | Dy III ion coordination number and geometry | U eff (cm –1 ) | τ 0 (s) |
|---|---|---|---|
| Dy III Al III 4 [12-MC-4] | 8; square antiprism (D 4d) | n.d. | n.d. |
| DyIIIGaIII 4 [12-MC-4] | 9; tricapped trigonal prism geometry (D 3h) | 18.3 | 2.2 × 10–6 |
| Dy III 2 Al III 8 [12-MC-4] 2 | 8; square antiprism (D 4d) | n.d. | n.d. |
| DyIIIGaIII 8 [12-MC-4]2 | 8; square antiprism (D 4d) | 27.1 | 2.27 × 10–8 |
| Dy III Al III 6 [3.3.1] MCr | 9; capped square antiprism geometry (C 4v) | 59 | 2.5 × 10–8 |
| DyIIIGaIII 6 [3.3.1] MCr | 9; tricapped trigonal prism geometry (D 3h) | 8.83 | 3.6 × 10–6 |
| Dy III 2 Al III 6 [18-MC-6] | 9; muffin (C s) | 38 | 2.4 × 10–8 |
| DyIII 2GaIII 4 [16-MC-6] | 8; triangular dodecahedral geometry (D 2d) | 13 to 18 | 3.6 to 6.8 × 10–6 |
The central DyIII ion binds to the oxime oxygen atoms of four napt3–, the oxime oxygen atoms of two H2napt–, the carbonyl oxygen atom of one H2napt–, and a bidentate nitrate ion. In Dy III Al III 4 [12-MC-4], the central DyIII ion is eight-coordinate with a square antiprism geometry (D 4d). As in Dy III Al III 4 [12-MC-4], the DyIIIGaIII 4 [12-MC-4] complexes do not exhibit a response in the out-of-phase magnetic susceptibility with zero applied magnetic field; however, a peak maximum was observed under applied DC magnetic fields between 200 and 2000 Oe. Under an applied DC magnetic field of 800 Oe, the authors determined the U eff and τ0 values considering an Orbach-only process as 18.3 cm–1 and 2.2 × 10–6 s, respectively, and a process that considered all relaxation mechanisms (direct, QTM, Orbach, and Raman), as 18.18 cm–1 and 2.97 × 10–6 s, respectively. In 2019, Athanasopoulou, Rentschler, and co-workers reported a DyIIIGaIII 8 [12-MC-4]2 dimer where two GaIII 4 12-MC-4 units, based on shi3– ligands, bind to one DyIII ion. The DyIII ion is captured in the central cavity of each MC to form a sandwich complex. The eight-coordinate DyIII ion has a square antiprism geometry (D 4d), similar to the Dy III Al III 4 [12-MC-4] and Dy III 2 Al III 8 [12-MC-4] 2 complexes. For the DyIIIGaIII 8 [12-MC-4]2 dimer, the ground state is mainly composed by the |11/2> microstate with the first excited state being |13/2>, where instead for Dy III Al III 4 [12-MC-4] and Dy III 2 Al III 8 [12-MC-4] 2 the ground states are |13/2> and |13/2,13/2⟩, respectively, and the first excited states are |11/2> and |13/2,11/2⟩, |11/2,13/2⟩, respectively. The DyIIIGaIII 8 [12-MC-4]2 dimer exhibited an out-of-phase magnetic susceptibility signal in zero applied DC magnetic field; however, peak maxima are not observed. Under a 1000 Oe applied DC magnetic field, peak maxima are observed, and considering both Orbach and Raman relaxation processes, the authors determined U eff to be 27.1 cm–1 with a τ0 of 2.27 × 10–8 s. In 2018, Lutter, Pecoraro, and co-workers reported several LnIIIGaIII 6 [3.3.1] metallacryptates (LnIII = PrIII, NdIII, and SmIII–YbIII) that are structurally similar to the Dy III Al III 6 [3.3.1] MCr. The LnIIIGaIII 6 [3.3.1] MCrs contains one H2shi–, one Hshi2–, and 7 shi3– ligands and three triethylammonium countercations. The LnIII ion is nine-coordinated with a spherical tricapped trigonal prism geometry (D 3h). Conversely, the Dy III Al III 6 [3.3.1] MCr contains two H2shi– and 7 shi3– ligands and two pyridinium countercations, and the nine-coordinate DyIII ion is in a spherical capped square antiprism geometry (C 4v). However, the overall MCr connectivity is similar between both structures. The NdIIIGaIII 6, DyIIIGaIII 6, and YbIIIGaIII 6 [3.3.1] MCrs displayed a slow magnetization relaxation in the presence of a 1000 Oe (NdIII and YbIII) or 750 Oe (DyIII) applied DC magnetic field; however, only the DyIIIGaIII 6 [3.3.1] MCr possessed a slow magnetization relaxation in zero applied field, though it is a weak response without a peak maximum in the out-of-phase magnetic susceptibility as in the Dy III Al III 6 [3.3.1] MCr. Using the 750 Oe applied DC field, the authors determined for the DyIIIGaIII 6 [3.3.1] MCr that the molecule contains one barrier to magnetization relaxation that follows an Orbach process. The U eff value was 8.83 cm–1 with a τ0 of 3.6 × 10–6 s. In 2015, Pecoraro, Mallah, and co-workers reported a series of LnIII 2GaIII 4 MC-like complexes (LnIII = GdIII, TbIII, DyIII, ErIII, YIII, and YIII 0.9DyIII 0.1) built with H3shi ligands. The molecule is not an archetype MC as both N–O and oxygen-only connectivity exist between the metal ions, yet the complex can be considered a collapsed 16-MC-6 as it does not have a central cavity. Each DyIII ion is eight-coordinated with a triangular dodecahedral geometry (D 2d). Of the presented DyIII–AlIII MCs, the closest comparable structure would be Dy III 2 Al III 6 [18-MC-6]; however, the DyIII ions in Dy III 2 Al III 6 [18-MC-6] are nine-coordinated with muffin geometry (C s). As in Dy III 2 Al III 6 [18-MC-6], the DyIII 2GaIII 4 [16-MC-6] exhibited an out-of-phase magnetic susceptibility signal in zero applied DC magnetic field. However, the DyIII 2GaIII 4 [16-MC-6] complex possesses two relaxation processes: one process at lower temperatures (2–5 K) and one process at higher temperatures (10–14 K). At lower temperatures, the DyIII ions are antiferromagnetically coupled, but an excited ferromagnetic state is accessible. The slow magnetization relaxation at lower temperatures is due to this populated excited state, and the authors determined U eff to be 13 cm–1 with a τ0 of 3.6 × 10–6 s. At higher temperatures, the DyIII ions are uncoupled and behave as isolated DyIII ions with a U eff of 18 cm–1 and a τ0 of 6.8 × 10–6 s. Though direct comparisons between the aluminum and gallium MC analogues are difficult, as the ligand sets, MC shape, DyIII ground states, and experimental conditions are sometimes different, the DyIII–AlIII MC U eff values tend to be larger than those of their GaIII counterpart. For Dy III Al III 6 [3.3.1] MCr and Dy III 2 Al III 6 [18-MC-6], the DyIII ions have geometries with a greater axial component, which may be one of the contributing factors leading to the higher U eff values.
4. Conclusions
The four MC classes investigated reveal how systematic synthetic control through manipulation of stoichiometric ratios of the starting components can lead to the reliable formation of certain MCs, and within each structure, the coordination environment and geometries of the central DyIII ions can be controlled by ligand scaffolding. For Dy III Al III 4 [12-MC-4] and Dy III 2 Al III 8 [12-MC-4] 2 , the DyIII ions are sequestered in an eight-coordinate ligand environment with square antiprism geometry with little axial contribution to the ligand field. However, for Dy III Al III 6 [3.3.1] MCr, and Dy III 2 Al III 6 [18-MC-6], the nine-coordinate DyIII ions with spherical capped square antiprism or muffin geometry, respectively, have a greater axial component to the ligand field, and this axial contribution has a direct effect on the dynamic magnetic properties of the molecules.
From the perspective of SMM applications, all of the systems studied in this work reveal a slow relaxation of the magnetization at very low temperature and in an applied static field. However, only Dy III Al III 6 [3.3.1] and Dy III 2 Al III 6 [18-MC-6] display a slow relaxation of the magnetization in the absence of a static magnetic field. In addition, a peak in the out-of-phase susceptibility is observed in 800 Oe applied field with U eff equal to 59 ± 2 and 38 ± 4 cm–1, and τ0 equal to 2.5 × 10–8 and 2.4 × 10–8 s for Dy III Al III 6 [3.3.1] and Dy III 2 Al III 6 [18-MC-6], respectively. Thus, a key factor in determining the increased relaxation rate is attributed here to the axial nature of the crystal field on DyIII ions. Therefore, to further improve the performance of these complexes, a valuable strategy would be to further engineer the Dy-ligand field by chemical substitution to deploy an axial structural modification of the molecule. Future AC susceptibility measurements at different applied fields and higher frequencies could also enable the disentangling of the contribution of different relaxation mechanisms. In addition, to gain greater insight into the magnetization relaxation mechanisms of the Dy III 2 Al III 6 [18-MC-6] system, diluted compounds where one of the DyIII ions is replaced with a diamagnetic center, such as YIII or LuIII, could also be performed. For example, diluted Dy/Y 18-MC-6 could better elucidate the relaxation processes occurring in the MC and allow for a better comparison of the molecules. Future work in our laboratories is directed toward the diluted versions of these MCs. Moreover, ab initio calculations of the complete crystal field tensors, phonon spectra, and vibrational density of states could serve as a valuable guide for rational structural design. Such a computational analysis would enable a more comprehensive understanding of the individual contributions of various relaxation mechanisms to the overall magnetization dynamics.
Supplementary Material
Acknowledgments
This work received financial support from the European Union_NextGenerationEU, PNRR MUR project PE0000023-NQSTI (EG, MS, SC), the European Union’s Horizon 2020 program under grant agreement no. 862893 (FET-OPEN project FATMOLS) (EG, SC), the Novo Nordisk Foundation under grant NNF21OC0070832 in the call “Exploratory Interdisciplinary Synergy Programme 2021” (SCh, EG, SC). Additional support was provided by the Summer Undergraduate Research Experience (SURE) program, the Faculty Professional Development Council (FPDC) Grant Program, and the Undergraduate Research Program at Shippensburg University (CMZ). The single crystal X-ray diffractometer and upgrades were provided by the National Science Foundation through the Major Research Instrumentation Program under Grant No. CHE 1625543 and the 2020 Laboratory and University Core Facility Research Equipment Program, respectively (MZ).
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.5c07724.
Crystallographic details, continuous shapes measures, and Dy and Al first-coordination sphere diagrams of DyIII 2AlIII 6 [18-MC-6]; crystal field parameters, energy levels, eigenstates, and Cole–Cole parameters for investigated compounds (PDF)
The authors declare no competing financial interest.
Published as part of ACS Omega special issue “Undergraduate Research as the Stimulus for Scientific Progress in the USA”.
Footnotes
Bruker. Apex3 v2018.1-0, SAINT V8.38A; Bruker AXS Inc.: Madison, WI, 2018.
Sheldrick, G. M. SHELXL2018; University of Göttingen, Germany, 2018.
The main factors contributing to the experimental error are the sample mass (measured with a ±0.1 mg error) and the effective amount of solvent still incorporated in the crystal structure, which significantly influence the normalization into μB /f.u. and cm3 K/mol units.
The implementation of a descent of symmetry approach to account for the small distortion found here would have caused an unnecessary overparametrization of the fitting Hamiltonian and to higher uncertainty on all of the obtained CF parameters.
The axial component fit is solid, while the accuracy in determining the nonaxial CF parameters is significantly reduced (surpassing the model sensitivity) and relies heavily on the initial fit parameters.
The lower bound for the fitted Raman exponent has been fixed to the high-temperature limit , (high-energy Debye cutoff) of n = 2. It is worth noting that a rate approaching the linear dependence with the system temperature could also be associated with the establishment of direct relaxation processes (τ–1 = AH 2 T, with H static field and A prefactor), promoted by the applied static field and the small dipolar interaction, responsible for the admixture of excited states and the Kramers degeneracy breaking.
References
- Caneschi A., Gatteschi D., Sessoli R., Barra A. L., Brunel L. C., Guillot M.. Alternating Current Susceptibility, High Field Magnetization, and Millimeter Band EPR Evidence for a Ground S = 10 State in [Mn12O12(Ch3COO)16(H2O)4].2CH3COOH.4H2O. J. Am. Chem. Soc. 1991;113(15):5873–5874. doi: 10.1021/ja00015a057. [DOI] [Google Scholar]
- Sessoli R., Tsai H. L., Schake A. R., Wang S., Vincent J. B., Folting K., Gatteschi D., Christou G., Hendrickson D. N.. High-Spin Molecules: [Mn12O12(O2CR)16(H2O)4] J. Am. Chem. Soc. 1993;115(5):1804–1816. doi: 10.1021/ja00058a027. [DOI] [Google Scholar]
- Sessoli R., Gatteschi D., Caneschi A., Novak M. A.. Magnetic Bistability in a Metal-Ion Cluster. Nature. 1993;365(6442):141–143. doi: 10.1038/365141a0. [DOI] [Google Scholar]
- Chiesa A., Guidi T., Carretta S., Ansbro S., Timco G. A., Vitorica-Yrezabal I., Garlatti E., Amoretti G., Winpenny R. E. P., Santini P.. Magnetic Exchange Interactions in the Molecular Nanomagnet Mn 12. Phys. Rev. Lett. 2017;119(21):217202. doi: 10.1103/PhysRevLett.119.217202. [DOI] [PubMed] [Google Scholar]
- Gil Y., Aravena D.. Understanding Single-Molecule Magnet Properties of Lanthanide Complexes from 4f Orbital Splitting. Dalton Transactions. 2024;53(5):2207–2217. doi: 10.1039/D3DT04179D. [DOI] [PubMed] [Google Scholar]
- Gil Y., Castro-Alvarez A., Fuentealba P., Spodine E., Aravena D.. Lanthanide SMMs Based on Belt Macrocycles: Recent Advances and General Trends. Chem. – Eur. J. 2022;28(48):e202200336. doi: 10.1002/chem.202200336. [DOI] [PubMed] [Google Scholar]
- Mironov V. S., Bazhenova T. A., Manakin Yu. V., Yagubskii E. B.. Pentagonal-Bipyramidal 4d and 5d Complexes with Unquenched Orbital Angular Momentum as a Unique Platform for Advanced Single-Molecule Magnets: Current State and Perspectives. Dalton Transactions. 2023;52(3):509–539. doi: 10.1039/D2DT02954E. [DOI] [PubMed] [Google Scholar]
- Guo F.-S., Day B. M., Chen Y.-C., Tong M.-L., Mansikkamäki A., Layfield R. A.. Magnetic Hysteresis up to 80 K in a Dysprosium Metallocene Single-Molecule Magnet. Science. 2018;362(6421):1400–1403. doi: 10.1126/science.aav0652. [DOI] [PubMed] [Google Scholar]
- Goodwin C. A. P., Ortu F., Reta D., Chilton N. F., Mills D. P.. Molecular Magnetic Hysteresis at 60 K in Dysprosocenium. Nature. 2017;548(7668):439–442. doi: 10.1038/nature23447. [DOI] [PubMed] [Google Scholar]
- Georgiev M., Chamati H.. Single-Ion Magnets with Giant Magnetic Anisotropy and Zero-Field Splitting. ACS Omega. 2022;7(47):42664–42673. doi: 10.1021/acsomega.2c06119. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Feng M., Tong M.. Single Ion Magnets from 3d to 5f: Developments and Strategies. Chemistry – A . European Journal. 2018;24(30):7574–7594. doi: 10.1002/chem.201705761. [DOI] [PubMed] [Google Scholar]
- Chicco S., Allodi G., Chiesa A., Garlatti E., Buch C. D., Santini P., De Renzi R., Piligkos S., Carretta S.. Proof-of-Concept Quantum Simulator Based on Molecular Spin Qudits. J. Am. Chem. Soc. 2024;146(1):1053–1061. doi: 10.1021/jacs.3c12008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chiesa A., Macaluso E., Carretta S.. Quantum Information Processing with Molecular Nanomagnets: An Introduction. Contemp Phys. 2023;64:253. doi: 10.1080/00107514.2024.2381952. [DOI] [Google Scholar]
- Santini P., Carretta S., Troiani F., Amoretti G.. Molecular Nanomagnets as Quantum Simulators. Phys. Rev. Lett. 2011;107(23):230502. doi: 10.1103/PhysRevLett.107.230502. [DOI] [PubMed] [Google Scholar]
- Gabarró-Riera G., Sañudo E. C.. Challenges for Exploiting Nanomagnet Properties on Surfaces. Commun. Chem. 2024;7(1):99. doi: 10.1038/s42004-024-01183-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gaita-Ariño A., Luis F., Hill S., Coronado E.. Molecular Spins for Quantum Computation. Nat. Chem. 2019;11:301–309. doi: 10.1038/s41557-019-0232-y. [DOI] [PubMed] [Google Scholar]
- Chiesa A., Santini P., Garlatti E., Luis F., Carretta S.. Molecular Nanomagnets: A Viable Path toward Quantum Information Processing? Rep. Prog. Phys. 2024;87(3):034501. doi: 10.1088/1361-6633/ad1f81. [DOI] [PubMed] [Google Scholar]
- Moreno-Pineda E., Godfrin C., Balestro F., Wernsdorfer W., Ruben M.. Molecular Spin Qudits for Quantum Algorithms. Chem. Soc. Rev. 2018;47(2):501–513. doi: 10.1039/C5CS00933B. [DOI] [PubMed] [Google Scholar]
- Christou G., Gatteschi D., Hendrickson D. N., Sessoli R.. Single-Molecule Magnets. MRS Bull. 2000;25(11):66–71. doi: 10.1557/mrs2000.226. [DOI] [Google Scholar]
- Lunghi A.. Toward Exact Predictions of Spin-Phonon Relaxation Times: An Ab Initio Implementation of Open Quantum Systems Theory. Sci. Adv. 2022;8(31):eabn7880. doi: 10.1126/sciadv.abn7880. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Briganti M., Santanni F., Tesi L., Totti F., Sessoli R., Lunghi A.. A Complete Ab Initio View of Orbach and Raman Spin–Lattice Relaxation in a Dysprosium Coordination Compound. J. Am. Chem. Soc. 2021;143(34):13633–13645. doi: 10.1021/jacs.1c05068. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Reta D., Kragskow J. G. C., Chilton N. F.. Ab Initio Prediction of High-Temperature Magnetic Relaxation Rates in Single-Molecule Magnets. J. Am. Chem. Soc. 2021;143(15):5943–5950. doi: 10.1021/jacs.1c01410. [DOI] [PubMed] [Google Scholar]
- Garlatti E., Chiesa A., Bonfà P., Macaluso E., Onuorah I. J., Parmar V. S., Ding Y.-S., Zheng Y.-Z., Giansiracusa M. J., Reta D., Pavarini E., Guidi T., Mills D. P., Chilton N. F., Winpenny R. E. P., Santini P., Carretta S.. A Cost-Effective Semi-Ab Initio Approach to Model Relaxation in Rare-Earth Single-Molecule Magnets. J. Phys. Chem. Lett. 2021;12(36):8826–8832. doi: 10.1021/acs.jpclett.1c02367. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mezei G., Zaleski C. M., Pecoraro V. L.. Structural and Functional Evolution of Metallacrowns. Chem. Rev. 2007;107(11):4933–5003. doi: 10.1021/cr078200h. [DOI] [PubMed] [Google Scholar]
- Pecoraro V. L., Stemmler A. J., Gibney B. R., Bodwin J. J., Wang H., Kampf J. W., Barwinski A.. Metallacrowns: A New Class of Molecular Recognition Agents. 1996;45:83–177. doi: 10.1002/9780470166468.ch2. [DOI] [Google Scholar]
- Chow C. Y., Trivedi E. R., Pecoraro V., Zaleski C. M.. Heterometallic Mixed 3d - 4f Metallacrowns: Structural Versatility, Luminescence, and Molecular Magnetism. Comments on Inorganic Chemistry. 2015;35(4):214–253. doi: 10.1080/02603594.2014.981811. [DOI] [Google Scholar]
- Lutter J. C., Zaleski C. M., Pecoraro V. L.. Metallacrowns: Supramolecular Constructs With Potential in Extended Solids, Solution-State Dynamics. Molecular Magnetism, and Imaging. 2018;71:177–246. doi: 10.1016/bs.adioch.2017.11.007. [DOI] [Google Scholar]
- Boron, T. T. Magnetic Metallacrowns: From Randomness to Rational Design. In Advances in Metallacrown Chemistry; Springer International Publishing: Cham, 2022; pp 157–219. [Google Scholar]
- Zaleski C. M., Depperman E. C., Kampf J. W., Kirk M. L., Pecoraro V. L.. Synthesis, Structure, and Magnetic Properties of a Large Lanthanide–Transition-Metal Single-Molecule Magnet. Angew. Chem., Int. Ed. 2004;43(30):3912–3914. doi: 10.1002/anie.200454013. [DOI] [PubMed] [Google Scholar]
- Rauguth A., Alhassanat A., Elnaggar H., Athanasopoulou A. A., Luo C., Ryll H., Radu F., Mashoff T., de Groot F. M. F., Rentschler E., Elmers H. J.. Anisotropy of 4f States in 3d–4f Single-Molecule Magnets. Phys. Rev. B. 2022;105(13):134415. doi: 10.1103/PhysRevB.105.134415. [DOI] [Google Scholar]
- Wang J., Lu G., Liu Y., Wu S.-G., Huang G.-Z., Liu J.-L., Tong M.-L.. Building Block and Directional Bonding Approaches for the Synthesis of {DyMn 4 } n (n = 2, 3) Metallacrown Assemblies. Cryst. Growth Des. 2019;19(3):1896–1902. doi: 10.1021/acs.cgd.8b01879. [DOI] [Google Scholar]
- Lou T., Yang H., Zeng S., Li D., Dou J.. A New Family of Heterometallic LnIII[12-MCFeIIIN(Shi)-4] Complexes: Syntheses, Structures and Magnetic Properties. Crystals (Basel) 2018;8(5):229. doi: 10.3390/cryst8050229. [DOI] [Google Scholar]
- Yang W., Yang H., Zeng S.-Y., Li D.-C., Dou J.-M.. Unprecedented Family of Heterometallic Ln III [18-Metallacrown-6] Complexes: Syntheses, Structures, and Magnetic Properties. Dalton Transactions. 2017;46(38):13027–13034. doi: 10.1039/C7DT02735D. [DOI] [PubMed] [Google Scholar]
- Deng W., Wu S., Ruan Z., Gong Y., Du S., Wang H., Chen Y., Zhang W., Liu J., Tong M.. Spin-State Control in Dysprosium(III) Metallacrown Magnets via Thioacetal Modification. Angew. Chem., Int. Ed. 2024;63(31):e20240427. doi: 10.1002/anie.202404271. [DOI] [PubMed] [Google Scholar]
- Wan R.-C., Wu S.-G., Liu J.-L., Jia J.-H., Huang G.-Z., Li Q.-W., Tong M.-L.. Modulation of Slow Magnetic Relaxation for Tb(III)-Metallacrown Complexes by Controlling Axial Halide Coordination. Acta Chimi Sin. 2020;78(5):412. doi: 10.6023/A20030077. [DOI] [Google Scholar]
- Wang J., Ruan Z.-Y., Li Q.-W., Chen Y.-C., Huang G.-Z., Liu J.-L., Reta D., Chilton N. F., Wang Z.-X., Tong M.-L.. Slow Magnetic Relaxation in a {EuCu 5 } Metallacrown. Dalton Transactions. 2019;48(5):1686–1692. doi: 10.1039/C8DT04814B. [DOI] [PubMed] [Google Scholar]
- Wang J., Li Q., Wu S., Chen Y., Wan R., Huang G., Liu Y., Liu J., Reta D., Giansiracusa M. J., Wang Z., Chilton N. F., Tong M.. Opening Magnetic Hysteresis by Axial Ferromagnetic Coupling: From Mono-Decker to Double-Decker Metallacrown. Angew. Chem., Int. Ed. 2021;60(10):5299–5306. doi: 10.1002/anie.202014993. [DOI] [PubMed] [Google Scholar]
- Jankolovits J., Kampf J. W., Pecoraro V. L.. Solvent Dependent Assembly of Lanthanide Metallacrowns Using Building Blocks with Incompatible Symmetry Preferences. Inorg. Chem. 2014;53(14):7534–7546. doi: 10.1021/ic500832u. [DOI] [PubMed] [Google Scholar]
- Melegari M., Marzaroli V., Polisicchio R., Seletti D., Marchiò L., Pecoraro V. L., Tegoni M.. Insights on the Structure in Solution of Paramagnetic Ln III /Ga III 12-Metallacrown-4 Complexes Using 1D 1 H NMR and Model Structures. Inorg. Chem. 2023;62(27):10645–10654. doi: 10.1021/acs.inorgchem.3c00983. [DOI] [PubMed] [Google Scholar]
- Salerno E. V., Eliseeva S. V., Petoud S., Pecoraro V. L.. Tuning White Light Emission Using Single-Component Tetrachroic Dy 3+ Metallacrowns: The Role of Chromophoric Building Blocks. Chem. Sci. 2024;15(21):8019–8030. doi: 10.1039/D4SC00389F. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Eliseeva S. V., Salerno E. V., Lopez Bermudez B. A., Petoud S., Pecoraro V. L.. Dy 3+ White Light Emission Can. Be Finely Controlled by Tuning the First Coordination Sphere of Ga 3+ /Dy 3+ Metallacrown Complexes. J. Am. Chem. Soc. 2020;142(38):16173–16176. doi: 10.1021/jacs.0c07198. [DOI] [PubMed] [Google Scholar]
- Nguyen T. N., Eliseeva S. V., Chow C. Y., Kampf J. W., Petoud S., Pecoraro V. L.. Peculiarities of Crystal Structures and Photophysical Properties of Ga III /Ln III Metallacrowns with a Non-Planar [12-MC-4] Core. Inorg. Chem. Front. 2020;7(7):1553–1563. doi: 10.1039/C9QI01647C. [DOI] [Google Scholar]
- Karns J. P., Eliseeva S. V., Ward C. L., Allen M. J., Petoud S., Lutter J. C.. Near-Infrared Lanthanide-Based Emission from Fused Bis[Ln(III)/Zn(II) 14-Metallacrown-5] Coordination Compounds. Inorg. Chem. 2022;61(15):5691–5695. doi: 10.1021/acs.inorgchem.2c00084. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rauguth A., Kredel A., Carrella L. M., Rentschler E.. 3d/4f Sandwich Complex Based on Metallacrowns. Inorg. Chem. 2021;60(18):14031–14037. doi: 10.1021/acs.inorgchem.1c01356. [DOI] [PubMed] [Google Scholar]
- Zaleski C. M., Tricard S., Depperman E. C., Wernsdorfer W., Mallah T., Kirk M. L., Pecoraro V. L.. Single Molecule Magnet Behavior of a Pentanuclear Mn-Based Metallacrown Complex: Solid State and Solution Magnetic Studies. Inorg. Chem. 2011;50(22):11348–11352. doi: 10.1021/ic2008792. [DOI] [PubMed] [Google Scholar]
- Lah M. S., Pecoraro V. L.. Isolation and Characterization of {MnII[MnIII(Salicylhydroximate)]4(Acetate)2(DMF)6}.Cntdot.2DMF: An Inorganic Analog of M2+(12-Crown-4) J. Am. Chem. Soc. 1989;111(18):7258–7259. doi: 10.1021/ja00200a054. [DOI] [Google Scholar]
- Lewis A. J., Garlatti E., Cugini F., Solzi M., Zeller M., Carretta S., Zaleski C. M.. Slow Magnetic Relaxation of a 12-Metallacrown-4 Complex with a Manganese(III)-Copper(II) Heterometallic Ring Motif. Inorg. Chem. 2020;59(17):11894–11900. doi: 10.1021/acs.inorgchem.0c01410. [DOI] [PubMed] [Google Scholar]
- Azar M. R., Boron T. T., Lutter J. C., Daly C. I., Zegalia K. A., Nimthong R., Ferrence G. M., Zeller M., Kampf J. W., Pecoraro V. L., Zaleski C. M.. Controllable Formation of Heterotrimetallic Coordination Compounds: Systematically Incorporating Lanthanide and Alkali Metal Ions into the Manganese 12-Metallacrown-4 Framework. Inorg. Chem. 2014;53(3):1729–1742. doi: 10.1021/ic402865p. [DOI] [PubMed] [Google Scholar]
- Boron T. T., Lutter J. C., Daly C. I., Chow C. Y., Davis A. H., Nimthong-Roldán A., Zeller M., Kampf J. W., Zaleski C. M., Pecoraro V. L.. The Nature of the Bridging Anion Controls the Single-Molecule Magnetic Properties of DyX 4 M 12-Metallacrown-4 Complexes. Inorg. Chem. 2016;55(20):10597–10607. doi: 10.1021/acs.inorgchem.6b01832. [DOI] [PubMed] [Google Scholar]
- Chow C. Y., Eliseeva S. V., Trivedi E. R., Nguyen T. N., Kampf J. W., Petoud S., Pecoraro V. L.. Ga3+/Ln3+ Metallacrowns: A Promising Family of Highly Luminescent Lanthanide Complexes That Covers Visible and Near-Infrared Domains. J. Am. Chem. Soc. 2016;138(15):5100–5109. doi: 10.1021/jacs.6b00984. [DOI] [PubMed] [Google Scholar]
- Nguyen T. N., Chow C. Y., Eliseeva S. V., Trivedi E. R., Kampf J. W., Martinić I., Petoud S., Pecoraro V. L.. One-Step Assembly of Visible and Near-Infrared Emitting Metallacrown Dimers Using a Bifunctional Linker. Chem. – Eur. J. 2018;24(5):1031–1035. doi: 10.1002/chem.201703911. [DOI] [PubMed] [Google Scholar]
- Travis J. R., Smihosky A. M., Kauffman A. C., Ramstrom S. E., Lewis A. J., Nagy S. G., Rheam R. E., Zeller M., Zaleski C. M.. Syntheses and Crystal Structures of Two Classes of Aluminum-Lanthanide-Sodium Heterotrimetallic 12-Metallacrown-4 Compounds: Individual Molecules and Dimers of Metallacrowns. J. Chem. Crystallogr. 2021;51(3):372–393. doi: 10.1007/s10870-020-00861-2. [DOI] [Google Scholar]
- Eliseeva S. V., Travis J. R., Nagy S. G., Smihosky A. M., Foley C. M., Kauffman A. C., Zaleski C. M., Petoud S.. Visible and Near-Infrared Emitting Heterotrimetallic Lanthanide–Aluminum–Sodium 12-Metallacrown-4 Compounds: Discrete Monomers and Dimers. Dalton Transactions. 2022;51(15):5989–5996. doi: 10.1039/D1DT04277G. [DOI] [PubMed] [Google Scholar]
- Zaleski C. M., Depperman E. C., Kampf J. W., Kirk M. L., Pecoraro V. L.. Using Ln III [15-MC Cu II (N)(S)‑pheHA −5] 3+ Complexes To Construct Chiral Single-Molecule Magnets and Chains of Single-Molecule Magnets. Inorg. Chem. 2006;45(25):10022–10024. doi: 10.1021/ic061326x. [DOI] [PubMed] [Google Scholar]
- Wu S.-G., Ruan Z.-Y., Huang G.-Z., Zheng J.-Y., Vieru V., Taran G., Wang J., Chen Y.-C., Liu J.-L., Ho L. T. A., Chibotaru L. F., Wernsdorfer W., Chen X.-M., Tong M.-L.. Field-Induced Oscillation of Magnetization Blocking Barrier in a Holmium Metallacrown Single-Molecule Magnet. Chem. 2021;7(4):982–992. doi: 10.1016/j.chempr.2020.12.022. [DOI] [Google Scholar]
- Rheam R. E., Zeller M., Zaleski C. M.. Crystal Structures of Three Anionic Lanthanide-Aluminium [3.3.1] Metallacryptate Complexes. Acta Crystallogr. E Crystallogr. Commun. 2020;76:1458–1466. doi: 10.1107/S2056989020010725. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Krause L., Herbst-Irmer R., Sheldrick G. M., Stalke D.. Comparison of Silver and Molybdenum Microfocus X-Ray Sources for Single-Crystal Structure Determination. J. Appl. Crystallogr. 2015;48(1):3–10. doi: 10.1107/S1600576714022985. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sheldrick G. M.. A Short History of SHELX . Acta Crystallogr. A. 2008;64(1):112–122. doi: 10.1107/S0108767307043930. [DOI] [PubMed] [Google Scholar]
- Hübschle C. B., Sheldrick G. M., Dittrich B.. ShelXle: A Qt Graphical User Interface for SHELXL . J. Appl. Crystallogr. 2011;44(6):1281–1284. doi: 10.1107/S0021889811043202. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sheldrick G. M.. Crystal Structure Refinement with SHELXL . Acta Crystallogr. C Struct Chem. 2015;71(1):3–8. doi: 10.1107/S2053229614024218. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Alvarez S., Alemany P., Casanova D., Cirera J., Llunell M., Avnir D.. Shape Maps and Polyhedral Interconversion Paths in Transition Metal Chemistry. Coord. Chem. Rev. 2005;249(17–18):1693–1708. doi: 10.1016/j.ccr.2005.03.031. [DOI] [Google Scholar]
- Casanova D., Cirera J., Llunell M., Alemany P., Avnir D., Alvarez S.. Minimal Distortion Pathways in Polyhedral Rearrangements. J. Am. Chem. Soc. 2004;126(6):1755–1763. doi: 10.1021/ja036479n. [DOI] [PubMed] [Google Scholar]
- Pinsky M., Avnir D.. Continuous Symmetry Measures. 5. The Classical Polyhedra. Inorg. Chem. 1998;37(21):5575–5582. doi: 10.1021/ic9804925. [DOI] [PubMed] [Google Scholar]
- Casanova D., Llunell M., Alemany P., Alvarez S.. The Rich Stereochemistry of Eight-Vertex Polyhedra: A Continuous Shape Measures Study. Chemistry – A . European Journal. 2005;11(5):1479–1494. doi: 10.1002/chem.200400799. [DOI] [PubMed] [Google Scholar]
- Alvarez S., Avnir D., Llunell M., Pinsky M.. Continuous Symmetry Maps and Shape Classification. The Case of Six-Coordinated Metal Compounds. New J. Chem. 2002;26(8):996–1009. doi: 10.1039/b200641n. [DOI] [Google Scholar]
- Travis J. R., Van Trieste G. P. III, Zeller M., Zaleski C. M.. Crystal Structures of Two Dysprosium–Aluminium–Sodium [3.3.1] Metallacryptates That Form Two-Dimensional Sheets. Acta Crystallogr., Sect. E: Crystallogr. Commun. 2020;76(8):1378–1390. doi: 10.1107/S2056989020010130. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lutter J. C., Eliseeva S. V., Kampf J. W., Petoud S., Pecoraro V. L.. A Unique Ln III {[3.3.1]Ga III Metallacryptate} Series That Possesses Properties of Slow Magnetic Relaxation and Visible/Near-Infrared Luminescence. Chem. – Eur. J. 2018;24(42):10773–10783. doi: 10.1002/chem.201801355. [DOI] [PubMed] [Google Scholar]
- Athanasopoulou A. A., Baldoví J. J., Carrella L. M., Rentschler E.. Field-Induced Slow Magnetic Relaxation in the First Dy(Iii)-Centered 12-Metallacrown-4 Double-Decker. Dalton Transactions. 2019;48(41):15381–15385. doi: 10.1039/C9DT02432H. [DOI] [PubMed] [Google Scholar]
- Ruiz-Martínez A., Casanova D., Alvarez S.. Polyhedral Structures with an Odd Number of Vertices: Nine-Atom Clusters and Supramolecular Architectures. Dalton Trans. 2008;(19):2583. doi: 10.1039/b718821h. [DOI] [PubMed] [Google Scholar]
- Ruiz-Martínez A., Casanova D., Alvarez S.. Polyhedral Structures with an Odd Number of Vertices: Nine-Coordinate Metal Compounds. Chemistry – A . European Journal. 2008;14(4):1291–1303. doi: 10.1002/chem.200701137. [DOI] [PubMed] [Google Scholar]
- Lutter J. C., Boron T. T., Chadwick K. E., Davis A. H., Kleinhaus S., Kampf J. W., Zaleski C. M., Pecoraro V. L.. Identification of Slow Magnetic Relaxation and Magnetocoolant Capabilities of Heterobimetallic Lanthanide-Manganese Metallacrown-like Compounds. Polyhedron. 2021;202:115190. doi: 10.1016/j.poly.2021.115190. [DOI] [Google Scholar]
- Zaleski C. M., Kampf J. W., Mallah T., Kirk M. L., Pecoraro V. L.. Assessing the Slow Magnetic Relaxation Behavior of Ln III 4 Mn III 6 Metallacrowns. Inorg. Chem. 2007;46(6):1954–1956. doi: 10.1021/ic0621648. [DOI] [PubMed] [Google Scholar]
- Abragam, A. ; Bleaney, B. . Electron Paramagnetic Resonance of Transition Ions; Oxford Classic Texts in the Physical Sciences; Oxford University Press: 2012. [Google Scholar]
- Chilton N. F., Anderson R. P., Turner L. D., Soncini A., Murray K. S.. PHI: A Powerful New Program for the Analysis of Anisotropic Monomeric and Exchange-coupled Polynuclear d - and f -block Complexes. J. Comput. Chem. 2013;34(13):1164–1175. doi: 10.1002/jcc.23234. [DOI] [PubMed] [Google Scholar]
- Eyring, L. ; Gschneidner, K. A. ; Lander, G. H. . Handbook on the Physics and Chemistry of Rare Earths; Elsevier Science: 2002; Vol. 23. [Google Scholar]
- Ishikawa N., Mizuno Y., Takamatsu S., Ishikawa T., Koshihara S.. Effects of Chemically Induced Contraction of a Coordination Polyhedron on the Dynamical Magnetism of Bis(Phthalocyaninato)Disprosium, a Single-4f-Ionic Single-Molecule Magnet with a Kramers Ground State. Inorg. Chem. 2008;47(22):10217–10219. doi: 10.1021/ic8014892. [DOI] [PubMed] [Google Scholar]
- Lannes A., Luneau D.. New Family of Lanthanide-Based Complexes with Different Scorpionate-Type Ligands: A Rare Case Where Dysprosium and Ytterbium Analogues Display Single-Ion-Magnet Behavior. Inorg. Chem. 2015;54(14):6736–6743. doi: 10.1021/acs.inorgchem.5b00432. [DOI] [PubMed] [Google Scholar]
- Li Q. W., Liu J. L., Jia J. H., Chen Y. C., Liu J., Wang L. F., Tong M. L.. “half-Sandwich” YbIII Single-Ion Magnets with Metallacrowns. Chem. Commun. 2015;51(51):10291–10294. doi: 10.1039/C5CC03389F. [DOI] [PubMed] [Google Scholar]
- Li Z., Deng K., Li D., Kong H., Ruan Z.-Y., Wu S.-G., Tong M.-L.. Luminescence and Magnetic Properties of {Ln III Cd II 16 } Metallacrown Complexes Centered with Compressed Tetragonal Antiprismatic Ln(III) Units. Cryst. Growth Des. 2025;25(8):2580–2587. doi: 10.1021/acs.cgd.5c00084. [DOI] [Google Scholar]
- Li B., Shi W., Du J., Zhang Y., Zhang H., Yang H., Sun L., Zhang Y., Li M.. Structures and Single-Molecule Magnet Behavior of Dy 3 and Dy 4 Clusters Constructed by Different Dysprosium(III) Salts. Inorg. Chem. 2024;63(34):15667–15678. doi: 10.1021/acs.inorgchem.4c01582. [DOI] [PubMed] [Google Scholar]
- Chiesa A., Cugini F., Hussain R., Macaluso E., Allodi G., Garlatti E., Giansiracusa M., Goodwin C. A. P., Ortu F., Reta D., Skelton J. M., Guidi T., Santini P., Solzi M., De Renzi R., Mills D. P., Chilton N. F., Carretta S.. Understanding Magnetic Relaxation in Single-Ion Magnets with High Blocking Temperature. Phys. Rev. B. 2020;101(17):174402. doi: 10.1103/PhysRevB.101.174402. [DOI] [Google Scholar]
- Sran B. S., Gonzalez J. F., Montigaud V., Le Guennic B., Pointillart F., Cador O., Hundal G.. Structural and Magnetic Investigations of a Binuclear Coordination Compound of Dysprosium(iii) Dinitrobenzoate. Dalton Trans. 2019;48(12):3922–3929. doi: 10.1039/C8DT04253E. [DOI] [PubMed] [Google Scholar]
- Rinehart J. D., Long J. R.. Exploiting Single-Ion Anisotropy in the Design of f-Element Single-Molecule Magnets. Chem. Sci. 2011;2(11):2078. doi: 10.1039/c1sc00513h. [DOI] [Google Scholar]
- Randall McClain K., Gould C. A., Chakarawet K., Teat S. J., Groshens T. J., Long J. R., Harvey B. G.. High-Temperature Magnetic Blocking and Magneto-Structural Correlations in a Series of Dysprosium(iii) Metallocenium Single-Molecule Magnets. Chem. Sci. 2018;9(45):8492–8503. doi: 10.1039/C8SC03907K. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cole K. S., Cole R. H.. Dispersion and Absorption in Dielectrics I. Alternating Current Characteristics. J. Chem. Phys. 1941;9(4):341–351. doi: 10.1063/1.1750906. [DOI] [Google Scholar]
- Shao F., Cahier B., Guihéry N., Rivière E., Guillot R., Barra A.-L., Lan Y., Wernsdorfer W., Campbell V. E., Mallah T.. Tuning the Ising-Type Anisotropy in Trigonal Bipyramidal Co(ii) Complexes. Chem. Commun. 2015;51(92):16475–16478. doi: 10.1039/C5CC07741A. [DOI] [PubMed] [Google Scholar]
- Rechkemmer Y., Breitgoff F. D., van der Meer M., Atanasov M., Hakl M., Orlita M., Neugebauer P., Neese F., Sarkar B., van Slageren J.. A Four-Coordinate Cobalt(II) Single-Ion Magnet with Coercivity and a Very High Energy Barrier. Nat. Commun. 2016;7(1):10467. doi: 10.1038/ncomms10467. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Novikov V. V., Pavlov A. A., Nelyubina Y. V., Boulon M.-E., Varzatskii O. A., Voloshin Y. Z., Winpenny R. E. P.. A Trigonal Prismatic Mononuclear Cobalt(II) Complex Showing Single-Molecule Magnet Behavior. J. Am. Chem. Soc. 2015;137(31):9792–9795. doi: 10.1021/jacs.5b05739. [DOI] [PubMed] [Google Scholar]
- Jiang X. F., Chen M. G., Tong J. P., Shao F.. A Mononuclear Dysprosium(Iii) Single-Molecule Magnet with a Non-Planar Metallacrown. New J. Chem. 2019;43(22):8704–8710. doi: 10.1039/C9NJ01662G. [DOI] [Google Scholar]
- Shrivastava K. N.. Theory of Spin–Lattice Relaxation. physica status solidi (b) 1983;117(2):437–458. doi: 10.1002/pssb.2221170202. [DOI] [Google Scholar]
- Singh A., Shrivastava K. N.. Optical-acoustic Two-phonon Relaxation in Spin Systems. physica status solidi (b) 1979;95(1):273–277. doi: 10.1002/pssb.2220950131. [DOI] [Google Scholar]
- Chow C. Y., Bolvin H., Campbell V. E., Guillot R., Kampf J. W., Wernsdorfer W., Gendron F., Autschbach J., Pecoraro V. L., Mallah T.. Assessing the Exchange Coupling in Binuclear Lanthanide(iii) Complexes and the Slow Relaxation of the Magnetization in the Antiferromagnetically Coupled Dy2 Derivative. Chem. Sci. 2015;6(7):4148–4159. doi: 10.1039/C5SC01029B. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
