Next Article in Journal
R-Phase Transformation Evolution in NiTi SMA Wires Studied via the Internal Friction Technique
Previous Article in Journal
Reactive Magnetron Sputtering for Y-Doped Barium Zirconate Electrolyte Deposition in a Complete Protonic Ceramic Fuel Cell
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Chiral 4f and 3d-4f Complexes from Enantiopure Salen-Type Schiff Base Ligands

by
Catherine P. Raptopoulou
Institute of Nanoscience and Nanotechnology, NCSR “Demokritos”, 15310 Aghia Paraskevi Attikis, Greece
Crystals 2024, 14(5), 474; https://doi.org/10.3390/cryst14050474
Submission received: 11 April 2024 / Revised: 13 May 2024 / Accepted: 16 May 2024 / Published: 18 May 2024
(This article belongs to the Section Crystal Engineering)

Abstract

:
This review summarizes the structural characteristics and physicochemical properties of chiral 4f and 3d-4f complexes based on enantiopure salen-type Schiff base ligands. The chirality originates from the enantiopure diamines and is imparted to the Schiff base ligands and complexes and finally to the crystal structures. The reported enantiopure Schiff base ligands derive from the condensation of aromatic aldehydes, such as salicylaldehyde and its various derivatives, and the enantiopure diamines, (1R,2R) or (1S,2S)-1,2-diamino-cyclohexane, (1R,2R) or (1S,2S)-1,2-diamino-1,2-diphenylethane, (R) or (S)-2,2′-diamino-1,1′-binaphthalene, and 1,2-diaminopropane.

1. Introduction

Molecular materials based on metal coordination complexes exhibit versatile electronic and structural features, which arise from the metal ions and the different types and numbers of ligands. These materials are usually prepared by wet chemistry methods, in some cases under environmentally friendly processes, and most importantly, combine different characteristics, such as metal oxidation states and versatile coordination geometries, with redox, catalytic, magnetic, electrical, and optical properties. The physicochemical properties of these materials depend on their exact structure and can be totally altered by small structural changes. Molecular materials with nonlinear optical, molecular magnetic, conducting, and superconducting properties have been proposed for potential applications [1,2]. Over the last decade, the field of molecular materials that combine different physical properties and are subjected to various external stimuli [3] has grown very rapidly. The rational design of these multifunctional molecular materials is of prime importance because of their potential applications in information storage, sensors, spintronics, biology, etc. [4].
Chirality is an intrinsic property of matter and is important in chemistry, biology, and physics, as it can affect the properties and reactivity of molecules. Chiral objects or molecules are not superimposable in their mirror image. The distinct left and right-handed forms are called enantiomers and exhibit different properties and activities. Eminent examples of chiral molecules are amino acids, sugars, and, therefore, proteins and nucleic acids, as well as certain pharmaceutical drugs. Metal coordination complexes can exhibit two types of chirality: (a) chirality at metal, which arises by the certain arrangement of ligands around the metal ion and disappears upon the separation of the metal and the ligands, and (b) chirality, which is imparted by the presence of chiral ligands around the metal ions [5]. Examples of metal complexes that belong in type (a) are the tetrahedral metal complexes with four different ligands, the octahedral tris-bidentate metal complexes, the octahedral bis-bidentate-bis-monodentate metal complexes, the octahedral metal complexes with three different monodentate ligands (MA2B2C2), the octahedral metal complexes with four different ligands in the fac-arrangement (MABCL3), and the octahedral metal complexes with linear tetradentate ligands in the cis-isomer (MD4L2). These chiral complexes are referred to as chiral-at-metal or ‘stereogenic-at-metal’ complexes [6]. From the synthetic point of view, the use of enantiopure ligands for the enantioselective synthesis of chiral coordination complexes is the most efficient method [7].
Chiral metal complexes are important in modern medicine as potential therapeutics (metallodrugs) [8] and can be selectively activated at the target site of a biomolecule by an external stimulus. These compounds can act as ‘prodrugs’ that interact with the biomolecules via various mechanisms such as redox processes, metal coordination, photo- and thermal-activation, and radiation [9]. Besides their importance in the pharmaceutical industry, chiral metal complexes, which combine chirality with magnetic, optical, and electrical properties, have received great attention in recent years and constitute a new and fast-growing class of multifunctional molecular materials. Coordination complexes that combine chirality and single-molecule magnet behavior, and in addition, chirality-induced properties, e.g., ferroelectricity [10,11], second-order nonlinear optical properties [12,13], magnetochiral dichroism [14,15], magneto-optical Faraday effect [16,17], and circularly polarized luminescence [18,19], represent a special class of multifunctional molecular materials.
Lanthanide(III) ions display large spin ground states and magnetic anisotropy, therefore, have been widely used for the synthesis of homo- and hetero-metallic complexes which show single-molecule magnet (SMM) [20,21,22], and single-ion magnet (SIM) [23,24,25,26,27,28] behavior. Schiff bases are formed by the condensation reaction between a primary amine and an aldehyde or ketone. These compounds are widely used in organic synthesis and coordination chemistry as ligands for metal complexes. They are also important in the production of various pharmaceuticals, dyes, and pesticides, in the food and agrochemical industry, in catalysis and energy storage, and for environmental and biomedical applications [29,30,31]. Schiff base ligands can be easily synthesized by various combinations of carbonyl compounds and primary amines and can be easily functionalized by choosing starting materials with the desired side groups. The combination of chiral Schiff base ligands with lanthanide(III) ions has been an effective strategy for the development of multifunctional molecular materials. The present minireview summarizes the chiral 4f and 3d-4f coordination complexes with enantiopure Schiff base ligands derived from the condensation of aromatic aldehydes, such as salicylaldehyde and its various derivatives, and the enantiopure diamines, (1R,2R) or (1S,2S)-1,2-diamino-cyclohexane, (1R,2R) or (1S,2S)-1,2-diamino-1,2-diphenylethane, (R) or (S)-2,2′-diamino-1,1′-binaphthalene, and 1,2-diaminopropane. The complexes reported herein were published up to December 2023. The Schiff base ligands discussed herein are depicted in Scheme 1, Scheme 2 and Scheme 3.

2. Homometallic 4f Complexes

Structural, CD, and magnetic data for the homometallic 4f complexes are listed in Table 1.
A pair of enantiopure complexes (Et3NH)[DyIII(L4)2] (1R/1S) were reported which contain ligand H2L4RR,SS, N,N’-(1,2-cyclohexanediylethylene)bis(3-nitrosalicylideneiminato), Scheme 1 [32]. The asymmetric unit of complex 1R contains two independent [DyIII(L4RR)2] anions and four (Et3NH)+ cations. Each DyIII ion is eight-coordinated to two Ophenoxo and two Nimino atoms from two (L4RR)2− ligands with a triangular dodecahedron (TDD-8) geometry (CShM (Continuous Shape Measures) = 0.82632 (Dy1) and 1.14706 (Dy2)), Figure 1. The CD spectra of both 1R/1S in the MeCN solution display mirror image patterns as expected for a pair of enantiomers. The magnetic properties of mononuclear complexes in which the LnIII ion is encapsulated by two salen-type ligands are reported for the first time. Both complexes exhibit field-induced SIM behavior combined with chirality. Dual magnetic relaxations are observed under a weak DC field as a result of both single-ion anisotropy and a direct process due to intermolecular interactions. Under a higher dc field, a single thermally activated Orbach relaxation is observed due to single-ion anisotropy. Thus, a field-induced switch from dual relaxation to a single relaxation process is achieved.
Pairs of enantiomerically pure chiral complexes with H2L5RR,SS were reported with formula [(LOEt)Ln(L5)] (LnIII = Dy (2R/2S, Tb (3R/3S), Ho (4R/4S); LOEt = [(Cp)Co(P(O)(OEt)2)3]) which crystallize with two independent molecules in the asymmetric unit [33]. Both LnIII ions are seven-coordinated to the two Ophenoxo and two Nimino of (L5)2− and to three O-atoms from LOEt in distorted monocapped triangular prism geometry. The enantiomeric nature of all complexes was confirmed by CD spectra. The temperature dependence of the magnetic susceptibility for 2R-4R shows a gradual decrease of the χMT product in the range 300–100 K and a further sharp decrease at 2 K, which can be attributed to the progressive depopulation of the Stark levels and/or possible antiferromagnetic interactions between the metal ions. The magnetization of 2R-4R at 2 K does not reach saturation at 70 kOe due to the ligand-field-induced splitting of the Stark levels and to the magnetic anisotropy. Complex 2R shows field-induced slow relaxation of magnetization, which follows a thermally activated mechanism. The energy barrier and pre-exponential factor τ0 (Table 1) were calculated by the Arrhenius law, τ = τ 0 e x p ( Δ k B T ) . Complex 2R represents a rare example of a seven-coordinated chiral SIM. Complexes 2R-4R show second-order nonlinear optical (NLO) effects with responses of 0.3 times that of urea. Unfortunately, no ferroelectric properties were observed for 2R-4R at room temperature.
The CeIV ion in complex [CeIV(L7RR)2] (5R) is eight-coordinated to the two Ophenoxo and the two Nimino atoms from two (L7RR)2− ligands in distorted square antiprism geometry [34]. The dihedral angle between the best mean planes of the two ligands, excluding the C-atoms of the cyclohexane ring, is 47.1°, and this is a unique architecture for CeIV complexes with two Schiff base ligands (Figure 2). The Ce-O and Ce-N bond distances are 2.212(5)-2.241(5) Å and 2.550(7)-2.570(6) Å, respectively. The complex was studied by CD, 1H NMR, 13C NMR, and NOE (nuclear Overhauser effect) studies; the latter showed that the azomethine H-atom is in proximity to the phenyl ring H-atom and the cyclohexane CH2. The spectral analysis suggests CeIV ion in a distorted planarity of the N2O2 chromophore group.
The same ligand gave complex (cation)[DyIII(L7RR)2] (6R) [35]. The DyIII ion is eight-coordinated to two Ophenoxo and two Nimino atoms from two (L7RR)2− ligands and displays square antiprism geometry with Δ absolute configuration. The countercation was formed in situ from the decomposition of H2L7RR with Dy(NO3)3 acting as a Lewis acid catalyst. Single crystal samples of 6R at r.t. show ferroelectric behavior with remnant polarization Pr~4.51 μCcm−2 and Ec~28.11 kVcm−1 and characteristic emission 4F9/26H15/2 and 4F9/26H13/2 transitions of DyIII at the solid luminescent spectrum (Figure 3). The enantiopure ligand H2L7SS (Scheme 1) gave complex (cation)[ErIII(L7SS)2] (7S), which is isomorphous to complex 6R, but in enantiomeric configuration [36]. The ErIII ion shows an eight-coordinate square antiprism geometry with Λ absolute configuration due to the S,S-enantiomer of the diamine used. Single crystal samples of 7S at r.t. show ferroelectric behavior with remnant polarization Pr~7.98 μCcm−2 and Ec~22.12 kVcm−1 at electric field 24 kVcm−1. The value of Pr is close to that of Ps~8.04 μCcm−2, which is much larger than that of other ferroelectrics, such as KH2PO4, triglycine sulfate, and NaKC4H4O6·4H2O. Both H2L7SS and 7S show second harmonic generation (SHG) efficiency, 0.6 and 1.2 times that of urea, respectively, thus confirming their potential as second-order NLO materials.
A pair of enantiopure complexes, namely [(LOEt)Dy(L9)] (8R/8S) was also reported (LOEt = [(Cp)Co(P(O)(OEt)2)3]) [33]. In both complexes, the DyIII ions are seven-coordinated to the (OphenoxoNimino)2 donor set of (L9)2− and to three O-atoms from LOEt ligand, with distorted monocapped triangular prism geometry (Figure 4). The CD spectra of 8R/8S in KBr pellets confirm the enantiomeric nature of both compounds (Figure 5, Table 1). The complex exhibits field-induced SIM behavior. The relaxation follows a thermally activated mechanism (Table 1). Complex 8R is similar to complex 2R-4R and displays similar second-order NLO effects, with no ferroelectric behavior at r.t.
The zwitterion of enantiopure ligand H2L10SS (Scheme 1) gave complex [SmIII(H2L10SS)(NO3)3] (9S), which crystallizes with two independent molecules in the asymmetric unit [37]. The SmIII ion is ten-coordinated to three chelate nitrato groups and to the two Ophenoxo and two Omethoxo atoms of the zwitterion H2L10SS. Both Nimino atoms are protonated and form intramolecular N-H···Ophenoxo hydrogen bonds. Hirshfeld surface analysis shows that the crystal packing is dominated by H···H, O···H and C···H contacts.
The enantiopure ligand H2L12RR (Scheme 1) gave the chiral 2D coordination polymer [CeIII2(H2L12RR)6(NO3)6(H2O)]n (10R) [38]. Ce(2) is nine-coordinated to three chelate nitrato groups and three Ophenoxo from three different H2L12RR ligands. Ce(1) is ten-coordinated, exhibiting the same coordination environment as Ce(2) and also an additional water molecule. Complex 10R shows 2D square grid network architecture. Both H2L12RR and complex 10R were tested on MDA-MB-231 breast cancer cells by MTT assay and showed antiproliferative activity, which is higher for complex 10R compared with H2L12RR.
Figure 5. The molecular structure of one of the independent molecules in 17R (CCDC-1046379, left) and 20R (CCDC-1046384, right) [39]. Color code: Dy pink, Yb tan, O red, N blue, C grey.
Figure 5. The molecular structure of one of the independent molecules in 17R (CCDC-1046379, left) and 20R (CCDC-1046384, right) [39]. Color code: Dy pink, Yb tan, O red, N blue, C grey.
Crystals 14 00474 g005

3. Heterometallic 3d-4f Complexes

The structural, CD, and magnetic data for the heterometallic 3d-4f complexes are summarized in Table 2, Table 3 and Table 4. Other spectroscopic data are listed in Table 5.
The enantiopure ligand H2L2RR (Scheme 1) gave a family of CuLn dinuclear complexes, [CuLn(L2RR)(NO3)3(H2O)] (LnIII = Ce (11R), Nd (12R)), and [CuLn(L2RR)(NO3)3] (Ln = Sm (13R), Eu (14R), Gd (15R), Tb (16R), Dy (17R), Ho (18R), Er (19R) and Yb (20R)), which crystallize with two independent dimers in the asymmetric unit (Figure 5) [39]. Temperature-controlled reversible conversion of one chiral crystal (15R-17R) to another chiral crystal (15′R-17′R) was observed, which leads to polymorphism and space group transformation from P1 to P21. The metal ions are linked through two Ophenoxo bridges. The CuII ion in 11R-17R is coordinated to the two Ophenoxo and the two Nimino atoms of (L2RR)2− in a square planar geometry, whereas in 15′R-17′R and 18R-20R a weak bond to one Onitrato leads to square pyramidal geometry. The LnIII ion in 13R-17R is ten-coordinated to two bridging Ophenoxo, two Omethoxo, and three chelate nitrato groups, whereas, in 11R and 12R, the LnIII ion is eleven-coordinated due to an additional coordinated water molecule. In 15′R-17′R and 18R-20R, one LnIII ion is ten-coordinated as in 13R-17R, and the second LnIII ion is nine-coordinated because one of the nitrato groups is monodentate. The Cu···Ln distances are in the range 3.2287(7)-3.4382(13) Å in going from 20R to 11R. The CD spectra of the aforementioned complexes confirmed their optical activity and enantiomeric nature. Complexes 15R-17R and 15′R-17′R show ferromagnetic coupling between the metal ions. The TbIII, DyIII, and HoIII compounds, 16R/16′R, 17R/17′R, and 18R display field-induced magnetic relaxation phenomena (Table 3). The magnetic difference between these complexes results from the lanthanide contraction effect of the LnIII ions. Complex 15′R and [CuLu(L2RR)(NO3)3] (21R) were previously reported [40].
Complexes [NiLn(L2RR)(NO3)3(H2O)] (Ln = Ce (22R), Nd (23R)) and [NiLn(L2RR)(NO3)3] (Ln = Nd (24R), Sm (25R), Eu (26R), Gd (27R), Tb (28R), Dy (29R), Yb (30R), and Lu (31R)) are isomorphous to abovementioned complexes 11R-14R, 15′R-17′R, 20R and 21R [40,41,42]. Complexes 28R and 29R exhibit field-induced SMM behavior due to the strong anisotropy and crystal field effect of the TbIII or DyIII ions (Figure 6, Table 2).
Complex [ZnLu(L2RR)(NO3)2(CH3CO2)] (32R) consists of a five-coordinate ZnII ion with square pyramidal geometry defined by the two Ophenoxo and the two Nimino atoms of the ligand and one O-atom from the μ2-CH3CO2, and a nine-coordinate LuIII ion bound to two bridging Ophenoxo, two Omethoxo, two chelate nitrato groups and one O-atom from the μ2-CH3CO2 [40]. The enantiomeric nature of all the abovementioned complexes was confirmed by solid-state CD spectroscopy.
Figure 6. Temperature dependence of the χac at different frequencies with Hdc = 2 kOe and Hac = 2.5 kOe and the least-squares fit of the experimental data to the Arrhenius equation for 28R (a) and 29R (b) [41]. Copyright © 2015 Elsevier B.V. All rights reserved.
Figure 6. Temperature dependence of the χac at different frequencies with Hdc = 2 kOe and Hac = 2.5 kOe and the least-squares fit of the experimental data to the Arrhenius equation for 28R (a) and 29R (b) [41]. Copyright © 2015 Elsevier B.V. All rights reserved.
Crystals 14 00474 g006
A pair of enantiomer tetranuclear complexes [Cu2Tb2(L2)(N3)6(MeOH)2] (33R/33S) crystallize with one cluster in the asymmetric unit (Figure 7) [43]. The CuTb subunits are related to each other through a pseudo inversion center and are bridged by two end-on azides. The two CuII ions are octahedral and coordinated to the two Ophenoxo and the two Nimino atoms of (L2)2−, one μ1,3-N3, and one MeOH. The two TbIII ions are eight-coordinated to the two bridging Ophenoxo and the two Omethoxo of (L2)2−, one μ1,3-N3, one terminal N3 and two μ1,1-N3 in bicapped trigonal prism geometry. The ac susceptibility studies under zero dc field showed slow relaxation of magnetization (Table 3). The SMM behavior was confirmed by the hysteresis loops observed on field-oriented single crystals of 33R below 4 K (Figure 8). This is the first time that angular resolved magnetometry measurements and ab initio calculations were performed on polynuclear SMMs in order to determine the magnetic anisotropy axis of each individual TbIII ion.
A family of Zn2Dy chiral SIMs were reported by using enantiopure H2L2RR and H2L2SS ligands, complexes [Zn2Dy(L2RR)2(X)2(H2O)](anion) (X = Cl, anion = ClO4 (34R); X = Cl, anion = CF3SO3 (35R); X = Br, anion = CF3SO3 (36R)) and [Zn2Dy(L2SS)2(X)2(H2O)][Zn2Dy(L2SS)2(X)2(MeOH)](anion)2 (X = Cl, anion = ClO4 (37S); X = Cl, anion = CF3SO3 (38S); X = Br, anion = CF3SO3 (39S)) [44]. All complexes are isomorphous. The asymmetric unit of 34R-36R contains two independent molecules, whereas, in 37S-39S, the two molecules differ in the coordinated solvent molecule. Each ZnII ion is five-coordinated to the two Ophenoxo and the two Nimino atoms of (L2)2− and to the X ligand, which occupies the apex of the square pyramid. Each DyIII ion is nine-coordinated to the two bridging Ophenoxo and the two Omethoxo atoms of each (L2)2− and one solvent molecule in muffin (MFF-9) geometry (Dy1, Dy2 (34R, 37S), Dy1 (35R, 36R, 38S, 39S)) and spherical tricapped trigonal prism, TTP-9 (Dy2 (35R, 36R, 38S, 39S)). The molecular structure of complex 34R is given in Figure 9. Solid-state CD spectra of all complexes confirmed their chiral nature and optical activity (Table 4). Both the coordinated anions and the counterions, as well as the coordinated solvent molecules, have important effects on the SIM behavior and on the second-harmonic generation and third-harmonic generation NLO properties of the chiral complexes (Table 5). All complexes display field-induced magnetic relaxation as indicated by the peaks in the χ″ vs. T curves. The data in the lnτ vs. 1/T plots are not linear, which suggests that the relaxation has an Orbach process in addition to the Raman process. The curves were fitted with the equation τ 1 = C T n + τ 0 1 e x p ( U e f f k T ) to consider both processes and give the values listed in Table 4. Theoretical calculation further confirmed the experimental data. The study of the NLO properties showed that the SHG signal intensity of the 37S-39S complexes is much stronger than that of the 34R-36R because of the MeOH coordination in the former (Figure 10). Moreover, the bromide coordination and the perchlorate counteranion favor the SHG properties between the 37S-39S complexes. The third harmonic generation (THG) signal intensity of the 34R-36R complexes is much stronger than that of the 37S-39S, which suggests that the coordination of the H2O molecule favors the THG properties. Moreover, the triflate counteranion favors the THG properties between the 34R-36R complexes.
Enantiopure ligands H2L2RR and H2L2SS gave a family of 3d-3d′-4f chiral complexes, [(Tp*)Fe(CN)3Cu(L2)Ln(NO3)3(H2O)3][(Tp*)Fe(CN)3] (LnIII = Gd (40R/40S), Tb (41R/41S), Dy (42R/42S); Tp* = hydridotris(3,5-dimethylpyrazol-1-yl)borate) which are isomorphous (Figure 11, Table 3) [45]. The FeIII ions are six-coordinated to the three C-atoms from CN and the three N-atoms from pyrazolyl groups in a distorted octahedral geometry. The CuII ions are five-coordinated to the two Ophenoxo and the two Nimino atoms of (L2)2− and the bridging cyanide group in the apex of a square pyramid. The LnIII ions are nine-coordinated to the two bridging Ophenoxo and the two Omethoxo of (L2)2−, one chelate nitrate, and three aqua ligands in a geometry between spherical capped square antiprism (CSAPR-9) and capped square antiprism J10 (JCSAPR-9). CD spectra confirm the enantiomeric nature of the complexes. Magnetic studies revealed ferromagnetic interactions between the metal ions without signs of SMM behavior under zero or 1 kOe dc field. This indicates fast QTM relaxation as a result of the observed distorted coordination symmetry, which disfavors the axial anisotropy of the TbIII and DyIII ions.
The enantiopure ligand H2L3RR (Scheme 1) gave the chiral complex [CuEu(L3RR)(NO3)3] (43R), which crystallizes with two independent molecules in the asymmetric unit (Table 3). The CuII ion is square planar, bound to the (OphenoxoNimino)2 donor set of (L3RR)2−, and the EuIII is ten-coordinated bound to the (OphenoxoOethoxo)2 donor set of (L3RR)2− and to three chelate nitrato groups [46]. The chiral 1D zigzag chains of [CuGd(L3)(NO3)3]n (44R/44S) were also reported (Table 3). The polymer structure is promoted by the semicoordination of nitrate to CuII. Magnetic studies showed strong ferromagnetic coupling between the metal ions [47].
The enantiopure ligand H2L6RR (Scheme 1) was synthesized by the condensation reaction of 4-formyl-3-hydroxybenzoic acid and (1R,2R)-(-)-1,2-diaminocyclohexane. The ligand gave four chiral 1D coordination polymers, [Mn2Ln2(L6RR)2Cl2(NO3)2(dmf)6(H2O)2]n (LnIII = Pr (45R), Nd (46R), Sm (47R), Gd (48R)), (Table 2) [48]. The MnIII ion is coordinated to the two Ophenoxo and the two Nimino of (L6RR)2−, one chlorido, and one monodentate nitrato group in a distorted octahedral geometry. Each carboxylato group of (L6RR)2− links two LnIII ions in μ2-fashion, thus forming a distorted square paddle-wheel unit LnIII2. Each LnIII ion is eight-coordinated to four Ocarboxylato, three dmf, and one H2O molecules (Figure 12). The activity of 45R-48R as heterogeneous enantioselective catalysts for the sulfoxidation reaction of various alkyl and aryl sulfides revealed that 45R-47R show comparable activity for the oxidation of methyl p-tolylsulfide, whereas 48R displays maximum conversion of 100% methyl phenylsulfoxide with 88% selectivity after 16 h and could be reused three times without loss of activity.
The enantiopure ligands H2L8RR and H2L8SS (Scheme 1) gave pairs of chiral complexes [ZnLn(L8)(N(SiHMe2)2)(AABP)] (LnIII = Y (49R/49S), Lu (50R/50S), Dy (51R/51S), Sm (52R/52S), La (53R/53S); AABP = alkoxy-amino-bis(phenolato)), (Table 4) [49]. The ZnII ion is five-coordinate to the two Ophenoxo and the two Nimino atoms of (L8)2− and to the nitrogen of N(SiHMe2)2 in distorted square pyramidal geometry. The LnIII ion is six-coordinated to the two bridging Ophenoxo atoms of (L8)2− and to the three O- and the N-atoms of AABP in distorted octahedral geometry (Figure 13). Complexes 51R/51S and 52R/52S are highly efficient catalysts for copolymerization of CO2 with cyclohexene oxide. The transformation of CO2 to other chemicals is very important, both economically and environmentally, and can be achieved via the ring-opening copolymerization of CO2 with epoxide to produce polycarbonates. The reaction requires selectivity in order to avoid other byproducts. It was shown that the ionic radii of the lanthanide ions influence the carbonate linkage, thus leading to excellent selectivity for the isolation of polymers only.
Table 2. Structural, CD, and magnetic data for the heterometallic Mn/4f and Ni/4f complexes.
Table 2. Structural, CD, and magnetic data for the heterometallic Mn/4f and Ni/4f complexes.
Complex3d/4f IonsSpace GroupCoordination NumberCD, λmax
(nm)
χMT (r.t.) (cm3Kmol−1)Ueff (K)t0 (s)Ref.
22RNi/CeP111−280, −400
+558 a
0.86--[41]
23RNi/NdP111−280, −400
+558 a
2.36--[41]
24RNi/NdP110----[42]
25RNi/SmP110−280, −400
+558 a
0.18--[41]
26RNi/EuP110−280, −400
+558 a
1.05--[41]
27RNi/GdP110−280, −400
+558 a
8.12--[41]
28RNi/TbP110−280, −400
+558 a
12.2629.12 b3.21 × 10−9[41]
29RNi/DyP110−280, −400
+558 a
14.8718.40 b7.39 × 10−6[41]
30RNi/YbP110−280, −400
+558 a
1.46--[41]
31RNi/LuP110----[40]
54RNi/CeC210--8.5 c7.7 × 10−5[50]
55RNi/NdC210--9.2 b1.9 × 10−5[50]
56R/56SNi/EuC210−325, +290, +370, +410 a---[50]
57RNi/DyC210--9.3 b2.1 × 10−5[50]
58RNi/ErC210--18.4 b1.7 × 10−6[50]
59SNi/YbC210--18.1 b2.1 × 10−6[50]
45RMn/PrC28----[48]
46RMn/NdC28----[48]
47RMn/SmC28----[48]
48RMn/GdC28----[48]
a the values correspond to the R-enantiomer, same values with opposite signs correspond to the S-enantiomer; b under 1.0 kOe external dc field; c under 0.5 kOe external dc field.
The pair of enantiomers, ligands H2L10RR and H2L10SS (Scheme 1), gave the chiral complexes [NiLn(L10)(NO3)3] (LnIII = Ce (54R), Nd (55R), Eu (56R/56S), Dy (57R), Er (58R), Yb (59S)) and [ZnLn(L10)(NO3)3(MeOH)] (LnIII = Ce (60R), Nd (61R), Eu (62R/62S), Dy (63S), Er (64S), Yb (65S)) [50]. Both complexes 56R and 56S crystallize with two independent molecules in the asymmetric unit (Table 2). The NiII ion is bound to the two Ophenoxo and the two Nimino atoms of (L10)2− in a square planar geometry, and the EuIII ion is ten-coordinated to the two bridging Ophenoxo and the two Omethoxo of (L10)2− and to three bidentate nitrato groups with sphenocorona geometry (CShM = 3.24). Complexes 62R and 62S crystallize with two independent molecules in the asymmetric unit (Table 4). Complexes 62R/62S are similar to 56R/56S, with the only difference being that the ZnII ion in the former is bound to a MeOH molecule in the apex of a square pyramid. The solution CD spectra of the enantiomer pairs display mirror image patterns as expected (Figure 14). Complexes 54R, 55R, 59S, 60R, and 61R, i.e., complexes with LnIII = Ce, Nd, and Yb, show slow relaxation of magnetization under an externally applied dc field, which is applied in order to avoid quantum tunneling of magnetization (QTM), and represent scarce examples of molecules with SIM behavior from ‘uncommon magnetic lanthanides’. The fitting of the data by using the generalized Debye model according to the equation ln χ M χ M = ln ω τ 0 U e f f / ( k B T ) gave the values listed in Table 2 and Table 4. Additionally, complexes with LnIII = Dy and Er, i.e., 57R, 58R and 63S show field-induced SIM behavior, whereas ZnEr complex 64S does not show signs of slow relaxation of the magnetization. The Ueff0) values for 57R were calculated according to the generalized Debye model (Table 2). For 58R and 63S, the data were fitted according to the Arrhenius equation ln 1 2 π ω = ln 1 τ 0 U e f f / ( k B T ) , which gave the values listed in Table 2 and Table 4, respectively.
The pair of enantiomers [ZnDy(L10)(NO3)3(H2O)] 66R/66S are isomorphous and crystallize with two independent molecules in the asymmetric unit (Table 4) [51]. The molecular structure of both complexes is analogous to that of the 63S above. The molecules of 66R are linked through hydrogen bonds between the coordinated H2O molecule and one of the nitrato O-atoms. Ac susceptibility measurements for 66R under 4 kOe dc field showed two maxima at the χ″ vs. T plots, which are consistent with two relaxation processes following the Orbach pathway. The data were fitted to the Arrhenius equation for the low- and high-temperature relaxations (Table 4). The complexes show chiroptical activity and circularly polarized luminescence (CPL) in dmf solution, and also magnetic circular dichroism (MCD) signals at r.t. in the range 280–600 nm with gmax(MCD)~6.7 × 10−2 T−1 attributed to the large orbital angular momentum of aromatic π-conjugated systems (Figure 15). These results suggest that the complexes may have potential applications in magnetochiral devices. The CPL spectra of both complexes show mirror images in the range 430–600 nm, with the gPL for the two enantiomers being ±2.5 × 10−3 (Figure 16).
Complexes [ZnLn(L10)(NO3)2(O2CMe)] (LnIII = Nd (67S), Dy (68R/68S), Yb (69R/69S)) were also reported (Table 4) [52,53,54]. In all complexes, the ZnII ion is located in the (OphenoxoNimino)2 pocket of (L10)2− and the LnIII ion is coordinated to the (OphenoxoOmethoxo)2 donor set and to two chelate nitrato groups. In addition, a bridging μ2-O2CMe links the two metal ions. The ZnII ion displays square pyramidal geometry, and the LnIII ion is nine-coordinated with intermediate geometry between the spherical capped square antiprism and the spherical tricapped trigonal prism. Complex 67S displays near-infrared luminescence in MeOH solution [52]. The pair of enantiomers 68R/68S were justified by solid-state CD spectroscopy. Both enantiomers display field-induced SIM behavior. Under 1.5 kOe external dc field, two relaxation processes are observed in the frequency and temperature dependences of the ac susceptibilities (Figure 17a). The fit of the data to the Arrhenius law gave the values given in Table 4 for the low- and the high-temperature relaxation, respectively. Polarization measurements as a function of the applied electric field on single crystals showed a hysteresis loop in the P vs. E curve at 400 K with a large coercitive field of ~17 kVcm−1 with polarization Ps~9.1 μCcm−2, which is comparable to other molecular ferroelectrics (Figure 17b, Table 5) [53]. The molecular crystals of 69R/69S display electrical bistability up to 573 K, which is the highest temperature reported for molecular ferroelectric materials. Moreover, this complex also exhibits optical activity, high- and low-temperature luminescence, para- and super-paramagnetic behavior at high- and low-temperature, respectively. The chiral complexes 69R/69S also display ferroelectric behavior and room temperature magnetoelectric coupling, therefore combining magnetic and electric polarizabilities in the same phase [54].
The enantiopure ligand H2L11RR (Scheme 1) gave the chiral complex [CuGd(L11RR)(NO3)3] (70R), Table 3, [55]. The molecular structure consists of a square planar CuII ion located in the (OphenoxoNimino)2 compartment of (L11RR)2− and a ten-coordinated GdIII ion bound to the (OphenoxoOmethoxo)2 donor set of (L11RR)2− and to three chelate nitrato groups. The chiral complex exhibits efficiency in second harmonic generation 0.3 times that of urea. The magnetic susceptibility studies revealed ferromagnetic coupling between the metal ions with exchange parameter J = 1.3 cm−1 (gCu = 2.10, gGd = 2.01).
The enantiopure ligand H2L13SS (Scheme 2) gave the chiral complex [ZnNd(L13SS)(NO3)3(dmf)2] (71S), Table 4, [52]. The ZnII ion is five-coordinated to the two Ophenoxo and two Nimino atoms of (L13SS)2− and one dmf molecule, whereas the NdIII is nine-coordinated to three O-atoms of the ligand, two chelate nitrates, one monodentate nitrate, and one dmf molecule. Complex 71S displays near-infrared luminescence in MeOH solution.
A chiral 1D coordination polymer [CuGd(L14SS)(thd)2]n (72S) was reported (Hthd = 2,2,6,6-tetramethyl-3,5-heptanedione), Table 3, [56]. The ligand H2L14SS in shown in Scheme 2. The CuII ion is square planar and is coordinated to the (OphenoxoNimino)2 donor atoms of the ligand. This subunit acts as a metalloligand to the two GdIII ions, Gd(1) and Gd(2). The two Ophenoxo and the one Omethoxo atoms coordinate at one side of the GdIII ion as a tridentate ligand. The Oamido coordinates with another GdIII ion as a monodentate ligand. The eight-coordination of each GdIII ion is completed by two bidentate thd ligands (Figure 18). The chiral 1D chains extend along the b axis. The magnetic studies revealed ferromagnetic coupling between the metal ions, compatible with spin S = 4 in the ground state.
The macrocycle enantiopure ligand H6L16RR (Scheme 3) gave the chiral complex [Zn3Er(L16RR)(O2CMe)(NO3)2(H2O)1.5(MeOH)0.5] (73R), Table 4, [57]. Each ZnII ion is five-coordinated to two Ophenoxo and two Nimino atoms from (L16RR)6− and one monodentate acetate, nitrate, or methanol ligands at the apical position of a square pyramid. The ErIII ion is nine-coordinated to the six Ophenoxo atoms of the ligand at equatorial positions, one bidentate nitrate, and a H2O molecule at apical sites. Complex 73R displays field-induced SIM behavior under 1.0 kOe external dc field (Figure 19).
Table 3. Structural, CD, and magnetic data for the heterometallic Cu/4f complexes.
Table 3. Structural, CD, and magnetic data for the heterometallic Cu/4f complexes.
Complex3d/4f IonsSpace GroupCoordination NumberCD, λmax
(nm)
χMT (r.t.) (cm3Kmol−1)Ueff (K)t0 (s)Ref.
11RCu/CeP111−290, −380, +600 a1.05--[39]
12RCu/NdP111−290, −380, +600 a1.52--[39]
13RCu/SmP2110−290, −380, +600 a0.64--[39]
14RCu/EuP110−290, −380, +600 a1.54--[39]
15R/15RCu/GdP1, P2110−290, −380, +600 a,b8.67, 8.22--[39]
16R/16RCu/TbP1, P2110−290, −380, +600 a,b12.35, 13.1030.02, 26.16 c9.81 × 10−9, 2.81 × 10−9[39]
17R/17RCu/DyP1, P2110−290, −380, +600 a,b14.76, 14.455.96, 15.72 c3.66 × 10−6, 1.70 × 10−7[39]
18RCu/HoP219, 10−300, −375, +60014.058.46 d3.03 × 10−7[39]
19RCu/ErP219, 10300, −375, +60010.36--[39]
20RCu/YbP219, 10300, −375, +6002.59--[39]
33R/33SCu/TbP18-25.1527.60 c2.03 × 10−5[43]
40R/40SCu, Fe/GdP219−380, −530, +470, +610 a9.11--[45]
41R/41SCu, Fe/TbP219−380, −530, +470, +610 a14.13--[45]
42R/42SCu, Fe/DyP219−380, −530, +470, +610 a17.69--[45]
43RCu/EuP2110----[46]
44R/44SCu/GdP21212110−237, −285, −390, +219, +261, +334 a8.38--[47]
70RCu/GdP21212110-8.20--[55]
72SCu/GdP32218-8.14--[56]
79Cu, Mo/Gd---9.11--[58]
80Cu, Mo/Tb---13.50--[58]
81Cu, Mo/DyP219----[58]
82Cu, W/LaP219-0.70--[58]
83Cu, W/GdP219-8.90--[58]
84Cu, W/TbP219-13.20--[58]
85Cu, W/DyP219----[58]
86Cu, Mo/LaC2222110-0.76--[58]
87Cu, Mo/Pr------[58]
a the values correspond to the R-enantiomer, same values with opposite sings correspond to the S-enantiomer; b for 15R-17R the λmax values are −300, −375, +600 nm; c under 2.0 kOe external dc field; d under zero dc field.
Table 4. Structural, CD, and magnetic data for the heterometallic Zn/4f complexes.
Table 4. Structural, CD, and magnetic data for the heterometallic Zn/4f complexes.
Complex3d/4f IonsSpace GroupCoordination NumberCD, λmax
(nm)
χMT (r.t.) (cm3Kmol−1)Ueff (K)t0 (s)Ref.
32RZn/LuP19----[40]
34RZn/DyP19−305, −37214.19212.1 b7.0 × 10−10[44]
35RZn/DyP19−310, −38614.21203.5 c1.0 × 10−10[44]
36RZn/DyP19−290, −40914.18207.3 c1.5 × 10−5[44]
37SZn/DyP19+305, +37228.36194.5 b3.1 × 10−9[44]
38SZn/DyP19+310, +38628.3370.1/231.6 c8.0 × 10−7/5.4 × 10−10[44]
39SZn/DyP19+290, +40928.40218.1 c3.5 × 10−11[44]
49R/49SZn/YP216----[49]
50R/50SZn/LuP216----[49]
51R/51SZn/DyP216----[49]
52R/52SZn/SmP216----[49]
53R/53SZn/LaP216----[49]
60RZn/CeP2110--4.7 d2.5 × 10−5[50]
61RZn/NdP2110--15.9 e3.7 × 10−6[50]
62R/62SZn/EuP2110−235, −300, +218, +265, +342, +395 a---[50]
63SZn/DyP2110--17.7 e8.3 × 10−7[50]
64SZn/ErP2110----[50]
65SZn/YbP2110----[50]
66R/66SZn/DyP2110−301, −360, +406 a14.4911.9/46.1 f4.23 × 10−5/8.85 × 10−8[51]
67SZn/NdP219--- [52]
68R/68SZn/DyP219−262, −303, +382 a13.4419.40/51.82 g
20.48/51.72 h
1.23 × 10−8/3.75 × 10−9
8.97 × 10−9/3.55 × 10−9
[53]
69R/69SZn/TbP219−261, −303, +386 a---[54]
71SZn/NdC29----[52]
73RZn/ErP2121219--8.1 i5.3 × 10−7[57]
a the values correspond to the R-enantiomer, same values with opposite sings correspond to the S-enantiomer; b under 1.8 kOe external dc field, n = 5.49, C = 1.6 × 10−3 s−1K−5.49 (34R), n = 5.72, C = 7.6 × 10−4 s−1K−5.72 (37S); c under 2.0 kOe external dc field, n = 5.06, C = 1.2 × 10−2 s−1K−5.06 (35R), n = 5.29, C = 1.9 × 10−2 s−1K−5.29 (36R), n = 4.54, C = 3.9 × 10−2 s−1K−4.54 for the slow relaxation and n = 5.98, C = 1.8 × 10−4 s−1K−5.98 for the fast relaxation (38S), n = 4.78, C = 5.4 × 10−2 s−1K−4.78 (39S); d under 0.5 kOe external dc field; e under 1.0 kOe external dc field; f under 4.0 kOe external dc field, for the low- and high-temperature relaxation; g under 1.5 kOe external dc field for the R-enantiomer, for the low- and high-temperature relaxation; h under 1.5 kOe external dc field for the S-enantiomer, for the low- and high-temperature relaxation; i under 1.0 kOe external dc field.
Table 5. Ferroelectric and spectroscopic data for the chiral 4f and 3d/4f complexes.
Table 5. Ferroelectric and spectroscopic data for the chiral 4f and 3d/4f complexes.
ComplexPr
(μCcm−2)
Ec
(kVcm−1)
UV-Vis
(nm)
Fluorescence (nm)MCD
gmax (T−1)
CPL
gPL
SHG, χR(2) (pmV−1)THG, χR(3) (pm2V−2)Ref.
6R4.5128.11------[35]
7S7.9822.12------[36]
32R---460----[40]
34R------0.04581[44]
35R------0.01821[44]
36R------0.07848[44]
37S------0.39485[44]
38S------0.02703[44]
39S------0.40718[44]
66R/66S--279, 3724906.7 × 10−2±2.5 × 10−3--[51]
68R9.117.0------[53]
70R--310-----[55]

4. Chiral Materials from Racemic Salen-Type Schiff Base Ligands

The condensation reaction of 2-hydroxy-benzaldehyde (salicylaldehyde) with 1,2-diamino-cyclohexane gave the Schiff base ligand H2L1 (Scheme 1) which upon reaction with Ce(NO3)3·6H2O in EtOH gave complex [CeIV(L1)2] (74) which crystallizes in the chiral monoclinic space group P21 [34]. The absolute structure was determined by the Flack parameter, x = 0.53(4), which indicates racemic twinning. The absolute configuration of both chiral centers in L(1)2− is R,S- and due to racemic twinning, the S,R- also exist in the crystal structure. The CeIV ion is eight-coordinated to the two Ophenoxo and the two Nimino atoms from two (L1)2− ligands with Ce-O and Ce-N distances in range 2.148(7)-2.241(9) Å and 2.525(11)-2.619(9) Å, respectively. The best planes of the coordination spheres of the two ligands form a dihedral angle of 87.7(3)°. The coordination geometry is best described as snub diphenoid (JSD-8, CShM = 1.31927).
The zwitterion of racemic H2L1 forms 3D coordination polymers [LnIIICl(NO3)2(H2L1)]n (LnIII = La (75), Ce (76), Nd (77)) [59,60] which crystallize in the tetragonal chiral space group P43212 with diamond (dia) network topology (Figure 20). The LnIII ion is nine-coordinated to two chelate nitrato and one chlorido group, as well as to four Ophenoxo atoms from four different H2L1 zwitterions. The coordination polyhedron around the LnIII ion is a tricapped trigonal prism (JTCTPR-9). The 3D structure is built by [LnIII(H2L1)]3+ subunits, with the LnIII ion and the H2L1 serving as the head and tail, respectively, of a tadpole-like model, with all the tails being right-handed. In this head-to-tail ligation mode, the tadpole-like models link each other and form two 1D right-handed helical chains, each one directed along the a and b axis. Two such helixes interweave each other by sharing the LnIII ions, thus forming the dia 3D network. Compounds 75–77 contain the enantiopure H2L1SS ligand due to the trans orientation of the ligand, which shows higher stability and lower steric hindrance, and also due to hydrogen bonds which stabilize the structure. Compounds 75–77 represent the first 3D chiral lanthanide MOFs with dia topology and opened new ways for the synthesis of enantiopure materials from racemate precursors. The chirality of these MOFs results in the SHG effects, and they are the first examples of NLO-active enantiopure Ln-salen frameworks.
The use of the zwitterion of racemic ligand H2L2 (Scheme 1) gave the 3D polymer {[CeIII(H2L2)(NO3)2](PF6)}n (78), which crystallizes in the hexagonal chiral space group P6422 with quartz (qtz) topology [61]. The structure consists of right-handed helixes of [CeIII(H2L2)]n. The CeIII ion is eight-coordinated to four Ophenoxo atoms from four different H2L2 ligands and to two chelate nitrato groups in a hexagonal bipyramid geometry (HB PY-8, CShM = 7.53410, Figure 21).
The racemic ligand H2L15 (Scheme 2) gave the chiral 1D coordination polymers [{CuLn(L15)(H2O)4}{M(CN)8}]n (MV = Mo, LnIII = Gd (79), Tb (80), Dy (81); MV = W, LnIII = La (82), Gd (83), Tb (84), Dy (85)) and [{Cu2Ln(L15)2(H2O)4}{Mo(CN)8}] (LnIII = La (86), Pr (87)) [58]. The polymeric structure of 81–85 is based on {M(CN)8} units coordinated through one cyano ligand to the CuII ion of a dinuclear {CuLn} moiety. The {CuIILnIIIMV} fragment is further linked through a second cyano group to the LnIII ion of an adjacent trimetallic fragment (Figure 22, Table 3). The (L15)2− ligand is coordinated via the two Ophenoxo and the Nimino atoms to the CuII ion and through the two bridging Ophenoxo and the two Omethoxo to the LnIII. The nine-coordination of the latter is completed by four aqua ligands and one nitrogen atom from the cyano bridge. The coordination geometry is distorted monocapped square antiprism. Intermolecular O···O and N···O type hydrogen bonds, involving the free cyano groups, the crystallization, and coordinated H2O molecules, establish a 3D supramolecular network. The structure of 86 is based on trinuclear {CuII2(L15)2LaIII} moieties linked by {Mo(CN)8}3− anions, resulting in 1D chains running along the crystallographic c axis. There are two C2 axes passing through LaIII and MoV; thus, half of the [{CuLa(L15)(H2O)4}{Mo(CN)8}] repeating unit is independent (Figure 23). The structure of the trinuclear {CuII2(L15)2LaIII} moiety consists of a central LaIII ion, which is bound to the (OphenoxoOmethoxo)2 donor set from the two (L15)2− ligands, whereas each of the CuII ions is hosted in the (OphenoxoNimino)2 pocket of each (L15)2− ligand. The ten-coordination around the LaIII ion is completed from cyano bridges and is described as distorted bicapped square antiprism. Each CuII ion is five-coordinated with one cyano bridge in the apical position of the square pyramid. The chains of 86 are linked through C-H···π intermolecular interactions and form a 3D supramolecular network.

5. Concluding Comments

The enantiopure Schiff base ligands depicted in Scheme 1, Scheme 2 and Scheme 3 gave homometallic 4f and heterometallic 3d-4f chiral complexes with diverse nuclearities, metal topologies, coordination geometries, and physical properties. In the homometallic 4f complexes 1–8, the LnIII ion is coordinated to the Ophenoxo/Nimino atoms of the Schiff base ligand, whereas in complexes 9 and 10, which contain the ligand in the zwitterion form, the lanthanide is bound to the Ophenoxo and/or Omethoxo atoms. In the 4f complexes 1–10, the LnIII ions display coordination numbers, which range from seven to ten. In the heterometallic 3d-4f complexes 11–73, the 3d metal ions are coordinated to the Schiff base ligands through the Ophenoxo/Nimino atoms in square planar geometry. In some cases, five- or six-coordination is observed from coordinated co-ligands, such as solvent molecules, organic or inorganic anions, and the 3d ions display square planar or octahedral geometries. In complexes 11–73, the lanthanide ions are coordinated via the Ophenoxo/Oalkoxo atoms of the ligands; the Ophenoxo atoms serve as bridging atoms between the 3d and 4f ions. The coordination environment of the LnIII ions is completed by co-ligands, such as nitrates, azides, and carboxylates, and gives coordination numbers from six to eleven. In the chiral complexes 74–86 derived from racemic Schiff bases, the LnIII ions are coordinated to the Ophenoxo/Nimino atoms, or to the Ophenoxo/Oalkoxo atoms, or to the Ophenoxo atoms, depending on the simultaneous presence of other metal ions, or the coordination diversity of the ligand which promotes the formation of network structures.
The chiral complexes discussed herein constitute multifunctional molecular materials that combine chirality with various physical properties. The majority of the complexes display field-induced slow magnetic relaxation compatible with single-ion magnetic behavior. This behavior was observed in complexes of oblate ions, such as TbIII, DyIII, and HoIII, prolate ions, such as YbIII, and of ‘uncommon magnetic lanthanides’, CeIII, and NdIII. Other physical properties, such as ferroelectric, optical, catalytic, and anticancer behavior, were reported.

Funding

This research received no external funding.

Data Availability Statement

Data available in a publicly accessible repository.

Conflicts of Interest

The author declares no conflicts of interest.

References

  1. Bruce, D.W.; O’Hare, D.; Walton, R.I. (Eds.) Molecular Materials; Wiley: Hoboken, NJ, USA, 2011. [Google Scholar]
  2. Coronado, E. Molecular magnetism: From chemical design to spin control in molecules, materials and devices. Nat. Rev. Mater. 2020, 5, 87–194. [Google Scholar] [CrossRef]
  3. Atzori, M.; Artizzu, F. (Eds.) Functional Molecular Materials: An Introductory Textbook; Pan Stanford Publishing, Pte. Ltd.: Singapore, 2018. [Google Scholar]
  4. Ouahab, L. Multifunctional Molecular Materials; Pan Stanford Publishing, Pte. Ltd.: Singapore, 2013. [Google Scholar]
  5. Mamula, O.; von Zelewsky, A. Supramolecular coordination compounds with chiral pyridine and polypyridine ligands derived from terpenes. Coord. Chem. Rev. 2003, 242, 87–95. [Google Scholar] [CrossRef]
  6. Bauer, E.B. Chiral-at-metal complexes and their catalytic applications in organic synthesis. Chem. Soc. Rev. 2012, 41, 3153–3167. [Google Scholar] [CrossRef] [PubMed]
  7. Knof, U.; von Zelewsky, A. Predetermined chirality at metal centers. Angew. Chem. Int. Ed. Engl. 1999, 38, 302–322. [Google Scholar] [CrossRef]
  8. Mukhtar, S.D.; Suhail, M. Chiral metallic anticancer drugs: A brief-review. Eur. J. Chem. 2022, 13, 483–490. [Google Scholar] [CrossRef]
  9. Anthony, E.J.; Bolitho, E.M.; Bridgewater, H.E.; Carter, O.W.L.; Donnelly, J.M.; Imberti, C.; Lant, E.C.; Lermyte, F.; Needham, R.J.; Palau, M.; et al. Metallodrugs are unique: Opportunities and challenges of discovery and development. Chem. Sci. 2020, 11, 12888–12917. [Google Scholar] [CrossRef]
  10. Li, D.-P.; Wang, T.-W.; Li, C.-H.; Liu, D.-S.; Li, Y.-Z.; You, X.-Z. Single-ion magnets based on mononuclear lanthanide complexes with chiral Schiff base ligands [Ln(FTA)3L] (Ln = Sm, Eu, Gd, Tb and Dy). Chem. Commun. 2010, 46, 2929–2931. [Google Scholar] [CrossRef]
  11. Guo, P.-H.; Liu, J.-L.; Jia, J.-H.; Wang, J.; Guo, F.-S.; Chen, Y.-C.; Lin, W.-Q.; Leng, J.-D.; Bao, D.-H.; Zhang, X.-D.; et al. Multifunctional DyIII4 cluster exhibiting white-emitting, ferroelectric and single-molecule magnet behavior. Chem. Eur. J. 2013, 19, 8769–8773. [Google Scholar] [CrossRef]
  12. Wen, H.-R.; Hu, J.-J.; Yang, K.; Zhang, J.-L.; Liu, S.-J.; Liao, J.-S.; Liu, C.-M. Family of chiral ZnII-LnIII (Ln = Dy and Tb) heterometallic complexes derived from the amine-phenol ligand showing multifunctional properties. Inorg. Chem. 2020, 59, 2811–2824. [Google Scholar] [CrossRef]
  13. Liu, C.-M.; Zhang, D.-Q.; Xiong, R.-G.; Hao, X.; Zhu, D.-B. A homochiral Zn-Dy heterometallic left-handed helical chain complex without chiral ligands: Anion-induced assembly and multifunctional integration. Chem. Commun. 2018, 54, 13379–13382. [Google Scholar] [CrossRef]
  14. Wang, K.; Zeng, S.; Wang, H.; Dou, J.; Jiang, J. Magneto-chiral dichroism in chiral mixed (phthalocyaninato)(porphyrinato) rare earth triple-decker SMMs. Inorg. Chem. Front. 2014, 1, 167–171. [Google Scholar] [CrossRef]
  15. Atzori, M.; Dhbaibi, K.; Douib, H.; Grasser, M.; Dorcet, V.; Breslavetz, I.; Paillot, K.; Cador, O.; Rikken, G.L.J.A.; Le Guennic, B.; et al. Helicene-based ligands enable strong magneto-chiral dichroism in a chiral ytterbium complex. J. Am. Chem. Soc. 2021, 143, 2671–2675. [Google Scholar] [CrossRef] [PubMed]
  16. Liu, C.-M.; Sun, R.; Wang, B.-W.; Wu, F.; Hao, X.; Shen, Z. Homochiral Ferromagnetic Coupling Dy2 single-molecule magnets with strong magneto-optical Faraday effects at room temperature. Inorg. Chem. 2021, 60, 12039–12048. [Google Scholar] [CrossRef] [PubMed]
  17. Wang, X.; Du, M.-H.; Xu, H.; Long, L.-S.; Kong, X.-J.; Zheng, L.-S. Cocrystallization of chiral 3d-4f clusters {Mn10Ln6} and {Mn6Ln2}. Inorg. Chem. 2021, 60, 5925–5930. [Google Scholar] [CrossRef] [PubMed]
  18. Lefeuvre, B.; Mattei, C.A.; Gonzalez, J.F.; Gendron, F.; Dorcet, V.; Riobé, F.; Lalli, C.; Guennic, B.L.; Cador, O.; Maury, O.; et al. Solid-state near-infrared circularly polarized luminescence from chiral YbIII-single-molecule magnet. Chem. Eur. J. 2021, 27, 7362–7366. [Google Scholar] [CrossRef] [PubMed]
  19. El Rez, B.; Liu, J.; Béreau, V.; Duhayon, C.; Horino, Y.; Suzuki, T.; Coolen, L.; Sutter, J.-P. Concomitant emergence of circularly polarized luminescence and single-molecule magnet behavior in chiral-at-metal Dy complex. Inorg. Chem. Front. 2020, 7, 4527–4534. [Google Scholar] [CrossRef]
  20. Woodruff, D.N.; Winpenny, R.E.P.; Layfield, R.A. Lanthanide single-molecule magnets. Chem. Rev. 2013, 113, 5110–5148. [Google Scholar] [CrossRef] [PubMed]
  21. Meng, Y.-S.; Jiang, S.-D.; Wang, B.-W.; Gao, S. Understanding the magnetic anisotropy toward single-ion magnets. Acc. Chem. Res. 2016, 49, 2381–2389. [Google Scholar] [CrossRef]
  22. Liu, K.; Shi, W.; Cheng, P. Toward heterometallic single-molecule magnets: Synthetic strategy, structures and properties of 3d-4f discrete complexes. Coord. Chem. Rev. 2015, 289–290, 74–122. [Google Scholar] [CrossRef]
  23. Gupta, S.K.; Rajeshkumar, T.; Rajaraman, G.; Murugavel, R. An air-stable Dy(III) single-ion magnet with high anisotropy barrier and blocking temperature. Chem. Sci. 2016, 7, 5181–5191. [Google Scholar] [CrossRef]
  24. Liu, J.; Chen, Y.-C.; Liu, J.-L.; Vieru, V.; Ungur, L.; Jia, J.-H.; Chibotaru, L.F.; Lan, Y.; Wernsdorfer, W.; Gao, S.; et al. A stable pentagonal bipyramidal Dy(III) single-ion magnet with a record magnetization reversal barrier over 1000 K. J. Am. Chem. Soc. 2016, 138, 5441–5450. [Google Scholar] [CrossRef] [PubMed]
  25. Ding, Y.-S.; Chilton, N.F.; Winpenny, R.E.P.; Zheng, Y.-Z. On approaching the limit of molecular magnetic anisotropy: A near-perfect pentagonal bipyramidal dysprosium(III) single-molecule magnet. Angew. Chem. Int. Ed. 2016, 55, 16071–16074. [Google Scholar] [CrossRef] [PubMed]
  26. Meng, Y.-S.; Xu, L.; Xiong, J.; Yuan, Q.; Liu, T.; Wang, B.-W.; Gao, S. Low-coordinate single-ion magnets by intercalation of lanthanides into a phenol matrix. Angew. Chem. Int. Ed. 2018, 57, 4763–4766. [Google Scholar] [CrossRef]
  27. Goodwin, C.A.P.; Ortu, F.; Reta, D.; Chilton, N.F.; Mills, D.P. Molecular magnetic hysteresis at 60 kelvin in dysprosocenium. Nature 2017, 548, 439–442. [Google Scholar] [CrossRef] [PubMed]
  28. Guo, F.-S.; Day, B.M.; Chen, Y.-C.; Tong, M.-L.; Mansikkamäki, A.; Layfield, R.A. Magnetic hysteresis up to 80 kelvin in a dysprosium metallocene single-molecule magnet. Science 2018, 362, 1400–1403. [Google Scholar] [CrossRef] [PubMed]
  29. Meena, R.; Meena, P.; Kumari, A.; Sharma, N.; Fahmi, N. Schiff Base in Organic, Inorganic and Physical Chemistry; Chapter 4; Akitsu, T., Ed.; IntechOpen: London, UK, 2023. [Google Scholar]
  30. Boulechfar, C.; Ferkous, H.; Delimi, A.; Djedouani, A.; Kahlouche, A.; Boublia, A.; Darwish, A.S.; Lemaoui, T.; Verma, R.; Benguerba, Y. Schiff bases and their metal complexes: A review on the history, synthesis, and applications. Inorg. Chem. Commun. 2023, 150, 110451. [Google Scholar] [CrossRef]
  31. Ceramella, J.; Iacopetta, D.; Catalano, A.; Cirillo, F.; Lappano, R.; Sinicropi, M.S. A review on the antimicrobial activity of Schiff bases: Data collection and recent studies. Antibiotics 2022, 11, 191. [Google Scholar] [CrossRef]
  32. Ren, M.; Xu, Z.-L.; Wang, T.-T.; Bao, S.-S.; Zheng, Z.-H.; Zhang, Z.-C.; Zheng, L.-M. Homochiral mononuclear Dy-Schiff base complexes showing field-induced double magnetic relaxation processes. Dalton Trans. 2016, 45, 690–695. [Google Scholar] [CrossRef]
  33. Yao, M.-X.; Zheng, Q.; Gao, F.; Li, Y.-Z.; Song, Y.; Zuo, J.-L. Field-induced slow magnetic relaxation in chiral seven-coordinated mononuclear lanthanide complexes. Dalton Trans. 2012, 41, 13682–13690. [Google Scholar] [CrossRef]
  34. Szłyk, E.; Wojtczak, A.; Dobrzańska, L.; Barwiołek, M. X-ray crystal structure and nuclear Overhauser effect studies of cerium(IV) complexes with Schiff bases obtained from N,N′-(1R,2R)(-)-1,2-cyclohexanediamine and benzaldehyde derivatives. Polyhedron 2008, 27, 765–776. [Google Scholar] [CrossRef]
  35. Sui, Y.; Fang, X.-N.; Hu, R.-H.; Li, J.; Liu, D.-S. A new type of multifunctional single ionic dysprosium complex based on chiral salen-type Schiff base ligand. Inorg. Chim. Acta 2014, 423, 540–544. [Google Scholar] [CrossRef]
  36. Sui, Y.; Hu, R.-H.; Luo, Z.-G.; Lin, W.-H.; Liu, D.-S. Synthesis, structure and properties of an erbium(iii) complex with chiral salen-type Schiff base ligand. Z. Anorg. Allg. Chem. 2015, 641, 1566–1570. [Google Scholar] [CrossRef]
  37. Okumura, Y.; Takiguchi, Y.; Nakame, D.; Akitsu, T. Crystal structure and Hirshfeld surface analysis of ((S,S)-2,2′-{(1,2-diphenylethane-1,2-diyl)bis[(azaniumylylidene) methanylylidene]}bis(6-methoxyphenolato))trinitratosamarium(III). Acta Cryst. 2021, E77, 579–582. [Google Scholar] [CrossRef] [PubMed]
  38. Fan, Y.-H.; Wang, A.-D.; Bi, C.-F.; Xiao, Y.; Bi, S.-Y.; Zhang, X.; Wang, Q. Synthesis, crystal structure and anticancer activity of 2D-coordination polymer of cerium(III) with chiral Schiff base trans-N,N-bis-(2-hydroxy-1-naphthalidehydene)-(1R,2R)-cyclohexanediamine. Synth. Met. 2011, 161, 1552–1556. [Google Scholar] [CrossRef]
  39. Wen, H.-R.; Bao, J.; Liu, S.-J.; Liu, C.-M.; Zhang, C.-W.; Tang, Y.-Z. Temperature-controlled polymorphism of chiral CuII-LnIII dinuclear complexes exhibiting slow magnetic relaxation. Dalton Trans. 2015, 44, 11191–11201. [Google Scholar] [CrossRef] [PubMed]
  40. Orita, S.; Akitsu, T. Variety of crystal structures of chiral Schiff base Lu(III)-Ni(II)/Cu(II)/Zn(II) and the related complexes. Open Chem. 2014, 1, 1–14. [Google Scholar] [CrossRef]
  41. Wen, H.-R.; Liu, S.-J.; Xie, X.-R.; Bao, J.; Liu, C.-M.; Chen, J.-L. A family of nickel-lanthanide heterometallic dinuclear complexes derived from a chiral Schiff-base ligand exhibiting single-molecule magnet behaviors. Inorg. Chim. Acta 2015, 435, 274–282. [Google Scholar] [CrossRef]
  42. Okamoto, Y.; Nidaira, K.; Akitsu, T. Environmental dependence of artifact CD peaks of chiral Schiff base 3d-4f complexes in soft mater PMMA matrix. Int. J. Mol. Sci. 2011, 12, 6966–6979. [Google Scholar] [CrossRef] [PubMed]
  43. Huang, X.-C.; Vieru, V.; Chibotaru, L.F.; Wernsdorfer, W.; Jiang, S.-D.; Wang, X.-Y. Determination of magnetic anisotropy in a multinuclear TbIII-based single-molecule magnet. Chem. Commun. 2015, 51, 10373–10376. [Google Scholar] [CrossRef]
  44. Liu, C.-M.; Sun, R.; Wang, B.-W.; Hao, X.; Li, X.-L. Effects of counterions, coordination anions, and coordination solvent molecules on single-molecule magnetic behaviors and nonlinear optical properties of chiral Zn2Dy Schiff base complexes. Inorg. Chem. 2022, 61, 18510–18523. [Google Scholar] [CrossRef]
  45. Deng, X.-W.; Cai, L.-Z.; Zhu, Z.-X.; Gao, F.; Zhou, Y.-L.; Yao, M.-X. Synthesis, structures and magnetic properties of chiral 3d–3d’–4f heterotrimetallic complexes based on [(Tp*)Fe(CN)3]. New J. Chem. 2017, 41, 5988–5994. [Google Scholar] [CrossRef]
  46. Constable, E.C.; Zhang, G.; Housecroft, C.E.; Neuburger, M.; Zampese, J.A. The mononuclear-dinuclear dance: Twisting the backbone in metalloligands operates a coordination switch. Inorg. Chim. Acta 2010, 363, 4207–4213. [Google Scholar] [CrossRef]
  47. Sui, Y.; Liu, D.-S.; Hu, R.-H.; Huang, J.-G. One-dimensional zigzag chain of Cu-Gd coordination polymers derived from chiral hexadentate Schiff base ligands: Synthesis, structure and magnetic properties. Inorg. Chim. Acta 2013, 395, 225–229. [Google Scholar] [CrossRef]
  48. Yadav, M.; Bhunia, A.; Jana, S.K.; Roesky, P.W. Manganese- and lanthanide-based 1D chiral coordination polymers as an enantioselective catalyst for sulfoxidation. Inorg. Chem. 2016, 55, 2701–2708. [Google Scholar] [CrossRef] [PubMed]
  49. Xu, R.; Hua, L.; Li, X.; Yao, Y.; Leng, X.; Chen, Y. Rare-earth/zinc heterometallic complexes containing both alkoxy-amino-bis(phenolato) and chiral salen ligands: Synthesis and catalytic application for copolymerization of CO2 with cyclohexene oxide. Dalton Trans. 2019, 48, 10565–10573. [Google Scholar] [CrossRef] [PubMed]
  50. Mayans, J.; Saez, Q.; Font-Bardia, M.; Escuer, A. Enhancement of magnetic relaxation properties with 3d diamagnetic cations in [ZnIILnIII] and [NiIILnIII], LnIII = Kramers lanthanides. Dalton Trans. 2019, 48, 641–652. [Google Scholar] [CrossRef] [PubMed]
  51. Huang, H.; Sun, R.; Wu, X.-F.; Liu, Y.; Zhan, J.-Z.; Wang, B.-W.; Gao, S. Circularly polarized luminescence and magnetooptic effects from chiral Dy(III) single molecule magnets. Dalton Trans. 2023, 52, 7646–7651. [Google Scholar] [CrossRef] [PubMed]
  52. Bi, W.-Y.; Lu, X.-Q.; Chai, W.-L.; Song, J.-R.; Wong, W.-Y.; Wong, W.-K.; Jones, R.A. Construction and NIR luminescent property of hetero-bimetallic Zn-Nd complexes from two chiral salen-type Schiff-base ligands. J. Mol. Struct. 2008, 891, 450–455. [Google Scholar] [CrossRef]
  53. Long, J.; Rouquette, J.; Thibaud, J.-M.; Ferreira, R.A.S.; Carlos, L.D.; Donnadieu, B.; Vieru, V.; Chibotaru, L.F.; Konczewicz, L.; Haines, J.; et al. A high-temperature molecular ferroelectric Zn/Dy complex exhibiting single-ion-magnet behavior and lanthanide luminescence. Angew. Chem. Int. Ed. Engl. 2015, 54, 2236–2240. [Google Scholar] [CrossRef]
  54. Long, J.; Ivanov, M.S.; Khomchenko, V.A.; Mamontova, E.; Thibaud, J.-M.; Rouquette, J.; Beaudhiun, M.; Granier, D.; Ferreira, R.A.S.; Carlos, L.D.; et al. Room temperature magnetoelectric coupling in a molecular ferroelectric ytterbium(III) complex. Science 2020, 367, 671–676. [Google Scholar] [CrossRef]
  55. Margeat, O.; Lacroix, P.G.; Costes, J.P.; Donnadieu, B.; Lepetit, C. Synthesis, structures, and physical properties of copper(II)-gadolinium(III) complexes combining ferromagnetic coupling and quadratic nonlinear optical properties. Inorg. Chem. 2004, 43, 4743–4750. [Google Scholar] [CrossRef] [PubMed]
  56. Hamamatsu, T.; Matsumoto, N.; Re, N.; Mrozinski, J. Chiral ferromagnetic chain of copper(II)-gadolinium(III) complex. Chem. Lett. 2009, 38, 762–763. [Google Scholar] [CrossRef]
  57. Yamashita, A.; Watanabe, A.; Akine, S.; Nabeshima, T.; Nakano, M.; Yamamura, T.; Kajiwara, T. Wheel-shaped ErIIIZnII3 single-molecule magnet: A macrocyclic approach to designing magnetic anisotropy. Angew. Chem. Int. Ed. Engl. 2011, 50, 4016–4019. [Google Scholar] [CrossRef] [PubMed]
  58. Visinescu, D.; Alexandru, M.-G.; Madalan, A.M.; Pichon, C.; Duhayon, C.; Sutter, J.-P.; Andruh, M. Magneto-structural variety of new 3d-4f-4(5)d heterotrimetallic complexes. Dalton Trans. 2015, 44, 16713–16727. [Google Scholar] [CrossRef] [PubMed]
  59. Fustero, S.; Román, R.; Asensio, A.; Maesto, M.A.; Aceña, J.L.; Simón-Fuentes, A. An approach to 2,4-substituted pyrazolo[1,5-a]pyridines and pyrazolo[1,5-a]azepines by ring-closing metathesis. Eur. J. Org. Chem. 2013, 7164–7174. [Google Scholar] [CrossRef]
  60. Sun, J.-W.; Zhu, J.; Song, H.-F.; Li, G.-M.; Yao, X.; Yan, P.-F. Spontaneous resolution of racemic salen-type ligand in the construction of 3d homochiral lanthanide frameworks. Cryst. Growth Des. 2014, 14, 5356–5360. [Google Scholar] [CrossRef]
  61. Gao, B.; Zhang, Q.; Yan, P.; Hou, G.; Li, G. Crystal engineering of salen type cerium complexes tuned by various cerium counterions. CrystEngComm 2013, 15, 4167–4175. [Google Scholar] [CrossRef]
Scheme 1. The ligands H2L1–H2L12.
Scheme 1. The ligands H2L1–H2L12.
Crystals 14 00474 sch001
Scheme 2. The ligands H2L13–H2L15.
Scheme 2. The ligands H2L13–H2L15.
Crystals 14 00474 sch002
Scheme 3. The ligand H2L16.
Scheme 3. The ligand H2L16.
Crystals 14 00474 sch003
Figure 1. The molecular structure of one of the independent [DyIII(L4RR)2] anions in 1R, CCDC-1424808 [32]. Color code: Dy pink, O red, N blue, C grey. The ligand in the front is shown with darker grey bonds for clarity purposes.
Figure 1. The molecular structure of one of the independent [DyIII(L4RR)2] anions in 1R, CCDC-1424808 [32]. Color code: Dy pink, O red, N blue, C grey. The ligand in the front is shown with darker grey bonds for clarity purposes.
Crystals 14 00474 g001
Figure 2. The labeled molecular structure of 5R. The ligands are displayed with filled and open bonds to visualize their spatial arrangement around the central ion (C1-C20 and C21-C40, respectively) [34]. Copyright © 2008 Elsevier Ltd. All rights reserved.
Figure 2. The labeled molecular structure of 5R. The ligands are displayed with filled and open bonds to visualize their spatial arrangement around the central ion (C1-C20 and C21-C40, respectively) [34]. Copyright © 2008 Elsevier Ltd. All rights reserved.
Crystals 14 00474 g002
Figure 3. (a) Electric hysteresis loop of 6R at r.t. for single crystal sample. (b) Temperature dependence of the dielectric constant of 6R at various frequencies. Inset: dielectric constant at high frequencies [35]. Copyright © 2014 Elsevier B.V. All rights reserved.
Figure 3. (a) Electric hysteresis loop of 6R at r.t. for single crystal sample. (b) Temperature dependence of the dielectric constant of 6R at various frequencies. Inset: dielectric constant at high frequencies [35]. Copyright © 2014 Elsevier B.V. All rights reserved.
Crystals 14 00474 g003
Figure 4. (a) CD spectra of 8R/8S and H2L9RR,SS in KBr pellets. (b) Partially labeled plot of the molecular structure of 8R and local coordination geometry of the DyIII ions; ethyl groups of the phosphates and H-atoms are omitted for clarity. Reproduced from Ref. [33] with permission from the Royal Society of Chemistry.
Figure 4. (a) CD spectra of 8R/8S and H2L9RR,SS in KBr pellets. (b) Partially labeled plot of the molecular structure of 8R and local coordination geometry of the DyIII ions; ethyl groups of the phosphates and H-atoms are omitted for clarity. Reproduced from Ref. [33] with permission from the Royal Society of Chemistry.
Crystals 14 00474 g004
Figure 7. The molecular structure of 33R, CCDC-1000153, [43]. Color code: Tb pink, Cu green, O red, N blue, C grey.
Figure 7. The molecular structure of 33R, CCDC-1000153, [43]. Color code: Tb pink, Cu green, O red, N blue, C grey.
Crystals 14 00474 g007
Figure 8. The hysteresis loops of 33R measured on a single crystal on a μ-SQUID at the indicated temperatures and field sweep rates. Reproduced from Ref. [43] with permission from the Royal Society of Chemistry.
Figure 8. The hysteresis loops of 33R measured on a single crystal on a μ-SQUID at the indicated temperatures and field sweep rates. Reproduced from Ref. [43] with permission from the Royal Society of Chemistry.
Crystals 14 00474 g008
Figure 9. The molecular structure of one of the independent molecules in 34R, CCDC-2182147, [44]. Color code: Dy pink, Zn lime, Cl yellow, O red, N blue, C grey.
Figure 9. The molecular structure of one of the independent molecules in 34R, CCDC-2182147, [44]. Color code: Dy pink, Zn lime, Cl yellow, O red, N blue, C grey.
Crystals 14 00474 g009
Figure 10. SHG spectra of crystalline samples of 37S-39S (a) and 34R-36R (b) with KH2PO4 (KDP) under excitation at λ = 1550 nm (Tint = 0.5 s). Reprinted with permission from Inorg. Chem. 2022, 61, 18510–18523 [44]. Copyright © 2022, American Chemical Society.
Figure 10. SHG spectra of crystalline samples of 37S-39S (a) and 34R-36R (b) with KH2PO4 (KDP) under excitation at λ = 1550 nm (Tint = 0.5 s). Reprinted with permission from Inorg. Chem. 2022, 61, 18510–18523 [44]. Copyright © 2022, American Chemical Society.
Crystals 14 00474 g010
Figure 11. The molecular structure of the cation in 42R, CCDC-1537066, [45]. Color code: Dy pink, Cu green, Fe orange, B light grey, O red, N blue, C dark grey.
Figure 11. The molecular structure of the cation in 42R, CCDC-1537066, [45]. Color code: Dy pink, Cu green, Fe orange, B light grey, O red, N blue, C dark grey.
Crystals 14 00474 g011
Figure 12. Part of the chiral 1D chain of 48R, CCDC-1471181, [48]. Color code: Gd pink, Mn orange, Cl yellow, O red, N blue, C grey. Symmetry operations: (′) x, y, 1 + z; (″) 2 − x, y, 1 − z; (‴) 2 − x, y, 2 − z.
Figure 12. Part of the chiral 1D chain of 48R, CCDC-1471181, [48]. Color code: Gd pink, Mn orange, Cl yellow, O red, N blue, C grey. Symmetry operations: (′) x, y, 1 + z; (″) 2 − x, y, 1 − z; (‴) 2 − x, y, 2 − z.
Crystals 14 00474 g012
Figure 13. The molecular structure of the cation in 51R, CCDC-1870215, [49]. Color code: Dy pink, Zn yellow, Si light grey, O red, N blue, C dark grey.
Figure 13. The molecular structure of the cation in 51R, CCDC-1870215, [49]. Color code: Dy pink, Zn yellow, Si light grey, O red, N blue, C dark grey.
Crystals 14 00474 g013
Figure 14. Solution CD spectra for the 56R/56S (top) and 62R/62S (bottom) pairs of enantiomers. R-enantiomers, red lines; S-enantiomers, blue lines. Reproduced from Ref. [50] with permission from the Royal Society of Chemistry.
Figure 14. Solution CD spectra for the 56R/56S (top) and 62R/62S (bottom) pairs of enantiomers. R-enantiomers, red lines; S-enantiomers, blue lines. Reproduced from Ref. [50] with permission from the Royal Society of Chemistry.
Crystals 14 00474 g014
Figure 15. (a) UV–visible absorption spectra of 66R and 66S (298 K, DMF). (b) CD spectra of 66R and 66S (H = 0, +1.6, and −1.6 T) in the ranges of 270–600 nm at a concentration of 0.05 g L−1 (298 K, DMF). (c) CD spectra of 66R and 66S at 0.05 g L−1 (298 K, DMF). (d) PL spectra of 66R. Reproduced from Ref. [51] with permission from the Royal Society of Chemistry.
Figure 15. (a) UV–visible absorption spectra of 66R and 66S (298 K, DMF). (b) CD spectra of 66R and 66S (H = 0, +1.6, and −1.6 T) in the ranges of 270–600 nm at a concentration of 0.05 g L−1 (298 K, DMF). (c) CD spectra of 66R and 66S at 0.05 g L−1 (298 K, DMF). (d) PL spectra of 66R. Reproduced from Ref. [51] with permission from the Royal Society of Chemistry.
Crystals 14 00474 g015
Figure 16. Optical properties of 66R/66S measured at room temperature. (a) Circularly polarized luminescence emission spectra. (b) gPL vs. wavelength curves. Reproduced from Ref. [51] with permission from the Royal Society of Chemistry.
Figure 16. Optical properties of 66R/66S measured at room temperature. (a) Circularly polarized luminescence emission spectra. (b) gPL vs. wavelength curves. Reproduced from Ref. [51] with permission from the Royal Society of Chemistry.
Crystals 14 00474 g016
Figure 17. (a) Top: Frequency dependence of the out-of-phase (χ″) susceptibility at different temperatures performed under a 1.5 kOe dc field for 68R. Bottom: Cole-Cole plot for 68R obtained using the ac susceptibility data (1.5 kOe). The solid lines correspond to the best fit obtained with a generalized Debye model. (b) Dielectric hysteresis loop (1 Hz) for 68R obtained from a single crystal at different temperatures [53]. © 2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Figure 17. (a) Top: Frequency dependence of the out-of-phase (χ″) susceptibility at different temperatures performed under a 1.5 kOe dc field for 68R. Bottom: Cole-Cole plot for 68R obtained using the ac susceptibility data (1.5 kOe). The solid lines correspond to the best fit obtained with a generalized Debye model. (b) Dielectric hysteresis loop (1 Hz) for 68R obtained from a single crystal at different temperatures [53]. © 2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Crystals 14 00474 g017
Figure 18. Part of the chiral 1D chain of 72S, CCDC-733874, [56]. Color code: Gd pink, Cu green, O red, N blue, C grey. Symmetry operations: (′) 1 − x, 1 − x + y, 5/3 − z; (″) 1 − x, −x + y, 5/3 − z; (‴) x, −1 + y, z.
Figure 18. Part of the chiral 1D chain of 72S, CCDC-733874, [56]. Color code: Gd pink, Cu green, O red, N blue, C grey. Symmetry operations: (′) 1 − x, 1 − x + y, 5/3 − z; (″) 1 − x, −x + y, 5/3 − z; (‴) x, −1 + y, z.
Crystals 14 00474 g018
Figure 19. (a) Top view (top) and side view (bottom) of the molecular structure of 73R. Color code: Er gree, Zn yellow, O red, N blue, C light grey. (b) (A) Temperature and (B) frequency dependence of χM″ values of 73R measured under 1.0 kOe external field. The solid curves are guides for the eye [57]. Copyright © 2011 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Figure 19. (a) Top view (top) and side view (bottom) of the molecular structure of 73R. Color code: Er gree, Zn yellow, O red, N blue, C light grey. (b) (A) Temperature and (B) frequency dependence of χM″ values of 73R measured under 1.0 kOe external field. The solid curves are guides for the eye [57]. Copyright © 2011 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Crystals 14 00474 g019
Figure 20. Crystal structure of 75: (a) the structural unit, #1, y, x, 2 − z; #2, 0.5 − x, −0.5 + y, 1.75 − z; #3, −0.5 + y, 0.5 − x, 0.25 + z; (b) tadpole-like metal-organic subunit; (c) combined ball and stick and cartoon representation of the 1D right-handed helical chain (hydrogen atoms are omitted for clarity); (d) combined ball and stick and cartoon representation of the 3D open metal-organic framework [60]. Copyright © 2014, American Chemical Society.
Figure 20. Crystal structure of 75: (a) the structural unit, #1, y, x, 2 − z; #2, 0.5 − x, −0.5 + y, 1.75 − z; #3, −0.5 + y, 0.5 − x, 0.25 + z; (b) tadpole-like metal-organic subunit; (c) combined ball and stick and cartoon representation of the 1D right-handed helical chain (hydrogen atoms are omitted for clarity); (d) combined ball and stick and cartoon representation of the 3D open metal-organic framework [60]. Copyright © 2014, American Chemical Society.
Crystals 14 00474 g020
Figure 21. (top) The coordination environment of CeIII ion (left) and the simplified ball and stick model (right, the blue lines represent the salen type ligand) in the cation of 78 (H-atoms are omitted for clarity). Symmetry codes: (i) x, xy + 1, −z + 3/2; (ii) −x, −x + y, -z + 3/2; (iii) −z, −y + 1, z. (bottom) 3D coordination structure of 78 along the c axis (left) and schematic illustration of quartz topology with the Schläfli symbol {64·82} along the c axis (right, the free PF6 ions are omitted for clarity). Reproduced from Ref. [61] with permission from the Royal Society of Chemistry.
Figure 21. (top) The coordination environment of CeIII ion (left) and the simplified ball and stick model (right, the blue lines represent the salen type ligand) in the cation of 78 (H-atoms are omitted for clarity). Symmetry codes: (i) x, xy + 1, −z + 3/2; (ii) −x, −x + y, -z + 3/2; (iii) −z, −y + 1, z. (bottom) 3D coordination structure of 78 along the c axis (left) and schematic illustration of quartz topology with the Schläfli symbol {64·82} along the c axis (right, the free PF6 ions are omitted for clarity). Reproduced from Ref. [61] with permission from the Royal Society of Chemistry.
Crystals 14 00474 g021
Figure 22. Part of the chiral zigzag chain of 81, CCDC-1060179, [58]. Color code: Dy pink, Mo yellow, Cu green, O red, N blue, C grey. Symmetry operations: (′) 1 − x, 0.5 + y, 1 − z; (″) 1 − x, −0.5 + y, 1 − z; (‴) x, −1 + y, z.
Figure 22. Part of the chiral zigzag chain of 81, CCDC-1060179, [58]. Color code: Dy pink, Mo yellow, Cu green, O red, N blue, C grey. Symmetry operations: (′) 1 − x, 0.5 + y, 1 − z; (″) 1 − x, −0.5 + y, 1 − z; (‴) x, −1 + y, z.
Crystals 14 00474 g022
Figure 23. Part of the chiral zigzag chain of 86, CCDC-1060184, [58]. Color code: La pink, Mo yellow, Cu green, O red, N blue, C grey. Symmetry operations: (′) −x, −y, −z; (″) −x, y, 0.5 − z; (‴) −x, −y, 0.5 + z; (*) −x, −y, −0.5 + x; (**) −x, y, −0.5 + z; (***) x, y, −1 + z.
Figure 23. Part of the chiral zigzag chain of 86, CCDC-1060184, [58]. Color code: La pink, Mo yellow, Cu green, O red, N blue, C grey. Symmetry operations: (′) −x, −y, −z; (″) −x, y, 0.5 − z; (‴) −x, −y, 0.5 + z; (*) −x, −y, −0.5 + x; (**) −x, y, −0.5 + z; (***) x, y, −1 + z.
Crystals 14 00474 g023
Table 1. Structural, CD, and magnetic data for the homometallic 4f complexes 1–10.
Table 1. Structural, CD, and magnetic data for the homometallic 4f complexes 1–10.
Complex4f IonSpace GroupCoordination NumberCD, λmax
(nm)
χMT (r.t.) (cm3Kmol−1)Ueff (K)t0 (s)Ref.
1R/1SDyC28-14.4039.90 b3.62 × 10−6[32]
2R/2SDyP217-13.8413.27 c2.02 × 10−6[33]
3R/3STbP217−291, −391
+350 a
11.80--[33]
4R/4SHoP217-14.12--[33]
5RCeP438−343, +320 a---[34]
6RDyP218----[35]
7SErP218----[36]
8R/8SDyC27−281, −389, +348 a13.4724.61 c8.49 × 10−8[33]
9SSmC210----[37]
10RCeP19, 10----[38]
a the values correspond to the R-enantiomer, same values with opposite signs correspond to the S-enantiomer; b under 1.5 kOe external dc field; c under 2.0 kOe external dc field.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Raptopoulou, C.P. Chiral 4f and 3d-4f Complexes from Enantiopure Salen-Type Schiff Base Ligands. Crystals 2024, 14, 474. https://doi.org/10.3390/cryst14050474

AMA Style

Raptopoulou CP. Chiral 4f and 3d-4f Complexes from Enantiopure Salen-Type Schiff Base Ligands. Crystals. 2024; 14(5):474. https://doi.org/10.3390/cryst14050474

Chicago/Turabian Style

Raptopoulou, Catherine P. 2024. "Chiral 4f and 3d-4f Complexes from Enantiopure Salen-Type Schiff Base Ligands" Crystals 14, no. 5: 474. https://doi.org/10.3390/cryst14050474

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop