Next Article in Journal
Functional Characterization of Lobularia maritima LmTrxh2 Gene Involved in Cold Tolerance in Tobacco through Alleviation of ROS Damage to the Plasma Membrane
Next Article in Special Issue
Substituent-Guided Cluster Nuclearity for Tetranuclear Iron(III) Compounds with Flat {Fe43-O)2} Butterfly Core
Previous Article in Journal
A Functional Polymorphism Downstream of Vitamin A Regulator Gene CYP26B1 Is Associated with Hand Osteoarthritis
Previous Article in Special Issue
Investigations on the Influence of the Axial Ligand in [Salophene]iron(III) Complexes on Biological Activity and Redox Behavior
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

New Dinuclear Macrocyclic Copper(II) Complexes as Potentially Fluorescent and Magnetic Materials

by
Magdalena Barwiolek
1,*,
Dominika Jankowska
1,
Anna Kaczmarek-Kędziera
1,
Iwona Lakomska
1,
Jedrzej Kobylarczyk
2,
Robert Podgajny
3,
Paweł Popielarski
4,
Joanna Masternak
5,
Maciej Witwicki
6 and
Tadeusz M. Muzioł
1
1
Faculty of Chemistry, Nicolaus Copernicus University in Torun, Gagarina 11, 87-100 Torun, Poland
2
Institute of Nuclear Physics PAN, Radzikowskiego 152, 31-342 Kraków, Poland
3
Faculty of Chemistry, Jagiellonian University, Gronostajowa 2, 30-387 Kraków, Poland
4
Faculty of Physics, Kazimierz Wielki University, Powstancow Wielkopolskich 2, 85-090 Bydgoszcz, Poland
5
Institute of Chemistry, Jan Kochanowski University of Kielce, Uniwersytecka 7, 25-406 Kielce, Poland
6
Faculty of Chemistry, University of Wrocław, Joliot Curie 14, 50-383 Wrocław, Poland
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(3), 3017; https://doi.org/10.3390/ijms24033017
Submission received: 30 December 2022 / Revised: 25 January 2023 / Accepted: 26 January 2023 / Published: 3 February 2023
(This article belongs to the Special Issue Oligonuclear Metal Complexes with Schiff Base Ligands 2.0)

Abstract

:
Two dinuclear copper(II) complexes with macrocyclic Schiff bases K1 and K2 were prepared by the template reaction of (R)-(+)-1,1′-binaphthalene-2,2′-diamine and 2-hydroxy-5-methyl-1,3-benzenedicarboxaldehyde K1, or 4-tert-butyl-2,6-diformylphenol K2 with copper(II) chloride dihydrate. The compounds were characterized by spectroscopic methods. X-ray crystal structure determination and DFT calculations confirmed their geometry in solution and in the solid phase. Moreover, intermolecular interactions in the crystal structure of K2 were analyzed using 3D Hirshfeld surfaces and the related 2D fingerprint plots. The magnetic study revealed very strong antiferromagnetic CuII-CuII exchange interactions, which were supported by magneto-structural correlation and DFT calculations conducted within a broken symmetry (BS) framework. Complexes K1 and K2 exhibited luminescent properties that may be of great importance in the search for new OLEDs. Both K1 and K2 complexes showed emissions in the range of 392–424 nm in solutions at various polarities. Thin materials of the studied compounds were deposited on Si(111) by the spin-coating method or by thermal vapor deposition and studied by scanning electron microscopy (SEM/EDS), atomic force microscopy (AFM), and fluorescence spectroscopy. The thermally deposited K1 and K2 materials showed high fluorescence intensity in the range of 318–531 nm for K1/Si and 326–472 nm for the K2/Si material, indicating that they could be used in optical devices.

1. Introduction

Macrocyclic metal complexes with Schiff bases are of great interest to researchers, thanks to various properties of importance in design and synthesis of functional molecular materials [1,2,3,4,5,6,7]. Complexes with optically active Schiff bases, particularly binaphthyl-based macrocycles, stand out significantly, revealing optical properties to find applications as organic compound sensors, colourimetric sensors, or fluorescent sensors for chiral recognition [8,9,10,11,12]. One of the greatest applications of these compounds is catalytic activity in various organic reactions [13,14]. Moreover, the Schiff bases and their complexes exhibit biological properties, e.g., anticancer and antibacterial activity [15,16].
A large number of donor atoms, such as O, N, S, P, and supramolecular construction, afford sufficient space for coordination atoms to create mono-, di-, and polynuclear complexes [17,18,19]. Macrocycle architecture consisting of a combination of binaphthyl groups can also show fluorescent properties. 1,1’-Binaphthyl units can be functionalized at the 2- and 2’-positions and are often chiral due to restricted rotation about the transannular bond [20]. Thus, their role in asymmetric synthesis and as a ligand in supramolecular chemistry is not to be underestimated [21]. The number of donor atoms can be changed, and thus various types of complexes can be achieved. Recently, E. Chinnaraja et al. reported the one-pot synthesis of a series the macrocyclic complexes [2+2] consisting of dialdehyde: 4-methyl-2,6-diformylphenol or 4-tert-butyl-2,6-diformylphenol and chiral binaphthyl amine as building blocks. Cu(II) metal salt was used in these syntheses. The obtained compounds were enantiomerically pure, showed chiroptical properties, and exhibited catalytic activity in many reactions of organic transformations [22,23,24]. Great interest was shown in the design, synthesis, and modification of complexes due to their possible impact on the field of molecular magnetism [25,26]. The study of magnetic interaction between the central Cu(II) ions in multinuclear complexes is the main purpose of magneto-structural investigations [27,28]. Considering magnetic properties, dinuclear macrocyclic complexes with Cu(II) deserve to be recognized [29,30,31]. Consequently, new compounds with fluorescent and magnetic properties that could be used in thin films are still sought for. It is also known that Schiff base complexes could create thin materials by using wet methods, e.g., dip or spin coating, as well as vapor deposition techniques. The properties of the materials depend on the presence of substituents, their electronic nature, or the presence and size of the aromatic rings. The choice of suitable substrates for designing new materials is one of the most important factors that can determine the unusual properties of the received films and their application. There are still many questions regarding the fluorescence and magnetic properties of the compounds and thin materials. The appropriate methods for creating new films should also be developed. New materials can improve crucial parameters such as life and durability of the new films, high luminescence, and designs, as well as provide new unique characteristics of the new devices such as smartphones, OLEDS, or solar batteries. Therefore, we report the synthesis, structure, and magnetic and spectroscopic properties of binuclear macrocyclic Cu(II) complexes obtained in the one-pot reaction of (R)-(+)-1,1′-binaphthalene-2,2′-diamine and 2-hydroxy-5-methyl-1,3-benzenedicarboxaldehyde K1, or 4-tert-butyl-2,6-diformylphenol K2 with copper(II) chloride dihydrate. DFT calculations were also carried out to support the magnetic behavior and optical properties of the complexes. Hirshfeld analysis extensively shows intramolecular interactions. The new copper(II) complexes were used as precursors of thin layers in the spin coating technique and thermal vapor deposition. The morphology and composition of layers were analyzed by AFM and SEM/EDX microscopy, and the fluorescence properties of the thin materials were also studied. Both of the presented copper(II) complexes exhibited strong antiferromagnetic CuII-CuII exchange interactions.

2. Results and Discussion

2.1. Dinuclear Copper(II) Complexes: Synthesis and Characterization

Reactions of template synthesis in a molar ratio of 1:1:1 resulted in dinuclear macrocyclic copper(II) complex formation (Figure 1). The IR spectra exhibited the -C=N- stretching bands, characteristic for the Schiff bases (Figures S1 and S2). The elemental analysis and X-ray studies confirmed the purity of the obtained compounds (Figure 2, Figure 3, Figure 4 and Figure 5). The thermal stability of K1 and K2 complexes was studied by thermogravimetric analysis from ambient temperature to 1000 °C under air (Figures S3–S5). The final decomposition product was a mixture of copper(II) oxide and copper(I) oxide that was additionally confirmed by the XRD analysis (Figures S6 and S7).

2.2. Crystal Structure Description

2.2.1. Crystal Structure Description of [Cu2Cl2(L’)] K1

The structure of the [Cu2Cl2(L’)] (K1) (L’ = (R)-(+)-1,1′-binaphthalene-2,2′-diamine and 2-hydroxy-5-methyl-1,3-benzenedicarboxaldehyde) complex crystallized in the orthorhombic space group P212121 with all atoms in the general positions. Copper(II) cations are found in a heavily distorted pentacoordinated environment (τ5 = 0.47 and 0.51 for Cu1 and Cu2, respectively [32]) (Figure 2). The analysis of polyhedra carried out in SHAPE [33] showed that for Cu1, we cannot discriminate between trigonal bipyramid (STBPY =3.224) and square pyramid (SSPY = 3.522), whereas for Cu2, trigonal bipyramid (STBPY =1.882, SSPY = 2.859) should be selected. Coordination spheres of both copper(II) ions consist of two nitrogen atoms and two oxygen atoms from the macrocyclic ring and apical chloride anion. Both copper polyhedra share the common edge (O22 – O49) with deprotonated hydroxyl groups bridging both copper(II) ions separated by 3.109 Å. The longest bonds occurred for apical chloride anions (2.495(3) and 2.363(2) Å) (Table 1). Cu-N/O bonds ranged from 1.917(11) to 2.034(10) and from 1.923(7) to 2.069(9) Å for Cu1 and Cu2, respectively. They were similar to those found in the polymorph (1.916(2)–2.078(4) Å) with significantly larger cells and lower density [24]. This is the only structural report for copper complexed in binaphthyl macrocyclic system. Mutual orientation of C1 and C11 (69.1°) as well as C28 and C38 (75.1°) naphthyl rings imposes the chirality of the complex. The Flack parameter shows that pure enantiomer was obtained. The dihedral angle between C22a and C49a phenyl rings was 33.1°.
In the crystal network, there are voids accounting for 1231.5 Å3 (22.7% of the volume cell) (Figure 3). The solvent density was poorly defined, and hence the bypass procedure was performed in Olex2 [34], resulting in a much better model. Nevertheless, the interaction model suffered from missing complex–solvent interactions. Therefore, we did not apply the Hirshfeld approach for intermolecular interactions. We limited our description only to complex–complex contacts. They were assured mainly by π–π and C–H…π interactions. The pore structure was also stabilized by a C71–H71B…Cl1[-x, 1-y, z] hydrogen bond.

2.2.2. Crystal Structure Description of [Cu2Cl2(L)]·H2O K2

The structure of the [Cu2Cl2(L″)]·H2O K2 (L″ = (R)-(+)-1,1′-binaphthalene-2,2′-diamine and 4-tert-butyl-2,6-diformylphenol) complex crystallized in the chiral orthorhombic P212121 space group with all atoms in general positions and the whole molecule given by the formula in the asymmetric unit. Copper(II) cations were found in a pentacoordinated environment with geometrical indices pointing at a very significantly distorted coordination sphere (τ5 = 0.61 and 0.46 [32] for Cu1 and Cu2, respectively) (Figure 4). The analysis performed in SHAPE [33] software indicated that Cu1 should rather be described as a trigonal bipyramid (STBPY = 1.542, SSPY = 3.424), but for Cu2, it did not discriminate unequivocally between the trigonal bipyramid (STBPY 2.255) and square pyramid (SSPY = 2.236). Both coordination polyhedra shared the common edge due to bridging O22 and O49 oxygen atoms, with Cu1-Cu2 separation (3.090 Å) being even slightly shorter than in K1. Hence, both for K1 and K2 we can expect significant magnetic interactions. The coordination spheres consisted of two nitrogen atoms and two oxygen atoms from the macrocyclic ring and an apical chloride anion. Cu-N bond lengths ranged from 1.951(6) to 2.077(5) Å, Cu-O bonds were from 1.903(5) to 2.045(5) Å, and Cu-Cl were from 2.302(3) to 2.370(2) Å (Table 2). The naphthyl rings were inclined by 68.1 and 77.5° for C1/C11 and C28/C38 angles, respectively, whereas the angle between C22a and C49a moieties was 30.0°. The mutual orientation of these systems imposed the chirality, and the Flack parameter indicated a pure enantiomer.
The packing show channels running along the b axis accounted for 461.2 Å3 (8.3% of the cell volume) (Figure 5). They were filled with water molecules forming C52-H52…O81[1-x, -1/2+y, 1/2-z] hydrogen bonds. However, there were still some voids (319.1 Å3, 5.8%), indicating that only the O81 water molecule was well defined, whereas there was probably disordered solvent filling those cavities. Hence, fingerprints spanned high di and de values (Figure 6). The wall of the cavity was formed by two complex molecules. In the network, weak dispersion interactions (H…H and C…H) prevailed due to C-H…π and π…π interactions between strongly inclined aromatic rings (Figure 6). The interaction landscape was completed by Cl…H interactions, with chloride being involved in intermolecular C6-H6…Cl1[-1/2+x, 1/2-y, 1-z] and C32-H32…Cl2[1+x, y, z] hydrogen bonds (red spot on the Hirshfeld surface).

2.3. DC Magnetic Measurements, BS DFT Computations, and EPR Spectra

The χMT and M(H) curves for compounds K1 and K2 are presented in Figure 7. At T = 300 K, the χMT product for both compounds was close to 0.17 cm3 K mol–1, being a value around five times smaller than the 0.867 cm3 K mol–1 expected for two uncoupled Cu(II) centers assuming gCu = 2.15 within the range typical for 5-coordinative Cu(II) complexes. With the temperature decreasing to around 100 K, the χMT products decreased to reach a steady value of 0.044 and 0.047 cm3 K mol–1 for K1 and K2, respectively. Below 10 K, both curves decreased afresh to reach values slightly above 0.040 cm3 K mol–1. The shapes of χMT were interpreted in terms of very strong magnetic exchange coupling between the Cu(II) centers in the Cu2O2 coordination cores through the phenoxo-bridges. The low-temperature part of χMT, visibly larger than 0, indicated the presence of paramagnetic contribution, which was also confirmed by the M(H) curves.
The χMT versus T curves were fitted in the 1.8–300 K temperature range using the latest PHI software (version 3.1.6, simplex method) [35] and the following Hamiltonian:
H ^ = 2 J C u C u ( S ^ C u 1 · S ^ C u 2 ) + μ B H g C u , a v g ( S ^ C u 1 · S ^ C u 2 )
where only the magnetic exchange constant between Cu(II) centers, JCu-Cu, and the average gCu,avg factor for both Cu sites were fitted. The additional magnetic impurities, IMP, were included, employing analytical expression for the temperature-dependent magnetic susceptibility, assuming the mononuclear Cu2+ complex with g = 2.0. The impurity value represents the fraction of the measured sample K1 or K2, χ = (1-IMP) χsample + (IMP) χIMP. The obtained values are presented in Table 3.
The obtained JCu-Cu values were in line with the Cu-O-Cu angles, 101.12 and 107.28 deg (av. 104.20 deg) for K1, and 99.92 and 107.96 deg (av. 103.94 deg) for K2 according to archetypal magneto-structural correlation for rhombus Cu2O2 cores involving hydroxo-, alkoxo-, and phenoxo-bridges [36]. Rather a minor difference of around 25 cm–1 between JCu-Cu values of K1 and K2 (according to the frames of the above correlation) might conveniently be interpreted in terms of minute differences in the Cu-O distances, Cu-O-Cu angles, and torsion angles within the Cu2O2 cores (see crystallographic tables). These results were further nicely supported by the broken symmetry and spin flip DFT computational data (Table 3 and Table S1), wherein the J(SUP) [37] values close to −337.0 cm−1 (K1) and −346.5 cm−1 (K2) estimated according to the spin unprojected (SUP) approach, recommended for the strong coupling regime, provided the best illustration of the JCu-Cu convergence.
The presence of paramagnetic contribution was confirmed by EPR experiments. The powder spectra recorded for K1 and K2 at 77 K, shown in Figure S8, were characteristic for paramagnetic Cu(II) centers. They were successfully simulated with gz = 2.230, gy = 2.072, gx = 2.072, and Az = 177 × 10−4 cm−1 for K1 and gz = 2.232, gy = 2.091, gx = 2.052, and Az = 173 × 10−4 cm−1 for K2. The parameters observed for the two complexes were similar and indicated the axial N2O2 coordination environment, which was far less distorted in comparison with dinuclear units, as well as the unpaired electron occupying the molecular orbital with strong contribution from dx2-y2 [38,39,40]. Thus, it is suggested that the paramagnetic contribution was due to the fractional incorporation of the monometallic complex of the CuII[CuIIvacancy]L2 or CuIICl[CuIIvacancy]L(LH) composition in the crystal, represented by the IMP contribution to the simulation of χMT(T) data. In comparison to the full model CuII2Cl2L2 complex, such a moiety might contain one empty N2O2 pocket, denoted as [Cuvacancy]. Such a solution relies on the incorporation of around one monometallic complex per eight to nine CuII2Cl2L2 moieties, and is pretty plausible, considering the fact that topological shapes of both moieties are sufficiently similar to each other to accommodate the resultant structural defects without the loss of the crystal stability. The above assumption is supported by the fact that optical inspection of the crystals of K1 and K2 did not indicate the presence of a separate phase, other than the examined crystals. Moreover, crystal structure solution and refinement allow for the presence of some level of such defects without the loss or even with a minute increase in the solution and refinement quality. The presence of paramagnetic “impurities” was previously observed in dinuclear S(1/2)-S(1/2) Cu2 systems [41,42,43].
The powder EPR spectra of the K1 and K2 complexes recorded at 350 K, at which temperature the S = 1 state is highly populated, showed broad, nearly isotropic lines centered at geff = 2.14 and 2.13, respectively. The peak-to-peak line widths were about 360 G for K1 and 340 G for K2 (Figure S9). The broadening of EPR spectra and lack of the resolved g tensor anisotropy were caused by spin–spin interactions between relatively close Cu(II) ions arranged in dimeric units in both complexes. Because the spectra did not undergo the resolution of the signals resulting from the resonance transition between the spin states ΔMS = ±1.0, neither the tensor g components nor the zero-field splitting D parameter can be determined.

2.4. UV–VIS and Fluorescence Spectroscopy

The UV–VIS absorption and fluorescence spectra of the copper(II) complexes were recorded at room temperature in solvents of different polarities: chloroform (εr = 4.9), acetone (εr = 20.6), methanol (εr = 32.7), acetonitrile (εr = 35.9), and DMSO (εr = 46.5) [44] (Figure 8, Table S2). Due to the limited solubility of K1 in acetonitrile, we were unable to record the UV–VIS spectrum in MeCN.
In the UV–VIS spectra of the K1 and K2 complexes, bands in the range of 272–290 nm connected with π → π* transitions in the aromatic rings and the bands from the n → π* transitions of the azomethine group between 322 and 324 nm for K1 and 322 and 328 nm for K2 were recorded. Bands originating from π → π* transitions in the aromatic rings are typical for this kind of compound and were present in the ligand expected range, and thus we do not show them in Figure 7; only the bands from the charge transfer and d-d transitions are shown.
Band maxima related to the ligand-to-metal charge transfer transition (LMCT) were observed in the range of 388 - 408 nm in a solvent at different polarity: methanol (388 nm), DMSO (398 nm), acetone (402 nm), and chloroform (408 nm) for K1, and methanol (388 nm), acetonitrile (392 nm), DMSO (394 nm), acetone (400 nm), and chloroform (404 nm) for K2 [30]. For both compounds, the bathochromic shift of the LMCT band with a decrease in the solvent polarity (except methanol) was noted (Figure 8, Table S2). This can be connected with the polar character of methanol, which is a protic solvent and causes a different type of interaction with compounds in comparison with the rest of the aprotic solvents [32,45]. Moreover, this can also originate from the distortion of the molecule geometry in the excited state, which implies an increase in the resonance energy and bathochromic shifts.
Additionally, in the absorption spectra, the presence of low-intensity bands at 660 nm for K1 and 664 nm for K2 absorption bands from d-d transitions [46] only in the less polar solvent chloroform were noted. This results from the square pyramidal or trigonal bipyramid environment of the copper atoms shifted towards the apical chloride anion [23,47]. The same was observed previously for similar copper(II) compounds [41,48].
Furthermore, the UV–VIS spectra of the compounds were also registered at room temperature in the solid state (Figure 9). The spectra show band maxima in the range of 265–266 nm related to π → π* transitions of the aromatic groups and between 385 and 387 nm related to n → π* transitions of the azomethine groups. Moreover, at 468 nm and 461 nm, bands connected with the ligand-to-metal charge transfer LMCT π → d transitions were observed. Above 700 nm for K1 and 720 nm for K2, bands from d-d transitions appeared. The band maxima were shifted towards higher wavelengths by 57–80 nm (LMCT π → d), 63–65 nm (n → π*C=N), and 40- 56 nm (d-d) in comparison to the solvents. This shift is connected to the increase in the rigidity of the complexes’ structures in the solid state in comparison to the solution. A red shift of the bands was noted in the spectra of other copper(II) complexes [49].

Emission Studies

The excitation of K1 and K2 in all solvents at 350 nm resulted in blue emission between 398 and 424 nm for K1 and 392 and 424 nm for K2 (Figure 10). Emission band maxima of K1 and K2 exhibited bathochromic shifts with increasing solvent polarity. Moreover, the highest fluorescence intensity in most polar solvents such as DMSO and methanol was noted (Figure 10, Table S2). A similar situation also occurred in the case of other metal(II) complexes [23]. When the emission spectra registered in a solution and in the solid state were compared, it was possible to infer that the solvent destroyed the π–π interactions, and thus the transition energy was increased in the solution, as was observed previously [50].

2.5. Circular Dichroism

The chiral character of the studied complexes is reflected in their CD spectra (Figure 11).
The two studied complexes had similar CD spectra with the signals at 314 nm(+24) and 365 nm(−198) for K1, and at 311 nm (+18) and 358 nm (−198) for K2. The Cotton effect noted above 400 nm at 406(+200) and 444(−174) for K1, and at 403(+198) and 440(−180) for K2 were connected with n → π*C=N transitions. Conversely, the bands observed at 589 nm for K1 and 584 nm for K2 from d-d transitions in Cu ions were a result of a distorted square-pyramidal or trigonal pyramidal geometry [24,51]. The low intense band corresponding to the d–d transitions largely exhibited the same position (≈585 nm) and even a similar intensity for both complexes. This feature can be explained by the existence of an identical {CuN2O2Cl} chromophore. The replacement of a tert-butyl group with a methyl group in the ring did not affect the optical activity of the complexes, which was a consequence of the similarity of the geometry of the studied compounds [52]. Moreover, the signs of the CD bands in the spectra of both complexes K1 and K2 were the same because they were derivatives of the same enantiomer of diamine ((R)-(+)-1,1′-binaphthalene-2,2′-diamine) [52,53], as was observed for a series of copper(II) complexes with chiral tri- and tetradentate Schiff base ligands derived from 1,1’-binaphthyl-2,2’-diamine [14].

2.6. Theoretical Calculations

The vertical absorption spectrum estimated within the ωB97X-D/def2-SVP/PCM(CHCl3) approach exhibited small intensity signals above 640 nm, corresponding to the d-d* transition forbidden by the Laporte rule (see Figure 12 and the corresponding natural transition orbitals in Figure 13a). The band of about 480 nm involved the transition between the metal and the ligand (the corresponding natural orbitals are presented in Figure 13b). The signals appearing below 400 nm solely arose from the π→π* excitations in the ligand. The shape of the spectrum was only mildly affected by the change of the substituent in the ligand macrocyclic ring from methyl in K1 to tert-butyl in K2. The computational results confirmed the assignment provided above on the basis of the experimental measurements in solution.

2.7. Thin Materials of Copper(II) Complexes

The morphology and roughness of the thin layers were examined by SEM and AFM techniques. To test the chemical composition of materials, the EDS analysis was recorded (Figure 14d, Figure 15, Figure 16, Figures S10 and S11). The optimal parameters of the layers (roughness, parameters, and homogeneity) were obtained in a multi-stage centrifugation spin-coated process using particular parameters: 2500 rpm to 3000 rpm, time of coating 5 or 10 s. Moreover, the thermal vapor deposition method was also used as a second technique to achieve thin materials. The two-dimensional (2D) and three-dimensional (3D) AFM images scanned over a surface area of 1 × 1 µm2 are shown in Figure 14, Figure 17 and Figure S12. The values of roughness parameters of materials obtained by thermal deposition were as follows: K1/Si Ra = 10.7–13.1 nm and Rq = 13.2–17.1 nm, and K2/Si Ra = 3.91–4.96 nm and Rq = 4.02–6.38 nm. The roughness of the spin-coated films was similar to that obtained by thermal deposition. However, the spin-coated materials were thinner than those achieved by thermal vapor deposition. A similar situation was noted by us in the case of the layers L1/Si or L2/Si of the macrocyclic Schiff bases derived from o-phenylenediamine and 2-hydroxy-5-methylisophthalaldehyde L1 or 2-hydroxy-5-tert-butyl-1,3-benzenedicarboxaldehyde L2 [54].
The values of the roughness parameters for thin materials obtained by both spin-coating and by thermal vapor deposition indicate the achievement of smooth, thin films of copper(II) complexes. Moreover, SEM/EDS analysis showed the presence of carbon, nitrogen, oxygen, and copper in the layer (Figures S10 and S11). SEM/EDS, together with mapping analysis, confirmed the composition of the new materials.
The new films obtained by thermal vapor deposition were also characterized by IR DRIFT (Figure S13). The analysis of the IR DRIFT data showed the presence of the characteristic for the Schiff base peaks between 1653 and 1647 cm−1 from stretching frequencies of the azomethine group, and bands from stretching vibrations of aromatic rings νC=CAr in the region 1568–1485 cm−1 were registered. The above-described bands confirmed the presence of the deposited compounds in the obtained materials.

Fluorescence Properties of the Materials

The fluorescence properties of the thin materials were also studied. The height (thickness) of the spin-coated materials was less than half that of thermally deposited films (70 nm K1/Si and 24 nm K2/Si), which influenced the emission properties of the composites. The spin-coated K1/Si materials did not exhibit fluorescence. Conversely, in the case of K2/Si film (Figure 18), the high intensity of the emission bands (λex = 320 nm) was observed.
The thin thermally deposited materials showed fluorescence in the range of 318–531 nm for K1/Si and 326–472 nm for the K2 material, λex = 250 nm. Furthermore, the emission bands were broad and split into three components (Figure 19, Table S3). The highest intensity of the emission bands for the smooth, thin layer with the equally distributed complexes on the Si surface was noted (Figure 18 and Figure 19). The bathochromic shift of the emission bands of the films in comparison to the solutions was noted. Red shifting of emission maxima was observed for most fluorescent compounds in the solid state, probably due to π–π stacking of the aromatic rings in the molecules [49,54]. An influence of molecular packing in the solid phase on the optical properties can therefore be concluded. This can arise from a different pathway of non-radiative transitions. It can be related to the reduction of the ligand conformational flexibility in the complex. This reduction results from the restraints imposed by the substrate surface. Layers obtained by thermal deposition methods are intended to be used as optical materials.

2.8. XAS Analysis

The normalized spectra of K1 and K2 were very similar due to the structural similarity of copper coordination spheres in both complexes. Those spectra showed two features for L3 and another two for the L2 edge (Figure 20, Table 4). In the case of the L3 edge, they occurred at 931.0 and 934.8 eV for K1 and 931.0 and 934.7 eV for K2, and the L2 features were shifted by around 20 eV towards higher energies. Those values were common for Cu(II) compounds [55,56,57]. According to the theory, the intensity of L3 peaks was around twofold stronger than in the case of L2, and only one peak related to 2p→3d transitions was expected [58,59,60]. However, additional features are very often observed due to 2p→4s transitions [61,62,63]. Those additional peaks are usually around 25 times weaker due to the smaller cross-section for such X-ray absorption. DeBeer George et al. showed for copper dimers with macrocyclic thiolate ligands and copper ions separated by around 2.9 Å that the L energy pattern depends on the effective nuclear charge the ligand field splitting and also possible metal–metal bonding [64]. Muzioł et al. suggested also that a more complex pattern of the L edge might result from strongly distorted coordination spheres [56]. In the reported structures, we observed such a deformation (see Section 2.2), and the Cu-Cu distance was around 3.1 Å. Hence, we observed much weaker signals following the main features of the L3 and L2 edges.

3. Materials and Methods

2-Hydroxy-5-methylisophthalaldehyde (97%), 2-hydroxy-5-tert-butyl-1,3-benzenedicarboxaldehyde (97%), (R)-(+)-1,1′-binaphthyl-2,2′-diamine (99%), and trimethylamine (99.5%) were purchased from Aldrich (Warsaw, Poland), and used without further purification. Copper(II) chloride dihydrate (analytical grade) was supplied by POCh Gliwice, Poland.

3.1. Methods and Instrumentation

UV–VIS absorption spectra were recorded in chloroform, acetone, DMSO, acetonitrile, and methanol (3.323 × 10−6 M) solutions on a Hitachi spectrophotometer. The fluorescence spectra were recorded on a spectrofluorometer Gildenpλotonics 700 (Dublin, Ireland) in the range 900–200 nm (grating 1, bandpass 5 and 8, integration time 100 ms, chloroform, acetone, DMSO, acetonitrile, and methanol solution of compounds the same as in the case of the UV–VIS studies or silicon slides). The elemental analysis was carried out using a Vario EL III Elemental analyzer. The thermal analysis (TG, DTG, DTA) was performed on an SDT 2960 TA analyzer under air, a heating rate of 10 °C min−1, and a heating range of up to 1000 °C and a Jupiter STA 449 F5 thermoanalyzer from Netzsch (Selb, Germany) with an automatic sample feeder coupled to a Vertex 70V FT-IR spectrometer from Bruker Optik (Ettlingen, Germany). After combustion, the residue of the sample was analyzed by an XRD analysis performed with a Philips X’Pert equipped with an X’Celerator Scientific detector. The IR spectra were recorded on the Bruker instrument using the ATR technique in the range of 70–4000 cm−1. Circular dichroism spectra were recorded with a Jasco J-815 spectropolarimeter (Jasco Inc.) in the range of 310–700 nm wavelengths. The solution of K1 and K2 complexes (≈1 × 10−4 M) was prepared by dissolving it in a CHCl3 solution.

3.1.1. Spin Coating

Layers of the complexes were deposited on Si(111) wafers (10 nm × 10 mm) that were ≈500 nm thick using the spin coating technique. Precursors were dissolved in chloroform and deposited on Si using a spin coater (Laurell 650 SZ). The spin speed varied from 2500 to 3000 rpm, and the coating time was 5 or 10 s.

3.1.2. Thermal Vapor Deposition

The thin layer of K1 and K2 was deposited on n-type silicon substrate. The orientation of the silicon substrate was (100) with electrical resistivity (ρ) equal to 6.2 × 10−3 Ω cm. The silicon wafer was first degreased in acetone, ethanol, and finally in deionized water using an ultrasonic bath. On the front side (polished side) of the silicon wafer, a K1 and K2 layers of 24–70 nm thickness were deposited in a vacuum (p = 2 × 10−4 Pa) by a thermal evaporation method, with an evaporation rate of 0.2 nm/s, without heating of the substrate.
The morphology and composition of the obtained films were analyzed with a scanning electron microscope (SEM; LEO Electron Microscopy Ltd, Cambridge, UK), the 21430 VP model equipped with secondary electrons (SE) detectors, and an energy-dispersive X-ray spectrometer (EDX) Quantax with an XFlash 4010 detector (Bruker AXS microanalysis GmbH, Berlin, Germany). The atomic force microscopy (AFM) images were performed in the tapping mode with a Multi Mode Nano Scope IIIa (Veeco Digital Instrument, SB, US) microscope. The structure of the produced layers was estimated using diffuse reflectance infrared Fourier transform spectroscopy (DRIFT, Spectrum 2000, PerkinElmer Inc., Waltham, MA, USA). The absorption spectra of the solid samples K1 and K2 were recorded on the Jasco V-750 spectrophotometer in the 200–800 nm (bandwidth 1.0 nm, data interval 0.2 nm).

3.2. Crystal Structure Determination

The diffraction data of the studied compounds were collected for the single crystal at 100 K using an XtaLAB Synergy Dualflex equipped with a HyPix detector and CuKα source (λ = 1.54184 Å) for K1 and on BL14.2 beamline (Helmholtz Zentrum Berlin, Bessy II) operating at λ = 0.7999 Å for K2. For K1, the full process of the data reduction was performed in CrysAlis Pro [65], whereas for K2, the data reduction and space group determination were performed with xdsapp [66,67], and then CrysAlis Pro was used for final data reduction. For K2, the absorption correction was introduced by the diffabs method implemented into WinGx [68]. The structure was solved by the direct methods and refined with the full-matrix least-squares procedure on F2 (SHELXL-2018/1) [69]. All heavy atoms were refined with anisotropic displacement parameters. Hydrogen atoms were located at calculated positions with thermal displacement parameters fixed to a value of 20% or 50% higher than those of the corresponding carbon atoms. It should be noted that for K1, the bypass procedure implemented into Olex2 [34] was applied due to poorly defined density in the solvent region of the porous complex. It resulted in significant amelioration of the final model. However, the whole interaction model suffered from missing solvent–complex and solvent–solvent interactions. In K2, the O81 water molecule was refined with partial occupancy (0.5), and the O21 molecule was located very close to the inversion center. In the final model of K2, there were missing hydrogen atoms from the O81 water molecule. The stable refinement was achieved with ISOR and DELU restraints for the positionally disordered C(61)H3 methyl group in K1 and for (ISOR) for C62, C64, and C74 atoms from tert-butyl groups of K2. All figures were prepared in DIAMOND [70] and ORTEP-3 [71]. The results of the data collection and refinement are summarized in Table 5.

3.3. SQUID Measurements

Magnetic measurements were investigated using a Quantum Design MPMS 3 SQUID magnetometer. For the measurements, powdered compounds K1 and K2 were sealed in two foil bags. The dc magnetic susceptibilities were measured in the 300–1.8K temperature range with an applied field of 5 kOe. The measurements were carried out in sweep mode with a scan rate of 2 K min–1. The magnetization curves were measured in the 0–70 kOe range with dc field stabilization. Magnetic data were corrected for the diamagnetic contribution of the plastic bags and sample by empirical and Pascal’s constants, respectively [72]. Moreover, the TIP contribution of 5.9 × 10−5 cm3 mol−1 for Cu(II) metal ion was applied.

3.4. EPR Measurements

EPR experiments were carried out for powder samples using a Bruker Elexsys E500 spectrometer operating at ∼9.6 GHz (X-band) frequency. The spectrometer was equipped with an NMR teslameter and a frequency counter. The temperature was controlled by using a finger-Dewar for measurements at 77 K and a Bruker ER 4131VT variable temperature accessory for 350 K. We set the amplitude and frequency of the modulating field to 10 G and 100 kHz, respectively, and we set the microwave power to 20 mW. The EPR spectra were simulated using EasySpin 5.2.35 [73,74].

3.5. Theoretical Calculations

The full geometry optimization of K1 and K2 was performed within the ωB97X-D/def2-SVP approach in a vacuum for magnetic coupling analysis and in the PCM model of chloroform for photophysical properties, starting from the crystal structure. The character of the stationary points was confirmed with the harmonic vibration analysis. Vertical absorption was investigated within the same approach, and ECD spectra were depicted for the analyzed enantiomers. The corresponding natural transition orbitals were examined in order to determine the character of the most intensive transitions. All these calculations were carried out with the Gaussian16 program [75]. The magnetic coupling parameters of the complexes were estimated within the broken symmetry approach and spin-flip formalism in B3LYP functional, according to the recommendations of Neese et al. [76] for the gas phase optimized geometry of both complexes. Moreover, the moderate-size def2-SVP basis set was applied for these calculations for the light atoms, as it has been proven to provide a good cost-to-quality balance for large systems containing copper atoms, and the basis set for copper was extended to the def2-QZVPP one [76]. Moreover, the spin-flip formalism was also employed with the same functional and basis set, for comparison. The magnetic coupling constants J were estimated according to spin-unprojected (SUP) scheme proposed by Ruiz for the strong coupling regime [37,77]. The corresponding calculations are based on the total spin-coupling Hamiltonian of the form -2J(Cu1-Cu2)SCu1SCu2, where J(Cu1-Cu2) is the exchange coupling constant between the two copper atoms with total spins SCu1 and SCu2, individually. Additionally, the spin-projected scheme and approximated spin-projected scheme were also applied for comparison, and the corresponding data are presented in the Supporting Information. All the magnetic coupling constant calculations were carried out with the Orca program package [78].

3.6. XAS

X-ray absorption spectra were recorded at the National Synchrotron Radiation Centre SOLARIS at the bending magnet PIRX beamline equipped with a collimated Plane Grating Monochromator for a copper L2,3-edge (910–1040 eV). The sample was finely ground and attached to double-sided adhesive conductive graphite tape. The measurements were performed with the step size of 0.25 eV for the pre-edge region, 0.15 eV for the edge regions, and 0.5 eV for the high energy part. The data sets were collected at room temperature in an ultra-high vacuum (UHV) using total electron yield mode (TEY). The data were processed using the ATHENA program from the Demeter package [79].

4. Experimental

Synthesis of Complexes

  • K1
A total of 0.5 mmol of 2-hydroxy-5-methylisophthalaldehyde, 0.5 mmol of (R)-(+)-1,1′-binaphthalene-2,2′-diamine, 0.5 mmol of copper(II) chloride dihydrate, and excess of triethylamine were dissolved in 80 cm3 of methanol. The synthesis was carried out under reflux for 1 hour. The product was dried under air, and single brown crystals were received. (yield: 94%). m. p. > 350 °C. C58H38Cl2Cu2N4O2x4 H2O (calc./found %): C 63.73/64.08, N 5.12/5.16.
Selected FT-IR (data reflectance, crystal) (cm−1), 3053, 3039, 2951, 2921 νC-HAr, 1616 νC=N, 1540, 1502, 1467νC=CAr, 1319 νPh-O (Figure S1).
  • K2
A total of 0.5 mmol of 4-tert-butyl-2,6-diformylphenol, 0.5 mmol of (R)-(+)-1,1′-binaphthalene-2,2′-diamine, 0.5 mmol of copper(II) chloride dihydrate, and excess of triethylamine were dissolved in 80 cm3 of methanol. The synthesis was carried out under reflux for 1 hour. The product was dried under air, and single brown crystals were crystalized in slow evaporation using a mixture of acetonitrile/chloroform and determined by crystal analysis. (yield: 90%). m. p. > 350 °C. C64H52Cl2Cu2N4O3 (calc./found %): C 68.44/68.69, N 4.99/5.12.
Selected FT-IR (data reflectance, crystal) (cm−1), 3050, 3040, 2952, 2865 νC-HAr, 1675 νC=N, 1581, 1502, 1466νC=CAr, 1322 νPh-O (Figure S2).

5. Conclusions

The research presented herein describes the synthesis of two new macrocyclic dinuclear copper(II) complexes, with N4O2 donor atoms. The Cu(II) complexes K1 and K2 were prepared by the template reaction of (R)-(+)-1,1′-binaphthalene-2,2′-diamine and 2-hydroxy-5-methyl-1,3-benzenedicarboxaldehyde K1, or 4-tert-butyl-2,6-diformylphenol K2 with copper(II) chloride dihydrate. The X-ray data show that copper(II) cations are in a heavily distorted pentacoordinated environment. The coordination spheres consist of two nitrogen atoms and two oxygen atoms from the macrocyclic ring and apical chloride anion. In the case of K1, the voids accounting for 1231.5 Å3 (22.7% of the volume cell) were noted, whereas for K2, channels running along the b axis accounted for 461.2 Å3 (8.3% of the cell volume) filled with water molecules forming C52-H52…O81[1-x, -1/2+y, 1/2-z] hydrogen bonds existed. The magnetic studies showed very strong antiferromagnetic CuII-CII exchange interactions represented by JCu-Cu = −305.66 ± 0.05 cm−1 for K1 and JCu-Cu = −328.45 ± 0.16 cm−1 for K2, which were nicely correlated with structural data and confirmed by the broken symmetry DFT calculations. The EPR spectra of the complexes K1 and K2 showed broad, nearly isotropic lines centered at geff = 2.14 and 2.13, respectively. The broadening of EPR spectra and lack of the resolved g tensor anisotropy were caused by spin–spin interactions between relatively close Cu(II) ions arranged in dimeric units in both complexes. The compliance between experimental and theoretical results validated the developed calculation method, which will be used to design new binuclear copper(II) complexes. The obtained complexes exhibited emission in the various polarity solvents and in the solid state. The blue emission between 398 and 424 nm for K1 and 392 and 424 nm for K2 in a solvent at different polarities was observed. The bathochromic shift with increasing solvent polarity for K2 was also observed. The highly ordered materials were obtained by spin-coating and thermal vapor methods, exhibiting fluorescence in the range of 318–531 nm for K1/Si and 326–472 nm for the K2/Si material, λex = 250 nm (thermally deposited films). The emission bands were broad and split into three components. The bathochromic shift of the fluorescence bands of films in comparison to that of the solution was noted. This can result from molecular packing in the solid state being different to that in the solution. The fluorescence emission of the layers makes these films potentially suitable for application in light-emitting devices.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms24033017/s1.

Author Contributions

Performing all the theoretical calculations, description of the applied computational methodology, and discussion of the obtained DFT results, A.K.-K.; crystallography, T.M.M.; performing the experiments, recording the UV–VIS and fluorescence spectra in various solvents, and UV–VIS and fluorescence data collection, D.J.; thermal deposition of the films, P.P.; magnetic measurements and interpretation of the data, R.P.; magnetic measurements and interpretation of the data, J.K.; EPR measurements and interpretation of the data, M.W.; financial support, I.L.; description of the obtained results and conceptualization, supervision, project administration, and manuscript writing, M.B.; manuscript editing (all authors). All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by grant4students IDUB project 4101.00000070 Dominika_Jankowska 20/2022 and Debiuty_3_Barwiołek_Magdalena IDUB project 4101.00000066.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in Supplementary Materials.

Acknowledgments

Part of this study was realized under grant4students IDUB project 4101.00000070 Dominika_Jankowska 20/2022 and Debiuty_3_Barwiołek_Magdalena IDUB project 4101.00000066. Tadeusz M. Muzioł thanks the University Centre of Excellence “Astrophysics and Astrochemistry” for research funding. Diffraction data were collected on MX14-2 at the BESSY II electron storage ring operated by the Helmholtz Zentrum Berlin [80]. We would particularly like to acknowledge the help and support of Frank Lennartz during the experiment. Special thanks are expressed to all members of the PIRX beamline team, especially M. Zając and Ewa Partyka-Jankowska, at the National Synchrotron Radiation Centre Solaris, Jagiellonian University, ul. Gronostajowa 7/P-1.6, 30-387 Cracow, Poland, for sharing their knowledge and technology. The maintenance and service costs of the SQUID magnetometer has been supported by a grant of the Faculty of Chemistry under the Strategic Programme Excellence Initiative at Jagiellonian University.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liu, X.; Hamon, J.R. Recent developments in penta-, hexa- and heptadentate Schiff base ligands and their metal complexes. Coord. Chem. Rev. 2019, 389, 94–118. [Google Scholar] [CrossRef]
  2. Gregoliński, J.; Ślepokura, K.; Kłak, J.; Witwicki, M. Multinuclear Ni(II) and Cu(II) complexes of a meso 6 + 6 macrocyclic amine derived from trans-1,2-diaminocyclopentane and 2,6-diformylpyridine. Dalton Trans. 2022, 51, 9735. [Google Scholar] [CrossRef] [PubMed]
  3. Vigato, P.; Tamburini, S.; Bertolo, L. The development of compartmental macrocyclic Schiff bases and related polyamine derivatives. Coord. Chem. Rev. 2007, 251, 1311–1492. [Google Scholar] [CrossRef]
  4. Chang, F.-F.; Zhang, K.; Huang, W. Schiff-base macrocyclic Zn(II) complexes based upon flexible pendant-armed extended dialdehydes. Dalton Trans. 2018, 48, 363–369. [Google Scholar] [CrossRef] [PubMed]
  5. Löffler, M.; Gregoliñski, J.; Korabik, M.; Lis, T.; Lisowski, J. Multinuclear Ni(II), Cu(II) and Zn(II) complexes of chiral macrocyclic nonaazamine. Dalton Trans. 2016, 45, 15586–15594. [Google Scholar] [CrossRef] [PubMed]
  6. Das, M.; Mukherjee, S.; Koley, B.; Choudhuri, I.; Bhattacharyya, N.; Roy, P.; Samanta, B.C.; Baraif, M.; Maity, T. Developing novel zinc(II) and copper(II) Schiff base complexes: Combined experimental and theoretical investigation on their DNA/protein binding efficacy and anticancer activity. New J. Chem. 2020, 44, 18347–18361. [Google Scholar] [CrossRef]
  7. Zhang, J.; Xu, L.; Wong, W.-Y. Energy materials based on metal Schiff base complexes. Coord. Chem. Rev. 2018, 355, 180–198. [Google Scholar] [CrossRef]
  8. Janczak, J.; Prochowicz, D.; Lewiński, J.; Fairen-Jimenez, D.; Bereta, T.; Lisowski, J. Trinuclear cage-like ZnII macrocyclic complexes: Enantiomeric recognition and gas adsorption properties. Eur. J. Chem. 2016, 22, 598–609. [Google Scholar] [CrossRef]
  9. Zhang, X.; Yin, J.; Yoon, J. Recent Advances in Development of Chiral Fluorescent and Colorimetric Sensors. Chem. Rev. 2014, 114, 4918–4959. [Google Scholar] [CrossRef]
  10. Li, Z.-B.; Lin, J.; Sabat, M.; Hyacinth, M.; Pu, L. Enantioselective Fluorescent Recognition of Chiral Acids by Cyclohexane-1,2-diamine-Based Bisbinaphthyl Molecules. J. Org. Chem. 2007, 72, 4905–4916. [Google Scholar] [CrossRef] [Green Version]
  11. Liu, X.; Manzur, C.; Novoa, N.; Celedón, S.; Carrillo, D.; Hamon, J.R. Multidentate unsymmetrically-substituted Schiff bases and their metal complexes: Synthesis, functional materials properties, and applications to catalysis. Coord. Chem. Rev. 2018, 357, 144–172. [Google Scholar] [CrossRef]
  12. Padnya, P.; Shibaeva, K.; Arsenyev, M.; Baryshnikova, S.; Terenteva, O.; Shiabiev, I.; Khannanov, A.; Boldyrev, A.; Gerasimov, A.; Grishaev, D.; et al. Catechol-Containing Schiff Bases on Thiacalixarene: Synthesis, Copper (II) Recognition, and Formation of Organic-Inorganic Copper-Based Materials. Molecules 2021, 26, 2334. [Google Scholar] [CrossRef]
  13. Che, C.-M.; Huang, J.-S. Metal complexes of chiral binaphthyl Schiff-base ligands and their application in stereoselective organic transformations. Coord. Chem. Rev. 2003, 242, 97–113. [Google Scholar] [CrossRef]
  14. Shockravi, A.; Javadi, A.; Abouzari-Lotf, E. Binaphthyl-based macromolecules: A review. RSC Adv. 2013, 3, 6717–6746. [Google Scholar] [CrossRef]
  15. Hajari, S.; Keypour, H.; Rezaei, M.T.; Farida, S.H.M.; Gable, R.W. New 15-membered macrocyclic Schiff base ligand; synthesis some Cd(II), Mn(II) and Zn(II) complexes, crystal structure, cytotoxicity, antibacterial and antioxidant activity. J. Mol. Struct. 2022, 1251, 132049. [Google Scholar] [CrossRef]
  16. El-Gammal, O.A.; Mohamed, F.S.; Rezk, G.N.; El-Bindary, A.A. Synthesis, characterization, catalytic, DNA binding and antibacterial activities of Co(II), Ni(II) and Cu(II) complexes with new Schiff base ligand. J. Mol. Liq. 2021, 1230, 129908. [Google Scholar] [CrossRef]
  17. Lee, E.; Lee, S.Y.; Lindoy, L.F.; Lee, S.S. Metallacycles derived from metal complexes of exo-coordinated macrocyclic ligands. Coord. Chem. Rev. 2013, 257, 3125–3138. [Google Scholar] [CrossRef]
  18. Tokunaga, H.; Kazama, K.; Tsuboi, M.; Miyasaka, M. A novel Schiff base macrocycle based on 1,10 -binaphthyl for fluorescence recognition. Luminescence 2021, 36, 1561–1568. [Google Scholar] [CrossRef]
  19. Korupoju, S.R.; Mangayarkarasi, N.; Ameerunisha, S.; Valente, E.J.; Zacharias, P.S. Formation of dinuclear macrocyclic and mononuclear acyclic complexes of a new trinucleating hexaaza triphenolic Schiffbase macrocycle: Structure and NLO properties. J. Chem. Soc. Dalton Trans. 2000, 16, 2845–2852. [Google Scholar] [CrossRef]
  20. Telfer, S.G.; Sato, T.; Harada, T.; Kuroda, R.; Lefebvre, J.; Leznoff, D.B. Mono- and Dinuclear Complexes of Chiral Tri- and Tetradentate Schiff-Base Ligands Derived from 1,1’-Binaphthyl-2,2’-diamine. Inorg. Chem. 2004, 43, 6168–6176. [Google Scholar] [CrossRef]
  21. Telfer, S.G.; Kuroda, R. 1,1′-Binaphthyl-2,2′-diol and 2,2′-diamino-1,1′-binaphthyl: Versatile frameworks for chiral ligands in coordination and metallosupramolecular chemistry. Coord. Chem. Rev. 2003, 242, 33–46. [Google Scholar] [CrossRef]
  22. Chinnaraja, E.; Arunachalam, R.; Suresh, E.; Sen, S.K.; Natarajan, R.; Subramanian, P.S. Binuclear Double-Stranded Helicates and Their Catalytic Applications in Desymmetrization of Mesodiols. Inorg. Chem. 2019, 58, 4465–4479. [Google Scholar] [CrossRef]
  23. Chinnaraja, E.; Arunachalam, R.; Samanta, K.; Natarajan, R.; Subramaniana, P.S. Enantioselective Michael Addition Reaction Catalysed by Enantiopure Binuclear Nickel(II) Close-Ended Helicates. Adv. Synth. Catal. 2020, 362, 1144–1155. [Google Scholar] [CrossRef]
  24. Chinnaraja, E.; Arunachalam, R.; Pillai, R.S.; Peuronen, A.; Rissanen, K.; Subramanian, P.S. One-pot synthesis of [2+2]-helicate-like macrocycle and 2+4-μ4-oxo tetranuclear open frame complexes: Chiroptical properties and asymmetric oxidative coupling of 2-naphthols. Appl. Organomet. Chem. 2020, 34, e5666. [Google Scholar] [CrossRef]
  25. Castellano, M.; Ruiz-García, R.; Cano, J.; Ferrando-Soria, J.; Pardo, E.; Fortea-Pérez, F.R.; Stiriba, S.-E.; Julve, M.; Lloret, F. Dicopper(II) Metallacyclophanes as Multifunctional Magnetic Devices: A Joint Experimental and Computational Study. Acc. Chem. Res. 2015, 48, 510–520. [Google Scholar] [CrossRef] [PubMed]
  26. Ferrando-Soria, J.; Vallejo, J.; Castellano, M.; Martínez-Lillo, J.; Pardo, E.; Cano, J.; Castro, I.; Lloret, F.; Ruiz-García, R.; Julve, M. Molecular magnetism, quo vadis? A historical perspective from a coordination chemist viewpoint. Coord. Chem. Rev. 2017, 339, 17–103. [Google Scholar] [CrossRef]
  27. Žilić, D.; Rakvin, B.; Milić, D.; Pajić, D.; Đilović, I.; Camettid, M.; Džolić, Z. Crystal structures and magnetic properties of a set of dihalo-bridged oxalamidato copper(II) dimers. Dalton Trans. 2014, 43, 11877–11887. [Google Scholar] [CrossRef] [PubMed]
  28. Brown, S.J.; Tao, X.; Wark, T.A.; Stephan, D.W.; Mascharak, P.K. Synthetic analog approach to metallobleomycins. 4. New halobridged dimeric and polymeric (infinite zigzag chain) complexes of copper(II) with peptide ligands related to bleomycins. Inorg. Chem. 1988, 27, 1581–1587. [Google Scholar] [CrossRef]
  29. Sena, N.; Butcher, J.R.; Jasinski, J.P.; Gupta, S.K. A novel single-pot synthesis of dinuclear and mononuclear copper(II) complexes with sterically demanding Schiff bases: Structural, spectral, magnetic, electrochemical, DNA binding and theoretical investigation. J. Mol. Struct. 2021, 1231, 129955. [Google Scholar] [CrossRef]
  30. Chattopadhyay, T.; Banu, K.S.; Banerjee, A.; Ribas, J.; Majee, A.; Nethaji, M.; Das, D. A novel single pot synthesis of binuclear copper(II) complexes of macrocyclic and macrocyclic compartmental ligands: Structures and magnetic properties. J. Mol. Struct. 2007, 833, 13–22. [Google Scholar] [CrossRef]
  31. Thompson, K.; Mandal, S.K.; Tandon, S.S.; Bridson, J.N.; Park, M.K. Magnetostructural Correlations in Bis(µ2-phenoxide)-Bridged Macrocyclic Dinuclear Copper(II) Complexes. Influence of Electron-Withdrawing Substituents on Exchange Coupling Laurence. Inorg. Chem. 1996, 35, 3117–3125. [Google Scholar] [CrossRef]
  32. Addison, A.W.; Rao, T.N.; Reedijk, J.; van Rijn, J.; Verschoor, G.C. Synthesis, structure, and spectroscopic properties of copper(II) compounds containing nitrogen–sulphur donor ligands; the crystal and molecular structure of aqua[1,7-bis(N-methylbenzimidazol-2′-yl)-2,6-dithiaheptane]copper(II) perchlorate. J. Chem. Soc. Dalton Trans. 1984, 7, 1349. [Google Scholar] [CrossRef]
  33. Llunel, M.; Casanova, D.; Cirera, J.; Bofill, J.M.; Alemany, P.; Alvarez, S.; Pinsky, M.; Avnir, D. SHAPE, Version 2.1; Program for the Calculations of Continuous Shape Measures of Polygonal and Polyhedral Molecular Fragments; University of Barcelona: Barcelona, Spain, 2013. [Google Scholar]
  34. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.K.; Puschmann, H. OLEX2: A complete structure solution, refinement and analysis program. J. Appl. Cryst. 2009, 42, 339–341. [Google Scholar] [CrossRef]
  35. Chilton, N.F.; Anderson, R.P.; Turner, L.D.; Soncini, A.; Murray, K.S.J. PHI: A powerful new program for the analysis of anisotropic monomeric and exchange-coupled polynuclear d- and f-block complexes. Comput. Chem. 2013, 34, 1164–1175. [Google Scholar] [CrossRef]
  36. Kahn, O. Molecular Magnetism; VCH: New York, NY, USA, 1993; Chapter 8; p. 160. [Google Scholar]
  37. Ruiz, E.; Cano, J.; Alvarez, S.; Alemany, P. Broken symmetry approach to calculation of exchange coupling constants for homobinuclear and heterobinuclear transition metal complexes. J. Comput Chem. 1999, 20, 1391–1400. [Google Scholar] [CrossRef]
  38. Maślewski, P.; Wyrzykowski, D.; Witwicki, M.; Dołęga, A. Histaminol and Its Complexes with Copper(II)—Studies in Solid State and Solution. Eur. J. Inorg. Chem. 2018, 2018, 1399–1408. [Google Scholar] [CrossRef]
  39. Mielcarek, A.; Bieńko, A.; Saramak, P.; Jezierska, J.; Dołęga, A. A Cu/Zn heterometallic complex with solvent-binding cavity, catalytic activity for the oxidation of 1-phenylethanol and unusual magnetic properties. Dalton Trans. 2019, 48, 17780. [Google Scholar] [CrossRef]
  40. Peisach, J.; Blumberg, W.E. Structural implications derived from the analysis of electron paramagnetic resonance spectra of natural and artificial copper proteins. Arch. Biochem. Biophys. 1974, 165, 691. [Google Scholar] [CrossRef]
  41. Barwiolek, M.; Kaczmarek-Kędziera, A.; Muziol, T.M.; Jankowska, D.; Jezierska, J.; Bieńko, A. Dinuclear Copper(II) Complexes with Schiff Bases Derived from 2-Hydroxy-5-Methylisophthalaldehyde and Histamine or 2-(2-Aminoethyl)pyridine and Their Application as Magnetic and Fluorescent Materials in Thin Film Deposition. Int. J. Mol. Sci. 2020, 21, 4587. [Google Scholar] [CrossRef]
  42. Buvailo, H.I.; Makhankova, V.G.; Kokozay, V.N.; Babaryk, A.A.; Omelchenko, I.V.; Shishkina, S.V.; Bienko, D.C.; Jezierska, J.; Bienko, A. Hybrid Cu-Containing Compounds Based on Lacunary Strandberg Anions: Synthesis under Mild Conditions, Crystal Structure, and Magnetic Properties. Inorg. Chem. 2022, 61, 5701–5714. [Google Scholar] [CrossRef]
  43. Buvaylo, E.A.; Kokozay, V.N.; Makhankova, V.G.; Melnyk, A.K.; Korabik, M.; Witwicki, M.; Skelton, B.W.; Vassilyeva, O.Y. Synthesis, Characterization, and Magnetic Properties of a Series of Copper(II) Chloride Complexes of Pyridyliminebenzoic Acids. Eur. J. Inorg. Chem. 2018, 2018, 1603–1619. [Google Scholar] [CrossRef]
  44. Reichardt, C. Solvents and Solvent Effects in Organic Chemistry; Wiley: Hoboken, NJ, USA, 2003; ISBN 3-527-30618-8. [Google Scholar]
  45. Fuenteabla, P.; Seron, L.; Sachez, C.; Manzur, J.; Paredes-Garcia, V.; Pizzaro, N.; Cepeda, M.; Venegaz-Yazigi, D.; Spodine, E. Macrocyclic ZnII and CuII complexes as guests of the hybrid composites based on the layered MnPS3 phase. Comparison of spectroscopic properties. J. Coord. Chem. 2014, 67, 23–24, 3894–3908. [Google Scholar] [CrossRef]
  46. Brudenell, S.J.; Spiccia, L.; Tiekink, E.R.T. Binuclear Copper(II) Complexes of Bis(pentadentate) Ligands Derived from Alkyl-Bridged Bis(1,4,7-triazacyclonane) Macrocycles. Inorg. Chem. 1996, 35, 1974–1979. [Google Scholar] [CrossRef]
  47. Chinnaraja, E.; Arunachalam, R.; Subramanian, P.S. Enantio- and Diastereoselective Synthesis of b-Nitroalcohol via Henry Reaction Catalyzed by Cu(II), Ni(II), Zn(II) Complexes of Chiral BINIM Ligands. Chem. Select 2016, 1, 5331–5338. [Google Scholar] [CrossRef]
  48. Grzybowski, J.J.; Merrell, P.H.; Urbach, F.L. Dicopper(II) Complexes with Binucleating Ligands Derived from 2-Hydroxy-5-methylisophthaldehyde and 2-(2-aminoethyl)pyridine. Inorg. Chem. 1978, 17, 3078–3082. [Google Scholar] [CrossRef]
  49. Roznyatovsky, V.V.; Borisova, N.E.; Reshetova, M.D.; Ustynyuk, Y.A.; Aleksandrov, G.G.; Eremenko, I.; Moiseev, I.I. Dinuclear and polynuclear transition metal complexes with macrocyclic ligands. 6. New dinuclear copper(II) complexes with macrocyclic Schiff bases derived from 4-tert-butyl-2,6-diformylphenol. Russ. Chem. Bull. 2004, 53, 1208–1217. [Google Scholar] [CrossRef]
  50. Qing, S.; Qiao-Lin, W.; Guang-Hua, L.; Xiao-Ming, L.; Ying, M. Bis-salicylaldiminato zinc complexes: Syntheses, characterization and luminescent properties. Polyhedron 2007, 26, 5053–5060. [Google Scholar] [CrossRef]
  51. Fan, J.; Autschbach, J.; Ziegler, T. Electronic Structure and Circular Dichroism of Tris(bipyridyl) Metal Complexes within Density Functional Theory. Inorg. Chem. 2010, 49, 1355. [Google Scholar] [CrossRef]
  52. Enamullah, M.; Uddin, A.K.M.R.; Pescitelli, G.; Berardozzi, R.; Makhloufi, G.; Vasylyeva, V.; Chamayou, A.-C.; Janiak, C. Induced chirality-at-metal and diastereoselectivity at Δ/Λ-configured distorted square-planar copper complexes by enantiopure Schiff base ligands: Combined circular dichroism, DFT and X-ray structural studies. Dalton Trans. 2014, 43, 3313. [Google Scholar] [CrossRef]
  53. Ahmed, N.; Tripathi, S.; Sarkar, A.; Ansari, K.U.; Das, C.; Prajesh, N.; Horike, S.; Boomishankar, R.; Shanmugam, M. Chiral tetranuclear copper(II) complexes: Synthesis, optical and magnetic properties. New J. Chem. 2020, 44, 16845. [Google Scholar] [CrossRef]
  54. Barwiolek, M.; Jankowska, D.; Kaczmarek-Kędziera, A.; Wojtulewski, S.; Skowroński, L.; Rerek, T.; Popielarski, P.; Muziol, T.M. Experimental and Theoretical Studies of the Optical Properties of the Schiff Bases and Their Materials Obtained from o-Phenylenediamine. Molecules 2022, 27, 7396. [Google Scholar] [CrossRef] [PubMed]
  55. Basumallick, L.; George, S.D.; Randall, D.W.; Hedman, B.; Hodgson, K.O.; Fujisawa, K.; Solomon, E.I. Spectroscopic comparison of the five-coordinate [Cu(SMeIm)(HB(3,5-iPr2pz)3)] with the four-coordinate [Cu(SCPh3)(HB(3,5-iPr2pz)3)]: Effect of coordination number increase on a blue copper type site. Inorg. Chim. Acta 2002, 33, 357–365. [Google Scholar] [CrossRef]
  56. Muzioł, T.M.; Tereba, N.; Podgajny, R.; Kedziera, D.; Wrzeszcz, G. Solvent-assisted structural conversion involving bimetallic complexes based on the Tris(Oxalato)Ferrate(III) unit with the green blue red crystal color sequence. Dalton Trans. 2019, 48, 11536–11546. [Google Scholar] [CrossRef] [PubMed]
  57. Muzioł, T.M.; Tereba, N.; Podgajny, R.; Pełka, R.; Czernia, D.; Wiśniewski, M.; Koter, S.; Wrzeszcz, G. Sorption and Magnetic Properties of Oxalato-Based Trimetallic Open Framework Stabilized by Charge-Assisted Hydrogen Bonds. Int. J. Mol. Sci. 2022, 23, 1556. [Google Scholar] [CrossRef] [PubMed]
  58. Baker, M.L.; Mara, M.W.; Yan, J.J.; Hodgson, K.O.; Hedman, B.; Solomon, E.I. K- and L-edge X-ray Absorption Spectroscopy (XAS) and Resonant Inelastic X-ray Scattering (RIXS) Determination of Differential Orbital Covalency (DOC) of Transition Metal Sites. Coord. Chem. Rev. 2017, 345, 182–208. [Google Scholar] [CrossRef]
  59. Chainani, A.; Sicot, M.; Fagot-Revurat, Y.; Vasseur, G.; Granet, J.; Kierren, B.; Moreau, L.; Oura, M.; Yamamoto, A.; Tokura, Y.; et al. Evidence for Weakly Correlated Oxygen Holes in the Highest-Tc Cuprate Superconductor HgBa2Ca2Cu3O8+δ. Phys. Rev. Lett. 2017, 119, 57001. [Google Scholar] [CrossRef]
  60. Pearce, C.I.; Pattrick, R.A.D.; Vaughan, D.J.; Henderson, C.M.B.; van der Laan, G. Copper oxidation state in chalcopyrite: Mixed Cu d9 and d10 characteristics. Geochim. Cosmochim. Acta 2006, 70, 4635–4642. [Google Scholar] [CrossRef]
  61. Wang, H.; Friedrich, S.; Li, L.; Mao, Z.; Ge, P.; Balasubramanian, M.; Patil, D.S. L-edge sum rule analysis on 3d transition metal sites: From d10 to d0 and towards application to extremely dilute metallo-enzymes. Phys. Chem. Chem. Phys. 2018, 20, 8166. [Google Scholar] [CrossRef]
  62. Pattrick, R.A.D.; Coker, V.S.; Pearce, C.I. The oxidation state of copper and cobalt in carrollite, CuCo2S4. Can. Mineral. 2008, 46, 1317–1322. [Google Scholar] [CrossRef]
  63. Sarangi, R.; Aboelella, N.; Fujisawa, K.; Tolman, W.B.; Hedman, B.; Hodgson, K.O.; Solomon, E. X-ray Absorption Edge Spectroscopy and Computational Studies on LCuO2 Species:  Superoxide−CuII versus Peroxide−CuIII Bonding. J. Am. Chem. Soc. 2006, 128, 8286–8296. [Google Scholar] [CrossRef] [Green Version]
  64. George, S.D.; Metz, M.; Szilagyi, R.K.; Wang, H.; Cramer, S.P.; Lu, Y.; Tolman, W.B.; Hedman, B.; Hodgson, K.O.; Solomon, E.I. A Quantitative Description of the Ground-State Wave Function of CuA by X-ray Absorption Spectroscopy: Comparison to Plastocyanin and Relevance to Electron Transfer. J. Am. Chem. Soc. 2001, 123, 5757–5767. [Google Scholar] [CrossRef] [PubMed]
  65. Oxford Diffraction Ltd. CrysAlis RED and CrysAlis CCD; Oxford Diffraction Ltd.: Abingdon, UK, 2000. [Google Scholar]
  66. Kabsch, W. XDS. Acta Cryst. 2010, D66, 125–132. [Google Scholar] [CrossRef]
  67. Krug, M.; Weiss, M.S.; Heinemann, U.; Mueller, U. XDSAPP: A graphical user interface for the convenient processing of diffraction data using XDS. J. Appl. Crystallogr. 2012, 45, 568–572. [Google Scholar] [CrossRef]
  68. Walker, N.; Stuart, D. An empirical method for correcting diffractometer data for absorption effects. Acta. Cryst. 1983, A39, 158–166. [Google Scholar] [CrossRef]
  69. Sheldrick, G.M. A short history of SHELX. Acta Crystallogr. Sect. A 2008, 64, 112–122. [Google Scholar] [CrossRef]
  70. Brandenburg, K. DIAMOND, release 2.1e; Crystal Impact GbR: Bonn, Germany, 2001. [Google Scholar]
  71. Farrugia, L.J.J. ORTEP-3 for Windows—A version of ORTEP-III with a Graphical User Interface (GUI). Appl. Crystallogr. 1997, 30, 565. [Google Scholar] [CrossRef]
  72. Bain, G.A.; Berry, J.F. Diamagnetic corrections and Pascal’s constants. J. Chem. Educ. 2008, 85, 532. [Google Scholar] [CrossRef]
  73. Stoll, S.; Schweiger, A. EasySpin, a Comprehensive Software Package for Spectral Simulation and Analysis in EPR. J. Magn. Reson. 2006, 178, 42–55. [Google Scholar] [CrossRef]
  74. Stoll, S. CW-EPR Spectral Simulations: Solid State. Methods Enzymol. 2015, 563, 121–142. [Google Scholar] [CrossRef]
  75. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16, revision B.01; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  76. David, G.; Wennmohs, F.; Neese, F.; Ferre, N. Chemical Tuning of Magnetic Exchange Couplings Using Broken-Symmetry Density Functional Theory. Inorg. Chem. 2018, 57, 12769–12776. [Google Scholar] [CrossRef]
  77. Bencini, A.; Gatteschi, D.X. Alpha.-SW calculations of the electronic structure and magnetic properties of weakly coupled transition-metal clusters. The [Cu2Cl6]2-dimers. J. Am. Chem. Soc. 1980, 108, 5763. [Google Scholar] [CrossRef] [PubMed]
  78. Neese, F.; Wennmohs, F.; Becker, U.; Riplinger, C. The ORCA quantum chemistry program package. J. Chem. Phys. 2020, 152, 224108. [Google Scholar] [CrossRef]
  79. Ravel, B.; Newville, M. ATHENA, ARTEMIS, HEPHAESTUS: Data analysis for x-ray absorption spectroscopy using IFEFFIT. J. Synchrotron Radiat. 2005, 12, 537–541. [Google Scholar] [CrossRef] [PubMed]
  80. Mueller, U.; Förster, R.; Hellmig, M.; Huschmann, F.U.; Kastner, A.; Malecki, P.; Pühringer, S.; Röwer, M.; Sparta, K.; Steffien, M.; et al. The macromolecular crystallography beamlines at BESSY II of the helmholtz-zentrum Berlin: Current status and perspectives. Eur. Phys. J. Plus 2015, 130, 141–150. [Google Scholar] [CrossRef]
Figure 1. Scheme of the copper(II) complex K1 and K2 synthesis.
Figure 1. Scheme of the copper(II) complex K1 and K2 synthesis.
Ijms 24 03017 g001
Figure 2. Structure of [Cu2Cl2(L’)] K1 with a numbering scheme and thermal ellipsoids at 30% probability.
Figure 2. Structure of [Cu2Cl2(L’)] K1 with a numbering scheme and thermal ellipsoids at 30% probability.
Ijms 24 03017 g002
Figure 3. The crystal network of K1 showed channels running along the c axis. The unit cell is given in blue box.
Figure 3. The crystal network of K1 showed channels running along the c axis. The unit cell is given in blue box.
Ijms 24 03017 g003
Figure 4. Structure of [Cu2Cl2(L)]·H2O K2 with a numbering scheme and thermal ellipsoids at 30% probability.
Figure 4. Structure of [Cu2Cl2(L)]·H2O K2 with a numbering scheme and thermal ellipsoids at 30% probability.
Ijms 24 03017 g004
Figure 5. The crystal network of K2 showing channels running along the b axis filled with water molecules. The unit cell is given in blue box.
Figure 5. The crystal network of K2 showing channels running along the b axis filled with water molecules. The unit cell is given in blue box.
Ijms 24 03017 g005
Figure 6. Hirshfeld surfaces and fingerprints of selected interactions created in the crystal network of K2: Hirshfeld surface (a) and fingerprint (b) for H…H (54.2%), Hirshfeld surface (c) and fingerprint (d) for H…C (28.4%), Hirshfeld surface (e) and fingerprint (f) for H…Cl (9.0%), and Hirshfeld surface (g) and fingerprint (h) for H…O (3.6%). In brackets, there is a given surface area included as a percentage of the total surface area.
Figure 6. Hirshfeld surfaces and fingerprints of selected interactions created in the crystal network of K2: Hirshfeld surface (a) and fingerprint (b) for H…H (54.2%), Hirshfeld surface (c) and fingerprint (d) for H…C (28.4%), Hirshfeld surface (e) and fingerprint (f) for H…Cl (9.0%), and Hirshfeld surface (g) and fingerprint (h) for H…O (3.6%). In brackets, there is a given surface area included as a percentage of the total surface area.
Ijms 24 03017 g006aIjms 24 03017 g006b
Figure 7. Magnetic properties of K1 and K2: (a) χMT plots at Hdc = 5 kOe together with the PHI fits and values of JCu-Cu and gCu,avg; (b) M(H) plots at T = 1.8 K.
Figure 7. Magnetic properties of K1 and K2: (a) χMT plots at Hdc = 5 kOe together with the PHI fits and values of JCu-Cu and gCu,avg; (b) M(H) plots at T = 1.8 K.
Ijms 24 03017 g007
Figure 8. Absorption spectra (a) of K1 and (b) K2 complexes in solvents: chloroform (red), acetone (blue), methanol (black), acetonitrile (green), and DMSO (purple) (3.323 × 10−6 mol/dm3, RT).
Figure 8. Absorption spectra (a) of K1 and (b) K2 complexes in solvents: chloroform (red), acetone (blue), methanol (black), acetonitrile (green), and DMSO (purple) (3.323 × 10−6 mol/dm3, RT).
Ijms 24 03017 g008
Figure 9. Solid state absorption spectra of black K1 and red K2., A—absorbance.
Figure 9. Solid state absorption spectra of black K1 and red K2., A—absorbance.
Ijms 24 03017 g009
Figure 10. Emission spectra of the (a) K1 and (b) K2 complexes’ solutions; λex = 350 nm chloroform, acetone, methanol, acetonitrile, DMSO (3.3 × 10−6 mol/dm3, RT).
Figure 10. Emission spectra of the (a) K1 and (b) K2 complexes’ solutions; λex = 350 nm chloroform, acetone, methanol, acetonitrile, DMSO (3.3 × 10−6 mol/dm3, RT).
Ijms 24 03017 g010
Figure 11. CD spectra of K1 and K2 in chloroform 1 × 10−4 M.
Figure 11. CD spectra of K1 and K2 in chloroform 1 × 10−4 M.
Ijms 24 03017 g011
Figure 12. Vertical absorption spectrum for K1, K2, and K2 with one water molecule, estimated within the ωB97X-D/def2-SVP/PCM(CHCl3) approach (the sticks represent the position and the corresponding oscillator strength for the subsequent excitations, and the curve is the fit to these data).
Figure 12. Vertical absorption spectrum for K1, K2, and K2 with one water molecule, estimated within the ωB97X-D/def2-SVP/PCM(CHCl3) approach (the sticks represent the position and the corresponding oscillator strength for the subsequent excitations, and the curve is the fit to these data).
Ijms 24 03017 g012
Figure 13. Natural transition orbitals for K1, corresponding to the most intensive transitions, estimated within the ωB97X-D/def2-SVP/PCM(CHCl3) approach ((a) correspond to the d-d* transition, (b) to LMCT, and (c) to intraligand ππ* transition).
Figure 13. Natural transition orbitals for K1, corresponding to the most intensive transitions, estimated within the ωB97X-D/def2-SVP/PCM(CHCl3) approach ((a) correspond to the d-d* transition, (b) to LMCT, and (c) to intraligand ππ* transition).
Ijms 24 03017 g013
Figure 14. AFM of K2 material, spin-coating, 3000 rpm, 10 s, chloroform scan size 10 μm, height (thickness) 7 nm, and EDS mapping of K1/Si.
Figure 14. AFM of K2 material, spin-coating, 3000 rpm, 10 s, chloroform scan size 10 μm, height (thickness) 7 nm, and EDS mapping of K1/Si.
Ijms 24 03017 g014
Figure 15. SEM of K1/Si thermal deposition. (a) K1/Si; (b) EDS; and (c) mapping of K1/Si. Magn 500×.
Figure 15. SEM of K1/Si thermal deposition. (a) K1/Si; (b) EDS; and (c) mapping of K1/Si. Magn 500×.
Ijms 24 03017 g015
Figure 16. SEM of K2/Si thermal deposition. (a) K1/Si; (b) EDS; and (c) mapping of K1/Si. Magn 500×.
Figure 16. SEM of K2/Si thermal deposition. (a) K1/Si; (b) EDS; and (c) mapping of K1/Si. Magn 500×.
Ijms 24 03017 g016
Figure 17. AFM of K2/Si, thermal deposition, scan size 10 µm, Ra = 4.02 nm, Rq = 5.36 nm.
Figure 17. AFM of K2/Si, thermal deposition, scan size 10 µm, Ra = 4.02 nm, Rq = 5.36 nm.
Ijms 24 03017 g017aIjms 24 03017 g017b
Figure 18. Emission spectra of the K2 material, λex = 320 nm, spin coating.
Figure 18. Emission spectra of the K2 material, λex = 320 nm, spin coating.
Ijms 24 03017 g018
Figure 19. Emission spectra of K1/Si and K2/Si thermal deposition.
Figure 19. Emission spectra of K1/Si and K2/Si thermal deposition.
Ijms 24 03017 g019
Figure 20. Normalized XANES spectra of copper L2,3-edge with peaks corresponding to 2p63d9 → 2p53d10. The signal for K1 was shifted by 2 to separate both spectra.
Figure 20. Normalized XANES spectra of copper L2,3-edge with peaks corresponding to 2p63d9 → 2p53d10. The signal for K1 was shifted by 2 to separate both spectra.
Ijms 24 03017 g020
Table 1. Selected bond length (Å) and valence angles (°) for the complex K1.
Table 1. Selected bond length (Å) and valence angles (°) for the complex K1.
(K1)
Cu(1)-N(1) 1.917(11)Cu(2)-O(49) 2.053(8)
Cu(1)-O(22) 1.944(8)Cu(2)-N(28) 2.069(9)
Cu(1)-O(49) 1.974(6)Cu(2)-Cl(2) 2.363(2)
Cu(1)-N(20) 2.034(10)
Cu(1)-Cl(1) 2.495(3)
Cu(2)-O(22) 1.923(7)
Cu(2)-N(47) 1.954(9)
Valence angles (°)
N(1)-Cu(1)-O(22)168.5(3)O(49)-Cu(1)-Cl(1)121.48(18)
N(1)-Cu(1)-O(49)92.4(3)N(20)-Cu(1)-Cl(1)95.0(3)
O(22)-Cu(1)-O(49)76.2(3)O(22)-Cu(2)-N(47)165.5(3)
N(1)-Cu(1)-N(20)101.8(4)O(22)-Cu(2)-O(49)74.8(3)
O(22)-Cu(1)-N(20)87.5(4)N(47)-Cu(2)-O(49)91.0(3)
O(49)-Cu(1)-N(20)140.3(3)O(22)-Cu(2)-N(28)89.3(3)
N(1)-Cu(1)-Cl(1)92.7(2)N(47)-Cu(2)-N(28)98.8(3)
O(22)-Cu(1)-Cl(1)93.2(2)O(49)-Cu(2)-N(28)134.7(3)
Table 2. Selected bond length (Å) and valence angles (°) for the complex K2.
Table 2. Selected bond length (Å) and valence angles (°) for the complex K2.
K2
Cu(1)-O(22) 1.920(5)Cu(2)-O(49) 1.989(5)
Cu(1)-N(1) 1.953(6)Cu(2)-N(28) 2.061(6)
Cu(1)-O(49) 2.047(5)Cu(2)-Cl(2) 2.370(3)
Cu(1)-N(20) 2.078(5)
Cu(1)-Cl(1) 2.302(3)
Cu(2)-O(22) 1.900(5)
Cu(2)-N(47) 1.954(7)
Valence angles (°)
O(22)-Cu(1)-N(1)165.2(2)O(49)-Cu(1)-Cl(1)115.4(2)
O(22)-Cu(1)-O(49)75.1(2)N(20)-Cu(1)-Cl(1)113.60(19)
N(1)-Cu(1)-O(49)90.4(2)O(22)-Cu(2)-N(47)165.0(3)
O(22)-Cu(1)-N(20)87.7(2)O(22)-Cu(2)-O(49)77.0(2)
N(1)-Cu(1)-N(20)99.3(2)N(47)-Cu(2)-O(49)89.5(3)
O(49)-Cu(1)-N(20)128.7(3)O(22)-Cu(2)-N(28)87.4(2)
O(22)-Cu(1)-Cl(1)93.74(19)N(47)-Cu(2)-N(28)98.5(3)
N(1)-Cu(1)-Cl(1)95.4(2)O(49)-Cu(2)-N(28)137.4(3)
Table 3. Fitted values of JCu-Cu, gCu,avg, and IMP contribution. The J(SUP) a was extracted from the broken symmetry and spin-flip DFT results (for details, see SI section).
Table 3. Fitted values of JCu-Cu, gCu,avg, and IMP contribution. The J(SUP) a was extracted from the broken symmetry and spin-flip DFT results (for details, see SI section).
CompoundK1K2
JCu-Cu/cm−1−305.66 ± 0.05−328.45 ± 0.16
J(SUP) a−337.0−346.5
gCu,avg2.020 ± 0.0012.187 ± 0.002
IMP b/%11.7412.46
Residual c6.35 ± 10−43.79 ± 10−4
a Spin-unprojected approximation. b Paramagnetic contribution represented by the option of “impurities” in PHI software. c The value of the fit quality test function embedded in PHI software.
Table 4. Summary of spectral features of the Cu absorption at the L-edges (peak energies and intensities).
Table 4. Summary of spectral features of the Cu absorption at the L-edges (peak energies and intensities).
CompoundEnergy of Maximum L3-Edge (eV) (L3-Edge Intensity)Energy of Maximum L2-Edge (eV) (L2-Edge Intensity)
K1931.0 (4.17)
934.8 (0.33)
951.0 (1.46)
954.5 (0.65)
K2931.0 (4.47)
934.7 (0.35)
950.9 (1.51)
954.5 (0.68)
Table 5. Crystal data and structure refinement for K1 and K2.
Table 5. Crystal data and structure refinement for K1 and K2.
Identification CodeK1K2
Empirical formulaC58 H38 Cl2 Cu2 N4 O2C64 H52 Cl2 Cu2 N4 O3
Formula weight1020.901123.07
Temperature (K)100(2)100(2)
Wavelength (Å)1.541840.7999
Crystal system, space groupOrthorhombic, P21212Orthorhombic, P212121
Unit cell dimensions (Å) and (°)a = 10.6153(3)  α = 90
b = 22.3313(15)  β = 90
c = 22.9053(11)  γ = 90
a = 10.922(2)  α = 90
b = 19.448(4)  β = 90
c = 26.041(5)  γ = 90
Volume (Å3)5429.8(5)5531.4(19)
Z, Calculated density (Mg⋅m–3)4, 1.2494, 1.349
Absorption coefficient (mm–1)2.2071.262
F(000)20882302
Crystal size (mm3)0.150 × 0.050 × 0.0300.170 × 0.070 × 0.070
Theta range for data collection (°)2.764 to 68.2431.471 to 28.430
Limiting indices−10 ≤ h ≤ 12
−26 ≤ k ≤ 26
−27 ≤ l ≤ 27
−12 ≤ h ≤ 12
−23 ≤ k ≤ 23
−30 ≤ l ≤ 30
Reflections collected/unique36173/9932 (R(int) = 0.0997)61117/9714 (R(int) = 0.0538)
Completeness to theta = 29.732° (%)99.899.5
Max. and min. transmission1.000 and 0.6720.8864 and 0.2903
Refinement methodFull-matrix least-squares on F2Full-matrix least-squares on F2
Data/restraints/parameters9932/13/6179714/18/679
Goodness-of-fit on F20.9791.068
Final R indices (I > 2sigma(I))R1 = 0.0795, wR2 = 0.1930R1 = 0.0581, wR2 = 0.1449
R indices (all data)R1 = 0.1161, wR2 = 0.2166R1 = 0.0664, wR2 = 0.1501
Largest diff. peak and hole (eÅ–3)1.420 and −0.4340.602 and −0.481
CCDC 2233358 and 2233362 contain the supplementary crystallographic data for K1 and K2. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif (accessed on 29 December 2022).
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

Barwiolek, M.; Jankowska, D.; Kaczmarek-Kędziera, A.; Lakomska, I.; Kobylarczyk, J.; Podgajny, R.; Popielarski, P.; Masternak, J.; Witwicki, M.; Muzioł, T.M. New Dinuclear Macrocyclic Copper(II) Complexes as Potentially Fluorescent and Magnetic Materials. Int. J. Mol. Sci. 2023, 24, 3017. https://doi.org/10.3390/ijms24033017

AMA Style

Barwiolek M, Jankowska D, Kaczmarek-Kędziera A, Lakomska I, Kobylarczyk J, Podgajny R, Popielarski P, Masternak J, Witwicki M, Muzioł TM. New Dinuclear Macrocyclic Copper(II) Complexes as Potentially Fluorescent and Magnetic Materials. International Journal of Molecular Sciences. 2023; 24(3):3017. https://doi.org/10.3390/ijms24033017

Chicago/Turabian Style

Barwiolek, Magdalena, Dominika Jankowska, Anna Kaczmarek-Kędziera, Iwona Lakomska, Jedrzej Kobylarczyk, Robert Podgajny, Paweł Popielarski, Joanna Masternak, Maciej Witwicki, and Tadeusz M. Muzioł. 2023. "New Dinuclear Macrocyclic Copper(II) Complexes as Potentially Fluorescent and Magnetic Materials" International Journal of Molecular Sciences 24, no. 3: 3017. https://doi.org/10.3390/ijms24033017

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