Next Article in Journal
Synthesis and Crystal Structures of Mn(II) and Co(II) Complexes as Catalysts for Oxidation of Cyclohexanone
Next Article in Special Issue
Structure and Vibrational Spectra of Pyridine Solvated Solid Bis(Pyridine)silver(I) Perchlorate, [Agpy2ClO4]·0.5py
Previous Article in Journal
Nucleophilic Substitution at a Coordinatively Saturated Five-Membered NHC∙Haloborane Centre
Previous Article in Special Issue
Synthesis, Physicochemical, Thermal and Antioxidative Properties of Zn(II) Coordination Compounds with Pyrazole-Type Ligand
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Investigations on the Spin States of Two Mononuclear Iron(II) Complexes Based on N-Donor Tridentate Schiff Base Ligands Derived from Pyridine-2,6-Dicarboxaldehyde

1
Department of Industrial Chemistry, College of Applied Sciences, Addis Ababa Science and Technology University, Addis Ababa P.O. Box 16417, Ethiopia
2
Nanotechnology Center of Excellence, Addis Ababa Science and Technology University, Addis Ababa P.O. Box 16417, Ethiopia
3
Department of Inorganic Chemistry, Faculty of Natural Sciences, Comenius University, Mlynská dolina CH-2, Ilkovičova 6, 84215 Bratislava, Slovakia
4
Institute of Quantum Materials and Technologies, Karlsruhe Institute of Technology, Hermann-von-Helmholtz-Platz 1, 76344 Karlsruhe, Germany
5
Institute of Nanotechnology, Karlsruhe Institute of Technology, Hermann-von-Helmholtz-Platz 1, 76344 Karlsruhe, Germany
6
Department of Inorganic Chemistry, Faculty of Science, Charles University in Prague, Hlavova 2030/8, 12843 Prague, Czech Republic
7
Institute of Applied Physics, Vienna University of Technology, Wiedner Hauptstraße 8-10, 1040 Vienna, Austria
8
Centre Européen de Science Quantique (CESQ), Institut de Science et d’Ingénierie Supramoléculaires (ISIS, UMR 7006), CNRS-Université de Strasbourg, 8 Allée Gaspard Monge, BP 70028, CEDEX, 67083 Strasbourg, France
*
Authors to whom correspondence should be addressed.
Submission received: 31 May 2022 / Revised: 6 July 2022 / Accepted: 6 July 2022 / Published: 8 July 2022
(This article belongs to the Special Issue Metal Complexes with N-donor Ligands)

Abstract

:
Iron(II)-Schiff base complexes are a well-studied class of spin-crossover (SCO) active species due to their ability to interconvert between a paramagnetic high spin-state (HS, S = 2, 5T2) and a diamagnetic low spin-state (LS, S = 0, 1A1) by external stimuli under an appropriate ligand field. We have synthesized two mononuclear FeII complexes, viz., [Fe(L1)2](ClO4)2.CH3OH (1) and [Fe(L2)2](ClO4)2.2CH3CN (2), from two N6–coordinating tridentate Schiff bases derived from 2,6-bis[(benzylimino)methyl]pyridine. The complexes have been characterized by elemental analysis, electrospray ionization–mass spectrometry (ESI-MS), Fourier-transform infrared spectroscopy (FTIR), solution state nuclear magnetic resonance spectroscopy, 1H and 13C NMR (both theoretically and experimentally), single-crystal diffraction and magnetic susceptibility studies. The structural, spectroscopic and magnetic investigations revealed that 1 and 2 are with Fe–N6 distorted octahedral coordination geometry and remain locked in LS state throughout the measured temperature range from 5–350 K.

1. Introduction

Ever since the very first report on spin-crossover (SCO) compounds in early 1931 [1], numerous reports have been devoted to this spectacular field of molecular magnetism [2,3,4]. Among these, octahedral FeII compounds has received special attention due to the clear discrimination between the paramagnetic high spin state (S = 2, 5T2) and diamagnetic low spin state (S = 0, 1A1), occurring with external stimuli in an appropriate ligand field [5,6,7,8]. Intermolecular interactions, such as π-π stacking or hydrogen bonding, usually enhance the SCO behavior with abrupt transitions and hysteresis loops [9,10].
Nano-sized SCO complexes are highly relevant for considering future applications and offer diverse pathways towards multifunctional systems for molecular memory, switches or display devices [10,11,12,13,14,15,16,17,18]. For acquiring a desired property, it is key to tune the ligand field, which could eventually modulate the magnetic properties. It has been observed that a N6– [19,20] or N4O2– [21] coordination environment around FeII can bring about sharp SCO with appreciable hysteresis width [22,23]. The effects of halogen substitution on the SCO behavior have been reported in FeII–N6 compounds [24,25,26,27,28,29,30], and the spin transition temperature, T1/2, is found to increase upon moving from fluoride to bromide substitution, thereby proving the size effect on the spin state of the complexes [20,24,31].
Previously, we have reported on how N6–coordination and variation in the ligand field in a series of bispyrazolone derivatives brought abrupt SCO at around room temperature [32]. Additionally, in our previous reviews, the role of azomethine and substituent effects in tuning the SCO behavior and attaining the desired architecture have been inferred [8,33]. In this context, Schiff bases are ideal candidates on account of their fine tunability to the ligand field by varying the substituents in both amine and aldehydic precursors.
Recently, we have observed that N6–coordination in FeII complexes from four azomethine and two pyridine nitrogens locks the spin state completely to a low spin condition even with an electron-donating methyl group attached to the meta position [34].
On account of the above facts and considering ligand-field and electronic effects, we are herein reporting our investigations on the structural and spin-state of two mononuclear FeII Schiff base complexes having N6–coordination. The ligands were designed without a substituent in L1 (2,6-bis[(benzylimino)methyl]pyridine) and with an electron-withdrawing chloro substituent at the meta position in L2 (2,6-bis[(3_chlorobenzylimino)methyl]pyridine) in comparison with the electron-donating methyl group of the previous reports [34,35].

2. Results and Discussion

2.1. Schiff Bases L1 and L2 and Their Complexes 1 and 2

Both ligands were prepared by condensation reaction, where benzylamine or 3-chlorobenzylamine were allowed to condense with pyridine-2,6-dicarboxaldehyde in ethanol under reflux (Scheme 1). Single crystals of L1, appropriate for X-ray diffraction studies, were grown by the slow evaporation of the solvent from a methanolic solution of L1 at room temperature. However, we were unsuccessful in generating single crystals for L2.
The corresponding complexes 1 and 2 were synthesized by the reaction of the tridentate ligands L1 and L2 with Fe(ClO4)2·6H2O in methanol for 1 and acetonitrile for 2, respectively (Scheme 1). Upon slow diffusion of diethylether into mother liquor at room temperature, black block crystals were obtained.
The 1H and 13C NMR spectra exhibit key resonances at 8.42 and 162.4 ppm, respectively, for L1, and 8.43 and 163.0 ppm, respectively, for L2, assigned to the imine (CH=N) group. The 1H and 13C NMR resonances of pyridine moiety are observed between 7.97–7.79 and 154.6–121.8 ppm, respectively, for L1, whereas for L2, they are between 8.00–7.81 and 154.4–122.0 ppm, respectively (see Table 1 and Table 2 for more detailed assignments and Tables S1 and S2 in Supplementary Materials for DFT computed NMR chemical shifts).
Upon coordination of L1 or L2 with FeII, the imine 1H NMR peaks are low-frequency shifted to 7.73 ppm (1) and 7.84 ppm (2), while the imine carbons are deshielded to 170.1 ppm (1) and 171.1 ppm (2), thus by more than +7.7 ppm. Even larger 1H shielding effects upon FeII complexation are observed for benzylic CH2 groups (Δδ(1H) = ca. −1.1 ppm) and ortho-hydrogens (Δδ(1H) = ca. −0.8 ppm). Contrarily, the largest coordination-induced 1H deshieldings (Δδ(1H)~+0.6 ppm) are seen for hydrogens of the pyridine moiety at the position 4 (py-4). Apart from the imine 13C nuclei, the coordination-induced deshieldings are also observed for pyridine-2/-6 and pyridine-3/-5 carbons, while benzylic C-1 carbons on the phenyl ring possess the most pronounced coordination-induced shielding (Δδ(13C) = −6 to −7 ppm).
The IR peaks were observed at 1644 cm−1 and 1651 cm−1 for L1 and L2, respectively, confirming the presence of CH=N. The IR spectra of L1 and L2 exhibit peaks at 1570 and 1594 cm−1, respectively, corresponding to the pyridine C=N stretching vibration [36]. Additionally, the IR spectra of 1 and 2 show a strong band at 1603, 1528 cm−1 (1) and 1600, 1529 cm−1 (2), confirming the coordination of the azomethine and pyridine nitrogen atoms to the metal centers [37,38]. Moreover, the elemental analysis and ESI-MS measurements are also in conformity with the molecular formulae assigned, [Fe(L1)2](ClO4)2·CH3OH and [Fe(L2)2](ClO4)2·2CH3CN for 1 and 2, respectively.

2.2. X-ray Crystallographic Analysis

The crystallographic data of the ligand L1 and complexes 1 and 2 are collated in Table 3; bond lengths (Table 4) and bond angles (Table S3) of 1 and 2 are also presented. For L1, X_ray quality crystals were grown by slow evaporation of its methanolic solution. The compound crystallizes in a triclinic lattice and space group P 1 ¯ with two symmetrically independent molecules located in the asymmetric part of the unit cell (Z = 4). The crystal structure is shown in Figure 1.
The single crystal X-ray diffraction data for the complexes 1 and 2 were collected at 120 K. Compound 1 crystallizes in a monoclinic crystal lattice with space group C2/c, whereas 2 crystalizes in a triclinic lattice with space group P 1 ¯ . An asymmetric unit of 1 consists of one discrete [FeL2]2+ cation, two ClO4 ions and one methanol solvent molecule (Figure 2a). However, the asymmetric units of 2 contain one discrete [FeL2]2+ cation, two ClO4 ions and two acetonitrile solvent molecules (Figure 2b). The unit cell of 1 contains eight complex units, sixteen counter anions and eight solvent molecules, whereas 2 contains two complex units, four counter anions and four solvent molecules (Figure S2c,d).
In both 1 and 2, the coordination sphere is composed of six nitrogen donors. Two pyridine nitrogen atoms of both L1 and L2 occupy the axial positions and the four azomethine nitrogens occupy the equatorial plane (Scheme 1). Two tridentate ligands are wrapped around the FeII metal centers in a distorted octahedral geometry, with average Fe–N bond lengths of 1.9505 Å for 1 and 1.9643 Å for 2 (Table 4). These distances are indicative of LS FeII centers, which is in accordance with magnetic investigations, vide infra [39,40,41]. The average Fe–Npy and Fe–Nimine bond distances are 1.8797(14) Å and 1.9860(15) Å, respectively, for 1, and 1.878(5) Å and 1.987(5) Å, respectively, for 2 (Table 4), which coincide very well with DFT optimized parameters for [Fe(L1)2]2+ (d(Fe–Npy)avrgd = 1.881 Å; d(Fe–Nimine)avrgd = 1.983 Å) and [Fe(L2)2]2+ (d(Fe–Npy)avrgd = 1.884 Å; d(Fe–Nimine)avrgd = 1.989 Å) in singlet (S = 0) states (see Supplementary Materials for TPSSh-D3(BJ)/def2-TZVP optimized geometries). The axial angles N(2)–Fe(1)–N(5) 178.92(6)° for 1 and 178.8(2)° for 2 indicate a clear distortion from the linear arrangement in both molecules (Table S3).
Hydrogen bonding and π-π stacking were found to have a marked influence on the magnetic properties of crystalline FeII complexes when they directly bridge individual ligands [42]. Upon inspection of the intermolecular interaction in the molecular packing of 1 and 2, it was found that complex units form weak π-π interaction through phenyl rings in the ligands on the bc plane for 1 and ab plane for 2, with an angle, centroid–centroid distances and shift distances of 3.140°, 3.715 and 1.810 Å, respectively, for 1, and 2.321°, 3.840 and 1.676 Å, respectively, for 2 (Figure S2a,b). Moreover, hydrogen bonding occurs between the oxygen atom of perchlorate counteranion and hydrogen atom of methanol, ClO1⋯H9 with a distance of 1.969(13) Å for 1, whereas for 2, the disordered perchlorate counteranions interact via hydrogen bonding with the CH2 group of the ligand moiety, O1E⋯H7A with a distance of 2.17(3) Å (Figure S2c,d). Hydrogen bonding interactions that are mediated by counteranions, as in our case, show negligible effects on the spin state of the metal center and are thereby locked in the LS state [43,44]. As regards the intermolecular interaction of 1, the solvent CH3OH forms short contacts with ClO4 ions with an average distance of 2.65(3) Å and the molecular packing of 2 shows that weak contact exists through the CH2 group of the ligand moiety and ClO4 ion with an average distance of 2.58(2) Å [45] (Figure S2c,d).

2.3. Magnetic Studies

The temperature dependence of the molar magnetic susceptibility for 1 and 2 was measured with a SQUID magnetometer (MPMS XL, Quantum Design) in the DC mode at BDC = 0.1 T. It was converted to the dimensionless product function. Its temperature dependence is shown in Figure 3a (1) and Figure 3b (2). Magnetic susceptibility measurements show that both complexes are completely locked in spin-paired diamagnetic states in the whole temperature range measured. This observation is in accordance with the average Fe–N bond observed—1.9505(15) Å for 1 and 1.950(5) Å for 2—which is characteristic of FeII in a low spin state [20,27]. However, the very small positive susceptibility values observed in the temperature dependance of the molar susceptibility curves for 1 and 2 may be due to the temperature-independent paramagnetism. The diamagnetic nature of both FeII complexes is also supported in solution by characteristic, well-resolved 1H and 13C NMR peaks. These chemical shifts are in excellent accord with those computed at the TPSSh-D3(BJ)/def2-TZVP level for closed-shell (S = 0) species (see Tables S1 and S2 and Figure S1 in Supplementary Materials). High spin complexes in a quintet (S = 2) state are computed at the same level to be energetically disfavored by 76.0 kJ/mol and 72.9 kJ/mol for 1 and 2, respectively. This relatively large energy gap can explain the locking of these complexes in the closed-shell state (S = 0) over a wide temperature range. A similar spin-paired, diamagnetic state has been observed for a N6–coordination (two azomethine and one pyridyl nitrogen from each ligand) in [FeL2]2+ Schiff base complexes [35]. Our previous investigations revealed that the spin state cannot be changed by introducing an electron-donating methyl group to the ligand system mentioned above with N6–coordination [34]. With the same type of hexa coordinate ligand systems, it has been observed that, if the pyridyl nitrogen coordination is replaced by imidazole nitrogen, the compound exhibits SCO behavior [46]. It is worth noting that, with the similar type of coordination environment reported by Alberto et al. [47] and Ishida et al. [25], the HS state and SCO become stabilized by increasing the size of the halogen substituent. Moreover, Gu et al. have shown that the electron-donating methyl substitution, counteranion or solvent have a negligible influence on the SCO behavior [20,39], with the same type of N6–ligand field, which is in accordance with the results that we have obtained.

3. Materials and Methods

All chemicals and reagents were purchased from commercial sources and were of analytical reagent grade, used without further purification. All complexation reactions were carried out under nitrogen atmosphere. FTIR spectra were measured on an Agilent Technologies Cary 630 FTIR spectrometer in the 4000–400 cm−1 range. NMR spectra were recorded with Bruker AscendTM (Billerica, MA, USA) 400 (400 MHz for 1H and 101 MHz for 13C) instruments in CD3CN with tetramethylsilane (TMS) as an internal reference. Elemental analyses were carried out on a FLASH elemental analyzer 1112 CHNS-O (Thermo Finnigan Italia, Rodano, Italy). Melting points were determined on a Büchi Melting Point M-565 apparatus. Single crystal XRD measurements were performed with a Bruker D8 VENTURE Kappa Duo diffractometer equipped with a PHOTON III detector and using monochromatic CuKα primary radiation. The phase problem was solved by intrinsic phasing (SHELXT) [48] and the structure model was refined by full-matrix least-squares on F2 values (SHELXL) [49]. Magnetic susceptibility measurements were carried out on a SQUID magnetometer (Quantum Design MPMS XL, San Diego, CA, USA) operating between 5 and 350 K with magnetic fields 1 kOe. The data were corrected for the intrinsic diamagnetic contributions of the sample and the sample holder. Electrospray ionization–mass spectrometry was performed using ESI-ToF Mass spectrometer (Bruker Daltonics micrOTOF-Q II).
CAUTION. Handling with metal–organic perchlorates is potentially dangerous due to their explosive properties. It should be handled with care in small quantities. Especially high temperature magnetic measurements are risky.

3.1. Synthesis

3.1.1. Synthesis of the Schiff Bases L1 and L2

The tridentate Schiff base ligands L1 ((E,E)-2,6-bis[(benzylimino)methyl)pyridine) were prepared by modified literature procedures [50,51], where a solution of benzylamine (0.65 mL, 6 mmol) and L2 ((E,E)-2,6-bis[(3-chlorobenzylimino)methyl]pyridine) 3-chlorobenzylamine (0.90 mL, 6 mmol) in ethanol (15 mL) were mixed with a stirred solution of pyridine-2,6-dicarbaldehyde (0.41 g, 3 mmol) in hot ethanol (15 mL). The mixtures were refluxed for 4 h (Scheme 1), while no precipitation occurred. The resultant solutions were concentrated to 10 mL and agitated thoroughly with a small amount of petroleum ether at 70–80 °C, which led to the precipitation of the ligands. These were then filtered, washed with petroleum ether and dried in vacuum. Pale orange (L1) and off-white (L2) needle-like crystalline products were collected.
L1: Yield; 85%; M. P. 75 °C; Anal. C21H19N3: Calcd. C 80.51, H 6.07, N 13.42; found C 80.46, H 6.08, N 13.51%; 1H NMR (CD3CN): δ(ppm) = 8.42 (s, 2H, CH-azomethine), 7.97 (d, J = 7.8 Hz, 2H, pyridine), 7.79 (t, J = 7.8 Hz, 1H, pyridine), 7.30 (m, 4H, Ar), 7.29 (m, 4H, Ar), 7.21 (m, 2H, Ar) and 4.89 (s, 4H, 2 × CH2). 13C NMR (CD3CN): δ(ppm) = 162.35 (C=N, azomethine), 154.55 (C-N, pyridine), 139.32, 137.49, 128.51, 128.19, 127.05, 121.83, 64.31 (Figure S3a,b). IR: ν ¯ (cm−1) = 3055(w), 3026(w), 2839(w), 1644(s), 1570(s), 1493(m), 1450(s), 1354(w), 1154(m), 1073(m), 1027(s), 991(m), 806(m) and 730(s).
L2: Yield; 47.3%; M. P. 74.3 °C; Anal. C21H17N3Cl2: Calcd. C 65.96, H 4.45, N 10.99; found C 65.80, H 4.61 N 10.87%; M. Pt. 63.4; 1H NMR (CD3CN): δ(ppm) = 8.43 (s, 2H, azomethine), 8.00 (d, J = 7.8 Hz, 2H, pyridine), 7.81 (t, J = 7.8 Hz, 1H, pyridine), 7.34 (bs, 2H, Ar), 7.29–7.20 (m, 6H, Ar) and 4.77 (s, 4H, 2 × CH2). 13C NMR (CD3CN): δ(ppm) = 162.99 (C=N, azomethine), 154.43 (C-N, pyridine), 141.83, 137.56, 133.77, 130.13, 127.97, 127.00, 126.56, 122.03, 63.41 (Figure S4a,b). IR: ν ¯ (cm−1) =3060(w), 3026(w), 2839(w), 1651(s), 1594(s), 1571(s), 1472(s), 1429(s), 1335(w), 1200(m), 1157(m), 1074(s), 995(m), 864(m) and 744(s).

3.1.2. Synthesis of the Complex [Fe(L1)2](ClO4)2·CH3OH (1)

To a stirred solution of L1 (0.313 g, 1.0 mmol) in methanol (20 mL), Fe(ClO4)2·6H2O (0.181 g, 0.5 mmol) was added under nitrogen atmosphere. The resultant purple solution was refluxed for 2 h under continuous stirring (Scheme 1). It was then cooled and filtered. Slow diffusion of diethyl ether vapor into the filtered solution for two weeks afforded block-shaped black crystals suitable for X-ray diffraction measurements. The crystals were then filtered and washed with cold methanol and dried subsequently under vacuum overnight. Yield: 0.266 g, 58.3%; Anal. C43H42Cl2FeN6O9, Calcd.: C 56.55; H 4.60; N 9.21; Found: C 56.77; H 4.78; N 9.16%. 1H NMR(CD3CN): δ(ppm) = 8.38 (t, J = 7.9 Hz, 2H, pyridine), 8.13 (d, J = 7.9 Hz, 4H, pyridine), 7.73 (s, 4H, azomethine), 7.29 (t, J = 7.4 Hz, 4H, Ar), 7.16 (t, J = 7.7 Hz, 8H, Ar), 6.47 (d, J = 7.6 Hz, 8H, Ar) and 3.68 (s, 8H, 4 × CH2). 13C NMR(CD3CN): δ(ppm) = 170.05, 160.27, 137.43, 133.27, 129.34, 129.28, 128.70, 128.46 and 62.39 (Figure S5a,b). FTIR: (cm−1) = 3021(w),1603(s), 1585(m), 1528(s), 1485(s), 1382(s), 1208(m), 1159(m), 1072(s), 980(m), 821(m), 740(s), 688(s), 620(s) and 450(s). ESI-MS: 781.20 (M+-ClO4), and 681.24 (M2+) (Figure S7a).

3.1.3. Synthesis of the Complex [Fe(L2)2](ClO4)2·2CH3CN (2)

The ligand L2 (0.191 g, 0.5 mmol) was dissolved in acetonitrile (20 mL) and this solution was mixed with Fe(ClO4)2·6H2O (0.090 g, 0.25 mmol) under nitrogen atmosphere. The resultant purple solution was refluxed for 2 h under continuous stirring (Scheme 1). The reaction mixture was then cooled and filtered. Slow diffusion of diethyl ether into the filtered solution yielded blocks of black crystals appropriate for X-ray diffraction. These were then filtered off, washed with cold acetonitrile and subsequently dried under vacuum. Yield: 0.213 g, 77.2%; Anal. C46H40Cl6FeN8O8, Calcd.: C, 50.12; H, 3.63; N, 10.17; Found: C, 49.98; H, 3.74; N, 10.18%. 1H NMR(CD3CN): δ(ppm) = 8.50 (t, J = 7.9 Hz, 2H, pyridine), 8.23 (d, J = 7.9 Hz, 4H, pyridine), 7.84 (s, 4H, azomethine), 7.33 (d, J = 9.5 Hz, 4H, Ar), 7.17 (t, J = 9.5 Hz, 4H, Ar), 6.48 (bs, 8H, Ar) and 3.74 (s, 8H, 4 × CH2). 13C NMR(CD3CN): δ(ppm) = 171.14, 160.36, 137.75, 135.49, 134.50, 130.94, 129.61, 128.74, 128.62, 127.26 and 61.79 (Figure S6a,b). FTIR: ν ¯ (cm−1) = 3044(w), 1600(s), 1564(s), 1529(s), 1474(s), 1396(s), 1355(m), 1194(m), 1075(s), 928(m), 852(m), 789(s), 705(s), 678(s), 614(s) and 442(s). ESI-MS: 917.04 (M+–ClO4_) and 818.09 (M2+) (Figure S7b).

3.2. Computational Details

The structures of all systems under investigation were fully optimized (without counteranion) in Turbomole [52] at the TPSSh level of theory, [53] including an atom-pairwise correction for dispersion forces (Grimme’s D3 model) with Becke–Johnson (BJ) damping [54,55] and employing the def2-TZVP basis set for all atoms [56]. The optimized structures were characterized as true minima on the potential energy hypersurface by harmonic vibrational frequency analyses. Calculations of NMR nuclear shieldings were performed in the Gaussian 16 program package [57] using gauge-including atomic orbitals (GIAO) at the same level as structure optimization (TPSSh-D3(BJ)/def2-TZVP). In these calculations, bulk solvent effects were simulated by means of the integral equation formalism of the polarizable continuum model (IEF-PCM) [58]. The calculated 1H and 13C shieldings were converted to chemical shifts (δ in ppm) relative to the shieldings of tetramethylsilane (TMS).

4. Conclusions

Two mononuclear FeII complexes, [Fe(L1)2](ClO4)2·CH3OH (1) and [Fe(L1)2](ClO4)2·2CH3CN (2), based on two unsymmetrical tridentate Schiff base ligands, were synthesized and characterized. Both complexes show distorted octahedral coordination geometries. Spectroscopic, magnetic and structural studies revealed that the spin states of both complexes remain diamagnetic throughout the measured temperature range. The ligand field created by N6-coordination was comprised of four azomethine and two pyridine nitrogen favors, thus showing a low spin FeII state. It is notable that the variation of solvents (acetonitrile and/or methanol) did not influence the magnetic properties of 1 and 2. Introducing an electron-withdrawing chlorine substituent into the meta position of L2 did not alter the ligand field, and hence, there is no change in the spin state. We hope the observed results are of particular importance for further design of molecular magnetic materials and encourage the development of new FeII Schiff base SCO systems. Further investigations by varying the ligand field and making substitutions in the ligand moiety are under way.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/inorganics10070098/s1, Figure S1: Comparison of calculated and experimental NMR shifts in [FeL2]2+ (S = 0) complexes (cf. Tables S1 and S2 for numeric data); Figure S2: Projection of the π-π interaction through phenyl rings of 1 (a) along the bc plane and 2 (b) along the ab plane. ClO4-ions, solvents and hydrogen atoms have been omitted for clarity. Short intermolecular contacts and H-bonding 1 (c) and 2 (d) both in b-direction; Figure S3: (a) 1H NMR and (b) 13C NMR of L1; Figure S4: (a) 1H NMR and (b) 13C NMR of L2; Figure S5: (a) 1H NMR and (b) 13C NMR of 1; Figure S6: (a) 1H NMR and (b) 13C NMR of 2; Figure S7: ESI-MS molecular ion peaks of 1 (a) and 2 (b); Table S1: Experimental and computed 1H NMR shifts (in ppm vs. TMS) in free ligands L1, L2 and corresponding [Fe(L1)2]2+ and [Fe(L2)2]2+ complexes (all in CD3CN); Table S2: Experimental and computed 13C NMR shifts (in ppm vs. TMS) in free ligands L1, L2 and corresponding [Fe(L1)2]2+ and [Fe(L2)2]2+ complexes (all in CD3CN); Table S3: Coordination bond angles for 1 and 2 at 120 K.

Author Contributions

The major work for this article, designing, execution and writing of the original draft, was done by the first author (Y.B.), which is part of his Ph.D. program. The second and third authors, N.S. and S.S. have done SQUID and ESI-MS measurements. The fourth author, R.G. solved the structures. The other authors, W.L., M.R., A.S., P.H. and M.T. were responsible for supervision, editing, and reviewing of this article. All authors have read and agreed to the published version of the manuscript.

Funding

This research project was supported by the National Scholarship Programme of the Slovak Republic 2021 and Addis Ababa Science and Technology University, Addis Ababa, Ethiopia. This work was also supported by the Slovak Research and Development Agency (Grant No. APVV-17-0324) and the Grant Agency of the Ministry of Education of the Slovak Republic (VEGA Project No. 1/0669/22).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We are thankful to National Scholarship Programme of the Slovak Republic 2021 for a research stay and Woldia University, Ethiopia, for a Ph.D studentship to Y.B. Thanks are due to Patrik Osuský, Department of Inorganic Chemistry, Comenius University, Slovakia, for spectroscopic measurements. We are also grateful to Jozef Noga, Department of Inorganic Chemistry, Comenius University, Slovakia, for his keen interest and support during this work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cambi, L.; Szego, L. The Magnetic Susceptibility of Complex Compounds. Ber. Dtsch. Chem. Ges 1931, 64, 2591–2598. [Google Scholar] [CrossRef]
  2. Gutlich, P.; Goodwin, H. Spin Crossover in Transition Metal Compounds I–III. In Topics in Current Chemistry; Springer: New York, NY, USA; Berlin/Heidelberg, Germany, 2004; pp. 2–356. ISBN 3540403949. [Google Scholar]
  3. Miller, J.S.; Gatteschi, D. Molecule-Based Magnets Themed Issue. Chem. Soc. Rev. 2011, 40, 3313–3335. [Google Scholar] [CrossRef]
  4. Senthil Kumar, K.; Ruben, M. Emerging Trends in Spin Crossover (SCO) Based Functional Materials and Devices. Coord. Chem. Rev. 2017, 346, 176–205. [Google Scholar] [CrossRef]
  5. Sugimoto, K.; Okubo, T.; Maekawa, M.; Kuroda-Sowa, T. Visualization of Weak Interaction Effects on N2O Schiff Base Ligands in Iron(II) Spin Crossover Complexes. Cryst. Growth Des. 2021, 21, 4178–4183. [Google Scholar] [CrossRef]
  6. Aromi, G.; Real, J.A. Special Issue “Spin Crossover (SCO) Research”. Magnetochemistry 2016, 2, 2–4. [Google Scholar] [CrossRef]
  7. Swart, M.; Costas, M. Spin States in Biochemistry and Inorganic Chemistry Influence on Structure and Reactivity, 1st ed.; Swart, M.C., Ed.; John Wiley & Sons, Ltd.: Chichester, UK, 2016; ISBN 9781118898314. [Google Scholar]
  8. Senthil Kumar, K.; Bayeh, Y.; Gebretsadik, T.; Elemo, F.; Gebrezgiabher, M.; Thomas, M.; Ruben, M. Spin Crossover in Iron(II) Schiff Base Complexes. Dalton Trans. 2019, 48, 15321–15337. [Google Scholar] [CrossRef]
  9. Zhong, Z.J.; Tao, J.Q.; Zhi, Y.; Dun, C.Y.; Liu, Y.J.; You, X.Z. A Stacking Spin Crossover Iron(II) Compound with a Large Hysteresis. J. Chem. Soc. Dalton Trans. 1998, 2, 327–328. [Google Scholar] [CrossRef]
  10. Real, J.A.; Gaspar, A.B.; Carmen Muñoz, M. Thermal, Pressure and Light Switchable Spin Crossover Materials. Dalton Trans. 2005, 12, 2062–2079. [Google Scholar] [CrossRef]
  11. Halcrow, M.A. Spin-Crossover Materials Properties and Applications; John Wiley & Sons, Ltd.: Chichester, UK, 2013. [Google Scholar]
  12. Antonio, J.; Ana, R.; Carmen, B.G.M.; Gütlich, P.; Ksenofontov, V.; Spiering, H.; Chemie, A. Bipyrimidine-Bridged Dinuclear Iron(II) Spin Crossover Compounds. Top. Curr. Chem. 2004, 233, 167–193. [Google Scholar] [CrossRef]
  13. Brooker, S. Spin Crossover with Thermal Hysteresis: Practicalities and Lessons Learnt. Chem. Soc. Rev. 2015, 44, 2880–2892. [Google Scholar] [CrossRef] [Green Version]
  14. Halcrow, M.A. Structure: Function Relationships in Molecular Spin Crossover Complexes W. Chem. Soc. Rev. 2011, 40, 4119–4142. [Google Scholar] [CrossRef] [PubMed]
  15. Gamez, P.; Costa, J.S.; Quesada, M.; Aromi, G. Iron Spin Crossover Compounds: From Fundamental Studies to Practical Applications. Dalton Trans. 2009, 38, 7845–7853. [Google Scholar] [CrossRef] [PubMed]
  16. Murray, K.S.; Kepert, C.J. Cooperativity in Spin Crossover Systems: Memory, Magnetism and Microporosity. Top. Curr. Chem. 2006, 233, 195–228. [Google Scholar] [CrossRef]
  17. Weber, B.; Bauer, W.; Obel, J. An Iron (II) Spin Crossover Complex with a 70 K Wide Thermal. Angew. Chem. Int. Ed. 2008, 47, 10098–10101. [Google Scholar] [CrossRef] [PubMed]
  18. Gaspar, A.B.; Mun, M.C. Dinuclear Iron (II) Spin Crossover Compounds: Singular Molecular Materials for Electronics. J. Mater. Chem. 2006, 16, 2522–2533. [Google Scholar] [CrossRef]
  19. Schäfer, B.; Bauer, T.; Faus, I.; Wolny, J.A.; Dahms, F.; Fuhr, O.; Lebedkin, S.; Wille, H.C.; Schlage, K.; Chevalier, K.; et al. A Luminescent Pt2Fe Spin Crossover Complex. Dalton Trans. 2017, 46, 2289–2302. [Google Scholar] [CrossRef] [PubMed]
  20. Wang, Y.T.; Li, S.T.; Wu, S.Q.; Cui, A.L.; Shen, D.Z.; Kou, H.Z. Spin Transitions in Fe(II) Metallogrids Modulated by Substituents, Counteranions, and Solvents. J. Am. Chem. Soc. 2013, 135, 5942–5945. [Google Scholar] [CrossRef] [PubMed]
  21. Weber, B. Spin Crossover Complexes with N4O2 Coordination Sphere-The Influence of Covalent Linkers on Cooperative Interactions. Coord. Chem. Rev. 2009, 253, 2432–2449. [Google Scholar] [CrossRef]
  22. Halcrow, M.A. Spin-Crossover Compounds with Wide Thermal Hysteresis. Chem. Lett. 2014, 43, 178–1188. [Google Scholar] [CrossRef]
  23. Lochenie, C.; Bauer, W.; Railliet, A.P.; Schlamp, S.; Garcia, Y.; Weber, B. Large Thermal Hysteresis for Iron(II) Spin Crossover Complexes with N-(Pyrid-4-Yl)Isonicotinamide. Inorg. Chem. 2014, 53, 11563–11572. [Google Scholar] [CrossRef]
  24. Phonsri, W.; Macedo, D.S.; Vignesh, K.R.; Rajaraman, G.; Davies, C.G.; Jameson, G.N.L.; Moubaraki, B.; Ward, J.S.; Kruger, P.E.; Chastanet, G.; et al. Halogen Substitution Effects on N2O Schiff Base Ligands in Unprecedented Abrupt Fe(II) Spin Crossover Complexes. Chem. Eur. J. 2017, 23, 7052–7065. [Google Scholar] [CrossRef] [PubMed]
  25. Kimura, A.; Ishida, T. Spin Crossover Temperature Predictable from DFT Calculation for Iron(II) Complexes with 4-Substituted Pybox and Related Heteroaromatic Ligands. ACS Omega 2018, 3, 6737–6747. [Google Scholar] [CrossRef]
  26. Mochida, N.; Kimura, A.; Ishida, T. Spin Crossover Hysteresis of [FeII (LHiPr)2(NCS)2](LHiPr = N-2-Pyridylmethylene-4 Isopropylaniline) Accompanied by Isopropyl Conformation Isomerism. Magnetochemistry 2015, 1, 17–27. [Google Scholar] [CrossRef]
  27. Halcrow, M.A. The Spin-States and Spin Transitions of Mononuclear Iron(II) Complexes of Nitrogen-Donor Ligands. Polyhedron 2007, 26, 3523–3576. [Google Scholar] [CrossRef]
  28. Kershaw Cook, L.J.; Kulmaczewski, R.; Mohammed, R.; Dudley, S.; Barrett, S.A.; Little, M.A.; Deeth, R.J.; Halcrow, M.A. A Unified Treatment of the Relationship between Ligand Substituents and Spin State in a Family of Iron (II) Complexes. Angew. Chem. Int. Ed. 2016, 55, 4327–4331. [Google Scholar] [CrossRef]
  29. Létard, J.F.; Guionneau, P.; Nguyen, O.; Costa, J.S.; Marcén, S.; Chastanet, G.; Marchivie, M.; Goux-Capes, L. A Guideline to the Design of Molecular-Based Materials with Long-Lived Photomagnetic Lifetimes. Chem. Eur. J. 2005, 11, 4582–4589. [Google Scholar] [CrossRef] [PubMed]
  30. Yamasaki, M.; Ishida, T. Heating-Rate Dependence of Spin Crossover Hysteresis Observed in an Iron(II) Complex Having Tris (2-Pyridyl)Methanol. J. Mater. Chem. C 2015, 3, 7784–7787. [Google Scholar] [CrossRef]
  31. Phonsri, W.; Harding, D.J.; Harding, P.; Murray, K.S.; Moubaraki, B.; Gass, I.A.; Cashion, J.D.; Jameson, G.N.L.; Adams, H. Stepped Spin Crossover in Fe(III) Halogen Substituted Quinolylsalicylaldimine Complexes. Dalton Trans. 2014, 43, 17509–17518. [Google Scholar] [CrossRef]
  32. Madhu, N.T.; Salitros, I.; Schramm, F.; Klyatskaya, S.; Fuhr, O.; Ruben, M. Above Room Temperature Spin Transition in a Series of Iron(II) Bis(Pyrazolyl)Pyridine Compounds. C. R. Chim. 2008, 11, 1166–1174. [Google Scholar] [CrossRef]
  33. Salitros, I.; Madhu, N.T.; Boca, R.; Pavlik, J.; Ruben, M. Room-Temperature Spin-Transition Iron Compounds. Monatsh. Chem. 2009, 140, 695–733. [Google Scholar] [CrossRef]
  34. Bayeh, Y.; Osusky, P.; Gyepes, R.; Yutronkie, N.J.; Sergawie, A.; Hrobarik, P.; Clerac, R.; Madhu, T. Low-Spin Mononuclear Iron(II) Complexes Based on Tridentate N-Donor Schiff Base Ligand Derived from Pyridine-2,6-dicarboxaldehyde. Polyhedron, 2022; under preparation. [Google Scholar]
  35. Stratton, W.J.; Busch, D.H. The Complexes of Pyridinaldazine with Iron(II) and Nickel(II). J. Am. Chem. Soc. 1958, 80, 1286–1289. [Google Scholar] [CrossRef]
  36. Nakamoto, K. Infrared and Raman Spectra of Inorganic and Coordination Compounds Part B: Theory and Applications; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 1978; Volume 156, ISBN 9780471744931. [Google Scholar]
  37. Nakamoto, K. Infrared and Raman Spectra of Inorganic and Coordination Compounds: Part A: Theory and Applications in Inorganic Chemistry, 6th ed.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2008. [Google Scholar]
  38. Hagiwara, H.; Tanaka, T.; Hora, S. Synthesis, Structure, and Spin Crossover above Room Temperature of a Mononuclear and Related Dinuclear Double Helicate Iron(II) Complexes. Dalton Trans. 2016, 45, 17132–17140. [Google Scholar] [CrossRef] [PubMed]
  39. Han, W.K.; Li, Z.H.; Zhu, W.; Li, T.; Li, Z.; Ren, X.; Gu, Z.G. Molecular Isomerism Induced Fe(II) Spin State Difference Based on the Tautomerization of the 4(5)-Methylimidazole Group. Dalton Trans. 2017, 46, 4218–4224. [Google Scholar] [CrossRef] [PubMed]
  40. Milin, E.; Benaicha, B.; El Hajj, F.; Patinec, V.; Triki, S.; Marchivie, M.; Gómez-García, C.J.; Pillet, S. Magnetic Bistability in Macrocycle-Based Fe(II) Spin Crossover Complexes: Counter Ion and Solvent Effects. Eur. J. Inorg. Chem. 2016, 2016, 5282. [Google Scholar] [CrossRef] [Green Version]
  41. Marchivie, M.; Guionneau, P.; Létard, J.F.; Chasseau, D. Photo-Induced Spin-Transition: The Role of the Iron(II) Environment Distortion. Acta Crystallogr. Sect. B Struct. Sci. 2005, 61, 25–28. [Google Scholar] [CrossRef] [Green Version]
  42. Létard, J.F.; Guionneau, P.; Rabardel, L.; Howard, J.A.K.; Goeta, A.E.; Chasseau, D.; Kahn, O. Structural, Magnetic, and Photomagnetic Studies of a Mononuclear Iron(II) Derivative Exhibiting an Exceptionally Abrupt Spin Transition. Light-Induced Thermal Hysteresis Phenomenon. Inorg. Chem. 1998, 37, 4432–4441. [Google Scholar] [CrossRef]
  43. Ikuta, Y.; Ooidemizu, M.; Yamahata, Y.; Yamada, M.; Osa, S.; Matsumoto, N.; Iijima, S.; Sunatsuki, Y.; Kojima, M.; Dahan, F.; et al. A New Family of Spin Crossover Complexes with a Tripod Ligand Containing Three Imidazoles. Inorg. Chem. 2003, 42, 7001–7017. [Google Scholar] [CrossRef]
  44. Matouzenko, G.S.; Molnar, G.; Bréfuel, N.; Perrin, M.; Bousseksou, A.; Borshch, S.A. Spin Crossover Iron(II) Coordination Polymer with Zigzag Chain Structure. Chem. Mater. 2003, 15, 550–556. [Google Scholar] [CrossRef]
  45. Buron-Le Cointe, M.; Hébert, J.; Baldé, C.; Moisan, N.; Toupet, L.; Guionneau, P.; Létard, J.F.; Freysz, E.; Cailleau, H.; Collet, E. Intermolecular Control of Thermoswitching and Photoswitching Phenomena in Two Spin-Crossover Polymorphs. Phys. Rev. B 2012, 85, 39–41. [Google Scholar] [CrossRef] [Green Version]
  46. Craze, A.R.; Howard-Smith, K.J.; Bhadbhade, M.M.; Mustonen, O.; Kepert, C.J.; Marjo, C.E.; Li, F. Investigation of the High-Temperature Spin-Transition of a Mononuclear Iron(II) Complex Using X-Ray Photoelectron Spectroscopy. Inorg. Chem. 2018, 57, 6503–6510. [Google Scholar] [CrossRef]
  47. Rodríguez-Velamazán, J.A.; Carbonera, C.; Castro, M.; Palacios, E.; Kitazawa, T.; Létard, J.F.; Burriel, R. Two-Step Thermal Spin Transition and LIESST Relaxation of the Polymeric Spin Crossover Compounds Fe(X-Py)2[Ag(CN)2]2 (X = H, 3-Methyl, 4-Methyl, 3,4-Dimethyl, 3-Cl). Chem. Eur. J. 2010, 16, 8785–8796. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Sheldrick, G.M. SHELXT—Integrated Space-Group and Crystal-Structure Determination. Acta Crystallogr. Sect. A Found. Crystallogr. 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Sheldrick, G.M. Crystal Structure Refinement with SHELXL. Acta Crystallogr. Sect. C Struct. Chem. 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  50. Lewkowski, J.; Dziȩgielewski, M. Stereochemistry of Phosphite Addition to Azomethine Bond of Achiral 2,6-Pyridinedicarbaldimines and Isophthalaldimines—A Comparative Study. Phosphorus Sulfur Silicon Relat. Elem. 2013, 188, 995–1006. [Google Scholar] [CrossRef]
  51. Scrimin, P.; Tecilla, P.; Tonellato, U.; Valle, G.; Veronese, A. Metal Ions Co-Operativity in the Catalysis of the Hydrolysis of a β-Amino Ester by a Macrocyclic Dinuclear Cu (II) Complex. Tetrahedron 1995, 51, 527–538. [Google Scholar] [CrossRef]
  52. TURBOMOLE, Version 7.5; TURBOMOLE GmbH, a Development of the University of Karlsruhe and Forschungszentrum Karlsruhe. 2020. Available online: http://www.turbomole.com (accessed on 29 May 2022).
  53. Tao, J.; Perdew, J.P.; Staroverov, V.N.; Scuseria, G.E. Climbing the Density Functional Ladder: Nonempirical Meta–Generalized Gradient Approximation Designed for Molecules and Solids. Phys. Rev. Lett. 2003, 91, 3–6. [Google Scholar] [CrossRef] [Green Version]
  54. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A Consistent and Accurate Ab Initio Parametrization of Density Functional Dispersion Correction (DFT-D) for the 94 Elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef] [Green Version]
  55. Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the Damping Function in Dispersion Corrected Density Functional Theory. J. Comput. Chem. 2011, 32, 1456–1465. [Google Scholar] [CrossRef]
  56. Weigend, F.; Ahlrichs, R. Balanced Basis Sets of Split Valence, Triple Zeta Valence and Quadruple Zeta Valence Quality for H to Rn: Design and Assessment of Accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297–3305. [Google Scholar] [CrossRef]
  57. 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.; Petersson, H.; et al. Gaussian 16; revision C.01; Gaussian, Inc.: Wallingford, UK, 2016. [Google Scholar]
  58. Tomasi, J.; Mennucci, B.; Cammi, R. Quantum Mechanical Continuum Solvation Models. Chem. Rev. 2005, 105, 2999–3093. [Google Scholar] [CrossRef]
Scheme 1. Synthetic route for ligands L1, L2 and complexes 1 and 2.
Scheme 1. Synthetic route for ligands L1, L2 and complexes 1 and 2.
Inorganics 10 00098 sch001
Figure 1. X-ray crystal structure of L1 molecule in bc-axis; Color code: blue, N; grey, C; and white, H.
Figure 1. X-ray crystal structure of L1 molecule in bc-axis; Color code: blue, N; grey, C; and white, H.
Inorganics 10 00098 g001
Figure 2. Labelled ORTEP drawing of 1 (a) and 2 (b) both in b-axis. Thermal ellipsoids are drawn on 30% probability level. Hydrogen atoms have been omitted for clarity. Color code: Yellow-green, Fe; blue, N; grey, C; green, Cl and red, O.
Figure 2. Labelled ORTEP drawing of 1 (a) and 2 (b) both in b-axis. Thermal ellipsoids are drawn on 30% probability level. Hydrogen atoms have been omitted for clarity. Color code: Yellow-green, Fe; blue, N; grey, C; green, Cl and red, O.
Inorganics 10 00098 g002
Figure 3. The χMT vs. T magnetic plots for 1 (a) and 2 (b) between 5–350 K.
Figure 3. The χMT vs. T magnetic plots for 1 (a) and 2 (b) between 5–350 K.
Inorganics 10 00098 g003
Table 1. Experimental 1H NMR shifts (in ppm vs. TMS) in free ligands L1, L2 and corresponding [Fe(L1)2]2+ and [Fe(L2)2]2+ complexes (all measured in CD3CN) a.
Table 1. Experimental 1H NMR shifts (in ppm vs. TMS) in free ligands L1, L2 and corresponding [Fe(L1)2]2+ and [Fe(L2)2]2+ complexes (all measured in CD3CN) a.
H-Iminepy-3,5py-4CH2H-2/H-6H-3/H-5H-4
L18.427.977.794.797.307.297.21
[Fe(L1)2]2+7.738.138.383.686.477.167.29
Δδ(1H) b−0.69+0.16+0.59−1.11−0.83−0.13+0.08
H-Iminepy-3,5py-4CH2H-2H-4H-5H-6
L28.438.007.814.777.347.257.257.25
[Fe(L2)2]2+7.848.238.503.746.487.177.336.48
Δδ(1H) b−0.59+0.23+0.69−1.03−0.86−0.08+0.08−0.77
a See SI for corresponding NMR spectra and computed chemical shifts. b 1H NMR coordination shifts as a difference between resonance of given 1H nuclei in complex and the free ligand. See Scheme 1 for atom numbering.
Table 2. Experimental 13C NMR shifts (in ppm vs. TMS) in free ligands L1, L2 and corresponding [Fe(L1)2]2+ and [Fe(L2)2]2+ complexes (all measured in CD3CN) a.
Table 2. Experimental 13C NMR shifts (in ppm vs. TMS) in free ligands L1, L2 and corresponding [Fe(L1)2]2+ and [Fe(L2)2]2+ complexes (all measured in CD3CN) a.
C-Iminepy-2,6py-3,5py-4CH2C-1C-2/C-6C-3/C-5C-4
L1162.4154.6121.8137.564.3139.3128.5128.2127.1
[Fe(L1)2]2+170.1160.3128.5137.462.4133.3128.7129.3129.3
Δδ(13C) b+7.7+5.7+6.6−0.1−1.9−6.0+0.2+1.2+2.2
C-Iminepy-2,6py-3,5py-4CH2C-1C-2C-3C-4C-5C-6
L2163.0154.4122.0133.863.4141.8128.0137.6127.0130.1126.6
[Fe(L2)2]2+171.1160.4128.6135.561.8134.5128.7137.8130.9129.6127.3
Δδ(13C) b+8.1+5.9+6.6+1.7−1.6−7.3+0.8+0.2+3.9−0.5+0.7
a See SI for corresponding NMR spectra and computed chemical shifts. b 13C NMR coordination shifts as a difference between resonance of given 13C nuclei in complex and the free ligand. See Scheme 1 for atom numbering.
Table 3. Collated crystal parameters data for L1, 1 and 2.
Table 3. Collated crystal parameters data for L1, 1 and 2.
ParameterL112
Empirical formulaC21H19N3C43H42Cl2FeN6O9C46H40Cl6FeN8O8
Formula weight313.39912.441101.41
Temperature120(2) K
Wavelength0.71073 Å
Crystal systemTriclinicMonoclinicTriclinic
Space group P 1 ¯ C2/c P 1 ¯
Unit cell dimensionsa = 8.9002(19) Å
b = 10.289(2) Å
c = 19.181(4)Å
α = 92.691(10)°
β = 99.378(9)°
γ = 100.078(10)°
a = 37.8848(9) Å
b = 10.5130(3) Å
c = 21.4945(5) Å
α = 90°
β = 109.7270(10)°γ = 90°
a = 10.1735(10) Å
b = 10.2669(9) Å
c = 23.149(2) Å
α = 92.685(3)°
β = 101.466(3)°
γ = 90.261(3)°
Volume1701.1(7) Å38058.5(4) Å32366.9(4) Å3
Z482
Calculated density1.224 g/cm31.509 g/cm31.545 g/cm3
Absorption coefficient0.073 mm−10.573 mm−10.721 mm−1
Crystal size0.321 × 0.299 × 0.050 mm30.248 × 0.220 × 0.104 mm30.427 × 0.358 × 0.309 mm3
Theta range for data collection2.016° to 26.570°2.020° to 27.493°2.043° to 27.573°
Limiting indices−11 ≤ h ≤ 10;
−12 ≤ k ≤ 12;
0 ≤ l ≤ 24
−48 ≤ h ≤ 49;
−13 ≤ k ≤ 13;
−23 ≤ l ≤ 27
−13 ≤ h ≤ 13;
−12 ≤ k ≤ 13;
−30 ≤ l ≤ 30
Reflections collected/unique7284/728454461/924271645/10877
Completeness to θ: fraction25.242°: 100.0%25.242°: 99.9%25.242°: 99.7%
Absorption correctionSemi-empirical from equivalents
Max. and
min. transmission
1.00 and 0.650.95 and 0.870.91 and 0.77
Refinement methodFull-matrix least-squares on F2
Data/restraints/parameters7284/0/4349242/6/63910877/25/683
Goodness-of-fit on F21.0201.0471.144
Final R indices (I > 2σ(I))R1 = 0.0717;
wR2 = 0.1517
R1 = 0.0354;
wR2 = 0.0779
R1 = 0.1089
wR2 = 0.2256
R indices (all data)R1 = 0.1317;
wR2 = 0.1819
R1 = 0.0434;
wR2 = 0.0818
R1 = 0.1308;
wR2 = 0.2359
Largest diff. peak and hole0.221 and −0.251 × 10−3 Å0.572 and −0.463 × 10−3 Å0.859 and −0.894 × 10−3 Å
Spin states-LSLS
CCDC number216584321549042128830
Table 4. Coordination bond lengths for 1 and 2 at 120 K.
Table 4. Coordination bond lengths for 1 and 2 at 120 K.
Fe1–N Bond Lengths (Å) at 120 K
12
Fe(1)–N(2)1.8794(14)Fe(1)–N(2)1.875(5)
Fe(1)–N(5)1.8799(14)Fe(1)–N(5)1.880(5)
Fe(1)–N(3)1.9763(15)Fe(1)–N(4)1.978(5)
Fe(1)–N(4)1.9846(14)Fe(1)–N(3)1.981(5)
Fe(1)–N(6)1.9891(16)Fe(1)–N(6)1.990(5)
Fe(1)–N(1)1.9938(15)Fe(1)–N(1)1.998(5)
Average Fe1–N1.9505(15) 1.950(5)
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bayeh, Y.; Suryadevara, N.; Schlittenhardt, S.; Gyepes, R.; Sergawie, A.; Hrobárik, P.; Linert, W.; Ruben, M.; Thomas, M. Investigations on the Spin States of Two Mononuclear Iron(II) Complexes Based on N-Donor Tridentate Schiff Base Ligands Derived from Pyridine-2,6-Dicarboxaldehyde. Inorganics 2022, 10, 98. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics10070098

AMA Style

Bayeh Y, Suryadevara N, Schlittenhardt S, Gyepes R, Sergawie A, Hrobárik P, Linert W, Ruben M, Thomas M. Investigations on the Spin States of Two Mononuclear Iron(II) Complexes Based on N-Donor Tridentate Schiff Base Ligands Derived from Pyridine-2,6-Dicarboxaldehyde. Inorganics. 2022; 10(7):98. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics10070098

Chicago/Turabian Style

Bayeh, Yosef, Nithin Suryadevara, Sören Schlittenhardt, Róbert Gyepes, Assefa Sergawie, Peter Hrobárik, Wolfgang Linert, Mario Ruben, and Madhu Thomas. 2022. "Investigations on the Spin States of Two Mononuclear Iron(II) Complexes Based on N-Donor Tridentate Schiff Base Ligands Derived from Pyridine-2,6-Dicarboxaldehyde" Inorganics 10, no. 7: 98. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics10070098

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