Next Article in Journal
Tyrosol Derivatives, Bearing 3,5-Disubstituted Isoxazole and 1,4-Disubstituted Triazole, as Potential Antileukemia Agents by Promoting Apoptosis
Next Article in Special Issue
Sodium β-Diketonate Glyme Adducts as Precursors for Fluoride Phases: Synthesis, Characterization and Functional Validation
Previous Article in Journal
Pressurized Liquid Extraction as a Novel Technique for the Isolation of Laurus nobilis L. Leaf Polyphenols
Previous Article in Special Issue
Metalloporphyrin Metal–Organic Frameworks: Eminent Synthetic Strategies and Recent Practical Exploitations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

High-Temperature Spin Crossover in Iron(II) Complexes with 2,6-Bis(1H-imidazol-2-yl)pyridine

by
Ludmila G. Lavrenova
1,*,
Olga G. Shakirova
1,2,*,
Evgeniy V. Korotaev
1,
Svetlana V. Trubina
1,
Alexsei Ya. Tikhonov
3,
Irina A. Os’kina
3,
Sergey A. Petrov
4,
Konstantin Yu. Zhizhin
5 and
Nikolay T. Kuznetsov
5
1
Nikolaev Institute of Inorganic Chemistry, Siberian Branch, Russian Academy of Sciences, Academic Lavrentyev Avenue 3, 630090 Novosibirsk, Russia
2
Department of Chemistry and Chemical Technologies, Faculty of Machinery and Chemical Technologies, Federal State Budget Institution of Higher Education, Komsomolsk-na-Amure State University, Lenin Avenue 27, 681013 Komsomolsk-on-Amur, Russia
3
N.N. Vorozhtsov Novosibirsk Institute of Organic Chemistry Siberian Branch, Russian Academy of Sciences, Academic Lavrentyev Avenue 9, 630090 Novosibirsk, Russia
4
Institute of Solid State Chemistry, Siberian Branch, Russian Academy of Sciences, Kutateladze Street 18, 630128 Novosibirsk, Russia
5
Kurnakov Institute of General and Inorganic Chemistry, Russian Academy of Sciences, Leninskii Avenue 31, 119991 Moscow, Russia
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(16), 5093; https://doi.org/10.3390/molecules27165093
Submission received: 29 June 2022 / Revised: 4 August 2022 / Accepted: 5 August 2022 / Published: 10 August 2022

Abstract

:
Novel iron(II) coordination compounds containing a ligand 2,6-bis(1H-imidazol-2-yl)pyridine (L), having such a composition as [FeL2]SO4·0.5H2O, [FeL2]Br2·H2O, [FeL2](ReO4)2, [FeL2]B10H10∙H2O, [FeL2]B12H12∙1.5H2O had been synthesized and studied using UV-Vis (diffuse reflectance), infrared, extended X-ray absorption fine structure (EXAFS), and Mössbauer spectroscopy methods, as well as X-ray diffraction and static magnetic susceptibility methods. The analysis of the μeff(T) dependence in the temperature range of 80–600 K have shown that all the obtained complexes exhibit a high-temperature spin crossover 1A15T2.

Graphical Abstract

1. Introduction

Searching for novel coordination compounds wherein the spin state of the central atom can be switched by an external action is an urgent problem. This is indicated by regularly appearing publications in the literature devoted to this topic [1,2,3,4,5,6,7,8]. These compounds include iron(II) complexes with spin crossover (SCO) 1A15T2. The change in spin multiplicity occurs owing to affecting temperature, pressure, irradiation with light of a certain wavelength (LIESST effect), high-frequency magnetic or electric field, and other factors. The transition from a low-spin (LS) to a high-spin (HS) state causes a change in the magnetic, optical, and vibrational properties of the complexes. An important property associated with the spin transition consists in changing metal–donor interatomic distance amounting to 0.2 Å in the case of Fe(II) complexes. Owing to the universal properties of complexes with SCO they have a wide range of potential applications in making optoelectronic, molecular electronic, and spintronic devices [9,10,11,12,13,14,15]. At present, polyfunctional materials combining SCO and other properties are under active study [16,17,18,19].
2,6-Bis(1H-imidazol-2-yl)pyridines represent a promising class of potential ligands for the synthesis of complexes with SCO [20]. Earlier, we have reported the studies on iron(II) complexes with 2,6-bis(benzimidazol-2-yl)- and 2,6-bis(4,5-dimethyl-1H-imidazol-2-yl)pyridine [21,22,23,24], wherein SCO is exhibited. It seemed worthwhile to continue these investigations into synthesizing and studying Fe(II) complexes with 2,6-bis(1H-imidazol-2-yl)pyridine (Scheme 1), which does not contain substituents at the imidazole fragment. Earlier, this ligand was used in order to synthesize complexes with Ru(II), and with a number of metals belonging to the first transition series [25,26,27,28,29]. The X-ray diffraction data have shown that L is coordinated to the metal ion in a tridentate-cyclic manner by two nitrogen atoms belonging to two imidazole rings and one nitrogen atom belonging to pyridine.

2. Materials and Methods

2.1. Materials

Commercial metal salts and solvents without further purification were used in the synthesis. 2,6-Bis(imidazol-2-yl)pyridine was synthesized as described in [30] (NMR spectra of ligand given in the Supplementary Materials, Figures S1–S4); K2B10H10·2H2O, K2B12H12 were obtained according to the procedure [31].

2.2. Synthesis of [FeL2]SO4 0.5H2O (1·0.5H2O)

A 0.28 g (1 mmol) sample of FeSO4·7H2O with the addition of 0.1 g of ascorbic acid was dissolved in 10 mL of water under heating; a 0.51 g (2 mmol) L was dissolved in 10 mL of ethanol, and the solutions were then heated and mixed. The resulting solution was evaporated until a red-violet precipitate began to form. After the solution with the precipitate was cooled in a crystallizer with ice, the precipitate was filtered off, washed twice with small portions of water, and dried in air. Yield: 55%. Anal. Calc. for C22H19FeN10O9/2S, (583.4): C, 45.3; H, 3.3; N, 24.0. Found: C, 46.2; H, 3.5; N, 23.4.

2.3. Synthesis of [FeL2]Br2·H2O (2·H2O) and [FeL2](ReO4)2 (3)

A 0.28 g (1 mmol) sample of FeSO4·7H2O was dissolved in 5 mL of distilled water acidified with 0.1 g of ascorbic acid. An excess (0.43 g, 1.5 mmol) of KReO4 or KBr (0.36 g, 3 mmol) in 10 mL of water and a solution of L (0.51 g, 2 mmol) in 10 mL of ethanol were successively added to the resulting solution under stirring. The resulting solution was evaporated until a red-violet precipitate began to form. After the solution with the precipitate was cooled in a crystallizer with ice, the precipitate was filtered off, washed twice with small portions of water, and dried in air. Yield: 70% (2·H2O), 30% (3). Anal. Calc. for C22H20Br2FeN10O, (656.1): C, 40.3; H, 3.1; N, 21.3. Found: C, 41.3; H, 3.2; N, 21.0. Anal. Calc. for C22H18FeN10O8Re2, (978.7): C, 27.0; H, 1.9; N, 14.3. Found: C, 27.5; H, 2.2; N, 14.6.

2.4. Synthesis of [FeL2]B10H10∙H2O (4∙H2O) and [FeL2]B12H12∙1.5H2O (5·1.5H2O)

A 0.14 g (0.5 mmol) sample of FeSO4·7H2O was dissolved in 3 mL of distilled water acidified with 0.05 g of ascorbic acid. An excess (0.23 g, 1 mmol) of K2B10H10∙2H2O or K2B12H12 (0.22 g, 1 mmol) in 10 mL of water and a solution of L (0.21 g, 1 mmol) in 5 mL of ethanol were added to the resulting solution under stirring. Red-brown precipitates began to form immediately. Each precipitate was filtered off, washed twice with small portions of water and ethanol, and dried in air. Yield: 45% (4∙H2O), 68% (5∙1.5H2O). Anal. Calc. for C22H30B10FeN10O, (614.5): C, 43.0; H, 4.9; N, 22.8. Found: C, 42.5; H, 5.0; N, 22.1. Anal. Calc. for C22H33B12FeN10O3/2, (647.1): C, 40.8; H, 5.1; N, 21.6. Found: C, 40.6; H, 5.3; N, 20.9.

2.5. Measurement and Characterization

The data for elemental analysis of the complexes was acquired on a EURO EA 3000 analyzer (EuroVector, Pavia, Italy).
X-ray absorption spectra (XAS) in the Fe K edge region (150 eV before and 800 eV after) were measured in the transmission mode with the use of synchrotron radiation on the 8 beamline, VEPP-3 storage ring at the Siberian Synchrotron and Terahertz Radiation Center (Budker Institute of Nuclear Physics, Siberian Branch, Russian Academy of Sciences, Novosibirsk, Russia) [32]. A Si(111) cut-off crystal was used as a two-crystal monochromator. The operating mode of the storage ring during measurement: energy—2 GeV; current—70–140 mA. For measurements, the samples were mixed with cellulose powder as a filler and pressed into tablets. The mass of the sample was calculated so that the absorption jump at the Fe K-edge was 0.8–1. The preprocessing of the absorption spectra (selection of the oscillating part—EXAFS spectra) was performed with the use of the VIPER 10.17 program [33]. The “radial pair distribution function” (Figure 1) was obtained by the Fourier transform of the k3-weighted EXAFS function in the range of wave vectors k = 2.0–11.0 Å−1.
The local environment of the Fe ion [interatomic distances (Ri) and angles (Qi), coordination numbers (Ni), and Debye–Waller factors (σ2)] was modeled using the EXCURVE 98 cod [34]. In this program phase and amplitude characteristics were calculated in the von-Bart and Hedin approximation. The amplitude suppression factor S02 due to multielectron processes was determined for the crystallographically characterized compound (S02 = 0.85) and fixed during further modeling of the studied compounds spectra. The Debye–Waller factor was the same separately for nitrogen and carbon atoms.
IR spectra were taken on a Scimitar FTS 2000 in the range of 4000–400 cm−1 and a Vertex 80 in the range of 600–100 cm−1. Compound samples were prepared as suspensions in vaseline and fluorinated oils and in polyethylene.
The Kubelka-Munk diffuse reflectance spectra were obtained on a Shimadzu UV-3101 PC scanning spectrometer.
The XRD investigation of polycrystalline samples was performed using a Shimadzu XRD 7000 diffractometer (CuKα radiation).
The static magnetic susceptibility was measured using Faraday balance type setup equipped with electromagnetic compensating torsion quartz microbalance. The Delta DTB9696 temperature controller (Delta Electronics Inc., Taipei, Taiwan) was used for the investigated compounds temperature stabilization (~1 K) in the range of temperatures 80–600 K. The temperature scanning rate for the process of heating or cooling samples was ~1 K/min. The magnetic field strength (7300 Oe) stabilization precision was ~2%. The compounds studied were sealed in quartz cellules filled with atmospheric air at 760 Torr. In order to study the effect of crystallization water, the samples were placed in open quartz ampoules and vacuumed at 10−2 Torr, after which the helium atmosphere at 5 Torr was formed.
The effective magnetic moment of studied compounds was calculated as:
μ e f f = ( 3 k N A μ B 2 χ T ) 1 2 ( 8 χ T ) 1 2 ,
k—Boltzmann constant, NA—Avogadro constant, and µB—Bohr magneton. In the above formula, the diamagnetic contribution in total magnetic susceptibility (χ) was taken into account using the method of Pascal’s constants. The direct (Tc↑) and reverse (Tc↓) transitions temperatures were obtained using condition the magnetic moment second derivative zero value (d2(µeff(T))/dT2 = 0).
The Mössbauer spectra were collected with a spectrometer NP-610 (KFKI, Budapest, Hungary), using 57Co in a metal Rh matrix as a radioactive source. The spectra were measured at a room temperature. The spectra were processed to find the values of isomer shift δ and quadrupole splitting ε. The isomer shifts are given relative to metal iron.

3. Results and Discussion

Iron(II) complexes with 2,6-bis(1H-imidazol-2-yl)pyridine were isolated from aqueous-ethanol solutions. To maintain the oxidation state of iron(II), ascorbic acid was added to the solution.
Complex 1·0.5H2O was synthesized by reacting stoichiometric amounts of FeSO4 and L. Syntheses 2·H2O, 3, 4∙H2O, 5·1.5H2O were carried out in two stages. At the first stage, the corresponding iron(II) salts were obtained by adding a 1.5- or 2-fold excess of KBr, KReO4, K2B10H10 or K2B12H12 into a solution of FeSO4 salt. At the second stage, a solution of the ligand in ethanol was added to the resulting salt solution. The obtained diffraction patterns show that the synthesized compounds are crystalline, while [FeL2]SO4·0.5H2O and [FeL2]Br2·H2O, as well as [FeL2]B10H10∙H2O and [FeL2]B12H12∙1.5H2O are isotypical.
The structure of the whole molecule for all complexes in the LS state (T = 297 °C) was established from the EXAFS data using the multiple scattering approximation within the software package EXCURV (except for [FeL2](ReO4)2 (3)) excluding hydrogen atoms, and anions that exert only a weak effect on the shape of the EXAFS spectrum owing to their spatial separation from the central iron ion. For complex 3, the signal-to-noise ratio in the EXAFS spectrum is worse than that for the spectra of other complexes owing to the presence of a heavy anion ReO4. Simulation of the whole molecule in the multiple scattering approximation gives too large errors in determining the parameters. Therefore, the simulation of the EXAFS spectrum for complex 3 could be carried out only in a single scattering approximation. The structure of the complexes in the LS state (by the example of complex 4∙H2O) obtained by means of EXAFS spectra simulation is presented in Figure 2.
Table 1 lists the microstructure parameters of the coordination site for complexes 1·0.5H2O, 2·H2O, 4∙H2O, 5·1.5H2O.
The structure of the coordination site of complex [FeL2](ReO4)2 (3) in the LS state was obtained from modeling the spectrum filtered in real space (ΔR = 0.9 to 3.0 Å). The simulation data are presented in Table 2. The coordination numbers of the nearest spheres of the iron ion environment were fixed in accordance with the data obtained in the simulation of complexes in the LS state in the multiple scattering approximation.
Figure 3 illustrates a comparison of the experimental and simulated radial distribution function for [FeL2]B10H10∙H2O. The simulation was carried out using a multiple scattering approximation.
Table 3 and Figures S5–S7 show the main vibration frequencies in the spectra of L and Fe(II) complexes. In the high-frequency spectral region for 1·0.5H2O, 2·H2O, 4∙H2O, 5·1.5H2O, ν(OH) vibrations are observed; in the wave number range of 3200–3050 cm−1 for all the complexes there are stretching vibrations of NH-groups, and in the range of 3100–2850 cm–1 there are vibrations of ν(CH). In the wave number range of 1650–1450 cm−1, there are stretching and bending vibration bands inherent in heterocyclic rings. The spectra of the complexes in the range of ring vibrations exhibit a change in the number and position of the imidazole and pyridine bands in comparison with the spectrum of the ligand, which indicates the coordination of nitrogen atoms of the rings to Fe(II) ions. The presence of non-split bands inherent in stretching vibrations characteristic of the SO42− and ReO4 anions indicates the fact that they are in outer-sphere position. The vibration bands of the B–H bonds of the outer-sphere anions B10H102− and B12H122− are centered at the wave numbers of 2470 (ν(BH)) and 1075 cm−1 (δ(BBH)). In the case of the spectra for 4∙H2O, 5·1.5H2O they are shifted with respect to those observed in the spectra of the initial salts, which could be, to all appearance, caused by the formation of H2Oδ− … δ+H–B bonds. In the far region of the spectra of all the complexes, Fe(3d6)–ligand(π) charge-transfer transition bands and vibration bands (M-N) are observed. The position of these bands is typical for the spectra of low-spin octahedral iron(II) complexes [35].
The diffuse reflectance spectra (DRS, Figures S8–S12) of all the obtained complexes exhibit intense metal-ligand charge transfer bands ν1(eg π L * ) in the wavelength range of 300–350 nm (λmax ≈ 324–326 nm). The DRS of 4∙H2O and 5∙1.5H2O exhibit three absorption bands, too, (see Table 4); they correspond to the 1A11T2, 1A11T1 and 1A11A2 transitions in the strong octahedral field of the ligands. For low-spin axially distorted octahedral iron(II) complexes, the term 1T1 is transformed according to the representation of 1E + 1A2, whereas term 1T2 is transformed according to the representation of 1E + 1B2 [36]. Therefore, the 1A11B2, 1A11A2 and 1A11E wide transition bands (see Table 4) observed in the spectra of 1·0.5H2O, 2·H2O and 3, indicate the fact that an axial distortion of the octahedral coordination of these complexes occur. Low-intensity transitions and overlapping three wide bands do not make it possible to perform reliable quantitative calculations of crystal field parameters in this case. The spectra of all the complexes do not exhibit the 5T25E band, which is caused by the HS state of iron(II). We were able to calculate the splitting parameters based on the difference between 1A11T2 and 1A11T1 absorption frequencies [36] for low-spin (LS) forms of complexes with closo-borate anions. The B values were computed using the formula 16B = [ν (1A11T2) − ν (1A11T1)]. The values C and ΔLS (Table 4) were calculated using the equations: νLS = ΔLS − C + 86B2LS and C = 4.41·B [36,37,38]. The obtained data show that 2,6-bis(1H-imidazol-2-yl)pyridine is a strong field ligand. In addition, these data, as well as the values obtained for a number of previously synthesized Fe(II) complexes with 2,6-bis(benzimidazol-2-yl)pyridine and 2,6-bis(4,5-dimethyl-1H-imidazole) [22,23,24], obey the inequality that reflects the condition for the manifestation of SCO [37]: 19.000 cm−1 ≤ ΔLS ≤ 22.000 cm−1.
The Mössbauer spectra of complexes 1·0.5H2O, 2·H2O, 3 and 5·1.5H2O represent quadrupole doublets whose parameters correspond to the LS state of iron(II) (Figure 4). The spectrum of 4·H2O also exhibits a broadened doublet related to the HS form of the complex (36%). The parameters of the Mössbauer spectra are presented in Table 5.
The temperature dependences of the effective magnetic moment for the analyzed complexes are shown in Figure 5. Spin crossover (SCO) is observed for all the compounds under investigation. In the case of complexes 1·0.5H2O, 2·H2O, 3 the µeff values observed in the thermal stability range (2.35, 2.1 and 3.35 µΒ, respectively), are significantly lower than the theoretical spin-only value of 4.9 µΒ for the Fe(II) ion. The 4∙H2O and 5∙1.5H2O complexes are more stable and thus a temperature of 600 K could be reached. For these compounds, a complete spin-crossover is observed. However, the µeff values observed in the HS state of these compounds are also lower than the theoretical value for Fe(II). It should be noted that the experimental µeff values for 4∙H2O and 5∙1.5H2O complexes are in the range of 4.6–5.7 observed for Fe(II) compounds [39,40]. In the case of 1·0.5H2O and 2·H2O the residual effective magnetic moment in LS state (0.65 and 0.4 µΒ, respectively), is presented by µeff(T) dependences. This fact could be connected with temperature-independent paramagnetism. Complexes 3, 4∙H2O and 5∙1.5H2O in LS state exhibit diamagnetism with a zero µeff value. Despite the fact that low µeff values are achieved in the investigated temperature range for 1·0.5H2O and 3, the condition of d2eff(T))/dT2 = 0 could be satisfied and some values of SCO temperature could be determined. Table 6 shows the temperature values of direct (Tc↑) and inverse (Tc↓) transitions. For the [FeL2]Br2·H2O one could suggest that the SCO temperature is higher than 420 K. The determined temperature of the inverse transition increases in a series of [FeL2](ReO4)2 → [FeL2]SO4·0.5H2O → FeL2]B10H10·H2O → [FeL2]B12H12·1.5H2O.
The effect of the crystallization of water has been studied for 4·H2O and 5·1.5H2O complexes (Figure 6). It should be noted that in the case of rarefied atmosphere the decomposition of the dehydrated complexes occurs in a lower temperature range than it is observed for initial compounds. Nevertheless, the SCO has been observed in this case, too. The µeff value of 4.65 µΒ achieved in HS state for 4 corresponds to the value observed for the initial complex. In the case of 5 complex, the µeff value (4.6 µΒ) exhibits an increase after dehydration. Residual µeff values (~1–1.5 µΒ) have been registered to occur for both complexes in the LS state. The SCO temperature increases after dehydration; however, the 5 complex demonstrates the highest SCO temperature values as it is observed in the case of the initial compound. Thus, the dehydration of 4·H2O and 5·1.5H2O complexes lead to appearing residual µeff value and to an increase in the temperature of SCO.

4. Conclusions

In this work, we have synthesized and investigated five novel coordination compounds of various iron(II) salts with 2,6-bis(1H-imidazol-2-yl)pyridine (L). The structure of the coordination core of the complexes has been determined by means of EXAFS spectra simulation. Two ligand molecules are coordinated to the iron(II) ion in a tridentate-cyclic fashion by the nitrogen atom belonging to pyridine and two nitrogen atoms belonging to imidazole rings. Thus, the complexes have the distorted-octahedral structure of the coordination polyhedron, the FeN6 core. The studies on the µeff(T) dependence have shown that complexes having such a composition as [FeL2]SO4·0.5H2O, [FeL2]Br2·H2O, [FeL2](ReO4)2, [FeL2]B10H10·H2O, [FeL2]B12H12·1.5H2O exhibit an 1A15T2 high-temperature spin crossover. A comparison of the data obtained for the synthesized compounds with those obtained by us earlier [21,22,23,24] shows that spin-crossover 1A15T2 is observed in all complexes of Fe(II) with 2,6-bis(imidazole-2-yl)pyridines. The temperatures of direct SCO, Tc↑, in most compounds are significantly higher than room temperature.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27165093/s1, Figure S1: NMR 1H of 2,6-di(1H-imidazol-2-yl)pyridine, 100 MHz, NS 24; Figure S2: NMR 13C of 2,6-di(1H-imidazol-2-yl)pyridine, 100 MHz, NS 96; Figure S3: NMR 13C of 2,6-di(1H-imidazol-2-yl)pyridine, 100 MHz, NS 7376; Figure S4: NMR 13C of 2,6-di(1H-imidazol-2-yl)pyridine, 100 MHz, NS 176; Figure S5: IR spectrum of 2,6-bis(1H-imidazol-2-yl)pyridine (L); Figure S6: IR spectrum of [FeL2]B10H10∙H2O (4∙H2O); Figure S7: IR spectrum of [FeL2]B12H12∙1.5H2O (5·1.5H2O); Figure S8: Comparison of the experimental DRS and model for complex [FeL2]SO4∙0.5H2O; Figure S9: Comparison of the experimental DRS and model for complex [FeL2]Br2∙H2O; Figure S10: Comparison of the experimental DRS and model for complex [FeL2](ReO4)2; Figure S11: DRS for complex [FeL2]B10H10∙H2O; Figure S12: DRS for complex [FeL2]B12H12∙1.5H2O.

Author Contributions

Manuscript conception, writing and original draft preparation, data analysis and interpretation, L.G.L.; synthesis of the studied complexes, diffuse reflectance and IR spectroscopy, O.G.S.; studies of magnetic properties, E.V.K.; EXAFS spectroscopy, S.V.T.; synthesis of ligand, A.Y.T. and I.A.O.; Mössbauer spectroscopy methods, S.A.P.; synthesis of the closo-borates, data analysis, K.Y.Z.; editing, supervision, N.T.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research work was partially supported by the Russian Science Foundation (grant No. 20-63-46026) and by the Ministry of Science and Higher Education of the Russian Federation (projects No. 121031700313-8, 121031700314-5, 121071500036-4, 1021051403061-8-1.4.1).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding authors.

Acknowledgments

All the EXAFS procedures were performed at the Siberian Synchrotron and Terahertz Radiation Center of the Novosibirsk VEPP-3 complex at Budker Institute of Nuclear Physics of the SB RAS (Novosibirsk, Russia). The authors thank V.V. Kriventsov for his assistance in XAS measurements. Authors would like to acknowledge the Multi Access Chemical Research Center SB RAS for spectral and analytical measurements.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are available from the authors on request.

References

  1. Gutlich, P.; Goodwin, H. Spin Crossover in Transition Metal Compounds I–III; Springer: Berlin/Heidelberg, Germany, 2004; pp. 233–235. [Google Scholar]
  2. Halcrow, M.A. Spin-Crossover Materials Properties and Applications; John Wiley & Sons Ltd.: Chichester, UK, 2013; p. 562. [Google Scholar]
  3. Shatruk, M.; Phan, H.; Chrisostomo, B.A.; Suleimenova, A. Symmetry-breaking structural phase transitions in spin crossover complexes. Coord. Chem. Rev. 2015, 289–290, 62–73. [Google Scholar] [CrossRef]
  4. Feltham, H.L.C.; Barltrop, A.S.; Brooker, S. Spin crossover in iron(II) complexes of 3,4,5-tri-substituted-1,2,4-triazole (Rdpt), 3,5-di-substituted-1,2,4-triazolate (dpt), and related ligands. Coord. Chem. Rev. 2017, 344, 26–53. [Google Scholar] [CrossRef]
  5. Scott, H.S.; Staniland, R.W.; Kruger, P.E. Spin crossover in homoleptic Fe(II) imidazolylimine complexes. Coord. Chem. Rev. 2018, 362, 24–43. [Google Scholar] [CrossRef]
  6. Kumar, K.S.; Ruben, M. Emerging trends in spin crossover (SCO) based functional materials and devices. Coord. Chem. Rev. 2017, 346, 176–205. [Google Scholar] [CrossRef]
  7. Lavrenova, L.G. Spin crossover in homo- and heteroligand iron(II) complexes with tris(pyrazol-1-yl)methane derivatives. Rus. Chem. Bull. 2018, 67, 1142–1152. [Google Scholar] [CrossRef]
  8. Shakirova, O.G.; Lavrenova, L.G. Spin crossover in new iron(II) coordination compounds with tris(pyrazol-1-yl)methane. Crystals 2020, 10, 843. [Google Scholar] [CrossRef]
  9. Kahn, O.; Krober, J.; Jay, C. Spin transition molecular materials for displays and data recording. Adv. Mater. 1992, 4, 718–728. [Google Scholar] [CrossRef]
  10. Gamez, P.; Costa, J.S.; Quesada, M.; Aromí, G. Iron Spin-Crossover compounds: From fundamental studies to practical applications. Dalton Trans. 2009, 7845–7853. [Google Scholar] [CrossRef] [PubMed]
  11. Hayami, S.; Holmes, S.M.; Halcrow, M.A. Spin-state switches in molecular materials chemistry. J. Mater. Chem. C 2015, 3, 7775–7778. [Google Scholar] [CrossRef]
  12. Kumar, K.S.; Bayeh, Y.; Gebretsadik, T.; Elemo, F.; Gebrezgiabher, M.; Thomas, M.; Ruben, M. Spin-crossover in iron(II)-Schiff base complexes. Dalton Trans. 2019, 48, 15321–15377. [Google Scholar] [CrossRef] [PubMed]
  13. Halcrow, M.A. Manipulating metal spin states for biomimetic, catalytic and molecular materials chemistry. Dalton Trans. 2020, 49, 15560–15567. [Google Scholar] [CrossRef] [PubMed]
  14. Letard, J.-F.; Daro, N.; Aymonier, C.; Cansell, F.; Saint-Martin, S. Nouveau Materiau a Transition de spin, son Procede de Preparation. European Patent 2391631, 7 December 2011. [Google Scholar]
  15. Bousseksou, A.; Vieu, C.; Letard, J.-F.; Demont, P.; Tuchagues, J.-P.; Malaquin, L.; Menegotto, J.; Salmon, L. Molecular Memory and Method for Making Same. European Patent 1430552, 2004. Available online: https://data.epo.org/gpi/EP1430552A1-MOLECULAR-MEMORY-AND-METHOD-FOR-MAKING-SAME (accessed on 28 June 2022).
  16. Piedrahita-Bello, M.; Angulo-Cervera, J.E.; Courson, R.; Molnár, G.; Malaquin, L.; Thibault, C.; Tondu, B.; Salmon, L.; Bousseksou, A. 4D printing with spin-crossover polymer composites. J. Mater. Chem. C 2020, 8, 6001–6005. [Google Scholar] [CrossRef]
  17. Nguyen, T.D.; Veauthier, J.M.; Angles-Tamayo, G.F.; Chavez, D.E.; Lapsheva, E.L.; Myers, T.W.; Nelson, T.R.; Schelter, E.J. Correlating mechanical sensitivity with spin transition in the explosive spin crossover complex [Fe(Htrz)3]n[ClO4]2n. J. Am. Chem. Soc. 2020, 142, 4842–4851. [Google Scholar] [CrossRef] [PubMed]
  18. Yuan, J.; Wu, S.Q.; Liu, M.J.; Sato, O.; Kou, H.Z. Rhodamine 6G-labeled pyridyl aroylhydrazone Fe(II) complex exhibiting synergetic spin crossover and fluorescence. J. Am. Chem. Soc. 2018, 140, 9426–9433. [Google Scholar] [CrossRef] [PubMed]
  19. Benaicha, B.; Van Do, K.; Yangui, A.; Pittala, N.; Lusson, A.; Sy, M.; Bouchez, G.; Fourati, H.; Gomez-Garcia, C.J.; Triki, S.; et al. Interplay between spin-crossover and luminescence in a multifunctional single crystal iron(II) complex: Towards a new generation of molecular sensors. Chem. Sci. 2019, 10, 6791–6798. [Google Scholar] [CrossRef] [Green Version]
  20. Boča, M.; Jameson, R.F.; Linert, W. Fascinating variability in the chemistry and properties of 2,6-bis-(benzimidazol-2-yl)-pyridine and 2,6-bis-(benzthiazol-2-yl)-pyridine and their complexes. Coord. Chem. Rev. 2011, 255, 290–317. [Google Scholar] [CrossRef]
  21. Lavrenova, L.G.; Dyukova, I.I.; Korotaev, E.V.; Sheludyakova, L.A.; Varnek, V.A. Spin crossover in new iron(II) complexes with 2,6-bis(benzimidazole-2-yl)pyridine. Rus. J. Inorg. Chem. 2020, 65, 30–35. [Google Scholar] [CrossRef]
  22. Ivanova, A.D.; Korotaev, E.V.; Komarov, V.Y.; Sheludyakova, L.A.; Varnek, V.A.; Lavrenova, L.G. Spin-crossover in iron(ii) coordination compounds with 2,6-bis(benzimidazol-2-yl)pyridine. New J. Chem. 2020, 44, 5834–5840. [Google Scholar] [CrossRef]
  23. Ivanova, A.D.; Lavrenova, L.G.; Korotaev, E.V.; Trubina, S.V.; Sheludyakova, L.A.; Petrov, S.A.; Zhizhin, K.Y.; Kuznetsov, N.T. High-temperature spin crossover in complexes of iron(II) closo-borates with 2,6-bis(benzimidazol-2-yl)pyridine. Russ. J. Inorg. Chem. 2020, 65, 1687–1694. [Google Scholar] [CrossRef]
  24. Ivanova, A.D.; Korotaev, E.V.; Komarov, V.Y.; Sukhikh, T.S.; Trubina, S.V.; Sheludyakova, L.A.; Petrov, S.A.; Tikhonov, A.Y.; Lavrenova, L.G. Spin crossover in iron(II) complexes with new ligand 2,6-bis(4,5-dimethyl-1H-imidazole-2-yl)pyridine. Inorg. Chim. Acta 2022, 532, 120746. [Google Scholar] [CrossRef]
  25. Stupka, G.; Gremaud, L.; Bernardinelli, G.; Williams, A.F. Redox state switching of transition metals by deprotonation of the tridentate ligand 2,6-bis(imidazol-2-yl)pyridine. Dalton Trans. 2004, 407–412. [Google Scholar] [CrossRef]
  26. Hammes, B.S.; Damiano, B.J.; Tobash, P.H.; Hidalgo, M.J.; Yap, G.P.A. Tuning the redox properties of a 1-D supramolecular array via selective deprotonation of coordinated imidazoles around a Mn(II) center. Inorg. Chem. Commun. 2005, 8, 513–516. [Google Scholar] [CrossRef]
  27. Arruda, E.G.R.; Farias, M.A.; Jannuzzi, S.A.V.; Almeida Gonsales, S.; Timm, R.A.; Sharma, S.; Zoppellaro, G.; Kubota, L.T.; Knobel, M.; Formiga, A.L.B. Synthesis, structural and magnetic characterization of a copper(II) complex of 2,6-di(1H-imidazol-2-yl)pyridine and its application in copper-mediated polymerization catalysis. Inorg. Chim. Acta 2017, 466, 456–463. [Google Scholar] [CrossRef]
  28. Ren, C.-X.; Li, S.-Y.; Yin, Z.-Z.; Lu, X.; Ding, Y.-Q. [2,6-Bis(4,5-dihydro-1H-imidazol-2-yl)pyridine]dichloridomanganese(II). Acta Cryst. E 2009, 65, m572–m573. [Google Scholar] [CrossRef] [Green Version]
  29. Shang, S.-M.; Ren, C.-X.; Wang, X.; Lu, L.-D.; Yang, X.-J. Bis[2,6-bis (4,5-dihydro-1H-imidazol-2-yl)pyridine]manganese(II) bis-(per-chlorate) acetonitrile solvate. Acta Cryst. E 2009, 65, m1023–m1024. [Google Scholar] [CrossRef] [Green Version]
  30. Voss, M.E.; Beer, C.M.; Mitchell, S.A.; Blomgren, P.A.; Zhichkin, P.E. A simple and convenient one-pot method for the preparation of heteroaryl-2-imidazoles from nitriles. Tetrahedron 2008, 64, 645–651. [Google Scholar] [CrossRef]
  31. Miller, H.C.; Muetterties, E.L.; Boone, J.L.; Garrett, P.; Hawthorne, M.F. Borane anions. Inorg. Synth. 1967, 10, 81–91. [Google Scholar] [CrossRef]
  32. Piminov, P.A.; Baranov, G.N.; Bogomyagkov, A.V.; Berkaev, D.E.; Borin, V.M.; Dorokhov, V.L.; Karnaev, S.E.; Kiselev, V.A.; Levichev, E.B.; Meshkov, O.I.; et al. Synchrotron radiation research and application at VEPP-4. Phys. Procedia 2016, 84, 19–26. [Google Scholar] [CrossRef]
  33. Klementev, K.V. Extraction of the fine structure from x-ray absorption spectra. J. Phys. D Appl. Phys. 2001, 34, 209. [Google Scholar] [CrossRef]
  34. Gurman, S.J.; Binsted, N.; Ross, I. A rapid, exact curved-wave theory for EXAFS calculations. J. Phys. C Solid State Phys. 1984, 17, 143–151. [Google Scholar] [CrossRef]
  35. Nakamoto, K. Infrared and Raman Spectra of Inorganic and Coordination Compounds; John Wiley & Sons Inc.: New York, NY, USA, 1986. [Google Scholar]
  36. Lever, A.B.P. Inorganic Electronic Spectroscopy, 2nd ed.; Elsevier: Amsterdam, The Netherlands, 1985; ISBN 978-0-444-42389-4. [Google Scholar]
  37. Hauser, A. Ligand Field Theoretical Considerations. Spin Crossover in Transition Metal Compounds I; Springer: Berlin/Heidelberg, Germany, 2004; Volume 233, pp. 49–58. [Google Scholar] [CrossRef]
  38. Sugano, S.; Tanabe, Y.; Kamimura, H. Multiplets of Transition—Metal ions in Crystals; Academic Press: New York, NY, USA; Pure and Applied Physics: London, UK, 1970. [Google Scholar]
  39. Selwood, P.W. Magnetochemistry; Interscience Publishers: New York, NY, USA, 1956. [Google Scholar]
  40. Rakitin, Y.V.; Kalinnikov, V.T. Modern Magnetochemistry; Nauka: St. Petersburg, Russia, 1994. [Google Scholar]
Scheme 1. 2,6-Bis(1H-imidazol-2-yl)pyridine (L).
Scheme 1. 2,6-Bis(1H-imidazol-2-yl)pyridine (L).
Molecules 27 05093 sch001
Figure 1. Fe K–edge EXAFS spectra (left) and their Fourier transform modules without consideration of the phase shift (right) of complexes 15.
Figure 1. Fe K–edge EXAFS spectra (left) and their Fourier transform modules without consideration of the phase shift (right) of complexes 15.
Molecules 27 05093 g001
Figure 2. General view of the structure of the coordination core obtained by means of EXAFS spectra simulation for such complexes as 1·0.5H2O, 2·H2O, 4∙H2O, 5·1.5H2O.
Figure 2. General view of the structure of the coordination core obtained by means of EXAFS spectra simulation for such complexes as 1·0.5H2O, 2·H2O, 4∙H2O, 5·1.5H2O.
Molecules 27 05093 g002
Figure 3. Comparison of the experimental and model radial distribution function for complex [FeL2]B10H10∙H2O.
Figure 3. Comparison of the experimental and model radial distribution function for complex [FeL2]B10H10∙H2O.
Molecules 27 05093 g003
Figure 4. Mössbauer spectra of complexes 1·0.5H2O, 2·H2O, 3, 4∙H2O, 5·1.5H2O (1–5).
Figure 4. Mössbauer spectra of complexes 1·0.5H2O, 2·H2O, 3, 4∙H2O, 5·1.5H2O (1–5).
Molecules 27 05093 g004
Figure 5. Dependences of µeff(T) (ae) and d2eff(T))/dT2 (fj) for 1·0.5H2O (a,f), 2·H2O (b,g), 3 (c,h), 4·H2O (d,i), 5·1.5H2O (e,j). The white and black triangles in the figure correspond to the samples cooling and heating processes respectively.
Figure 5. Dependences of µeff(T) (ae) and d2eff(T))/dT2 (fj) for 1·0.5H2O (a,f), 2·H2O (b,g), 3 (c,h), 4·H2O (d,i), 5·1.5H2O (e,j). The white and black triangles in the figure correspond to the samples cooling and heating processes respectively.
Molecules 27 05093 g005aMolecules 27 05093 g005b
Figure 6. Dependences of µeff(T) (a,b) and d2eff(T))/dT2 (c,d) for dehydrated complexes 4 (a,c), 5 (b,d). The white and black triangles in the figure correspond to the samples cooling and heating processes respectively.
Figure 6. Dependences of µeff(T) (a,b) and d2eff(T))/dT2 (c,d) for dehydrated complexes 4 (a,c), 5 (b,d). The white and black triangles in the figure correspond to the samples cooling and heating processes respectively.
Molecules 27 05093 g006
Table 1. Microstructure parameters of the Fe coordination site for the complexes at T = 300 K (LS) obtained by EXAFS fitting with multiple scattering approximation (Ri—interatomic distance, 2σi2—Debye–Waller factor, Fi—the statistical error of the fitting).
Table 1. Microstructure parameters of the Fe coordination site for the complexes at T = 300 K (LS) obtained by EXAFS fitting with multiple scattering approximation (Ri—interatomic distance, 2σi2—Debye–Waller factor, Fi—the statistical error of the fitting).
BondsRi, ÅAnglesΩ, Deg.
1·0.5H2O2·H2O4∙H2O5·1.5H2O1·0.5H2O2·H2O4∙H2O5·1.5H2O
Fe(1)-N(1)1.961.971.981.94N(1)Fe(1)N(8)107.9102.6105.9102.6
Fe(1)-N(3)1.961.911.931.95N(1)Fe(1)N(6)95.392.595.1100.1
Fe(1)-N(4)1.951.971.981.93N(1)Fe(1)N(9)97.192.196.6102.2
Fe(1)-N(6)1.951.961.981.96N(1)Fe(1)N(3)75.279.875.685.3
Fe(1)-N(8)1.861.881.901.86N(3)Fe(1)N(9)97.8102.1100.3102.9
Fe(1)-N(9)1.961.951.961.96N(3)Fe(1)N(4)75.881.280.673.7
2σ2 (Fe-N), Å20.0140.0130.0130.012Fi (*)2.72.51.81.6
* The determination accuracy for parameters (interatomic distances and angles) based on the EXAFS data is ±1% (for the nearest sphere of the environment). F i = i N w i 2 ( χ i e x p ( k ) χ i t h ( k ) ) 2 , w i = k i n i N k i n χ j e x p ( k ) .
Table 2. Microstructure parameters for complex [FeL2](ReO4)2 in the LS state obtained by single-scattering EXAFS data fitting.
Table 2. Microstructure parameters for complex [FeL2](ReO4)2 in the LS state obtained by single-scattering EXAFS data fitting.
CompoundCentral Ion–Scattering AtomNiRi, Å2σi2, Å2Fi
[FeL2](ReO4)2Fe–N61.960.0102.7
Fe–C82.780.014
Fe–C43.20
Table 3. The main vibrational frequencies (cm−1) in the spectra of L and complexes.
Table 3. The main vibrational frequencies (cm−1) in the spectra of L and complexes.
L1·0.5H2O2·H2O34·H2O5·1.5H2OAssignment
34853433 32253269ν(OH)
3100w3020w3147
3116
3050w3151
3130
3148
3129
ν(NH)
3079
3035
3060
3021
306730672954
2924
2854
2988
2040
2849
ν(CH)
1598
1573
1552
1595
1570
1550
1569
1551
1479
1593
1566
1552
1622
1558
1490
1469
1622
1558
1490
1471
Rring
1118
1085
1070
ν(SO4)
903 ν(ReO4)
24732493ν(B-H)
1086
1036
1105
1054
δ(BBH)
496491485485484Fe(3d6)–ligand (π) charge-transfer transition
385
356
330
376
334
381
321
334333ν(M-N)
Table 4. Parameters of diffuse reflectance spectra.
Table 4. Parameters of diffuse reflectance spectra.
Complexλ(1A11T2)λ(1A11B2)λ(1A11T1)λ(1A11A2)λ(1A11E)Calculated Parameter
BCLS
1·0.5H2O 454 613536
2·H2O 454 624539
3 432 603503
4·H2O475 518620 109.3482.01.97 × 104
5·1.5H2O465 518620 137.5606.51.98 × 104
Table 5. Mössbauer spectra parameters of the complexes.
Table 5. Mössbauer spectra parameters of the complexes.
Complexδ, mm/sε, mm/sΓ, mm/s
1·0.5H2O0.2690.3640.33
2·H2O0.2660.3420.25
30.2880.4640.28
4∙H2O0.278 (64%)
0.925 (36%)
0.420
2.224
0.26
0.80
5·1.5H2O0.2820.4650.25
Table 6. The temperatures of direct (Tc↑) and inverse (Tc↓) SCO for the studied complexes.
Table 6. The temperatures of direct (Tc↑) and inverse (Tc↓) SCO for the studied complexes.
ComplexTc↑, KTc↓, K
[FeL2]SO4·0.5H2O>420409
[FeL2]Br2·H2O>420>420
[FeL2](ReO4)2340340
[FeL2]B10H10·H2O436436
[FeL2]B12H12·1.5H2O455455
[FeL2]B10H10447440
[FeL2]B12H12458458
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lavrenova, L.G.; Shakirova, O.G.; Korotaev, E.V.; Trubina, S.V.; Tikhonov, A.Y.; Os’kina, I.A.; Petrov, S.A.; Zhizhin, K.Y.; Kuznetsov, N.T. High-Temperature Spin Crossover in Iron(II) Complexes with 2,6-Bis(1H-imidazol-2-yl)pyridine. Molecules 2022, 27, 5093. https://doi.org/10.3390/molecules27165093

AMA Style

Lavrenova LG, Shakirova OG, Korotaev EV, Trubina SV, Tikhonov AY, Os’kina IA, Petrov SA, Zhizhin KY, Kuznetsov NT. High-Temperature Spin Crossover in Iron(II) Complexes with 2,6-Bis(1H-imidazol-2-yl)pyridine. Molecules. 2022; 27(16):5093. https://doi.org/10.3390/molecules27165093

Chicago/Turabian Style

Lavrenova, Ludmila G., Olga G. Shakirova, Evgeniy V. Korotaev, Svetlana V. Trubina, Alexsei Ya. Tikhonov, Irina A. Os’kina, Sergey A. Petrov, Konstantin Yu. Zhizhin, and Nikolay T. Kuznetsov. 2022. "High-Temperature Spin Crossover in Iron(II) Complexes with 2,6-Bis(1H-imidazol-2-yl)pyridine" Molecules 27, no. 16: 5093. https://doi.org/10.3390/molecules27165093

Article Metrics

Back to TopTop