Next Article in Journal
Reciprocating Thermal Behavior in Multichannel Relaxation of Cobalt(II) Based Single Ion Magnets
Next Article in Special Issue
Crystal-to-Crystal Transformation from K2[Co(C2O4)2(H2O)2]·4H2O to K2[Co(μ-C2O4)(C2O4)]
Previous Article in Journal / Special Issue
Neutron Studies of a High Spin Fe19 Molecular Nanodisc
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Solvent-Induced Hysteresis Loop in Anionic Spin Crossover (SCO) Isomorph Complexes

1
Univ Brest, CNRS, CEMCA, 6 Avenue Le Gorgeu, C.S. 93837, CEDEX 3, 29238 Brest, France
2
Instituto de Ciencia Molecular (ICMol), Departamento de Química Inorgánica, Universidad de Valencia, C/Catedrático José Beltrán 2, 46980 Paterna, Spain
*
Authors to whom correspondence should be addressed.
Magnetochemistry 2021, 7(6), 75; https://doi.org/10.3390/magnetochemistry7060075
Submission received: 17 April 2021 / Revised: 19 May 2021 / Accepted: 20 May 2021 / Published: 23 May 2021

Abstract

:
Reaction of Fe(II) with the tris-(pyridin-2-yl)ethoxymethane (py3C-OEt) tripodal ligand in the presence of the pseudohalide ancillary NCSe (E = S, Se, BH3) ligand leads to the mononuclear complex [Fe(py3C-OEt)2][Fe(py3C-OEt)(NCSe)3]2·2CH3CN (3), which has been characterised as an isomorph of the two previously reported complexes, Fe(py3C-OEt)2][Fe(py3C-OEt)(NCE)3]2·2CH3CN, with E = S (1), BH3 (2). X-ray powder diffraction of the three complexes (13), associated with the previously reported single crystal structures of 12, revealed a monomeric isomorph structure for 3, formed by the spin crossover (SCO) anionic [Fe(py3C-OEt)(NCSe)3] complex, associated with the low spin (LS) [Fe(py3C-OEt)2]2+ cationic complex and two solvent acetonitrile molecules. In the [Fe(py3C-OEt)2]2+ complex, the metal ion environment involves two py3C-OEt tridentate ligands, while the [Fe(py3C-OEt)(NCSe)3] anion displays a hexacoordinated environment involving three N-donor atoms of one py3C-OEt ligand and three nitrogen atoms arising from the three (NCSe) coligands. The magnetic studies for 3 performed in the temperature range 300-5-400 K, indicated the presence of a two-step SCO transition centred around 170 and 298 K, while when the sample was heated at 400 K until its complete desolvation, the magnetic behaviour of the high temperature transition (T1/2 = 298 K) shifted to a lower temperature until the two-step behaviour merged with a gradual one-step transition at ca. 216 K.

1. Introduction

The spin crossover (SCO) materials are by far the most investigated molecular systems among switchable systems during the last decade due to their many possible applications for the development of new generations of electronic devices, such as displays [1,2,3,4], memory devices [4,5,6,7,8], and sensors [9,10,11,12,13,14]. Although the SCO behaviour can be essentially observed in octahedral complexes based on metal ions allowing spin state changes between the low spin (LS) and high spin (HS) states under external stimulus, such as temperature, pressure, light irradiation, or magnetic field, those based on Fe(II) ion exhibiting d6 electronic configuration remain the most studied systems [15,16,17,18,19,20,21,22,23,24,25,26,27,28,29]. Nevertheless, such complexes are mostly either cationic or neutral, and the Fe(II) anionic complexes exhibiting SCO behaviour have been relatively scarcely reported [21,22,23,24,25,26,27,28,29]. Furthermore, the few anionic SCO examples are restricted to only three different systems. The first one is the series [FeIIH3L][FeIIL]X, (X = AsF6, BF4, ClO4, PF6 and SbF6), based on the ligand tris-(2-(((2-methylimidazol-4-yl)methylidene)amino)ethyl) amine (H3L) and on its deprotonated anionic form (L3) [21]. The second one consists of the trinuclear [FeII3(μ-L)6(H2O)6]6 complex involving the 4-(1,2,4-triazol-4-yl)ethanedisulfonate anion (L2) [22], which displays a HS-HS-HS to HS-LS-HS transition around room temperature and a large hysteresis loop (>85 K). The last system concerns the series of mononuclear complexes involving the tris(2-pyridyl)methane (py3C-R, R = CnH2n+1, aryl group, O-CnH2n+1, O-aryl, O-CO-CnH2n+1), tridentate functionalized ligands (Scheme 1a) [23,24,25,26,27,28,29]. Such complexes, of general formula {A[Fe((py3C-R)(NCE)3)]m} (A = [(CnH2n+1)4N]+, [Fe(py3C-R)2]2+, E = S, Se, BH3), are based on the mononuclear [Fe((py3C-R)(NCE)3)] anion composed by an Fe(II) metal centre, one py3C-R tridentate ligand, and three terminal κN-SCE linear coligands (Scheme 1b). The different studies, reported essentially by Ishida et al. and some of us [23,24,25,26,27,28,29], have concerned the study of different chemical effects, such as those of the cationic counter ion or of the functional group (R) covalently linked to the tripodal py3C motif, on the transition temperatures and the cooperativity. More recently, some of us extended such effects to that of crystal packing by designing a series of polymorph complexes [29].
In order to determine the effect of the ancillary anionic coligands (NCE with E = S, BH3, Se) on the transition temperatures and the cooperativity, we have reported recently two isomorphic complexes of general formula, [Fe(py3C-OEt)2][Fe(py3C-OEt)(NCE)3]2·2CH3CN (NCE = NCS, NCBH3), based on the SCO [Fe(py3C-OEt)(NCE)3] anion and on the cationic LS complex, [Fe(py3C-OEt)2]2+, as counter ion [28]. At the same time, the NCSe analogue complex (E = Se), which completes such isomorphic series, has been also prepared. However, this complex, of a presumably chemical formula of [Fe(py3C-OEt)2][Fe(py3C-OEt)(NCSe)3]2·2CH3CN (3), could not be obtained as single crystals due to their instability. In addition, while magnetic behaviours of complexes 1 and 2 remained unchanged in the heating and cooling scan modes, complex 3 showed significant changes during the cooling/warming scan modes. These unexpected observations pushed us to explore in detail the peculiar switching behaviour of this compound. Here, we report the syntheses, structural characterization, infrared spectroscopy, and magnetic properties of the new isomorph [Fe(py3C-OEt)2][Fe(py3C-OEt)(NCSe)3]2·2CH3CN (3) exhibiting solvent-induced hysteresis loop of 50 K.

2. Results and Discussion

2.1. Syntheses

The py3C-OEt (tris-(pyridin-2-yl)ethoxymethane) tripodal ligand was prepared as previously described [28,29,30,31]. Compound 3 was obtained as a red polycrystalline powder and as single crystals by mixing a solution of [N(C2H5)4](NCSe) with a solution of FeCl2 and tris(pyridin-2-yl)ethoxymethane at −32 °C (see details in Section 3).

2.2. Structural Characterization and Magnetic Properties

In contrast to complexes 1 and 2, for which the crystal structures were determined using single crystal X-ray diffraction, complex 3, which was expected to be isomorph to the structure observed for 1 and 2, showed poor quality single crystal diffraction patterns that clearly precluded any correct single crystal structural characterization. However, after several attempts at 100 K, we succeeded in collecting some intensities, which led to unit cell parameters depicted in Table S1. Comparison of these parameters to those of the two isomorph complexes 1 and 2, indicated that the structure of complex 3 was isomorphic to complexes 1 and 2 (Table S1). This conclusion was supported by the experimental X-ray powder diffraction pattern observed for the polycrystalline powder of complex 3, which was very similar to the one observed for complex 1, as well as to the simulated pattern derived from the single crystal structure of complex 1 (Figure 1 and Figure S1).
It is also worth mentioning that the elemental analyses of complex 3 agreed with the chemical formula, [Fe(py3C-OEt)2][Fe(py3C-OEt)(NCSe)3]2·2CH3CN, expected for a complex isomorph to 1 and 2. Therefore, all the data strongly support that the crystal structure of complex 3 is isomorphic to those of complexes 1 and 2 [28]; and therefore its crystal structure consists of a low spin (LS) [Fe(py3C-OEt)2]2+ cationic complex (Figure 2b), an anionic [Fe(py3C-OEt)(NCE)3] (E = S (1), BH3 (2) Se (3)) complex (Figure 2a) and two CH3CN solvent molecules.
As previously described, complex 1 exhibited an incomplete SCO transition (T1/2 = 205 K), while 2 displayed a complete two-step transition at 245 K and 380 K (Figure 3). Their magnetic properties were studied in warming and cooling modes (300-2-400-2 K for 1; 300-2-500-2 K for 2), but no significant hysteretic effects or any change due to possible desolvation were detected. Also, irradiation at 10 K with a green light for several hours revealed no noticeable increase of the thermal variation of the product of the molar magnetic susceptibility times the temperature (χmT) for both compounds. As for complex 1, susceptibility measurements for complex 3 were performed in the 300-5 K and 5-400 K temperature ranges. The thermal variation of the χmT product for the three complexes (13) are shown in Figure 3. For compound 3, the χmT value per formula at 400 K (≈6.71 cm3 K mol−1) was in agreement with the value expected for two isolated Fe(II) ions (S = 2 and g ≈ 2.1), revealing the presence of two magnetically isolated HS Fe(II) ions [15,16,17,18,19,20,21,22,23,24,25,26,27,28,29]. On cooling, the χmT value of 3 showed an initial abrupt drop, at around 298 K, reaching a value of ca. 3.40 cm3 K mol−1 at 265 K. On further cooling, we observed a second drop, at around 170 K, to reach a plateau at ca. 0.56 cm3 K mol−1 below 116 K. This low temperature χmT value implies the presence of a residual HS fraction of ca. 8 %. This behaviour indicates the presence of an incomplete HS to LS two-step transition centred at around 170 and 298 K.
In contrast to isomorphs 1 and 2, where the magnetic properties did not show any change after successive cooling and heating scans in the ranges 2–400 K for 1 and 2–500 K for 2, the two-step behaviour described above for complex 3 (Figure 3) was irreversible due to a gradual desolvation of the sample at 400 K, as previously observed in several solvated systems [32,33,34,35,36,37,38,39,40]. As a matter of fact, when the sample was maintained at 400 K until its complete desolvation, the second cycle (400-50-400 K) in 3 produced a shift to lower temperatures for the high temperature step, while the transition temperature of the low temperature step remained unchanged (See Figure 4). Similar trends were observed for the third and fourth cycles, until the fifth cycle where the initial high temperature step merged with the low temperature step (Figure 4), to lead to the gradual one-step transition depicted in Figure 5a. There were no hysteretic effects and the HS fraction of ca. 8% remained unchanged after the different heating and cooling cycles.
Irradiation of the sample with a green laser (λ = 532 nm) at 10 K produced an increase of the χmT product, indicative of the presence of a light-induced excited spin state trapping at low temperatures (LIESST effect). After ca. 4 h, the χmT product reached saturation at a value of ca. 1.6 cm3 K mol−1 (Figure S3). This value indicates that around 1/3 of one of the two LS Fe(II) centres in the [Fe(py3C-OEt)(NCS)3] anions were excited to the HS state. After switching off the light irradiation, heating the sample further increased the χmT value up to ca. 2.0 cm3 K mol−1 (representing ca. 44 % of one of the two Fe(II) centres). On further heating, the sample relaxed to the LS state at a TLIESST of ca. 58 K (Figure 5b).
One of the major points, deserving special attention in regards to the three isomorphic complexes, concerns the origin of the unexpected and singular magnetic behaviour that only occurred for isomorph 3 (see Figure 3 and Figure 5a). To try to understand the process that occurred at high temperatures, we performed thermogravimetric analysis (TGA) and X-ray powder diffraction on the three isomorphs to know more about the desolvation and solvation processes of this system. Thus, TGA measurements were performed for the three isomorphs, which were heated at 5 °C min−1, under nitrogen atmosphere, from room temperature to 390 K. In Figure 6, the mass evolution with temperature for the three complexes were gathered, showing clearly that the two isomorphs 1 and 2 remained stable and retained their solvent molecules up to 390 K, while complex 3 started to lose weight from room temperature and lost 4.37 % of its mass when heated up to 370 K, corresponding to two CH3CN solvent molecules per formula unit. These measurements revealed that despite their isomorphic structures, complexes 1 and 2 retained their crystallization solvent molecules while complex 3 lost them even at moderate temperatures, suggesting that the crystal packing in 3 has larger cavities that allow an easy desolvation (see below).
To check for the reversibility of this desolvation process, we performed successive desolvation and resolvation cycles, by heating the solvated sample and by adding two drops of CH3CN on the desolvated sample, respectively. After resolvation, magnetic measurements (Figure S4) and X-ray powder diffraction (Figure 7) showed that the sample recovered its original behaviour, supporting the reversibility of the desolvation/resolvation process.

2.3. Variable Temperature Magnetic Properties and Infrared Spectroscopy

In order to confirm the Fe(II) spin state at high and low temperatures and the presence of the incomplete HS to LS transition for 3 (see Figure 3), we measured the infrared spectrum at 100 and 300 K in the range of the fundamental stretching vibration of the NCSe units (1975–2130 cm−1), since it has been clearly established that the intensities of these stretching vibrations are very sensitive to the spin state of the Fe(II) metal ion [28,29,41,42,43,44,45,46,47]. We thus recorded the infrared spectra for 3 at 350 and 100 K, according to the thermal evolution of the χmT product depicted in Figure 5a. The infrared spectra for 3 in the C≡N frequency region (1975–2130 cm−1) at 350 and 100 K are displayed in Figure 8. At 350 K, two νC≡N stretching broad bands, characteristic of the HS state, appeared at 2050 and 2075 cm−1, while at 100 K, four strong bands, characteristic of the LS state, appeared at 2044, 2075, 2082, and 2109 cm−1. In agreement with the presence of an 8 % HS fraction at low temperatures (see magnetic section), the three bands observed at 2044, 2075, and 2082 cm−1 can be viewed as resulting from the decrease in intensity of the two broad and strong bands observed for the HS state, while the band observed at a higher frequency (2109 cm−1) appeared as the specific band of the LS state.

2.4. Magneto-Spectroscopic Relationships

Based on the magnetic behaviours (Figure 5a) and on the main bands that were temperature sensitive (Figure 8) of both solvated and desolvated phases of complex 3, we investigated the thermal evolution of the infrared spectra of the stretching vibration of the NCSe in the range 1975–2130 cm−1. For both phases (solvated and desolvated), we recorded the infrared spectra in the vicinity of the SCO transitions from 100 to 350 K. First, we recorded the infrared spectra for the freshly prepared complex 3, heating the sample from 100 to 350 K to avoid the partial desolvation of the sample (Figure 9a). Then, the same sample was heated during one hour at 400 K to ensure its complete desolvation, and the corresponding infrared spectra were then recorded cooling the sample from 350 to 100 K (Figure 9b).
The intensities of the two ν(NCSe) broad bands (2050 and 2075 cm−1) attributed to the HS state decreased gradually with decreasing temperature from 350 to 100 K, but persisted even at 100 K, supporting the presence of a fraction of HS Fe(II) centres, as revealed by the magnetic data. In parallel, a new band, characteristic of the LS state, appeared at higher frequencies (2109 cm−1), whose intensity gradually increased with decreasing the temperature. However, as can be easily observed in both Figure 9a,b, the infrared spectra did not show any clear difference between the infrared band evolutions of the solvated (Figure 9a) and desolvated (Figure 9b) phases of complex 3, as revealed by the magnetic data. Thus, in order to show more clearly this difference and to appreciate, at least qualitatively, the consistency of the experimental infrared data, we correlated the thermal variation of the χmT product derived from magnetic study and the thermal evolution of the intensity of the infrared bands. The results, depicted in Figure 10, showed that the intensities of the characteristic infrared bands (2109, 2075, 2050 cm−1) fit perfectly with the thermal evolution of the χmT product, in agreement with the presence of a two-step SCO transition in the solvated sample (Figure 10a) and a gradual one-step SCO behaviour in the desolvated sample (Figure 10b).
The final question to be answered by this study was why the isomorph based on the NCSe ligand (complex 3) displayed a reversible solvation/desolvation process with CH3CN while the two other isomorphs (complexes 1 and 2) retained the solvent molecules at temperatures exceeding 390 K. The answer could only come from the crystal packing of this triad of isomorphs. Indeed, examination of the crystal packing revealed that the CH3CN solvent molecules are located in tetragonal-like channels running along the [010] direction, which are generated by eclipsed stacks of the -Fe-NCE…Fe…ECN-Fe-NCE…Fe- metallacycles (see Figure 11). Even though the three isomorphs display, as expected, similar crystal packing, the fact that they differ by the nature of the NCE ancillary ligands (E = S (1), BH3 (2), Se (3)), induces strong differences in the sizes of the metallacycles, due to the different lengths of the three linear anions (the N-E distance increases from 2.72 Å for E = BH3, to 2.80 Å for E = S and 2.94 Å for E = Se). The largest tetragonal-like channels are expected to be those of the isomorph with the longest ancillary linear ligand (i.e., compound 3). Unfortunately, this difference could not be quantified due to the lack of single crystal structural data of complex 3. However, in order to provide a reasonable estimation of the effect of the NCE ligands on the size of the metallacycles, we have calculated the relative increase of their dimensions when passing from complex 2 to complex 1 (see Figure 11 and Table 1 with the different Fe···Fe distances (d1–d4) determining the size of the channels).
As can be seen in Table 1, when passing from complex 2 (with NCBH3 and N···B distance of 2.72 Å) to complex 1 (with NCS and N···S distance of 2.80 Å), there is an increase of +2.9 % in the size of the anion, which led to increases in the Fe···Fe distances in the metallacycle of up to 28.6 % (see d1 to d4 in Table 1). Therefore, if we consider complex 3 based on the NCSe linear ligand, which corresponds to the biggest linear anion (with N···Se distance of 2.94 Å), the expected metallacycle should be significantly larger than those observed for the isomorphs 1 and 2. These observations explain clearly why complex 3 involves larger tetragonal-like channels, allowing the easy and reversible solvation and desolvation processes.

3. Experimental Section

3.1. Starting Materials

All starting reagents and solvents were purchased and used as received. The tris-(pyridin-2-yl)ethoxymethane (py3C-OEt) ligand was prepared under nitrogen atmosphere as described previously [28,29,30,31].

3.2. Synthesis of [Fe(py3C-OEt)2][Fe(py3C-OEt)(NCSe)3]2·2CH3CN (3)

Tris-(pyridin-2-yl)ethoxymethane (50.0 mg, 0.17 mmol) and FeCl2 (20.0 mg, 0.16 mmol) were dissolved in methanol (5 mL) in the presence of a few mg of ascorbic acid. The mixture was stirred at room temperature for 15 min. To the resulting solution was added a solution of acetonitrile (5 mL) containing the [(C2H5)4N]NCSe salt (162 mg, 0.69 mmol). After 30 min stirring, the resulting solution was filtered and quickly cooled at −32 °C. After three days, the bright red polycristalline powder of (3) as well as a few red single crystals were obtained. Anal. Calcd. (%) for [Fe(py3C-OEt)2][Fe(py3C-OEt)(NCSe)3]2·2CH3CN (C82H74Fe3N20O4Se6) 3: C, 48.2; H, 3.7; N, 13.7; found (%): C, 47.9; H, 3.8; N, 14.0. IR data (ν/cm−1) for the freshly filtered sample (powder and single crystals, Figure S5): 410 (w), 423 (w), 477 (w), 500 (w), 513 (w), 530 (w), 659 (m), 726 (w), 758 (w), 886 (w), 1011 (m), 1086 (w), 1108 (m), 1143 (m), 1205 (w), 1252 (w), 1291 (w), 1389 (w), 1434 (m), 1462 (s), 1593 (m), 2060 (s), 2244 (w), 2871 (w), 2901 (w), 2972.11 (w), 3076 (w), 3442 (br).

3.3. Characterization of the Materials

Infrared spectra of complex 3 were performed using a platinum ATR Vertex 70 BRUKER spectrometer with variable temperature cell holder (VT Cell Holder typer P/N GS21525). 1H and 13C NMR spectra were performed using BRUKER DRX 300 MHz, Advance 400 MHz and Advance III HD 500 MHz equipment. TGA measurements were performed on ATG-LabsysTM, Setaram (see details in Supplementary Information).

3.4. Magnetic Measurements

Magnetic susceptibility measurements were performed using a Quantum Design MPMS-XL-5 SQUID susceptometer (San Diego, CA, USA). The susceptibility data were corrected for the diamagnetic contributions using Pascal’s constant tables [48]. The photomagnetic studies were performed by irradiating the sample at 10 K with a green Diode Pumped Solid State Laser DPSS-532-20 from Chylas (see details in Supplementary Information).

4. Conclusions

We have shown that the compound [Fe(py3C-OEt)2][Fe(py3C-OEt)(NCSe)3]2·2CH3CN (3), based on the [Fe(py3C-OEt)2]2+ LS cation and the SCO [Fe(py3C-OEt)(NCSe)3] anion, displays a reversible desolvation process that affects the SCO behaviour. This compound has been prepared as a polycrystalline powder and as single crystals using a similar protocol to that used previously for the syntheses of the two isomorphic complexes [Fe(py3C-OEt)2][Fe(py3C-OEt)(NCE)3]2·2CH3CN (E = S (1), BH3 (2)) based on the two ancillary linear ligands NCS and NCBH3 [28]. Despite being obtained in the form of prismatic-shaped single crystals, complex 3 could not be characterized by X-ray single crystal diffraction, because of the low stability of its single crystals, in contrast to complexes 1 and 2, which have been structurally characterized. However, combined X-ray powder diffraction, infrared spectra, and CHN elemental analyses clearly revealed that complex 3 exhibits an isomorphic structure to those of complexes 1 and 2. TGA analyses performed on the single crystal samples of the three isomorphs revealed clearly that the two isomorphs 1 and 2, for which the corresponding single crystals are stable, retained their solvent molecules up to 390 K, while complex 3 began to lose its CH3CN solvent molecules from room temperature. The magnetic studies for 3 performed in cooling and heating scans in the temperature ranges 300–5 K and 5–400 K, respectively, indicated the presence of an incomplete HS to LS two-step like transition centred around 170 and 298 K, while when the sample was heated at 400 K until its complete desolvation, the magnetic behaviour of the high temperature transition (T1/2 = 298 K) shifted to a lower temperature until the two-step behaviour merged with a gradual one-step transition at ca. 216 K. Such behaviour, which can be viewed as a solvent-induced hysteresis loop of 50 K [41,42,43,44,45,46,47], was confirmed by infrared spectra recorded in the vicinity of the SCO transition for both solvated and desolvated samples. Furthermore, successive desolvation and solvation cycles, tracked by SQUID measurements and X-ray powder diffraction, showed that the desolvation process is fully reversible. As shown by the TGA analysis, compound 3 exhibited a different magnetic behaviour due to its easy desolvation and solvation process, in contrast to the two other isomorphs, which retained their solvent molecule up to 390 K, despite their isomorphic structures. This unexpected behaviour was elucidated by careful examination of the crystal packing of these isomorph complexes, which clearly revealed that the solvent molecules are located in tetragonal-like channels generated by the eclipsed stacks of the “-Fe-NCE…Fe…ECN-Fe-NCE…Fe-” metallacycles. Therefore, the isomorph based on the bigger NCSe ancillary ligand, should display the larger tetragonal-like channels, allowing easier solvation and desolvation as observed in isomorph 3.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/magnetochemistry7060075/s1, Table S1: Crystal data of [Fe(py3C-OEt)2][Fe(py3C-OEt)(NCE)3]2·2CH3CN (E = S (1), BH3 (2), Se (3)), Figure S1: Experimental and simulated XRPD patterns for compounds 1 and 2, and experimental one for complex 3, Figure S2: Thermal variation of χm of a freshly prepared sample of compound 3 for different consecutive heating and cooling scans, Figure S3: Time dependence of χmT at 10 K for compound 3 under laser irradiation with green light (λ = 532 nm, switched on at time ca. 10 min), Figure S4: Different heating and cooling scans of the χmT product performed on the de-solvated compound 3 after re-solvation with two drops of CH3CN, Figure S5: IR spectra of single crystals and polycrystalline powder of complex 3.

Author Contributions

E.C. performed the chemical syntheses (ligand and the complex) under supervision of S.T. and F.C., and recorded and analysed the infrared spectra under supervision of S.T. and N.C.; F.C. interpreted the NMR spectra; S.B. performed the X-ray powder diffraction; C.J.G.-G. performed and interpreted the magnetic measurements; S.T. supervised the experimental work and wrote the manuscript on which all the authors have contributed. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by “Mission pour les Initiatives Transverses et Interdisciplinaires (MITI, CNRS)”, Mol-CoSM ANR Project N° ANR-20-CE07-0028-01 the “Université de Brest” (IBSAM institute), the Région Bretagne (EC), the Generalidad Valenciana (Prometeo2019/076 project) and the Spanish MINECO (Project CTQ2017-87201-P AEI/FEDER, EU).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article or Supplementary Material.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cobo, S.; Molnár, G.; Real, J.-A.; Bousseksou, A. Multilayer Sequential Assembly of Thin Films That Display Room-Temperature Spin Crossover with Hysteresis. Angew. Chem. Int. Ed. 2006, 45, 5786–5789. [Google Scholar] [CrossRef]
  2. Kahn, O.; Martinez, C.J. Spin-transition polymers: From molecular materials toward memory devices. Science 1998, 279, 44–48. [Google Scholar] [CrossRef]
  3. Kumar, K.S.; Ruben, M. Sublimable Spin-Crossover Complexes: From Spin-State Switching to Molecular Devices. Angew. Chem. Int. Ed. 2021, 60, 7502–7521. [Google Scholar] [CrossRef] [Green Version]
  4. Murray, K.S.; Kepert, C.J. Cooperativity in Spin Crossover Systems: Memory, Magnetism and Microporosity. In Spin Crossover in Transition Metal Compounds I; Gütlich, P., Goodwin, H.A., Eds.; Topics in Current Chemistry; Springer: Berlin/Heidelberg, Germany, 2004; Volume 233, pp. 195–228. [Google Scholar]
  5. Létard, J.-F.; Guionneau, P.; Goux-Capes, L. Towards Spin Crossover Applications. In Spin Crossover in Transition Metal Compounds III; Gütlich, P., Goodwin, H.A., Eds.; Topics in Current Chemistry; Springer: Berlin/Heidelberg, Germany, 2004; Volume 235, pp. 221–249. [Google Scholar]
  6. Bousseksou, A.; Molnar, G.; Salmon, L.; Nicolazzi, W. Molecular spin crossover phenomenon: Recent achievements and prospects. Chem. Soc. Rev. 2011, 40, 3313–3335. [Google Scholar] [CrossRef] [PubMed]
  7. Rueckes, T.; Kim, K.; Joselevich, E.; Tseng, G.Y.; Cheung, C.-L.; Lieber, C.M. Carbon Nanotube-Based Nonvolatile Random Access Memory for Molecular Computing. Science 2000, 289, 94–97. [Google Scholar] [CrossRef] [Green Version]
  8. Galet, A.; Gaspar, A.B.; Carmen Muñoz, M.; Bukin, G.V.; Levchenko, G.; Real, J.-A. Tunable Bistability in a Three-Dimensional Spin-Crossover Sensory- and Memory-Functional Material. Adv. Mater. 2005, 17, 2949–2953. [Google Scholar] [CrossRef]
  9. Bartual-Murgui, C.; Akou, A.; Thibault, C.; Molnar, G.; Vieu, C.; Salmon, L.; Bousseksou, A. Spin-crossover metal–organic frameworks: Promising materials for designing gas sensors. J. Mater. Chem. C 2015, 3, 1277–1285. [Google Scholar] [CrossRef]
  10. Lapresta-Fernandez, A.; Titos-Padilla, S.; Herrera, J.M.; Salinas-Castillo, A.; Colacio, E.; Capitan-Vallvey, L.F. Photographing the synergy between magnetic and colour properties in spin crossover material [Fe(NH2trz)3](BF4)2: A temperature sensor perspective. Chem. Commun. 2013, 49, 288–290. [Google Scholar] [CrossRef] [PubMed]
  11. Linares, J.; Codjovi, E.; Garcia, Y. Pressure and temperature spin crossover sensors with optical detection. Sensors 2012, 12, 4479–4492. [Google Scholar] [CrossRef] [PubMed]
  12. Cuéllar, M.P.; Lapresta-Fernández, A.; Herrera, J.M.; Salinas-Castillo, A.; del Carmen Pegalajar, M.; Titos-Padilla, S.; Colacio, E.; Capitán-Vallvey, L.F. Thermochromic sensor design based on Fe(II) spin crossover/polymers hybrid materials and artificial neural networks as a tool in modelling. Sens. Actuators B 2015, 208, 180–187. [Google Scholar] [CrossRef]
  13. Rodriguez-Jimenez, S.; Feltham, H.L.C.; Brooker, S. Non-Porous Iron(II)-Based Sensor: Crystallographic Insights into a Cycle of Colorful Guest-Induced Topotactic Transformations. Angew. Chem. Int. Ed. 2016, 55, 15067–15071. [Google Scholar] [CrossRef]
  14. Benaicha, B.; Van Do, K.; Yangui, A.; Pittala, N.; Lusson, A.; Sy, M.; Bouchez, G.; Fourati, H.; Gómez-García, 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]
  15. Halcrow, M.A. (Ed.) Spin-Crossover Materials, Properties and Applications; John Wiley & Sons Ltd: Oxford, UK, 2013. [Google Scholar]
  16. Phan, H.; Hrudka, J.J.; Igimbayeva, D.; Lawson Daku, L.M.; Shatruk, M. A Simple Approach for Predicting the Spin State of Homoleptic Fe(II) Tris-diimine Complexes. J. Am. Chem. Soc. 2017, 139, 6437–6447. [Google Scholar] [CrossRef] [PubMed]
  17. Pittala, N.; Thétiot, F.; Charles, C.; Triki, S.; Boukheddaden, K.; Chastanet, G.; Marchivie, M. An unprecedented trinuclear FeII triazole-based complex exhibiting a concerted and complete sharp spin transition above room temperature. Chem. Commun. 2017, 53, 8356–8359. [Google Scholar] [CrossRef] [PubMed]
  18. 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]
  19. El Hajj, F.; Sebki, G.; Patinec, V.; Marchivie, M.; Triki, S.; Handel, H.; Yefsah, S. Macrocycle-based spin-crossover materials. Inorg. Chem. 2009, 48, 10416–10423. [Google Scholar] [CrossRef]
  20. Setifi, F.; Milin, E.; Charles, C.; Thétiot, F.; Triki, S.; Gómez-García, C.G. Spin Crossover Iron(II) Coordination Polymer Chains: Syntheses, Structures, and Magnetic Characterizations of [Fe(aqin)2(μ-M(CN)4)] (M = Ni(II), Pt(II), aqin = Quinolin-8-amine). Inorg. Chem. 2014, 53, 97–104. [Google Scholar] [CrossRef] [PubMed]
  21. Yamada, M.; Ooidemizu, M.; Ikuta, Y.; Osa, S.; Matsumoto, N.; Iijima, S.; Kojima, M.; Dahan, F.; Tuchagues, J.-P. Interlayer Interaction of Two-Dimensional Layered Spin Crossover Complexes [FeIIH3LMe][FeIILMe]X (X = ClO4, BF4, PF6, AsF6, and SbF6; H3LMe = Tris[2-(((2-methylimidazol-4-yl)methylidene)amino)ethyl]amine). Inorg. Chem. 2003, 42, 8406–8416. [Google Scholar] [CrossRef]
  22. Gómez, V.; Sáenz de Pipaón, C.; Maldonado-Illescas, P.; Waerenborgh, J.C.; Martin, E.; Benet-Buchholz, J.; Galán-Mascarós, J.-R. Easy Excited-State Trapping and Record High TTIESST in a Spin-Crossover Polyanionic Fe(II) Trimer. J. Am. Chem. Soc. 2015, 137, 11924–11927. [Google Scholar] [CrossRef] [PubMed]
  23. Hirosawa, N.; Oso, Y.; Ishida, T. Spin crossover and light-induced excited spin-state trapping observed for an iron (II) complex chelated with tripodal tetrakis(2-pyridyl) methane. Chem. Lett. 2012, 41, 716–718. [Google Scholar] [CrossRef]
  24. Yamasaki, M.; Ishida, T. Spin-crossover thermal hysteresis and light-induced effect on iron (II) complexes with tripodal tris (2-pyridyl) methanol. Polyhedron 2015, 85, 795–799. [Google Scholar] [CrossRef]
  25. 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]
  26. Ishida, T.; Kaneto, T.; Yamasaki, M. An iron(II) complex tripodally chelated with 1,1,1-tris(pyridine-2-yl)ethane showing room-temperature spin-crosssover behaviour. Acta Cryst. Sect. C 2016, 72, 797–801. [Google Scholar] [CrossRef] [PubMed]
  27. Kashiro, A.; Some, K.; Kobayashi, Y.; Ishida, T. Iron(II) and 1,1,1-Tris(2-pyridyl)nonadecane Complex Showing an Order–Disorder-Type Structural Transition and Spin-Crossover Synchronized over Both Conformers. Inorg. Chem. 2019, 58, 7672–7676. [Google Scholar] [CrossRef]
  28. Cuza, E.; Benmansour, S.; Cosquer, N.; Conan, F.; Pillet, S.; Gómez-García, C.J.; Triki, S. Spin Cross-Over (SCO) Anionic Fe(II) Complexes Based on the Tripodal Ligand Tris(2-pyridyl)ethoxymethane. Magnetochemistry 2020, 6, 26. [Google Scholar] [CrossRef]
  29. Cuza, E.; Mekuimemba, C.D.; Cosquer, N.; Conan, F.; Pillet, S.; Chastanet, C.; Triki, S. Spin Crossover and High-Spin State in Fe(II) Anionic Polymorphs Based on Tripodal Ligands. Inorg. Chem. 2021, 60, 6536–6549. [Google Scholar] [CrossRef]
  30. White, D.L.; Faller, J.W. Preparation and Reactions of the C3v Ligand Tris(2-pyridyl)methane and Its Derivatives. Inorg. Chem. 1982, 21, 3119–3122. [Google Scholar] [CrossRef]
  31. Jonas, R.T.; Stack, T.D.P. Synthesis and Characterization of a Family of Systematically Varied Tris(2-pyridyl)methoxymethane Ligands: Copper(I) and Copper(II) Complexes. Inorg. Chem. 1998, 37, 6615–6629. [Google Scholar] [CrossRef]
  32. Benmansour, S.; Gómez-Claramunt, P.; Gómez-García, C.G. Effects of water removal on the structure and spin-crossover in an anilato-based compound. J. Appl. Phys. 2021, 129, 123904. [Google Scholar] [CrossRef]
  33. Kulmaczewski, R.; Bamiduro, F.; Shahid, N.; Cespedes, O.; Halcrow, M.A. Structural Transformations and Spin-Crossover in [FeL2]2+ Salts (L=4-{tert-Butylsulfanyl}-2,6-di{pyrazol-1-yl}pyridine): The Influence of Bulky Ligand Substituents. Chem. Eur. J. 2021, 27, 2082–2092. [Google Scholar] [CrossRef]
  34. Mondal, D.J.; Roy, S.; Yadav, J.; Zeller, M.; Konar, S. Solvent-Induced Reversible Spin-Crossover in a 3D Hofmann-Type Coordination Polymer and Unusual Enhancement of the Lattice Cooperativity at the Desolvated State. Inorg. Chem. 2020, 59, 13024–13028. [Google Scholar] [CrossRef] [PubMed]
  35. Phonsri, W.; Davies, C.G.; Jameson, G.N.L.; Moubaraki, B.; Murray, K.S. Spin Crossover, Polymorphism and Porosity to Liquid Solvent in Heteroleptic Iron(III) {Quinolylsalicylaldimine/Thiosemicarbazone-Salicylaldimine} Complexes. Chem. Eur. J. 2016, 22, 1322–1333. [Google Scholar] [CrossRef]
  36. Lennartson, A.; Southon, P.; Sciortino, N.F.; Kepert, C.J.; Frandsen, C.; Mørup, S.; Piligkos, S.; McKenzie, C. Reversible Guest Binding in a Non-Porous FeII Coordination Polymer Host Toggles Spin Crossover. J. Chem. Eur. J. 2015, 21, 16066–16072. [Google Scholar] [CrossRef] [PubMed]
  37. Roberts, T.D.; Tuna, F.; Malkin, T.L.; Kilner, C.A.; Halcrow, M.A. An iron(ii) complex exhibiting five anhydrous phases, two of which interconvert by spin-crossover with wide hysteresis. Chem. Sci. 2012, 3, 349–354. [Google Scholar] [CrossRef]
  38. Clemente-León, M.; Coronado, E.; Giménez-López, M.C.; Romero, F.M. Structural, Thermal, and 52Magnetic Study of Solvation Processes in Spin-Crossover [Fe(bpp)2][Cr(L)(ox)2]2·nH2O Complexes. Inorg. Chem. 2007, 46, 11266–11276. [Google Scholar] [CrossRef]
  39. Hayami, S.; Gu, Z.-Z.; Yoshiki, H.; Fujishima, A.; Sato, O. Iron(III) Spin-Crossover Compounds with a Wide Apparent Thermal Hysteresis around Room Temperature. J. Am. Chem. Soc. 2001, 123, 11644–11650. [Google Scholar] [CrossRef]
  40. Garcia, Y.; Van Koningsbruggen, P.J.; Codjovi, E.; Lapouyade, R.; Kahn, O.; Rabardel, L. Non-classical FeII spin-crossover behaviour leading to an unprecedented extremely large apparent thermal hysteresis of 270 K: Application for displays. J. Mater. Chem. 1997, 7, 857–858. [Google Scholar] [CrossRef]
  41. Sorai, M.; Seki, S. Phonon coupled cooperative low-spin 1A1 ↔ high-spin 5T2 transition in [Fe(phen)2(NCS)2] and [Fe(phen)2(NCSe)2] crystals. J. Phys. Chem. Solids 1974, 35, 555–570. [Google Scholar] [CrossRef]
  42. Brehm, G.; Reiher, M.; Le Guennic, B.; Leibold, M.; Schindler, S.; Heinemann, F.W.; Schneider, S. Investigation of the low-spin to high-spin transition in a novel [Fe(pmea)(NCS)2] complex by IR and Raman spectroscopy and DFT calculations. J. Raman Spectrosc. 2006, 37, 108–122. [Google Scholar] [CrossRef]
  43. Park, Y.; Jung, Y.M.; Sarker, S.; Lee, J.-J.; Lee, Y.; Lee, K.; Oh, J.J.; Joo, S.-W. Temperature-dependent infrared spectrum of (Bu4N)2[Ru(dcbpyH)2(NCS)2] on nanocrystalline TiO2 surfaces. Solar Energy Mater. Solar Cells 2010, 94, 857–864. [Google Scholar] [CrossRef]
  44. Varma, V.; Fernandes, J.-R. An Infrared Spectroscopic Study of the Low-Spin-High-spin transition in in FexMn1−x(Phen)2(NCS)2: A Composition-Induced Change in the Order of the Spin-State Transition. Chem. Phys. Let. 1990, 167, 367–370. [Google Scholar] [CrossRef]
  45. Sankar, G.; Thomas, J.M.; Varma, V.; Kulkani, G.U.; Rao, C.N.R. An investigation of the first-order spin-state transition in the Fe(phen)2(NCS)2 EXAFS and infrared spectroscopy. Chem. Phys. Lett. 1996, 251, 79–83. [Google Scholar] [CrossRef]
  46. Smit, E.; de Waal, D.; Heyns, A.M. The spin-transition complexes [Fe(Htrz)3](ClO4)2 and [Fe(NH2trz)3](ClO4)2 I. FT-IR spectra of a low pressure and a low temperature phase transition. Mater. Res. Bull. 2000, 35, 1697–1707. [Google Scholar] [CrossRef]
  47. Durand, P.; Pillet, S.; Bendeif, E.-E.; Carteret, C.; Bouazaoui, M.; El Hamzaoui, H.; Capoen, B.; Salmon, L.; Hébert, S.; Ghanbaja, J.; et al. Room temperature bistability with wide thermal hysteresis in a spin crossover silica nanocomposite. J. Mater. Chem. C 2013, 1, 1933–1942. [Google Scholar] [CrossRef]
  48. Bain, G.A.; Berry, J.F. Diamagnetic corrections and Pascal’s constants. J. Chem. Educ. 2008, 85, 532–536. [Google Scholar] [CrossRef]
Scheme 1. (a) The tris-(2-pyridyl)methane (py3C-R) ligands and their tridentate coordination mode (b).
Scheme 1. (a) The tris-(2-pyridyl)methane (py3C-R) ligands and their tridentate coordination mode (b).
Magnetochemistry 07 00075 sch001
Figure 1. X-ray powder diffraction patterns for [Fe(py3C-OEt)2][Fe(py3C-OEt)(NCE)3]2·2CH3CN (E = S (1), Se (3)) and the simulated pattern derived from the crystal structure of complex 1.
Figure 1. X-ray powder diffraction patterns for [Fe(py3C-OEt)2][Fe(py3C-OEt)(NCE)3]2·2CH3CN (E = S (1), Se (3)) and the simulated pattern derived from the crystal structure of complex 1.
Magnetochemistry 07 00075 g001
Figure 2. View of the anionic [Fe(py3C-OEt)(NCS)3] (a) and of the cationic [Fe(py3C-OEt)2]2+ (b) complexes in [Fe(py3C-Oet)2][Fe(py3C-OEt)(NCS)3]2·2CH3CN (1) [28]. Codes of equivalent positions: (a) = -x, -y, -z.
Figure 2. View of the anionic [Fe(py3C-OEt)(NCS)3] (a) and of the cationic [Fe(py3C-OEt)2]2+ (b) complexes in [Fe(py3C-Oet)2][Fe(py3C-OEt)(NCS)3]2·2CH3CN (1) [28]. Codes of equivalent positions: (a) = -x, -y, -z.
Magnetochemistry 07 00075 g002
Figure 3. (a) Temperature dependence of the χmT product of 1 (), 2 (), and 3 (o).
Figure 3. (a) Temperature dependence of the χmT product of 1 (), 2 (), and 3 (o).
Magnetochemistry 07 00075 g003
Figure 4. Thermal variation of χmT product of a freshly prepared sample of compound 3 for different consecutive heating and cooling scans.
Figure 4. Thermal variation of χmT product of a freshly prepared sample of compound 3 for different consecutive heating and cooling scans.
Magnetochemistry 07 00075 g004
Figure 5. Magnetic properties of compound 3: (a) thermal evolution of the χmT product for the solvated (o) and desolvated (●) samples; (b) thermal evolution of the χmT product for the desolvated sample in the dark (●), under 532 nm () light irradiation at 10 K, and the subsequent thermal relaxation () in the dark. Inset shows the thermal variation of the derivative of χmT.
Figure 5. Magnetic properties of compound 3: (a) thermal evolution of the χmT product for the solvated (o) and desolvated (●) samples; (b) thermal evolution of the χmT product for the desolvated sample in the dark (●), under 532 nm () light irradiation at 10 K, and the subsequent thermal relaxation () in the dark. Inset shows the thermal variation of the derivative of χmT.
Magnetochemistry 07 00075 g005
Figure 6. TGA analyses of compounds 13.
Figure 6. TGA analyses of compounds 13.
Magnetochemistry 07 00075 g006
Figure 7. Experimental X-ray powder diffraction patterns of solvated and desolvated samples of 3, confirming the reversibility of the desolvation/resolvation process.
Figure 7. Experimental X-ray powder diffraction patterns of solvated and desolvated samples of 3, confirming the reversibility of the desolvation/resolvation process.
Magnetochemistry 07 00075 g007
Figure 8. The principal infrared bands in the 1975–2130 cm−1 region for complex 3 at 350 and 100 K.
Figure 8. The principal infrared bands in the 1975–2130 cm−1 region for complex 3 at 350 and 100 K.
Magnetochemistry 07 00075 g008
Figure 9. Temperature dependence of the infrared spectra, in the temperature range 350–100 K, for compound 3: solvated (a) and desolvated (b) samples.
Figure 9. Temperature dependence of the infrared spectra, in the temperature range 350–100 K, for compound 3: solvated (a) and desolvated (b) samples.
Magnetochemistry 07 00075 g009
Figure 10. Temperature dependences of the χmT product and of the relative absorbance of the ν(CN) bands observed at 2109 cm−1 (Δ), 2075 cm−1 (), and at 2050 cm−1 (-●-), for the solvated (a) and desolvated (b) samples of 3.
Figure 10. Temperature dependences of the χmT product and of the relative absorbance of the ν(CN) bands observed at 2109 cm−1 (Δ), 2075 cm−1 (), and at 2050 cm−1 (-●-), for the solvated (a) and desolvated (b) samples of 3.
Magnetochemistry 07 00075 g010
Figure 11. Structure and dimensions of the tetragonal-like channels where the CH3CN molecules are located in compounds 13.
Figure 11. Structure and dimensions of the tetragonal-like channels where the CH3CN molecules are located in compounds 13.
Magnetochemistry 07 00075 g011
Table 1. Relative increase Fe···Fe distances (d1 and d2) as function of the nature of the NCE ancillary ligands (dN….E) in the two isomorphs 1 and 2.
Table 1. Relative increase Fe···Fe distances (d1 and d2) as function of the nature of the NCE ancillary ligands (dN….E) in the two isomorphs 1 and 2.
2 (E = BH3)1 (E = S)3 (E = Se)
dN….E (relative increase)2.72 Å2.80 Å (+2.9 %)2.94 Å (+8.1 %)
T/Spin State200 K/LS100 K/LS
d1 (relative increase)12.75714.774 Å (+15.8%)
d2 (relative increase)11.683 Å12.653 Å (+8.3%)
d3 (relative increase)8.146 Å10.481 Å (+28.6%)
d4 (relative decrease)9.125 Å8.907 Å (−2.4%)
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Cuza, E.; Benmansour, S.; Cosquer, N.; Conan, F.; Gómez-García, C.J.; Triki, S. Solvent-Induced Hysteresis Loop in Anionic Spin Crossover (SCO) Isomorph Complexes. Magnetochemistry 2021, 7, 75. https://doi.org/10.3390/magnetochemistry7060075

AMA Style

Cuza E, Benmansour S, Cosquer N, Conan F, Gómez-García CJ, Triki S. Solvent-Induced Hysteresis Loop in Anionic Spin Crossover (SCO) Isomorph Complexes. Magnetochemistry. 2021; 7(6):75. https://doi.org/10.3390/magnetochemistry7060075

Chicago/Turabian Style

Cuza, Emmelyne, Samia Benmansour, Nathalie Cosquer, Françoise Conan, Carlos J. Gómez-García, and Smail Triki. 2021. "Solvent-Induced Hysteresis Loop in Anionic Spin Crossover (SCO) Isomorph Complexes" Magnetochemistry 7, no. 6: 75. https://doi.org/10.3390/magnetochemistry7060075

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