Next Article in Journal / Special Issue
Halogen Substituent Effect on the Spin-Transition Temperature in Spin-Crossover Fe(III) Compounds Bearing Salicylaldehyde 2-Pyridyl Hydrazone-Type Ligands and Dicarboxylic Acids
Previous Article in Journal / Special Issue
Heteroleptic and Homoleptic Iron(III) Spin-Crossover Complexes; Effects of Ligand Substituents and Intermolecular Interactions between Co-Cation/Anion and the Complex
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Pybox-Iron(II) Spin-Crossover Complexes with Substituent Effects from the 4-Position of the Pyridine Ring (Pybox = 2,6-Bis(oxazolin-2-yl)pyridine)

Department of Engineering Science, The University of Electro-Communications, Chofu, Tokyo 182-8585, Japan
*
Author to whom correspondence should be addressed.
Inorganics 2017, 5(3), 52; https://doi.org/10.3390/inorganics5030052
Submission received: 30 June 2017 / Revised: 27 July 2017 / Accepted: 31 July 2017 / Published: 8 August 2017
(This article belongs to the Special Issue Spin-Crossover Complexes)

Abstract

:
Spin-crossover (SCO) behavior of a series of [Fe(X-pybox)2](ClO4)2 was investigated, where X-pybox stands for 4-X-substituted 2,6-bis(oxazolin-2-yl)pyridine with X = H, Cl, Ph, CH3O, and CH3S. We confirmed that the mother compound [Fe(H-pybox)2](ClO4)2 underwent SCO above room temperature. After X was introduced, the SCO temperatures (T1/2) were modulated as 310, 230, and 330 K for X = Cl, Ph, and CH3S, respectively. The CH3O derivative possessed the high-spin state down to 2 K. Crystallographic analysis for X = H, Cl, CH3O, and CH3S was successful, being consistent with the results of the magnetic study. Distorted coordination structures stabilize the HS (high-spin) state, and the highest degree of the coordination structure distortion is found in the CH3O derivative. A plot of T1/2 against the Hammett substituent constant σp showed a positive relation. Solution susceptometry was also performed to remove intermolecular interaction and rigid crystal lattice effects, and the T1/2’s were determined as 260, 270, 240, 170, and 210 K for X = H, Cl, Ph, CH3O, and CH3S, respectively, in acetone. The substituent effect on T1/2 became very distinct, and it is clarified that electron-donating groups stabilize the HS state.

Graphical Abstract

1. Introduction

Spin-crossover (SCO) is a reversible transition between low-spin (LS) and high-spin (HS) states by external stimuli like heat, light, pressure, or magnetic field [1,2,3,4,5,6,7]. Iron(II) (3d6) coordination compounds attract a great deal of attention in various SCO complexes, because SCO takes place between dia- and paramagnetic states to show drastic change in magnetic and chromic properties. The six-nitrogen donor structures (i.e., FeIIN6) have been studied extensively, and in particular diimines like 2-pyridylmethyleneimine and bipyridyl derivatives [8,9,10,11,12,13] and triimines like 2,6-bis(azaaryl)pyridine derivatives and tripodal tris(azaaryl) compounds [14,15,16,17,18,19] are the most popular ligands for this purpose. The structural similarity found between the meridional SCO ligands 1-bpp [14] and 3-bpp [15] (2,6-bis(pyrazol-1-yl)- and 2,6-bis(pyrazol-3-yl)pyridine, respectively) affords us a clue for the development of robust SCO ligands with respect to the ligand-field engineering.
In this study, the synthesis and physical properties of Fe2+ complexes were investigated (Scheme 1), in which pybox (2,6-bis(oxazolin-2-yl)pyridine) ligands were involved as a triimine ligand. The pybox coordination compounds have been intensively studied for various reaction catalysts often carrying a chiral center in the oxazoline rings [20,21] and luminescent lanthanoid complexes as relatively recent topics [22,23,24]. More recently, Gao et al. explored the SCO application of pybox-iron(II) systems; the mother X = H compound [Fe(pybx)2](ClO4)2 [25] and a tetramethylated compound [Fe(L)2](ClO4)2 (L = 2,6-bis(4,4-dimethyloxazolin-2-yl)pyridine) [26] underwent SCO around T1/2 = 345 and 162–176 K, respectively, with attention concentrated on the solvent and counter anion dependence. Furthermore, Halcrow et al. reported that chiral (S,S)- or (R,R)-4,4’-dimethylated derivatives were subjected to the SCO study [27]. On the other hand, we are now focusing on intramolecular substituent effects, because the introduction of substituents would bring about drastic SCO tuning through the covalent bonds, rather than through van der Waals interaction. Furthermore, the physical properties should be designed and controlled in a non-serendipitous way.
We planned to modulate the SCO temperature with the aid of electronic and steric effects from a substituent group which is bound at the 4-position of the pyridine ring. The electronic substituent effects will be mainly discussed in this work, since the steric effects are hardly parameterized. After the difference of crystal packing motif is ignored, the Hammett substituent constants were applied to the relationship analysis. A role of the coordination structure distortion will be clarified. At the next stage, solution susceptometry was performed to remove intermolecular interaction and rigid crystal lattice effects. The intrinsic substituent dependence on the SCO behavior will be revealed, as discussed in connection with the known parallel results on the SCO complexes having a related meridional ligand.

2. Results

2.1. Preparation

New compounds [Fe(X-pybox)2](ClO4)2 (X = Cl, Ph, CH3O, CH3S) were prepared according to the conventional method [23,28,29,30]. All the ligands were known, except for X = CH3S. The complex formation using Fe(ClO4)2·6H2O was conducted in methanol. The resulting dark red polycrystalline compound was isolated on a filter, and the product was purified by recrystallization from methanol. The elemental and spectroscopic analyses supported the formula of [Fe(X-pybox)2](ClO4)2. Solvated molecules were found in the X = Ph and CH3S case.

2.2. Crystal Structures

The X-ray crystallographic analysis on [Fe(X-pybox)2](ClO4)2 (X = H, Cl, CH3O, CH3S) was successful at 100 and/or 400 K (Table 1 and Figure 1). The crystal structure of the mother compound [Fe(H-pybox)2](ClO4)2 at 173 K has already been reported [25]; the crystal system is monoclinic P21/n with a = 15.386(3), b = 10.700(2), c = 16.934(3) Å, β = 103.31(3)°, and V = 2713.2(10) Å3. We determined the crystal structure of this compound at 400 K as well as 100 K. The space group was kept at 400 K, and the cell volume was enlarged by 8.4% (Table 1). The Fe–N distances are listed in Table 2, and suggest that the population is almost shifted to the HS state at 400 K.
The space group of the [Fe(Cl-pybox)2](ClO4)2 crystal is P21/c, and there is a unique crystallographically-independent molecule in a unit cell. The cell parameters (Table 1) were converted to a space group P21/n to give a = 19.314(3), b = 11.170(2), c = 14.787(2) Å, and β = 103.31(3)° at 100 K. The molecular arrangement motif is different from that of [Fe(H-pybox)2](ClO4)2. The Fe–N bond lengths (Table 2) guarantee the LS state at 100 K. The crystal structure at 400 K was also determined. The space group was maintained, and the cell volume was somewhat enlarged by 7.5% (Table 1). The Fe–N bond lengths were elongated by 10% (Table 2). This finding suggests that most of the population occupies the HS state at 400 K.
The space group of the crystal of [Fe(CH3O-pybox)2](ClO4)2 is monoclinic Cc, and the Flack parameter was satisfactorily reduced. There is a unique crystallographically-independent molecule in a unit cell. The molecular packing motif is completely different from those of the H and Cl derivatives. From the Fe–N bond lengths (Table 2), the state is suggested to be HS. Table 3 summarizes the N–Fe–N bond angles. As a common feature, the N–Fe–N bite angles are largely deviated from the right angle owing to the five-membered chelate rings. However, the molecular structure of the CH3O derivative clearly exhibits a bent N(py)–Fe–N(py) axis (the angle of 156.74(12)°) (Figure 1c). In contrast, the 400 K structures of the X = H and Cl derivatives show practically linear N(py)–Fe–N(py) axes (177.6(2) and 176.50(15)°, respectively).
Compound [Fe(CH3S-pybox)2](ClO4)2 crystallizes in a space group orthorhombic P212121. The Flack parameter was satisfactorily reduced. There is a unique crystallographically-independent molecule. The crystal involves two molar methanol molecules; namely, the composition formula is [Fe(CH3S-pybox)2](ClO4)2(CH3OH)2, being consistent with the results of the elemental analysis and magnetic measurements (see below). From the Fe–N bond lengths (Table 2), the state is suggested to be LS at 100 K. The molecular structure (Figure 1d) displays a relatively linear N(py)–Fe–N(py) axis (176.62(19)°), as usually found in the LS molecules of this family. At high temperatures (e.g., 400 K), the crystals were broken down and the crystal structure analysis was unsuccessful. The elemental analysis suggests that some methanol molecules would be liberated from the crystal lattice.

2.3. Mangetic Properties of Polycrystalline Specimens

The magnetic susceptibilities of polycrystalline specimens of [Fe(X-pybox)2](ClO4)2 were measured on a SQUID magnetometer in a temperature range of 10–400 K. As Figure 1 shows, the high temperature limit of the χmT value was 3.5–4.0 cm3·K·mol−1, being compatible with the S = 2 HS state of the iron(II) ion. On cooling, the χmT value of [Fe(H-pybox)2](ClO4)2 gradually decreased and reached practically null below 270 K. Though the χmT value did not reach the high-temperature limit even at 400 K, the present SCO behavior is supposed to reproduce well the reported data (T1/2 = 345 K [25]). Similarly, [Fe(X-pybox)2](ClO4)2 (X = Cl, Ph, CH3S) showed SCO, and the transition temperatures (T1/2) were determined to be 310, 230, and 330 K for X = Cl, Ph, and CH3S, respectively. The CH3O derivative possessed the HS state in all the temperature range investigated here.
The CH3S derivative exhibited a χmT jump at 350 K. After the jump, the χmT profile never reproduced the original one because of the loss of the solvent molecules. The crystal structure analysis of the specimen after the thermal cycle was unsuccessful, and the solvent loss was confirmed by the elemental and spectroscopic analyses. We can exclude possibility that the χmT increase would dominantly come from the solvent discharge “reaction”. In the present case, from a close look, the χmT reached the half level before the jump, so the T1/2 can be correctly defined to be 330 K before the reaction. However, only the solvated form was structurally characterized and will be incorporated to the magneto-structure discussion, implying that the T1/2 data of [Fe(CH3S-pybox)2](ClO4)2 involved the effect of the solvation.
Figure 2b depicts the plot of T1/2 against the Hammett substituent constant σp (−0.27, −0.01, 0, 0.00, 0.23 for the CH3O, Ph, H, CH3S, and Cl groups, respectively) [31]. No SCO appeared in the χmT(T) profile of the CH3O derivative. To incorporate this result into the present discussion, it is enough to suppose that the G(T) level crossing temperature between the HS and LS states must be located much below those of the H, Cl, Ph, and CH3S derivatives (230–350 K). Therefore, we can safely conclude a positive relationship, which indicates that electron-donating groups suppress T1/2 while electron-withdrawing groups raise T1/2. In Figure 2b, The CH3O data point has tentatively been placed at 0 K. This trend is qualitatively compatible with that of the 1-bpp series [Fe(X-1-bpp)2](BF4)2 (X-1-bpp = 4-X-substituted 1-bpp) [32]. The rigid crystal lattice may play a role of a spin-transition inhibitor. The conclusion seems to have an approximate meaning, because the various space groups and molecular packing motifs were found but neglected in the above discussion.

2.4. Mangetic Properties of Solution Specimens

In solution, many of the problems from intermolecular interactions in various crystal lattices are removed, so the intrinsic intramolecular substituent effect can be extracted from variable-temperature experiments. We carried out solution susceptometry for the present complexes, inspired by the work on [Fe(X-1-bpp)2](BF4)2 by Halcrow and co-workers [32]. They utilized the Evans method, but we applied the conventional SQUID apparatus. The results of the acetone-solution magnetic susceptibility measurements are summarized in Figure 3a. The data were acquired on cooling until the solution was solidified. We have found no indication of possible ligand-iron ion dissociation reaction. The T1/2’s were determined as 260, 270, 240, 170, and 210 K for [Fe(X-pybox)2](ClO4)2 with X = H, Cl, Ph, CH3O, and CH3S, respectively. The CH3O derivatives exhibited a remarkable difference between the solid-state and solution data; namely, the HS state at any temperature in the solid state, whereas T1/2 = 170 K was characterized in the acetone solution. The broad transition widths observed are consistent with the complete absence of cooperativity.
Similarly to the solid-state experiments, we plotted the solution T1/2 against the Hammett substituent constant σp [31] (Figure 3b). We can easily find a positive relationship, which indicates that electron-donating groups suppress T1/2 while electron-withdrawing groups raise T1/2. An empirical relationship equation is described with Equation (1) (a line in Figure 3b). This trend is similar to that of the 1-bpp series [32]. The solution data have the advantage of negligible intermolecular interaction. However, no structural information is afforded from the solution experiment.
T 1 / 2 = 200 ( 60 ) K σ p + 230 ( 10 ) K

3. Discussion

3.1. Substituent Effect

In the present study, the Hammett substituent constants are available to explain the static electronic effect from substituents [33]. There are two major electronic effects from substituents: inductive effect along the σ electron network and mesomeric effect along the π electron network. The 4-position in a pyridine ring is preferable for the mesomeric effect from the pyridine nitrogen site (i.e., the 1-position). Applying the Hammett σp constant is reasonable in this system. The π-electron-withdrawing group stabilizes the LS state. This is because the increase of π electron delocalization into the ligand leads to an increase in the ligand-field splitting parameter [34,35]. In other words, the iron(II) t2g orbitals are stabilized owing to the dπ–pπ orbital interaction, and the ligand-field splitting is enhanced.

3.2. Coordination Structure Effect

The HS states are known to favor distorted coordination geometry in general [34,35,36,37,38,39], which is related to long Fe–N distances and also accommodation of steric congestion during transition. From the N–Fe–N bond angles summarized in Table 3, we can calculate the Σ value [37] for estimation of distortion degree (Table 4), according to Equation (2). An ideal octahedron (Oh) possesses Σ = 0°. Other popular parameters are also listed in Table 4. The Θ value represents the deviation of the coordination geometry from an octahedron to a trigonal prism [38]. An ideal trigonal antiprism viewed from the principal axis contributes null, and an ideal Oh leads to Θ = 0°. The α value is the average of the four N–Fe–N bite angles [39]. By using the SHAPE software [40], the continuous shape measures (CShM) are calculated with respect to an Oh. An ideal Oh returns null. The ϕ (Npy–Fe–Npy angle) and θ (dihedral angle between two pybox systems) values are measures of the angular Jahn–Teller distortion, as Halcrow et al. proposed to describe the criterion of SCO [Fe(X-1-bpp)2]2+ and related compounds [14]. Very recently, a new empirical rule has been proposed based on the interatomic N–N distance in the chelatable diimine structure [36]. Since there is no structural data of the free ligands, the distances were calculated with the density functional theory on the b3lyp/6-311+G(2d,p) level in Gaussian 03 [41].
= i = 1 12 | ( c i s N - Fe - N ) i 90 ° ) |
The geometrical parameters of the CH3O derivative could be obtained only for the HS state. Accordingly, the X = CH3O coordination structure at 100 K was compared with the X = H and Cl structures at 400 K, and the highest degree of distortion was found in the CH3O derivative, as indicated with the distortion parameters (Table 4) as well as the determined molecular structure (Figure 1c). As for the LS states, when the X = Cl and CH3S structures at 100 K were compared with the X = H structure at 173 K, their distortions were comparable to each other. These findings are confirmed by the parameters; in particular, Σ and CShM seem to be sensitive and convenient metrics. The ϕ and θ values of the HS [Fe(CH3O-pybox)2](ClO4)2 predict that this compound is not an SCO compound, and in fact it was found to be an HS compound in the whole temperature range (Figure 2a). On the other hand, the calculated N–N distance is insensitive to the substitution. As Figure 4a exemplifies, the CShM vs. σp plot has a relation with a negative slope, implying that the electron-donating group would favor a distorted structure. Such distortion would bring about preference for the HS states due to reduction of the ligand-field strength [42,43,44,45] and suppress the SCO temperature. Gao and co-workers have already reported the SCO study on a [Fe(H-pybox)2]2+ series with counter anion and solvent variation [25] with the substituent X fixed to H, and the magneto-structural relation was clarified to give a partially similar conclusion: highly distorted structures are favorable for the HS state and only intermediately distorted compounds show SCO.
As Figure 2 displays, although the difference of molecular packing motif is ignored, the intramolecular substituent effects seem to be dominant and approximately regulate the overall T1/2 trend in the solid state. As for [Fe(CH3O-pybox)2](ClO4)2, the CH3O substituent brings about an excessive effect through the coordination structure distortion, as indicated by the drastic difference from the solution result. The substituents regulate both geometry and ligand-field strength. The plot of T1/2 for [Fe(X-1-bpp)2](BF4)2 against T1/2 for [Fe(X-pybox)2](ClO4)2 displayed a positive correlation (Figure 4b). The platforms are different, but the common substituent effect is still observable. Consequently, the SCO characteristics can be discussed in connection with the substituent effect. We have to pay attention to an indirect mechanism, where the substituent effect is operative through geometrical modification and coordination structure distortion, together with the direct mechanism from the electronic substituent effect.

3.3. Electronic Substituent Effect

The intrinsic intramolecular substituent effect can be extracted from the solution data. The results on the solution SQUID susceptometry for [Fe(X-pybox)2](ClO4)2 (X = H, Cl, Ph, CH3O, CH3S) are summarized in Figure 3a. The solution T1/2 vs. σp plot displays a positive relationship (Figure 3b). The solution data have no structural information, and the coordination structures might be modified by the substituent effects. However, we can assume that structural distortion would be minor from thinking of the vast conformational freedom in solution. Thus, the solution results seem to be more reliable than those of the solid-state experiments to evaluate substituent effect. The conclusion is derived here, which is basically the same as the solid-state experiments. Electron-donating groups stabilize the HS state, whereas electron-withdrawing groups stabilize the LS one.
The plot of T1/2 of [Fe(X-1-bpp)2](BF4)2 in acetone solutions against the solution T1/2 of [Fe(X-pybox)2](ClO4)2 displayed a positive correlation (Figure 5). This relationship is more evident than that of Figure 4b, because both were acquired in solutions. An empirical relationship is formulated as Equation (3) and superposed in Figure 5. In the doubly meridional chelated systems involving 4-X-1-dpp and 4-X-pybox, the structural distortion effects would be similar to each other. Such distortion effects are assumed to be cancelled out in the comparison analysis; the linear relationship observed here is accounted for in terms of the essential electronic substituent effect represented by the Hammett substituent constant σp.
T 1 / 2 ( pybox ) = 1.10 ( 4 ) T 1 / 2 ( bpp )

4. Experimental Section

4.1. Materials

Caution! We have not yet encountered any hazard, but the perchlorate salts should be handled with care.
The ligands X-pybox (X = H [23], Cl [28], Ph [29], and CH3O [30]) were prepared according to the known procedure. A new derivative CH3S-pybox was prepared according to the known procedure for the CH3O-pybox [30] with modification. A mixture of Cl-pybox (0.3707 g; 1.47 mmol) and sodium methylthiolate (0.1212 g; 1.73 mmol) in dry N,N-dimethylformamide was stirred at 40 °C for 12 h. After being cooled, the mixture was poured into aqueous NaHCO3, and organic substrates were extracted with ethyl acetate. The organic layer was washed with aqueous NaHCO3, dried over MgSO4, and filtered. Concentration of the filtrate gave CH3S-pybox (0.4266 g; 1.34 mmol) in 91% yield as a colorless solid. m.p. 193–195 °C. 1H NMR (ECA-500, JEOL, Tokyo, Japan) (500 MHz, CDCl3) δ 7.95 (2H, s), 4.53 (4H, t, J = 9.6 Hz), 4.11 (4H, t, J = 9.6 Hz), 2.18 (3H, s). 13C NMR (126 MHz, CDCl3) δ 163.53, 152.87, 146.27, 121.26, 68.47, 55.09, 13.96. MS (ESI+) (JMS-T100 AccuTOF, JEOL, Tokyo, Japan) m/z 286.03 (M + Na+), 264.05 (M + H+). IR (neat, attenuated total reflection (ATR)) (Nicolet 6700, Thermo Scientific, San Jose, CA, USA) 1638, 1572, 1382, 1122, 942, 868, 786, 670, 533 cm−1.
The target complexes were prepared as follows. A mixture of Cl-pybox (0.0559 g; 0.226 mmol), Fe(ClO4)2·6H2O (0.0211 g; 0.113 mmol), and l-ascorbic acid (11 mg) in methanol (18 mL) was allowed to stand in a refrigerator for 24 h. The filtration gave 0.039 g (0.052 mmol) of [Fe(Cl-pybox)2](ClO4)2. Yield 47%. m.p. 204 °C (dec.). The product was purified by recrystallization from methanol prior to the elemental and spectroscopic analyses, crystallographic analysis, and magnetic study. IR (neat, ATR) 1612, 1499, 1274, 1033, 926, 760, 651 cm−1. Anal. Calcd. for C22H20Cl4FeN6O12: C, 34.86%; H, 2.66%; N, 11.09%. Found: C, 35.02%; H, 2.55%; N, 11.37%.
Similarly, compounds [Fe(X-pybox)2](ClO4)2 (X = Ph, CH3O, and CH3S) were prepared, and the yields were 44%, 26%, and 90%, respectively. [Fe(Ph-pybox)2](ClO4)2·H2O: m.p. 209 °C (dec.). IR (neat, ATR) 1584, 1374, 1269, 1068, 913, 757, 620 cm−1. Anal. Calcd. for C34H30Cl2FeN6O12·H2O: C, 47.16%; H, 4.07%; N, 9.43%. Found: C, 47.04%; H, 3.55%; N, 9.48%. [Fe(CH3O-pybox)2](ClO4)2: m.p. 289 °C (dec.). IR (neat, ATR) 1587, 1372, 1239, 1072, 916, 859, 620, 578 cm−1. Anal. Calcd. for C24H26Cl2FeN6O14: C, 38.47%; H, 3.50%; N, 11.22%. Found: C, 38.53; H, 3.29; N, 11.16%. [Fe(CH3S-pybox)2](ClO4)2·2CH3OH: m.p. 208 °C (dec.). IR (neat, ATR) 1574, 1485, 1372, 1270, 1069, 915, 796, 619 cm−1. Anal. Calcd. for C24H26Cl2FeN6O12S2·CH4O·H2O: C, 36.11%; H, 3.88%; N, 10.11%; S, 7.71%. Found: C, 36.14%; H, 3.58%; N, 10.08%; S, 7.70%. After thermal treatment at 400 K in a SQUID magnetometer. Anal. Calcd. for C24H26Cl2FeN6O12S2: C, 36.89%; H, 3.35%; N, 10.76%; S, 8.21%. Found: C, 36.56%; H, 3.18%; N, 10.60%; S, 8.19%.

4.2. Crystallographic Analysis

X-Ray diffraction data of [Fe(X-pybox)2](ClO4)2 (X = H, Cl, CH3O, CH3S) were collected on a Saturn70 CCD diffractometer (Rigaku, Tokyo, Japan) with graphite monochromated Mo Kα radiation (λ = 0.71073 Å). The structures were directly solved by a heavy-atom method and expanded using Fourier techniques in the CrystalStructure [46]. Numerical absorption correction was used. Hydrogen atoms were located at calculated positions, and their parameters were refined as a riding model. The thermal displacement parameters of non-hydrogen atoms were refined anisotropically. Selected crystallographic data are given in Table 1, and selected bond distances and angles are listed in Table 2 and Table 3, respectively. CCDC numbers 1559025, 1559026, 1559027, 1559028, and 1559029 for [Fe(X-pybox)2](ClO4)2 (X = H, Cl (100 K), Cl (400 K), CH3O, CH3S, respectively) include the experimental details and full geometrical parameter tables. These data can be obtained free of charge via http://www.ccdc.cam.ac.uk/conts/retrieving.html.

4.3. Magnetic Study

Magnetic susceptibilities of [Fe(X-pybox)2](ClO4)2 (X = H, Cl, Ph, CH3O, CH3S) were measured on a Quantum Design MPMS-XL7 SQUID magnetometer (San Diego, CA, USA) with a static field of 0.5 T. The magnetic responses were corrected with diamagnetic blank data of the sample holder measured separately. The diamagnetic contribution of the sample itself was estimated from Pascal’s constants. The solution magnetic susceptibilities were measured in the same SQUID apparatus. The specimen was dissolved in acetone, and the resultant clear solution was transferred into a 5 mm·ϕ NMR sample tube. After the solution was degassed with argon gas bubbling, the sample tube was sealed and mounted in a SQUID probe. The susceptibility data were acquired on cooling until the acetone solution was solidified (typically around 160 K). In the analysis, the diamagnetic contribution was numerically estimated as a temperature-independent susceptibility term, so that the χmT values of the LS and/or HS regions should be almost constant. Owing to the ambiguity of the sample amount, the S-shaped profile was drawn on the molar fraction basis, γHS = x(HS)/(x(HS) + x(LS)) against temperature.

4.4. DFT Calculation Study

Density-functional-theory (DFT) calculation was performed by using the Gaussian 03 package [41]. The geometry was optimized after the b3lyp Hamiltonian and the 6-311+G(2d,p) basis set were chosen. The convergence criterion was below 10−8 a.u. in the self-consistent field energy.

5. Conclusions

We synthesized four new iron(II) complexes (Cl, Ph, CH3O, and CH3S). The SCO transition temperature T1/2 can be changed by introducing a substituent at the 4-position of the pyridine ring. The T1/2 around or slightly above room temperature may be very attractive for future application of the SCO materials. In the solid-state study, there seems to be a correlation between σp and Σ or σp and CShM, and furthermore between σp and T1/2. The coordination structure distortion depends on the substituents, and the distortion also indirectly regulates T1/2. Distorted structures stabilize the HS state. In the solution study, the substituent dependence on T1/2 became very obvious, and electron-donating groups stabilize the HS state. The SCO temperature is regulated by the substituents, being similar to the known parallel work on [Fe(X-1-bpp)2](BF4)2. The platforms are different, but the SCO characteristics can be discussed in connection with the substituent effect in a generalized manner.

Acknowledgments

This work was financially supported from KAKENHI (JSPS/15H03793).

Author Contributions

Akifumi Kimura participated in the preparation, X-ray structural analysis, and magnetic study. Takayuki Ishida designed the study and wrote the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gütlich, P.; Goodwin, H.A. (Eds.) Spin Crossover in Transition Metal Compounds I, II, and III; Springer: Berlin, Germany, 2004. [Google Scholar]
  2. Gütlich, P.; Gaspar, A.B.; Garcia, Y. Spin state switching in iron coordination compounds. Beilstein J. Org. Chem. 2013, 9, 342–391. [Google Scholar] [CrossRef] [PubMed]
  3. Halcrow, M.A. Spin-Crossover Materials: Properties and Applications; John Wiley & Sons, Ltd.: Oxford, UK, 2013. [Google Scholar]
  4. Halcrow, M.A. Spin-Crossover compounds with wide thermal hysteresis. Chem. Lett. 2014, 43, 1178–1188. [Google Scholar] [CrossRef]
  5. Kahn, O. Chapter 4. In Molecular Magnetism; VCH: Weinhein, Germany, 1993. [Google Scholar]
  6. Harding, D.J.; Harding, P.; Phonsri, W. Spin crossover in iron(III) complexes. Coord. Chem. Rev. 2016, 313, 38–61. [Google Scholar] [CrossRef]
  7. Letard, J.-F.; Guionneau, P.; Nguyen, O.; Costa, J.S.; Marcen, 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]
  8. Letard, J.-F.; Guionneau, P.; Godjovi, E.; Lavastre, O.; Bravic, G.; Chasseau, D.; Kahn, O. Wide Thermal Hysteresis for the Mononuclear Spin-Crossover Compound cis-Bis(thiocyanato)bis[N-(2’-pyridylmethylene)-4-(phenylethynyl)anilino]iron(II). J. Am. Chem. Soc. 1997, 119, 10861–10862. [Google Scholar] [CrossRef]
  9. Yamada, M.; Hagiwara, H.; Torigoe, H.; Matsumoto, N.; Kojima, M.; Dahan, F.; Tuchagues, J.P.; Re, N.; Iijima, S. A variety of spin-crossover behaviors depending on the counter anion: Two-dimensional complexes constructed by NH···Cl hydrogen bonds, [(FeIIH3LMe)]Cl·X (X = PF6, AsF6, SbF6, CF3SO3; H3LMe = tris[2-{[(2-methylimidazol-4-yl)methylidene]amino}ethyl]amine). Chem. Eur. J. 2006, 12, 4536–4549. [Google Scholar] [PubMed]
  10. Takahashi, K.; Kawakami, T.; Gu, Z.; Einaga, Y.; Fujishima, A.; Sato, O. An abrupt spin transition based on short S...S contacts in a novel Fe(II) complex whose ligand contains a 1,3-dithiole ring. Chem. Commun. 2003, 2374–2375. [Google Scholar] [CrossRef]
  11. 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]
  12. Oso, Y.; Ishida, T. Spin-crossover transition in a mesophase iron(II) thiocyanate complex chelated with 4-hexadecyl-N-(2-pyridylmethylene)aniline. Chem. Lett. 2009, 38, 604–605. [Google Scholar] [CrossRef]
  13. Oso, Y.; Kanatsuki, D.; Saito, S.; Nogami, T.; Ishida, T. Spin-crossover transition coupled with another solid-solid phase transition for iron(II) thiocyanate complexes chelated with alkylated N-(di-2-pyridylmethylene)anilines. Chem. Lett. 2008, 37, 760–761. [Google Scholar] [CrossRef]
  14. Halcrow, M.A. Iron(II) complexes of 2,6-di(pyrazol-1-yl)pyridines—A versatile system for spin-crossover research. Coord. Chem. Rev. 2009, 253, 2493–2514. [Google Scholar] [CrossRef]
  15. Craig, G.A.; Roubeau, O.; Aromi, G. Spin state switching in 2,6-bis(pyrazol-3-yl)pyridine (3-bpp) based Fe(II) complexes. Coord. Chem. Rev. 2014, 269, 13–31. [Google Scholar] [CrossRef]
  16. Krober, J.; Codjovi, E.; Kahn, O.; Groliere, F.; Jay, C. A Spin Transition System with a Thermal Hysteresis at Room Temperature. J. Am. Chem. Soc. 1993, 115, 9810–9811. [Google Scholar] [CrossRef]
  17. 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]
  18. 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]
  19. 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]
  20. Johnson, J.S.; Evans, D.A. Chiral Bis(oxazoline) Copper(II) Complexes: Versatile Catalysts for Enantioselective Cycloaddition, Aldol, Michael, and Carbonyl Ene Reactions. Acc. Chem. Res. 2000, 33, 325–335. [Google Scholar] [CrossRef] [PubMed]
  21. Desimoni, G.; Faita, G.; Quadrelli, P. Pyridine-2,6-bis(oxazolines), Helpful Ligands for Asymmetric Catalysts. Chem. Rev. 2003, 103, 3119–3154. [Google Scholar] [CrossRef] [PubMed]
  22. Yuasa, J.; Ohno, T.; Miyata, K.; Tsumatori, H.; Hasegawa, Y.; Kawai, T. Noncovalent Ligand-to-Ligand Interactions Alter Sense of Optical Chirality in Luminescent Tris(β-diketonate) Lanthanide(III) Complexes Containing a Chiral Bis(oxazolinyl) Pyridine Ligand. J. Am. Chem. Soc. 2011, 133, 9892–9902. [Google Scholar] [CrossRef] [PubMed]
  23. De Bettencourt-Dias, A.; Barber, P.S.; Viswanathan, S.; de Lill, D.T.; Rollett, A.; Ling, G.; Altun, S. Para-Derivatized Pybox Ligands As Sensitizers in Highly Luminescent Ln(III) Complexes. Inorg. Chem. 2010, 49, 8848–8861. [Google Scholar] [CrossRef] [PubMed]
  24. De Bettencourt-Dias, A.; Barber, P.S.; Bauer, S. A Water-Soluble Pybox Derivatives and Its Highly Luminescent Lanthanide Ion Complexes. J. Am. Chem. Soc. 2012, 134, 6987–6994. [Google Scholar] [CrossRef] [PubMed]
  25. Zhu, Y.-Y.; Li, H.-Q.; Ding, Z.-Y.; Lu, X.-J.; Zhao, L.; Meng, Y.-S.; Liu, T.; Gao, S. Spin transition in a series of [Fe(pybox)2]2+ complexes modulated by ligand structures, counter anions, and solvents. Inorg. Chem. Front. 2016, 3, 1624–1636. [Google Scholar] [CrossRef]
  26. Zhu, Y.-Y.; Liu, C.-W.; Yin, J.; Meng, Z.-S.; Yang, Q.; Wang, J.; Liu, T.; Gao, S. Structural phase transition in a multi-induced mononuclear FeII spin-crossover complex. Dalton Trans. 2015, 44, 20906–20912. [Google Scholar] [CrossRef] [PubMed]
  27. Burrows, K.E.; McGrath, S.E.; Kulmaczewski, R.; Cespedes, O.; Barrett, S.A.; Halcrow, M.A. Spin State of Homochiral and Heterochiral Isomers of [Fe(PyBox)2]2+ Derivatives. Chem. Eur. J. 2017, 23, 9067–9075. [Google Scholar] [CrossRef] [PubMed]
  28. De Bettencourt-Dias, A.; Rossini, J.S.K. Ligand Design for Luminescent Lanthanide-Containing Metallopolymers. Inorg. Chem. 2016, 55, 9954–9963. [Google Scholar] [CrossRef] [PubMed]
  29. Yu, X.; Yang, T.; Wang, S.; Xu, H.; Gong, H. Nickel-Catalyzed Reductive Cross-Coupling of Unactivated Alkyl Halides. Org. Lett. 2011, 13, 2138–2141. [Google Scholar] [CrossRef] [PubMed]
  30. Vermonden, T.; Branowska, D.; Marcelis, A.T.M.; Sudholter, E.J.R. Synthesis of 4-functionalized terdendate pyridine-based ligads. Tetrahedron 2003, 59, 5039–5045. [Google Scholar] [CrossRef]
  31. Hansch, C.; Leo, A.; Taft, R.W. A survey of Hammett substituent constants and resonance and field parameters. Chem. Rev. 1991, 91, 165–195. [Google Scholar] [CrossRef]
  32. Cook, L.J.K.; Rafal, K.; 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] [PubMed]
  33. Isaccs, N.S. Chapter 4. In Physical Organic Chemistry; Wiley: New York, NY, USA, 1987. [Google Scholar]
  34. Tweedle, M.F.; Wilson, L.J. Variable Spin Iron(III) Chelates with Hexadentate Ligands Derived from Triethylenetetramine and Various Salicylaldehydes. Synthesis, Characterization, and Solution State Studies of a New 2T ↔ 6A Spin Equilibrium System. J. Am. Chem. Soc. 1976, 98, 4824–4834. [Google Scholar] [CrossRef]
  35. Takahashi, K.; Hasegawa, Y.; Sakamoto, R.; Nishikawa, M.; Kume, S.; Nishibori, E.; Nishihara, H. Solid-State Ligand-Driven Light-Induced Spin Change at Ambient Temperatures in Bis(dipyrazolylstyrylpyridine)iron(II) Complexes. Inorg. Chem. 2012, 51, 5188–5198. [Google Scholar] [CrossRef] [PubMed]
  36. Phan, H.; Hrudka, J.J.; Igimbayeva, D.; Daku, L.M.L.; 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]
  37. Guionneau, P.; Marchivie, M.; Bravic, G.; Létard, J.-F.; Chasseau, D. Structural Aspects of Spin Crossover. Example of the [FeIILn(NCS)2] Complexes. Top. Curr. Chem. 2004, 234, 97–128. [Google Scholar]
  38. Marchivie, M.; Guionneau, P.; Letard, J.F. Photo-induced spin-transition: the role of the iron(II) environment distortion. Acta Crystallogr. Sect. B Struct. Sci. 1991, 15, 181–190. [Google Scholar] [CrossRef] [PubMed]
  39. Halcrow, M.A. Structure:function relationships in molecular spin-crossover complexes. Chem. Soc. Rev. 2011, 40, 4119–4142. [Google Scholar] [CrossRef] [PubMed]
  40. Lluncll, M.; Casanova, D.; Circra, J.; Bofill, J.M.; Alcmany, P.; Alvarez, S.; Pinsky, M.; Avnir, D. SHAPE; v2.1; University of Barcelona and The Hebrew University of Jerusalem: Barcelona, Spain, 2005. [Google Scholar]
  41. Gaussian 03; revision C.02; Gaussian Inc.: Wallingford, CT, USA, 2004.
  42. Kroll, N.; Theilacker, K.; Schoknecht, M.; Baabe, D.; Wiedemann, D.; Kaupp, M.; Grohmann, A.; Hörner, G. Controlled ligand distortion and its consequences for structure, symmetry, conformation and spin-state preferences of iron(II) complexes. Dalton Trans. 2015, 44, 19232–19247. [Google Scholar] [CrossRef] [PubMed]
  43. Matouzenko, G.S.; Jeanneau, E.; Verat, A.Y.; de Gaetano, Y. The Nature of Spin Crossover and Coordination Core Distortion in a Family of Binuclear Iron(II) Complexes with Bipyridyl-Like Bridging Ligands. Eur. J. Inorg. Chem. 2012, 969–977. [Google Scholar] [CrossRef]
  44. Cook, L.J.K.; Thorp-Greenwood, F.L.; Comyn, T.P.; Cespedes, O.; Chastanet, G.; Halcrow, M.A. Unexpected Spin-Crossover and a Low-Pressure Phase Change in an Iron(II) Dipyrazolylpyridine Complex Exhibiting a High-Spin Jahn–Teller Distortion. Inorg. Chem. 2015, 54, 6319–6330. [Google Scholar] [CrossRef] [PubMed]
  45. Yang, Q.; Cheng, X.; Gao, C.; Wang, B.; Wang, Z.; Gao, S. Structural Distortion Controlled Spin-Crossover Behavior. Cryst. Growth Des. 2015, 15, 2565–2567. [Google Scholar] [CrossRef]
  46. CRYSTALSTRUCTURE, version 4.2.1; Rigaku/MSC: The Woodlands, TX, USA, 2015.
Scheme 1. Synthesis of [Fe(X-pybox)2](ClO4)2 (X = H, Cl, Ph, CH3O, CH3S). Pybox: (2,6-bis(oxazolin-2-yl)pyridine.
Scheme 1. Synthesis of [Fe(X-pybox)2](ClO4)2 (X = H, Cl, Ph, CH3O, CH3S). Pybox: (2,6-bis(oxazolin-2-yl)pyridine.
Inorganics 05 00052 sch001
Figure 1. X-ray crystal structures of [Fe(X-pybox)2](ClO4)2 for X = (a) H (400 K); (b) Cl (100 K); (c) CH3O (100 K); and (d) CH3S (100 K). The thermal ellipsoids are drawn at the 20%, 50%, 50%, and 50% probability levels, respectively. Hydrogen atoms and counter anions are omitted for clarity.
Figure 1. X-ray crystal structures of [Fe(X-pybox)2](ClO4)2 for X = (a) H (400 K); (b) Cl (100 K); (c) CH3O (100 K); and (d) CH3S (100 K). The thermal ellipsoids are drawn at the 20%, 50%, 50%, and 50% probability levels, respectively. Hydrogen atoms and counter anions are omitted for clarity.
Inorganics 05 00052 g001
Figure 2. (a) Temperature dependence of χmT for polycrystalline [Fe(X-pybox)2](ClO4)2 measured at 5000 Oe; (b) The spin transition temperature T1/2 as a function of the Hammett substituent constant σp. The CH3O data point is tentatively placed at 0 K.
Figure 2. (a) Temperature dependence of χmT for polycrystalline [Fe(X-pybox)2](ClO4)2 measured at 5000 Oe; (b) The spin transition temperature T1/2 as a function of the Hammett substituent constant σp. The CH3O data point is tentatively placed at 0 K.
Inorganics 05 00052 g002
Figure 3. (a) Temperature dependence of the HS (high-spin) molar fraction γHS for acetone solutions of [Fe(X-pybox)2](ClO4)2, measured at 5000 Oe. (b) The spin transition temperature T1/2 as a function of the Hammett substituent constant σp. A dotted line represents the best linear fit.
Figure 3. (a) Temperature dependence of the HS (high-spin) molar fraction γHS for acetone solutions of [Fe(X-pybox)2](ClO4)2, measured at 5000 Oe. (b) The spin transition temperature T1/2 as a function of the Hammett substituent constant σp. A dotted line represents the best linear fit.
Inorganics 05 00052 g003
Figure 4. Relations between (a) the distortion parameter CShM (continuous shape measures) and Hammett substituent constant σp. (b) Plot of T1/2 for polycrystalline [Fe(X-1-bpp)2](ClO4)2 vs. T1/2 for solution [Fe(X-pybox)2](BF4)2. The [Fe(X-1-bpp)2](BF4)2 data are taken from Ref. [32].
Figure 4. Relations between (a) the distortion parameter CShM (continuous shape measures) and Hammett substituent constant σp. (b) Plot of T1/2 for polycrystalline [Fe(X-1-bpp)2](ClO4)2 vs. T1/2 for solution [Fe(X-pybox)2](BF4)2. The [Fe(X-1-bpp)2](BF4)2 data are taken from Ref. [32].
Inorganics 05 00052 g004
Figure 5. Plot of T1/2 for solution [Fe(X-1-bpp)2](ClO4)2 vs. T1/2 for solution [Fe(X-pybox)2](BF4)2. The [Fe(X-1-bpp)2](BF4)2 data are taken from Ref. [32]. A dotted line represents the best linear fit.
Figure 5. Plot of T1/2 for solution [Fe(X-1-bpp)2](ClO4)2 vs. T1/2 for solution [Fe(X-pybox)2](BF4)2. The [Fe(X-1-bpp)2](BF4)2 data are taken from Ref. [32]. A dotted line represents the best linear fit.
Inorganics 05 00052 g005
Table 1. Selected crystallographic parameters of [Fe(X-pybox)2](ClO4)2 (X = H, Cl, CH3O, CH3S).
Table 1. Selected crystallographic parameters of [Fe(X-pybox)2](ClO4)2 (X = H, Cl, CH3O, CH3S).
XHClClCH3OCH3S
T/K400100400100100
Formula weight689.20758.09758.09749.25845.46
Crystal systemmonoclinicmonoclinicmonoclinicmonoclinicorthorhombic
Space groupP21/nP21/cP21/cCcP212121
a16.140(3)14.787(2)15.857(4)12.084(3)12.656(3)
b10.9603(17)11.170(2)10.588(3)12.822(3)15.676(4)
c17.049(3)18.164(3)19.084(4)19.328(5)16.897(4)
β/°102.736(9)109.071(8)107.947(11)101.716(12)90.
V32941.8(9)2835.4(8)3048.1(13)2932.3(12)3352.1(13)
Z44444
dcalcd./g·cm−31.5561.7761.6521.6971.675
μ (MoKα)/mm−10.7630.9830.9140.7780.811
No. of unique reflections57606481696664887660
R (F) (I > 2σ (I)) a0.08300.06980.08300.03610.0503
wR (F2) (all reflections) b0.29590.19780.28630.09080.1193
Goodness-of-fit parameter0.9811.0471.0091.0521.068
Flack parameter---−0.008(4)0.002(9)
a R = Σ[|Fo| − |Fc|]/Σ|Fo|; b wR = [Σw(Fo2Fc2)/ΣwFo4]1/2.
Table 2. Fe–N bond distances (d) in Å for [Fe(X-pybox)2](ClO4)2 (X = H, Cl, CH3O, CH3S).
Table 2. Fe–N bond distances (d) in Å for [Fe(X-pybox)2](ClO4)2 (X = H, Cl, CH3O, CH3S).
XH (400 K)Cl (100 K)Cl (400 K)CH3O (100 K)CH3S (100 K)
d (Fe1–N1)2.134(6)1.956(4)2.153(4)2.209(3)1.991(4)
d (Fe1–N2)2.060(5)1.896(4)2.098(4)2.115(3)1.907(4)
d (Fe1–N3)2.161(5)1.963(4)2.159(5)2.203(3)1.981(4)
d (Fe1–N4)2.129(6)1.960(4)2.149(4)2.159(3)1.956(4)
d (Fe1–N5)2.067(5)1.901(4)2.088(4)2.136(3)1.901(4)
d (Fe1–N6)2.160(6)1.977(4)2.194(4)2.218(3)1.978(5)
Average2.121.942.142.171.95
Table 3. N–Fe–N bond angles (ϕ) in ° for [Fe(X-pybox)2](ClO4)2 (X = H, Cl, CH3O, CH3S).
Table 3. N–Fe–N bond angles (ϕ) in ° for [Fe(X-pybox)2](ClO4)2 (X = H, Cl, CH3O, CH3S).
XH (400 K)Cl (100 K)Cl (400 K)CH3O (100 K)CH3S (100 K)
ϕ (N1–Fe1–N2)75.2(2)79.55(17)74.14(16)74.73(12)78.90(18)
ϕ (N1–Fe1–N3)150.8(2)159.31(17)148.07(18)147.31(12)158.41(18)
ϕ (N1–Fe1–N4)96.0(2)92.36(15)92.48(15)103.25(12)93.7(2)
ϕ (N1–Fe1–N5)105.6(2)97.50(16)102.35(16)116.23(12)104.09(18)
ϕ (N1–Fe1–N6)91.5(2)93.41(16)95.74(16)88.78(12)89.80(19)
ϕ (N2–Fe1–N3)75.5(2)79.76(17)73.93(17)73.65(12)79.53(18)
ϕ (N2–Fe1–N4)105.9(2)101.08(16)105.09(15)125.05(12)98.96(19)
ϕ (N2–Fe1–N5)177.6(2)177.03(16)176.50(15)156.74(12)176.62(19)
ϕ (N2–Fe1–N6)102.8(2)99.94(17)105.62(16)87.91(12)102.61(19)
ϕ (N3–Fe1–N4)91.17(19)91.28(15)95.20(16)88.33(12)89.0(2)
ϕ (N3–Fe1–N5)103.7(2)103.19(16)109.57(18)96.25(12)97.45(18)
ϕ (N3–Fe1–N6)95.7(2)90.46(16)93.30(17)98.11(12)95.55(19)
ϕ (N4–Fe1–N5)76.4(2)79.34(16)74.76(15)74.30(12)79.4(2)
ϕ (N4–Fe1–N6)158.89(18)158.89(18)149.28(16)146.72(12)158.42(18)
ϕ (N5–Fe1–N6)79.78(18)79.78(18)74.58(15)72.55(12)79.10(19)
Table 4. Distortion parameters for [Fe(X-pybox)2](ClO4)2 (X = H, Cl, CH3O, CH3S).
Table 4. Distortion parameters for [Fe(X-pybox)2](ClO4)2 (X = H, Cl, CH3O, CH3S).
Parameters aH bClCH3OCH3S
T/K173400 c100400 c100100
Σ/°90.1130.390.8142.0158.896.6
Θ/°293417310455596326
α/°79.775.579.674.473.879.2
CShM (Oh)2.2244.3272.2835.0196.9192.484
ϕ/°179.14(11)177.6(2)177.03(16)176.50(15)156.74(12)176.6(2)
θ/°92.595.12(7)91.06(4)93.04(5)93.41(3)94.99(5)
dcalc. (N–N)/Å d2.8552.8542.8542.852
a For the definition, see the text. b The 173 K structural data are taken from Ref. [25]. c The χmT vs. T plot indicates the spin-crossover is not completed even at 400 K, so the 400 K data imply the “almost” high-spin (HS) data. d An averaged value from two distances in a metal-free ligand. CShM: continuous shape measure; Oh: octahedron.

Share and Cite

MDPI and ACS Style

Kimura, A.; Ishida, T. Pybox-Iron(II) Spin-Crossover Complexes with Substituent Effects from the 4-Position of the Pyridine Ring (Pybox = 2,6-Bis(oxazolin-2-yl)pyridine). Inorganics 2017, 5, 52. https://doi.org/10.3390/inorganics5030052

AMA Style

Kimura A, Ishida T. Pybox-Iron(II) Spin-Crossover Complexes with Substituent Effects from the 4-Position of the Pyridine Ring (Pybox = 2,6-Bis(oxazolin-2-yl)pyridine). Inorganics. 2017; 5(3):52. https://doi.org/10.3390/inorganics5030052

Chicago/Turabian Style

Kimura, Akifumi, and Takayuki Ishida. 2017. "Pybox-Iron(II) Spin-Crossover Complexes with Substituent Effects from the 4-Position of the Pyridine Ring (Pybox = 2,6-Bis(oxazolin-2-yl)pyridine)" Inorganics 5, no. 3: 52. https://doi.org/10.3390/inorganics5030052

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