ACS Publications. Most Trusted. Most Cited. Most Read
Widening the Window of Spin-Crossover Temperatures in Bis(formazanate)iron(II) Complexes via Steric and Noncovalent Interactions
My Activity
  • Open Access
Article

Widening the Window of Spin-Crossover Temperatures in Bis(formazanate)iron(II) Complexes via Steric and Noncovalent Interactions
Click to copy article linkArticle link copied!

  • Francesca Milocco
    Francesca Milocco
    Stratingh Institute for Chemistry, University of Groningen, Nijenborgh 4, 9747 AG Groningen, The Netherlands
  • Folkert de Vries
    Folkert de Vries
    Stratingh Institute for Chemistry, University of Groningen, Nijenborgh 4, 9747 AG Groningen, The Netherlands
  • Harmke S. Siebe
    Harmke S. Siebe
    Stratingh Institute for Chemistry, University of Groningen, Nijenborgh 4, 9747 AG Groningen, The Netherlands
  • Silène Engbers
    Silène Engbers
    Stratingh Institute for Chemistry, University of Groningen, Nijenborgh 4, 9747 AG Groningen, The Netherlands
  • Serhiy Demeshko
    Serhiy Demeshko
    Institut für Anorganische Chemie, Universität Göttingen, Tammannstraße 4, 37077 Göttingen, Germany
  • Franc Meyer
    Franc Meyer
    Institut für Anorganische Chemie, Universität Göttingen, Tammannstraße 4, 37077 Göttingen, Germany
    More by Franc Meyer
  • Edwin Otten*
    Edwin Otten
    Stratingh Institute for Chemistry, University of Groningen, Nijenborgh 4, 9747 AG Groningen, The Netherlands
    *Email [email protected]
    More by Edwin Otten
Open PDFSupporting Information (1)

Inorganic Chemistry

Cite this: Inorg. Chem. 2021, 60, 3, 2045–2055
Click to copy citationCitation copied!
https://doi.org/10.1021/acs.inorgchem.0c03593
Published January 19, 2021

Copyright © 2021 American Chemical Society. This publication is licensed under CC-BY-NC-ND.

Abstract

Click to copy section linkSection link copied!

Bis(formazanate)iron(II) complexes undergo a thermally induced S = 0 to S = 2 spin transition in solution. Here we present a study of how steric effects and π-stacking interactions between the triarylformazanate ligands affect the spin-crossover behavior, in addition to electronic substituent effects. Moreover, the effect of increasing the denticity of the formazanate ligands is explored by including additional OMe donors in the ligand (7). In total, six new compounds (27) have been synthesized and characterized, both in solution and in the solid state, via spectroscopic, magnetic, and structural analyses. The series spans a broad range of spin-crossover temperatures (T1/2) for the LS ⇌ HS equilibrium in solution, with the exception of compound 6 which remains high-spin (S = 2) down to 210 K. In the solid state, 6 was shown to exist in two distinct forms: a tetrahedral high-spin complex (6a, S = 2) and a rare square-planar structure with an intermediate-spin state (6b, S = 1). SQUID measurements, 57Fe Mössbauer spectroscopy, and differential scanning calorimetry indicate that in the solid state the square-planar form 6b undergoes an incomplete spin-change-coupled isomerization to tetrahedral 6a. The complex that contains additional OMe donors (7) results in a six-coordinate (NNO)2Fe coordination geometry, which shifts the spin-crossover to significantly higher temperatures (T1/2 = 444 K). The available experimental and computational data for 7 suggest that the Fe···OMe interaction is retained upon spin-crossover. Despite the difference in coordination environment, the weak OMe donors do not significantly alter the electronic structure or ligand-field splitting, and the occurrence of spin-crossover (similar to the compounds lacking the OMe groups) originates from a large degree of metal–ligand π-covalency.

Copyright © 2021 American Chemical Society

Synopsis

A series of Fe(II) complexes with formazanate ligands are reported, and ligand substituent effects on structure and spin-crossover properties are examined. These ligand modifications allow isolation of compounds with tetrahedral geometries in both low- and high-spin ground states as well as an intermediate-spin square-planar complex. Steric properties, π-stacking interactions, and additional donor substituents lead to a wide range of spin-crossover temperatures (T1/2) in this class of compounds.

Introduction

Click to copy section linkSection link copied!

The geometry of transition metal complexes is dependent on the electronic structure, (1) and it is often the case that the geometry preferred on steric grounds is overridden in favor of a different one by electronic effects. (2) In four-coordinate complexes two extreme geometries can be observed: the sterically favored tetrahedral and the electronically stabilized square-planar structure. While complexes with a d8 configuration have been thoroughly investigated, the balance between steric and electronic effects on the geometry of compounds with a lower d-electron count is not well established. In the case of first-row transition metals such as Fe(II), the electronic stabilization is typically small, and therefore these compounds tend to adopt a tetrahedral configuration. (1a,2,3) Therefore, to observe square-planar Fe(II) complexes, specific requirements are usually needed that result in intermediate-spin (S = 1) compounds: (1b) (i) macrocyclic ligands that enforce a planar geometry around the metal center (4) or (ii) strong field ligands, e.g., phosphines, that provide a greater ligand field stabilization energy compared to nitrogen and oxygen donors (in the case of mono- and bidentate ligands), often in combination with ortho-substituted aryl coligands. (1b,5) Exceptions to this where Fe(II) square-planar structures were observed have been sporadically reported. (6) Furthermore, while isomerization between tetrahedral and square-planar geometries is a well-established phenomenon for cobalt(II), (7) nickel(II), (7a) and copper(II), (7a,8) it is rare for iron(II). (6e)
In the simplistic terms of crystal field theory, the spin state of a complex in a certain geometry is determined by the orbital splitting (Δ) and the pairing energy (PE). (1a) When the values of these two parameters are comparable, various electronic configurations, differing in the spin state, may be accessible. This opens the possibility of switching between different spin states by using external stimuli (e.g., temperature, pressure, or light), leading to the phenomenon of spin-crossover (SCO). (9) While the major representatives in the category of spin-crossover compounds are six-coordinate Fe(II) complexes with nitrogen donor ligands, (9c,10) pioneering work on four-coordinate Fe(II) compounds has been conducted by the groups of Chirik, (11) Smith, (12) Peters, (13) and ours. (14) To illustrate the relationship between geometry and spin state in Fe(II) complexes, a comparison of the expected splitting of the d-orbital manifold in common coordination geometries is provided in Figure 1A–C. The role of ligand design in tuning the SCO properties, such as the spin-crossover temperature (T1/2), is well recognized. (15) However, predicting the effect of changes in steric/electronic properties of the ligand and spin-crossover energetics remains very challenging due to the small energy differences involved.

Figure 1

Figure 1. Common ligand field splitting diagrams for octahedral (A), square-planar (B), and tetrahedral (C) geometries and unusual ligand field splitting for the pseudo-tetrahedral geometries found in bis(formazanate)iron(II) complexes (D). (14)

Following our report of a four-coordinate Fe(II) spin-crossover complex with formazanate ligands, (14a) we recently established that spin-crossover is a general feature of this class of compounds. (14b) The stability of the low-spin (S = 0) state for these compounds is ascribed to an unusual splitting pattern of the d-orbitals in this geometry. Specifically, the formazanate ligands, which are good π-acceptor ligands, are engaged in π-backdonation with the metal, and this allows the formation of a highly covalent metal–ligand bond, stabilizing one of the d-orbitals (the antibonding dyz orbital that belongs to the t2 set in a conventional tetrahedral complex), which gives rise to an “inverted” ligand field with an approximate “two-over-three” splitting pattern (Figure 1D). We demonstrated that it is possible to tune the SCO properties of bis(formazanate)iron(II) complexes by substituent effects that are purely electronic in nature. (14b) In the present work, we extend these studies to include steric effects as well as π-stacking interactions between the triarylformazanate ligands. Included in this analysis are nonsymmetric ligands that have two different N–Ar substituents. In addition, we describe the effect of additional OMe donor groups in the ligand. The aim of this work is to obtain comprehensive insight into how the spin-crossover properties of this class of compounds may be modulated via modification of the ligand.

Results and Discussion

Click to copy section linkSection link copied!

The bis(formazanate) iron complexes 27 were synthesized following a procedure previously reported by us (14b) starting from the iron precursor Fe[N(SiMe3)2]2 as depicted in Scheme 1. Complex 1 has already been extensively studied in our previous work, (14a) and it is therefore included in the discussion as reference compound. Besides compounds 1 and 2, all the others feature nonsymmetric ligands that have two different N–Ar substituents. The effect of an electron-withdrawing perfluorinated ring (Ar = C6F5) is studied either in the C–Ar3 position (2 and 4), as the N–Ar1 group (5), or in both positions (6). At the same time the influence of the electron-donating, sterically demanding mesityl group (Ar5 = Mes) is investigated in the N–Ar position either alone (3) or in combination with the perfluorinated ring (4, 5, and 6). Furthermore, the ortho-anisyl group (Ar1 = o-An) is introduced in the N–Ar position in compound 7, increasing the coordination ability of the formazanate to a tridentate monoanionic ligand.

Scheme 1

Scheme 1. Synthesis of Compounds 17

Solid-State Characterization

While attempts to obtain single crystals suitable for X-ray diffraction were not successful for 2 and 5, the other compounds could be obtained in crystalline form. Single-crystal X-ray diffraction studies for complexes 3, 4, 6, and 7 allowed determination of their molecular structure, and pertinent metrical parameters are collected in Table 1. Overall, the structure of compound 3 is very similar to 1: it has relatively short Fe–N distances averaging to 1.831 Å and a flattened tetrahedral geometry around the Fe center (angle between the ligand coordination planes of 64.06(9)°), features that are indicative of a low-spin Fe(II) center. (14a) The steric pressure exerted by the N-Mes groups is evinced by the N(Mes)–Fe–N(Mes) angle of 109.19(6)°, which is noticeably larger than the N(Ph)–Fe–N(Ph) angle (100.68(6)°). The N-mesityl rings in 3 are engaged in off-center π-stacking interactions (Figure 2) both within the same molecule (interplanar angle of 2.77°; distance between Mes centroids and the least-squares plane of the other Mes ring of 3.200/3.229 Å) and between neighboring molecules (centroid-to-plane distance of 3.604/3.702 Å, Figure 2). Complex 4 shows similar intramolecular interactions between N-Mes groups (interplanar angle of 2.22°), but in this case the π-stacking does not extend to adjacent molecules (Figure S2). In contrast to 3, the Fe–N bonds are long (1.9610(12)–1.9946(11) Å), and the angle between the formazanate coordination planes is increased to 83.21(7)°, indicating that 4 is high-spin in the solid.

Figure 2

Figure 2. Crystal structure of compound 3 showing the π-stacking interactions between the mesityl rings. The Fe center, ligand backbone, and the mesityl rings are shown as 50% probability ellipsoids and the remaining atoms as wireframe; hydrogen atoms are removed for clarity.

Table 1. Selected Bond Lengths (Å) and Angles (deg) in Compounds 1, 3, 4, 6, and 7 at 100 K (Unless Stated otherwise)
 1a346ab6b7
Fe(1)–N(1)1.8278(15)1.8192(14)1.9946(11)2.030(2)1.9259(9)1.877(2)
Fe(1)–N(4)1.8207(15)1.8351(13)1.9610(12)1.9851(19)1.9461(9)1.883(2)
Fe(1)–N(5)1.8330(16)1.8242(13)1.9864(12)2.035(2) 1.874(2)
Fe(1)–N(8)1.8174(16)1.8449(13)1.9616(12)1.9966(19) 1.895(2)
Fe(1)–O(1)     2.1128(18)
Fe(1)–O(2)     2.1029(19)
∠(NFeN)/(NFeN)c60.97(10)64.06(9)83.21(7)89.31(12)0.00(0)81.67(14)
Fe out-of-planed0.0010.0180.1160.5820.7000.220
0.0460.1190.1160.580 0.224
a

Data taken from ref (14a).

b

Structure measured at 200 K.

c

Dihedral angle between the coordination planes defined by the N–Fe–N atoms.

d

Displacement of the Fe atom out of the plane defined by the four N atoms of each ligand backbone.

Compound 6, in which the ligands are highly asymmetric from an electronic point of view (N–C6F5 and N-Mes), was obtained in two distinctly different forms (6a/b) depending on the crystallization conditions. A batch of crystals was obtained from hot hexane (6a) and analyzed by single-crystal X-ray diffraction. It shows a distorted tetrahedral environment around Fe, with off-center π-stacking between the N-Mes rings (interplanar angle of 2.87°) (Figure 3). The Fe–N distances of 6a (1.9851(19)–2.030(2) Å) and the angle between the formazanate coordination planes (89.31(12)°) indicate a high-spin Fe center. Surprisingly, crystals obtained by diffusion of hexane into a THF solution of 6 show that under these conditions it crystallizes as a square-planar complex (6b, Figure 3).

Figure 3

Figure 3. Molecular structure of compounds 6a and 6b showing 50% probability ellipsoids; hydrogen atoms omitted for clarity. The inset for each shows the Fe(NNCNN)2 core of the structure with the N–Fe–N planes and the dihedral angle.

The square-planar geometry has the N–Ar groups in an anti-relationship, which allows off-center parallel intramolecular π-stacking interactions between the electron-rich Mes and the electron-deficient C6F5 groups (interplanar angle = 10.14°; distance = 3.259 Å, Figure 4). In addition, off-center parallel stacking between the C–C6F5 rings of neighboring molecules (centroid-to-plane distance of 3.222 Å) and a weaker intermolecular interaction between the N–Ar groups (interplanar angle of 10.14° and centroid-to-plane distance of 3.459 Å) are observed. (16) While the FeN4 fragment is planar (enforced by the crystallographic symmetry), the FeNNCNN six-membered chelate rings are puckered with the Fe center displaced out of the ligand plane. The Fe–N bonds in square-planar 6b (1.9259(9)–1.9461(9) Å) are shorter than those found in high-spin FeN4 complexes, such as tetrahedral 6a and in distorted-planar iron bis(amidinate) complexes reported by Hessen et al. (2.0528–2.0697 Å). (6a) The similarity of the metrical parameters in 6b to those in intermediate-spin Fe(II) porphyrins (e.g., 1.972(4) Å in Fe(TPP) (4a)) suggests that 6b also has an S = 1 ground state. To the best of our knowledge, this is the first example of an intermediate-spin Fe(II) complex with bidentate nitrogen donor ligands which adopts a square-planar geometry in the solid state. Although solution studies (vide infra) indicate that 6 is high-spin in toluene, the accessibility of a square-planar polymorph for 6 suggests that controlling the strength of π-stacking interactions is a viable approach to change the geometric preference and thus spin state in this class of compounds.

Figure 4

Figure 4. Crystal structure of compound 6b illustrating the π-stacking interactions between the aromatic rings, showing 50% probability ellipsoids. Parts of the molecule are shown as wireframe, and hydrogen atoms are removed for clarity.

57Fe Mössbauer spectroscopy was employed to elucidate the electronic structure of 6. A quadrupole doublet with isomer shift δ = 0.75 mm/s and quadrupole splitting |ΔEq| = 1.21 mm/s was observed for a batch of crystals for the tetrahedral complex 6a (Figure 5 B). In contrast, crystals of square-planar 6b have a lower isomer shift (δ = 0.54 mm/s) and a higher quadrupole splitting (|ΔEq| = 2.73 mm/s) (Figure 5A). The Mössbauer spectra of both batches differ significantly from low-spin (S = 0) bis(formazanate)iron compounds, which have isomer shifts (δ) around 0 mm/s and an |ΔEq| of ca. 2 mm/s. (14) The isomer shift of 6a is indicative of a high-spin state (S = 2) (17) and indeed is comparable to that in the high-spin bis(formazanate)iron complex Fe(PhNNCPhNNPh) (δ = 0.60 mm/s). (14b) On the other hand, the isomer shift for 6b is in agreement with an intermediate spin state (S = 1), similar to the one reported for Fe(TPP) (δ = 0.50 mm/s). (4a) A crude powder of a pristine sample of 6 (i.e., not purified by crystallization) shows a Mössbauer spectrum identical with that of 6b and remains unchanged between 7 and 300 K (Figure 5C and Figure S4a–c). The magnetic susceptibility measurement of the powder sample of 6 recorded on a SQUID magnetometer gave χMT ≈ 1.1 cm3 mol–1 K, supporting the assignment of an intermediate-spin state (Figure 6). The magnetic susceptibility in the solid state stays constant up to 390 K, and then it suddenly increases, approaching a value of 2.5 cm3 mol–1 K at 400 K, which is lower than the expected value for a high-spin state S = 2 but could be an indication of an incomplete spin transition. To further probe this, the sample used for the SQUID measurement was subsequently analyzed by Mössbauer spectroscopy (Figure 5D). After 6b was heated to 400 K, the major species (82%) has a quadrupole doublet with δ = 0.74 mm/s and |ΔEq| = 1.17 mm/s, which are in good agreement with the values obtained for 6a. Thus, this indicates that square-planar, intermediate-spin 6b switches at least partially to tetrahedral, high-spin 6a in the solid state. Differential scanning calorimetry analysis of a fresh powder sample of 6 shows an endothermic transition at 412 K with an onset temperature around 397 K (followed by subsequent decomposition) and corroborates a spin transition in the solid state at high temperature.

Figure 5

Figure 5. 57Fe Mössbauer spectra at 80 K in the solid state of 6b (A), 6a (B), and a powder sample of 6 before (C) and after (D) heating to 400 K for SQUID measurements. The red line in the spectrum of heated 6 represents the main species with 82% area, and the gray subspectra are unknown impurities.

Figure 6

Figure 6. Magnetic susceptibility data for a powder sample of 6 in the solid state (heating to 400 K and subsequent cooling). The solid black line shows the best fit curve for S = 1 with the parameters g = 2.10 and D = 11.2 cm–1 (100% IS). The dashed red line shows the spin-only value for an S = 2 system.

Lastly, compound 7 containing formazanate ligands with an additional OMe donor moiety was characterized. Single-crystal X-ray diffraction allowed determination of the molecular structure as shown in Figure 7. It shows a distorted octahedral geometry where both the formazanate moieties act as tridentate ligands. The Fe–N bond lengths, which average 1.882 Å, are shorter than those reported for an octahedral monoformazanate iron(II) cationic complex (Fe–N average of 1.974 Å), (18) which reflects the relatively poor donor ability of the OMe groups. Nevertheless, the Fe–O bonds in 7 are relatively short (2.1128(18) and 2.1029(19) Å) and in agreement with it having a low-spin ground state. These metrical parameters stand in marked contrast to those reported by Hannedouche for an iron complex with related β-diketiminate ligands, which interact with only one o-OMe group (Fe–O distance = 2.465 Å) to form a five-coordinate complex that has a high-spin ground state based on the metrical data. (19)
Table 2. 57Fe Mössbauer Parameters (δ = Isomer Shift in mm s–1; |ΔEq| = Quadrupole Splitting in mm s–1) for Compounds 1, 3, and 6a
 136a6b6b
δ0.030.050.750.540.55
Eq|2.051.991.212.732.72
a

Measured in the solid state at 80 K.

b

Powder sample of the crude product before crystallization.

Figure 7

Figure 7. Molecular structures of 7 showing 50% probability ellipsoids. One of the N–Ph rings is shown as wireframe, and hydrogen atoms are omitted for clarity.

Variable-Temperature NMR and UV/Vis Spectroscopy in Solution

We subsequently studied the spin-crossover behavior in solution by monitoring the spectral changes as a function of temperature. The NMR chemical shifts for all compounds are found to be temperature dependent but at low temperature do not follow the Curie behavior that is expected for a paramagnet: instead, the NMR resonances of all compounds except 6 converge into the diamagnetic range of the spectrum, suggestive of population of the S = 0 state. With 1 as reference, the changes induced by the different ligand substituents are discussed below. The enthalpy and entropy differences (ΔHS) that describe the LS ⇌ HS equilibrium as well as the spin-crossover temperature (T1/2) for the series of compounds are collected in Table 3, and a plot of the high-spin fraction as a function of temperature is shown in Figure 8.

Figure 8

Figure 8. Temperature dependence of the high-spin fraction (γHS) of compounds 15 and 7 in toluene-d8, including error bars for T1/2HS = 0.5). The liquid range for toluene is indicated with the color gradient at the temperature axis.

Table 3. Thermodynamic Parameters for the Equilibrium between S = 0 and S = 2 Spin States in Toluene-d8 Solution for Compounds 17
 1a2b34b567b
ΔH (kJ mol–1)22.2 ± 0.38.5 ± 0.426.3 ± 0.112.6 ± 1.019.0 ± 0.437.5 ± 1.6
ΔS (J mol–1 K–1)64 ± 145 ± 478 ± 167 ± 570 ± 185 ± 5
T1/2c (K)345 ± 7192 ± 18340 ± 2188 ± 21271 ± 8444 ± 34
a

Data reproduced from ref (14a).

b

Estimated from fitting a limited temperature range.

c

The uncertainty in T1/2 is obtained by using error propagation from ΔH and ΔS.

For compound 2, which has a symmetrical ligand with a highly electron-withdrawing C–Ar3 group, the variable-temperature 1H and 19F NMR spectra in toluene-d8 are indicative of an equilibrium between the high- and low-spin states. Although the former is predominant even at 207 K (the lowest temperature that could be reached inside the NMR probe), and fitting the temperature dependence of the chemical shifts thus is somewhat less accurate, it is clear from the data that the thermodynamic values that describe the spin equilibrium are much decreased in 2H = 8.5 ± 0.4 kJ mol–1, ΔS = 45 ± 4 J mol–1 K–1) compared to 1. This can be attributed to the decrease in σ-donor strength of the ligands, which results in a smaller ligand-field splitting and destabilization of the low-spin state.
The introduction of an electron-rich, sterically demanding mesityl ring as an N–Ar group in compound 3 resulted in larger differences between both spin states, with ΔH = 26.3 ± 0.1 kJ mol–1 and ΔS = 78 ± 1 J mol–1 K–1 from fitting the NMR data. The increase in these values stands in contrast to the expected effect of electron-donating substituents at that position, since the N–Ar groups predominantly influence metal–ligand π-bonding. (14b) However, it is clear from the crystallographic data of 3 (vide supra) that the N–Mes rings are engaged in noncovalent interactions (stacking), and we conclude that these attractive forces act to stabilize the more compact low-spin state.
The two effects discussed above were subsequently combined in compound 4. While the crystallographic data indicate that 4 is high spin in the solid state, the solution data clearly indicate that the S = 0 state is populated at low temperature. The combination of two opposing effects on the relative stability of the low-spin state results in thermodynamic parameters for the spin-state equilibrium in 4H = 12.6 ± 1.0 kJ mol–1; ΔS = 67 ± 5 J mol–1 K–1) that are intermediate between those of compounds 2 and 3.
Subsequently, we evaluated the influence of a highly electron-withdrawing N–C6F5 substituent that is present in compounds 5 and 6. Changing the N–Ph group in 3 to N–C6F5 in 5 results in a noticeable decrease in ΔH and ΔS to values of 19.0 ± 0.4 kJ mol–1 and 70 ± 1 J mol–1 K–1, respectively. In the absence of structural data for 5, we refrain from a detailed interpretation of these values. It is noted, however, that this result runs counter to the expectation that an electron-withdrawing N–Ar group leads to increased ΔHS due to stronger metal–ligand π-bonding.
For compound 6, which has an additional C6F5 substituent at the C–Ar3 position, the solution characterization data are indicative of a high-spin ground state also at low temperature, as shown by a magnetic moment of 4.9–5.1 μB across the temperature range studied (217–348 K). The spectral changes in the VT 1H NMR studies provide evidence for a spin equilibrium at the lowest temperatures (a departure from Curie behavior), which may indicate population of either the intermediate-spin (S = 1) state that is observed by crystallography or a low-spin (S = 0) complex similar to that present for the other compounds. However, the population of a different spin state is too small at these temperatures to allow an unambiguous interpretation of the changes that occur.
Finally, the spin-crossover properties of compound 7 were evaluated in solution. At room temperature, the 1H NMR spectrum of 7 shows resonances in the diamagnetic range, and the number of signals is indicative of C2v symmetry. While most peaks are sharp, those corresponding to the o-CH (N–Ph) and the OMe groups appear broadened, suggesting that also 7 may show a temperature-dependent equilibrium between a LS (S = 0) diamagnetic state and a HS (S = 2) paramagnetic state. Indeed, when the temperature was increased, the resonances of 7 broaden substantially and shift away from their diamagnetic values (Figure 9).

Figure 9

Figure 9. 1H NMR spectra of 7 recorded between 247 and 397 K (toluene-d8, 500 MHz).

The variable-temperature NMR data can be modeled with the equilibrium parameters ΔH = 37.5 ± 1.6 kJ mol–1 and ΔS = 85 ± 5 J mol–1 K–1. The increase in ΔH compared to the other compounds discussed above indicates that there is a substantial additional enthalpic penalty upon changing the spin state from singlet to quintet. A key question surrounding the spin-crossover in 7 is whether or not the Fe···OMe interaction is retained in solution; that is, does it involve a change in the coordination sphere around the Fe center, or does the ligand maintain the same coordination mode in both spin states? Several lines of experimental and computational evidence point toward retention of the tridentate NNO coordination mode of the ligand in both spin states, resulting in an octahedral geometry for 7 throughout. First, although 7 is predominantly low-spin at room temperature, its OMe resonance is somewhat broadened. This is likely because it is in close proximity to the paramagnetic center and is thus noticeably affected, also when the population of high-spin 7 is still very low. In addition, the spin-state equilibrium in 7 is characterized by a value of ΔS (85 ± 5 J mol–1 K–1) that is only marginally larger than that of the others; loss of the Fe···OMe interaction in the high-spin state is expected to lead to a much larger entropy change. Finally, we performed density functional theory calculations on 7 in both spin states, with and without the Fe···OMe interaction (7calc and 7′calc, respectively; see the Supporting Information for details). The results of geometry optimizations with a def2-TZVP basis set (20) using either pure (BP86) (21) or hybrid functionals (TPSSh, (22) B3LYP (23)) all indicate that structures with the Fe···OMe interaction are favored over those in which the OMe group points away from the metal center. The optimized geometries for 7calc in the low-spin state have short Fe–O bonds of 2.14–2.21 Å, which are elongated to 2.43–2.51 Å in the S = 2 minima; the shortest bonds are found for the TPSSh geometries and the longest for BP86. Although, as expected, there are large differences between these functionals for the computed energy differences between the different spin states, (24) it is important to note that the calculations indicate that coordination of the OMe groups is stabilizing in both spin states, regardless of the functional used (ΔGcalc > 23.7 kJ mol–1). Analysis of the frontier molecular orbitals of (low-spin) 7calc shows that the additional interaction with the weak OMe donor groups does not lead to a substantial change in ligand-field strength in comparison to a structure in which the OMe groups are rotated away from the metal center (7′calc; see Figures S54 and S55 for a comparison of the canonical DFT orbitals). In fact, the HOMO–LUMO energy gap at the BP86/def2-TZVP level is somewhat smaller in the structure with the Fe···OMe interaction (LS-7calc: 9227 cm–1) compared to without (LS-7′calc: 10466 cm–1). (25) Analysis of the intrinsic bonding orbitals (26) at the BP86/def2-TZVP geometry shows that while the change from a four- to six-coordinate environment does result in somewhat different localized Fe orbitals, in both cases there clearly is a substantial degree of metal–ligand π-covalency (Figure 10). A similar analysis using the B3LYP and TPSSh functionals at the corresponding minima for 7calc shows that the intrinsic bonding orbitals are qualitatively similar for TPSSh. On the other hand, the B3LYP results show less covalent Fe–N bonds, which is reflected by a lower N-contribution to the relevant IBOs and a smaller Wiberg bond index for the Fe–N bonds (Table S6).

Figure 10

Figure 10. Representation of intrinsic bonding orbitals at the BP86/def2-TZVP minima, both with (7calc; (A)) and without Fe–O interaction (7′calc; (B)).

To corroborate the NMR data, we subsequently performed variable-temperature UV–vis spectroscopic measurements on all compounds. Dilute solutions in toluene (ca. 10–5 M) were analyzed at temperatures down to 183 K. The fact that we could access lower temperatures in the UV–vis spectrometer was particularly helpful in the analysis of compounds 2 and 4, for which spin-crossover has a relatively low T1/2. The thermodynamic parameters obtained from the fitting of the UV–vis data are congruent with those found from the NMR analysis (see Table S4).
Although the UV–vis spectra of the compounds in this series often are equilibrium mixtures that contain both spin states, the data at the extremes of the temperature range represent predominantly low- or high-spin (at low or high temperature, respectively), and these were taken to extract the absorption maxima of the other spin state by scaled subtraction (see the Supporting Information for details). The only exception is compound 6, the UV–vis spectrum of which does not change appreciably with temperature, and 6 is predominantly found in the high-spin state (Figure 11A). The LS spectra for 25 show two intense bands in the visible range with absorption maxima between 375–445 and 515–575 nm which are assigned to ligand-based π–π* transitions. (14) In the HS state, the two bands are bathochromically shifted (around 390–510 and 580–630 nm, respectively), and the lowest energy band shows a significantly lower intensity (Figure S24).

Figure 11

Figure 11. UV/vis spectra in toluene for (A) compound 6 recorded between 183 and 293 K and (B) compound 7 recorded between 293 and 383 K.

The UV–vis spectrum of compound 7 is distinct from the others as it shows three absorption maxima in the LS state (λ = 459, 608, and 828 nm; see Figure 11B). The lowest-energy transition in 7 is much broader and occurs at significantly lower energy than in the other compounds. Thus, the presence of the additional OMe donor groups in 7 results in an additional low-lying excited state that is a distinguishing feature of this compound.

Conclusions

Click to copy section linkSection link copied!

In this work, we have extended the series of bis(formazanate)iron complexes to systems featuring nonsymmetric ligands with two different N–Ar substituents. We have demonstrated that the spin-crossover behavior of this class of compounds may be modulated via modification of the ligand using different strategies: electronic effects, steric effects, π-stacking interactions, and ligand denticity. The formazanate ligands reported in this work allowed crystallographic characterization of structures with different coordination geometries and spin states: pseudo-tetrahedral low-spin (3), tetrahedral high-spin (4, 6a), square-planar intermediate-spin (6b), and octahedral low-spin (7). Moreover, 6b is shown to thermally switch in the solid state to 6a, undergoing an incomplete spin-change-coupled square-planar–tetrahedral isomerization, which is rare for iron(II) compounds. The combination of sterics, π-stacking interactions, and electronic effects provides a plethora of tools that can be used to substantially affect spin-crossover behavior in this class of compounds. Overall, we were able to tune the system to obtain solution spin-crossover properties that range from very low T1/2 (∼190 K in 2 and 4) to well above room temperature (444 K in 7). Computational data suggest that the spin-crossover in the six-coordinate bis(formazanate)iron(II) complex (7) is of similar nature to that previously described for the four-coordinate derivatives (14) and originates from a large degree of covalency in the Fe–N bonds due to metal → ligand π-back-donation. Given the relevance of understanding and tuning spin-state-dependent reactivity, we anticipate that this study provides useful insight into ways to fine-tune the spin-state energetics in Fe(II) complexes.

Experimental Section

Click to copy section linkSection link copied!

General Considerations

All manipulations were performed under nitrogen or argon by using standard glovebox, Schlenk, and vacuum-line techniques. THF (Aldrich, anhydrous, 99.8%) was dried by percolation over columns of Al2O3 (Fluka); toluene, hexane, and pentane (Aldrich, anhydrous, 99.8%) were passed over columns of Al2O3 (Fluka), BASF R3-11-supported Cu oxygen scavenger, and molecular sieves (Aldrich, 4 Å). THF-d8 (Euriso-top) and Tol-d8 (Aldrich) were vacuum transferred from Na/K alloy and stored under nitrogen.
The compounds 2H, (27)3H, (28)4H, (29)5H, (28)6H, (30) and Fe[N(SiMe3)2]2 (31) were synthesized according to the literature procedures. Ligand 7H was prepared according to a slightly adapted version of a literature method (28,32) (see the Supporting Information for a detailed description). Sodium carbonate (Merck), tetrabutylammonium bromide (Sigma-Aldrich, 99%), o-anisidine (Sigma-Aldrich, >99%), hydrochloric acid (Boom B.V., 37–38%), and sodium nitrite (Sigma-Aldrich, 99%) were used as received.
NMR spectra were recorded on a Varian Mercury 400, Inova 500, or Bruker 600 MHz spectrometer. The 1H and 13C NMR spectra were referenced internally by using the residual solvent resonances and reported in ppm relative to TMS (0 ppm). The assignments of NMR resonances were aided by COSY, HMQC, HSQC, and HMBC experiments using standard pulse sequences.
Elemental analyses were performed by the analytical laboratory of the Institute of Inorganic Chemistry at the University of Göttingen using an Elementar Vario EL III instrument.

Synthesis of Fe[PhNNC(C6F5)NNPh]2 (2)

A dark-orange solution of L2H (516.5 mg, 1.32 mmol) in THF (20 mL) was added to a green solution of Fe[N(SiMe3)2]2 (247.9 mg, 0.66 mmol) in THF (10 mL). The reaction mixture was stirred overnight at room temperature, leading to a red-brick-colored solution. The volatiles were removed under a vacuum, and the product was extracted in THF. Slow diffusion of hexane into the THF solution at −30 °C resulted in 303.8 mg of dark-brown powder (0.36 mmol, 55%). 1H NMR (500 MHz, THF-d8, 25 °C): δ 24.61 (4H, Ph m-CH), −9.31 (2H, Ph p-CH), −13.40 (br, 4H, Ph o-CH) ppm. 19F NMR (470 MHz, THF-d8, −65 °C): δ – 111.79 (2F, C6F5, o-CF), −130.81 (1F, C6F5, p-CF), −157.34 ppm (2F, C6F5, m-CF). HMQC-NMR (125 MHz, THF-d8, +25 °C): δ 107.2 (Ph m-CH), 3.98 ppm (Ph p-CH). The signal of Ph o-CH in the HMQC spectrum was not visible due to line broadening.

Synthesis of Fe[PhNNC(p-Tol)NNMes]2 (3)

Fe[N(SiMe3)2]2 (0.71 g 1.88 mmol) was dissolved in THF (15 mL), and a solution of L3H (1.31 g, 3.66 mmol) in THF (15 mL) was added. The reaction mixture was stirred for 2 days at room temperature, leading to a dark-red solution. The solution was filtered, and the volatiles were removed under a vacuum. Recrystallization by slow diffusion of hexane into a THF solution gave 0.62 g of brown powder (0.83 mmol, 44% yield). 1H NMR (600 MHz, THF-d8, 25 °C): δ 12.48 (2H, Ph m-CH), 10.94 (2H, p-Tol m-CH), 10.56 (1H, Mes m-CHA), 9.76 (3H, p-Tol p-CH3), 8.66 (6H, Mes o-CH3), 8.45 (1H, Mes m-CHB), 3.96 (2H, p-Tol o-CH), 0.42 (1H, Ph p-CH), −0.70 (3H, Mes p-CH3), −1.44 ppm (2H, Ph o-CH). 1H NMR (500 MHz, THF-d8, −55 °C): δ 8.06 (2H, p-Tol o-CH), 7.37 (2H, p-Tol m-CH), 7.25 (2H, Ph m-CH), 7.11 (1H, Mes m-CHA), 7.05 (1H, Ph p-CH), 6.65 (1H, Mes m-CHB), 6.47 (2H, Ph o-CH), 2.58 (6H, p-Tol p-CH3 and Mes o-CH3A), 1.69 (3H, Mes o-CH3B), 0.41 ppm (3H, Mes p-CH3). 13C NMR (125 MHz, THF-d8, −55 °C): δ 144.5 (Ph ipso-C), 142.3 (Mes ipso-C), 141.4 (NCN), 137.8 (ipso-C), 135.6 (ipso-C), 134.2 (ipso-C), 133.9 (Mes m-C), 133.7 (Ph p-C), 133.3 (p-tol o-C), 133.2 (Ph o-C), 132.9 (p-tol m-C), 131.9 (Ph m-C), 24.7 (Mes o-CH3), 24.2 (p-tol p-CH3), 22.8 (Mes p-CH3), 22.0 ppm (Mes o-CH3). Anal. Calcd for C46H46N8Fe: C, 72.06; H, 6.05; N, 14,61. Found: C, 72.54; H 5.82; N, 14.12.

Synthesis of Fe[PhNNC(C6F5)NNMes]2 (4)

A dark-orange solution of L4H (478.3 mg, 1.11 mmol) in toluene (40 mL) was added to a green solution of Fe[N(SiMe3)2]2 (190.5 mg, 0.506 mmol) in toluene (10 mL). The reaction mixture was stirred for 2 days at room temperature, leading to a brown solution. The volatiles were removed under a vacuum, and the product was extracted into toluene. Slow diffusion of hexane into the toluene solution resulted in 189.6 mg of dark-brown crystals (0.206 mmol, 41%). 1H NMR (500 MHz, toluene-d8, 25 °C): δ 35.51 (3H), 30.98 (3H), 28.44 (2H Ph m-CH), 21.38 (1H), 12.61 (1H), −8.71 (3H), −19.26 (1H, Ph p-CH), −23.52 (2H Ph o-CH) ppm. 19F NMR (470 MHz, toluene-d8, 25 °C): δ −103.93 (2F, C6F5, o-CF), −125.22 (1F, C6F5, p-CF), −147.38 ppm (2F, C6F5, m-CF).

Synthesis of Fe[C6F5NNC(p-Tol)NNMes]2 (5)

A red-brick-colored solution of L5H (1.41 g, 3.16 mmol) in toluene (40 mL) was added to a green solution of Fe[N(SiMe3)2]2 (0.60 g, 1.58 mmol) in toluene (10 mL). The reaction mixture was stirred overnight at room temperature, leading to a brown solution. The solution was filtered, and the volatiles were removed under a vacuum; the obtained dark solid was quickly washed with cold hexane, giving 1.28 g (1.35 mmol, 85%) of crude product. Any attempt to recrystallize the product was unsuccessful. 1H NMR (400 MHz, C6D6, 25 °C): δ 24.23 (3H), 17.22 (2H), 15.52 (3H), 14.15 (2H), 12.84 (6H, Mes o-CH3), −6.20 ppm (2H). 19F NMR (375 MHz, C6D6, 25 °C): δ −91.63 (1F, p-CF), −163.25 ppm (2F, m-CF). The signal of C6F5o-CF was not visible due to line broadening.

Synthesis of Fe[C6F5NNC(C6F5)NNMes]2 (6)

A dark-orange solution of 6H (1.352 g, 2.587 mmol) in THF (40 mL) was added to a green solution of Fe[N(SiMe3)2]2 (0.489 g 1.299 mmol) in THF (20 mL). The reaction mixture was stirred for 5 h, leading to a brown solution. The volatiles were removed under a vacuum, and a brown solid was collected in 70% yield (1.009 g, 0.919 mmol). The solid was recrystallized from refluxing hexane which afforded crystals of 6a suitable for X-ray diffraction. Alternatively, diffusion of hexane into a THF solution afforded single crystals of 6b. 1H NMR (500 MHz, toluene-d8, 25 °C): δ 20.94 (3H, Mes p-CH3), 14.77 (2H, Mes m-CH), 11.22 ppm (6H, Mes o-CH3). 19F NMR (375 MHz, toluene-d8, 25 °C): δ −93.78 (1F, p-CF), −106.37 (2F, o-CF), −130.67 (1F, p-CF), −154.58 (2F, m-CF), −160.23 ppm (2F, m-CF). The signal of o-CF of N–C6F5 was not visible due to line broadening. Note: NMR spectra are identical for 6a and 6b. Anal. Calcd for C44H22N8F20Fe: C, 48.11; H, 2.02; N, 10,20. Found: C, 48.25; H 1.85; N, 10.04.

Synthesis of Fe[PhNNC(p-Tol)NN(o-An)]2·0.5(THF) (7)

A fuchsia THF (10 mL) solution of L7H (96.4 mg, 0.28 mmol) was added to a green solution of Fe[N(SiMe3)2]2 (52.7 mg, 0.14 mmol) in 5 mL of THF. The reaction mixture was stirred for 3 days at room temperature, leading to a brown solution that was filtered through a 0.2 μm syringe filter, and slow diffusion of hexane into the THF solution afforded 7 as dark needles in 70% yield (76.6 mg, 0.098 mmol). 1H NMR (400 MHz, C6D6, 25 °C): δ = 8.43 (d, J = 7.6 Hz, 1H, o-An δCH), 8.36 (d, J = 7.9 Hz, 2H, p-tolyl o-CH), 7.37 (d, J = 7.7 Hz, 2H, p-tolyl m-CH), 7.08 (t, J = 7.2 Hz, 1H, o-An γCH), 6.84 (d, J = 6.7 Hz, 2H, Ph m-CH), 6.75 (t, J = 7.5 Hz, 1H, o-An βCH), 6.64 (t, J = 7.3 Hz, 1H, Ph p-CH), 6.41 (m, 1H, Ph o-CH), 6.30 (d, J = 7.9 Hz, 1H, o-An αCH), 3.58 (m, 1H, THF),a 2.78 (s, 3H, o-An OCH3), 2.41 (s, 3H, p-tolyl CH3), 1.42 ppm (m, 1H, THF).a13C NMR (151 MHz, C6D6, 25 °C): δ = 169.7 (Ph ipso-C), 152.8 (NCN), 151.6 (o-An ipso-COCH3), 150.4 (o-An ipso-C), 137.0 (p-tolyl ipso-CCH3), 136.4 (p-tolyl ipso-C), 128.8 (p-tolyl m-CH), 127.4 (p-tolyl o-CH), 127.3 (Ph m-CH), 126.6 (Ph p-CH) 124.6 (Ph o-CH), 124.4 (o-An βCH), 123.2 (o-An γCH), 118.6 (o-An δCH), 112.0 (o-An αCH), 67.8 (THF),a 56.5 (o-An OCH3), 25.8 (THF),a 21.0 ppm (p-tolyl CH3). Anal. Calcd for C44H42N8O2.5Fe: C 67.87, H 5.44, N 14.39. Found: C 68.02, H 5.43, N 14.02.

X-ray Crystallography

Single crystals of compounds 3, 4, 6a, 6b, and 7 (directly obtained from the mother liquor) were mounted on top of a cryoloop and transferred into the cold nitrogen stream (100 K; 200 K for 6a) of a Bruker-AXS D8 Venture diffractometer. Data collection and reduction was done by using the Bruker software suite APEX3. (33) The final unit cell was obtained from the xyz centroids of 9772 (3), 9845 (4), 9892 (6a), 9813 (6b), and 9876 (7) reflections after integration. A multiscan absorption correction was applied for compounds 3, 4, 6a, and 6b based on the intensities of symmetry-related reflections measured at different angular settings (SADABS). For compound 7, a numerical absorption correction was applied after indexing of the crystal faces in APEX3. The structures were solved by direct methods using SHELXS, (34) and refinement of the structure was performed by using SHLELXL. (35) From the refinement of 4 it was clear that the hexane solvent molecule was disordered. A two-site disorder model was used to describe this. The site-occupancy factor for the major disorder component refined to 0.72. Several of the atoms in the disordered solvent molecule gave nonpositive definite displacement parameters when refined freely, and ultimately DFIX and ISOR instructions were applied. The structure of 6a was measured at 200 K because the data at 100 K indicated an (incomplete) phase transition (not further investigated). For all structures, the hydrogen atoms were generated by geometrical considerations, constrained to idealized geometries and allowed to ride on their carrier atoms with an isotropic displacement parameter related to the equivalent displacement parameter of their carrier atoms. Crystal data and details on data collection and refinement are presented in Table 4.
Table 4. Crystallographic Data for Compounds 3, 4, 6a, 6b, and 7
 346a6b7
chem formulaC46H46N8FeC50H46F10FeN8C44H22F20FeN8C44H22F20FeN8C46H46FeN8O3
Mr766.761004.801098.541098.54814.76
cryst systtriclinictriclinicmonoclinictriclinicmonoclinic
color, habitred, blockred, blockbrown, blockred, blockgreen, needle
size (mm)0.23 × 0.18 × 0.110.42 × 0.26 × 0.090.30 × 0.17 × 0.150.34 × 0.20 × 0.090.49 × 0.06 × 0.02
space groupP-1P-1P21/cP-1P21/c
a (Å)8.5063(5)12.3467(8)12.7118(6)7.2786(6)12.0477(5)
b (Å)11.3217(7)12.4167(8)20.5795(10)12.5743(10)24.0684(9)
c (Å)21.5733(13)15.4105(10)17.4336(7)12.6798(9)14.6588(5)
α (deg)93.547(2)86.354(2)90118.595(2)90
β (deg)94.372(2)88.205(2)95.991(2)99.168(3)109.140(2)
γ (deg)109.254(2)83.349(2)9094.529(3)90
V3)1947.3(2)2341.2(3)4535.8(4)989.67(13)4015.6(3)
Z22414
ρcalc, g cm–31.3081.4251.6091.8431.348
radiation [Å]Mo Kα 0.71073Mo Kα 0.71073Mo Kα 0.71073Mo Kα 0.71073Cu Kα 1.54178
μ(Mo Kα), mm–10.4320.4070.4580.525 
μ(Cu Kα), mm–1    3.433
F(000)808103621925481712
temp (K)100(2)100(2)200(2)100(2)100(2)
θ range (deg)2.76–27.163.04–27.922.88–26.382.88–27.943.68–65.14
data collected (h, k, l)–10:10; –13:14; –27:27–16:16; –16:16; –20:20–15:15; –25:25; –21:21–9:9; –16:16; –16:16–14:14; –28:28; –17:16
no. of rflns collected6024070899549184768532490
no. of indpndt collected853811211908947556743
observed reflns Fo ≥ 2.0σ(Fo)72389251634645575373
R(F) (%)3.463.364.322.554.61
wR(F2) (%)8.038.059.932.5512.01
GooF1.0451.0401.0307.071.037
weighting a, b0.0267, 1.52600.0311, 1.26770.0324, 3.46910.0372, 0.61250.0552, 2.9707
params refined504687664334527
min, max resid dens–0.427, 0.295–0.280, 0.375–0.260, 0.264–0.451, 0.379–0.346, 0.689

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.inorgchem.0c03593.

  • Full experimental and characterization data, computational details (PDF)

Accession Codes

CCDC 20361482036152 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Author
  • Authors
    • Francesca Milocco - Stratingh Institute for Chemistry, University of Groningen, Nijenborgh 4, 9747 AG Groningen, The Netherlands
    • Folkert de Vries - Stratingh Institute for Chemistry, University of Groningen, Nijenborgh 4, 9747 AG Groningen, The Netherlands
    • Harmke S. Siebe - Stratingh Institute for Chemistry, University of Groningen, Nijenborgh 4, 9747 AG Groningen, The Netherlands
    • Silène Engbers - Stratingh Institute for Chemistry, University of Groningen, Nijenborgh 4, 9747 AG Groningen, The Netherlands
    • Serhiy Demeshko - Institut für Anorganische Chemie, Universität Göttingen, Tammannstraße 4, 37077 Göttingen, Germany
    • Franc Meyer - Institut für Anorganische Chemie, Universität Göttingen, Tammannstraße 4, 37077 Göttingen, GermanyOrcidhttp://orcid.org/0000-0002-8613-7862
  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

Financial support from The Netherlands Organization for Scientific Research (NWO) is gratefully acknowledged (VIDI grant to E.O.). We thank the Center for Information Technology of the University of Groningen for their support and for providing access to the Peregrine high performance computing cluster as well as Prof. Wesley Browne and Dr. Johannes Klein (University of Groningen) for access to VT-UV/vis spectroscopy. S.D. and F.M. acknowledge support from the Universität Göttingen.

Additional Note

Click to copy section linkSection link copied!

a Crystals of 7 contain one THF per iron complex, but drying results in loss of part of the THF solvate molecules.

References

Click to copy section linkSection link copied!

This article references 35 other publications.

  1. 1
    (a) Poli, R. Open-Shell Organometallics as a Bridge between Werner-Type and Low-Valent Organometallic Complexes. The Effect of the Spin State on the Stability, Reactivity, and Structure. Chem. Rev. 1996, 96, 21352204,  DOI: 10.1021/cr9500343
    (b) Hawrelak, E. J.; Bernskoetter, W. H.; Lobkovsky, E.; Yee, G. T.; Bill, E.; Chirik, P. J. Square planar vs tetrahedral geometry in four coordinate iron(II) complexes. Inorg. Chem. 2005, 44, 31033111,  DOI: 10.1021/ic048202+
  2. 2
    Hartwig, J. F. Organotransition Metal Chemistry: From Bonding to Catalysis; University Science Books: Sausalito, CA, 2010.
  3. 3
    Cirera, J.; Ruiz, E.; Alvarez, S. Stereochemistry and spin state in four-coordinate transition metal compounds. Inorg. Chem. 2008, 47, 28712889,  DOI: 10.1021/ic702276k
  4. 4
    (a) Collman, J. P.; Hoard, J. L.; Kim, N.; Lang, G.; Reed, C. A. Synthesis, stereochemistry, and structure-related properties of alpha, beta, gamma, delta-tetraphenylporphinatoiron(II). J. Am. Chem. Soc. 1975, 97, 26762681,  DOI: 10.1021/ja00843a015
    (b) Kirner, J. F.; Dow, W.; Scheidt, W. R. Molecular stereochemistry of two intermediate-spin complexes. Iron(II) phthalocyanine and manganese(II) phthalocyanine. Inorg. Chem. 1976, 15, 16851690,  DOI: 10.1021/ic50161a042
    (c) Strauss, S. H.; Silver, M. E.; Ibers, J. A. Iron(II) octaethylchlorine: structure and ligand affinity comparison with its porphyrin and isobacteriochlorin homologs. J. Am. Chem. Soc. 1983, 105, 41084109,  DOI: 10.1021/ja00350a069
  5. 5
    Chatt, J.; Shaw, B. L. Alkyls and aryls of transition metals. Part IV. Cobalt(II) and iron(II) derivatives. J. Chem. Soc. 1961, 285,  DOI: 10.1039/jr9610000285
  6. 6
    (a) Nijhuis, C. A.; Jellema, E.; Sciarone, T. J. J.; Meetsma, A.; Budzelaar, P. H. M.; Hessen, B. First-Row Transition Metal Bis(amidinate) Complexes; Planar Four-Coordination of FeII Enforced by Sterically Demanding Aryl Substituents Eur. Eur. J. Inorg. Chem. 2005, 2005, 20892099,  DOI: 10.1002/ejic.200500094
    (b) Wurzenberger, X.; Piotrowski, H.; Klufers, P. A stable molecular entity derived from rare iron(II) minerals: the square-planar high-spin d6 Fe(II)O4 chromophore Angew. Angew. Chem., Int. Ed. 2011, 50, 49744978,  DOI: 10.1002/anie.201006898
    (c) Cantalupo, S. A.; Fiedler, S. R.; Shores, M. P.; Rheingold, A. L.; Doerrer, L. H. High-spin square-planar Co(II) and Fe(II) complexes and reasons for their electronic structure. Angew. Chem., Int. Ed. 2012, 51, 10001005,  DOI: 10.1002/anie.201106091
    (d) Pinkert, D.; Demeshko, S.; Schax, F.; Braun, B.; Meyer, F.; Limberg, C. A Dinuclear Molecular Iron(II) Silicate with Two High-Spin Square-Planar FeO4 Units. Angew. Chem., Int. Ed. 2013, 52, 51555158,  DOI: 10.1002/anie.201209650
    (e) Liu, Y.; Luo, L.; Xiao, J.; Wang, L.; Song, Y.; Qu, J.; Luo, Y.; Deng, L. Four-Coordinate Iron(II) Diaryl Compounds with Monodentate N-Heterocyclic Carbene Ligation: Synthesis, Characterization, and Their Tetrahedral-Square Planar Isomerization in Solution. Inorg. Chem. 2015, 54, 47524760,  DOI: 10.1021/acs.inorgchem.5b00138
  7. 7
    (a) Holm, R. H.; Chakravorty, A.; Theriot, L. J. The Synthesis, Structures, and Solution Equilibria of Bis(pyrrole-2-aldimino)metal(II) Complexes. Inorg. Chem. 1966, 5, 625635,  DOI: 10.1021/ic50038a028
    (b) Wolny, J. A.; Rudolf, M. F.; Ciunik, Z.; Gatner, K.; Wołowiec, S. Cobalt(II) triazene 1-oxide bis(chelates). A case of planar (low spin)–tetrahedral (high spin) isomerism. J. Chem. Soc., Dalton Trans. 1993, 16111622,  DOI: 10.1039/DT9930001611
    (c) Ingleson, M. J.; Pink, M.; Fan, H.; Caulton, K. G. Exploring the reactivity of four-coordinate PNPCoX with access to three-coordinate spin triplet PNPCo. Inorg. Chem. 2007, 46, 1032110334,  DOI: 10.1021/ic701171p
  8. 8
    Gaazo, J. Plasticity of the coordination sphere of copper(II) complexes, its manifestation and causes. Coord. Chem. Rev. 1976, 19, 253297,  DOI: 10.1016/S0010-8545(00)80317-3
  9. 9
    (a) Cambi, L.; Szegö, L. Über die magnetische Susceptibilität der komplexen Verbindungen. Ber. Dtsch. Chem. Ges. B 1931, 64, 25912598,  DOI: 10.1002/cber.19310641002
    (b) Kahn, O.; Martinez, C. J. Spin-Transition Polymers: From Molecular Materials Toward Memory Devices. Science 1998, 279, 4448,  DOI: 10.1126/science.279.5347.44
    (c) Gütlich, P.; Garcia, Y.; Goodwin, H. A. Spin crossover phenomena in Fe(ii) complexes. Chem. Soc. Rev. 2000, 29, 419427,  DOI: 10.1039/b003504l
    (d) Sato, O.; Tao, J.; Zhang, Y. Z. Control of magnetic properties through external stimuli. Angew. Chem., Int. Ed. 2007, 46, 21522187,  DOI: 10.1002/anie.200602205
    (e) Bousseksou, A.; Molnar, G.; Salmon, L.; Nicolazzi, W. Molecular spin crossover phenomenon: recent achievements and prospects. Chem. Soc. Rev. 2011, 40, 33133335,  DOI: 10.1039/c1cs15042a
  10. 10
    Halcrow, M. A. The spin-states and spin-transitions of mononuclear iron(II) complexes of nitrogen-donor ligands. Polyhedron 2007, 26, 35233576,  DOI: 10.1016/j.poly.2007.03.033
  11. 11
    Bowman, A. C.; Milsmann, C.; Bill, E.; Turner, Z. R.; Lobkovsky, E.; DeBeer, S.; Wieghardt, K.; Chirik, P. J. Synthesis and Electronic Structure Determination of N-Alkyl-Substituted Bis(imino)pyridine Iron Imides Exhibiting Spin Crossover Behavior. J. Am. Chem. Soc. 2011, 133, 1735317369,  DOI: 10.1021/ja205736m
  12. 12
    (a) Scepaniak, J. J.; Harris, T. D.; Vogel, C. S.; Sutter, J.; Meyer, K.; Smith, J. M. Spin Crossover in a Four-Coordinate Iron(II) Complex. J. Am. Chem. Soc. 2011, 133, 38243827,  DOI: 10.1021/ja2003473
    (b) Mathoniere, C.; Lin, H. J.; Siretanu, D.; Clerac, R.; Smith, J. M. Photoinduced single-molecule magnet properties in a four-coordinate iron(II) spin crossover complex. J. Am. Chem. Soc. 2013, 135, 1908319086,  DOI: 10.1021/ja410643s
    (c) Lin, H. J.; Siretanu, D.; Dickie, D. A.; Subedi, D.; Scepaniak, J. J.; Mitcov, D.; Clerac, R.; Smith, J. M. Steric and electronic control of the spin state in three-fold symmetric, four-coordinate iron(II) complexes. J. Am. Chem. Soc. 2014, 136, 1332613332,  DOI: 10.1021/ja506425a
  13. 13
    Creutz, S. E.; Peters, J. C. Spin-State Tuning at Pseudo-tetrahedral d6 Ions: Spin Crossover in [BP3]FeII–X Complexes. Inorg. Chem. 2016, 55, 38943906,  DOI: 10.1021/acs.inorgchem.6b00066
  14. 14
    (a) Travieso-Puente, R.; Broekman, J. O. P.; Chang, M.-C.; Demeshko, S.; Meyer, F.; Otten, E. Spin-Crossover in a Pseudo-tetrahedral Bis(formazanate) Iron Complex. J. Am. Chem. Soc. 2016, 138, 55035506,  DOI: 10.1021/jacs.6b01552
    (b) Milocco, F.; de Vries, F.; Bartels, I. M. A.; Havenith, R. W. A.; Cirera, J.; Demeshko, S.; Meyer, F.; Otten, E. Electronic Control of Spin-Crossover Properties in Four-Coordinate Bis(formazanate) Iron(II) Complexes. J. Am. Chem. Soc. 2020, 142, 2017020181,  DOI: 10.1021/jacs.0c10010
  15. 15
    (a) 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, 2653,  DOI: 10.1016/j.ccr.2016.10.006
    (b) Constable, E. C.; Baum, G.; Bill, E.; Dyson, R.; van Eldik, R.; Fenske, D.; Kaderli, S.; Morris, D.; Neubrand, A.; Neuburger, M.; Smith, D. R.; Wieghardt, K.; Zehnder, M.; Zuberbühler, A. D. Control of Iron(II) Spin States in 2,2′:6′,2″-Terpyridine Complexes through Ligand Substitution. Chem. - Eur. J. 1999, 5, 498508,  DOI: 10.1002/(SICI)1521-3765(19990201)5:2<498::AID-CHEM498>3.0.CO;2-V
  16. 16
    (a) Bacchi, S.; Benaglia, M.; Cozzi, F.; Demartin, F.; Filippini, G.; Gavezzotti, A. X-ray Diffraction and Theoretical Studies for the Quantitative Assessment of Intermolecular Arene–Perfluoroarene Stacking Interactions. Chem. - Eur. J. 2006, 12, 35383546,  DOI: 10.1002/chem.200501248
    (b) Salonen, L. M.; Ellermann, M.; Diederich, F. Aromatic rings in chemical and biological recognition: energetics and structures. Angew. Chem., Int. Ed. 2011, 50, 48084842,  DOI: 10.1002/anie.201007560
    (c) Wheeler, S. E. Local Nature of Substituent Effects in Stacking Interactions. J. Am. Chem. Soc. 2011, 133, 1026210274,  DOI: 10.1021/ja202932e
  17. 17
    Gütlich, P. Fifty Years of Mössbauer Spectroscopy in Solid State Research - Remarkable Achievements, Future Perspectives. Z. Anorg. Allg. Chem. 2012, 638, 1543,  DOI: 10.1002/zaac.201100416
  18. 18
    Milocco, F.; Demeshko, S.; Meyer, F.; Otten, E. Ferrate(II) complexes with redox-active formazanate ligands. Dalton Trans. 2018, 47, 88178823,  DOI: 10.1039/C8DT01597J
  19. 19
    Lepori, C.; Guillot, R.; Hannedouche, J. C1-symmetric β-Diketiminatoiron(II) Complexes for Hydroamination of Primary Alkenylamines. Adv. Synth. Catal. 2019, 361, 714719,  DOI: 10.1002/adsc.201801464
  20. 20
    Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 32973305,  DOI: 10.1039/b508541a
  21. 21
    Becke, A. D. Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A 1988, 38, 30983100,  DOI: 10.1103/PhysRevA.38.3098
  22. 22
    (a) Staroverov, V. N.; Scuseria, G. E.; Tao, J.; Perdew, J. P. Comparative assessment of a new nonempirical density functional: Molecules and hydrogen-bonded complexes. J. Chem. Phys. 2003, 119, 1212912137,  DOI: 10.1063/1.1626543
    (b) Tao, J.; Perdew, J. P.; Staroverov, V. N.; Scuseria, G. E. Climbing the Density Functional Ladder: Nonempirical Meta--Generalized Gradient Approximation Designed for Molecules and Solids. Phys. Rev. Lett. 2003, 91, 146401,  DOI: 10.1103/PhysRevLett.91.146401
  23. 23
    Becke, A. D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 56485652,  DOI: 10.1063/1.464913
  24. 24
    (a) Harvey, J. N. In Principles and Applications of Density Functional Theory in Inorganic Chemistry I; Springer: Berlin, 2004; pp 151184.
    (b) Swart, M.; Gruden, M. Spinning around in Transition-Metal Chemistry. Acc. Chem. Res. 2016, 49, 26902697,  DOI: 10.1021/acs.accounts.6b00271
  25. 25
    These orbital splitting energies are relatively small; spin-crossover complexes commonly have significantly larger values for Δ. See for example:Hauser, A. Ligand Field Theoretical Considerations. In Spin Crossover in Transition Metal Compounds I; Gütlich, P., Goodwin, H. A., Eds.; Springer Berlin Heidelberg: Berlin, Heidelberg, 2004, p 49.
  26. 26
    (a) Knizia, G. Intrinsic Atomic Orbitals: An Unbiased Bridge between Quantum Theory and Chemical Concepts. J. Chem. Theory Comput. 2013, 9, 48344843,  DOI: 10.1021/ct400687b
    (b) Knizia, G.; Klein, J. E. M. N. Electron Flow in Reaction Mechanisms—Revealed from First Principles. Angew. Chem., Int. Ed. 2015, 54, 55185522,  DOI: 10.1002/anie.201410637
  27. 27
    Kamphuis, A. J.; Milocco, F.; Koiter, L.; Pescarmona, P. P.; Otten, E. Highly Selective Single-Component Formazanate Ferrate(II) Catalysts for the Conversion of CO2 into Cyclic Carbonates. ChemSusChem 2019, 12, 36353641,  DOI: 10.1002/cssc.201900740
  28. 28
    Chang, M.-C.; Roewen, P.; Travieso-Puente, R.; Lutz, M.; Otten, E. Formazanate Ligands as Structurally Versatile, Redox-Active Analogues of β-Diketiminates in Zinc Chemistry. Inorg. Chem. 2015, 54, 379388,  DOI: 10.1021/ic5025873
  29. 29
    Chang, M. C.; Otten, E. Synthesis and ligand-based reduction chemistry of boron difluoride complexes with redox-active formazanate ligands. Chem. Commun. 2014, 50, 74317433,  DOI: 10.1039/C4CC03244F
  30. 30
    (a) Chang, M. C.; Chantzis, A.; Jacquemin, D.; Otten, E. Boron difluorides with formazanate ligands: redox-switchable fluorescent dyes with large stokes shifts. Dalton Trans. 2016, 45, 94779484,  DOI: 10.1039/C6DT01226D
    (b) Chang, M.-C. Formazanate as redox-active, structurally versatile ligand platform: Zinc and boron chemistry. PhD thesis, University of Groningen, 2016.
  31. 31
    Broere, D. L.; Coric, I.; Brosnahan, A.; Holland, P. L. Quantitation of the THF Content in Fe[N(SiMe3)2]2.xTHF. Inorg. Chem. 2017, 56, 31403143,  DOI: 10.1021/acs.inorgchem.7b00056
  32. 32
    Gilroy, J. B.; Ferguson, M. J.; McDonald, R.; Patrick, B. O.; Hicks, R. G. Formazans as β-diketiminate analogues. Structural characterization of boratatetrazines and their reduction to borataverdazyl radical anions. Chem. Commun. 2007, 126128,  DOI: 10.1039/B609365E
  33. 33
    Bruker. APEX3, SAINT and SADABS; Bruker AXS Inc.: Madison, WI, 2016.
  34. 34
    Sheldrick, G. A short history of SHELX. Acta Crystallogr., Sect. A 2008, 64, 112122,  DOI: 10.1107/S0108767307043930
  35. 35
    Sheldrick, G. Crystal structure refinement with SHELXL. Acta Crystallogr., Sect. C 2015, 71, 38,  DOI: 10.1107/S2053229614024218

Cited By

Click to copy section linkSection link copied!
Citation Statements
Explore this article's citation statements on scite.ai

This article is cited by 28 publications.

  1. Bijoy Dey, Jordi Cirera, Liliana P. Ferreira, Paulo N. Martinho, Vadapalli Chandrasekhar. Peripheral Ligand Modification to Tune Spin Crossover Behavior of Fe(III) Complexes. Crystal Growth & Design 2025, 25 (9) , 3065-3074. https://doi.org/10.1021/acs.cgd.5c00112
  2. Fangqing Gao, Binbin Wang, Yan Liu, Tengli Wang, Liang Wang, Li-Fei Zou, Shufang Xue, Yunnan Guo. Aggregation-Induced Spin-Crossover Switching in a Spin-Labile Iron(II) Complex. Inorganic Chemistry 2025, 64 (12) , 5904-5912. https://doi.org/10.1021/acs.inorgchem.4c04701
  3. Amelia M. Wheaton, Jill A. Chipman, Rebecca K. Walde, Heike Hofstetter, John F. Berry. Chemically Separable Co(II) Spin-State Isomers. Journal of the American Chemical Society 2024, 146 (39) , 26926-26935. https://doi.org/10.1021/jacs.4c08097
  4. Marco E. Reinhard, Baldeep K. Sidhu, Issiah B. Lozada, Natalia Powers-Riggs, Robert J. Ortiz, Hyeongtaek Lim, Rachel Nickel, Johan van Lierop, Roberto Alonso-Mori, Matthieu Chollet, Leland B. Gee, Patrick L. Kramer, Thomas Kroll, Sumana L. Raj, Tim B. van Driel, Amy A. Cordones, Dimosthenis Sokaras, David E. Herbert, Kelly J. Gaffney. Time-Resolved X-ray Emission Spectroscopy and Synthetic High-Spin Model Complexes Resolve Ambiguities in Excited-State Assignments of Transition-Metal Chromophores: A Case Study of Fe-Amido Complexes. Journal of the American Chemical Society 2024, 146 (26) , 17908-17916. https://doi.org/10.1021/jacs.4c02748
  5. Bijoy Dey, Jordi Cirera, Sakshi Mehta, Liliana P. Ferreira, Ramar Arumugam, Abhishake Mondal, Paulo N. Martinho, Vadapalli Chandrasekhar. Steric Effects on Spin States in a Series of Fe(III) Complexes. Crystal Growth & Design 2023, 23 (9) , 6668-6678. https://doi.org/10.1021/acs.cgd.3c00555
  6. Zoufeng Xu, Yu-Shien Sung, Elisa Tomat. Design of Tetrazolium Cations for the Release of Antiproliferative Formazan Chelators in Mammalian Cells. Journal of the American Chemical Society 2023, 145 (28) , 15197-15206. https://doi.org/10.1021/jacs.3c02033
  7. Le Shi, Jedrzej Kobylarczyk, Katarzyna Dziedzic-Kocurek, Jan J. Stanek, Barbara Sieklucka, Robert Podgajny. Site Selectivity for the Spin States and Spin Crossover in Undecanuclear Heterometallic Cyanido-Bridged Clusters. Inorganic Chemistry 2023, 62 (18) , 7032-7044. https://doi.org/10.1021/acs.inorgchem.3c00325
  8. Chenggang Jiang, Thomas S. Teets. Trimetallic Iridium–Nickel–Iridium Bis(formazanate) Assemblies. Inorganic Chemistry 2022, 61 (23) , 8788-8796. https://doi.org/10.1021/acs.inorgchem.2c00726
  9. Alexandra C. Brown, Niklas B. Thompson, Daniel L. M. Suess. Evidence for Low-Valent Electronic Configurations in Iron–Sulfur Clusters. Journal of the American Chemical Society 2022, 144 (20) , 9066-9073. https://doi.org/10.1021/jacs.2c01872
  10. Akram Ali, Saumitra Bhowmik, Suman K. Barman, Narottam Mukhopadhyay, Christine E. Glüer (nee Schiewer), Francesc Lloret, Franc Meyer, Rabindranath Mukherjee. Iron(III) Complexes of a Hexadentate Thioether-Appended 2-Aminophenol Ligand: Redox-Driven Spin State Switchover. Inorganic Chemistry 2022, 61 (13) , 5292-5308. https://doi.org/10.1021/acs.inorgchem.1c03992
  11. Yifeng Qin, Qiangui Zeng, Juan Zhu, Ziwen Liu, Guojie Zhang, Xiaohui Duan, Bo Wu. Synthesis, thermodynamics and catalytic properties of novel energetic coordination compounds based on nitrotriazolate-formazan ligand. Inorganic Chemistry Communications 2025, 179 , 114679. https://doi.org/10.1016/j.inoche.2025.114679
  12. Ivan Nemec, Radovan Herchel. The Role of Methyl Substitution in Spin Crossover of Fe(III) Complexes with Pentadentate Schiff Base Ligands. Inorganics 2025, 13 (2) , 57. https://doi.org/10.3390/inorganics13020057
  13. Anayely Cruz-Galván, Juan Olguín. Spin crossover iron(II) complexes in non-hexacoordinate geometries. Trends in Chemistry 2025, 7 (1) , 26-42. https://doi.org/10.1016/j.trechm.2024.11.001
  14. Kevin Tran, Patrick Sander, Maximilian Seydi Kilic, Jules Brehme, Ralf Sindelar, Franz Renz. Facile Approach for the Fabrication of Vapor Sensitive Spin Transition Composite Nanofibers. European Journal of Inorganic Chemistry 2024, 27 (33) https://doi.org/10.1002/ejic.202400363
  15. Oleksandr Markin, Iurii Gudyma, Kamel Boukheddaden. Role of harmonic and anharmonic interactions in controlling the cooperativity of spin-crossover molecular crystals. Physical Review B 2024, 110 (17) https://doi.org/10.1103/PhysRevB.110.174416
  16. Qian-Qian Jia, Xue-Jie Zhang, Liandong Zhu, Li-Zhi Huang. Fe(II) coordination transition regulates reductive dechlorination: The overlooked abiotic role of lactate. Water Research 2024, 254 , 121342. https://doi.org/10.1016/j.watres.2024.121342
  17. W. K. Damdoom, O. H. R. Al-Jeilawi. Synthesis, Characterization of Formazan Derivatives from Isoniazid and Study Their Antioxidant Activity and Molecular Docking. Russian Journal of Bioorganic Chemistry 2024, 50 (1) , 86-94. https://doi.org/10.1134/S1068162024010163
  18. Sunita Birara, Shalu Saini, Moumita Majumder, Prem Lama, Shree Prakash Tiwari, Ramesh K. Metre. Design and synthesis of a solution-processed redox-active bis(formazanate) zinc complex for resistive switching applications. Dalton Transactions 2023, 52 (48) , 18429-18441. https://doi.org/10.1039/D3DT02809G
  19. Maysaa A.M. Alhar. Preparation of Chalcone-Azo reagents and study of their chemical analysis, thermal studies, and biological effects against fungi. Bionatura 2023, 8 (CSS 4) , 1-13. https://doi.org/10.21931/RB/CSS/2023.08.04.99
  20. Xiu-e Jiang, Bo Wu, Bo Yang, Zai-chao Zhang, Ya-lin Yang, Hui-ying Du, Cong-ming Ma. Base-promoted decarboxylative gem-diazo-C-coupling: Synthesis, characterization and performance of nitrotriazolylformazans. Energetic Materials Frontiers 2023, 4 (2) , 68-76. https://doi.org/10.1016/j.enmf.2023.06.002
  21. Dmitriy S. Yambulatov, Julia K. Voronina, Alexander S. Goloveshkin, Roman D. Svetogorov, Sergey L. Veber, Nikolay N. Efimov, Anna K. Matyukhina, Stanislav A. Nikolaevskii, Igor L. Eremenko, Mikhail A. Kiskin. Change in the Electronic Structure of the Cobalt(II) Ion in a One-Dimensional Polymer with Flexible Linkers Induced by a Structural Phase Transition. International Journal of Molecular Sciences 2023, 24 (1) , 215. https://doi.org/10.3390/ijms24010215
  22. Sergey K. Dedushenko, Yury D. Perfiliev. On the correlation of the 57Fe Mӧssbauer isomer shift and some structural parameters of a substance. Hyperfine Interactions 2022, 243 (1) https://doi.org/10.1007/s10751-022-01795-1
  23. N. A. Protasenko, E. V. Baranov, I. A. Yakushev, A. S. Bogomyakov, V. K. Cherkasov. Cobalt(III) Bis-o-semiquinone Complexes with p-Tolyl-Substituted Formazan Ligands: Synthesis, Structure, and Magnetic Properties. Russian Journal of Coordination Chemistry 2022, 48 (12) , 819-827. https://doi.org/10.1134/S1070328422700129
  24. Maysaa A. M. Alhar. A Study of the Spectral, Chromatographic and Solubility Characteristics of New Analytical Reagents from Anil-Azo and Their Biological Effects on Bacteria. Biomedical and Pharmacology Journal 2022, 15 (2) , 1115-1126. https://doi.org/10.13005/bpj/2447
  25. Natalia A. Protasenko, Maxim V. Arsenyev, Evgeny V. Baranov, Alyona A. Starikova, Artem S. Bogomyakov, Vladimir K. Cherkasov. Heteroligand o‐Semiquinonato Cobalt Complexes of 3‐Cyano and 3‐Nitroformazans. European Journal of Inorganic Chemistry 2022, 2022 (17) https://doi.org/10.1002/ejic.202200152
  26. Bingzhi Liu, Tingyu Pan, Jiajun Liu, Li Feng, Yuning Chen, Huaili Zheng. Taping into the super power and magic appeal of ultrasound coupled with EDTA on degradation of 2,4,6-TCP by Fe0 based advanced oxidation processes. Chemosphere 2022, 288 , 132650. https://doi.org/10.1016/j.chemosphere.2021.132650
  27. Jireh Joy D. Sacramento, Therese Albert, Maxime Siegler, Pierre Moënne‐Loccoz, David P. Goldberg. An Iron(III) Superoxide Corrole from Iron(II) and Dioxygen. Angewandte Chemie 2022, 134 (2) https://doi.org/10.1002/ange.202111492
  28. Jireh Joy D. Sacramento, Therese Albert, Maxime Siegler, Pierre Moënne‐Loccoz, David P. Goldberg. An Iron(III) Superoxide Corrole from Iron(II) and Dioxygen. Angewandte Chemie International Edition 2022, 61 (2) https://doi.org/10.1002/anie.202111492

Inorganic Chemistry

Cite this: Inorg. Chem. 2021, 60, 3, 2045–2055
Click to copy citationCitation copied!
https://doi.org/10.1021/acs.inorgchem.0c03593
Published January 19, 2021

Copyright © 2021 American Chemical Society. This publication is licensed under CC-BY-NC-ND.

Article Views

3460

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. Common ligand field splitting diagrams for octahedral (A), square-planar (B), and tetrahedral (C) geometries and unusual ligand field splitting for the pseudo-tetrahedral geometries found in bis(formazanate)iron(II) complexes (D). (14)

    Scheme 1

    Scheme 1. Synthesis of Compounds 17

    Figure 2

    Figure 2. Crystal structure of compound 3 showing the π-stacking interactions between the mesityl rings. The Fe center, ligand backbone, and the mesityl rings are shown as 50% probability ellipsoids and the remaining atoms as wireframe; hydrogen atoms are removed for clarity.

    Figure 3

    Figure 3. Molecular structure of compounds 6a and 6b showing 50% probability ellipsoids; hydrogen atoms omitted for clarity. The inset for each shows the Fe(NNCNN)2 core of the structure with the N–Fe–N planes and the dihedral angle.

    Figure 4

    Figure 4. Crystal structure of compound 6b illustrating the π-stacking interactions between the aromatic rings, showing 50% probability ellipsoids. Parts of the molecule are shown as wireframe, and hydrogen atoms are removed for clarity.

    Figure 5

    Figure 5. 57Fe Mössbauer spectra at 80 K in the solid state of 6b (A), 6a (B), and a powder sample of 6 before (C) and after (D) heating to 400 K for SQUID measurements. The red line in the spectrum of heated 6 represents the main species with 82% area, and the gray subspectra are unknown impurities.

    Figure 6

    Figure 6. Magnetic susceptibility data for a powder sample of 6 in the solid state (heating to 400 K and subsequent cooling). The solid black line shows the best fit curve for S = 1 with the parameters g = 2.10 and D = 11.2 cm–1 (100% IS). The dashed red line shows the spin-only value for an S = 2 system.

    Figure 7

    Figure 7. Molecular structures of 7 showing 50% probability ellipsoids. One of the N–Ph rings is shown as wireframe, and hydrogen atoms are omitted for clarity.

    Figure 8

    Figure 8. Temperature dependence of the high-spin fraction (γHS) of compounds 15 and 7 in toluene-d8, including error bars for T1/2HS = 0.5). The liquid range for toluene is indicated with the color gradient at the temperature axis.

    Figure 9

    Figure 9. 1H NMR spectra of 7 recorded between 247 and 397 K (toluene-d8, 500 MHz).

    Figure 10

    Figure 10. Representation of intrinsic bonding orbitals at the BP86/def2-TZVP minima, both with (7calc; (A)) and without Fe–O interaction (7′calc; (B)).

    Figure 11

    Figure 11. UV/vis spectra in toluene for (A) compound 6 recorded between 183 and 293 K and (B) compound 7 recorded between 293 and 383 K.

  • References


    This article references 35 other publications.

    1. 1
      (a) Poli, R. Open-Shell Organometallics as a Bridge between Werner-Type and Low-Valent Organometallic Complexes. The Effect of the Spin State on the Stability, Reactivity, and Structure. Chem. Rev. 1996, 96, 21352204,  DOI: 10.1021/cr9500343
      (b) Hawrelak, E. J.; Bernskoetter, W. H.; Lobkovsky, E.; Yee, G. T.; Bill, E.; Chirik, P. J. Square planar vs tetrahedral geometry in four coordinate iron(II) complexes. Inorg. Chem. 2005, 44, 31033111,  DOI: 10.1021/ic048202+
    2. 2
      Hartwig, J. F. Organotransition Metal Chemistry: From Bonding to Catalysis; University Science Books: Sausalito, CA, 2010.
    3. 3
      Cirera, J.; Ruiz, E.; Alvarez, S. Stereochemistry and spin state in four-coordinate transition metal compounds. Inorg. Chem. 2008, 47, 28712889,  DOI: 10.1021/ic702276k
    4. 4
      (a) Collman, J. P.; Hoard, J. L.; Kim, N.; Lang, G.; Reed, C. A. Synthesis, stereochemistry, and structure-related properties of alpha, beta, gamma, delta-tetraphenylporphinatoiron(II). J. Am. Chem. Soc. 1975, 97, 26762681,  DOI: 10.1021/ja00843a015
      (b) Kirner, J. F.; Dow, W.; Scheidt, W. R. Molecular stereochemistry of two intermediate-spin complexes. Iron(II) phthalocyanine and manganese(II) phthalocyanine. Inorg. Chem. 1976, 15, 16851690,  DOI: 10.1021/ic50161a042
      (c) Strauss, S. H.; Silver, M. E.; Ibers, J. A. Iron(II) octaethylchlorine: structure and ligand affinity comparison with its porphyrin and isobacteriochlorin homologs. J. Am. Chem. Soc. 1983, 105, 41084109,  DOI: 10.1021/ja00350a069
    5. 5
      Chatt, J.; Shaw, B. L. Alkyls and aryls of transition metals. Part IV. Cobalt(II) and iron(II) derivatives. J. Chem. Soc. 1961, 285,  DOI: 10.1039/jr9610000285
    6. 6
      (a) Nijhuis, C. A.; Jellema, E.; Sciarone, T. J. J.; Meetsma, A.; Budzelaar, P. H. M.; Hessen, B. First-Row Transition Metal Bis(amidinate) Complexes; Planar Four-Coordination of FeII Enforced by Sterically Demanding Aryl Substituents Eur. Eur. J. Inorg. Chem. 2005, 2005, 20892099,  DOI: 10.1002/ejic.200500094
      (b) Wurzenberger, X.; Piotrowski, H.; Klufers, P. A stable molecular entity derived from rare iron(II) minerals: the square-planar high-spin d6 Fe(II)O4 chromophore Angew. Angew. Chem., Int. Ed. 2011, 50, 49744978,  DOI: 10.1002/anie.201006898
      (c) Cantalupo, S. A.; Fiedler, S. R.; Shores, M. P.; Rheingold, A. L.; Doerrer, L. H. High-spin square-planar Co(II) and Fe(II) complexes and reasons for their electronic structure. Angew. Chem., Int. Ed. 2012, 51, 10001005,  DOI: 10.1002/anie.201106091
      (d) Pinkert, D.; Demeshko, S.; Schax, F.; Braun, B.; Meyer, F.; Limberg, C. A Dinuclear Molecular Iron(II) Silicate with Two High-Spin Square-Planar FeO4 Units. Angew. Chem., Int. Ed. 2013, 52, 51555158,  DOI: 10.1002/anie.201209650
      (e) Liu, Y.; Luo, L.; Xiao, J.; Wang, L.; Song, Y.; Qu, J.; Luo, Y.; Deng, L. Four-Coordinate Iron(II) Diaryl Compounds with Monodentate N-Heterocyclic Carbene Ligation: Synthesis, Characterization, and Their Tetrahedral-Square Planar Isomerization in Solution. Inorg. Chem. 2015, 54, 47524760,  DOI: 10.1021/acs.inorgchem.5b00138
    7. 7
      (a) Holm, R. H.; Chakravorty, A.; Theriot, L. J. The Synthesis, Structures, and Solution Equilibria of Bis(pyrrole-2-aldimino)metal(II) Complexes. Inorg. Chem. 1966, 5, 625635,  DOI: 10.1021/ic50038a028
      (b) Wolny, J. A.; Rudolf, M. F.; Ciunik, Z.; Gatner, K.; Wołowiec, S. Cobalt(II) triazene 1-oxide bis(chelates). A case of planar (low spin)–tetrahedral (high spin) isomerism. J. Chem. Soc., Dalton Trans. 1993, 16111622,  DOI: 10.1039/DT9930001611
      (c) Ingleson, M. J.; Pink, M.; Fan, H.; Caulton, K. G. Exploring the reactivity of four-coordinate PNPCoX with access to three-coordinate spin triplet PNPCo. Inorg. Chem. 2007, 46, 1032110334,  DOI: 10.1021/ic701171p
    8. 8
      Gaazo, J. Plasticity of the coordination sphere of copper(II) complexes, its manifestation and causes. Coord. Chem. Rev. 1976, 19, 253297,  DOI: 10.1016/S0010-8545(00)80317-3
    9. 9
      (a) Cambi, L.; Szegö, L. Über die magnetische Susceptibilität der komplexen Verbindungen. Ber. Dtsch. Chem. Ges. B 1931, 64, 25912598,  DOI: 10.1002/cber.19310641002
      (b) Kahn, O.; Martinez, C. J. Spin-Transition Polymers: From Molecular Materials Toward Memory Devices. Science 1998, 279, 4448,  DOI: 10.1126/science.279.5347.44
      (c) Gütlich, P.; Garcia, Y.; Goodwin, H. A. Spin crossover phenomena in Fe(ii) complexes. Chem. Soc. Rev. 2000, 29, 419427,  DOI: 10.1039/b003504l
      (d) Sato, O.; Tao, J.; Zhang, Y. Z. Control of magnetic properties through external stimuli. Angew. Chem., Int. Ed. 2007, 46, 21522187,  DOI: 10.1002/anie.200602205
      (e) Bousseksou, A.; Molnar, G.; Salmon, L.; Nicolazzi, W. Molecular spin crossover phenomenon: recent achievements and prospects. Chem. Soc. Rev. 2011, 40, 33133335,  DOI: 10.1039/c1cs15042a
    10. 10
      Halcrow, M. A. The spin-states and spin-transitions of mononuclear iron(II) complexes of nitrogen-donor ligands. Polyhedron 2007, 26, 35233576,  DOI: 10.1016/j.poly.2007.03.033
    11. 11
      Bowman, A. C.; Milsmann, C.; Bill, E.; Turner, Z. R.; Lobkovsky, E.; DeBeer, S.; Wieghardt, K.; Chirik, P. J. Synthesis and Electronic Structure Determination of N-Alkyl-Substituted Bis(imino)pyridine Iron Imides Exhibiting Spin Crossover Behavior. J. Am. Chem. Soc. 2011, 133, 1735317369,  DOI: 10.1021/ja205736m
    12. 12
      (a) Scepaniak, J. J.; Harris, T. D.; Vogel, C. S.; Sutter, J.; Meyer, K.; Smith, J. M. Spin Crossover in a Four-Coordinate Iron(II) Complex. J. Am. Chem. Soc. 2011, 133, 38243827,  DOI: 10.1021/ja2003473
      (b) Mathoniere, C.; Lin, H. J.; Siretanu, D.; Clerac, R.; Smith, J. M. Photoinduced single-molecule magnet properties in a four-coordinate iron(II) spin crossover complex. J. Am. Chem. Soc. 2013, 135, 1908319086,  DOI: 10.1021/ja410643s
      (c) Lin, H. J.; Siretanu, D.; Dickie, D. A.; Subedi, D.; Scepaniak, J. J.; Mitcov, D.; Clerac, R.; Smith, J. M. Steric and electronic control of the spin state in three-fold symmetric, four-coordinate iron(II) complexes. J. Am. Chem. Soc. 2014, 136, 1332613332,  DOI: 10.1021/ja506425a
    13. 13
      Creutz, S. E.; Peters, J. C. Spin-State Tuning at Pseudo-tetrahedral d6 Ions: Spin Crossover in [BP3]FeII–X Complexes. Inorg. Chem. 2016, 55, 38943906,  DOI: 10.1021/acs.inorgchem.6b00066
    14. 14
      (a) Travieso-Puente, R.; Broekman, J. O. P.; Chang, M.-C.; Demeshko, S.; Meyer, F.; Otten, E. Spin-Crossover in a Pseudo-tetrahedral Bis(formazanate) Iron Complex. J. Am. Chem. Soc. 2016, 138, 55035506,  DOI: 10.1021/jacs.6b01552
      (b) Milocco, F.; de Vries, F.; Bartels, I. M. A.; Havenith, R. W. A.; Cirera, J.; Demeshko, S.; Meyer, F.; Otten, E. Electronic Control of Spin-Crossover Properties in Four-Coordinate Bis(formazanate) Iron(II) Complexes. J. Am. Chem. Soc. 2020, 142, 2017020181,  DOI: 10.1021/jacs.0c10010
    15. 15
      (a) 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, 2653,  DOI: 10.1016/j.ccr.2016.10.006
      (b) Constable, E. C.; Baum, G.; Bill, E.; Dyson, R.; van Eldik, R.; Fenske, D.; Kaderli, S.; Morris, D.; Neubrand, A.; Neuburger, M.; Smith, D. R.; Wieghardt, K.; Zehnder, M.; Zuberbühler, A. D. Control of Iron(II) Spin States in 2,2′:6′,2″-Terpyridine Complexes through Ligand Substitution. Chem. - Eur. J. 1999, 5, 498508,  DOI: 10.1002/(SICI)1521-3765(19990201)5:2<498::AID-CHEM498>3.0.CO;2-V
    16. 16
      (a) Bacchi, S.; Benaglia, M.; Cozzi, F.; Demartin, F.; Filippini, G.; Gavezzotti, A. X-ray Diffraction and Theoretical Studies for the Quantitative Assessment of Intermolecular Arene–Perfluoroarene Stacking Interactions. Chem. - Eur. J. 2006, 12, 35383546,  DOI: 10.1002/chem.200501248
      (b) Salonen, L. M.; Ellermann, M.; Diederich, F. Aromatic rings in chemical and biological recognition: energetics and structures. Angew. Chem., Int. Ed. 2011, 50, 48084842,  DOI: 10.1002/anie.201007560
      (c) Wheeler, S. E. Local Nature of Substituent Effects in Stacking Interactions. J. Am. Chem. Soc. 2011, 133, 1026210274,  DOI: 10.1021/ja202932e
    17. 17
      Gütlich, P. Fifty Years of Mössbauer Spectroscopy in Solid State Research - Remarkable Achievements, Future Perspectives. Z. Anorg. Allg. Chem. 2012, 638, 1543,  DOI: 10.1002/zaac.201100416
    18. 18
      Milocco, F.; Demeshko, S.; Meyer, F.; Otten, E. Ferrate(II) complexes with redox-active formazanate ligands. Dalton Trans. 2018, 47, 88178823,  DOI: 10.1039/C8DT01597J
    19. 19
      Lepori, C.; Guillot, R.; Hannedouche, J. C1-symmetric β-Diketiminatoiron(II) Complexes for Hydroamination of Primary Alkenylamines. Adv. Synth. Catal. 2019, 361, 714719,  DOI: 10.1002/adsc.201801464
    20. 20
      Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 32973305,  DOI: 10.1039/b508541a
    21. 21
      Becke, A. D. Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A 1988, 38, 30983100,  DOI: 10.1103/PhysRevA.38.3098
    22. 22
      (a) Staroverov, V. N.; Scuseria, G. E.; Tao, J.; Perdew, J. P. Comparative assessment of a new nonempirical density functional: Molecules and hydrogen-bonded complexes. J. Chem. Phys. 2003, 119, 1212912137,  DOI: 10.1063/1.1626543
      (b) Tao, J.; Perdew, J. P.; Staroverov, V. N.; Scuseria, G. E. Climbing the Density Functional Ladder: Nonempirical Meta--Generalized Gradient Approximation Designed for Molecules and Solids. Phys. Rev. Lett. 2003, 91, 146401,  DOI: 10.1103/PhysRevLett.91.146401
    23. 23
      Becke, A. D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 56485652,  DOI: 10.1063/1.464913
    24. 24
      (a) Harvey, J. N. In Principles and Applications of Density Functional Theory in Inorganic Chemistry I; Springer: Berlin, 2004; pp 151184.
      (b) Swart, M.; Gruden, M. Spinning around in Transition-Metal Chemistry. Acc. Chem. Res. 2016, 49, 26902697,  DOI: 10.1021/acs.accounts.6b00271
    25. 25
      These orbital splitting energies are relatively small; spin-crossover complexes commonly have significantly larger values for Δ. See for example:Hauser, A. Ligand Field Theoretical Considerations. In Spin Crossover in Transition Metal Compounds I; Gütlich, P., Goodwin, H. A., Eds.; Springer Berlin Heidelberg: Berlin, Heidelberg, 2004, p 49.
    26. 26
      (a) Knizia, G. Intrinsic Atomic Orbitals: An Unbiased Bridge between Quantum Theory and Chemical Concepts. J. Chem. Theory Comput. 2013, 9, 48344843,  DOI: 10.1021/ct400687b
      (b) Knizia, G.; Klein, J. E. M. N. Electron Flow in Reaction Mechanisms—Revealed from First Principles. Angew. Chem., Int. Ed. 2015, 54, 55185522,  DOI: 10.1002/anie.201410637
    27. 27
      Kamphuis, A. J.; Milocco, F.; Koiter, L.; Pescarmona, P. P.; Otten, E. Highly Selective Single-Component Formazanate Ferrate(II) Catalysts for the Conversion of CO2 into Cyclic Carbonates. ChemSusChem 2019, 12, 36353641,  DOI: 10.1002/cssc.201900740
    28. 28
      Chang, M.-C.; Roewen, P.; Travieso-Puente, R.; Lutz, M.; Otten, E. Formazanate Ligands as Structurally Versatile, Redox-Active Analogues of β-Diketiminates in Zinc Chemistry. Inorg. Chem. 2015, 54, 379388,  DOI: 10.1021/ic5025873
    29. 29
      Chang, M. C.; Otten, E. Synthesis and ligand-based reduction chemistry of boron difluoride complexes with redox-active formazanate ligands. Chem. Commun. 2014, 50, 74317433,  DOI: 10.1039/C4CC03244F
    30. 30
      (a) Chang, M. C.; Chantzis, A.; Jacquemin, D.; Otten, E. Boron difluorides with formazanate ligands: redox-switchable fluorescent dyes with large stokes shifts. Dalton Trans. 2016, 45, 94779484,  DOI: 10.1039/C6DT01226D
      (b) Chang, M.-C. Formazanate as redox-active, structurally versatile ligand platform: Zinc and boron chemistry. PhD thesis, University of Groningen, 2016.
    31. 31
      Broere, D. L.; Coric, I.; Brosnahan, A.; Holland, P. L. Quantitation of the THF Content in Fe[N(SiMe3)2]2.xTHF. Inorg. Chem. 2017, 56, 31403143,  DOI: 10.1021/acs.inorgchem.7b00056
    32. 32
      Gilroy, J. B.; Ferguson, M. J.; McDonald, R.; Patrick, B. O.; Hicks, R. G. Formazans as β-diketiminate analogues. Structural characterization of boratatetrazines and their reduction to borataverdazyl radical anions. Chem. Commun. 2007, 126128,  DOI: 10.1039/B609365E
    33. 33
      Bruker. APEX3, SAINT and SADABS; Bruker AXS Inc.: Madison, WI, 2016.
    34. 34
      Sheldrick, G. A short history of SHELX. Acta Crystallogr., Sect. A 2008, 64, 112122,  DOI: 10.1107/S0108767307043930
    35. 35
      Sheldrick, G. Crystal structure refinement with SHELXL. Acta Crystallogr., Sect. C 2015, 71, 38,  DOI: 10.1107/S2053229614024218
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.inorgchem.0c03593.

    • Full experimental and characterization data, computational details (PDF)

    Accession Codes

    CCDC 20361482036152 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.