Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Symmetric and asymmetric ligands for FeIII spin crossover – the influence of the C2 axis

Conor T. Kelly*a, Emmelyne Cuzaa, Eoin Pasquettia, Niall Quinna, Michael Griffina, Peter Nockemannb, Helge Müller-Bunza, Julia Bruno-Colmenareza, Solveig Feltonc, Zoi G. Ladad and Grace G. Morgan*a
aSchool of Chemistry, University College Dublin, Belfield, Dublin 4, Ireland. E-mail: conor.kelly3@ucd.ie; grace.morgan@ucd.ie
bSchool of Chemistry and Chemical Engineering, The QUILL Research Centre, Queen's University Belfast, David Keir Building, Stranmillis Rd, Belfast, BT9 5AG, UK
cCentre for Quantum Materials and Technologies (CQMT), School of Mathematics and Physics, Queen's University Belfast, Belfast, BT7 1NN, UK
dDepartment of Chemistry, University of Patras, 26504 Patras, Greece

Received 31st July 2025 , Accepted 4th September 2025

First published on 10th September 2025


Abstract

Modulation of the local strain and geometry in FeIII Schiff base complexes has been shown to allow the stabilisation of both the high spin (HS) and low spin (LS) states, along with thermal spin crossover (SCO). Complexes with hexadentate Schiff base ligands can be readily modified by changing the length of the tetraamine backbone linker. We report here 34 complexes of the symmetric [FeIII(R-sal2232)]+ and asymmetric [FeIII(R-sal2223)]+ families, where the former typically support the HS state, along with a handful of SCO examples, and the latter only supports the HS state. Magnetic measurements reveal that one symmetric example, [FeIII(5-I-sal2232)]ClO4 1.5, undergoes thermal SCO close to room temperature. We compare the structural distortion and spectroscopic properties in these examples, to indentify the factors that influence spin state choice. This reveals the importance of molecular symmetry, by way of a C2 axis bisecting the complex which is present in the samples which stabilise the LS state so far. The aforementioned example and three others, one reported previously, have short metal–ligand bond lengths suggesting adoption of the LS state coupled with the presence of a C2 axis. The additional strain in the [FeIII(R-sal2223)]+ complexes due to the asymmetric nature of the backbone results in significantly greater distortion around the FeIII centre which inhibits the stabilisation of the less distorted LS state. Computational analysis of the [FeIII(5-I-sal2232)]+ and [FeIII(5-I-sal2223)]+ isomers reveals that the HS state is more stable in the asymmetric [FeIII(5-I-sal2223)]+ species, whereas the energy difference between the HS and LS state for the [FeIII(5-I-sal2232)]+ cation is sufficiently small to allow for SCO to occur.


Introduction

Spin state selection in transition metal ions drives many aspects of their reactivity and function including in redox cycles in biology1–4 and catalysis,2–7 and in spin crossover systems.8–10 Spin states are affected by many factors including the local ligand field strength, coordination constraints and lattice pressure. In FeIII the spin state choices are S = 1/2, S = 3/2 and S = 5/2, with thermal switching typically observed between the S = 1/2 and S = 5/2 states.11–13 The potential for thermal spin state switching in FeIII hexadentate Schiff base chelates is well known, with those of the [FeIII(R-sal2222)]+ form first realised,14–16 while the related [FeIII(R-sal2323)]+ series is known to favour the S = 1/2 state,17–22 where the numbers refer to the lengths of the alkylene chains between each of the nitrogens of the tetramine backbone (e.g., 232 refers to an ethylene, propylene, ethylene linkage). These ligand types can be easily modulated to manipulate the local coordination strain while maintaining a consistent donor set, albeit in different positions. The FeIII complexes shown in Fig. 1 have been studied to date, with the exception of those with the ‘223’ backbone, [FeIII(R-sal2223)]+. While FeIII complexes with the ‘222’ and ‘323’ backbones favour SCO and LS respectively, the remaining ‘232’,21,23,24 ‘223’, and ‘333’21,25,26 analogues are much less developed.
image file: d5dt01833a-f1.tif
Fig. 1 Molecular structures of FeIII complexes with hexadentate Schiff base ligands with an N4O22− coordination sphere. Relevant nomenclature is indicated beneath each molecule, whereby the numbers refer to the lengths of the alkylene chains between each of the nitrogens of the tetraamine backbone (e.g., 232 refers to an ethylene, propylene, ethylene linkage). The two ligand types investigated in this work are highlighted in green, with those with the asymmetric 223 ligand, dashed line, reported for the first time in this work.

We have previously reported the ‘232’ complex, [FeIII(4,6-diOMe-sal2232)]ClO4,24 which showed an incomplete two-step thermal SCO with complex structural symmetry-breaking phase transitions and associated changes in local point group symmetry. We now extend our structural investigation of FeIII complexes with the ‘232’ series and compare these with the spin state outcomes with the isomeric ‘223’ series in selected examples, reporting 34 new examples. While all the ‘223’ complexes have structural parameters (bond lengths and distortion parameters) consistent with a HS assignment, the ‘232’ series can stabilise either HS or LS arrangements. We examine the relationship between spin state configuration and ligand in both series using X-ray diffraction, spectroscopic data, and quantum chemistry calculations.

Synthetic procedures

Symmetric complexes 1.1–1.14 were synthesised by a facile one pot Schiff base condensation of a salicylaldehyde and N,N’-bis(2-aminoethyl)-1,3-propanediamine followed by complexation with either an FeII (undergoes aerial oxidation to FeIII) or FeIII salt. Asymmetric complexes 2.1–2.11 were synthesised starting from N-(2-aminoethyl)-N’-(3-aminopropyl)-ethylenediamine tetrahydrochloride, which was neutralised with triethylamine, followed by the Schiff base condensation with a salicylaldehyde. This was followed by complexation with either an FeII or FeIII salt. Further synthetic details, including synthetic scheme can be found in section S1 of the SI.

Results and discussion

Single crystal X-ray diffraction (SCXRD)

Structural characterisation of the symmetric [FeIII(R-sal2232)]+ family of complexes. Structural determination of the 1.1–1.14 complexes, reveals an N4O22− coordination sphere around the FeIII centre, with cis-phenolate donors, cis-amine donors and trans-imine donors, similar to that of the earlier [FeIII(R-sal2222)]+ complexes.14–16 We can utilise the metal–ligand bond lengths to infer the spin state of the FeIII centre, by comparison of values between these samples and similar structures in the literature.21,23,24 The bond lengths are longer in the HS state, due to lengthening caused by population of the antibonding orbitals. Bond lengths for a selected number of complexes are presented in Table 3, and for all complexes in Table S2.2. Comparison of the bond lengths reveal that in most cases the bond lengths are typically of the order of: Fe–O = 1.91–1.93 Å, Fe–Nimine = 2.10–2.12 Å, and Fe–Namine = 2.16–2.24 Å. These lengths are comparable to that of the previously published HS complexes, 1.1a,21 1.10a,21 and the HS sites in 1.12a.24 There are three outliers, 1.5, 1.12b and 1.13 with noticeably shorter bond lengths, Table 3, which are comparable to the LS site in the previously reported [FeIII(4,6-diOMe-sal2232)]ClO4 complex which undergoes two-step SCO.24 In these examples, 1.5, 1.12b and 1.13, and the LS site in 1.12a, we observe that only half of the cation is in the asymmetric unit, as the complex cation is bisected by a C2 axis, Fig. 2. The shortened bond lengths in the LS state, allow the ligand to be more symmetrical around the FeIII centre, which is hindered in the more asymmetrically distorted HS examples where often one O–Fe–Namine axis is elongated further than the other. The breaking of the molecular symmetry in the remaining [FeIII(R-sal2232)]+ complexes appears to prevent the complex from accessing the LS state. While the C2 axis appears necessary for the stabilisation of the LS state, there are a couple of examples where the C2 axis is present but the HS state is stabilised at the temperature of the diffraction experiment, 1.1a and 1.1b.
image file: d5dt01833a-f2.tif
Fig. 2 Structure of isomeric complexes SCO 1.5 (a) and HS 2.4 (b) at 100 K, ellipsoids are drawn at 50% atomic probability. Hydrogen atoms have been omitted for clarity. The asymmetric unit of 1.5 (a) is half of the complex as there is a C2 axis bisecting the complex.

We next used structural distortion parameters to quantify the distortion around the FeIII centres, notably the parameters Σ and Θ, which are commonly used to diagnose the spin state, Table 1.27–29

image file: d5dt01833a-t1.tif

The Σ parameter is the sum of the deviation of the 12 unique cis ligand–metal–ligand angles (ϕi) from 90°. The Θ parameter is the sum of the deviation of the 24 unique torsional angles (θi) between ligand atoms on opposite triangular faces of the octahedron, Fig. S2.1. In addition, we consider α as the dihedral angle between the least squares planes of the two phenolate rings, as defined by Halcrow et al.15 with the values summarised in Table 1.

Table 1 Summary of all complexes with molecular formulae, structural details and structural distortion parameters. Complexes which have been reported previously are indicated with the appropriate citation and CSD Refcode for structural determination
Complex Molecular Formula Ta (K) Siteb Σc (°) Θc (°) αc(°) Spin stated Space group Comment
a The temperature at which the diffraction experiment was performed.b Where more than one crystallographic site is present the sites are labelled.c The distortion parameters are defined in the main text.d Spin state of the FeIII sites at the temperature of the diffraction experiment was performed.e Where An = anilate.
[FeIII(R-sal2232)]+ Family
1.1a21 [FeIII(sal2232)]ClO421 295   116 509 108 HS Pbcn Z′ = 0.5, C2 axis bisecting cation. CSD Refcode: DABKAS
1.1b [FeIII(sal2232)]OTf 100 I 101 436 112 HS P42bc Two sites, each with a C2 axis bisecting cation.
      II 80.7 260 101 HS  
1.1c23 [FeIII(sal2232)] [MnIICrIII(Br2An)3]e(CH3CN)423 120   73.6 272 86.1 HS P22121 CSD Refcode: HOWHIM
1.2 [FeIII(5-F-sal2232)]ClO4 100   76.9 254 112 HS P21/c  
1.3a [FeIII(5-Cl-sal2232)]ClO4 150   92.4 355 104 HS P21/c  
1.3b [FeIII(5-Cl-sal2232)]BF4 100   90.0 341 102 HS P21/c  
1.4a [FeIII(5-Br-sal2232)]ClO4 100   88.2 326 105 HS P21/c  
1.4b [FeIII(5-Br-sal2232)]BF4 100   84.4 304 103 HS P21/c  
1.5 [FeIII(5-I-sal2232)]ClO4 100   53.8 141 69.8 LS C2/c Z′ = 0.5, C2 axis bisecting cation. SCO close to RT.
1.6 [FeIII(3,5-diF-sal2232)]ClO4 100   79.4 271 111 HS P21/n  
1.7 [FeIII(3,5-diCl-sal2232)]ClO4 100   80.5 300 113 HS P21/n  
1.8a [FeIII(3,5-diBr-sal2232)]ClO4 100   81.0 301 114 HS P21/n  
1.8b [FeIII(3,5-diBr-sal2232)]NO3 100   99.9 364 114 HS P21/c  
1.9a [FeIII(3,5-diI-sal2232)]ClO4 100   85.9 319 115 HS P21/n  
1.9b [FeIII(3,5-diI-sal2232)]NO3 100 I 97.1 351 116 HS P21/n Two sites
      II 118 518 116 HS    
1.10a21 [FeIII(3-OMe-sal2232)]ClO421 295   84.1 303 96.9 HS Pccn CSD Refcode: DABPUR
1.10b [FeIII(3-OMe-sal2232)]NO3 100   88.5 325 95.2 HS Pccn  
1.10c [FeIII(3-OMe-sal2232)]BF4 100   85.5 304 94.3 HS Pccn  
1.11 [FeIII(4-OH-sal2232)]ClO4 293   73.3 248 111 HS P21/c  
1.12a24 [FeIII(4,6-diOMe-sal2232)]ClO424 200   74.6 302 112 HS C2/c Two-step SCO from HS → HS[thin space (1/6-em)]:[thin space (1/6-em)]HS[thin space (1/6-em)]:[thin space (1/6-em)]LS → HS[thin space (1/6-em)]:[thin space (1/6-em)]LS[thin space (1/6-em)]:[thin space (1/6-em)]LS. LS site in the HS[thin space (1/6-em)]:[thin space (1/6-em)]HS[thin space (1/6-em)]:[thin space (1/6-em)]LS mixed phase is bisected by a C2 axis. CSD Refcode: ELANOV
    160 I 54.0 151 108 LS C2/c
      II 87.5 346 116 HS  
1.12b [FeIII(4,6-diOMe-sal2232)]PF6 100   58.9 222 107 LS C2/c Z′ = 0.5, C2 axis bisecting cation.
1.12c [FeIII(4,6-diOMe-sal2232)]OTf·0.27H2O 100   73.1 249 96.5 HS P[1 with combining macron]  
1.12d [FeIII(4,6-diOMe-sal2232)]BPh4 293   82.8 270 125 HS P[1 with combining macron]  
1.13 [FeIII(5-NO2-sal2232)]PF6 100   64.7 173 103 LS I2/a Z′ = 0.5, C2 axis bisecting cation.
1.14 [FeIII(3,5-ditBu-sal2232)]ClO4 100   85.0 286 98.8 HS P21/n  
[FeIII(R-sal2223)]+ Family
2.1 [FeIII(5-F-sal2223)]NO3 100   91.0 317 139 HS P212121  
2.2a [FeIII(5-Cl-sal2223)]ClO4 105   81.2 294 119 HS Ia  
2.2b [FeIII(5-Cl-sal2223)]NO3 100   82.7 256 123 HS Ia  
2.3 [FeIII(5-Br-sal2223)]ClO4 100   81.1 275 121 HS Ia  
2.4 [FeIII(5-I-sal2223)]ClO4 110   82.5 259 122 HS Ia  
2.5 [FeIII(3,5-diBr-sal2223)]NO3 105 I 93.0 330 115 HS P[1 with combining macron] Two sites
      II 99.9 429 112 HS
2.6a [FeIII(3-Br-5-Cl-sal2223)]ClO4·0.25H2O·0.25CH3OH 105 I 84.7 288 134 HS P[1 with combining macron] Two sites
      II 86.0 301 132 HS    
2.6b [FeIII(3-Br-5-Cl-sal2223)]NO3 105 I 102 440 113 HS P[1 with combining macron] Two sites
      II 93.9 332 117 HS  
2.7 [FeIII(3-Br-6-OMe-sal2223)]ClO4·0.5CH3OH 115 I 97.5 383 118 HS P21/c Two sites
      II 93.9 397 116 HS
2.8 [FeIII(3-OEt-sal2223)]NO3·CH3OH 115   83.2 322 125 HS P21/c  
2.9 [FeIII(3-MeOMe-5-NO2-sal2223)]ClO4 110   85.3 275 136 HS P21/c  
2.10 [FeIII(3-Allyl-sal2223)]ClO4 100 I 82.1 284 114 HS P[1 with combining macron] Two sites
      II 85.9 289 122 HS
2.11 [FeIII(napsal2223)]ClO4·C2H3N·0.33H2O 100   87.8 306 120 HS P21/c  


The distortion parameters vary quite widely, dependent on the substitution of the ligand, and also the packing of the molecules in the lattice. The HS complexes have values in the range: Σ = 70°–125°; Θ = 230°–520°; and α = 85°–125°, which are significantly larger than FeIII complexes of the LS [FeIII(R-sal2323)]+ family18 but of the order of those of HS [FeIII(R-sal2222)]+ complexes.14–16 Complexes with the ‘232’ and ‘222’ tetraamine backbone are significantly more strained due to the shorter tetraamine backbone than ‘323’, resulting in the cis-orientation of the phenolate donors and increased distortion around the FeIII centre. The structures of ‘232’ with Σ < 65°, Θ < 225° also have bond lengths indicative of the LS state, Fig. 3. While 1.5 has a noticeably smaller α angle than the other complexes reported, there does not appear to be a clear link between this angle and spin state.


image file: d5dt01833a-f3.tif
Fig. 3 (a) Comparison of the Σ and Θ distortion parameters for 1.1–1.14. The sites which show lower values of Σ and Θ exhibit LS or SCO behaviour. The red HS site indicated as a square is one capable of undergoing SCO. (b) Comparison of Θ and α distortion parameters. The sites highlighted in green represent sites assigned as LS, sites in orange represent sites assigned with an equilibrium of HS/LS as full conversion to LS did not occur, and sites in red represent HS.

The smaller α angle in 1.5 permits stronger hydrogen bonding interactions between the cations and perchlorate anion (N–H⋯O), Table 2, resulting in a denser packing where hydrogen bonded chains form along the c-axis with the cations arranged pointing the same direction along the chain, Fig. 4(a). Comparing this to the packing in 1.4a, as an example of a HS analogue, reveals a different packing arrangement, where the hydrogen bonding chains are offset with cations of alternating direction along the b-axis, Fig. 4(b). This arrangement in 1.4a is replicated in most of the other complexes in the ‘232’ family, except 1.2, 1.11 and 1.12c which form dimer-like hydrogen bonded units through moieties on the salicylaldehyde ring, and those with a C2 axis (1.1a, 1.1b, 1.12a, 1.12b and 1.13). The complexes with a C2 axis show a similar motif to 1.5 where the cations are arranged pointing the same direction along the hydrogen bonding chain, Fig. S2.2. The combination of the C2 axis and associated packing arrangement must therefore lead to favourable conditions to promote SCO and stabilise the LS state. Further hydrogen bonding information for 1.1a–1.14 is shown in Table S2.4.


image file: d5dt01833a-f4.tif
Fig. 4 Partial packing diagram for 1.5 (a) and 1.4a (b) with hydrogen bonding chain formation shown in bright blue. The labels for the hydrogen bonds correspond to the values in Table 2.
Table 2 Hydrogen bonding interactions for 1.5 SCO and 1.4a HS, with labels corresponding to Fig. 4
Label d(D–H) (Å) d(H⋯A) (Å) d(D⋯A) (Å) ∠DHA (°)
1.5 SCO
1 0.96(3) 2.00(4) 2.957(3) 174(3)
1.4a HS
2 0.82(2) 2.18(2) 2.993(2) 167(2)
3 0.84(2) 2.43(2) 3.130(2) 141(2)


Table 3 Summary of bond lengths for selected complexes 1.1a, 1.5, 1.12a, 1.12b, 1.13 and 2.4 at 100 K (unless otherwise indicated)
Complex 1.1a21 1.5 1.12a (160 K)24 1.12b 1.13 2.4
a Sites are indicated as I and II when more than one crystallographically independent site is present.b The bond lengths in this case are slightly longer than expected for a fully LS site.
Sitea     I II      
Spin state HS LS LS HS LS[thin space (1/6-em)]:[thin space (1/6-em)]HS (mixed)b LS HS
Bond lengths (Å)
Fe–O 1.924(3) 1.8805(16) 1.871(3) 1.915(3) 1.8939(13) 1.8921(10) 1.914(6)
  1.918(3) 1.917(6)
Fe–Nimine 2.134(3) 1.9407(18) 1.924(4) 2.077(4) 1.9879(18) 1.9603(12) 2.119(7)
  2.088(4) 2.105(7)
Fe–Namine 2.207(3) 2.024(2) 2.002(4) 2.185(4) 2.0888(16) 2.0279(13) 2.204(7)
  2.192(4) 2.167(8)


Structural characterisation of the asymmetric [FeIII(R-sal2223)]+ family of complexes. We also report here for the first time complexes with the asymmetric ‘223’ tetraamine backbone, which has the same chain length as the ‘232’ backbone, but no access to C2 symmetry in the complex due to the position of the propylene linker in this tetraamine. Similarly to the ‘232’ case, the ligand adopts an N4O22− coordination sphere with cis-phenolate donors, cis-amine donors and trans-imine donors. The bond lengths around the FeIII centres are long for all complexes, 2.1–2.11, suggesting the ‘223’ tetraamine backbone has a greater propensity to stabilise the HS state. The asymmetric nature of this ligand means that the Fe–Namine distances are typically longer for those of the amine between the two ethylene linkers than for the amine between the ethylene and propylene linker, e.g. 2.201(9) Å and 2.169(9) Å respectively for 2.4. In addition, the aromatic moiety on the propylene end of the backbone is more distorted. The imine nitrogen, phenolate oxygen and iron atoms deviate from the plane of the aromatic ring as compared to the [FeIII(R-sal2232)]+ examples where the aromatic ring, imine, phenolate oxygen and iron are all almost coplanar.

In the mono-halogenated examples 2.1–2.4 we notice a preference for crystallisation in non-centrosymmetric space groups, Ia (for 2.2–2.4) and P212121 (for 2.1), with those in the Sohncke space group P212121 having only one of the Δ or Λ isomers of the cationic complex present in the asymmetric unit.

The distortion parameters Σ, Θ, and α are similar in 2.1–2.11, with values of the same magnitude as the HS sites in the [FeIII(R-sal2232)]+ examples, Table 1 and Fig. 5. Due to the asymmetric nature of this ligand backbone, the α angle is much larger than that of the [FeIII(R-sal2232)]+ examples, with values falling in the 110° to 140° region. This is due to the aforementioned distortion of the aromatic ring on the propylene linker. This increased distortion due to the placement of the propylene linker beside the imine results in the complexes having a preference for stabilisation of the more distorted HS state.


image file: d5dt01833a-f5.tif
Fig. 5 (a) Comparison of the Σ and Θ distortion parameters for all complexes, with those of the [FeIII(R-sal2232)]+ family (1.1–1.14) shown in red, and those of the [FeIII(R-sal2223)]+ family (2.1–2.11) shown in blue. The red HS site indicated as a square is one capable of undergoing SCO. (b) Comparison of Θ and α distortion parameters, with the same colour scheme.

Magnetic measurements

In order to confirm the spin state assignment by SCXRD, and to record the change at different temperatures we performed magnetic susceptibility measurements on those complexes where there was sufficient mass of sample. Plots of χMT versus T for 1.5 [FeIII(5-I-sal2232)]ClO4 and isomeric 2.4 [FeIII(5-I-sal2223)]ClO4 are shown in Fig. 6(a) and (b). The data reveals that 1.5 is LS (χMT = 0.48 cm3 K mol−1) up to ∼250 K, with the onset of gradual SCO above this temperature with a maximum value of χMT = 3.36 cm3 K mol−1 at 400 K. For 2.4 the complex is stabilised in the HS state across the measured range with a value of χMT = 4.78 cm3 K mol−1 at 300 K. Additional magnetic measurements for 1.8a, 1.9a, 1.10c, 1.11, 2.2b, 2.5, 2.6a, 2.6b, 2.8 and 2.10 are shown in Fig. S3.1, which all show HS values of χMT across the measured temperature range.
image file: d5dt01833a-f6.tif
Fig. 6 Plot of χMT versus T for 1.5 (a) and 2.4 (b) recorded between 400 K or 350 K and 5 K, at a scan rate of 1 K min−1. EPR spectra for 1.5 (c) and 1.3a (d) recorded on a lightly ground powders at room temperature, using an X-band EPR spectrometer (9.4 GHz).

We also recorded solid state electron paramagnetic resonance (EPR) spectra at room temperature on a lightly ground sample of the SCO sample, 1.5, and a HS analogue, 1.3a, Fig. 6(c) and (d). For 1.5 we observe a characteristic axial LS FeIII EPR signal, Fig. 6(c), with an intense resonance centred geff ≈ 2.1. A minor resonance is also observed centred close to geff ≈ 4, which is indicative of the presence of a trace HS component.30,31 This matches the observed χMT value of 1.03 cm3 K mol−1 observed in the magnetic susceptibility measurement at RT which is slightly higher than the fully LS system. In the case of 1.3a, we observe a broad resonance centred at geff ≈ 4, indicating the HS state of this sample.

In solution, room temperature magnetic susceptibility in DMSO-d6 was determined using Evans' 1H NMR method and a value of χMT = 3.98 cm3 K mol−1 was obtained for 1.5 indicating that the complex adopts a spin state closer to that of a pure HS moment (χMTspin–only (S = 5/2) = 4.375 cm3 K mol−1) in solution. This suggests that in the absence of the intermolecular interactions and packing effects in the crystal the preferred spin state tends towards HS.

Optical and vibrational spectroscopy

UV-visible spectra were recorded at room temperature on isomeric complexes 1.5 and 2.4 for both solid samples and in acetonitrile solution. The spectra for SCO 1.5 and HS 2.4 are shown in Fig. 7, along with calculated spectra for the respective cationic species (B3LYP/def2-TZVP/D3BJ). For 1.5 we observe differences between the solution measurements and solid state, whereby in solution we observe prominent bands at 240 nm (εmax = 69[thin space (1/6-em)]500 L mol−1 K−1), 304 nm (εmax = 12[thin space (1/6-em)]000 L mol−1 K−1), 328 nm (εmax = 12[thin space (1/6-em)]100 L mol−1 K−1), and 525 nm (εmax = 6100 L mol−1 K−1). The solution spectra of 1.5 and 2.4 are very similar, which is to be expected as only the backbone of the ligand is changing and not the chromophoric aromatic salicylaldehyde, and the spin state is also remaining consistent, with prominent bands in 2.4 at 240 nm (εmax = 58[thin space (1/6-em)]300 L mol−1 K−1), 300 nm (εmax = 10[thin space (1/6-em)]500 L mol−1 K−1), 340 nm (εmax = 8700 L mol−1 K−1), and 512 nm (εmax = 5300 L mol−1 K−1). The solid-state sample of HS 2.4 has bands at similar frequencies as in solution. The difference between the solid-state spectra of HS 2.4 and spin equilibrium LS[thin space (1/6-em)]:[thin space (1/6-em)]HS 1.5 are in line with the spin state assignment from magnetic data.
image file: d5dt01833a-f7.tif
Fig. 7 (a) UV-visible spectra of 1.5 in solution and the solid state along with the calculated spectra for the optimised LS and HS [FeIII(5-I-sal2232)]+ cation (B3LYP/def2-TZVP/D3BJ). (b) UV-visible spectra of 2.4 in solution and the solid state along with the calculated spectra for the optimised HS [FeIII(5-I-sal2223)]+ cation (B3LYP/def2-TZVP/D3BJ). All spectra were recorded at room temperature.

Raman spectroscopy was recorded on crystals of HS 1.3a, HS 1.4a, SCO 1.5, HS 2.2, HS 2.3, and HS 2.4 at room temperature over a frequency range of 250 cm−1 to 1650 cm−1. Changes in spin state are clearly apparent in the spectra of the isomeric complexes LS[thin space (1/6-em)]:[thin space (1/6-em)]HS 1.5 and HS 2.4, Fig. 8, and are particularly marked in the region between 300–500 cm−1. These are likely due to changes in the ligand to metal bonds as a result of changing spin state and changing backbone. These low-frequency Raman bands at ∼304–340 cm−1 and ∼412–436 cm−1 can be attributed to Fe–O stretching and Fe–N stretching vibrations, respectively.32,33 When we compare LS[thin space (1/6-em)]:[thin space (1/6-em)]HS 1.5 with other ‘232’ examples which are HS, Fig. S5.1, a difference in this region is also observed. In the infrared spectrum, differences are observed in the region where metal–ligand vibrations would be expected, with a shift in one peak from 543 cm−1 in 1.5 to 527 cm−1 in 2.4. In addition, the peak at 1332 cm−1 in 1.5 is likely due to C–O vibrations shifted to higher wavelength due to the LS state. A strong IR band near 1100 cm−1, attributed to the ν3 (asymmetric Cl–O stretch) of perchlorate, and a shoulder around 625–640 cm−1 (ν2 mode) support the presence of non-coordinated ClO4 ions.34,35


image file: d5dt01833a-f8.tif
Fig. 8 (a) Raman spectra of 1.5 and 2.4 recorded on crystals of the samples at room temperature. (b) Infrared spectra of 1.5 and 2.4 recorded on crystals of the samples at room temperature. The regions highlighted in green represent where the spectra differ between the two complexes.

Quantum chemistry calculations

Quantum chemistry calculations were performed using the ORCA 6.0.1 software package.36 To further understand the energetics in these FeIII complexes we studied the SCO [FeIII(5-I-sal2232)]+ and HS [FeIII(5-I-sal2223)]+ cations in more detail. The experimentally obtained molecular coordinates from SCXRD were used as a starting point for the geometry optimisation of the cations, for both the S = 1/2 (LS) and S = 5/2 (HS) states. Comparison of the geometry reveals similarity to the experimentally obtained structures, Table 4. The optimised structure of LS [FeIII(5-I-sal2232)]+ has similar bond lengths and distortions around the FeIII centre to the experimentally obtained structure. While we cannot compare the HS optimised structure directly with experimental data, as the structural data were obtained at 100 K when the complex is fully LS, we note that the bond lengths around the FeIII centre are close to those obtained for HS complexes with different ‘R’ groups on the phenolate ring, Table 4. One of the most notable differences between the LS and HS optimised structures of [FeIII(5-I-sal2232)]+ is the large change in the α angle between the planes of the phenolate rings, from α = 69.20° (LS) to α = 113.4° (HS). The optimised structure of [FeIII(5-I-sal2223)]+ for the HS state has similar bond lengths and distortions to the experimentally obtained data for 2.4. The LS optimised structure has similar bond lengths to the ‘232’ example, however the phenolate and imine donor adjacent to the propylene linker are quite distorted from the plane of the phenolate ring, perhaps increasing the instability of this spin state in the ‘223’ series.
Table 4 Summary of experimentally determined bond lengths and distortion parameters compared with the parameters for the optimised geometries obtained using TPSSh/def2-TZVP/D3BJ
  LS 1.5 LS Opt HS Opt HS 2.4 LS Opt HS Opt
Bond lengths (Å)
Fe–O 1.8805(16) 1.890 1.935 1.915(6) 1.870 1.912
  1.886 1.928 1.913(7) 1.910 1.942
Fe–Nimine 1.9407(18) 1.936 2.122 2.100(8) 1.974 2.119
  1.939 2.122 2.120(9) 1.939 2.133
Fe–Namine 2.024(2) 2.020 2.215 2.169(9) 2.027 2.199
  2.055 2.210 2.201(9) 2.029 2.227
Distortion parameters (°)
Σ 53.80 56.53 79.70 82.46 56.21 91.90
Θ 140.79 149.62 270.33 259.40 170.36 286.52
α 69.76 69.20 113.4 122.03 117.8 130.21


We use a similar procedure to that benchmarked by Kepp37 and Kulik et al.38 against FeII and FeIII SCO complexes to determine the Gibb's free energy, ΔGSCO, with the B3LYP* (15% HF exchange) and TPSSh functional, which we also utilised in a previous study of [FeIII(R-sal2323)]+ complexes.18 While the absolute values vary depending on the functional used, we can directly compare the ΔGSCO for the [FeIII(5-I-sal2232)]+ and [FeIII(5-I-sal2223)]+ cations when using the same methods. We observe that the HS state is significantly more stabilised in the [FeIII(5-I-sal2223)]+ cation, which is in agreement with the experimentally obtained data where spectroscopic and structural data reveals that this complex only supports the HS state, Table 5. Comparison of the ΔGSCO to the previously reported values for the LS [FeIII(sal2323)]+ species,18 and a reference SCO [FeIII(sal2222)]+ compound show that stability of the LS state decreases with ‘323’ ≈ ‘222’ > ‘232’ > ‘223’, which is in line with experimental results for these types of complexes, Table S5.5.

Table 5 Calculated Gibbs free energies (ΔGSCO) upon SCO for the [FeIII(5-I-sal2232)]+ and [FeIII(5-I-sal2223)]+ cations, using both the TPSSh and B3LYP* functional. A positive value means the LS state is favoured and a negative value means the HS state is favoured
  [FeIII(5-I-sal2232)]+ [FeIII(5-I-sal2223)]+
ΔGSCO (HS−LS) (kJ mol−1)    
TPSSh +36.53 +22.04
B3LYP* +2.51 −11.11


Conclusions

In expanding the series of [FeIII(R-sal2232)]+ from the handful of previous examples, we revealed that of the 25 examples reported, the vast majority are stabilised in the HS state. We show one example where SCO occurs close to room temperature, 1.5, and two other new complexes, 1.12b and 1.13, where crystallographic data indicates the presence of a LS site. In all examples where a LS site has been observed the local symmetry around the FeIII centre resulted in a C2 axis bisecting the ligand, which appears to be essential for the stability of the less distorted LS state, by virtue of the fact that the decrease in distortion can allow a more symmetric arrangement of the ligand around the FeIII centre. We also report a new family of asymmetric FeIII complexes of the [FeIII(R-sal2223)]+ form. Changes in the local distortion due to the position of the propylene linker (asymmetric ‘223’ arrangement instead of ‘232’) result in these complexes only supporting the HS state. In addition, comparison of the ΔGSCO calculated using DFT showed that the HS state is more energetically favourable in the ‘223’ example, which may be why we only observe the stabilisation of the HS state in these examples.

Experimental details

Synthetic procedures

Reagents were purchased from Merck (Sigma-Aldrich), Fluorochem, TCI Chemicals or VWR and were used without further purification. Complete synthetic details for 1.1–1.14 and 2.1–2.11 can be found in Section S1.

Single crystal X-ray diffraction (SCXRD)

SCXRD data was collected on suitable crystals using a Rigaku Oxford Diffraction SuperNova diffractometer with an Atlas CCD detector. The temperature was controlled using an Oxford Instruments Cryojet. CrysAlisPRO software was used for the data collection, integration, reduction and finalisation.39 An analytical absorption correction based on the shape of the crystal was performed. Structure solution was performed using either direct methods with ShelXS40 or with ShelXT41 and refined by full matrix least-squares on F2 with ShelXL.42 Hydrogen atoms were geometrically constrained and refined riding on the parent atom. Their isotropic displacement parameters were fixed to 1.2 times the parent atom, except for terminal –CH3 groups where this was 1.5 times the parent atom. All non-hydrogen atoms were refined anisotropically. Further crystallographic details can be found in Section S2. Octahedral distortion parameters were determined using the OctaDist program.43

Magnetic measurements

Magnetic susceptibility measurements were performed using a Quantum Design MPMS-XL SQUID magnetometer (5 T or 7 T) operating between 5 K and 400 K. Magnetic susceptibility was recorded with a field of 0.1 T or 0.5 T on a polycrystalline sample packed in a gelatin capsule at a scan rate of 1 K min−1. Diamagnetic corrections were applied to correct for the contribution of the diamagnetism of the samples.

Electron paramagnetic resonance (EPR) spectroscopy

EPR spectra were recorded using a Magnettech MS200 X-band (9.4 GHz) spectrometer between 50 and 450 mT with a modulation amplitude of 0.2 mT and a microwave power of 10 mW. Measurements were performed on lightly ground powders of 1.5 and 1.3a.

UV-Visible spectroscopy

UV–visible absorption spectra were acquired using an Agilent Cary 60 UV–Vis spectrophotometer at room temperature. Measurements in solution were conducted in acetonitrile. For solid-state analyses, the samples were finely ground and homogenized with 200 mg of potassium bromide (KBr), then compressed into a transparent pellet using a 15 mm die. The resulting pellet was subsequently mounted onto a solid sample holder for spectral acquisition.

Raman spectroscopy

Raman measurements were performed using the T64000 Horiba Jobin Yvon micro-Raman setup at room temperature. The excitation wavelength for the complex was 632.8 nm, emitted from a He–Ne laser (Optronics Technologies SA Model HLA-20P, 20 mW). The laser power on the sample was 2 mW. The backscattered radiation was collected from a single configuration of the monochromator after passing through an appropriate edge filter (LP02-633RU-25, laser2000, UK, Ltd, Huntingdon, Cambridgeshire, UK). The calibration of the instrument was achieved via the standard Raman peak position of Si at 520.5 cm−1. The spectral resolution was 5 cm−1.

Elemental analysis

CHN analysis was performed using an Exeter Analytical CE-440 Elemental Analyser.

Infrared spectroscopy

Infrared spectra were recorded using a Bruker Alpha Platinum ATR-FTIR spectrometer fitted with a diamond anvil.

Evans' method 1H NMR

Magnetic susceptibility measurements in solution were obtained using Evans' method using an Agilent DD2 500 MHz spectrometer. A standard 5 mm NMR tube containing the sample was fitted with a 3 mm coaxial insert tube containing DMSO-d6 (6% TMS). The NMR tube was carefully sealed to avoid solvent evaporation. The molar magnetic susceptibility, χM, for a long cylindrical shape oriented parallel to the magnetic field was calculated according to the equation below, where Δν is the shift in the frequency of the reference solvent peak in Hz, χ0 is the molar susceptibility of the solvent, ν0 is the operating frequency of the spectrometer in Hz, [C] is the concentration of the sample in mol L−1, and MW is the molecular mass (of either the solvent or sample):
image file: d5dt01833a-t2.tif

Quantum chemistry calculations

Calculations were performed using the ORCA 6.0.1 software.36 Structural data obtained from SCXRD were used as the starting point for the geometry optimisations. Geometry optimisations were performed using either the B3LYP*44,45 or TPSSh46,47 functional and the triple ζ def2-TZVP basis set48 together with the atom-pairwise dispersion correction (D3BJ).49,50 Cartesian coordinates of all the optimised geometries can be found in Tables S6.1 and S6.2. Analytical frequencies were calculated at the same level in order to determine that the geometries represented stable energy minima. The thermodynamic properties, ΔGSCO, were computed with T = 298.15 K thermodynamic corrections. Calculated UV-visible spectra were determined using TDDFT using the B3LYP functional, which utilised the Tamm-Dancoff approximation (TDA).51

Author contributions

C. T. K. – conceptualisation, data curation, formal analysis, investigation, methodology, validation, writing – original draft, writing – review & editing. E. C. – formal analysis, investigation, data curation. E. P. – conceptualisation, formal analysis, investigation. M. G. – formal analysis, investigation. N. Q. – investigation. P. N. – investigation, resources. H. M.-B. – formal analysis, investigation, data curation. J. B.-C. – formal analysis, investigation. S. F. – investigation, resources. Z. G. L. – investigation, methodology, resources. G. G. M. – conceptualisation, funding acquisition, resources, writing – review & editing, supervision, project administration.

Conflicts of interest

There are no conflicts to declare.

Data availability

The data supporting this article have been included as part of the SI. Supplementary information: synthetic details for 1.1–1.14 and 2.1–2.11; crystallographic details; bond lengths and distortion parameters; additional magnetic measurements; additional UV-visible, infrared and Raman spectra; Cartesian coordinates of the optimised geometries of [FeIII(5-I-sal2232)]+ and [FeIII(5-I-sal2223)]+ cationic species. See DOI: https://doi.org/10.1039/d5dt01833a.

CCDC 2456518–2456538 and 2467729–2467741 contain the supplementary crystallographic data for this paper.52a–ah

Acknowledgements

This work was financially supported by Science Foundation Ireland (SFI) with a US-Ireland Award (19/US/3631 to G. G. M.) and a Frontiers for the Future Award (19/FFP/6909 to G. G. M.). The authors wish to thank Mr Rónán Crowley for elemental analysis. A portion of this work has been carried out using the ResearchIT Sonic cluster at University College Dublin which was funded by UCD IT Services and the Research Office.

References

  1. I. G. Denisov, T. M. Makris, S. G. Sligar and I. Schlichting, Chem. Rev., 2005, 105, 2253–2278 CrossRef CAS.
  2. M. Swart and M. Costas, in Spin States in Biochemistry and Inorganic Chemistry, 2015, pp. 1–5 Search PubMed.
  3. M. Swart, M. Güell and M. Solà, in Quantum Biochemistry, 2010, pp. 551–583 Search PubMed.
  4. M. Swart, Int. J. Quantum Chem., 2013, 113, 2–7 CrossRef CAS.
  5. M. P. Shaver, L. E. N. Allan, H. S. Rzepa and V. C. Gibson, Angew. Chem., Int. Ed., 2006, 45, 1241–1244 CrossRef CAS.
  6. M. P. Johansson and M. Swart, Dalton Trans., 2011, 40, 8419–8428 RSC.
  7. J. N. Harvey, R. Poli and K. M. Smith, Coord. Chem. Rev., 2003, 238–239, 347–361 CrossRef CAS.
  8. P. Gütlich and H. A. Goodwin, in Spin Crossover in Transition Metal Compounds I, ed. P. Gütlich and H. A. Goodwin, Springer Berlin Heidelberg, Berlin, Heidelberg, 2004, pp. 1–47 Search PubMed.
  9. K. S. Murray, in Spin–Crossover Materials, 2013, pp. 1–54 Search PubMed.
  10. A. Bousseksou, G. Molnár, L. Salmon and W. Nicolazzi, Chem. Soc. Rev., 2011, 40, 3313–3335 RSC.
  11. D. J. Harding, P. Harding and W. Phonsri, Coord. Chem. Rev., 2016, 313, 38–61 CrossRef CAS.
  12. M. Nihei, T. Shiga, Y. Maeda and H. Oshio, Coord. Chem. Rev., 2007, 251, 2606–2621 CrossRef CAS.
  13. P. J. van Koningsbruggen, Y. Maeda and H. Oshio, in Spin Crossover in Transition Metal Compounds I, ed. P. Gütlich and H. A. Goodwin, Springer Berlin Heidelberg, Berlin, Heidelberg, 2004, pp. 259–324 Search PubMed.
  14. M. F. Tweedle and L. J. Wilson, J. Am. Chem. Soc., 1976, 98, 4824–4834 CrossRef CAS.
  15. R. Pritchard, S. A. Barrett, C. A. Kilner and M. A. Halcrow, Dalton Trans., 2008, 3159–3168 RSC.
  16. S. Dorbes, L. Valade, J. A. Real and C. Faulmann, Chem. Commun., 2005, 69–71 RSC.
  17. K. J. Howard-Smith, A. R. Craze, R. Tokunaga, T. Taira, H. Min, M. J. Wallis, D. J. Fanna, S. Hayami and F. Li, Cryst. Growth Des., 2023, 23, 1219–1228 CrossRef CAS.
  18. C. T. Kelly, M. Griffin, K. Esien, S. Felton, H. Müller-Bunz and G. G. Morgan, Cryst. Growth Des., 2022, 22, 6429–6439 CrossRef CAS.
  19. R. Kannappan, S. Tanase, I. Mutikainen, U. Turpeinen and J. Reedijk, Polyhedron, 2006, 25, 1646–1654 CrossRef CAS.
  20. T. Ito, M. Sugimoto, H. Ito, K. Toriumi, H. Nakayama, W. Mori and M. Sekizaki, Chem. Lett., 2006, 12, 121–124 CrossRef.
  21. S. Hayami, T. Matoba, S. Nomiyama, T. Kojima, S. Osaki and Y. Maeda, Bull. Chem. Soc. Jpn., 1997, 70, 3001–3009 CrossRef CAS.
  22. C. T. Kelly, S. Dunne, I. A. Kühne, A. Barker, K. Esien, S. Felton, H. Müller-Bunz, Y. Ortin and G. G. Morgan, Angew. Chem., Int. Ed., 2023, 62, e202217388 CrossRef CAS PubMed.
  23. A. Abhervé, M. Clemente-León, E. Coronado, C. J. Gómez-García and M. Verneret, Inorg. Chem., 2014, 53, 12014–12026 CrossRef.
  24. M. Griffin, S. Shakespeare, H. J. Shepherd, C. J. Harding, J. F. Letard, C. Desplanches, A. E. Goeta, J. A. Howard, A. K. Powell, V. Mereacre, Y. Garcia, A. D. Naik, H. Muller-Bunz and G. G. Morgan, Angew. Chem., Int. Ed., 2011, 50, 896–900 CrossRef CAS.
  25. S. Sundaresan, I. A. Kühne, C. T. Kelly, A. Barker, D. Salley, H. Müller-Bunz, A. K. Powell and G. G. Morgan, Crystals, 2019, 9, 19 CrossRef.
  26. T. Ito, M. Sugimoto, H. Ito, K. Toriumi, H. Nakayama, W. Mori and M. Sekizaki, Chem. Lett., 1983, 12, 121–124 CrossRef.
  27. M. G. B. Drew, C. J. Harding, V. McKee, G. G. Morgan and J. Nelson, J. Chem. Soc., Chem. Commun., 1995, 1035–1038 RSC.
  28. P. Guionneau, C. Brigouleix, Y. Barrans, A. E. Goeta, J.-F. Létard, J. A. K. Howard, J. Gaultier and D. Chasseau, C. R. Acad. Sci., Ser. IIc:Chim., 2001, 4, 161–171 CrossRef CAS.
  29. M. Marchivie, P. Guionneau, J.-F. Létard and D. Chasseau, Acta Crystallogr., Sect. B:Struct. Sci., 2005, 61, 25–28 CrossRef.
  30. M. S. Shongwe, S. H. Al-Rahbi, M. A. Al-Azani, A. A. Al-Muharbi, F. Al-Mjeni, D. Matoga, A. Gismelseed, I. A. Al-Omari, A. Yousif, H. Adams, M. J. Morris and M. Mikuriya, Dalton Trans., 2012, 41, 2500–2514 RSC.
  31. J. Tang, J. Sánchez Costa, S. Smulders, G. Molnár, A. Bousseksou, S. J. Teat, Y. Li, G. A. van Albada, P. Gamez and J. Reedijk, Inorg. Chem., 2009, 48, 2128–2135 CrossRef CAS PubMed.
  32. Z. G. Lada, K. S. Andrikopoulos, A. Chrissanthopoulos, S. P. Perlepes and G. A. Voyiatzis, Inorg. Chem., 2019, 58, 5183–5195 CrossRef CAS.
  33. Z. G. Lada, A. Chrissanthopoulos, S. P. Perlepes, K. S. Andrikopoulos and G. A. Voyiatzis, Chem. Commun., 2022, 58, 521–524 RSC.
  34. J. L. Pascal, J. Potier, D. J. Jones, J. Roziere and A. Michalowicz, Inorg. Chem., 1985, 24, 238–241 CrossRef CAS.
  35. K. Nakamoto, in Infrared and Raman Spectra of Inorganic and Coordination Compounds, ed. K. Nakamoto, 2008, pp. 149–354 Search PubMed.
  36. F. Neese, Wiley Interdiscip. Rev.: Comput. Mol. Sci., 2022, 12, e1606 Search PubMed.
  37. K. P. Kepp, Inorg. Chem., 2016, 55, 2717–2727 CrossRef CAS PubMed.
  38. V. Vennelakanti, M. G. Taylor, A. Nandy, C. Duan and H. J. Kulik, J. Chem. Phys., 2023, 159, 024120 CrossRef CAS.
  39. CrysAlisPRO, Rigaku Oxford Diffraction, Rigaku Corporation, Wroclaw, Poland Search PubMed.
  40. G. M. Sheldrick, Acta Crystallogr., Sect. A:Found. Crystallogr., 2008, 64, 112–122 CrossRef CAS.
  41. G. M. Sheldrick, Acta Crystallogr., Sect. A:Found. Adv., 2015, 71, 3–8 CrossRef PubMed.
  42. G. M. Sheldrick, Acta Crystallogr., Sect. C:Struct. Chem., 2015, 71, 3–8 Search PubMed.
  43. R. Ketkaew, Y. Tantirungrotechai, P. Harding, G. Chastanet, P. Guionneau, M. Marchivie and D. J. Harding, Dalton Trans., 2021, 50, 1086–1096 RSC.
  44. M. Reiher, O. Salomon and B. Artur Hess, Theor. Chem. Acc., 2001, 107, 48–55 Search PubMed.
  45. O. Salomon, M. Reiher and B. A. Hess, J. Chem. Phys., 2002, 117, 4729–4737 CrossRef CAS.
  46. J. Tao, J. P. Perdew, V. N. Staroverov and G. E. Scuseria, Phys. Rev. Lett., 2003, 91, 146401 CrossRef PubMed.
  47. J. P. Perdew, J. Tao, V. N. Staroverov and G. E. Scuseria, J. Chem. Phys., 2004, 120, 6898–6911 CrossRef CAS.
  48. F. Weigend and R. Ahlrichs, Phys. Chem. Chem. Phys., 2005, 7, 3297–3305 RSC.
  49. S. Grimme, J. Comput. Chem., 2006, 27, 1787–1799 CrossRef CAS PubMed.
  50. S. Grimme, J. Antony, S. Ehrlich and H. Krieg, J. Chem. Phys., 2010, 132, 154104 CrossRef.
  51. S. Grimme, J. Chem. Phys., 2013, 138, 244104 CrossRef PubMed.
  52. (a) C. T. Kelly, E. Cuza, E. Pasquetti, N. Quinn, M. Griffin, P. Nockemann, H. Müller-Bunz, J. Bruno-Colmenarez, S. Felton, Z. G. Lada and G. G. Morgan, CCDC 2456518: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ng6jw; (b) C. T. Kelly, E. Cuza, E. Pasquetti, N. Quinn, M. Griffin, P. Nockemann, H. Müller-Bunz, J. Bruno-Colmenarez, S. Felton, Z. G. Lada and G. G. Morgan, CCDC 2456519: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ng6kx; (c) C. T. Kelly, E. Cuza, E. Pasquetti, N. Quinn, M. Griffin, P. Nockemann, H. Müller-Bunz, J. Bruno-Colmenarez, S. Felton, Z. G. Lada and G. G. Morgan, CCDC 2456520: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ng6ly; (d) C. T. Kelly, E. Cuza, E. Pasquetti, N. Quinn, M. Griffin, P. Nockemann, H. Müller-Bunz, J. Bruno-Colmenarez, S. Felton, Z. G. Lada and G. G. Morgan, CCDC 2456521: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ng6mz; (e) C. T. Kelly, E. Cuza, E. Pasquetti, N. Quinn, M. Griffin, P. Nockemann, H. Müller-Bunz, J. Bruno-Colmenarez, S. Felton, Z. G. Lada and G. G. Morgan, CCDC 2456522: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ng6n0; (f) C. T. Kelly, E. Cuza, E. Pasquetti, N. Quinn, M. Griffin, P. Nockemann, H. Müller-Bunz, J. Bruno-Colmenarez, S. Felton, Z. G. Lada and G. G. Morgan, CCDC 2456523: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ng6p1; (g) C. T. Kelly, E. Cuza, E. Pasquetti, N. Quinn, M. Griffin, P. Nockemann, H. Müller-Bunz, J. Bruno-Colmenarez, S. Felton, Z. G. Lada and G. G. Morgan, CCDC 2456524: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ng6q2; (h) C. T. Kelly, E. Cuza, E. Pasquetti, N. Quinn, M. Griffin, P. Nockemann, H. Müller-Bunz, J. Bruno-Colmenarez, S. Felton, Z. G. Lada and G. G. Morgan, CCDC 2456525: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ng6r3; (i) C. T. Kelly, E. Cuza, E. Pasquetti, N. Quinn, M. Griffin, P. Nockemann, H. Müller-Bunz, J. Bruno-Colmenarez, S. Felton, Z. G. Lada and G. G. Morgan, CCDC 2456526: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ng6s4; (j) C. T. Kelly, E. Cuza, E. Pasquetti, N. Quinn, M. Griffin, P. Nockemann, H. Müller-Bunz, J. Bruno-Colmenarez, S. Felton, Z. G. Lada and G. G. Morgan, CCDC 2456527: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ng6t5; (k) C. T. Kelly, E. Cuza, E. Pasquetti, N. Quinn, M. Griffin, P. Nockemann, H. Müller-Bunz, J. Bruno-Colmenarez, S. Felton, Z. G. Lada and G. G. Morgan, CCDC 2456528: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ng6v6; (l) C. T. Kelly, E. Cuza, E. Pasquetti, N. Quinn, M. Griffin, P. Nockemann, H. Müller-Bunz, J. Bruno-Colmenarez, S. Felton, Z. G. Lada and G. G. Morgan, CCDC 2456529: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ng6w7; (m) C. T. Kelly, E. Cuza, E. Pasquetti, N. Quinn, M. Griffin, P. Nockemann, H. Müller-Bunz, J. Bruno-Colmenarez, S. Felton, Z. G. Lada and G. G. Morgan, CCDC 2456530: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ng6x8; (n) C. T. Kelly, E. Cuza, E. Pasquetti, N. Quinn, M. Griffin, P. Nockemann, H. Müller-Bunz, J. Bruno-Colmenarez, S. Felton, Z. G. Lada and G. G. Morgan, CCDC 2456531: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ng6y9; (o) C. T. Kelly, E. Cuza, E. Pasquetti, N. Quinn, M. Griffin, P. Nockemann, H. Müller-Bunz, J. Bruno-Colmenarez, S. Felton, Z. G. Lada and G. G. Morgan, CCDC 2456532: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ng6zb; (p) C. T. Kelly, E. Cuza, E. Pasquetti, N. Quinn, M. Griffin, P. Nockemann, H. Müller-Bunz, J. Bruno-Colmenarez, S. Felton, Z. G. Lada and G. G. Morgan, CCDC 2456533: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ng70d; (q) C. T. Kelly, E. Cuza, E. Pasquetti, N. Quinn, M. Griffin, P. Nockemann, H. Müller-Bunz, J. Bruno-Colmenarez, S. Felton, Z. G. Lada and G. G. Morgan, CCDC 2456534: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ng71f; (r) C. T. Kelly, E. Cuza, E. Pasquetti, N. Quinn, M. Griffin, P. Nockemann, H. Müller-Bunz, J. Bruno-Colmenarez, S. Felton, Z. G. Lada and G. G. Morgan, CCDC 2456535: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ng72g; (s) C. T. Kelly, E. Cuza, E. Pasquetti, N. Quinn, M. Griffin, P. Nockemann, H. Müller-Bunz, J. Bruno-Colmenarez, S. Felton, Z. G. Lada and G. G. Morgan, CCDC 2456536: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ng73h; (t) C. T. Kelly, E. Cuza, E. Pasquetti, N. Quinn, M. Griffin, P. Nockemann, H. Müller-Bunz, J. Bruno-Colmenarez, S. Felton, Z. G. Lada and G. G. Morgan, CCDC 2456537: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ng74j; (u) C. T. Kelly, E. Cuza, E. Pasquetti, N. Quinn, M. Griffin, P. Nockemann, H. Müller-Bunz, J. Bruno-Colmenarez, S. Felton, Z. G. Lada and G. G. Morgan, CCDC 2456538: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ng75k; (v) C. T. Kelly, E. Cuza, E. Pasquetti, N. Quinn, M. Griffin, P. Nockemann, H. Müller-Bunz, J. Bruno-Colmenarez, S. Felton, Z. G. Lada and G. G. Morgan, CCDC 2467729: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ntw5k; (w) C. T. Kelly, E. Cuza, E. Pasquetti, N. Quinn, M. Griffin, P. Nockemann, H. Müller-Bunz, J. Bruno-Colmenarez, S. Felton, Z. G. Lada and G. G. Morgan, CCDC 2467730: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ntw6l; (x) C. T. Kelly, E. Cuza, E. Pasquetti, N. Quinn, M. Griffin, P. Nockemann, H. Müller-Bunz, J. Bruno-Colmenarez, S. Felton, Z. G. Lada and G. G. Morgan, CCDC 2467731: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ntw7m; (y) C. T. Kelly, E. Cuza, E. Pasquetti, N. Quinn, M. Griffin, P. Nockemann, H. Müller-Bunz, J. Bruno-Colmenarez, S. Felton, Z. G. Lada and G. G. Morgan, CCDC 2467732: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ntw8n; (z) C. T. Kelly, E. Cuza, E. Pasquetti, N. Quinn, M. Griffin, P. Nockemann, H. Müller-Bunz, J. Bruno-Colmenarez, S. Felton, Z. G. Lada and G. G. Morgan, CCDC 2467733: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ntw9p; (a a) C. T. Kelly, E. Cuza, E. Pasquetti, N. Quinn, M. Griffin, P. Nockemann, H. Müller-Bunz, J. Bruno-Colmenarez, S. Felton, Z. G. Lada and G. G. Morgan, CCDC 2467734: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ntwbq; (a b) C. T. Kelly, E. Cuza, E. Pasquetti, N. Quinn, M. Griffin, P. Nockemann, H. Müller-Bunz, J. Bruno-Colmenarez, S. Felton, Z. G. Lada and G. G. Morgan, CCDC 2467735: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ntwcr; (a c) C. T. Kelly, E. Cuza, E. Pasquetti, N. Quinn, M. Griffin, P. Nockemann, H. Müller-Bunz, J. Bruno-Colmenarez, S. Felton, Z. G. Lada and G. G. Morgan, CCDC 2467736: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ntwds; (a d) C. T. Kelly, E. Cuza, E. Pasquetti, N. Quinn, M. Griffin, P. Nockemann, H. Müller-Bunz, J. Bruno-Colmenarez, S. Felton, Z. G. Lada and G. G. Morgan, CCDC 2467737: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ntwft; (a e) C. T. Kelly, E. Cuza, E. Pasquetti, N. Quinn, M. Griffin, P. Nockemann, H. Müller-Bunz, J. Bruno-Colmenarez, S. Felton, Z. G. Lada and G. G. Morgan, CCDC 2467738: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ntwgv; (a f) C. T. Kelly, E. Cuza, E. Pasquetti, N. Quinn, M. Griffin, P. Nockemann, H. Müller-Bunz, J. Bruno-Colmenarez, S. Felton, Z. G. Lada and G. G. Morgan, CCDC 2467739: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ntwhw; (a g) C. T. Kelly, E. Cuza, E. Pasquetti, N. Quinn, M. Griffin, P. Nockemann, H. Müller-Bunz, J. Bruno-Colmenarez, S. Felton, Z. G. Lada and G. G. Morgan, CCDC 2467740: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ntwjx; (a h) C. T. Kelly, E. Cuza, E. Pasquetti, N. Quinn, M. Griffin, P. Nockemann, H. Müller-Bunz, J. Bruno-Colmenarez, S. Felton, Z. G. Lada and G. G. Morgan, CCDC 2467741: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2ntwky.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.