ACS Publications. Most Trusted. Most Cited. Most Read
Variation of Spin-Transition Temperature in the Iron(III) Complex Induced by Different Compositions of the Crystallization Solvent
My Activity

Figure 1Loading Img
  • Open Access
Communication

Variation of Spin-Transition Temperature in the Iron(III) Complex Induced by Different Compositions of the Crystallization Solvent
Click to copy article linkArticle link copied!

Open PDFSupporting Information (1)

Crystal Growth & Design

Cite this: Cryst. Growth Des. 2023, 23, 3, 1323–1329
Click to copy citationCitation copied!
https://doi.org/10.1021/acs.cgd.2c01411
Published February 16, 2023

Copyright © 2023 American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Abstract

Click to copy section linkSection link copied!

We crystallized the Schiff-base iron(III) spin-crossover complex [Fe(3,5Cl-L5)(NCSe)] from different two-component solvent mixtures containing methanol and chloroform (Φ = V(CH3OH)/V(solvent) = 0.05, 0.25, 0.50, 0.83, and 1.00). The obtained crystalline products were characterized by X-ray diffraction, and it was confirmed that they are all composed of the same crystalline phase, and they do not contain any crystal solvent. However, significant differences in magnetic properties were observed, and thermal hysteresis changed from (in K) 121T and 134T for Φ = 0.05 and 0.25, down to 72T and 96T for Φ = 1.00. The crystal structures of the low-spin and high-spin phases were studied theoretically and experimentally.

This publication is licensed under

CC-BY 4.0 .
  • cc licence
  • by licence
Copyright © 2023 American Chemical Society

Special Issue

Published as part of a Crystal Growth and Design virtual special issue on Molecular Magnets and Switchable Magnetic Materials

Synopsis

The iron(III) Schiff-base spin-crossover complex [Fe(3,5Cl-L5)(NCSe)] exhibits a large variation of spin transition temperature and hysteresis loop width stimulated by the composition of a two-component solvent mixture containing methanol and chloroform. The relationship between crystal quality and T1/2 has been extensively studied.

The tunability of the spin-crossover (SCO) behavior is an important assumption for versatile applications of SCO materials. Beneficially, SCO parameters such as thermal hysteresis width, transition temperature (T1/2), cooperativity degree, curve abruptness, and completeness can be manipulated by sample modifications. The effect of the presence (1−3) and loss (4) of a solvent molecule in the lattice, the selection of an anion, (5,6) as well as the choice of a side substituent (6−8) are well-known strategies for modifying the SCO behavior. Also, metal dilution represents a sophisticated approach for modification of the SCO behavior and observing of the cooperativity degree and transition temperature of SCO materials. The SCO behavior can be also tuned by a postsynthetic modification as was manifested by solid state performed anion metathesis. (9)

Our ongoing interest in the magnetic behavior of Fe(III) complexes with pentadentate Schiff base ligands has brought remarkable results on distinct magnetic behaviors of polymorphs within this class of compounds, (10) or hydrogen bonding induced modification of T1/2. (11) Recently, a new report on SCO with broad thermal hysteresis observed for an Fe(III) complex with a pentadentate Schiff base ligand H23,5Cl-L5 (N,N′-bis(1-hydroxy-3,5-dichloro-2-benzyliden)-1,6-diamino-4-azahexane) has been reported by Renz et al. which naturally caught our attention. (12) Complex [Fe(3,5-Cl-L5)(NCSe)] (1) exhibits thermally induced SCO with 24 K wide thermal hysteresis (T1/2 = 99↓ and 129↑ K). From the comparison of the low-spin (LS, S = 1/2) and high-spin (HS, S = 5/2) crystal structures, it is apparent that reorganization of noncovalent interactions (H···Cl and Cl···Cl) happens upon spin transition. This, if significant, could explain the observation of the wide thermal hysteresis. In the original report, (12) the authors did not investigate this possibility in greater detail, and therefore, our original motivation for studying this system was to extensively theoretically and experimentally investigate the LS and HS crystal structures of 1.

For the preparation of 1 we used a procedure similar to that in the original report, (12) but instead of an ultrasonic bath we used a standard magnetic stirrer. The reaction between [Fe(3,5-Cl-L5)Cl] and KNCSe in pure methanol led to precipitation of a brown microcrystalline powder, which was filtered off using a paper filter. For the preparation of the single-crystals suitable for X-ray diffraction experiments we used the remaining mother liquor which was crystallized isothermally. After several days thin needle-like crystals were obtained. First, we collected the diffraction data for the HS state (150 K), and then we attempted to measure the LS state crystal structure at 90 K as was done in the original report. (12) To our surprise, the measurement revealed the crystal structure with the very same metal–ligand bond lengths and unit cell parameters as was observed for the HS phase. Obviously, this did not match the previously reported magnetic properties. (12) Therefore, we measured the temperature dependence of the magnetic moment (μeffB) for two obtained fractions, needle-like crystals (1_Φ1) and microcrystalline powder precipitate (1_Φ1p). The measurements confirmed the presence of SCO with thermal hysteresis for both batches; however, the critical temperatures and profiles of magnetic functions were different (Figure 1). The 1_Φ1 batch exhibited much lower T1/2 (72↓ and 96↑ K, ΔT = 24 K) than 1_Φ1_p (106↓ and 129↑ K, ΔT = 23 K, Figure 1). Powder X-ray diffraction undoubtedly confirmed that both samples were the same crystallographic phase identical to the HS structure of 1 (Supporting Information, Figure S6–S7).

Figure 1

Figure 1. Temperature dependence of μeffB for 1_Φ1 and 1_Φ1p.

Inspired by this intriguing inconsistency, we investigated this phenomenon in greater detail. From magnetic measurements it is apparent that the low T1/2 was observed for the batch composed of crystals which grew more than 5 days, whereas the high T1/2 was observed for the precipitate. Therefore, one of the tested hypotheses was that different critical temperatures were related to the crystallinity of the samples, which is related to the time of the crystallization. In an attempt to prepare crystals, whose crystallization spans different time periods, we modified the composition of crystallization solutions from pure CH3OH to solutions with a growing portion of CHCl3. CHCl3 was chosen as the second solvent because it dissolves 1 very well; it does not enter the crystal structure of 1 (does not form a solvate) and has a similar boiling temperature as CH3OH. Therefore, it could be expected that the crystallization of 1 will be governed by slow evaporation of the solvent mixture. This mixture gradually loses slightly more CHCl3 than CH3OH molecules (because of higher vapor pressure of CHCl3), and thus the solubility of 1 should be continuously decreasing. Four different crystallization mixtures with a different volume fraction Φ of CH3OH were prepared (Φ = 0.83, 0.50, 0.25 and 0.05), and the corresponding crystalline samples (1_Φ0.83, 1_Φ0.50, 1_Φ0.25, 1_Φ0.05) were obtained. In the case of the solutions with Φ > 0.5 also microcrystalline precipitates 1_Φ0.83p (besides already prepared 1_Φ1_p) were obtained and were included into this study. The purity of all the prepared samples was confirmed by X-ray powder diffraction. It must be noted that in the case of one of the precipitates (1_Φ1_p), the presence of KCl impurity was detected as this is a side product of the ligand metathesis (Cl ligand was substituted by KNCSe). We decided not to wash the precipitates with water (which would dissolve the impurity), because it would introduce a third solvent into the studied system. Furthermore, the presence of the diamagnetic KCl should not affect the SCO temperatures.

1 crystallizes as very thin needle-shaped crystals. The symmetry of the crystals is monoclinic with the P21/n space group (see Supporting Information, Table S1). The crystal structures of LS and HS phases were well described in the original report, (12) and we will not add a new structural description here.

The best quality crystals were obtained for the batch 1_Φ0.25. Again, we opted to measure the LS and HS crystal structures using a selected single crystal from this batch. We monitored SCO using diffraction methods. Therefore, we started the single crystal X-ray diffraction experiments at 140 K, and we collected sets of diffraction data at selected temperatures on cooling and also on heating. At 116 K (on cooling), a dramatic change in the quality of diffractions occurred as this was the T1/2↓ temperature. The crystal structure of 1_Φ0.25@↓116K shows metal–ligand (ML) bond lengths significantly shorter (Table S2) than observed at higher temperatures and in very good agreement with the LS structure of 1 reported in the original paper. (12) Upon further cooling (down to 108 K), neither the M–L bond lengths nor the unit cell parameters changed significantly, and therefore, we can conclude that the full HS → LS spin conversion occurred between 116 and 117 K (Figure 2). Upon heating, we detected the LS → HS transition between 133 (LS) and 135 K (HS structure), whereas we were not able to get reasonable data from the measurement at 134 K. We attempted to also measure another SCO thermal cycle, but the crystal cracked upon another cooling. Here, we may conclude that the single crystal X-ray diffraction measurements confirmed that 1_Φ0.25 exhibits thermally induced SCO with hysteresis wide 18 K (if we assume T1/2↑ to be 134 K). This is not in agreement with the magnetic data reported in the original paper, (12) but it is also not in good agreement with the magnetic data measured for 1_Φ1p. Of note here is that the coordinates of the non-hydrogen atoms in the HS crystal structure of 1_Φ1 (at 90 K) are almost identical with those determined for 1_Φ0.25 (at 130 K, see Supporting Information, Figure S1) and thus, the difference in T1/2 cannot be assigned to changes in the crystal structures.

Figure 2

Figure 2. Temperature dependence of μeffB for 1_Φ0.25 (top) and the temperature dependence of the selected unit cell parameters as determined from the single crystal X-ray diffraction experiment for 1_Φ0.25 (bottom).

We performed magnetic measurements for all the prepared batches. The results showed that all the samples show thermally induced spin crossover (Figure S2 and Figure S3). Remarkably, it is obvious that the batches prepared from the mixture with the largest content of methanol showed hysteretic loops shifted to lower temperatures, whereas the batches from the higher CHCl3 content had hysteretic loops shifted to higher temperatures (Figure S2). The shift of hysteretic loops is observable also between 1_Φ0.83 and 1_Φ1; however, the data for their precipitates 1_Φ0.83p and 1_Φ1p did not show a significant difference in T1/2 (Figure S3). The hysteretic loops for the batches with the highest CHCl3 content (1_Φ0.25 and 1_Φ0.05) are very similar, and they exhibit the highest T1/2. The magnetic behavior of 1_Φ0.25 fits the temperature dependence of the unit cell parameters rather well (Figure 2). The observed behavior is reproducible.

The magnetic data were analyzed by using an Ising-like model (ISM) with Gaussian distribution of the cooperativity parameter derived by Boča et al. (13) Within the ISM defined by the Hamiltonian

H^=Δ2σ^Γσσ^
(1)
the transition between two spin states, σ = −1 for LS and σ = +1 for HS states, is governed by the energy difference between HS and LS states (Δ) and the cooperativeness of the system (Γ). Such a model can be modified to also include the distribution of the cooperativity parameter Γ in order to deal with the imperfections of the crystalline/powder samples by varying the standard deviation parameter σ of the Gaussian distribution function. The main advantage of such a model is its ability to reproduce the slope of the hysteresis loops, and hence better agreement with the experimental data can be achieved. Thus, the magnetic data were fitted by varying Δ, Γ, reff, and σ parameters within ISM, and subsequently calculated x′HS for given temperature was used to calculate total molar susceptibility as
xHS=xHS(1xrHS)
(2)
χmol=(xHS+xrHS)χHS+(1xHSxrHS)χHS
(3)
where xrHS is the residual molar fraction of the HS state at a low temperature due to incomplete spin crossover, and xHS is rescaled high-spin fraction calculated from ISM resulting in the total high-spin mole fraction xHS equal to xHS = xHS + xrHS. The molar susceptibility of the HS and LS species was calculated by the Curie law. The fitted parameters are listed in Table S3 and plotted in Figure 3 for crystal samples and in Figure S4 for microcrystalline precipitates. Evidently, there is small variation of the cooperativity Γ and entropy parameter reff within the prepared crystal batches of 1, and variations of T1/2↓ and T1/2↑ are mainly due to a variation of Δ. Furthermore, there is a significant increase of the distribution parameter σ for the microcrystalline precipitates, Table S3.

Figure 3

Figure 3. Temperature dependence the high-spin molar fraction xHS according to ISM (top), the ISM parameters (middle), and transition temperatures (bottom) for crystalline samples of 1.

We also theoretically attempted to investigate the impact of different solvents on the molecular geometry and the energies of the LS and HS isomer of 1 using density functional theory (DFT) and utilizing ORCA 5.0 software. (14) We selected three functionals based on published benchmark studies, (15−17) namely, OPBE, (18) r2SCAN, (19) and B3LYP* (B3LYP with reduced Hartree–Fock exchange to 15%) (20) and also included the atom-pairwise dispersion correction (D4). (21) The optimization was done in a vacuum, chloroform, and methanol with the C-PCM implicit solvation model. (22,23) Impact of solvents is demonstrated for the HS state of [Fe(3,5-Cl-L5)(NCSe)] in Figure 4. It seems that OPBE underestimates the Fe–NCSe bond, whereas B3LYP* and r2SCAN overestimates the bond lengths to amino-nitrogen of 3,5-Cl-L5. In the case of the LS molecular geometries, OPBE heavily underestimates all Fe–N bond lengths (Figure S5). Moreover, there is also significant variation in bond lengths induced by the implicit solvation model, which points to the importance of intermolecular interactions.

Figure 4

Figure 4. Graphical comparison of the donor–acceptor bond distances between X-ray data (dotted lines) and the respective DFT methods (B3LYP*, OPBE, and r2SCAN) for the high-spin state of 1.

The analysis of the electronic energy differences between the HS and LS isomers revealed positive values of ΔEel = EelHSEelLS for all three DFT functionals, and thus these functionals properly found the LS state with lower electronic energy (Eel), Table S4. As the molecular vibrations have a significant impact on the SCO properties, a better description is achieved with the energy difference corrected by the zero-temperature vibrational energy from the frequency calculation, ΔEel+ZPE, which is depicted in Figure 5.

Figure 5

Figure 5. Graphical comparison the energy difference corrected by the zero-temperature vibrational energy from the frequency calculation ΔEel+ZPE between the HS and LS states for different DFT methods applied to 1.

Apparently, ΔEel+ZPE is significantly affected by applying the implicit solvation model, and both B3LYP*and r2SCAN provided reasonable values of the HS-LS separation. Moreover, it seems that a more polar solvent like CH3OH tends to increase ΔEel+ZPE, and thus it stabilizes the LS state. This is in contradiction to the experimental finding (Figure 3 and Table S3), for which higher Φ(CH3OH) yielded lower Δ and T1/2 values. We can speculate that this discrepancy is caused by the implicit solvation approach which cannot grasp the effect of all intermolecular interactions properly, or the properties of 1 in the solid state simply cannot be encompassed by such an approach at all.

However, if the presented theoretical calculations are not deceptive, perhaps it most likely leads to the conclusion that the reported phenomenon is not governed by thermodynamics, but by the kinetics of the crystal growth of 1 under various contents of chloroform and methanol in crystallizing solutions. Hence, the quality of the crystalline material and SCO properties are affected by a solvent mixture, but the solvent molecules do not cocrystallize. One of the possible approaches to test this hypothesis is to correlate parameters such as crystal mosaicity with the observed magnetic behavior. Therefore, we decided to prepare several batches of 1 crystallized at different crystallization rates. As was mentioned above, the best crystals were obtained in the batch of 1_Φ0.25. However, the batches crystallized from solutions with a major chloroform fraction (Φ0.25 and Φ0.05) have practically the same magnetic properties (Figure 3). Thus, we decided to crystallize 1 from Φ0.5 and Φ0.83 solutions, because 1_Φ0.5 and 1_Φ0.83 differ in T1/2↓ (114 K for 1_Φ0.5 and 103 K for 1_Φ0.83), and their spin transition on cooling is still accessible using standard commercial cryogenic devices (above 80 K). For preparation of the batches, we used the same reaction procedures as are described in Supporting Information, but the mother liquors were crystallized using three different crystallization rates: fast (≈1day), slow (up to 4 days), and very slow (≥7 days). The majority of the used single crystals were obtained for all the batches crystallized from Φ0.5 solutions regardless of the rate of crystallization used. However, also very slow crystallization from the Φ0.83 solutions resulted in the production of a few suitable single crystals.

The crystal mosaicity was determined taking into consideration the results of previous work on mosaicity in SCO complexes. (24) We set experimental conditions to be as identical as possible for all the investigated specimens. The experiment was conducted at 150 K, which is well above the highest T1/2↓ (Table S3). The data were collected using ω-scans (width 0.5 deg), and the exposition time was adjusted for each crystal based on its size, aiming for 0.83 Å resolution, completeness above 99%, data redundancy > 3, and I/σ > 10. As was already mentioned, the crystals of 1 are very thin (typically ≈0.05 mm in two dimensions), needle-like shaped growing in clumps of overlapping specimens, which makes it challenging to investigate the statistically relevant number of crystallites within a reasonable measurement time. Thus, we were capable of performing measurements for a limited number (14) of single crystals. It is important to note that the vast majority of the prepared crystals were not suitable for single-crystal experiments, and thus the results obtained only represent those crystals that met the necessary criteria.

After collecting each set of data at 150 K, we determined T1/2↓ for each crystal by measuring unit cell parameters starting from 125 K down to 80 K in decrements of 5 K. The occurrence of the SCO phenomenon was recorded between two measured temperature points, and the average of these values was used for visualization purposes. The T1/2↓ value of crystals that remained in the high spin phase at 80 K was set to 70 K for visualization purposes. The distribution of the determined T1/2↓ values is shown (Figure 6A). The results supported our hypothesis that the speed of crystallization affects the quality of crystals and the T1/2↓ value. Crystals crystallized in the ”fast” and “slow” modes showed T1/2↓ values above 100 K. Crystals that crystallized “very slowly” showed T1/2↓ values below 90 K with most of them having T1/2↓ even below 80 K.

Figure 6

Figure 6. Plot of T1/2↓ versus crystallization rate (A). The colored boxes are used to highlight the crystallization rate, with a lighter color indicating slower crystallization. The plots of Rint vs T1/2↓ (B), e3 vs T1/2↓ (C), and eavg vs T1/2↓ (D). The value of eavg was calculated as the arithmetic average of components e1, e2, and e3. In all plots, the T1/2↓ value of crystals that remained in the high spin phase down to 80 K was assigned as 70 K for visualization purposes.

The results of the X-ray diffraction measurements were evaluated in the CrysAlisPro software. (25) As a first indication of the crystal quality, we inspected the equivalency of symmetry equivalent reflections (Rint) in the studied crystals. There is no strong correlation between T1/2↓ and Rint, but apparently, the crystals with lower T1/2↓ values tend to have larger Rint and vice versa (Figure 6B).

The CrysAlisPro software provides the mosaicity in three directions (e1, e2, and e3) by fitting a Gaussian function to the peak. (26,27) Although these parameters do not measure the mosaicity directly, variations in the values of e1, e2, and e3 can indicate changes in the mosaicity of the studied crystals and thus their quality. (28) Some authors only use the e3 parameter due to its correlation with mosaicity values obtained from other data integration methods. (26) The results revealed that the e2 component varied slightly among the measurements (0.76–0.96°), while larger variations were observed for e1 (0.75–2.61°) and e3 (0.71–1.66°). As per previous reports, for evaluation of mosaicity, we considered the values of e3 and the arithmetic average (eavg) (29) of all three components e1, e2, and e3 (Figure 6C–D, Table S5). As with Rint, the results for eavg and e3 did not show a strong linear correlation; however, crystals with higher T1/2↓ values tend to have lower values of e and e3 and vice versa.

In summary, to the best of our knowledge, such an unprecedented solvent-induced variation of SCO spin-transition temperature as was discussed for 1 has not been reported yet. We showed that the different crystallization rates produced crystals exhibiting different SCO critical temperatures. Very slow crystallization (≥7 days) of the Φ0.5 and Φ0.83 solutions resulted into crystals exhibiting low SCO critical temperatures (T1/2↓ < 90 K), while faster crystallization (shorter than 4 days) led to larger T1/2↓ values (>100 K). The performed diffraction experiments indicate that crystals with low T1/2↓ values tend to have larger mosaicity parameters and Rint, and thus they are of lower quality than those with larger T1/2↓ values.

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.cgd.2c01411.

  • Synthesis, crystallographic data, magnetic measurements, DFT calculation details, and powder diffraction patterns (PDF)

Accession Codes

CCDC 1909057 and 22234622223463 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 Authors
  • Author
  • Author Contributions

    The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript.

  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

The authors (I.N. and R.H.) acknowledge financial support from institutional sources of the Department of Inorganic Chemistry and Palacký University Olomouc, Czech Republic. We also acknowledge the CzechNanoLab Research Infrastructure supported by MEYS CR (LM2018110) for the measurement of some of the magnetic data. I.N. is grateful to Jakub Wojciechowski for discussions about mosaicity evaluation in CrysAlisPro software.

References

Click to copy section linkSection link copied!

This article references 29 other publications.

  1. 1
    Bari, R. A.; Sivardière, J. Low-Spin-High-Spin Transitions in Transition-Metal-Ion Compounds. Phys. Rev. B 1972, 5 (11), 44664471,  DOI: 10.1103/PhysRevB.5.4466
  2. 2
    Wajnflasz, J. Etude de La Transition “Low Spin”-“High Spin” Dans Les Complexes Octaédriques d’ion de Transition. Phys. status solidi 1970, 40 (2), 537545,  DOI: 10.1002/pssb.19700400212
  3. 3
    Boča, R.; Linert, W. Is There a Need for New Models of the Spin Crossover? In Molecular Magnets Recent Highlights; Springer Vienna: Vienna, 2002; pp 83100 DOI: 10.1007/978-3-7091-6018-3_6 .
  4. 4
    Shakirova, O. G.; Lavrenova, L. G.; Kurat’eva, N. V.; Naumov, D. Y.; Daletskii, V. A.; Sheludyakova, L. A.; Logvinenko, V. A.; Vasilevskii, S. F. Spin Crossover in Iron(II) Complexes with Tris(Pyrazol-1-Yl)Methane. Russ. J. Coord. Chem. 2010, 36 (4), 275283,  DOI: 10.1134/S1070328410040068
  5. 5
    Halcrow, M. A. Structure:Function Relationships in Molecular Spin-Crossover Complexes. Chem. Soc. Rev. 2011, 40 (7), 4119,  DOI: 10.1039/c1cs15046d
  6. 6
    Roubeau, O. Triazole-Based One-Dimensional Spin-Crossover Coordination Polymers. Chem. - A Eur. J. 2012, 18 (48), 1523015244,  DOI: 10.1002/chem.201201647
  7. 7
    Arcis-Castíllo, Z.; Zheng, S.; Siegler, M. A.; Roubeau, O.; Bedoui, S.; Bonnet, S. Tuning the Transition Temperature and Cooperativity of Bapbpy-Based Mononuclear Spin-Crossover Compounds: Interplay between Molecular and Crystal Engineering. Chem. - A Eur. J. 2011, 17 (52), 1482614836,  DOI: 10.1002/chem.201101301
  8. 8
    Feltham, H. L. C.; Johnson, C.; Elliott, A. B. S.; Gordon, K. C.; Albrecht, M.; Brooker, S. “Tail” Tuning of Iron(II) Spin Crossover Temperature by 100 K. Inorg. Chem. 2015, 54 (6), 29022909,  DOI: 10.1021/ic503040f
  9. 9
    Askew, J. H.; Shepherd, H. J. Post-Synthetic Anion Exchange in Iron(II) 1,2,4-Triazole Based Spin Crossover Materials via Mechanochemistry. Dalt. Trans. 2020, 49 (9), 29662971,  DOI: 10.1039/C9DT04700J
  10. 10
    Krüger, C.; Augustín, P.; Dlhán̆, L.; Pavlik, J.; Moncol’, J.; Nemec, I.; Boča, R.; Renz, F. Iron(III) Complexes with Pentadentate Schiff-Base Ligands: Influence of Crystal Packing Change and Pseudohalido Coligand Variations on Spin Crossover. Polyhedron 2015, 87, 194201,  DOI: 10.1016/j.poly.2014.11.014
  11. 11
    Nemec, I.; Herchel, R.; Trávníček, Z. The Relationship between the Strength of Hydrogen Bonding and Spin Crossover Behaviour in a Series of Iron(III) Schiff Base Complexes. Dalt. Trans. 2015, 44 (10), 44744484,  DOI: 10.1039/C4DT03400G
  12. 12
    Rajnak, C.; Mičová, R.; Moncol, J.; Dlháň, L.; Krüger, C.; Renz, F.; Boča, R. Spin-Crossover in an Iron(III) Complex Showing a Broad Thermal Hysteresis. Dalton Trans. 2021, 50 (2), 472475,  DOI: 10.1039/D0DT03610B
  13. 13
    Boča, R.; Boča, M.; Dlhán̆, L.; Falk, K.; Fuess, H.; Haase, W.; Jaroščiak, R.; Papánková, B.; Renz, F.; Vrbová, M.; Werner, R. Strong Cooperativeness in the Mononuclear Iron(II) Derivative Exhibiting an Abrupt Spin Transition above 400 K. Inorg. Chem. 2001, 40 (13), 30253033,  DOI: 10.1021/ic000807s
  14. 14
    Neese, F. Software Update: The ORCA Program System - Version 5.0. WIREs Comput. Mol. Sci. 2022, 12, e1606 DOI: 10.1002/wcms.1606 .
  15. 15
    Cirera, J.; Ruiz, E. Assessment of the SCAN Functional for Spin-State Energies in Spin-Crossover Systems. J. Phys. Chem. A 2020, 124 (24), 50535058,  DOI: 10.1021/acs.jpca.0c03758
  16. 16
    Cirera, J.; Via-Nadal, M.; Ruiz, E. Benchmarking Density Functional Methods for Calculation of State Energies of First Row Spin-Crossover Molecules. Inorg. Chem. 2018, 57 (22), 1409714105,  DOI: 10.1021/acs.inorgchem.8b01821
  17. 17
    Siig, O. S.; Kepp, K. P. Iron(II) and Iron(III) Spin Crossover: Toward an Optimal Density Functional. J. Phys. Chem. A 2018, 122 (16), 42084217,  DOI: 10.1021/acs.jpca.8b02027
  18. 18
    Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77 (18), 38653868,  DOI: 10.1103/PhysRevLett.77.3865
  19. 19
    Furness, J. W.; Kaplan, A. D.; Ning, J.; Perdew, J. P.; Sun, J. Accurate and Numerically Efficient r 2 SCAN Meta-Generalized Gradient Approximation. J. Phys. Chem. Lett. 2020, 11 (19), 82088215,  DOI: 10.1021/acs.jpclett.0c02405
  20. 20
    Reiher, M. Theoretical Study of the Fe(Phen)2(NCS)2 Spin-Crossover Complex with Reparametrized Density Functionals. Inorg. Chem. 2002, 41 (25), 69286935,  DOI: 10.1021/ic025891l
  21. 21
    Caldeweyher, E.; Ehlert, S.; Hansen, A.; Neugebauer, H.; Spicher, S.; Bannwarth, C.; Grimme, S. A Generally Applicable Atomic-Charge Dependent London Dispersion Correction. J. Chem. Phys. 2019, 150 (15), 154122,  DOI: 10.1063/1.5090222
  22. 22
    Garcia-Ratés, M.; Neese, F. Effect of the Solute Cavity on the Solvation Energy and Its Derivatives within the Framework of the Gaussian Charge Scheme. J. Comput. Chem. 2020, 41 (9), 922939,  DOI: 10.1002/jcc.26139
  23. 23
    Barone, V.; Cossi, M. Quantum Calculation of Molecular Energies and Energy Gradients in Solution by a Conductor Solvent Model. J. Phys. Chem. A 1998, 102 (11), 19952001,  DOI: 10.1021/jp9716997
  24. 24
    Lakhloufi, S.; Tailleur, E.; Guo, W.; Le Gac, F.; Marchivie, M.; Lemée-Cailleau, M.-H.; Chastanet, G.; Guionneau, P. Mosaicity of Spin-Crossover Crystals. Crystals 2018, 8 (9), 363,  DOI: 10.3390/cryst8090363
  25. 25
    CrysAlisPro, 1.171.42.49; Rigaku Oxford Diffraction, 2020.
  26. 26
    Harrison, K.; Wu, Z.; Juers, D. H. A Comparison of Gas Stream Cooling and Plunge Cooling of Macromolecular Crystals. J. Appl. Crystallogr. 2019, 52 (5), 12221232,  DOI: 10.1107/S1600576719010318
  27. 27
    Kabsch, W. Integration, Scaling, Space-Group Assignment and Post-Refinement. Acta Crystallogr. Sect. D Biol. Crystallogr. 2010, 66 (2), 133144,  DOI: 10.1107/S0907444909047374
  28. 28
    Madsen, S. R.; Overgaard, J.; Stalke, D.; Iversen, B. B. High-Pressure Single Crystal X-Ray Diffraction Study of the Linear Metal Chain Compound Co3(Dpa)4Br2·CH2Cl2. Dalt. Trans. 2015, 44 (19), 90389043,  DOI: 10.1039/C5DT00447K
  29. 29
    Farley, C.; Burks, G.; Siegert, T.; Juers, D. H. Improved Reproducibility of Unit-Cell Parameters in Macromolecular Cryocrystallography by Limiting Dehydration during Crystal Mounting. Acta Crystallogr. Sect. D Biol. Crystallogr. 2014, 70 (8), 21112124,  DOI: 10.1107/S1399004714012310

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 1 publications.

  1. Theerapoom Boonprab, Warisa Thammasangwan, Guillaume Chastanet, Mathieu Gonidec, Phimphaka Harding, David J. Harding. Halide Anion Effects and Magnetostructural Relationships in Iron(III) Spin Crossover Complexes. Crystal Growth & Design 2024, 24 (19) , 8145-8152. https://doi.org/10.1021/acs.cgd.4c01068

Crystal Growth & Design

Cite this: Cryst. Growth Des. 2023, 23, 3, 1323–1329
Click to copy citationCitation copied!
https://doi.org/10.1021/acs.cgd.2c01411
Published February 16, 2023

Copyright © 2023 American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Article Views

997

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. Temperature dependence of μeffB for 1_Φ1 and 1_Φ1p.

    Figure 2

    Figure 2. Temperature dependence of μeffB for 1_Φ0.25 (top) and the temperature dependence of the selected unit cell parameters as determined from the single crystal X-ray diffraction experiment for 1_Φ0.25 (bottom).

    Figure 3

    Figure 3. Temperature dependence the high-spin molar fraction xHS according to ISM (top), the ISM parameters (middle), and transition temperatures (bottom) for crystalline samples of 1.

    Figure 4

    Figure 4. Graphical comparison of the donor–acceptor bond distances between X-ray data (dotted lines) and the respective DFT methods (B3LYP*, OPBE, and r2SCAN) for the high-spin state of 1.

    Figure 5

    Figure 5. Graphical comparison the energy difference corrected by the zero-temperature vibrational energy from the frequency calculation ΔEel+ZPE between the HS and LS states for different DFT methods applied to 1.

    Figure 6

    Figure 6. Plot of T1/2↓ versus crystallization rate (A). The colored boxes are used to highlight the crystallization rate, with a lighter color indicating slower crystallization. The plots of Rint vs T1/2↓ (B), e3 vs T1/2↓ (C), and eavg vs T1/2↓ (D). The value of eavg was calculated as the arithmetic average of components e1, e2, and e3. In all plots, the T1/2↓ value of crystals that remained in the high spin phase down to 80 K was assigned as 70 K for visualization purposes.

  • References


    This article references 29 other publications.

    1. 1
      Bari, R. A.; Sivardière, J. Low-Spin-High-Spin Transitions in Transition-Metal-Ion Compounds. Phys. Rev. B 1972, 5 (11), 44664471,  DOI: 10.1103/PhysRevB.5.4466
    2. 2
      Wajnflasz, J. Etude de La Transition “Low Spin”-“High Spin” Dans Les Complexes Octaédriques d’ion de Transition. Phys. status solidi 1970, 40 (2), 537545,  DOI: 10.1002/pssb.19700400212
    3. 3
      Boča, R.; Linert, W. Is There a Need for New Models of the Spin Crossover? In Molecular Magnets Recent Highlights; Springer Vienna: Vienna, 2002; pp 83100 DOI: 10.1007/978-3-7091-6018-3_6 .
    4. 4
      Shakirova, O. G.; Lavrenova, L. G.; Kurat’eva, N. V.; Naumov, D. Y.; Daletskii, V. A.; Sheludyakova, L. A.; Logvinenko, V. A.; Vasilevskii, S. F. Spin Crossover in Iron(II) Complexes with Tris(Pyrazol-1-Yl)Methane. Russ. J. Coord. Chem. 2010, 36 (4), 275283,  DOI: 10.1134/S1070328410040068
    5. 5
      Halcrow, M. A. Structure:Function Relationships in Molecular Spin-Crossover Complexes. Chem. Soc. Rev. 2011, 40 (7), 4119,  DOI: 10.1039/c1cs15046d
    6. 6
      Roubeau, O. Triazole-Based One-Dimensional Spin-Crossover Coordination Polymers. Chem. - A Eur. J. 2012, 18 (48), 1523015244,  DOI: 10.1002/chem.201201647
    7. 7
      Arcis-Castíllo, Z.; Zheng, S.; Siegler, M. A.; Roubeau, O.; Bedoui, S.; Bonnet, S. Tuning the Transition Temperature and Cooperativity of Bapbpy-Based Mononuclear Spin-Crossover Compounds: Interplay between Molecular and Crystal Engineering. Chem. - A Eur. J. 2011, 17 (52), 1482614836,  DOI: 10.1002/chem.201101301
    8. 8
      Feltham, H. L. C.; Johnson, C.; Elliott, A. B. S.; Gordon, K. C.; Albrecht, M.; Brooker, S. “Tail” Tuning of Iron(II) Spin Crossover Temperature by 100 K. Inorg. Chem. 2015, 54 (6), 29022909,  DOI: 10.1021/ic503040f
    9. 9
      Askew, J. H.; Shepherd, H. J. Post-Synthetic Anion Exchange in Iron(II) 1,2,4-Triazole Based Spin Crossover Materials via Mechanochemistry. Dalt. Trans. 2020, 49 (9), 29662971,  DOI: 10.1039/C9DT04700J
    10. 10
      Krüger, C.; Augustín, P.; Dlhán̆, L.; Pavlik, J.; Moncol’, J.; Nemec, I.; Boča, R.; Renz, F. Iron(III) Complexes with Pentadentate Schiff-Base Ligands: Influence of Crystal Packing Change and Pseudohalido Coligand Variations on Spin Crossover. Polyhedron 2015, 87, 194201,  DOI: 10.1016/j.poly.2014.11.014
    11. 11
      Nemec, I.; Herchel, R.; Trávníček, Z. The Relationship between the Strength of Hydrogen Bonding and Spin Crossover Behaviour in a Series of Iron(III) Schiff Base Complexes. Dalt. Trans. 2015, 44 (10), 44744484,  DOI: 10.1039/C4DT03400G
    12. 12
      Rajnak, C.; Mičová, R.; Moncol, J.; Dlháň, L.; Krüger, C.; Renz, F.; Boča, R. Spin-Crossover in an Iron(III) Complex Showing a Broad Thermal Hysteresis. Dalton Trans. 2021, 50 (2), 472475,  DOI: 10.1039/D0DT03610B
    13. 13
      Boča, R.; Boča, M.; Dlhán̆, L.; Falk, K.; Fuess, H.; Haase, W.; Jaroščiak, R.; Papánková, B.; Renz, F.; Vrbová, M.; Werner, R. Strong Cooperativeness in the Mononuclear Iron(II) Derivative Exhibiting an Abrupt Spin Transition above 400 K. Inorg. Chem. 2001, 40 (13), 30253033,  DOI: 10.1021/ic000807s
    14. 14
      Neese, F. Software Update: The ORCA Program System - Version 5.0. WIREs Comput. Mol. Sci. 2022, 12, e1606 DOI: 10.1002/wcms.1606 .
    15. 15
      Cirera, J.; Ruiz, E. Assessment of the SCAN Functional for Spin-State Energies in Spin-Crossover Systems. J. Phys. Chem. A 2020, 124 (24), 50535058,  DOI: 10.1021/acs.jpca.0c03758
    16. 16
      Cirera, J.; Via-Nadal, M.; Ruiz, E. Benchmarking Density Functional Methods for Calculation of State Energies of First Row Spin-Crossover Molecules. Inorg. Chem. 2018, 57 (22), 1409714105,  DOI: 10.1021/acs.inorgchem.8b01821
    17. 17
      Siig, O. S.; Kepp, K. P. Iron(II) and Iron(III) Spin Crossover: Toward an Optimal Density Functional. J. Phys. Chem. A 2018, 122 (16), 42084217,  DOI: 10.1021/acs.jpca.8b02027
    18. 18
      Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77 (18), 38653868,  DOI: 10.1103/PhysRevLett.77.3865
    19. 19
      Furness, J. W.; Kaplan, A. D.; Ning, J.; Perdew, J. P.; Sun, J. Accurate and Numerically Efficient r 2 SCAN Meta-Generalized Gradient Approximation. J. Phys. Chem. Lett. 2020, 11 (19), 82088215,  DOI: 10.1021/acs.jpclett.0c02405
    20. 20
      Reiher, M. Theoretical Study of the Fe(Phen)2(NCS)2 Spin-Crossover Complex with Reparametrized Density Functionals. Inorg. Chem. 2002, 41 (25), 69286935,  DOI: 10.1021/ic025891l
    21. 21
      Caldeweyher, E.; Ehlert, S.; Hansen, A.; Neugebauer, H.; Spicher, S.; Bannwarth, C.; Grimme, S. A Generally Applicable Atomic-Charge Dependent London Dispersion Correction. J. Chem. Phys. 2019, 150 (15), 154122,  DOI: 10.1063/1.5090222
    22. 22
      Garcia-Ratés, M.; Neese, F. Effect of the Solute Cavity on the Solvation Energy and Its Derivatives within the Framework of the Gaussian Charge Scheme. J. Comput. Chem. 2020, 41 (9), 922939,  DOI: 10.1002/jcc.26139
    23. 23
      Barone, V.; Cossi, M. Quantum Calculation of Molecular Energies and Energy Gradients in Solution by a Conductor Solvent Model. J. Phys. Chem. A 1998, 102 (11), 19952001,  DOI: 10.1021/jp9716997
    24. 24
      Lakhloufi, S.; Tailleur, E.; Guo, W.; Le Gac, F.; Marchivie, M.; Lemée-Cailleau, M.-H.; Chastanet, G.; Guionneau, P. Mosaicity of Spin-Crossover Crystals. Crystals 2018, 8 (9), 363,  DOI: 10.3390/cryst8090363
    25. 25
      CrysAlisPro, 1.171.42.49; Rigaku Oxford Diffraction, 2020.
    26. 26
      Harrison, K.; Wu, Z.; Juers, D. H. A Comparison of Gas Stream Cooling and Plunge Cooling of Macromolecular Crystals. J. Appl. Crystallogr. 2019, 52 (5), 12221232,  DOI: 10.1107/S1600576719010318
    27. 27
      Kabsch, W. Integration, Scaling, Space-Group Assignment and Post-Refinement. Acta Crystallogr. Sect. D Biol. Crystallogr. 2010, 66 (2), 133144,  DOI: 10.1107/S0907444909047374
    28. 28
      Madsen, S. R.; Overgaard, J.; Stalke, D.; Iversen, B. B. High-Pressure Single Crystal X-Ray Diffraction Study of the Linear Metal Chain Compound Co3(Dpa)4Br2·CH2Cl2. Dalt. Trans. 2015, 44 (19), 90389043,  DOI: 10.1039/C5DT00447K
    29. 29
      Farley, C.; Burks, G.; Siegert, T.; Juers, D. H. Improved Reproducibility of Unit-Cell Parameters in Macromolecular Cryocrystallography by Limiting Dehydration during Crystal Mounting. Acta Crystallogr. Sect. D Biol. Crystallogr. 2014, 70 (8), 21112124,  DOI: 10.1107/S1399004714012310
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.cgd.2c01411.

    • Synthesis, crystallographic data, magnetic measurements, DFT calculation details, and powder diffraction patterns (PDF)

    Accession Codes

    CCDC 1909057 and 22234622223463 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.