Widening the Window of Spin-Crossover Temperatures in Bis(formazanate)iron(II) Complexes via Steric and Noncovalent InteractionsClick to copy article linkArticle link copied!
- Francesca MiloccoFrancesca MiloccoStratingh Institute for Chemistry, University of Groningen, Nijenborgh 4, 9747 AG Groningen, The NetherlandsMore by Francesca Milocco
- Folkert de VriesFolkert de VriesStratingh Institute for Chemistry, University of Groningen, Nijenborgh 4, 9747 AG Groningen, The NetherlandsMore by Folkert de Vries
- Harmke S. SiebeHarmke S. SiebeStratingh Institute for Chemistry, University of Groningen, Nijenborgh 4, 9747 AG Groningen, The NetherlandsMore by Harmke S. Siebe
- Silène EngbersSilène EngbersStratingh Institute for Chemistry, University of Groningen, Nijenborgh 4, 9747 AG Groningen, The NetherlandsMore by Silène Engbers
- Serhiy DemeshkoSerhiy DemeshkoInstitut für Anorganische Chemie, Universität Göttingen, Tammannstraße 4, 37077 Göttingen, GermanyMore by Serhiy Demeshko
- Franc MeyerFranc MeyerInstitut für Anorganische Chemie, Universität Göttingen, Tammannstraße 4, 37077 Göttingen, GermanyMore by Franc Meyer
- Edwin Otten*Edwin Otten*Email [email protected]Stratingh Institute for Chemistry, University of Groningen, Nijenborgh 4, 9747 AG Groningen, The NetherlandsMore by Edwin Otten
Abstract
Bis(formazanate)iron(II) complexes undergo a thermally induced S = 0 to S = 2 spin transition in solution. Here we present a study of how steric effects and π-stacking interactions between the triarylformazanate ligands affect the spin-crossover behavior, in addition to electronic substituent effects. Moreover, the effect of increasing the denticity of the formazanate ligands is explored by including additional OMe donors in the ligand (7). In total, six new compounds (2–7) have been synthesized and characterized, both in solution and in the solid state, via spectroscopic, magnetic, and structural analyses. The series spans a broad range of spin-crossover temperatures (T1/2) for the LS ⇌ HS equilibrium in solution, with the exception of compound 6 which remains high-spin (S = 2) down to 210 K. In the solid state, 6 was shown to exist in two distinct forms: a tetrahedral high-spin complex (6a, S = 2) and a rare square-planar structure with an intermediate-spin state (6b, S = 1). SQUID measurements, 57Fe Mössbauer spectroscopy, and differential scanning calorimetry indicate that in the solid state the square-planar form 6b undergoes an incomplete spin-change-coupled isomerization to tetrahedral 6a. The complex that contains additional OMe donors (7) results in a six-coordinate (NNO)2Fe coordination geometry, which shifts the spin-crossover to significantly higher temperatures (T1/2 = 444 K). The available experimental and computational data for 7 suggest that the Fe···OMe interaction is retained upon spin-crossover. Despite the difference in coordination environment, the weak OMe donors do not significantly alter the electronic structure or ligand-field splitting, and the occurrence of spin-crossover (similar to the compounds lacking the OMe groups) originates from a large degree of metal–ligand π-covalency.
Synopsis
A series of Fe(II) complexes with formazanate ligands are reported, and ligand substituent effects on structure and spin-crossover properties are examined. These ligand modifications allow isolation of compounds with tetrahedral geometries in both low- and high-spin ground states as well as an intermediate-spin square-planar complex. Steric properties, π-stacking interactions, and additional donor substituents lead to a wide range of spin-crossover temperatures (T1/2) in this class of compounds.
Introduction
Figure 1
Figure 1. Common ligand field splitting diagrams for octahedral (A), square-planar (B), and tetrahedral (C) geometries and unusual ligand field splitting for the pseudo-tetrahedral geometries found in bis(formazanate)iron(II) complexes (D). (14)
Results and Discussion
Scheme 1
Solid-State Characterization
Figure 2
Figure 2. Crystal structure of compound 3 showing the π-stacking interactions between the mesityl rings. The Fe center, ligand backbone, and the mesityl rings are shown as 50% probability ellipsoids and the remaining atoms as wireframe; hydrogen atoms are removed for clarity.
1a | 3 | 4 | 6ab | 6b | 7 | |
---|---|---|---|---|---|---|
Fe(1)–N(1) | 1.8278(15) | 1.8192(14) | 1.9946(11) | 2.030(2) | 1.9259(9) | 1.877(2) |
Fe(1)–N(4) | 1.8207(15) | 1.8351(13) | 1.9610(12) | 1.9851(19) | 1.9461(9) | 1.883(2) |
Fe(1)–N(5) | 1.8330(16) | 1.8242(13) | 1.9864(12) | 2.035(2) | 1.874(2) | |
Fe(1)–N(8) | 1.8174(16) | 1.8449(13) | 1.9616(12) | 1.9966(19) | 1.895(2) | |
Fe(1)–O(1) | 2.1128(18) | |||||
Fe(1)–O(2) | 2.1029(19) | |||||
∠(NFeN)/(NFeN)c | 60.97(10) | 64.06(9) | 83.21(7) | 89.31(12) | 0.00(0) | 81.67(14) |
Fe out-of-planed | 0.001 | 0.018 | 0.116 | 0.582 | 0.700 | 0.220 |
0.046 | 0.119 | 0.116 | 0.580 | 0.224 |
Data taken from ref (14a).
Structure measured at 200 K.
Dihedral angle between the coordination planes defined by the N–Fe–N atoms.
Displacement of the Fe atom out of the plane defined by the four N atoms of each ligand backbone.
Figure 3
Figure 3. Molecular structure of compounds 6a and 6b showing 50% probability ellipsoids; hydrogen atoms omitted for clarity. The inset for each shows the Fe(NNCNN)2 core of the structure with the N–Fe–N planes and the dihedral angle.
Figure 4
Figure 4. Crystal structure of compound 6b illustrating the π-stacking interactions between the aromatic rings, showing 50% probability ellipsoids. Parts of the molecule are shown as wireframe, and hydrogen atoms are removed for clarity.
Figure 5
Figure 5. 57Fe Mössbauer spectra at 80 K in the solid state of 6b (A), 6a (B), and a powder sample of 6 before (C) and after (D) heating to 400 K for SQUID measurements. The red line in the spectrum of heated 6 represents the main species with 82% area, and the gray subspectra are unknown impurities.
Figure 6
Figure 6. Magnetic susceptibility data for a powder sample of 6 in the solid state (heating to 400 K and subsequent cooling). The solid black line shows the best fit curve for S = 1 with the parameters g = 2.10 and D = 11.2 cm–1 (100% IS). The dashed red line shows the spin-only value for an S = 2 system.
1 | 3 | 6a | 6b | 6b | |
---|---|---|---|---|---|
δ | 0.03 | 0.05 | 0.75 | 0.54 | 0.55 |
|ΔEq| | 2.05 | 1.99 | 1.21 | 2.73 | 2.72 |
Measured in the solid state at 80 K.
Powder sample of the crude product before crystallization.
Figure 7
Figure 7. Molecular structures of 7 showing 50% probability ellipsoids. One of the N–Ph rings is shown as wireframe, and hydrogen atoms are omitted for clarity.
Variable-Temperature NMR and UV/Vis Spectroscopy in Solution
Figure 8
Figure 8. Temperature dependence of the high-spin fraction (γHS) of compounds 1–5 and 7 in toluene-d8, including error bars for T1/2 (γHS = 0.5). The liquid range for toluene is indicated with the color gradient at the temperature axis.
1a | 2b | 3 | 4b | 5 | 6 | 7b | |
---|---|---|---|---|---|---|---|
ΔH (kJ mol–1) | 22.2 ± 0.3 | 8.5 ± 0.4 | 26.3 ± 0.1 | 12.6 ± 1.0 | 19.0 ± 0.4 | – | 37.5 ± 1.6 |
ΔS (J mol–1 K–1) | 64 ± 1 | 45 ± 4 | 78 ± 1 | 67 ± 5 | 70 ± 1 | – | 85 ± 5 |
T1/2c (K) | 345 ± 7 | 192 ± 18 | 340 ± 2 | 188 ± 21 | 271 ± 8 | – | 444 ± 34 |
Data reproduced from ref (14a).
Estimated from fitting a limited temperature range.
The uncertainty in T1/2 is obtained by using error propagation from ΔH and ΔS.
Figure 9
Figure 9. 1H NMR spectra of 7 recorded between 247 and 397 K (toluene-d8, 500 MHz).
Figure 10
Figure 10. Representation of intrinsic bonding orbitals at the BP86/def2-TZVP minima, both with (7calc; (A)) and without Fe–O interaction (7′calc; (B)).
Figure 11
Figure 11. UV/vis spectra in toluene for (A) compound 6 recorded between 183 and 293 K and (B) compound 7 recorded between 293 and 383 K.
Conclusions
Experimental Section
General Considerations
Synthesis of Fe[PhNNC(C6F5)NNPh]2 (2)
Synthesis of Fe[PhNNC(p-Tol)NNMes]2 (3)
Synthesis of Fe[PhNNC(C6F5)NNMes]2 (4)
Synthesis of Fe[C6F5NNC(p-Tol)NNMes]2 (5)
Synthesis of Fe[C6F5NNC(C6F5)NNMes]2 (6)
Synthesis of Fe[PhNNC(p-Tol)NN(o-An)]2·0.5(THF) (7)
X-ray Crystallography
3 | 4 | 6a | 6b | 7 | |
---|---|---|---|---|---|
chem formula | C46H46N8Fe | C50H46F10FeN8 | C44H22F20FeN8 | C44H22F20FeN8 | C46H46FeN8O3 |
Mr | 766.76 | 1004.80 | 1098.54 | 1098.54 | 814.76 |
cryst syst | triclinic | triclinic | monoclinic | triclinic | monoclinic |
color, habit | red, block | red, block | brown, block | red, block | green, needle |
size (mm) | 0.23 × 0.18 × 0.11 | 0.42 × 0.26 × 0.09 | 0.30 × 0.17 × 0.15 | 0.34 × 0.20 × 0.09 | 0.49 × 0.06 × 0.02 |
space group | P-1 | P-1 | P21/c | P-1 | P21/c |
a (Å) | 8.5063(5) | 12.3467(8) | 12.7118(6) | 7.2786(6) | 12.0477(5) |
b (Å) | 11.3217(7) | 12.4167(8) | 20.5795(10) | 12.5743(10) | 24.0684(9) |
c (Å) | 21.5733(13) | 15.4105(10) | 17.4336(7) | 12.6798(9) | 14.6588(5) |
α (deg) | 93.547(2) | 86.354(2) | 90 | 118.595(2) | 90 |
β (deg) | 94.372(2) | 88.205(2) | 95.991(2) | 99.168(3) | 109.140(2) |
γ (deg) | 109.254(2) | 83.349(2) | 90 | 94.529(3) | 90 |
V (Å3) | 1947.3(2) | 2341.2(3) | 4535.8(4) | 989.67(13) | 4015.6(3) |
Z | 2 | 2 | 4 | 1 | 4 |
ρcalc, g cm–3 | 1.308 | 1.425 | 1.609 | 1.843 | 1.348 |
radiation [Å] | Mo Kα 0.71073 | Mo Kα 0.71073 | Mo Kα 0.71073 | Mo Kα 0.71073 | Cu Kα 1.54178 |
μ(Mo Kα), mm–1 | 0.432 | 0.407 | 0.458 | 0.525 | |
μ(Cu Kα), mm–1 | 3.433 | ||||
F(000) | 808 | 1036 | 2192 | 548 | 1712 |
temp (K) | 100(2) | 100(2) | 200(2) | 100(2) | 100(2) |
θ range (deg) | 2.76–27.16 | 3.04–27.92 | 2.88–26.38 | 2.88–27.94 | 3.68–65.14 |
data collected (h, k, l) | –10:10; –13:14; –27:27 | –16:16; –16:16; –20:20 | –15:15; –25:25; –21:21 | –9:9; –16:16; –16:16 | –14:14; –28:28; –17:16 |
no. of rflns collected | 60240 | 70899 | 54918 | 47685 | 32490 |
no. of indpndt collected | 8538 | 11211 | 9089 | 4755 | 6743 |
observed reflns Fo ≥ 2.0σ(Fo) | 7238 | 9251 | 6346 | 4557 | 5373 |
R(F) (%) | 3.46 | 3.36 | 4.32 | 2.55 | 4.61 |
wR(F2) (%) | 8.03 | 8.05 | 9.93 | 2.55 | 12.01 |
GooF | 1.045 | 1.040 | 1.030 | 7.07 | 1.037 |
weighting a, b | 0.0267, 1.5260 | 0.0311, 1.2677 | 0.0324, 3.4691 | 0.0372, 0.6125 | 0.0552, 2.9707 |
params refined | 504 | 687 | 664 | 334 | 527 |
min, max resid dens | –0.427, 0.295 | –0.280, 0.375 | –0.260, 0.264 | –0.451, 0.379 | –0.346, 0.689 |
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.inorgchem.0c03593.
Full experimental and characterization data, computational details (PDF)
CCDC 2036148–2036152 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.
Acknowledgments
Financial support from The Netherlands Organization for Scientific Research (NWO) is gratefully acknowledged (VIDI grant to E.O.). We thank the Center for Information Technology of the University of Groningen for their support and for providing access to the Peregrine high performance computing cluster as well as Prof. Wesley Browne and Dr. Johannes Klein (University of Groningen) for access to VT-UV/vis spectroscopy. S.D. and F.M. acknowledge support from the Universität Göttingen.
Additional Note
a Crystals of 7 contain one THF per iron complex, but drying results in loss of part of the THF solvate molecules.
References
This article references 35 other publications.
- 1(a) Poli, R. Open-Shell Organometallics as a Bridge between Werner-Type and Low-Valent Organometallic Complexes. The Effect of the Spin State on the Stability, Reactivity, and Structure. Chem. Rev. 1996, 96, 2135– 2204, DOI: 10.1021/cr9500343Google Scholar1aOpen-Shell Organometallics as a Bridge between Werner-Type and Low-Valent Organometallic Complexes. The Effect of the Spin State on the Stability, Reactivity, and StructurePoli, RinaldoChemical Reviews (Washington, D. C.) (1996), 96 (6), 2135-2204CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review with > 772 refs.(b) Hawrelak, E. J.; Bernskoetter, W. H.; Lobkovsky, E.; Yee, G. T.; Bill, E.; Chirik, P. J. Square planar vs tetrahedral geometry in four coordinate iron(II) complexes. Inorg. Chem. 2005, 44, 3103– 3111, DOI: 10.1021/ic048202+Google Scholar1bSquare Planar vs Tetrahedral Geometry in Four Coordinate Iron(II) ComplexesHawrelak, Eric J.; Bernskoetter, Wesley H.; Lobkovsky, Emil; Yee, Gordon T.; Bill, Eckhard; Chirik, Paul J.Inorganic Chemistry (2005), 44 (9), 3103-3111CODEN: INOCAJ; ISSN:0020-1669. (American Chemical Society)The geometric preferences of a family of four coordinate, iron(II) d6 complexes of the general form L2FeX2 have been systematically evaluated. Treatment of Fe2(Mes)4 (Mes = 2,4,6-Me3C6H2) with monodentate phosphine and phosphite ligands furnished square planar trans-P2Fe(Mes)2 derivs. Identification of the geometry has been accomplished by a combination of soln. and solid-state magnetometry and, in two cases (P = PMe3, PEt2Ph), x-ray diffraction. In contrast, both tetrahedral and square planar coordination has been obsd. upon complexation of chelating phosphine ligands. A combination of crystallog. and magnetic susceptibility data for (depe)Fe(Mes)2 (depe = 1,2-bis(diethylphosphino)ethane) established a tetrahedral mol. geometry whereas SQUID magnetometry and Moessbauer spectroscopy on samples of (dppe)Fe(Mes)2 (dppe = 1,2-bis(diphenylphosphino)ethane) indicated a planar mol. When dissolved in chlorinated solvents, the latter compd. promotes chlorine atom abstraction, forming tetrahedral (dppe)Fe(Mes)Cl and (dppe)FeCl2. Ligand substitution reactions have been studied for both structural types and are rapid on the NMR time scale at ambient temp.
- 2Hartwig, J. F. Organotransition Metal Chemistry: From Bonding to Catalysis; University Science Books: Sausalito, CA, 2010.Google ScholarThere is no corresponding record for this reference.
- 3Cirera, J.; Ruiz, E.; Alvarez, S. Stereochemistry and spin state in four-coordinate transition metal compounds. Inorg. Chem. 2008, 47, 2871– 2889, DOI: 10.1021/ic702276kGoogle Scholar3Stereochemistry and Spin State in Four-Coordinate Transition Metal CompoundsCirera, Jordi; Ruiz, Eliseo; Alvarez, SantiagoInorganic Chemistry (2008), 47 (7), 2871-2889CODEN: INOCAJ; ISSN:0020-1669. (American Chemical Society)A systematic DFT computational study of the stereochem. assocd. with each spin state for first transition series four-coordinate dn (n = 0-10) homoleptic metal complexes is presented. The stereochem. of [MMe4]x- complexes in the 21 spin configurations analyzed can be predicted from the d orbital occupation in the ideal tetrahedral geometry, grouped in three families with tetrahedral, square planar, or intermediate structures that can be described in some cases as sawhorses. The effect of the following factors on the spin state and stereochem. preferences has also been studied: (a) substitution of the σ-donor Me ligands by π-donor chlorides, (b) a high (+ 4) oxidn. state of the metal, and (c) substitution of the metal atom by a second transition series one. Through those factors, low-spin tetrahedral structures can be achieved, as summarized by a magic cube.
- 4(a) Collman, J. P.; Hoard, J. L.; Kim, N.; Lang, G.; Reed, C. A. Synthesis, stereochemistry, and structure-related properties of alpha, beta, gamma, delta-tetraphenylporphinatoiron(II). J. Am. Chem. Soc. 1975, 97, 2676– 2681, DOI: 10.1021/ja00843a015Google Scholar4aSynthesis, stereochemistry, and structure-related properties of α,β,γ,δ-tetraphenylporphinatoiron(II)Collman, James P.; Hoard, J. L.; Kim, Nancy; Lang, George; Reed, Christopher A.Journal of the American Chemical Society (1975), 97 (10), 2676-81CODEN: JACSAT; ISSN:0002-7863.Addnl. data considered in abstracting and indexing are available from a source cited in the original document. Redn. of chloro-α,β,γ,δ-tetraphenylporphinatoiron(III) by bis(acetylacetonato)chromium(II) under anaerobic conditions yielded α,β,γ,δ-tetraphenylporphinatoiron(II), Fe(TPP), which crystd. from benzene-EtOH soln. in the tetragonal system. Pertinent crystal data are: space group, I‾42d; α 15.080(9) and c 14.043(9) Å; d. (exptl.) = 1.38 and d. (calcd.) = 1.39 for Z = 4; required mol. symmetry S4. Intensities of 1164 reflections having (sin θ)/λ < 0.71 Å-1, recorded with Zr-filtered Mo Kα radiation on a Picker FACS-I diffractometer, were used in anisotropic full-matrix least-squares refinement of the 112 structural parameters. The length of the FeII-N bonds in the intermediate-spin (S = 1) Fe(TPP) mol. is 1.972(4) Å, as compared with 2.086(4) Å in high-spin (S = 2) 2-methylimidazole-α,β,γ,δ-tetraphenylporphinatoiron(II), with 2.004(3) Å in low-spin (S = 0) bis(piperidine)-α,β,γ,δ-tetraphenylporphinatoiron(II, and with Cu-N = 1.981(7) Å in Cu(TPP). The foregoing assignment of spin states to the 3 Fe(II) porphyrins is fully confirmed by the Moessbauer spectra recorded for them in zero and in large applied magnetic fields at 4.2-300°K. Recrystd. Fe(TPP) conforming to the theor. compn. has an effective magnetic moment of ∼4.4 μB at room temp. Detn. of the anisotropic magnetic susceptibilities as a function of temp. for the tetragonal crystal, when considered in conjunction with the Moessbauer data, should lead to a more detailed specification of the electronic configuration of the intermediate-spin iron(II) atom in Fe(TPP) than that given in this paper.(b) Kirner, J. F.; Dow, W.; Scheidt, W. R. Molecular stereochemistry of two intermediate-spin complexes. Iron(II) phthalocyanine and manganese(II) phthalocyanine. Inorg. Chem. 1976, 15, 1685– 1690, DOI: 10.1021/ic50161a042Google Scholar4bMolecular stereochemistry of two intermediate-spin complexes. Iron(II) phthalocyanine and manganese(II) phthalocyanineKirner, John F.; Dow, W.; Scheidt, W. RobertInorganic Chemistry (1976), 15 (7), 1685-90CODEN: INOCAJ; ISSN:0020-1669.The mol. stereochem. of Fe(II) phthalocyanine (Pc) and Mn(II) phthalocyanine was detd. by x-ray diffraction methods. The phthalocyanine ligand constrains the metal ion to effectively square-planar coordination and to an intermediate spin state. The FeII-N bond distance of 1.926(1) and the MnII-N bond length of 1.938(3) Å are wholly consistent with the assignment of an intermediate-spin ground state. Both complexes crystallize as the β polymorph. Crystal data are: for FePc, space group P21/a, a 19.392(5), b 4.786(2), c 14.604(4) Å, β 120.85(1)°, d.(exptl.) 1.61, d.(calcd.) 1.623, and Z = 2, required mol. symmetry ‾1; for MnPc, space group P21/a, a 19.400(4), b 4.761(2), c 14.613(3) Å, β 120.74(1)°, d.(exptl.) 1.61, d.(calcd.) 1.625, and Z = 2, required mol. symmetry ‾1. Final discrepancy indices are: FePc, R1 0.045, R2 0.057; MnPc, R1 0.066, R2 0.066.(c) Strauss, S. H.; Silver, M. E.; Ibers, J. A. Iron(II) octaethylchlorine: structure and ligand affinity comparison with its porphyrin and isobacteriochlorin homologs. J. Am. Chem. Soc. 1983, 105, 4108– 4109, DOI: 10.1021/ja00350a069Google Scholar4cIron(II) octaethylchlorine: structure and ligand affinity comparison with its porphyrin and isobacteriochlorin homologsStrauss, Steven H.; Silver, Michael E.; Ibers, James A.Journal of the American Chemical Society (1983), 105 (12), 4108-9CODEN: JACSAT; ISSN:0002-7863.trans-7,8-Dihydro-2,3,7,8,12,13,17,18-octaethylporphyrinatoiron(II), Fe(OEC), is orthorhombic, space group Pbcn, with a 21.880(9), b 15.795(6), and c 8.554(4) Å at -150°; Z = 4. The structure was refined anisotropically by least-squares to a final R(F2) = 0.105. The Fe and 4 N atoms are rigorously planar, while the rest of the chlorin macrocycle is significantly S4 ruffled. The affinity of Fe(OEC) and its octaethylporphyrin and octaethylisobacteriochlorin homologs for the weak σ-donor ligands THF and EtSH is strongly macrocycle dependent. This dependence may have a structural basis whereby reduced macrocycles more easily distort, adjusting their core size to accommodate the changing requirements of the metal upon complexation of a weak ligand.
- 5Chatt, J.; Shaw, B. L. Alkyls and aryls of transition metals. Part IV. Cobalt(II) and iron(II) derivatives. J. Chem. Soc. 1961, 285, DOI: 10.1039/jr9610000285Google Scholar5Alkyls and aryls of transition metals. IV. Cobalt(II) and iron(II) derivativesChatt, J.; Shaw, B. L.Journal of the Chemical Society (1961), (), 285-90CODEN: JCSOA9; ISSN:0368-1769.cf. CA 54, 20936d. The prepn. and properties of stable organometallic complexes of the types trans-[MR2-(PR2')2] (M = Co, Fe; R, R' = org. radicals) were described. The groups R were ortho-substituted aryl groups, where the substituents were somewhat bulky. These were the first planar complexes of Co(II) and Fe(II) that had only monodentate ligands. Their configurations were established by their magnetic and elec. dipole moments. Possible reasons for the unusual configurations and stabilities of these organometallic derivs. were discussed.
- 6(a) Nijhuis, C. A.; Jellema, E.; Sciarone, T. J. J.; Meetsma, A.; Budzelaar, P. H. M.; Hessen, B. First-Row Transition Metal Bis(amidinate) Complexes; Planar Four-Coordination of FeII Enforced by Sterically Demanding Aryl Substituents Eur. Eur. J. Inorg. Chem. 2005, 2005, 2089– 2099, DOI: 10.1002/ejic.200500094Google ScholarThere is no corresponding record for this reference.(b) Wurzenberger, X.; Piotrowski, H.; Klufers, P. A stable molecular entity derived from rare iron(II) minerals: the square-planar high-spin d6 Fe(II)O4 chromophore Angew. Angew. Chem., Int. Ed. 2011, 50, 4974– 4978, DOI: 10.1002/anie.201006898Google Scholar6bA Stable Molecular Entity Derived from Rare Iron(II) Minerals: The Square-Planar High-Spin-d6 FeIIO4 ChromophoreWurzenberger, Xaver; Piotrowski, Holger; Kluefers, PeterAngewandte Chemie, International Edition (2011), 50 (21), 4974-4978, S4974/1-S4974/4CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A strongly alk. (LiOH or NaOH) aq. soln. of furanose mimic anhydroerythritol (meso-oxolane-3,4-diol, H2L) and FeCl2 gave a reddish-blue soln., from which, depending on the conditions, orange-red crystals of the tetrahydrates Li2[FeL2]·4H2O (1) or Na2[FeL2]·4H2O (2), or violet crystals of the nonahydrate Na2[FeL2]·9H2O (3) were grown. The bidentate oxolanediolato ligand stabilizes the rare high-spin ferrous SP-4 coordination in 1 as revealed by x-ray crystallog. and SQUID magnetometer measurements. The structure of 2 is also SP-4 as revealed by x-ray crystallog. Compd. 3 is slightly distorted toward tetrahedron. Three related mononuclear high-spin [FeIIX4]2- (X = Cl, F, OH) species and their high-spin [MnIIX4]2- analogs were analyzed in a DFT approach using the unrestricted-B3LYP/tzvp level of theory. Enhanced Jahn-Teller flattening can stabilize square planar high-spin d6 centers despite their "wrong" spin state by a marked sepn. of the neg. charge on the ligand atoms and the stereochem. active d electrons. This result reveals square-planar high-spin centers to be building blocks of reasonable stability.(c) Cantalupo, S. A.; Fiedler, S. R.; Shores, M. P.; Rheingold, A. L.; Doerrer, L. H. High-spin square-planar Co(II) and Fe(II) complexes and reasons for their electronic structure. Angew. Chem., Int. Ed. 2012, 51, 1000– 1005, DOI: 10.1002/anie.201106091Google Scholar6cHigh-Spin Square-Planar CoII and FeII Complexes and Reasons for Their Electronic StructureCantalupo, Stefanie A.; Fiedler, Stephanie R.; Shores, Matthew P.; Rheingold, Arnold L.; Doerrer, Linda H.Angewandte Chemie, International Edition (2012), 51 (4), 1000-1005, S1000/1-S1000/18CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)The high-spin, square planar complexes [K(DME)2]2[M(ddfp)2] (M = CoII, FeII, ddfp = perfluoropinacolate) were prepd. and characterized by x-ray crystallog., temp.-dependent magnetic susceptibility, cyclic voltammetry, and other methods. The tetrahedral analog [K(DME)2]2[Zn(ddfp)2] was also prepd. and characterized crystallog. The combination of high-spin electron configuration and square-planar geometry is made possible by ligand constraints that generate five non-degenerate d-orbitals with ligand-based π-donation, a relatively weak ligand-field splitting, and no intervening ligand-based π-acceptor mol. orbitals.(d) Pinkert, D.; Demeshko, S.; Schax, F.; Braun, B.; Meyer, F.; Limberg, C. A Dinuclear Molecular Iron(II) Silicate with Two High-Spin Square-Planar FeO4 Units. Angew. Chem., Int. Ed. 2013, 52, 5155– 5158, DOI: 10.1002/anie.201209650Google Scholar6dA Dinuclear Molecular Iron(II) Silicate with Two High-Spin Square-Planar FeO4 UnitsPinkert, Denise; Demeshko, Serhiy; Schax, Fabian; Braun, Beatrice; Meyer, Franc; Limberg, ChristianAngewandte Chemie, International Edition (2013), 52 (19), 5155-5158CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)Dinuclear iron complex [L'2Fe2][Na(OEt2)2]2 (1) (H3L' = ligand with elimination of "Me2SiO" from 3-(1,1-dimethylethyl)-3-[(hydroxydimethylsilyl)oxy]-1,1,5,5-tetramethyl-1,5-trisiloxanediol) was synthesized and studied by x-ray diffraction anal., temp.-dependent magnetic susceptibility and Mossbauer spectroscopy.(e) Liu, Y.; Luo, L.; Xiao, J.; Wang, L.; Song, Y.; Qu, J.; Luo, Y.; Deng, L. Four-Coordinate Iron(II) Diaryl Compounds with Monodentate N-Heterocyclic Carbene Ligation: Synthesis, Characterization, and Their Tetrahedral-Square Planar Isomerization in Solution. Inorg. Chem. 2015, 54, 4752– 4760, DOI: 10.1021/acs.inorgchem.5b00138Google Scholar6eFour-Coordinate Iron(II) Diaryl Compounds with Monodentate N-Heterocyclic Carbene Ligation: Synthesis, Characterization, and Their Tetrahedral-Square Planar Isomerization in SolutionLiu, Yuesheng; Luo, Lun; Xiao, Jie; Wang, Lei; Song, You; Qu, Jingping; Luo, Yi; Deng, LiangInorganic Chemistry (2015), 54 (10), 4752-4760CODEN: INOCAJ; ISSN:0020-1669. (American Chemical Society)The salt elimination reactions of (IPr2Me2)2FeCl2 (IPr2Me2 = 1,3-diisopropyl-4,5-dimethylimidazol-2-ylidene) with the corresponding aryl Grignard reagents afford [(IPr2Me2)2FeAr2] (Ar = Ph, 3; C6H4-p-Me, 4; C6H4-p-tBu, 5; C6H3-3,5-(CF3)2, 6) in good yields. X-ray crystallog. studies revealed the presence of both tetrahedral and trans square planar isomers for 3 and 6 and the tetrahedral structures for 4 and 5. Magnetic susceptibility and 57Fe Mossbauer spectrum measurements on the solid samples indicated the high-spin (S = 2) and intermediate-spin (S = 1) nature of the tetrahedral and square planar structures, resp. Soln. property studies, including soln. magnetic susceptibility measurement, variable-temp. 1H and 19F NMR, and absorption spectroscopy, on 3-6, as well as an 57Fe Mossbauer spectrum study on a frozen THF soln. of tetrahedral [(IPr2Me2)257FePh2] suggest the coexistence of tetrahedral and trans square planar structures in soln. phase. D. functional theory calcns. on (IPr2Me2)2FePh2 disclosed that the tetrahedral and trans square planar isomers are close in energy and that the geometry isomerization can occur by spin-change-coupled geometric transformation on four-coordinate iron(II) center.
- 7(a) Holm, R. H.; Chakravorty, A.; Theriot, L. J. The Synthesis, Structures, and Solution Equilibria of Bis(pyrrole-2-aldimino)metal(II) Complexes. Inorg. Chem. 1966, 5, 625– 635, DOI: 10.1021/ic50038a028Google Scholar7aThe synthesis, structures, and solution equilibria of bis-(pyrrole-2-aldimino)metal(II) complexesHolm, R. H.; Chakravorty, A.; Theriot, L. J.Inorganic Chemistry (1966), 5 (4), 625-35CODEN: INOCAJ; ISSN:0020-1669.Synthesis of an extensive series of bis(pyrrole-2-aldimino)metal(II) complexes with M(II) = Co, Ni, Pd, Cu, and Zn and various alkyl groups (R) appended to the azomethine N was effected by a nonaq. chelation reaction in tetrahydrofuran. Preliminary single crystal x-ray results for complexes with R = tert-Bu reveal that Co, Ni, and Zn complexes are isomorphous, but appreciable differences in the cell consts. of the Ni complex indicate that it is not truly isostructural with the tetrahedral Co and Zn complexes. The Cu complex exists in 2 cryst. modifications, neither of which is isomorphous with the Co-Ni-Zn series. Spectral and magnetic studies in soln. show that the tert-Bu Co complex is tetrahedral whereas the corresponding Cu complex is distorted from planarity to an unknown extent. Cu complexes with less bulky R groups are planar. The tert-Bu Ni complex is pseudo-tetrahedral; complexes with sec-alkyl groups such as iso-Pr are involved in a configurational equil. between planar and pseudo-tetrahedral forms. The paramagnetic Ni complexes show large isotropic proton hyperfine contact shifts. Spin d. calcns. for the coordinated ligand system are used as the basis of proton resonance assignments. It is concluded that in the pseudo-tetrahedral form spin imbalance exists in the highest filled ligand π-mol. orbital, and that, in addn., there is an underlying spin imbalance in the highest filled σ-mol. orbital, the result of which is observable in the proton resonance spectra. Thermodynamic parameters characterizing the structural change were obtained for the Ni complexes from the temp. dependence of the proton contact shifts. A quant. comparison of the stabilization of tetrahedral Ni(II) by pyrrole-2-aldimine, salicylaldimine, and β-keto amine ligand systems is presented.(b) Wolny, J. A.; Rudolf, M. F.; Ciunik, Z.; Gatner, K.; Wołowiec, S. Cobalt(II) triazene 1-oxide bis(chelates). A case of planar (low spin)–tetrahedral (high spin) isomerism. J. Chem. Soc., Dalton Trans. 1993, 1611– 1622, DOI: 10.1039/DT9930001611Google Scholar7bCobalt(II) triazene 1-oxide bis(chelates). A case of planar (low spin)-tetrahedral (high spin) isomerismWolny, Juliusz A.; Rudolf, Mikolaj F.; Ciunik, Zbigniew; Gatner, Kazimierz; Wolowiec, StanislawJournal of the Chemical Society, Dalton Transactions: Inorganic Chemistry (1972-1999) (1993), (10), 1611-22CODEN: JCDTBI; ISSN:0300-9246.High- and low-spin [CoL2] (HL = 3-phenyl-1-triazene 1-oxide derivs.) were prepd. and isolated. The low-spin complexes possess square-planar structure and the high-spin complexes are tetrahedral. The mol. structure of high-spin [CoL2] [HL = 3-(4-methylphenyl)-1-methyl-1-triazene 1-oxide] was detd.; triclinic, space group P‾1, a 7.970(5), b 10.174(5), c 11.676(5) Å, α 87.18(4), β 74.31(4), γ 74.06(4)°, Z = 2, R = 0.041. For some complexes the isolation of both planar (low-spin) and tetrahedral (high-spin) isomers or of their conglomerates is possible depending on the synthesis conditions. The crystal structure of square-planar [Ni(ON(Me)NNC6H4Me-4)2], which is isomorphous with the low-spin isomer of [Co(ON(Me)NNC6H4Me-4)2], was detd.: triclinic, space group P‾1, a 7.495(2), b 7.694(5), c 8.612(3) Å, α 64.64(5), β 87.84(2), γ 78.64(3)°, Z = 1, R = 0.0271. The complexes exhibit a planar-tetrahedral equil. in noncoordinating solvents, the ΔHΘ and ΔSΘ values of which, detd. from soln. magnetic susceptibility measurements, at 1-15 kJ mol-1 and 5-30 J K-1 mol-1, resp. The electrochem. properties of the complexes are given.(c) Ingleson, M. J.; Pink, M.; Fan, H.; Caulton, K. G. Exploring the reactivity of four-coordinate PNPCoX with access to three-coordinate spin triplet PNPCo. Inorg. Chem. 2007, 46, 10321– 10334, DOI: 10.1021/ic701171pGoogle Scholar7cExploring the reactivity of four-coordinate PNPCoX with access to three-coordinate spin triplet PNPCoIngleson, Michael J.; Pink, Maren; Fan, Hongjun; Caulton, Kenneth G.Inorganic Chemistry (2007), 46 (24), 10321-10334CODEN: INOCAJ; ISSN:0020-1669. (American Chemical Society)(PNP)CoX, where PNP is (tBu2PCH2SiMe2)2N- and X is Cl, I, N3, OAr, OSO2CF3, and N(H)Ar, are reported. Some of these show magnetic susceptibility, color, and 1H NMR evidence of being in equil. between a blue, tetrahedral S = 3/2 state and a red, planar S = 1/2 state; the equil. populations are influenced by subtle solvent effects (e.g., benzene and cyclohexane are different), as well as by temp. Attempted oxidn. to Co(III) with O2 occurs instead at phosphorus, giving [P(O)NP(O)]CoX species. The single O-atom transfer reagent PhI:O likewise oxidizes P. Even I2 oxidizes P to give the pendant phosphonium species (tBu2P(I)CH2SiMe2NSiMe2CH2PtBu2)CoI2 with a tetrahedral S = 3/2 cobalt; the solid-state structure shows intermol. PI···ICo interactions. Attempted alkyl metathesis of PNPCoX inevitably results in redn., forming PNPCo, which is a spin triplet with planar T-shaped coordination geometry with no agostic interaction. Triplet PNPCo binds N2(weakly) and CO (whose low CO stretching frequency indicates strong PNP → Co donor power), but not ethene or MeCCMe.
- 8Gaazo, J. Plasticity of the coordination sphere of copper(II) complexes, its manifestation and causes. Coord. Chem. Rev. 1976, 19, 253– 297, DOI: 10.1016/S0010-8545(00)80317-3Google ScholarThere is no corresponding record for this reference.
- 9(a) Cambi, L.; Szegö, L. Über die magnetische Susceptibilität der komplexen Verbindungen. Ber. Dtsch. Chem. Ges. B 1931, 64, 2591– 2598, DOI: 10.1002/cber.19310641002Google ScholarThere is no corresponding record for this reference.(b) Kahn, O.; Martinez, C. J. Spin-Transition Polymers: From Molecular Materials Toward Memory Devices. Science 1998, 279, 44– 48, DOI: 10.1126/science.279.5347.44Google Scholar9bSpin-transition polymers: from molecular materials toward memory devicesKahn, O.; Martinez, C. JayScience (Washington, D. C.) (1998), 279 (5347), 44-48CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)A review with 53 refs. Some 3dn (4 ≤ n ≤ 7) transition metal compds. exhibit a cooperative transition between a low-spin (LS) and a high-spin (HS) state. This transition is abrupt and occurs with a thermal hysteresis, which confers a memory effect on the system. The intersite interactions and thus the cooperativity are magnified in polymeric compds. such as [Fe(Rtrz)3]A2·nH2O in which the Fe2+ ions are triply bridged by 4-R-substituted-1,2,4-triazole mols. Moreover, in these compds., the spin transition is accompanied by a well-pronounced change of color between violet in the LS state and white in the HS state. The transition temps. of these materials can be fine tuned, using an approach based on the concept of a mol. alloy. In particular, it is possible to design a compd. for which room temp. falls in the middle of the thermal hysteresis loop. These materials have many potential applications, for example, as temp. sensors, as active elements of various types of displays, and in information storage and retrieval.(c) Gütlich, P.; Garcia, Y.; Goodwin, H. A. Spin crossover phenomena in Fe(ii) complexes. Chem. Soc. Rev. 2000, 29, 419– 427, DOI: 10.1039/b003504lGoogle Scholar9cSpin crossover phenomena in Fe(II) complexesGutlich, Philipp; Garcia, Yann; Goodwin, Harold A.Chemical Society Reviews (2000), 29 (6), 419-427CODEN: CSRVBR; ISSN:0306-0012. (Royal Society of Chemistry)A review with 58 refs. The behavior of spin crossover compds. is among the most striking and fascinating shown by relatively simple mol. species. This review aims to draw attention to the various ways in which spin crossover phenomena are manifested in Fe(II) complexes, to offer some rationalization for these, and to highlight their possible applications. Typical examples have been selected along with more recent ones to give an overall view of the scope and development of the area. The article is structured to provide the basic material for those who wish to enter the field of spin crossover.(d) Sato, O.; Tao, J.; Zhang, Y. Z. Control of magnetic properties through external stimuli. Angew. Chem., Int. Ed. 2007, 46, 2152– 2187, DOI: 10.1002/anie.200602205Google Scholar9dControl of magnetic properties through external stimuliSato, Osamu; Tao, Jun; Zhang, Yuan-ZhuAngewandte Chemie, International Edition (2007), 46 (13), 2152-2187CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. The magnetic properties of many magnetic materials can be controlled by external stimuli. The principal focus here is on the thermal, photochem., electrochem., and chem. of phase transitions that involve changes in magnetization. Mol. compds. described herein range from metal complexes through pure org. compds. to composite materials. Most of the review is devoted to the properties. valence-tautomeric compds., mol. magnets, and spin-crossover complexes, which could find future applications in memory devices or optical switches.(e) Bousseksou, A.; Molnar, G.; Salmon, L.; Nicolazzi, W. Molecular spin crossover phenomenon: recent achievements and prospects. Chem. Soc. Rev. 2011, 40, 3313– 3335, DOI: 10.1039/c1cs15042aGoogle Scholar9eMolecular spin crossover phenomenon: Recent achievements and prospectsBousseksou, Azzedine; Molnar, Gabor; Salmon, Lionel; Nicolazzi, WilliamChemical Society Reviews (2011), 40 (6), 3313-3335CODEN: CSRVBR; ISSN:0306-0012. (Royal Society of Chemistry)A review. Recently we assisted a strong renewed interest in the fascinating field of mol. spin crossover complexes by (1) the emergence of nanosized spin crossover materials through direct synthesis of coordination nanoparticles and nanopatterned thin films as well as by (2) the use of novel sophisticated high spatial and temporal resoln. exptl. techniques and theor. approaches for the study of spatiotemporal phenomena in cooperative spin crossover systems. Besides generating new fundamental knowledge on size-redn. effects and the dynamics of the spin crossover phenomenon, this research aims also at the development of practical applications such as sensor, display, information storage and nanophotonic devices. In this crit. review, we discuss recent work in the field of mol.-based spin crossover materials with a special focus on these emerging issues, including chem. synthesis, phys. properties and theor. aspects as well (223 refs.).
- 10Halcrow, M. A. The spin-states and spin-transitions of mononuclear iron(II) complexes of nitrogen-donor ligands. Polyhedron 2007, 26, 3523– 3576, DOI: 10.1016/j.poly.2007.03.033Google Scholar10The spin-states and spin-transitions of mononuclear iron(II) complexes of nitrogen-donor ligandsHalcrow, Malcolm A.Polyhedron (2007), 26 (14), 3523-3576CODEN: PLYHDE; ISSN:0277-5387. (Elsevier B.V.)A review of mononuclear iron(II) complexes with heterocyclic N-donor ligation is presented. A brief introduction to spin-crossover chem. and low-temp. spin-trapping is provided, since many of these compds. undergo thermal spin-transitions upon cooling or heating. These are highlighted, and the structural changes underlying spin-crossover are discussed where this is known. Materials showing spin-trapping behavior following thermal quenching or irradn. at very low temps. are also described.
- 11Bowman, A. C.; Milsmann, C.; Bill, E.; Turner, Z. R.; Lobkovsky, E.; DeBeer, S.; Wieghardt, K.; Chirik, P. J. Synthesis and Electronic Structure Determination of N-Alkyl-Substituted Bis(imino)pyridine Iron Imides Exhibiting Spin Crossover Behavior. J. Am. Chem. Soc. 2011, 133, 17353– 17369, DOI: 10.1021/ja205736mGoogle Scholar11Synthesis and Electronic Structure Determination of N-Alkyl-Substituted Bis(imino)pyridine Iron Imides Exhibiting Spin Crossover BehaviorBowman, Amanda C.; Milsmann, Carsten; Bill, Eckhard; Turner, Zoe R.; Lobkovsky, Emil; DeBeer, Serena; Wieghardt, Karl; Chirik, Paul J.Journal of the American Chemical Society (2011), 133 (43), 17353-17369CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Three new N-alkyl substituted bis(imino)pyridine iron imide complexes, (iPrPDI)FeNR (iPrPDI = 2,6-(2,6-iPr2-C6H3-N:CMe)2C5H3N; R = 1-adamantyl (1Ad), cyclooctyl (CyOct), and 2-adamantyl (2Ad)) were synthesized by addn. of the appropriate alkyl azide to the iron bis(dinitrogen) complex, (iPrPDI)Fe(N2)2. SQUID magnetic measurements on the isomeric iron imides, (iPrPDI)FeN1Ad and (iPrPDI)FeN2Ad, established spin crossover behavior with the latter example having a more complete spin transition in the exptl. accessible temp. range. X-ray diffraction on all three alkyl-substituted bis(imino)pyridine iron imides established essentially planar compds. with relatively short Fe-Nimide bond lengths and two-electron redn. of the redox-active bis(imino)pyridine chelate. Zero- and applied-field Mossbauer spectroscopic measurements indicate diamagnetic ground states at cryogenic temps. and established low isomer shifts consistent with highly covalent mols. For (iPrPDI)FeN2Ad, Mossbauer spectroscopy also supports spin crossover behavior and allowed extn. of thermodn. parameters for the S = 0 to S = 1 transition. X-ray absorption spectroscopy and computational studies were also performed to explore the electronic structure of the bis(imino)pyridine alkyl-substituted imides. An electronic structure description with a low spin ferric center (S = 1/2) antiferromagnetically coupled to an imidyl radical (Simide = 1/2) and a closed-shell, dianionic bis(imino)pyridine chelate (SPDI = 0) is favored for the S = 0 state. An iron-centered spin transition to an intermediate spin ferric ion (SFe = 3/2) accounts for the S = 1 state obsd. at higher temps. Other possibilities based on the computational and exptl. data are also evaluated and compared to the electronic structure of the bis(imino)pyridine iron N-aryl imide counterparts.
- 12(a) Scepaniak, J. J.; Harris, T. D.; Vogel, C. S.; Sutter, J.; Meyer, K.; Smith, J. M. Spin Crossover in a Four-Coordinate Iron(II) Complex. J. Am. Chem. Soc. 2011, 133, 3824– 3827, DOI: 10.1021/ja2003473Google Scholar12aSpin Crossover in a Four-Coordinate Iron(II) ComplexScepaniak, Jeremiah J.; Harris, T. David; Vogel, Carola S.; Sutter, Jorg; Meyer, Karsten; Smith, Jeremy M.Journal of the American Chemical Society (2011), 133 (11), 3824-3827CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The four-coordinate iron(II) phosphoraniminato complex PhB(MesIm)3Fe-N=PPh3 undergoes an S = 0 to S = 2 spin transition with TC = 81 K, as detd. by variable-temp. magnetic measurements and Mossbauer spectroscopy. Variable-temp. single-crystal X-ray diffraction revealed that the S = 0 to S = 2 transition is assocd. with an increase in the Fe-C and Fe-N bond distances and a decrease in the N-P bond distance. These structural changes have been interpreted in terms of electronic structure theory.(b) Mathoniere, C.; Lin, H. J.; Siretanu, D.; Clerac, R.; Smith, J. M. Photoinduced single-molecule magnet properties in a four-coordinate iron(II) spin crossover complex. J. Am. Chem. Soc. 2013, 135, 19083– 19086, DOI: 10.1021/ja410643sGoogle Scholar12bPhotoinduced Single-Molecule Magnet Properties in a Four-Coordinate Iron(II) Spin Crossover ComplexMathoniere, Corine; Lin, Hsiu-Jung; Siretanu, Diana; Clerac, Rodolphe; Smith, Jeremy M.Journal of the American Chemical Society (2013), 135 (51), 19083-19086CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The four-coordinate Fe-(II) complex, PhB(MesIm)3FeNPPh3 (1) was previously reported to undergo a thermal spin-crossover (SCO) between high-spin (HS, S = 2) and low-spin (LS, S = 0) states. This complex is photoactive <20 K, undergoing a photoinduced LS to HS spin state change, as detd. by optical reflectivity and photomagnetic measurements. With continuous white light irradn., 1 displays slow relaxation of the magnetization, i.e. single-mol. magnet (SMM) properties, at temps. <5 K. This complex provides a structural template for the design of new photoinduced mononuclear SMMs based on the SCO phenomenon.(c) Lin, H. J.; Siretanu, D.; Dickie, D. A.; Subedi, D.; Scepaniak, J. J.; Mitcov, D.; Clerac, R.; Smith, J. M. Steric and electronic control of the spin state in three-fold symmetric, four-coordinate iron(II) complexes. J. Am. Chem. Soc. 2014, 136, 13326– 13332, DOI: 10.1021/ja506425aGoogle Scholar12cSteric and Electronic Control of the Spin State in Three-Fold Symmetric, Four-Coordinate Iron(II) ComplexesLin, Hsiu-Jung; Siretanu, Diana; Dickie, Diane A.; Subedi, Deepak; Scepaniak, Jeremiah J.; Mitcov, Dmitri; Clerac, Rodolphe; Smith, Jeremy M.Journal of the American Chemical Society (2014), 136 (38), 13326-13332CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The 3-fold sym., four-coordinate Fe(II) phosphoraminimato complexes PhB(MesIm)3Fe-N=PRR'R'' (PRR'R'' = PMePh2, PMe2Ph, PMe3, and PPr3) undergo a thermally induced S = 0 to S = 2 spin-crossover in fluid soln. Smaller phosphoraminimato ligands stabilize the low-spin state, and an excellent correlation is obsd. between the characteristic temp. of the spin-crossover (T1/2) and the Tolman cone angle (θ). Complexes with para-substituted triarylphosphoraminimato ligands (p-XC6H4)3P=N- (X = H, Me and OMe) also undergo spin-crossover in soln. These isosteric phosphoraminimato ligands reveal that the low-spin state is stabilized by more strongly donating ligands. This control over the spin state provides important insights for modulating the magnetic properties of four-coordinate Fe(II) complexes.
- 13Creutz, S. E.; Peters, J. C. Spin-State Tuning at Pseudo-tetrahedral d6 Ions: Spin Crossover in [BP3]FeII–X Complexes. Inorg. Chem. 2016, 55, 3894– 3906, DOI: 10.1021/acs.inorgchem.6b00066Google Scholar13Spin-State Tuning at Pseudo-tetrahedral d6 Ions: Spin Crossover in [BP3]FeII-X ComplexesCreutz, Sidney E.; Peters, Jonas C.Inorganic Chemistry (2016), 55 (8), 3894-3906CODEN: INOCAJ; ISSN:0020-1669. (American Chemical Society)Low-coordinate transition-metal complexes that undergo spin crossover remain rare. We report here a series of four-coordinate, pseudo-tetrahedral P3FeII-X complexes supported by tris(phosphine)borate P3 ([PhBPR3]-) and phosphiniminato X-type ligands (-N=PR'3) that, in combination, tune the spin-crossover behavior of the system. Most of the reported iron complexes undergo spin crossover at temps. near or above room temp. in soln. and in the solid state. The change in spin state coincides with a significant change in the degree of π-bonding between Fe and the bound N atom of the phosphiniminato ligand. Spin crossover is accompanied by striking changes in the UV-visible (UV-vis) absorption spectra, which allows for quant. modeling of the thermodn. parameters of the spin equil. These spin equil. have also been studied by numerous techniques including paramagnetic NMR (NMR), IR, and M.ovrddot.ossbauer spectroscopies; X-ray crystallog.; and solid-state superconducting quantum interference device (SQUID) magnetometry. These studies allow qual. correlations to be made between the steric and electronic properties of the ligand substituents and the enthalpy and entropy changes assocd. with the spin equil.
- 14(a) Travieso-Puente, R.; Broekman, J. O. P.; Chang, M.-C.; Demeshko, S.; Meyer, F.; Otten, E. Spin-Crossover in a Pseudo-tetrahedral Bis(formazanate) Iron Complex. J. Am. Chem. Soc. 2016, 138, 5503– 5506, DOI: 10.1021/jacs.6b01552Google Scholar14aSpin-Crossover in a Pseudo-tetrahedral Bis(formazanate) Iron ComplexTravieso-Puente, Raquel; Broekman, J. O. P.; Chang, Mu-Chieh; Demeshko, Serhiy; Meyer, Franc; Otten, EdwinJournal of the American Chemical Society (2016), 138 (17), 5503-5506CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Spin-crossover in a pseudo-tetrahedral bis(formazanate) iron(II) complex (1) is described. Structural, magnetic, and spectroscopic analyses indicate that this compd. undergoes thermal switching between an S = 0 and an S = 2 state, which is very rare in four-coordinate complexes. The transition to the high-spin state is accompanied by an increase in Fe-N bond lengths and a concomitant contraction of intraligand N-N bonds. The latter suggests that stabilization of the low-spin state is due to the π-acceptor properties of the ligand. One-electron redn. of 1 leads to the formation of the corresponding anion, which contains a low-spin (S = 1/2) Fe(I) center. The findings are rationalized by electronic structure calcns. using d. functional theory.(b) Milocco, F.; de Vries, F.; Bartels, I. M. A.; Havenith, R. W. A.; Cirera, J.; Demeshko, S.; Meyer, F.; Otten, E. Electronic Control of Spin-Crossover Properties in Four-Coordinate Bis(formazanate) Iron(II) Complexes. J. Am. Chem. Soc. 2020, 142, 20170– 20181, DOI: 10.1021/jacs.0c10010Google Scholar14bElectronic Control of Spin-Crossover Properties in Four-Coordinate Bis(formazanate) Iron(II) ComplexesMilocco, Francesca; de Vries, Folkert; Bartels, Imke M. A.; Havenith, Remco W. A.; Cirera, Jordi; Demeshko, Serhiy; Meyer, Franc; Otten, EdwinJournal of the American Chemical Society (2020), 142 (47), 20170-20181CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The transition between spin states in d-block metal complexes has important ramifications for their structure and reactivity, with applications ranging from information storage materials to understanding catalytic activity of metalloenzymes. Tuning the ligand field (ΔO) by steric and/or electronic effects has provided spin-crossover compds. for several transition metals in the periodic table, but this has mostly been limited to coordinatively satd. metal centers in octahedral ligand environments. Spin-crossover complexes with low coordination nos. are much rarer. Here we report a series of four-coordinate, (pseudo)tetrahedral Fe(II) complexes with formazanate ligands and demonstrate how electronic substituent effects can be used to modulate the thermally induced transition between S = 0 and S = 2 spin states in soln. All six compds. undergo spin-crossover in soln. with T1/2 above room temp. (300-368 K). While structural anal. by X-ray crystallog. shows that the majority of these compds. are low-spin in the solid state (and remain unchanged upon heating), we find that packing effects can override this preference and give rise to either rigorously high-spin (6) or gradual spin-crossover behavior (5) also in the solid state. D. functional theory calcns. are used to delineate the empirical trends in soln. spin-crossover thermodn. In all cases, the stabilization of the low-spin state is due to the π-acceptor properties of the formazanate ligand, resulting in an "inverted" ligand field, with an approx. "two-over-three" splitting of the d-orbitals and a high degree of metal-ligand covalency due to metal → ligand π-backdonation. The computational data indicate that the electronic nature of the para-substituent has a different influence depending on whether it is present at the C-Ar or N-Ar rings, which is ascribed to the opposing effect on metal-ligand σ- and π-bonding.
- 15(a) Feltham, H. L. C.; Barltrop, A. S.; Brooker, S. Spin crossover in iron(II) complexes of 3,4,5-tri-substituted-1,2,4-triazole (Rdpt), 3,5-di-substituted-1,2,4-triazolate (dpt – ), and related ligands. Coord. Chem. Rev. 2017, 344, 26– 53, DOI: 10.1016/j.ccr.2016.10.006Google Scholar15aSpin crossover in iron(II) complexes of 3,4,5-tri-substituted-1,2,4-triazole (Rdpt), 3,5-di-substituted-1,2,4-triazolate (dpt-), and related ligandsFeltham, Humphrey L. C.; Barltrop, Alexis S.; Brooker, SallyCoordination Chemistry Reviews (2017), 344 (), 26-53CODEN: CCHRAM; ISSN:0010-8545. (Elsevier B.V.)A review. A general introduction to spin crossover involving iron(II) is provided. Then the 26 examples of 3,4,5-tri-substituted-1,2,4-triazole (mostly Rdpt and Rppt) ligands and the 11 examples of 3,5-di-substituted-1,2,4-triazole (including Hdpt) ligands that are of interest to this review are introduced. The 26 examples of 3,4,5-tri-substituted-1,2,4-triazole ligands fall into main two categories: (a) the 13 examples of sym. 4-substituted 3,5-di(2-pyridyl)triazole (Rdpt) ligands and (b) the 13 ligands closely related to the Rdpt ligands but featuring Ph (known as Rppt ligands), pyrazine or pyrimidine rings in place of one of more of the pyridine rings. In contrast, of the 11 examples of 3,5-di-substituted-1,2,4-triazole (HL, including Hdpt) ligands, only 2 are sym. (3,5-di(2-pyridyl)triazole and 3,5-di(2-pyrazyl)triazole ligands); the other 9 are non-sym. as two different rings are attached to the 3 and 5 positions of the triazole ring. In all but one example, in iron complexes of the 3,5-di-substituted-1,2,4-triazole ligands (HL, including Hdpt) the ligand deprotonates and coordinates as an anionic 3,5-di-substituted-1,2,4-triazolate (L-) ligand. This field was reviewed in 2008, and considered the 17 such complexes published up until 2007. Hence the present review provides a comprehensive update on the substantial progress that has been made in this field since then. Specifically it covers the 92 new iron(II) complexes of these 37 ligands, published since then and before 1 Dec. 2015. Of the 92 of new iron(II) complexes reported, 72 have been structurally characterized, and almost half of them, 40, are SCO-active. A wide range of nuclearities, from mono- to di- to tri- to penta- and poly-nuclear, as well as a range of binding modes, are obsd., so this review is organised into sections accordingly. In each section, after briefly summarizing the findings of the 2008 review, the new developments are detailed. Finally, the findings of this rich avenue of investigation into such ligands are summarized, indicating a bright future ahead.(b) Constable, E. C.; Baum, G.; Bill, E.; Dyson, R.; van Eldik, R.; Fenske, D.; Kaderli, S.; Morris, D.; Neubrand, A.; Neuburger, M.; Smith, D. R.; Wieghardt, K.; Zehnder, M.; Zuberbühler, A. D. Control of Iron(II) Spin States in 2,2′:6′,2″-Terpyridine Complexes through Ligand Substitution. Chem. - Eur. J. 1999, 5, 498– 508, DOI: 10.1002/(SICI)1521-3765(19990201)5:2<498::AID-CHEM498>3.0.CO;2-VGoogle Scholar15bControl of iron(II) spin states in 2,2':6',2''-terpyridine complexes through ligand substitutionConstable, Edwin C.; Baum, Gerhard; Bill, Eckhard; Dyson, Raylene; Van Eldik, Rudi; Fenske, Dieter; Kaderli, Susan; Morris, Darrell; Neubrand, Anton; Neuburger, Markus; Smith, Diane R.; Wieghardt, Karl; Zehnder, Margareta; Zuberbuhler, Andreas D.Chemistry - A European Journal (1999), 5 (2), 498-508CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH)6- And 6,6''-aryl-substituted 2,2':6',2''-terpyridine ligands and their Fe(II) complexes were prepd. The introduction of Ph substituents at both the 6- and 6'' - positions leads exclusively to the formation of orange high-spin Fe(II) complexes while the presence of a single 6-Ph substituent results in spin-crossover systems. [Fe(2)2]X2 (2 = 4,6-diphenyl-2,2':6',2''-terpyridine; X = ClO4 or PF6) were studied in detail, and the solid-state x-ray structures of both the low- and high-spin forms are reported. Mossbauer spectroscopic and magnetic susceptibility measurements are reported, and the temp. and pressure dependence of the high-spin/low-spin transition were studied. An x-ray structural study of [Fe(2)2](ClO4)2 is also reported; this complex is highly distorted with two very long Fe···N contacts of over 2.4 Å and is best regarded as a four-coordinate Fe complex.
- 16(a) Bacchi, S.; Benaglia, M.; Cozzi, F.; Demartin, F.; Filippini, G.; Gavezzotti, A. X-ray Diffraction and Theoretical Studies for the Quantitative Assessment of Intermolecular Arene–Perfluoroarene Stacking Interactions. Chem. - Eur. J. 2006, 12, 3538– 3546, DOI: 10.1002/chem.200501248Google Scholar17aX-ray diffraction and theoretical studies for the quantitative assessment of intermolecular arene-perfluoroarene stacking interactionsBacchi, Sergio; Benaglia, Maurizio; Cozzi, Franco; Demartin, Francesco; Filippini, Giuseppe; Gavezzotti, AngeloChemistry - A European Journal (2006), 12 (13), 3538-3546CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH & Co. KGaA)The arene-perfluoroarene stacking interaction was studied by exptl. and theor. methods. Compds. with different possibilities for formation of this recognition motif in the solid state were synthesized, and their crystal structures detd. by single-crystal x-ray diffraction. The crystal packing of these compds., as well as the packing of related compds. retrieved from crystallog. databases, were analyzed with quant. crystal potentials: total lattice energies and the cohesive energies of closest mol. pairs in the crystals were calcd. The arene-perfluoroarene recognition motif emerges as a dominant interaction in the nonhydrogen-bonding compds. studied here, to the point that asym. dimers formed over the stacking motif carry over to asym. units made of two mols. in the crystal both for pure compds. and for mol. complexes; however, inter-ring distances and angles range from 3.70 to 4.85 Å and from 5 to 21°, resp. Pixel energy partitioning reveals that whenever arom. rings stack, the largest cohesive energy contribution comes from dispersion, which roughly amts. to 20 kJ mol-1 per Ph ring, while the coulombic term is minor but significant enough to make a difference between the arene-arene or perfluoroarene-perfluoroarene interactions on the one hand, and arene-perfluoroarene interactions on the other, whereby the latter are favored by ∼10 kJ mol-1 per Ph ring. No evidence of special interaction which can be attributed to H···F confrontation was recognizable.(b) Salonen, L. M.; Ellermann, M.; Diederich, F. Aromatic rings in chemical and biological recognition: energetics and structures. Angew. Chem., Int. Ed. 2011, 50, 4808– 4842, DOI: 10.1002/anie.201007560Google Scholar17bAromatic rings in chemical and biological recognition: energetics and structuresSalonen, Laura M.; Ellermann, Manuel; Diederich, FrancoisAngewandte Chemie, International Edition (2011), 50 (21), 4808-4842CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. This review describes a multidimensional treatment of mol. recognition phenomena involving arom. rings in chem. and biol. systems. It summarizes new results reported since the appearance of an earlier review in 2003 in host-guest chem., biol. affinity assays and biostructural anal., data base mining in the Cambridge Structural Database (CSD) and the Protein Data Bank (PDB), and advanced computational studies. Topics addressed are arene-arene, perfluoroarene-arene, S···arom., cation-π, and anion-π interactions, as well as hydrogen bonding to π systems. The generated knowledge benefits, in particular, structure-based hit-to-lead development and lead optimization both in the pharmaceutical and in the crop protection industry. It equally facilitates the development of new advanced materials and supramol. systems, and should inspire further utilization of interactions with arom. rings to control the stereochem. outcome of synthetic transformations.(c) Wheeler, S. E. Local Nature of Substituent Effects in Stacking Interactions. J. Am. Chem. Soc. 2011, 133, 10262– 10274, DOI: 10.1021/ja202932eGoogle Scholar17cLocal Nature of Substituent Effects in Stacking InteractionsWheeler, Steven E.Journal of the American Chemical Society (2011), 133 (26), 10262-10274CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Popular explanations of substituent effects in π-stacking interactions hinge upon substituent-induced changes in the aryl π-system. This entrenched view has been used to explain substituent effects in countless stacking interactions over the past 2 decades. However, for a broad range of stacked dimers, it is shown that substituent effects are better described as arising from local, direct interactions of the substituent with the proximal vertex of the other ring. Consequently, substituent effects in stacking interactions are additive, regardless of whether the substituents are on the same or opposite rings. Substituent effects are also insensitive to the introduction of heteroatoms on distant parts of either stacked ring. This local, direct interaction viewpoint provides clear, unambiguous explanations of substituent effects for myriad stacking interactions that are in accord with robust computational data, including DFT-D and new benchmark CCSD(T) results. Many of these computational results cannot be readily explained using traditional π-polarization-based models. Analyses of stacking interactions based solely on the sign of the electrostatic potential above the face of an arom. ring or the mol. quadrupole moment face a similar fate. The local, direct interaction model provides a simple means of analyzing substituent effects in complex arom. systems and also offers simple explanations of the crystal packing of fluorinated benzenes and the recently published dependence of the stability of protein-RNA complexes on the regiochem. of fluorinated base analogs [J. Am. Chem. Soc.2011, 133, 3687-3689].
- 17Gütlich, P. Fifty Years of Mössbauer Spectroscopy in Solid State Research - Remarkable Achievements, Future Perspectives. Z. Anorg. Allg. Chem. 2012, 638, 15– 43, DOI: 10.1002/zaac.201100416Google Scholar18Fifty Years of Moessbauer Spectroscopy in Solid State Research - Remarkable Achievements, Future PerspectivesGuetlich, PhilippZeitschrift fuer Anorganische und Allgemeine Chemie (2012), 638 (1), 15-43CODEN: ZAACAB; ISSN:0044-2313. (Wiley-VCH Verlag GmbH & Co. KGaA)Moessbauer spectroscopy was founded more than fifty years ago based on an outstanding discovery by the young German physicist Rudolf Ludwig Moessbauer while working on his Ph.D. thesis. He discovered the recoilless nuclear resonance fluorescence of gamma radiation and was awarded the Nobel Prize in Physics in 1961 as one of the youngest recipients of this most prestigious award. His discovery led to the development of a new technique for measurements of hyperfine interactions between nuclear moments and electromagnetic fields. This method, with highest sharpness of tuning of 10-13, yields information on valence state, symmetry, magnetic behavior, phase transition, lattice dynamics and other solid state properties.
- 18Milocco, F.; Demeshko, S.; Meyer, F.; Otten, E. Ferrate(II) complexes with redox-active formazanate ligands. Dalton Trans. 2018, 47, 8817– 8823, DOI: 10.1039/C8DT01597JGoogle Scholar19Ferrate(II) complexes with redox-active formazanate ligandsMilocco, Francesca; Demeshko, Serhiy; Meyer, Franc; Otten, EdwinDalton Transactions (2018), 47 (26), 8817-8823CODEN: DTARAF; ISSN:1477-9226. (Royal Society of Chemistry)The synthesis of mono(formazanate) Fe complexes is described. In the presence of Bu4N halides, salt metathesis reactions afford the ferrate(II) complexes [Bu4N][LFeX2] (L = PhNNC(p-tol)NNPh; X = Cl, Br) in good yield, and the products were characterized. The high-spin ferrate(II) complexes show cyclic voltammograms that are consistent with reversible, ligand-based 1-electron redn. The halides in these ferrate(II) compds. are labile, and are displaced by 4-methoxyphenyl isocyanide (4 equiv) as evidenced by formation of the low-spin, cationic octahedral complex [LFe(CNC6H4(p-OMe))4][Br]. Thus, a straightforward route to mono(formazanate) Fe(II) complexes was established.
- 19Lepori, C.; Guillot, R.; Hannedouche, J. C1-symmetric β-Diketiminatoiron(II) Complexes for Hydroamination of Primary Alkenylamines. Adv. Synth. Catal. 2019, 361, 714– 719, DOI: 10.1002/adsc.201801464Google Scholar20C1-symmetric β-Diketiminatoiron(II) Complexes for Hydroamination of Primary AlkenylaminesLepori, Clement; Guillot, Regis; Hannedouche, JeromeAdvanced Synthesis & Catalysis (2019), 361 (4), 714-719CODEN: ASCAF7; ISSN:1615-4150. (Wiley-VCH Verlag GmbH & Co. KGaA)The synthesis and solid-state characterization of an array of well-defined low-coordinate C1-sym. β-diketiminatoiron(II) alkyl complexes featuring steric and electronic variations on one of the N-aryl substituents of the β-diketiminate ligand scaffold are reported. All complexes display unique catalytic abilities of promoting the selective exo-cyclohydroamination of unprotected 2,2-diphenylpent-4-en-1-amine under mild reactions conditions. The incorporation of a potentially coordinative ortho-methoxy substituent on one of the N-aryl rings of the β-diketiminate skeleton, in conjunction with a more crowded 2,6-diisopropylphenyl group on the other, affords a far more active catalyst C1-sym. β-diketiminatoiron(II) alkyl complex than the authors' previously reported C2-sym. β-diketiminatoiron(II) alkyl complex (Ar1 = Ar2 = 2,4,6-(Me)3C6H2)/cyclopentylamine system. Comparative studies let the authors postulate that this superior activity of C1-sym. β-diketiminatoiron(II) alkyl complex , when compared with other complexes , is likely arising from steric effects and/or the coordinating ability of the ortho-methoxy substituent. The scope and limitations of this novel C1-sym. β-diketiminatoiron(II) alkyl complex are also presented.
- 20Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297– 3305, DOI: 10.1039/b508541aGoogle Scholar21Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracyWeigend, Florian; Ahlrichs, ReinhartPhysical Chemistry Chemical Physics (2005), 7 (18), 3297-3305CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)Gaussian basis sets of quadruple zeta valence quality for Rb-Rn are presented, as well as bases of split valence and triple zeta valence quality for H-Rn. The latter were obtained by (partly) modifying bases developed previously. A large set of more than 300 mols. representing (nearly) all elements-except lanthanides-in their common oxidn. states was used to assess the quality of the bases all across the periodic table. Quantities investigated were atomization energies, dipole moments and structure parameters for Hartree-Fock, d. functional theory and correlated methods, for which we had chosen Moller-Plesset perturbation theory as an example. Finally recommendations are given which type of basis set is used best for a certain level of theory and a desired quality of results.
- 21Becke, A. D. Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A 1988, 38, 3098– 3100, DOI: 10.1103/PhysRevA.38.3098Google Scholar22Density-functional exchange-energy approximation with correct asymptotic behaviorBecke, A. D.Physical Review A: Atomic, Molecular, and Optical Physics (1988), 38 (6), 3098-100CODEN: PLRAAN; ISSN:0556-2791.Current gradient-cor. d.-functional approxns. for the exchange energies of at. and mol. systems fail to reproduce the correct 1/r asymptotic behavior of the exchange-energy d. A gradient-cor. exchange-energy functional is given with the proper asymptotic limit. This functional, contg. only one parameter, fits the exact Hartree-Fock exchange energies of a wide variety of at. systems with remarkable accuracy, surpassing the performance of previous functionals contg. two parameters or more.
- 22(a) Staroverov, V. N.; Scuseria, G. E.; Tao, J.; Perdew, J. P. Comparative assessment of a new nonempirical density functional: Molecules and hydrogen-bonded complexes. J. Chem. Phys. 2003, 119, 12129– 12137, DOI: 10.1063/1.1626543Google Scholar23aComparative assessment of a new nonempirical density functional: Molecules and hydrogen-bonded complexesStaroverov, Viktor N.; Scuseria, Gustavo E.; Tao, Jianmin; Perdew, John P.Journal of Chemical Physics (2003), 119 (23), 12129-12137CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)A comprehensive study is undertaken to assess the nonempirical meta-generalized gradient approxn. (MGGA) of Tao, Perdew, Staroverov, and Scuseria (TPSS) against 14 common exchange-correlation energy functionals. Principal results are presented in the form of statistical summaries of deviations from expt. for the G3/99 test set (223 enthalpies of formation, 86 ionization potentials, 58 electron affinities, 8 proton affinities) and three addnl. test sets involving 96 bond lengths, 82 harmonic vibrational frequencies, and 10 hydrogen-bonded complexes, all computed using the 6-311++G(3df,3pd) basis. The TPSS functional matches, or exceeds in accuracy all prior nonempirical constructions and, unlike semiempirical functionals, consistently provides a high-quality description of diverse systems and properties. The computational cost of self-consistent MGGA is comparable to that of ordinary GGA, and exact exchange (unavailable in some codes) is not required. A one-parameter global hybrid version of the TPSS functional is introduced and shown to give further improvement for most properties.(b) Tao, J.; Perdew, J. P.; Staroverov, V. N.; Scuseria, G. E. Climbing the Density Functional Ladder: Nonempirical Meta--Generalized Gradient Approximation Designed for Molecules and Solids. Phys. Rev. Lett. 2003, 91, 146401, DOI: 10.1103/PhysRevLett.91.146401Google Scholar23bClimbing the Density Functional Ladder: Nonempirical Meta-Generalized Gradient Approximation Designed for Molecules and SolidsTao, Jianmin; Perdew, John P.; Staroverov, Viktor N.; Scuseria, Gustavo E.Physical Review Letters (2003), 91 (14), 146401/1-146401/4CODEN: PRLTAO; ISSN:0031-9007. (American Physical Society)The electron d., its gradient, and the Kohn-Sham orbital kinetic energy d. are the local ingredients of a meta-generalized gradient approxn. (meta-GGA). We construct a meta-GGA d. functional for the exchange-correlation energy that satisfies exact constraints without empirical parameters. The exchange and correlation terms respect two paradigms: one- or two-electron densities and slowly varying densities, and so describe both mols. and solids with high accuracy, as shown by extensive numerical tests. This functional completes the third rung of "Jacob's ladder" of approxns., above the local spin d. and GGA rungs.
- 23Becke, A. D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648– 5652, DOI: 10.1063/1.464913Google Scholar24Density-functional thermochemistry. III. The role of exact exchangeBecke, Axel D.Journal of Chemical Physics (1993), 98 (7), 5648-52CODEN: JCPSA6; ISSN:0021-9606.Despite the remarkable thermochem. accuracy of Kohn-Sham d.-functional theories with gradient corrections for exchange-correlation, the author believes that further improvements are unlikely unless exact-exchange information is considered. Arguments to support this view are presented, and a semiempirical exchange-correlation functional (contg. local-spin-d., gradient, and exact-exchange terms) is tested for 56 atomization energies, 42 ionization potentials, 8 proton affinities, and 10 total at. energies of first- and second-row systems. This functional performs better than previous functionals with gradient corrections only, and fits expt. atomization energies with an impressively small av. abs. deviation of 2.4 kcal/mol.
- 24(a) Harvey, J. N. In Principles and Applications of Density Functional Theory in Inorganic Chemistry I; Springer: Berlin, 2004; pp 151– 184.Google ScholarThere is no corresponding record for this reference.(b) Swart, M.; Gruden, M. Spinning around in Transition-Metal Chemistry. Acc. Chem. Res. 2016, 49, 2690– 2697, DOI: 10.1021/acs.accounts.6b00271Google Scholar25bSpinning around in Transition-Metal ChemistrySwart, Marcel; Gruden, MajaAccounts of Chemical Research (2016), 49 (12), 2690-2697CODEN: ACHRE4; ISSN:0001-4842. (American Chemical Society)A review. The great diversity and richness of transition metal chem., such as the features of an open d-shell, opened a way to numerous areas of scientific research and technol. applications. Depending on the nature of the metal and its environment, there are often several energetically accessible spin states, and the progress in accurate theor. treatment of this complicated phenomenon is presented in this Account. The spin state energetics of a transition metal complex can be predicted theor. on the basis of d. functional theory (DFT) or wave function based methodol., where DFT has advantages since it can be applied routinely to medium-to-large-sized mols. and spin-state consistent d. functionals are now available. Addnl. factors such as the effect of the basis set, thermochem. contributions, solvation, relativity, and dispersion, have been investigated by many researchers, but challenges in unambiguous assignment of spin states still remain. The first DFT studies showed intrinsic spin-state preferences of hybrid functionals for high spin and early generalized gradient approxn. functionals for low spin. Progress in the development of d. functional approxns. (DFAs) then led to a class of specially designed DFAs, such as OPBE, SSB-D, and S12g, and brought a very intriguing and fascinating observation that the spin states of transition metals and the SN2 barriers of org. mols. are somehow intimately linked. Among the many noteworthy results that emerged from the search for the appropriate description of the complicated spin state preferences in transition metals, we mainly focused on the examn. of the connection between the spin state and the structures or coordination modes of the transition metal complexes. Changes in spin states normally lead only to changes in the metal-ligand bond lengths, but to the best of our knowledge, the dapsox ligand showed the first example of a transition-metal complex where a change in spin state leads also to changes in the coordination, switching between pentagonal-bipyramidal and capped-octahedron. Moreover, we have summarized the results of the thorough study that cor. the exptl. assignment of the nature of the recently synthesized Sc3+ adduct of [FeIV(O)(TMC)]2+ (TMC = 1,4,8,11-tetramethylcyclam) and firmly established that the Sc3+-capped iron-oxygen complex corresponds to high-spin FeIII. Last, but not least, we have provided deeper insight and rationalization of the observation that unlike in metalloenzymes, where the FeIV-oxo is usually obsd. with high spin, biomimetic FeIV-oxo complexes typically have a intermediate spin state. Energy decompn. analyses on the trigonal-bypiramidal (TBP) and octahedral model systems with ammonia ligands have revealed that the interaction energy of the prepd. metal ion in the intermediate spin state is much smaller for the TBP structure. This sheds light on the origin of the intermediate spin state of the biomimetic TBP FeIV-oxo complexes.
- 25These orbital splitting energies are relatively small; spin-crossover complexes commonly have significantly larger values for Δ. See for example:Hauser, A. Ligand Field Theoretical Considerations. In Spin Crossover in Transition Metal Compounds I; Gütlich, P., Goodwin, H. A., Eds.; Springer Berlin Heidelberg: Berlin, Heidelberg, 2004, p 49.Google ScholarThere is no corresponding record for this reference.
- 26(a) Knizia, G. Intrinsic Atomic Orbitals: An Unbiased Bridge between Quantum Theory and Chemical Concepts. J. Chem. Theory Comput. 2013, 9, 4834– 4843, DOI: 10.1021/ct400687bGoogle Scholar27aIntrinsic Atomic Orbitals: An Unbiased Bridge between Quantum Theory and Chemical ConceptsKnizia, GeraldJournal of Chemical Theory and Computation (2013), 9 (11), 4834-4843CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Modern quantum chem. can make quant. predictions on an immense array of chem. systems. However, the interpretation of those predictions is often complicated by the complex wave function expansions used. Here we show that an exceptionally simple algebraic construction allows for defining at. core and valence orbitals, polarized by the mol. environment, which can exactly represent SCF wave functions. This construction provides an unbiased and direct connection between quantum chem. and empirical chem. concepts, and can be used, for example, to calc. the nature of bonding in mols., in chem. terms, from first principles. In particular, we find consistency with electronegativities (χ), C 1s core-level shifts, resonance substituent parameters (σR), Lewis structures, and oxidn. states of transition-metal complexes.(b) Knizia, G.; Klein, J. E. M. N. Electron Flow in Reaction Mechanisms—Revealed from First Principles. Angew. Chem., Int. Ed. 2015, 54, 5518– 5522, DOI: 10.1002/anie.201410637Google Scholar27bElectron Flow in Reaction Mechanisms-Revealed from First PrinciplesKnizia, Gerald; Klein, Johannes E. M. N.Angewandte Chemie, International Edition (2015), 54 (18), 5518-5522CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)The "curly arrow" of Robinson and Ingold is the primary tool for describing and rationalizing reaction mechanisms. Despite this approach's ubiquity and stellar success, its phys. basis has never been clarified and a direct connection to quantum chem. has never been found. Here we report that the bond rearrangements expressed by curly arrows can be directly obsd. in ab initio computations, as transformations of intrinsic bond orbitals (IBOs) along the reaction coordinate. Our results clarify that curly arrows are rooted in phys. reality-a notion which has been challenged before-and show how quantum chem. can directly establish reaction mechanisms in intuitive terms and unprecedented detail.
- 27Kamphuis, A. J.; Milocco, F.; Koiter, L.; Pescarmona, P. P.; Otten, E. Highly Selective Single-Component Formazanate Ferrate(II) Catalysts for the Conversion of CO2 into Cyclic Carbonates. ChemSusChem 2019, 12, 3635– 3641, DOI: 10.1002/cssc.201900740Google Scholar28Highly Selective Single-Component Formazanate Ferrate(II) Catalysts for the Conversion of CO2 into Cyclic CarbonatesKamphuis, Aeilke J.; Milocco, Francesca; Koiter, Luuk; Pescarmona, Paolo P.; Otten, EdwinChemSusChem (2019), 12 (15), 3635-3641CODEN: CHEMIZ; ISSN:1864-5631. (Wiley-VCH Verlag GmbH & Co. KGaA)The development of new families of active and selective single-component catalysts based on earth-abundant metal is of interest from a sustainable chem. perspective. In this context, anionic mono(formazanate) Fe(II) complexes bearing labile halide ligands, which possess both Lewis acidic and nucleophilic functionalities, were developed as novel single-component homogeneous catalysts for the reaction of CO2 with epoxides to produce cyclic carbonates. The influence of the halide ligand and the electronic properties of the formazanate ligand backbone on the catalytic activity are studied by employing the Fe(II) complexes with and without an addnl. nucleophile. Very high selectivity is achieved towards the formation of the cyclic carbonate products from various terminal and internal epoxides without the need of a cocatalyst. Crystallog. data are given.
- 28Chang, M.-C.; Roewen, P.; Travieso-Puente, R.; Lutz, M.; Otten, E. Formazanate Ligands as Structurally Versatile, Redox-Active Analogues of β-Diketiminates in Zinc Chemistry. Inorg. Chem. 2015, 54, 379– 388, DOI: 10.1021/ic5025873Google Scholar29Formazanate Ligands as Structurally Versatile, Redox-Active Analogues of β-Diketiminates in Zinc ChemistryChang, Mu-Chieh; Roewen, Peter; Travieso-Puente, Raquel; Lutz, Martin; Otten, EdwinInorganic Chemistry (2015), 54 (1), 379-388CODEN: INOCAJ; ISSN:0020-1669. (American Chemical Society)A range of tetrahedral bis(formazanate)zinc complexes with different steric and electronic properties of the formazanate ligands were synthesized. The solid-state structures for several of these were detd. by x-ray crystallog., which showed that complexes with sym., unhindered ligands prefer coordination to the Zn center via the terminal N atoms of the NNCNN ligand backbone. Steric or electronic modifications can override this preference and give rise to solid-state structures in which the formazanate ligand forms a 5-membered chelate by binding to the metal center via an internal N atom. In soln., these compds. show dynamic equil. that involve both 5- and 6-membered chelates. All compds. are intensely colored, and the effect of the ligand substitution pattern on the UV-visible absorption spectra was evaluated. Their cyclic voltammetry is reported, which shows that all compds. may be electrochem. reduced to radical anionic (L2Zn-) and dianionic (L2Zn2-) forms. While unhindered NAr substituents lie in the plane of the ligand backbone (Ar = Ph), the introduction of sterically demanding substituents (Ar = Mes) favors a perpendicular orientation in which the NMes group is no longer in conjugation with the backbone, resulting in hypsochromic shifts in the absorption spectra. The redox potentials in L2Zn compds. may be altered in a straightforward manner over a relatively wide range (∼700 mV) via the introduction of electron-donating or -withdrawing substituents on the formazanate framework.
- 29Chang, M. C.; Otten, E. Synthesis and ligand-based reduction chemistry of boron difluoride complexes with redox-active formazanate ligands. Chem. Commun. 2014, 50, 7431– 7433, DOI: 10.1039/C4CC03244FGoogle Scholar30Synthesis and ligand-based reduction chemistry of boron difluoride complexes with redox-active formazanate ligandsChang, M.-C.; Otten, E.Chemical Communications (Cambridge, United Kingdom) (2014), 50 (56), 7431-7433CODEN: CHCOFS; ISSN:1359-7345. (Royal Society of Chemistry)Mono(formazanate) boron difluoride complexes (LBF2), which show remarkably facile and reversible ligand-based redox-chem., were synthesized by transmetalation of bis(formazanate) zinc complexes with boron trifluoride. The one-electron redn. product [LBF2]-[Cp2Co]+ and a key intermediate for the transmetalation reaction, the six-coordinate zinc complex (L(BF3))2Zn were isolated and fully characterized.
- 30(a) Chang, M. C.; Chantzis, A.; Jacquemin, D.; Otten, E. Boron difluorides with formazanate ligands: redox-switchable fluorescent dyes with large stokes shifts. Dalton Trans. 2016, 45, 9477– 9484, DOI: 10.1039/C6DT01226DGoogle Scholar31aBoron difluorides with formazanate ligands: redox-switchable fluorescent dyes with large stokes shiftsChang, M.-C.; Chantzis, A.; Jacquemin, D.; Otten, E.Dalton Transactions (2016), 45 (23), 9477-9484CODEN: DTARAF; ISSN:1477-9226. (Royal Society of Chemistry)The synthesis of a series of (formazanate)boron difluorides and their 1-electron redn. products is described. The neutral compds. are fluorescent with large Stokes shifts. DFT calcns. suggest that a large structural reorganization accompanies photoexictation and accounts for the large Stokes shift. Redn. of the neutral boron difluorides occurs at the ligand and generates the corresponding radical anions. These complexes are non-fluorescent, allowing switching of the emission by changing the ligand oxidn. state.(b) Chang, M.-C. Formazanate as redox-active, structurally versatile ligand platform: Zinc and boron chemistry. PhD thesis, University of Groningen, 2016.Google ScholarThere is no corresponding record for this reference.
- 31Broere, D. L.; Coric, I.; Brosnahan, A.; Holland, P. L. Quantitation of the THF Content in Fe[N(SiMe3)2]2.xTHF. Inorg. Chem. 2017, 56, 3140– 3143, DOI: 10.1021/acs.inorgchem.7b00056Google Scholar32Quantitation of the THF Content in Fe[N(SiMe3)2]2·xTHFBroere, Daniel L. J.; Coric, Ilija; Brosnahan, Anna; Holland, Patrick L.Inorganic Chemistry (2017), 56 (6), 3140-3143CODEN: INOCAJ; ISSN:0020-1669. (American Chemical Society)The absence of residual solvent in metal precursors can be of key importance for the successful prepn. of metal complexes or materials. Herein, the authors describe methods for the quantitation of residual coordinated THF that binds to Fe[N(SiMe3)2]2, a commonly used iron synthon, when prepd. according to common literature procedures. A simple method for quantitation of the amt. of residual coordinated THF using 1H NMR spectroscopy is highlighted. Finally, a detailed synthetic procedure is described for the synthesis of THF-free Fe[N(SiMe3)2]2.
- 32Gilroy, J. B.; Ferguson, M. J.; McDonald, R.; Patrick, B. O.; Hicks, R. G. Formazans as β-diketiminate analogues. Structural characterization of boratatetrazines and their reduction to borataverdazyl radical anions. Chem. Commun. 2007, 126– 128, DOI: 10.1039/B609365EGoogle Scholar33Formazans as β-diketiminate analogues. Structural characterization of boratatetrazines and their reduction to borataverdazyl radical anionsGilroy, Joe B.; Ferguson, Michael J.; McDonald, Robert; Patrick, Brian O.; Hicks, Robin G.Chemical Communications (Cambridge, United Kingdom) (2007), (2), 126-128CODEN: CHCOFS; ISSN:1359-7345. (Royal Society of Chemistry)Formazans, R'NHN:CRN:NR' (R ≠ R' = p-tolyl, Ph) react with boron triacetate to produce boratatetrazines I, which can be reduced by cobaltocene to yield borataverdazyl radical anions, e.g., II-the first boron contg. verdazyl radicals.
- 33Bruker. APEX3, SAINT and SADABS; Bruker AXS Inc.: Madison, WI, 2016.Google ScholarThere is no corresponding record for this reference.
- 34Sheldrick, G. A short history of SHELX. Acta Crystallogr., Sect. A 2008, 64, 112– 122, DOI: 10.1107/S0108767307043930Google Scholar35A short history of SHELXSheldrick, George M.Acta Crystallographica, Section A: Foundations of Crystallography (2008), 64 (1), 112-122CODEN: ACACEQ; ISSN:0108-7673. (International Union of Crystallography)An account is given of the development of the SHELX system of computer programs from SHELX-76 to the present day. In addn. to identifying useful innovations that have come into general use through their implementation in SHELX, a crit. anal. is presented of the less-successful features, missed opportunities and desirable improvements for future releases of the software. An attempt is made to understand how a program originally designed for photog. intensity data, punched cards and computers over 10000 times slower than an av. modern personal computer has managed to survive for so long. SHELXL is the most widely used program for small-mol. refinement and SHELXS and SHELXD are often employed for structure soln. despite the availability of objectively superior programs. SHELXL also finds a niche for the refinement of macromols. against high-resoln. or twinned data; SHELXPRO acts as an interface for macromol. applications. SHELXC, SHELXD and SHELXE are proving useful for the exptl. phasing of macromols., esp. because they are fast and robust and so are often employed in pipelines for high-throughput phasing. This paper could serve as a general literature citation when one or more of the open-source SHELX programs (and the Bruker AXS version SHELXTL) are employed in the course of a crystal-structure detn.
- 35Sheldrick, G. Crystal structure refinement with SHELXL. Acta Crystallogr., Sect. C 2015, 71, 3– 8, DOI: 10.1107/S2053229614024218Google Scholar36Crystal structure refinement with SHELXLSheldrick, George M.Acta Crystallographica, Section C: Structural Chemistry (2015), 71 (1), 3-8CODEN: ACSCGG; ISSN:2053-2296. (International Union of Crystallography)The improvements in the crystal structure refinement program SHELXL have been closely coupled with the development and increasing importance of the CIF (Crystallog. Information Framework) format for validating and archiving crystal structures. An important simplification is that now only one file in CIF format (for convenience, referred to simply as 'a CIF') contg. embedded reflection data and SHELXL instructions is needed for a complete structure archive; the program SHREDCIF can be used to ext. the and files required for further refinement with SHELXL. Recent developments in SHELXL facilitate refinement against neutron diffraction data, the treatment of H atoms, the detn. of abs. structure, the input of partial structure factors and the refinement of twinned and disordered structures. SHELXL is available free to academics for the Windows, Linux and Mac OS X operating systems, and is particularly suitable for multiple-core processors.
Cited By
Smart citations by scite.ai include citation statements extracted from the full text of the citing article. The number of the statements may be higher than the number of citations provided by ACS Publications if one paper cites another multiple times or lower if scite has not yet processed some of the citing articles.
This article is cited by 28 publications.
- Bijoy Dey, Jordi Cirera, Liliana P. Ferreira, Paulo N. Martinho, Vadapalli Chandrasekhar. Peripheral Ligand Modification to Tune Spin Crossover Behavior of Fe(III) Complexes. Crystal Growth & Design 2025, 25
(9)
, 3065-3074. https://doi.org/10.1021/acs.cgd.5c00112
- Fangqing Gao, Binbin Wang, Yan Liu, Tengli Wang, Liang Wang, Li-Fei Zou, Shufang Xue, Yunnan Guo. Aggregation-Induced Spin-Crossover Switching in a Spin-Labile Iron(II) Complex. Inorganic Chemistry 2025, 64
(12)
, 5904-5912. https://doi.org/10.1021/acs.inorgchem.4c04701
- Amelia M. Wheaton, Jill A. Chipman, Rebecca K. Walde, Heike Hofstetter, John F. Berry. Chemically Separable Co(II) Spin-State Isomers. Journal of the American Chemical Society 2024, 146
(39)
, 26926-26935. https://doi.org/10.1021/jacs.4c08097
- Marco E. Reinhard, Baldeep K. Sidhu, Issiah B. Lozada, Natalia Powers-Riggs, Robert J. Ortiz, Hyeongtaek Lim, Rachel Nickel, Johan van Lierop, Roberto Alonso-Mori, Matthieu Chollet, Leland B. Gee, Patrick L. Kramer, Thomas Kroll, Sumana L. Raj, Tim B. van Driel, Amy A. Cordones, Dimosthenis Sokaras, David E. Herbert, Kelly J. Gaffney. Time-Resolved X-ray Emission Spectroscopy and Synthetic High-Spin Model Complexes Resolve Ambiguities in Excited-State Assignments of Transition-Metal Chromophores: A Case Study of Fe-Amido Complexes. Journal of the American Chemical Society 2024, 146
(26)
, 17908-17916. https://doi.org/10.1021/jacs.4c02748
- Bijoy Dey, Jordi Cirera, Sakshi Mehta, Liliana P. Ferreira, Ramar Arumugam, Abhishake Mondal, Paulo N. Martinho, Vadapalli Chandrasekhar. Steric Effects on Spin States in a Series of Fe(III) Complexes. Crystal Growth & Design 2023, 23
(9)
, 6668-6678. https://doi.org/10.1021/acs.cgd.3c00555
- Zoufeng Xu, Yu-Shien Sung, Elisa Tomat. Design of Tetrazolium Cations for the Release of Antiproliferative Formazan Chelators in Mammalian Cells. Journal of the American Chemical Society 2023, 145
(28)
, 15197-15206. https://doi.org/10.1021/jacs.3c02033
- Le Shi, Jedrzej Kobylarczyk, Katarzyna Dziedzic-Kocurek, Jan J. Stanek, Barbara Sieklucka, Robert Podgajny. Site Selectivity for the Spin States and Spin Crossover in Undecanuclear Heterometallic Cyanido-Bridged Clusters. Inorganic Chemistry 2023, 62
(18)
, 7032-7044. https://doi.org/10.1021/acs.inorgchem.3c00325
- Chenggang Jiang, Thomas S. Teets. Trimetallic Iridium–Nickel–Iridium Bis(formazanate) Assemblies. Inorganic Chemistry 2022, 61
(23)
, 8788-8796. https://doi.org/10.1021/acs.inorgchem.2c00726
- Alexandra C. Brown, Niklas B. Thompson, Daniel L. M. Suess. Evidence for Low-Valent Electronic Configurations in Iron–Sulfur Clusters. Journal of the American Chemical Society 2022, 144
(20)
, 9066-9073. https://doi.org/10.1021/jacs.2c01872
- Akram Ali, Saumitra Bhowmik, Suman K. Barman, Narottam Mukhopadhyay, Christine E. Glüer
(nee Schiewer), Francesc Lloret, Franc Meyer, Rabindranath Mukherjee. Iron(III) Complexes of a Hexadentate Thioether-Appended 2-Aminophenol Ligand: Redox-Driven Spin State Switchover. Inorganic Chemistry 2022, 61
(13)
, 5292-5308. https://doi.org/10.1021/acs.inorgchem.1c03992
- Yifeng Qin, Qiangui Zeng, Juan Zhu, Ziwen Liu, Guojie Zhang, Xiaohui Duan, Bo Wu. Synthesis, thermodynamics and catalytic properties of novel energetic coordination compounds based on nitrotriazolate-formazan ligand. Inorganic Chemistry Communications 2025, 179 , 114679. https://doi.org/10.1016/j.inoche.2025.114679
- Ivan Nemec, Radovan Herchel. The Role of Methyl Substitution in Spin Crossover of Fe(III) Complexes with Pentadentate Schiff Base Ligands. Inorganics 2025, 13
(2)
, 57. https://doi.org/10.3390/inorganics13020057
- Anayely Cruz-Galván, Juan Olguín. Spin crossover iron(II) complexes in non-hexacoordinate geometries. Trends in Chemistry 2025, 7
(1)
, 26-42. https://doi.org/10.1016/j.trechm.2024.11.001
- Kevin Tran, Patrick Sander, Maximilian Seydi Kilic, Jules Brehme, Ralf Sindelar, Franz Renz. Facile Approach for the Fabrication of Vapor Sensitive Spin Transition Composite Nanofibers. European Journal of Inorganic Chemistry 2024, 27
(33)
https://doi.org/10.1002/ejic.202400363
- Oleksandr Markin, Iurii Gudyma, Kamel Boukheddaden. Role of harmonic and anharmonic interactions in controlling the cooperativity of spin-crossover molecular crystals. Physical Review B 2024, 110
(17)
https://doi.org/10.1103/PhysRevB.110.174416
- Qian-Qian Jia, Xue-Jie Zhang, Liandong Zhu, Li-Zhi Huang. Fe(II) coordination transition regulates reductive dechlorination: The overlooked abiotic role of lactate. Water Research 2024, 254 , 121342. https://doi.org/10.1016/j.watres.2024.121342
- W. K. Damdoom, O. H. R. Al-Jeilawi. Synthesis, Characterization of Formazan Derivatives from Isoniazid and Study Their Antioxidant Activity and Molecular Docking. Russian Journal of Bioorganic Chemistry 2024, 50
(1)
, 86-94. https://doi.org/10.1134/S1068162024010163
- Sunita Birara, Shalu Saini, Moumita Majumder, Prem Lama, Shree Prakash Tiwari, Ramesh K. Metre. Design and synthesis of a solution-processed redox-active bis(formazanate) zinc complex for resistive switching applications. Dalton Transactions 2023, 52
(48)
, 18429-18441. https://doi.org/10.1039/D3DT02809G
- Maysaa A.M. Alhar. Preparation of Chalcone-Azo reagents and study of their chemical analysis,
thermal studies, and biological effects against fungi. Bionatura 2023, 8
(CSS 4)
, 1-13. https://doi.org/10.21931/RB/CSS/2023.08.04.99
- Xiu-e Jiang, Bo Wu, Bo Yang, Zai-chao Zhang, Ya-lin Yang, Hui-ying Du, Cong-ming Ma. Base-promoted decarboxylative gem-diazo-C-coupling: Synthesis, characterization and performance of nitrotriazolylformazans. Energetic Materials Frontiers 2023, 4
(2)
, 68-76. https://doi.org/10.1016/j.enmf.2023.06.002
- Dmitriy S. Yambulatov, Julia K. Voronina, Alexander S. Goloveshkin, Roman D. Svetogorov, Sergey L. Veber, Nikolay N. Efimov, Anna K. Matyukhina, Stanislav A. Nikolaevskii, Igor L. Eremenko, Mikhail A. Kiskin. Change in the Electronic Structure of the Cobalt(II) Ion in a One-Dimensional Polymer with Flexible Linkers Induced by a Structural Phase Transition. International Journal of Molecular Sciences 2023, 24
(1)
, 215. https://doi.org/10.3390/ijms24010215
- Sergey K. Dedushenko, Yury D. Perfiliev. On the correlation of the 57Fe Mӧssbauer isomer shift and some structural parameters of a substance. Hyperfine Interactions 2022, 243
(1)
https://doi.org/10.1007/s10751-022-01795-1
- N. A. Protasenko, E. V. Baranov, I. A. Yakushev, A. S. Bogomyakov, V. K. Cherkasov. Cobalt(III) Bis-o-semiquinone Complexes with p-Tolyl-Substituted Formazan Ligands: Synthesis, Structure, and Magnetic Properties. Russian Journal of Coordination Chemistry 2022, 48
(12)
, 819-827. https://doi.org/10.1134/S1070328422700129
- Maysaa A. M. Alhar. A Study of the Spectral, Chromatographic and Solubility Characteristics of New Analytical Reagents from Anil-Azo and Their Biological Effects on Bacteria. Biomedical and Pharmacology Journal 2022, 15
(2)
, 1115-1126. https://doi.org/10.13005/bpj/2447
- Natalia A. Protasenko, Maxim V. Arsenyev, Evgeny V. Baranov, Alyona A. Starikova, Artem S. Bogomyakov, Vladimir K. Cherkasov. Heteroligand o‐Semiquinonato Cobalt Complexes of 3‐Cyano and 3‐Nitroformazans. European Journal of Inorganic Chemistry 2022, 2022
(17)
https://doi.org/10.1002/ejic.202200152
- Bingzhi Liu, Tingyu Pan, Jiajun Liu, Li Feng, Yuning Chen, Huaili Zheng. Taping into the super power and magic appeal of ultrasound coupled with EDTA on degradation of 2,4,6-TCP by Fe0 based advanced oxidation processes. Chemosphere 2022, 288 , 132650. https://doi.org/10.1016/j.chemosphere.2021.132650
- Jireh Joy D. Sacramento, Therese Albert, Maxime Siegler, Pierre Moënne‐Loccoz, David P. Goldberg. An Iron(III) Superoxide Corrole from Iron(II) and Dioxygen. Angewandte Chemie 2022, 134
(2)
https://doi.org/10.1002/ange.202111492
- Jireh Joy D. Sacramento, Therese Albert, Maxime Siegler, Pierre Moënne‐Loccoz, David P. Goldberg. An Iron(III) Superoxide Corrole from Iron(II) and Dioxygen. Angewandte Chemie International Edition 2022, 61
(2)
https://doi.org/10.1002/anie.202111492
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.
Recommended Articles
Abstract
Figure 1
Figure 1. Common ligand field splitting diagrams for octahedral (A), square-planar (B), and tetrahedral (C) geometries and unusual ligand field splitting for the pseudo-tetrahedral geometries found in bis(formazanate)iron(II) complexes (D). (14)
Scheme 1
Scheme 1. Synthesis of Compounds 1–7Figure 2
Figure 2. Crystal structure of compound 3 showing the π-stacking interactions between the mesityl rings. The Fe center, ligand backbone, and the mesityl rings are shown as 50% probability ellipsoids and the remaining atoms as wireframe; hydrogen atoms are removed for clarity.
Figure 3
Figure 3. Molecular structure of compounds 6a and 6b showing 50% probability ellipsoids; hydrogen atoms omitted for clarity. The inset for each shows the Fe(NNCNN)2 core of the structure with the N–Fe–N planes and the dihedral angle.
Figure 4
Figure 4. Crystal structure of compound 6b illustrating the π-stacking interactions between the aromatic rings, showing 50% probability ellipsoids. Parts of the molecule are shown as wireframe, and hydrogen atoms are removed for clarity.
Figure 5
Figure 5. 57Fe Mössbauer spectra at 80 K in the solid state of 6b (A), 6a (B), and a powder sample of 6 before (C) and after (D) heating to 400 K for SQUID measurements. The red line in the spectrum of heated 6 represents the main species with 82% area, and the gray subspectra are unknown impurities.
Figure 6
Figure 6. Magnetic susceptibility data for a powder sample of 6 in the solid state (heating to 400 K and subsequent cooling). The solid black line shows the best fit curve for S = 1 with the parameters g = 2.10 and D = 11.2 cm–1 (100% IS). The dashed red line shows the spin-only value for an S = 2 system.
Figure 7
Figure 7. Molecular structures of 7 showing 50% probability ellipsoids. One of the N–Ph rings is shown as wireframe, and hydrogen atoms are omitted for clarity.
Figure 8
Figure 8. Temperature dependence of the high-spin fraction (γHS) of compounds 1–5 and 7 in toluene-d8, including error bars for T1/2 (γHS = 0.5). The liquid range for toluene is indicated with the color gradient at the temperature axis.
Figure 9
Figure 9. 1H NMR spectra of 7 recorded between 247 and 397 K (toluene-d8, 500 MHz).
Figure 10
Figure 10. Representation of intrinsic bonding orbitals at the BP86/def2-TZVP minima, both with (7calc; (A)) and without Fe–O interaction (7′calc; (B)).
Figure 11
Figure 11. UV/vis spectra in toluene for (A) compound 6 recorded between 183 and 293 K and (B) compound 7 recorded between 293 and 383 K.
References
This article references 35 other publications.
- 1(a) Poli, R. Open-Shell Organometallics as a Bridge between Werner-Type and Low-Valent Organometallic Complexes. The Effect of the Spin State on the Stability, Reactivity, and Structure. Chem. Rev. 1996, 96, 2135– 2204, DOI: 10.1021/cr95003431aOpen-Shell Organometallics as a Bridge between Werner-Type and Low-Valent Organometallic Complexes. The Effect of the Spin State on the Stability, Reactivity, and StructurePoli, RinaldoChemical Reviews (Washington, D. C.) (1996), 96 (6), 2135-2204CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review with > 772 refs.(b) Hawrelak, E. J.; Bernskoetter, W. H.; Lobkovsky, E.; Yee, G. T.; Bill, E.; Chirik, P. J. Square planar vs tetrahedral geometry in four coordinate iron(II) complexes. Inorg. Chem. 2005, 44, 3103– 3111, DOI: 10.1021/ic048202+1bSquare Planar vs Tetrahedral Geometry in Four Coordinate Iron(II) ComplexesHawrelak, Eric J.; Bernskoetter, Wesley H.; Lobkovsky, Emil; Yee, Gordon T.; Bill, Eckhard; Chirik, Paul J.Inorganic Chemistry (2005), 44 (9), 3103-3111CODEN: INOCAJ; ISSN:0020-1669. (American Chemical Society)The geometric preferences of a family of four coordinate, iron(II) d6 complexes of the general form L2FeX2 have been systematically evaluated. Treatment of Fe2(Mes)4 (Mes = 2,4,6-Me3C6H2) with monodentate phosphine and phosphite ligands furnished square planar trans-P2Fe(Mes)2 derivs. Identification of the geometry has been accomplished by a combination of soln. and solid-state magnetometry and, in two cases (P = PMe3, PEt2Ph), x-ray diffraction. In contrast, both tetrahedral and square planar coordination has been obsd. upon complexation of chelating phosphine ligands. A combination of crystallog. and magnetic susceptibility data for (depe)Fe(Mes)2 (depe = 1,2-bis(diethylphosphino)ethane) established a tetrahedral mol. geometry whereas SQUID magnetometry and Moessbauer spectroscopy on samples of (dppe)Fe(Mes)2 (dppe = 1,2-bis(diphenylphosphino)ethane) indicated a planar mol. When dissolved in chlorinated solvents, the latter compd. promotes chlorine atom abstraction, forming tetrahedral (dppe)Fe(Mes)Cl and (dppe)FeCl2. Ligand substitution reactions have been studied for both structural types and are rapid on the NMR time scale at ambient temp.
- 2Hartwig, J. F. Organotransition Metal Chemistry: From Bonding to Catalysis; University Science Books: Sausalito, CA, 2010.There is no corresponding record for this reference.
- 3Cirera, J.; Ruiz, E.; Alvarez, S. Stereochemistry and spin state in four-coordinate transition metal compounds. Inorg. Chem. 2008, 47, 2871– 2889, DOI: 10.1021/ic702276k3Stereochemistry and Spin State in Four-Coordinate Transition Metal CompoundsCirera, Jordi; Ruiz, Eliseo; Alvarez, SantiagoInorganic Chemistry (2008), 47 (7), 2871-2889CODEN: INOCAJ; ISSN:0020-1669. (American Chemical Society)A systematic DFT computational study of the stereochem. assocd. with each spin state for first transition series four-coordinate dn (n = 0-10) homoleptic metal complexes is presented. The stereochem. of [MMe4]x- complexes in the 21 spin configurations analyzed can be predicted from the d orbital occupation in the ideal tetrahedral geometry, grouped in three families with tetrahedral, square planar, or intermediate structures that can be described in some cases as sawhorses. The effect of the following factors on the spin state and stereochem. preferences has also been studied: (a) substitution of the σ-donor Me ligands by π-donor chlorides, (b) a high (+ 4) oxidn. state of the metal, and (c) substitution of the metal atom by a second transition series one. Through those factors, low-spin tetrahedral structures can be achieved, as summarized by a magic cube.
- 4(a) Collman, J. P.; Hoard, J. L.; Kim, N.; Lang, G.; Reed, C. A. Synthesis, stereochemistry, and structure-related properties of alpha, beta, gamma, delta-tetraphenylporphinatoiron(II). J. Am. Chem. Soc. 1975, 97, 2676– 2681, DOI: 10.1021/ja00843a0154aSynthesis, stereochemistry, and structure-related properties of α,β,γ,δ-tetraphenylporphinatoiron(II)Collman, James P.; Hoard, J. L.; Kim, Nancy; Lang, George; Reed, Christopher A.Journal of the American Chemical Society (1975), 97 (10), 2676-81CODEN: JACSAT; ISSN:0002-7863.Addnl. data considered in abstracting and indexing are available from a source cited in the original document. Redn. of chloro-α,β,γ,δ-tetraphenylporphinatoiron(III) by bis(acetylacetonato)chromium(II) under anaerobic conditions yielded α,β,γ,δ-tetraphenylporphinatoiron(II), Fe(TPP), which crystd. from benzene-EtOH soln. in the tetragonal system. Pertinent crystal data are: space group, I‾42d; α 15.080(9) and c 14.043(9) Å; d. (exptl.) = 1.38 and d. (calcd.) = 1.39 for Z = 4; required mol. symmetry S4. Intensities of 1164 reflections having (sin θ)/λ < 0.71 Å-1, recorded with Zr-filtered Mo Kα radiation on a Picker FACS-I diffractometer, were used in anisotropic full-matrix least-squares refinement of the 112 structural parameters. The length of the FeII-N bonds in the intermediate-spin (S = 1) Fe(TPP) mol. is 1.972(4) Å, as compared with 2.086(4) Å in high-spin (S = 2) 2-methylimidazole-α,β,γ,δ-tetraphenylporphinatoiron(II), with 2.004(3) Å in low-spin (S = 0) bis(piperidine)-α,β,γ,δ-tetraphenylporphinatoiron(II, and with Cu-N = 1.981(7) Å in Cu(TPP). The foregoing assignment of spin states to the 3 Fe(II) porphyrins is fully confirmed by the Moessbauer spectra recorded for them in zero and in large applied magnetic fields at 4.2-300°K. Recrystd. Fe(TPP) conforming to the theor. compn. has an effective magnetic moment of ∼4.4 μB at room temp. Detn. of the anisotropic magnetic susceptibilities as a function of temp. for the tetragonal crystal, when considered in conjunction with the Moessbauer data, should lead to a more detailed specification of the electronic configuration of the intermediate-spin iron(II) atom in Fe(TPP) than that given in this paper.(b) Kirner, J. F.; Dow, W.; Scheidt, W. R. Molecular stereochemistry of two intermediate-spin complexes. Iron(II) phthalocyanine and manganese(II) phthalocyanine. Inorg. Chem. 1976, 15, 1685– 1690, DOI: 10.1021/ic50161a0424bMolecular stereochemistry of two intermediate-spin complexes. Iron(II) phthalocyanine and manganese(II) phthalocyanineKirner, John F.; Dow, W.; Scheidt, W. RobertInorganic Chemistry (1976), 15 (7), 1685-90CODEN: INOCAJ; ISSN:0020-1669.The mol. stereochem. of Fe(II) phthalocyanine (Pc) and Mn(II) phthalocyanine was detd. by x-ray diffraction methods. The phthalocyanine ligand constrains the metal ion to effectively square-planar coordination and to an intermediate spin state. The FeII-N bond distance of 1.926(1) and the MnII-N bond length of 1.938(3) Å are wholly consistent with the assignment of an intermediate-spin ground state. Both complexes crystallize as the β polymorph. Crystal data are: for FePc, space group P21/a, a 19.392(5), b 4.786(2), c 14.604(4) Å, β 120.85(1)°, d.(exptl.) 1.61, d.(calcd.) 1.623, and Z = 2, required mol. symmetry ‾1; for MnPc, space group P21/a, a 19.400(4), b 4.761(2), c 14.613(3) Å, β 120.74(1)°, d.(exptl.) 1.61, d.(calcd.) 1.625, and Z = 2, required mol. symmetry ‾1. Final discrepancy indices are: FePc, R1 0.045, R2 0.057; MnPc, R1 0.066, R2 0.066.(c) Strauss, S. H.; Silver, M. E.; Ibers, J. A. Iron(II) octaethylchlorine: structure and ligand affinity comparison with its porphyrin and isobacteriochlorin homologs. J. Am. Chem. Soc. 1983, 105, 4108– 4109, DOI: 10.1021/ja00350a0694cIron(II) octaethylchlorine: structure and ligand affinity comparison with its porphyrin and isobacteriochlorin homologsStrauss, Steven H.; Silver, Michael E.; Ibers, James A.Journal of the American Chemical Society (1983), 105 (12), 4108-9CODEN: JACSAT; ISSN:0002-7863.trans-7,8-Dihydro-2,3,7,8,12,13,17,18-octaethylporphyrinatoiron(II), Fe(OEC), is orthorhombic, space group Pbcn, with a 21.880(9), b 15.795(6), and c 8.554(4) Å at -150°; Z = 4. The structure was refined anisotropically by least-squares to a final R(F2) = 0.105. The Fe and 4 N atoms are rigorously planar, while the rest of the chlorin macrocycle is significantly S4 ruffled. The affinity of Fe(OEC) and its octaethylporphyrin and octaethylisobacteriochlorin homologs for the weak σ-donor ligands THF and EtSH is strongly macrocycle dependent. This dependence may have a structural basis whereby reduced macrocycles more easily distort, adjusting their core size to accommodate the changing requirements of the metal upon complexation of a weak ligand.
- 5Chatt, J.; Shaw, B. L. Alkyls and aryls of transition metals. Part IV. Cobalt(II) and iron(II) derivatives. J. Chem. Soc. 1961, 285, DOI: 10.1039/jr96100002855Alkyls and aryls of transition metals. IV. Cobalt(II) and iron(II) derivativesChatt, J.; Shaw, B. L.Journal of the Chemical Society (1961), (), 285-90CODEN: JCSOA9; ISSN:0368-1769.cf. CA 54, 20936d. The prepn. and properties of stable organometallic complexes of the types trans-[MR2-(PR2')2] (M = Co, Fe; R, R' = org. radicals) were described. The groups R were ortho-substituted aryl groups, where the substituents were somewhat bulky. These were the first planar complexes of Co(II) and Fe(II) that had only monodentate ligands. Their configurations were established by their magnetic and elec. dipole moments. Possible reasons for the unusual configurations and stabilities of these organometallic derivs. were discussed.
- 6(a) Nijhuis, C. A.; Jellema, E.; Sciarone, T. J. J.; Meetsma, A.; Budzelaar, P. H. M.; Hessen, B. First-Row Transition Metal Bis(amidinate) Complexes; Planar Four-Coordination of FeII Enforced by Sterically Demanding Aryl Substituents Eur. Eur. J. Inorg. Chem. 2005, 2005, 2089– 2099, DOI: 10.1002/ejic.200500094There is no corresponding record for this reference.(b) Wurzenberger, X.; Piotrowski, H.; Klufers, P. A stable molecular entity derived from rare iron(II) minerals: the square-planar high-spin d6 Fe(II)O4 chromophore Angew. Angew. Chem., Int. Ed. 2011, 50, 4974– 4978, DOI: 10.1002/anie.2010068986bA Stable Molecular Entity Derived from Rare Iron(II) Minerals: The Square-Planar High-Spin-d6 FeIIO4 ChromophoreWurzenberger, Xaver; Piotrowski, Holger; Kluefers, PeterAngewandte Chemie, International Edition (2011), 50 (21), 4974-4978, S4974/1-S4974/4CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A strongly alk. (LiOH or NaOH) aq. soln. of furanose mimic anhydroerythritol (meso-oxolane-3,4-diol, H2L) and FeCl2 gave a reddish-blue soln., from which, depending on the conditions, orange-red crystals of the tetrahydrates Li2[FeL2]·4H2O (1) or Na2[FeL2]·4H2O (2), or violet crystals of the nonahydrate Na2[FeL2]·9H2O (3) were grown. The bidentate oxolanediolato ligand stabilizes the rare high-spin ferrous SP-4 coordination in 1 as revealed by x-ray crystallog. and SQUID magnetometer measurements. The structure of 2 is also SP-4 as revealed by x-ray crystallog. Compd. 3 is slightly distorted toward tetrahedron. Three related mononuclear high-spin [FeIIX4]2- (X = Cl, F, OH) species and their high-spin [MnIIX4]2- analogs were analyzed in a DFT approach using the unrestricted-B3LYP/tzvp level of theory. Enhanced Jahn-Teller flattening can stabilize square planar high-spin d6 centers despite their "wrong" spin state by a marked sepn. of the neg. charge on the ligand atoms and the stereochem. active d electrons. This result reveals square-planar high-spin centers to be building blocks of reasonable stability.(c) Cantalupo, S. A.; Fiedler, S. R.; Shores, M. P.; Rheingold, A. L.; Doerrer, L. H. High-spin square-planar Co(II) and Fe(II) complexes and reasons for their electronic structure. Angew. Chem., Int. Ed. 2012, 51, 1000– 1005, DOI: 10.1002/anie.2011060916cHigh-Spin Square-Planar CoII and FeII Complexes and Reasons for Their Electronic StructureCantalupo, Stefanie A.; Fiedler, Stephanie R.; Shores, Matthew P.; Rheingold, Arnold L.; Doerrer, Linda H.Angewandte Chemie, International Edition (2012), 51 (4), 1000-1005, S1000/1-S1000/18CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)The high-spin, square planar complexes [K(DME)2]2[M(ddfp)2] (M = CoII, FeII, ddfp = perfluoropinacolate) were prepd. and characterized by x-ray crystallog., temp.-dependent magnetic susceptibility, cyclic voltammetry, and other methods. The tetrahedral analog [K(DME)2]2[Zn(ddfp)2] was also prepd. and characterized crystallog. The combination of high-spin electron configuration and square-planar geometry is made possible by ligand constraints that generate five non-degenerate d-orbitals with ligand-based π-donation, a relatively weak ligand-field splitting, and no intervening ligand-based π-acceptor mol. orbitals.(d) Pinkert, D.; Demeshko, S.; Schax, F.; Braun, B.; Meyer, F.; Limberg, C. A Dinuclear Molecular Iron(II) Silicate with Two High-Spin Square-Planar FeO4 Units. Angew. Chem., Int. Ed. 2013, 52, 5155– 5158, DOI: 10.1002/anie.2012096506dA Dinuclear Molecular Iron(II) Silicate with Two High-Spin Square-Planar FeO4 UnitsPinkert, Denise; Demeshko, Serhiy; Schax, Fabian; Braun, Beatrice; Meyer, Franc; Limberg, ChristianAngewandte Chemie, International Edition (2013), 52 (19), 5155-5158CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)Dinuclear iron complex [L'2Fe2][Na(OEt2)2]2 (1) (H3L' = ligand with elimination of "Me2SiO" from 3-(1,1-dimethylethyl)-3-[(hydroxydimethylsilyl)oxy]-1,1,5,5-tetramethyl-1,5-trisiloxanediol) was synthesized and studied by x-ray diffraction anal., temp.-dependent magnetic susceptibility and Mossbauer spectroscopy.(e) Liu, Y.; Luo, L.; Xiao, J.; Wang, L.; Song, Y.; Qu, J.; Luo, Y.; Deng, L. Four-Coordinate Iron(II) Diaryl Compounds with Monodentate N-Heterocyclic Carbene Ligation: Synthesis, Characterization, and Their Tetrahedral-Square Planar Isomerization in Solution. Inorg. Chem. 2015, 54, 4752– 4760, DOI: 10.1021/acs.inorgchem.5b001386eFour-Coordinate Iron(II) Diaryl Compounds with Monodentate N-Heterocyclic Carbene Ligation: Synthesis, Characterization, and Their Tetrahedral-Square Planar Isomerization in SolutionLiu, Yuesheng; Luo, Lun; Xiao, Jie; Wang, Lei; Song, You; Qu, Jingping; Luo, Yi; Deng, LiangInorganic Chemistry (2015), 54 (10), 4752-4760CODEN: INOCAJ; ISSN:0020-1669. (American Chemical Society)The salt elimination reactions of (IPr2Me2)2FeCl2 (IPr2Me2 = 1,3-diisopropyl-4,5-dimethylimidazol-2-ylidene) with the corresponding aryl Grignard reagents afford [(IPr2Me2)2FeAr2] (Ar = Ph, 3; C6H4-p-Me, 4; C6H4-p-tBu, 5; C6H3-3,5-(CF3)2, 6) in good yields. X-ray crystallog. studies revealed the presence of both tetrahedral and trans square planar isomers for 3 and 6 and the tetrahedral structures for 4 and 5. Magnetic susceptibility and 57Fe Mossbauer spectrum measurements on the solid samples indicated the high-spin (S = 2) and intermediate-spin (S = 1) nature of the tetrahedral and square planar structures, resp. Soln. property studies, including soln. magnetic susceptibility measurement, variable-temp. 1H and 19F NMR, and absorption spectroscopy, on 3-6, as well as an 57Fe Mossbauer spectrum study on a frozen THF soln. of tetrahedral [(IPr2Me2)257FePh2] suggest the coexistence of tetrahedral and trans square planar structures in soln. phase. D. functional theory calcns. on (IPr2Me2)2FePh2 disclosed that the tetrahedral and trans square planar isomers are close in energy and that the geometry isomerization can occur by spin-change-coupled geometric transformation on four-coordinate iron(II) center.
- 7(a) Holm, R. H.; Chakravorty, A.; Theriot, L. J. The Synthesis, Structures, and Solution Equilibria of Bis(pyrrole-2-aldimino)metal(II) Complexes. Inorg. Chem. 1966, 5, 625– 635, DOI: 10.1021/ic50038a0287aThe synthesis, structures, and solution equilibria of bis-(pyrrole-2-aldimino)metal(II) complexesHolm, R. H.; Chakravorty, A.; Theriot, L. J.Inorganic Chemistry (1966), 5 (4), 625-35CODEN: INOCAJ; ISSN:0020-1669.Synthesis of an extensive series of bis(pyrrole-2-aldimino)metal(II) complexes with M(II) = Co, Ni, Pd, Cu, and Zn and various alkyl groups (R) appended to the azomethine N was effected by a nonaq. chelation reaction in tetrahydrofuran. Preliminary single crystal x-ray results for complexes with R = tert-Bu reveal that Co, Ni, and Zn complexes are isomorphous, but appreciable differences in the cell consts. of the Ni complex indicate that it is not truly isostructural with the tetrahedral Co and Zn complexes. The Cu complex exists in 2 cryst. modifications, neither of which is isomorphous with the Co-Ni-Zn series. Spectral and magnetic studies in soln. show that the tert-Bu Co complex is tetrahedral whereas the corresponding Cu complex is distorted from planarity to an unknown extent. Cu complexes with less bulky R groups are planar. The tert-Bu Ni complex is pseudo-tetrahedral; complexes with sec-alkyl groups such as iso-Pr are involved in a configurational equil. between planar and pseudo-tetrahedral forms. The paramagnetic Ni complexes show large isotropic proton hyperfine contact shifts. Spin d. calcns. for the coordinated ligand system are used as the basis of proton resonance assignments. It is concluded that in the pseudo-tetrahedral form spin imbalance exists in the highest filled ligand π-mol. orbital, and that, in addn., there is an underlying spin imbalance in the highest filled σ-mol. orbital, the result of which is observable in the proton resonance spectra. Thermodynamic parameters characterizing the structural change were obtained for the Ni complexes from the temp. dependence of the proton contact shifts. A quant. comparison of the stabilization of tetrahedral Ni(II) by pyrrole-2-aldimine, salicylaldimine, and β-keto amine ligand systems is presented.(b) Wolny, J. A.; Rudolf, M. F.; Ciunik, Z.; Gatner, K.; Wołowiec, S. Cobalt(II) triazene 1-oxide bis(chelates). A case of planar (low spin)–tetrahedral (high spin) isomerism. J. Chem. Soc., Dalton Trans. 1993, 1611– 1622, DOI: 10.1039/DT99300016117bCobalt(II) triazene 1-oxide bis(chelates). A case of planar (low spin)-tetrahedral (high spin) isomerismWolny, Juliusz A.; Rudolf, Mikolaj F.; Ciunik, Zbigniew; Gatner, Kazimierz; Wolowiec, StanislawJournal of the Chemical Society, Dalton Transactions: Inorganic Chemistry (1972-1999) (1993), (10), 1611-22CODEN: JCDTBI; ISSN:0300-9246.High- and low-spin [CoL2] (HL = 3-phenyl-1-triazene 1-oxide derivs.) were prepd. and isolated. The low-spin complexes possess square-planar structure and the high-spin complexes are tetrahedral. The mol. structure of high-spin [CoL2] [HL = 3-(4-methylphenyl)-1-methyl-1-triazene 1-oxide] was detd.; triclinic, space group P‾1, a 7.970(5), b 10.174(5), c 11.676(5) Å, α 87.18(4), β 74.31(4), γ 74.06(4)°, Z = 2, R = 0.041. For some complexes the isolation of both planar (low-spin) and tetrahedral (high-spin) isomers or of their conglomerates is possible depending on the synthesis conditions. The crystal structure of square-planar [Ni(ON(Me)NNC6H4Me-4)2], which is isomorphous with the low-spin isomer of [Co(ON(Me)NNC6H4Me-4)2], was detd.: triclinic, space group P‾1, a 7.495(2), b 7.694(5), c 8.612(3) Å, α 64.64(5), β 87.84(2), γ 78.64(3)°, Z = 1, R = 0.0271. The complexes exhibit a planar-tetrahedral equil. in noncoordinating solvents, the ΔHΘ and ΔSΘ values of which, detd. from soln. magnetic susceptibility measurements, at 1-15 kJ mol-1 and 5-30 J K-1 mol-1, resp. The electrochem. properties of the complexes are given.(c) Ingleson, M. J.; Pink, M.; Fan, H.; Caulton, K. G. Exploring the reactivity of four-coordinate PNPCoX with access to three-coordinate spin triplet PNPCo. Inorg. Chem. 2007, 46, 10321– 10334, DOI: 10.1021/ic701171p7cExploring the reactivity of four-coordinate PNPCoX with access to three-coordinate spin triplet PNPCoIngleson, Michael J.; Pink, Maren; Fan, Hongjun; Caulton, Kenneth G.Inorganic Chemistry (2007), 46 (24), 10321-10334CODEN: INOCAJ; ISSN:0020-1669. (American Chemical Society)(PNP)CoX, where PNP is (tBu2PCH2SiMe2)2N- and X is Cl, I, N3, OAr, OSO2CF3, and N(H)Ar, are reported. Some of these show magnetic susceptibility, color, and 1H NMR evidence of being in equil. between a blue, tetrahedral S = 3/2 state and a red, planar S = 1/2 state; the equil. populations are influenced by subtle solvent effects (e.g., benzene and cyclohexane are different), as well as by temp. Attempted oxidn. to Co(III) with O2 occurs instead at phosphorus, giving [P(O)NP(O)]CoX species. The single O-atom transfer reagent PhI:O likewise oxidizes P. Even I2 oxidizes P to give the pendant phosphonium species (tBu2P(I)CH2SiMe2NSiMe2CH2PtBu2)CoI2 with a tetrahedral S = 3/2 cobalt; the solid-state structure shows intermol. PI···ICo interactions. Attempted alkyl metathesis of PNPCoX inevitably results in redn., forming PNPCo, which is a spin triplet with planar T-shaped coordination geometry with no agostic interaction. Triplet PNPCo binds N2(weakly) and CO (whose low CO stretching frequency indicates strong PNP → Co donor power), but not ethene or MeCCMe.
- 8Gaazo, J. Plasticity of the coordination sphere of copper(II) complexes, its manifestation and causes. Coord. Chem. Rev. 1976, 19, 253– 297, DOI: 10.1016/S0010-8545(00)80317-3There is no corresponding record for this reference.
- 9(a) Cambi, L.; Szegö, L. Über die magnetische Susceptibilität der komplexen Verbindungen. Ber. Dtsch. Chem. Ges. B 1931, 64, 2591– 2598, DOI: 10.1002/cber.19310641002There is no corresponding record for this reference.(b) Kahn, O.; Martinez, C. J. Spin-Transition Polymers: From Molecular Materials Toward Memory Devices. Science 1998, 279, 44– 48, DOI: 10.1126/science.279.5347.449bSpin-transition polymers: from molecular materials toward memory devicesKahn, O.; Martinez, C. JayScience (Washington, D. C.) (1998), 279 (5347), 44-48CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)A review with 53 refs. Some 3dn (4 ≤ n ≤ 7) transition metal compds. exhibit a cooperative transition between a low-spin (LS) and a high-spin (HS) state. This transition is abrupt and occurs with a thermal hysteresis, which confers a memory effect on the system. The intersite interactions and thus the cooperativity are magnified in polymeric compds. such as [Fe(Rtrz)3]A2·nH2O in which the Fe2+ ions are triply bridged by 4-R-substituted-1,2,4-triazole mols. Moreover, in these compds., the spin transition is accompanied by a well-pronounced change of color between violet in the LS state and white in the HS state. The transition temps. of these materials can be fine tuned, using an approach based on the concept of a mol. alloy. In particular, it is possible to design a compd. for which room temp. falls in the middle of the thermal hysteresis loop. These materials have many potential applications, for example, as temp. sensors, as active elements of various types of displays, and in information storage and retrieval.(c) Gütlich, P.; Garcia, Y.; Goodwin, H. A. Spin crossover phenomena in Fe(ii) complexes. Chem. Soc. Rev. 2000, 29, 419– 427, DOI: 10.1039/b003504l9cSpin crossover phenomena in Fe(II) complexesGutlich, Philipp; Garcia, Yann; Goodwin, Harold A.Chemical Society Reviews (2000), 29 (6), 419-427CODEN: CSRVBR; ISSN:0306-0012. (Royal Society of Chemistry)A review with 58 refs. The behavior of spin crossover compds. is among the most striking and fascinating shown by relatively simple mol. species. This review aims to draw attention to the various ways in which spin crossover phenomena are manifested in Fe(II) complexes, to offer some rationalization for these, and to highlight their possible applications. Typical examples have been selected along with more recent ones to give an overall view of the scope and development of the area. The article is structured to provide the basic material for those who wish to enter the field of spin crossover.(d) Sato, O.; Tao, J.; Zhang, Y. Z. Control of magnetic properties through external stimuli. Angew. Chem., Int. Ed. 2007, 46, 2152– 2187, DOI: 10.1002/anie.2006022059dControl of magnetic properties through external stimuliSato, Osamu; Tao, Jun; Zhang, Yuan-ZhuAngewandte Chemie, International Edition (2007), 46 (13), 2152-2187CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. The magnetic properties of many magnetic materials can be controlled by external stimuli. The principal focus here is on the thermal, photochem., electrochem., and chem. of phase transitions that involve changes in magnetization. Mol. compds. described herein range from metal complexes through pure org. compds. to composite materials. Most of the review is devoted to the properties. valence-tautomeric compds., mol. magnets, and spin-crossover complexes, which could find future applications in memory devices or optical switches.(e) Bousseksou, A.; Molnar, G.; Salmon, L.; Nicolazzi, W. Molecular spin crossover phenomenon: recent achievements and prospects. Chem. Soc. Rev. 2011, 40, 3313– 3335, DOI: 10.1039/c1cs15042a9eMolecular spin crossover phenomenon: Recent achievements and prospectsBousseksou, Azzedine; Molnar, Gabor; Salmon, Lionel; Nicolazzi, WilliamChemical Society Reviews (2011), 40 (6), 3313-3335CODEN: CSRVBR; ISSN:0306-0012. (Royal Society of Chemistry)A review. Recently we assisted a strong renewed interest in the fascinating field of mol. spin crossover complexes by (1) the emergence of nanosized spin crossover materials through direct synthesis of coordination nanoparticles and nanopatterned thin films as well as by (2) the use of novel sophisticated high spatial and temporal resoln. exptl. techniques and theor. approaches for the study of spatiotemporal phenomena in cooperative spin crossover systems. Besides generating new fundamental knowledge on size-redn. effects and the dynamics of the spin crossover phenomenon, this research aims also at the development of practical applications such as sensor, display, information storage and nanophotonic devices. In this crit. review, we discuss recent work in the field of mol.-based spin crossover materials with a special focus on these emerging issues, including chem. synthesis, phys. properties and theor. aspects as well (223 refs.).
- 10Halcrow, M. A. The spin-states and spin-transitions of mononuclear iron(II) complexes of nitrogen-donor ligands. Polyhedron 2007, 26, 3523– 3576, DOI: 10.1016/j.poly.2007.03.03310The spin-states and spin-transitions of mononuclear iron(II) complexes of nitrogen-donor ligandsHalcrow, Malcolm A.Polyhedron (2007), 26 (14), 3523-3576CODEN: PLYHDE; ISSN:0277-5387. (Elsevier B.V.)A review of mononuclear iron(II) complexes with heterocyclic N-donor ligation is presented. A brief introduction to spin-crossover chem. and low-temp. spin-trapping is provided, since many of these compds. undergo thermal spin-transitions upon cooling or heating. These are highlighted, and the structural changes underlying spin-crossover are discussed where this is known. Materials showing spin-trapping behavior following thermal quenching or irradn. at very low temps. are also described.
- 11Bowman, A. C.; Milsmann, C.; Bill, E.; Turner, Z. R.; Lobkovsky, E.; DeBeer, S.; Wieghardt, K.; Chirik, P. J. Synthesis and Electronic Structure Determination of N-Alkyl-Substituted Bis(imino)pyridine Iron Imides Exhibiting Spin Crossover Behavior. J. Am. Chem. Soc. 2011, 133, 17353– 17369, DOI: 10.1021/ja205736m11Synthesis and Electronic Structure Determination of N-Alkyl-Substituted Bis(imino)pyridine Iron Imides Exhibiting Spin Crossover BehaviorBowman, Amanda C.; Milsmann, Carsten; Bill, Eckhard; Turner, Zoe R.; Lobkovsky, Emil; DeBeer, Serena; Wieghardt, Karl; Chirik, Paul J.Journal of the American Chemical Society (2011), 133 (43), 17353-17369CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Three new N-alkyl substituted bis(imino)pyridine iron imide complexes, (iPrPDI)FeNR (iPrPDI = 2,6-(2,6-iPr2-C6H3-N:CMe)2C5H3N; R = 1-adamantyl (1Ad), cyclooctyl (CyOct), and 2-adamantyl (2Ad)) were synthesized by addn. of the appropriate alkyl azide to the iron bis(dinitrogen) complex, (iPrPDI)Fe(N2)2. SQUID magnetic measurements on the isomeric iron imides, (iPrPDI)FeN1Ad and (iPrPDI)FeN2Ad, established spin crossover behavior with the latter example having a more complete spin transition in the exptl. accessible temp. range. X-ray diffraction on all three alkyl-substituted bis(imino)pyridine iron imides established essentially planar compds. with relatively short Fe-Nimide bond lengths and two-electron redn. of the redox-active bis(imino)pyridine chelate. Zero- and applied-field Mossbauer spectroscopic measurements indicate diamagnetic ground states at cryogenic temps. and established low isomer shifts consistent with highly covalent mols. For (iPrPDI)FeN2Ad, Mossbauer spectroscopy also supports spin crossover behavior and allowed extn. of thermodn. parameters for the S = 0 to S = 1 transition. X-ray absorption spectroscopy and computational studies were also performed to explore the electronic structure of the bis(imino)pyridine alkyl-substituted imides. An electronic structure description with a low spin ferric center (S = 1/2) antiferromagnetically coupled to an imidyl radical (Simide = 1/2) and a closed-shell, dianionic bis(imino)pyridine chelate (SPDI = 0) is favored for the S = 0 state. An iron-centered spin transition to an intermediate spin ferric ion (SFe = 3/2) accounts for the S = 1 state obsd. at higher temps. Other possibilities based on the computational and exptl. data are also evaluated and compared to the electronic structure of the bis(imino)pyridine iron N-aryl imide counterparts.
- 12(a) Scepaniak, J. J.; Harris, T. D.; Vogel, C. S.; Sutter, J.; Meyer, K.; Smith, J. M. Spin Crossover in a Four-Coordinate Iron(II) Complex. J. Am. Chem. Soc. 2011, 133, 3824– 3827, DOI: 10.1021/ja200347312aSpin Crossover in a Four-Coordinate Iron(II) ComplexScepaniak, Jeremiah J.; Harris, T. David; Vogel, Carola S.; Sutter, Jorg; Meyer, Karsten; Smith, Jeremy M.Journal of the American Chemical Society (2011), 133 (11), 3824-3827CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The four-coordinate iron(II) phosphoraniminato complex PhB(MesIm)3Fe-N=PPh3 undergoes an S = 0 to S = 2 spin transition with TC = 81 K, as detd. by variable-temp. magnetic measurements and Mossbauer spectroscopy. Variable-temp. single-crystal X-ray diffraction revealed that the S = 0 to S = 2 transition is assocd. with an increase in the Fe-C and Fe-N bond distances and a decrease in the N-P bond distance. These structural changes have been interpreted in terms of electronic structure theory.(b) Mathoniere, C.; Lin, H. J.; Siretanu, D.; Clerac, R.; Smith, J. M. Photoinduced single-molecule magnet properties in a four-coordinate iron(II) spin crossover complex. J. Am. Chem. Soc. 2013, 135, 19083– 19086, DOI: 10.1021/ja410643s12bPhotoinduced Single-Molecule Magnet Properties in a Four-Coordinate Iron(II) Spin Crossover ComplexMathoniere, Corine; Lin, Hsiu-Jung; Siretanu, Diana; Clerac, Rodolphe; Smith, Jeremy M.Journal of the American Chemical Society (2013), 135 (51), 19083-19086CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The four-coordinate Fe-(II) complex, PhB(MesIm)3FeNPPh3 (1) was previously reported to undergo a thermal spin-crossover (SCO) between high-spin (HS, S = 2) and low-spin (LS, S = 0) states. This complex is photoactive <20 K, undergoing a photoinduced LS to HS spin state change, as detd. by optical reflectivity and photomagnetic measurements. With continuous white light irradn., 1 displays slow relaxation of the magnetization, i.e. single-mol. magnet (SMM) properties, at temps. <5 K. This complex provides a structural template for the design of new photoinduced mononuclear SMMs based on the SCO phenomenon.(c) Lin, H. J.; Siretanu, D.; Dickie, D. A.; Subedi, D.; Scepaniak, J. J.; Mitcov, D.; Clerac, R.; Smith, J. M. Steric and electronic control of the spin state in three-fold symmetric, four-coordinate iron(II) complexes. J. Am. Chem. Soc. 2014, 136, 13326– 13332, DOI: 10.1021/ja506425a12cSteric and Electronic Control of the Spin State in Three-Fold Symmetric, Four-Coordinate Iron(II) ComplexesLin, Hsiu-Jung; Siretanu, Diana; Dickie, Diane A.; Subedi, Deepak; Scepaniak, Jeremiah J.; Mitcov, Dmitri; Clerac, Rodolphe; Smith, Jeremy M.Journal of the American Chemical Society (2014), 136 (38), 13326-13332CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The 3-fold sym., four-coordinate Fe(II) phosphoraminimato complexes PhB(MesIm)3Fe-N=PRR'R'' (PRR'R'' = PMePh2, PMe2Ph, PMe3, and PPr3) undergo a thermally induced S = 0 to S = 2 spin-crossover in fluid soln. Smaller phosphoraminimato ligands stabilize the low-spin state, and an excellent correlation is obsd. between the characteristic temp. of the spin-crossover (T1/2) and the Tolman cone angle (θ). Complexes with para-substituted triarylphosphoraminimato ligands (p-XC6H4)3P=N- (X = H, Me and OMe) also undergo spin-crossover in soln. These isosteric phosphoraminimato ligands reveal that the low-spin state is stabilized by more strongly donating ligands. This control over the spin state provides important insights for modulating the magnetic properties of four-coordinate Fe(II) complexes.
- 13Creutz, S. E.; Peters, J. C. Spin-State Tuning at Pseudo-tetrahedral d6 Ions: Spin Crossover in [BP3]FeII–X Complexes. Inorg. Chem. 2016, 55, 3894– 3906, DOI: 10.1021/acs.inorgchem.6b0006613Spin-State Tuning at Pseudo-tetrahedral d6 Ions: Spin Crossover in [BP3]FeII-X ComplexesCreutz, Sidney E.; Peters, Jonas C.Inorganic Chemistry (2016), 55 (8), 3894-3906CODEN: INOCAJ; ISSN:0020-1669. (American Chemical Society)Low-coordinate transition-metal complexes that undergo spin crossover remain rare. We report here a series of four-coordinate, pseudo-tetrahedral P3FeII-X complexes supported by tris(phosphine)borate P3 ([PhBPR3]-) and phosphiniminato X-type ligands (-N=PR'3) that, in combination, tune the spin-crossover behavior of the system. Most of the reported iron complexes undergo spin crossover at temps. near or above room temp. in soln. and in the solid state. The change in spin state coincides with a significant change in the degree of π-bonding between Fe and the bound N atom of the phosphiniminato ligand. Spin crossover is accompanied by striking changes in the UV-visible (UV-vis) absorption spectra, which allows for quant. modeling of the thermodn. parameters of the spin equil. These spin equil. have also been studied by numerous techniques including paramagnetic NMR (NMR), IR, and M.ovrddot.ossbauer spectroscopies; X-ray crystallog.; and solid-state superconducting quantum interference device (SQUID) magnetometry. These studies allow qual. correlations to be made between the steric and electronic properties of the ligand substituents and the enthalpy and entropy changes assocd. with the spin equil.
- 14(a) Travieso-Puente, R.; Broekman, J. O. P.; Chang, M.-C.; Demeshko, S.; Meyer, F.; Otten, E. Spin-Crossover in a Pseudo-tetrahedral Bis(formazanate) Iron Complex. J. Am. Chem. Soc. 2016, 138, 5503– 5506, DOI: 10.1021/jacs.6b0155214aSpin-Crossover in a Pseudo-tetrahedral Bis(formazanate) Iron ComplexTravieso-Puente, Raquel; Broekman, J. O. P.; Chang, Mu-Chieh; Demeshko, Serhiy; Meyer, Franc; Otten, EdwinJournal of the American Chemical Society (2016), 138 (17), 5503-5506CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Spin-crossover in a pseudo-tetrahedral bis(formazanate) iron(II) complex (1) is described. Structural, magnetic, and spectroscopic analyses indicate that this compd. undergoes thermal switching between an S = 0 and an S = 2 state, which is very rare in four-coordinate complexes. The transition to the high-spin state is accompanied by an increase in Fe-N bond lengths and a concomitant contraction of intraligand N-N bonds. The latter suggests that stabilization of the low-spin state is due to the π-acceptor properties of the ligand. One-electron redn. of 1 leads to the formation of the corresponding anion, which contains a low-spin (S = 1/2) Fe(I) center. The findings are rationalized by electronic structure calcns. using d. functional theory.(b) Milocco, F.; de Vries, F.; Bartels, I. M. A.; Havenith, R. W. A.; Cirera, J.; Demeshko, S.; Meyer, F.; Otten, E. Electronic Control of Spin-Crossover Properties in Four-Coordinate Bis(formazanate) Iron(II) Complexes. J. Am. Chem. Soc. 2020, 142, 20170– 20181, DOI: 10.1021/jacs.0c1001014bElectronic Control of Spin-Crossover Properties in Four-Coordinate Bis(formazanate) Iron(II) ComplexesMilocco, Francesca; de Vries, Folkert; Bartels, Imke M. A.; Havenith, Remco W. A.; Cirera, Jordi; Demeshko, Serhiy; Meyer, Franc; Otten, EdwinJournal of the American Chemical Society (2020), 142 (47), 20170-20181CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The transition between spin states in d-block metal complexes has important ramifications for their structure and reactivity, with applications ranging from information storage materials to understanding catalytic activity of metalloenzymes. Tuning the ligand field (ΔO) by steric and/or electronic effects has provided spin-crossover compds. for several transition metals in the periodic table, but this has mostly been limited to coordinatively satd. metal centers in octahedral ligand environments. Spin-crossover complexes with low coordination nos. are much rarer. Here we report a series of four-coordinate, (pseudo)tetrahedral Fe(II) complexes with formazanate ligands and demonstrate how electronic substituent effects can be used to modulate the thermally induced transition between S = 0 and S = 2 spin states in soln. All six compds. undergo spin-crossover in soln. with T1/2 above room temp. (300-368 K). While structural anal. by X-ray crystallog. shows that the majority of these compds. are low-spin in the solid state (and remain unchanged upon heating), we find that packing effects can override this preference and give rise to either rigorously high-spin (6) or gradual spin-crossover behavior (5) also in the solid state. D. functional theory calcns. are used to delineate the empirical trends in soln. spin-crossover thermodn. In all cases, the stabilization of the low-spin state is due to the π-acceptor properties of the formazanate ligand, resulting in an "inverted" ligand field, with an approx. "two-over-three" splitting of the d-orbitals and a high degree of metal-ligand covalency due to metal → ligand π-backdonation. The computational data indicate that the electronic nature of the para-substituent has a different influence depending on whether it is present at the C-Ar or N-Ar rings, which is ascribed to the opposing effect on metal-ligand σ- and π-bonding.
- 15(a) Feltham, H. L. C.; Barltrop, A. S.; Brooker, S. Spin crossover in iron(II) complexes of 3,4,5-tri-substituted-1,2,4-triazole (Rdpt), 3,5-di-substituted-1,2,4-triazolate (dpt – ), and related ligands. Coord. Chem. Rev. 2017, 344, 26– 53, DOI: 10.1016/j.ccr.2016.10.00615aSpin crossover in iron(II) complexes of 3,4,5-tri-substituted-1,2,4-triazole (Rdpt), 3,5-di-substituted-1,2,4-triazolate (dpt-), and related ligandsFeltham, Humphrey L. C.; Barltrop, Alexis S.; Brooker, SallyCoordination Chemistry Reviews (2017), 344 (), 26-53CODEN: CCHRAM; ISSN:0010-8545. (Elsevier B.V.)A review. A general introduction to spin crossover involving iron(II) is provided. Then the 26 examples of 3,4,5-tri-substituted-1,2,4-triazole (mostly Rdpt and Rppt) ligands and the 11 examples of 3,5-di-substituted-1,2,4-triazole (including Hdpt) ligands that are of interest to this review are introduced. The 26 examples of 3,4,5-tri-substituted-1,2,4-triazole ligands fall into main two categories: (a) the 13 examples of sym. 4-substituted 3,5-di(2-pyridyl)triazole (Rdpt) ligands and (b) the 13 ligands closely related to the Rdpt ligands but featuring Ph (known as Rppt ligands), pyrazine or pyrimidine rings in place of one of more of the pyridine rings. In contrast, of the 11 examples of 3,5-di-substituted-1,2,4-triazole (HL, including Hdpt) ligands, only 2 are sym. (3,5-di(2-pyridyl)triazole and 3,5-di(2-pyrazyl)triazole ligands); the other 9 are non-sym. as two different rings are attached to the 3 and 5 positions of the triazole ring. In all but one example, in iron complexes of the 3,5-di-substituted-1,2,4-triazole ligands (HL, including Hdpt) the ligand deprotonates and coordinates as an anionic 3,5-di-substituted-1,2,4-triazolate (L-) ligand. This field was reviewed in 2008, and considered the 17 such complexes published up until 2007. Hence the present review provides a comprehensive update on the substantial progress that has been made in this field since then. Specifically it covers the 92 new iron(II) complexes of these 37 ligands, published since then and before 1 Dec. 2015. Of the 92 of new iron(II) complexes reported, 72 have been structurally characterized, and almost half of them, 40, are SCO-active. A wide range of nuclearities, from mono- to di- to tri- to penta- and poly-nuclear, as well as a range of binding modes, are obsd., so this review is organised into sections accordingly. In each section, after briefly summarizing the findings of the 2008 review, the new developments are detailed. Finally, the findings of this rich avenue of investigation into such ligands are summarized, indicating a bright future ahead.(b) Constable, E. C.; Baum, G.; Bill, E.; Dyson, R.; van Eldik, R.; Fenske, D.; Kaderli, S.; Morris, D.; Neubrand, A.; Neuburger, M.; Smith, D. R.; Wieghardt, K.; Zehnder, M.; Zuberbühler, A. D. Control of Iron(II) Spin States in 2,2′:6′,2″-Terpyridine Complexes through Ligand Substitution. Chem. - Eur. J. 1999, 5, 498– 508, DOI: 10.1002/(SICI)1521-3765(19990201)5:2<498::AID-CHEM498>3.0.CO;2-V15bControl of iron(II) spin states in 2,2':6',2''-terpyridine complexes through ligand substitutionConstable, Edwin C.; Baum, Gerhard; Bill, Eckhard; Dyson, Raylene; Van Eldik, Rudi; Fenske, Dieter; Kaderli, Susan; Morris, Darrell; Neubrand, Anton; Neuburger, Markus; Smith, Diane R.; Wieghardt, Karl; Zehnder, Margareta; Zuberbuhler, Andreas D.Chemistry - A European Journal (1999), 5 (2), 498-508CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH)6- And 6,6''-aryl-substituted 2,2':6',2''-terpyridine ligands and their Fe(II) complexes were prepd. The introduction of Ph substituents at both the 6- and 6'' - positions leads exclusively to the formation of orange high-spin Fe(II) complexes while the presence of a single 6-Ph substituent results in spin-crossover systems. [Fe(2)2]X2 (2 = 4,6-diphenyl-2,2':6',2''-terpyridine; X = ClO4 or PF6) were studied in detail, and the solid-state x-ray structures of both the low- and high-spin forms are reported. Mossbauer spectroscopic and magnetic susceptibility measurements are reported, and the temp. and pressure dependence of the high-spin/low-spin transition were studied. An x-ray structural study of [Fe(2)2](ClO4)2 is also reported; this complex is highly distorted with two very long Fe···N contacts of over 2.4 Å and is best regarded as a four-coordinate Fe complex.
- 16(a) Bacchi, S.; Benaglia, M.; Cozzi, F.; Demartin, F.; Filippini, G.; Gavezzotti, A. X-ray Diffraction and Theoretical Studies for the Quantitative Assessment of Intermolecular Arene–Perfluoroarene Stacking Interactions. Chem. - Eur. J. 2006, 12, 3538– 3546, DOI: 10.1002/chem.20050124817aX-ray diffraction and theoretical studies for the quantitative assessment of intermolecular arene-perfluoroarene stacking interactionsBacchi, Sergio; Benaglia, Maurizio; Cozzi, Franco; Demartin, Francesco; Filippini, Giuseppe; Gavezzotti, AngeloChemistry - A European Journal (2006), 12 (13), 3538-3546CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH & Co. KGaA)The arene-perfluoroarene stacking interaction was studied by exptl. and theor. methods. Compds. with different possibilities for formation of this recognition motif in the solid state were synthesized, and their crystal structures detd. by single-crystal x-ray diffraction. The crystal packing of these compds., as well as the packing of related compds. retrieved from crystallog. databases, were analyzed with quant. crystal potentials: total lattice energies and the cohesive energies of closest mol. pairs in the crystals were calcd. The arene-perfluoroarene recognition motif emerges as a dominant interaction in the nonhydrogen-bonding compds. studied here, to the point that asym. dimers formed over the stacking motif carry over to asym. units made of two mols. in the crystal both for pure compds. and for mol. complexes; however, inter-ring distances and angles range from 3.70 to 4.85 Å and from 5 to 21°, resp. Pixel energy partitioning reveals that whenever arom. rings stack, the largest cohesive energy contribution comes from dispersion, which roughly amts. to 20 kJ mol-1 per Ph ring, while the coulombic term is minor but significant enough to make a difference between the arene-arene or perfluoroarene-perfluoroarene interactions on the one hand, and arene-perfluoroarene interactions on the other, whereby the latter are favored by ∼10 kJ mol-1 per Ph ring. No evidence of special interaction which can be attributed to H···F confrontation was recognizable.(b) Salonen, L. M.; Ellermann, M.; Diederich, F. Aromatic rings in chemical and biological recognition: energetics and structures. Angew. Chem., Int. Ed. 2011, 50, 4808– 4842, DOI: 10.1002/anie.20100756017bAromatic rings in chemical and biological recognition: energetics and structuresSalonen, Laura M.; Ellermann, Manuel; Diederich, FrancoisAngewandte Chemie, International Edition (2011), 50 (21), 4808-4842CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. This review describes a multidimensional treatment of mol. recognition phenomena involving arom. rings in chem. and biol. systems. It summarizes new results reported since the appearance of an earlier review in 2003 in host-guest chem., biol. affinity assays and biostructural anal., data base mining in the Cambridge Structural Database (CSD) and the Protein Data Bank (PDB), and advanced computational studies. Topics addressed are arene-arene, perfluoroarene-arene, S···arom., cation-π, and anion-π interactions, as well as hydrogen bonding to π systems. The generated knowledge benefits, in particular, structure-based hit-to-lead development and lead optimization both in the pharmaceutical and in the crop protection industry. It equally facilitates the development of new advanced materials and supramol. systems, and should inspire further utilization of interactions with arom. rings to control the stereochem. outcome of synthetic transformations.(c) Wheeler, S. E. Local Nature of Substituent Effects in Stacking Interactions. J. Am. Chem. Soc. 2011, 133, 10262– 10274, DOI: 10.1021/ja202932e17cLocal Nature of Substituent Effects in Stacking InteractionsWheeler, Steven E.Journal of the American Chemical Society (2011), 133 (26), 10262-10274CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Popular explanations of substituent effects in π-stacking interactions hinge upon substituent-induced changes in the aryl π-system. This entrenched view has been used to explain substituent effects in countless stacking interactions over the past 2 decades. However, for a broad range of stacked dimers, it is shown that substituent effects are better described as arising from local, direct interactions of the substituent with the proximal vertex of the other ring. Consequently, substituent effects in stacking interactions are additive, regardless of whether the substituents are on the same or opposite rings. Substituent effects are also insensitive to the introduction of heteroatoms on distant parts of either stacked ring. This local, direct interaction viewpoint provides clear, unambiguous explanations of substituent effects for myriad stacking interactions that are in accord with robust computational data, including DFT-D and new benchmark CCSD(T) results. Many of these computational results cannot be readily explained using traditional π-polarization-based models. Analyses of stacking interactions based solely on the sign of the electrostatic potential above the face of an arom. ring or the mol. quadrupole moment face a similar fate. The local, direct interaction model provides a simple means of analyzing substituent effects in complex arom. systems and also offers simple explanations of the crystal packing of fluorinated benzenes and the recently published dependence of the stability of protein-RNA complexes on the regiochem. of fluorinated base analogs [J. Am. Chem. Soc.2011, 133, 3687-3689].
- 17Gütlich, P. Fifty Years of Mössbauer Spectroscopy in Solid State Research - Remarkable Achievements, Future Perspectives. Z. Anorg. Allg. Chem. 2012, 638, 15– 43, DOI: 10.1002/zaac.20110041618Fifty Years of Moessbauer Spectroscopy in Solid State Research - Remarkable Achievements, Future PerspectivesGuetlich, PhilippZeitschrift fuer Anorganische und Allgemeine Chemie (2012), 638 (1), 15-43CODEN: ZAACAB; ISSN:0044-2313. (Wiley-VCH Verlag GmbH & Co. KGaA)Moessbauer spectroscopy was founded more than fifty years ago based on an outstanding discovery by the young German physicist Rudolf Ludwig Moessbauer while working on his Ph.D. thesis. He discovered the recoilless nuclear resonance fluorescence of gamma radiation and was awarded the Nobel Prize in Physics in 1961 as one of the youngest recipients of this most prestigious award. His discovery led to the development of a new technique for measurements of hyperfine interactions between nuclear moments and electromagnetic fields. This method, with highest sharpness of tuning of 10-13, yields information on valence state, symmetry, magnetic behavior, phase transition, lattice dynamics and other solid state properties.
- 18Milocco, F.; Demeshko, S.; Meyer, F.; Otten, E. Ferrate(II) complexes with redox-active formazanate ligands. Dalton Trans. 2018, 47, 8817– 8823, DOI: 10.1039/C8DT01597J19Ferrate(II) complexes with redox-active formazanate ligandsMilocco, Francesca; Demeshko, Serhiy; Meyer, Franc; Otten, EdwinDalton Transactions (2018), 47 (26), 8817-8823CODEN: DTARAF; ISSN:1477-9226. (Royal Society of Chemistry)The synthesis of mono(formazanate) Fe complexes is described. In the presence of Bu4N halides, salt metathesis reactions afford the ferrate(II) complexes [Bu4N][LFeX2] (L = PhNNC(p-tol)NNPh; X = Cl, Br) in good yield, and the products were characterized. The high-spin ferrate(II) complexes show cyclic voltammograms that are consistent with reversible, ligand-based 1-electron redn. The halides in these ferrate(II) compds. are labile, and are displaced by 4-methoxyphenyl isocyanide (4 equiv) as evidenced by formation of the low-spin, cationic octahedral complex [LFe(CNC6H4(p-OMe))4][Br]. Thus, a straightforward route to mono(formazanate) Fe(II) complexes was established.
- 19Lepori, C.; Guillot, R.; Hannedouche, J. C1-symmetric β-Diketiminatoiron(II) Complexes for Hydroamination of Primary Alkenylamines. Adv. Synth. Catal. 2019, 361, 714– 719, DOI: 10.1002/adsc.20180146420C1-symmetric β-Diketiminatoiron(II) Complexes for Hydroamination of Primary AlkenylaminesLepori, Clement; Guillot, Regis; Hannedouche, JeromeAdvanced Synthesis & Catalysis (2019), 361 (4), 714-719CODEN: ASCAF7; ISSN:1615-4150. (Wiley-VCH Verlag GmbH & Co. KGaA)The synthesis and solid-state characterization of an array of well-defined low-coordinate C1-sym. β-diketiminatoiron(II) alkyl complexes featuring steric and electronic variations on one of the N-aryl substituents of the β-diketiminate ligand scaffold are reported. All complexes display unique catalytic abilities of promoting the selective exo-cyclohydroamination of unprotected 2,2-diphenylpent-4-en-1-amine under mild reactions conditions. The incorporation of a potentially coordinative ortho-methoxy substituent on one of the N-aryl rings of the β-diketiminate skeleton, in conjunction with a more crowded 2,6-diisopropylphenyl group on the other, affords a far more active catalyst C1-sym. β-diketiminatoiron(II) alkyl complex than the authors' previously reported C2-sym. β-diketiminatoiron(II) alkyl complex (Ar1 = Ar2 = 2,4,6-(Me)3C6H2)/cyclopentylamine system. Comparative studies let the authors postulate that this superior activity of C1-sym. β-diketiminatoiron(II) alkyl complex , when compared with other complexes , is likely arising from steric effects and/or the coordinating ability of the ortho-methoxy substituent. The scope and limitations of this novel C1-sym. β-diketiminatoiron(II) alkyl complex are also presented.
- 20Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297– 3305, DOI: 10.1039/b508541a21Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracyWeigend, Florian; Ahlrichs, ReinhartPhysical Chemistry Chemical Physics (2005), 7 (18), 3297-3305CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)Gaussian basis sets of quadruple zeta valence quality for Rb-Rn are presented, as well as bases of split valence and triple zeta valence quality for H-Rn. The latter were obtained by (partly) modifying bases developed previously. A large set of more than 300 mols. representing (nearly) all elements-except lanthanides-in their common oxidn. states was used to assess the quality of the bases all across the periodic table. Quantities investigated were atomization energies, dipole moments and structure parameters for Hartree-Fock, d. functional theory and correlated methods, for which we had chosen Moller-Plesset perturbation theory as an example. Finally recommendations are given which type of basis set is used best for a certain level of theory and a desired quality of results.
- 21Becke, A. D. Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A 1988, 38, 3098– 3100, DOI: 10.1103/PhysRevA.38.309822Density-functional exchange-energy approximation with correct asymptotic behaviorBecke, A. D.Physical Review A: Atomic, Molecular, and Optical Physics (1988), 38 (6), 3098-100CODEN: PLRAAN; ISSN:0556-2791.Current gradient-cor. d.-functional approxns. for the exchange energies of at. and mol. systems fail to reproduce the correct 1/r asymptotic behavior of the exchange-energy d. A gradient-cor. exchange-energy functional is given with the proper asymptotic limit. This functional, contg. only one parameter, fits the exact Hartree-Fock exchange energies of a wide variety of at. systems with remarkable accuracy, surpassing the performance of previous functionals contg. two parameters or more.
- 22(a) Staroverov, V. N.; Scuseria, G. E.; Tao, J.; Perdew, J. P. Comparative assessment of a new nonempirical density functional: Molecules and hydrogen-bonded complexes. J. Chem. Phys. 2003, 119, 12129– 12137, DOI: 10.1063/1.162654323aComparative assessment of a new nonempirical density functional: Molecules and hydrogen-bonded complexesStaroverov, Viktor N.; Scuseria, Gustavo E.; Tao, Jianmin; Perdew, John P.Journal of Chemical Physics (2003), 119 (23), 12129-12137CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)A comprehensive study is undertaken to assess the nonempirical meta-generalized gradient approxn. (MGGA) of Tao, Perdew, Staroverov, and Scuseria (TPSS) against 14 common exchange-correlation energy functionals. Principal results are presented in the form of statistical summaries of deviations from expt. for the G3/99 test set (223 enthalpies of formation, 86 ionization potentials, 58 electron affinities, 8 proton affinities) and three addnl. test sets involving 96 bond lengths, 82 harmonic vibrational frequencies, and 10 hydrogen-bonded complexes, all computed using the 6-311++G(3df,3pd) basis. The TPSS functional matches, or exceeds in accuracy all prior nonempirical constructions and, unlike semiempirical functionals, consistently provides a high-quality description of diverse systems and properties. The computational cost of self-consistent MGGA is comparable to that of ordinary GGA, and exact exchange (unavailable in some codes) is not required. A one-parameter global hybrid version of the TPSS functional is introduced and shown to give further improvement for most properties.(b) Tao, J.; Perdew, J. P.; Staroverov, V. N.; Scuseria, G. E. Climbing the Density Functional Ladder: Nonempirical Meta--Generalized Gradient Approximation Designed for Molecules and Solids. Phys. Rev. Lett. 2003, 91, 146401, DOI: 10.1103/PhysRevLett.91.14640123bClimbing the Density Functional Ladder: Nonempirical Meta-Generalized Gradient Approximation Designed for Molecules and SolidsTao, Jianmin; Perdew, John P.; Staroverov, Viktor N.; Scuseria, Gustavo E.Physical Review Letters (2003), 91 (14), 146401/1-146401/4CODEN: PRLTAO; ISSN:0031-9007. (American Physical Society)The electron d., its gradient, and the Kohn-Sham orbital kinetic energy d. are the local ingredients of a meta-generalized gradient approxn. (meta-GGA). We construct a meta-GGA d. functional for the exchange-correlation energy that satisfies exact constraints without empirical parameters. The exchange and correlation terms respect two paradigms: one- or two-electron densities and slowly varying densities, and so describe both mols. and solids with high accuracy, as shown by extensive numerical tests. This functional completes the third rung of "Jacob's ladder" of approxns., above the local spin d. and GGA rungs.
- 23Becke, A. D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648– 5652, DOI: 10.1063/1.46491324Density-functional thermochemistry. III. The role of exact exchangeBecke, Axel D.Journal of Chemical Physics (1993), 98 (7), 5648-52CODEN: JCPSA6; ISSN:0021-9606.Despite the remarkable thermochem. accuracy of Kohn-Sham d.-functional theories with gradient corrections for exchange-correlation, the author believes that further improvements are unlikely unless exact-exchange information is considered. Arguments to support this view are presented, and a semiempirical exchange-correlation functional (contg. local-spin-d., gradient, and exact-exchange terms) is tested for 56 atomization energies, 42 ionization potentials, 8 proton affinities, and 10 total at. energies of first- and second-row systems. This functional performs better than previous functionals with gradient corrections only, and fits expt. atomization energies with an impressively small av. abs. deviation of 2.4 kcal/mol.
- 24(a) Harvey, J. N. In Principles and Applications of Density Functional Theory in Inorganic Chemistry I; Springer: Berlin, 2004; pp 151– 184.There is no corresponding record for this reference.(b) Swart, M.; Gruden, M. Spinning around in Transition-Metal Chemistry. Acc. Chem. Res. 2016, 49, 2690– 2697, DOI: 10.1021/acs.accounts.6b0027125bSpinning around in Transition-Metal ChemistrySwart, Marcel; Gruden, MajaAccounts of Chemical Research (2016), 49 (12), 2690-2697CODEN: ACHRE4; ISSN:0001-4842. (American Chemical Society)A review. The great diversity and richness of transition metal chem., such as the features of an open d-shell, opened a way to numerous areas of scientific research and technol. applications. Depending on the nature of the metal and its environment, there are often several energetically accessible spin states, and the progress in accurate theor. treatment of this complicated phenomenon is presented in this Account. The spin state energetics of a transition metal complex can be predicted theor. on the basis of d. functional theory (DFT) or wave function based methodol., where DFT has advantages since it can be applied routinely to medium-to-large-sized mols. and spin-state consistent d. functionals are now available. Addnl. factors such as the effect of the basis set, thermochem. contributions, solvation, relativity, and dispersion, have been investigated by many researchers, but challenges in unambiguous assignment of spin states still remain. The first DFT studies showed intrinsic spin-state preferences of hybrid functionals for high spin and early generalized gradient approxn. functionals for low spin. Progress in the development of d. functional approxns. (DFAs) then led to a class of specially designed DFAs, such as OPBE, SSB-D, and S12g, and brought a very intriguing and fascinating observation that the spin states of transition metals and the SN2 barriers of org. mols. are somehow intimately linked. Among the many noteworthy results that emerged from the search for the appropriate description of the complicated spin state preferences in transition metals, we mainly focused on the examn. of the connection between the spin state and the structures or coordination modes of the transition metal complexes. Changes in spin states normally lead only to changes in the metal-ligand bond lengths, but to the best of our knowledge, the dapsox ligand showed the first example of a transition-metal complex where a change in spin state leads also to changes in the coordination, switching between pentagonal-bipyramidal and capped-octahedron. Moreover, we have summarized the results of the thorough study that cor. the exptl. assignment of the nature of the recently synthesized Sc3+ adduct of [FeIV(O)(TMC)]2+ (TMC = 1,4,8,11-tetramethylcyclam) and firmly established that the Sc3+-capped iron-oxygen complex corresponds to high-spin FeIII. Last, but not least, we have provided deeper insight and rationalization of the observation that unlike in metalloenzymes, where the FeIV-oxo is usually obsd. with high spin, biomimetic FeIV-oxo complexes typically have a intermediate spin state. Energy decompn. analyses on the trigonal-bypiramidal (TBP) and octahedral model systems with ammonia ligands have revealed that the interaction energy of the prepd. metal ion in the intermediate spin state is much smaller for the TBP structure. This sheds light on the origin of the intermediate spin state of the biomimetic TBP FeIV-oxo complexes.
- 25These orbital splitting energies are relatively small; spin-crossover complexes commonly have significantly larger values for Δ. See for example:Hauser, A. Ligand Field Theoretical Considerations. In Spin Crossover in Transition Metal Compounds I; Gütlich, P., Goodwin, H. A., Eds.; Springer Berlin Heidelberg: Berlin, Heidelberg, 2004, p 49.There is no corresponding record for this reference.
- 26(a) Knizia, G. Intrinsic Atomic Orbitals: An Unbiased Bridge between Quantum Theory and Chemical Concepts. J. Chem. Theory Comput. 2013, 9, 4834– 4843, DOI: 10.1021/ct400687b27aIntrinsic Atomic Orbitals: An Unbiased Bridge between Quantum Theory and Chemical ConceptsKnizia, GeraldJournal of Chemical Theory and Computation (2013), 9 (11), 4834-4843CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Modern quantum chem. can make quant. predictions on an immense array of chem. systems. However, the interpretation of those predictions is often complicated by the complex wave function expansions used. Here we show that an exceptionally simple algebraic construction allows for defining at. core and valence orbitals, polarized by the mol. environment, which can exactly represent SCF wave functions. This construction provides an unbiased and direct connection between quantum chem. and empirical chem. concepts, and can be used, for example, to calc. the nature of bonding in mols., in chem. terms, from first principles. In particular, we find consistency with electronegativities (χ), C 1s core-level shifts, resonance substituent parameters (σR), Lewis structures, and oxidn. states of transition-metal complexes.(b) Knizia, G.; Klein, J. E. M. N. Electron Flow in Reaction Mechanisms—Revealed from First Principles. Angew. Chem., Int. Ed. 2015, 54, 5518– 5522, DOI: 10.1002/anie.20141063727bElectron Flow in Reaction Mechanisms-Revealed from First PrinciplesKnizia, Gerald; Klein, Johannes E. M. N.Angewandte Chemie, International Edition (2015), 54 (18), 5518-5522CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)The "curly arrow" of Robinson and Ingold is the primary tool for describing and rationalizing reaction mechanisms. Despite this approach's ubiquity and stellar success, its phys. basis has never been clarified and a direct connection to quantum chem. has never been found. Here we report that the bond rearrangements expressed by curly arrows can be directly obsd. in ab initio computations, as transformations of intrinsic bond orbitals (IBOs) along the reaction coordinate. Our results clarify that curly arrows are rooted in phys. reality-a notion which has been challenged before-and show how quantum chem. can directly establish reaction mechanisms in intuitive terms and unprecedented detail.
- 27Kamphuis, A. J.; Milocco, F.; Koiter, L.; Pescarmona, P. P.; Otten, E. Highly Selective Single-Component Formazanate Ferrate(II) Catalysts for the Conversion of CO2 into Cyclic Carbonates. ChemSusChem 2019, 12, 3635– 3641, DOI: 10.1002/cssc.20190074028Highly Selective Single-Component Formazanate Ferrate(II) Catalysts for the Conversion of CO2 into Cyclic CarbonatesKamphuis, Aeilke J.; Milocco, Francesca; Koiter, Luuk; Pescarmona, Paolo P.; Otten, EdwinChemSusChem (2019), 12 (15), 3635-3641CODEN: CHEMIZ; ISSN:1864-5631. (Wiley-VCH Verlag GmbH & Co. KGaA)The development of new families of active and selective single-component catalysts based on earth-abundant metal is of interest from a sustainable chem. perspective. In this context, anionic mono(formazanate) Fe(II) complexes bearing labile halide ligands, which possess both Lewis acidic and nucleophilic functionalities, were developed as novel single-component homogeneous catalysts for the reaction of CO2 with epoxides to produce cyclic carbonates. The influence of the halide ligand and the electronic properties of the formazanate ligand backbone on the catalytic activity are studied by employing the Fe(II) complexes with and without an addnl. nucleophile. Very high selectivity is achieved towards the formation of the cyclic carbonate products from various terminal and internal epoxides without the need of a cocatalyst. Crystallog. data are given.
- 28Chang, M.-C.; Roewen, P.; Travieso-Puente, R.; Lutz, M.; Otten, E. Formazanate Ligands as Structurally Versatile, Redox-Active Analogues of β-Diketiminates in Zinc Chemistry. Inorg. Chem. 2015, 54, 379– 388, DOI: 10.1021/ic502587329Formazanate Ligands as Structurally Versatile, Redox-Active Analogues of β-Diketiminates in Zinc ChemistryChang, Mu-Chieh; Roewen, Peter; Travieso-Puente, Raquel; Lutz, Martin; Otten, EdwinInorganic Chemistry (2015), 54 (1), 379-388CODEN: INOCAJ; ISSN:0020-1669. (American Chemical Society)A range of tetrahedral bis(formazanate)zinc complexes with different steric and electronic properties of the formazanate ligands were synthesized. The solid-state structures for several of these were detd. by x-ray crystallog., which showed that complexes with sym., unhindered ligands prefer coordination to the Zn center via the terminal N atoms of the NNCNN ligand backbone. Steric or electronic modifications can override this preference and give rise to solid-state structures in which the formazanate ligand forms a 5-membered chelate by binding to the metal center via an internal N atom. In soln., these compds. show dynamic equil. that involve both 5- and 6-membered chelates. All compds. are intensely colored, and the effect of the ligand substitution pattern on the UV-visible absorption spectra was evaluated. Their cyclic voltammetry is reported, which shows that all compds. may be electrochem. reduced to radical anionic (L2Zn-) and dianionic (L2Zn2-) forms. While unhindered NAr substituents lie in the plane of the ligand backbone (Ar = Ph), the introduction of sterically demanding substituents (Ar = Mes) favors a perpendicular orientation in which the NMes group is no longer in conjugation with the backbone, resulting in hypsochromic shifts in the absorption spectra. The redox potentials in L2Zn compds. may be altered in a straightforward manner over a relatively wide range (∼700 mV) via the introduction of electron-donating or -withdrawing substituents on the formazanate framework.
- 29Chang, M. C.; Otten, E. Synthesis and ligand-based reduction chemistry of boron difluoride complexes with redox-active formazanate ligands. Chem. Commun. 2014, 50, 7431– 7433, DOI: 10.1039/C4CC03244F30Synthesis and ligand-based reduction chemistry of boron difluoride complexes with redox-active formazanate ligandsChang, M.-C.; Otten, E.Chemical Communications (Cambridge, United Kingdom) (2014), 50 (56), 7431-7433CODEN: CHCOFS; ISSN:1359-7345. (Royal Society of Chemistry)Mono(formazanate) boron difluoride complexes (LBF2), which show remarkably facile and reversible ligand-based redox-chem., were synthesized by transmetalation of bis(formazanate) zinc complexes with boron trifluoride. The one-electron redn. product [LBF2]-[Cp2Co]+ and a key intermediate for the transmetalation reaction, the six-coordinate zinc complex (L(BF3))2Zn were isolated and fully characterized.
- 30(a) Chang, M. C.; Chantzis, A.; Jacquemin, D.; Otten, E. Boron difluorides with formazanate ligands: redox-switchable fluorescent dyes with large stokes shifts. Dalton Trans. 2016, 45, 9477– 9484, DOI: 10.1039/C6DT01226D31aBoron difluorides with formazanate ligands: redox-switchable fluorescent dyes with large stokes shiftsChang, M.-C.; Chantzis, A.; Jacquemin, D.; Otten, E.Dalton Transactions (2016), 45 (23), 9477-9484CODEN: DTARAF; ISSN:1477-9226. (Royal Society of Chemistry)The synthesis of a series of (formazanate)boron difluorides and their 1-electron redn. products is described. The neutral compds. are fluorescent with large Stokes shifts. DFT calcns. suggest that a large structural reorganization accompanies photoexictation and accounts for the large Stokes shift. Redn. of the neutral boron difluorides occurs at the ligand and generates the corresponding radical anions. These complexes are non-fluorescent, allowing switching of the emission by changing the ligand oxidn. state.(b) Chang, M.-C. Formazanate as redox-active, structurally versatile ligand platform: Zinc and boron chemistry. PhD thesis, University of Groningen, 2016.There is no corresponding record for this reference.
- 31Broere, D. L.; Coric, I.; Brosnahan, A.; Holland, P. L. Quantitation of the THF Content in Fe[N(SiMe3)2]2.xTHF. Inorg. Chem. 2017, 56, 3140– 3143, DOI: 10.1021/acs.inorgchem.7b0005632Quantitation of the THF Content in Fe[N(SiMe3)2]2·xTHFBroere, Daniel L. J.; Coric, Ilija; Brosnahan, Anna; Holland, Patrick L.Inorganic Chemistry (2017), 56 (6), 3140-3143CODEN: INOCAJ; ISSN:0020-1669. (American Chemical Society)The absence of residual solvent in metal precursors can be of key importance for the successful prepn. of metal complexes or materials. Herein, the authors describe methods for the quantitation of residual coordinated THF that binds to Fe[N(SiMe3)2]2, a commonly used iron synthon, when prepd. according to common literature procedures. A simple method for quantitation of the amt. of residual coordinated THF using 1H NMR spectroscopy is highlighted. Finally, a detailed synthetic procedure is described for the synthesis of THF-free Fe[N(SiMe3)2]2.
- 32Gilroy, J. B.; Ferguson, M. J.; McDonald, R.; Patrick, B. O.; Hicks, R. G. Formazans as β-diketiminate analogues. Structural characterization of boratatetrazines and their reduction to borataverdazyl radical anions. Chem. Commun. 2007, 126– 128, DOI: 10.1039/B609365E33Formazans as β-diketiminate analogues. Structural characterization of boratatetrazines and their reduction to borataverdazyl radical anionsGilroy, Joe B.; Ferguson, Michael J.; McDonald, Robert; Patrick, Brian O.; Hicks, Robin G.Chemical Communications (Cambridge, United Kingdom) (2007), (2), 126-128CODEN: CHCOFS; ISSN:1359-7345. (Royal Society of Chemistry)Formazans, R'NHN:CRN:NR' (R ≠ R' = p-tolyl, Ph) react with boron triacetate to produce boratatetrazines I, which can be reduced by cobaltocene to yield borataverdazyl radical anions, e.g., II-the first boron contg. verdazyl radicals.
- 33Bruker. APEX3, SAINT and SADABS; Bruker AXS Inc.: Madison, WI, 2016.There is no corresponding record for this reference.
- 34Sheldrick, G. A short history of SHELX. Acta Crystallogr., Sect. A 2008, 64, 112– 122, DOI: 10.1107/S010876730704393035A short history of SHELXSheldrick, George M.Acta Crystallographica, Section A: Foundations of Crystallography (2008), 64 (1), 112-122CODEN: ACACEQ; ISSN:0108-7673. (International Union of Crystallography)An account is given of the development of the SHELX system of computer programs from SHELX-76 to the present day. In addn. to identifying useful innovations that have come into general use through their implementation in SHELX, a crit. anal. is presented of the less-successful features, missed opportunities and desirable improvements for future releases of the software. An attempt is made to understand how a program originally designed for photog. intensity data, punched cards and computers over 10000 times slower than an av. modern personal computer has managed to survive for so long. SHELXL is the most widely used program for small-mol. refinement and SHELXS and SHELXD are often employed for structure soln. despite the availability of objectively superior programs. SHELXL also finds a niche for the refinement of macromols. against high-resoln. or twinned data; SHELXPRO acts as an interface for macromol. applications. SHELXC, SHELXD and SHELXE are proving useful for the exptl. phasing of macromols., esp. because they are fast and robust and so are often employed in pipelines for high-throughput phasing. This paper could serve as a general literature citation when one or more of the open-source SHELX programs (and the Bruker AXS version SHELXTL) are employed in the course of a crystal-structure detn.
- 35Sheldrick, G. Crystal structure refinement with SHELXL. Acta Crystallogr., Sect. C 2015, 71, 3– 8, DOI: 10.1107/S205322961402421836Crystal structure refinement with SHELXLSheldrick, George M.Acta Crystallographica, Section C: Structural Chemistry (2015), 71 (1), 3-8CODEN: ACSCGG; ISSN:2053-2296. (International Union of Crystallography)The improvements in the crystal structure refinement program SHELXL have been closely coupled with the development and increasing importance of the CIF (Crystallog. Information Framework) format for validating and archiving crystal structures. An important simplification is that now only one file in CIF format (for convenience, referred to simply as 'a CIF') contg. embedded reflection data and SHELXL instructions is needed for a complete structure archive; the program SHREDCIF can be used to ext. the and files required for further refinement with SHELXL. Recent developments in SHELXL facilitate refinement against neutron diffraction data, the treatment of H atoms, the detn. of abs. structure, the input of partial structure factors and the refinement of twinned and disordered structures. SHELXL is available free to academics for the Windows, Linux and Mac OS X operating systems, and is particularly suitable for multiple-core processors.
Supporting Information
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.inorgchem.0c03593.
Full experimental and characterization data, computational details (PDF)
CCDC 2036148–2036152 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.