Tris(cyclooctatetraene)triiron

From Wikipedia, the free encyclopedia
Tris(cyclooctatetraene)triiron
Names
Other names
tris(μ253-cyclo-octatetraene)-tri-iron
Identifiers
3D model (JSmol)
  • InChI=1S/3C8H8.3Fe/c3*1-2-4-6-8-7-5-3-1;;;/h3*1-8H;;;/b3*2-1-,3-1-,4-2-,5-3-,6-4-,7-5-,8-6-,8-7-;;;
    Key: NPBKKAOUFKDAOG-RKOWHLECSA-N
  • C\1=C\C=C/C=C\C=C1.C\1=C\C=C/C=C\C=C1.C\1=C\C=C/C=C\C=C1.[Fe]2[Fe][Fe]2
Properties
C24H24Fe3
Molar mass 479.991 g·mol−1
Appearance black rhomboidal crystals
Density 1.87 (from structure)
reacts
Solubility in other solvents insoluble: benzene, toluene, pentane
Structure
Monoclinic
P 21/n
Related compounds
Related compounds
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).

Tris(cyclooctatetraene)triiron or Fe3(COT)3, also referred to as the Lavallo-Grubbs compound (after its discoverers[1]) is an organoiron compound with the formula Fe3(C8H8)3. It is a pyrophoric, black crystalline solid, which is insoluble in common organic solvents.The compound represents a rare example of a hydrocarbon analogue of the well-known Triiron dodecacarbonyl (Fe3(CO)12), originally prepared by Dewar and Jones in the early 20th century.[2]

Preparation[edit]

Lavello and Grubbs discovered the compound unexpectedly when trying to prepare noncarbonyl, low coordinate, Fe(0) complexes of N-heterocyclic carbenes (NHCs). They found that reactions of Fe(COT)2 and the NHC, 1,3-bis(2,4,6-trimethylphenyl)-4,5-dihydroimidazol-2-ylidene (SIMes), produced tetrametallic, Fe(I)-Fe(0) mixed valent NHC-COT complexes. In an attempt to characterize intermediates of the unusual transformation, they employed the more sterically hindered NHC, 1,3-bis(2,6-diisopropylphenyl)-4,5-dihydroimidazol-2-ylidene (SIPr) (with Dipp substituents).[1]

Tris(cyclooctatetraene)triiron preparation from bis(cyclooctatetraene)iron mediated by catalytic amounts of NHC. (Dipp = 2,6-diisopropylphenyl)[1]

In benzene, the Dipp substituted NHC reacts with Fe(COT)2 to produce large black rhomboidal crystals of tris(cyclooctatetraene)triiron over 24 h at room temperature. Notably, the reaction was found to occur with catalytic amounts of NHC (10 mol%) yielding 67% of Fe3(COT)3 after 24 h (turn over number=9.5). The synthesis is optimized when the reaction is conducted at 45 °C, yielding 95% conversion to the tris(cyclooctatetraene)triiron cluster. They also highlighted that heating Fe(COT)2 in benzene without any NHC to 100 °C for 24 h forms trace amounts of Fe3(COT)3, but also large amounts of iron metal. Unsurprisingly, elemental analysis of the cluster affirms a 1:1 Fe:COT ratio.[1]

The formation of Fe3(COT)3 from Fe(COT)2 has been calculated to be slightly exothermic(15 kcal/mol).[3]

Other NHCs lead to other unique mixed NHC-COT low valent iron complexes. Lavallo and Grubbs rationalize the transformation by emphasizing the capacity of NHCs to catalytically induce the formation of metal-metal bonds,[4] where the steric hinderance of the NHC is essential, in particular, for the lability of the NHC (in coordination and dissociation) in the cycle. The bulky NHC is proposed to prevent reduction of COT by a bimetallic [(L)Fe2(COT)2] intermediate, where steric constraints block the bonding hapticity required to ligate a reduced form of COT. Another possibility put forward is that reduction of COT occurs only following coordination by a second carbene in the case of SIMes during the catalytic cycle. The sterically hindered NHC prevents such a transformation from occurring.[5]

Proposed intermediates by Lavallo and Grubbs in the formation of Fe3(C8H8)3 and alternative product with changes in NHC-substituents.[1][5]

Electronic structure and bonding[edit]

Example of intrinsic bonding orbitals from optimized singlet geometry of Fe3(COT)3 ; BP86/DZP/TZP(Fe).[3][6][7][8][9]

The discoverers were reluctant to assert an oxidation state of the iron centers in the compound, instead deferring the details of the electronic structure to computational studies.[1]

The crystal structure reveals that the three iron centers arrange in an equilateral triangle (nearly ideal; Fe1 = 59.67°, Fe2 = 60.15°, and Fe3 = 60.18°) The corresponding bond lengths are similar to one another, (Fe1–Fe2 = 2.829 Å, Fe1–Fe3 = 2.815 Å, and Fe2–Fe3 = 2.830 Å), and reflective of Fe-Fe single bonds.[10] As a trinuclear cluster, it would be thought to have a stable 48-electron closed-shell configuration (24 electrons from the three iron atoms and 24 electrons from the three COT rings).[11][12][13]

In the original depiction, each COT ligand acts as an η3 and η5 donor, and thus, some degree of π-allylic and pentadienyl bonding modes can be inferred – though the degree of metal-to-ligand electron transfer is uncertain.[14][15] Computational models suggest the binding mode to lie between η3 and η5, as small shifts in geometry make each mode effectively equivalent (see section on fluxional behavior). Furthermore, DFT calculations with the BLYP functional using a TZP basis set for iron and DZP for carbon and hydrogen estimate a Hirshfeld charge of 0.08 on the iron centers (and Voronoi deformation density of 0.00). Interestingly, all of the bond orders of the C-C ring lie between 1.26 and 1.33, sharply contrasting the discrete single and double bonds of free cyclooctatetraene, or COT complexes with non-bound olefins.[13][16]

Doubly reduced COT (dianion)[17] is known to adopt a planar (aromatic) comformation to metal centers, which is not observed in Fe3(COT)3. However, arguments also exist that such comformations are more related to binding efficiency than aromaticity.[18]

Comparison to Fe-Fe bonding theories of Fe3(CO)12 as in Shaefer's depictions.[19]

In computational studies (BP86), when Fe3(C8H8)3 is optimized as a singlet (gas phase), the iron centers are arranged in an ideal equilateral triangle, as experimentally observed in the crystal structure. In such an electronic configuration, each iron atom achieves an 18-electron configuration through pseudo η5 and η3 coordination to alternating COT ligands. However, if the compound is optimized as a triplet structure, the iron centers instead are a scalene triangle, featuring significant Jahn-Teller distortions.[3]

Additional NBO analysis of the singlet structure reveals Wiberg Bond Indices of 0.22 for the Fe-Fe bonds, closely reminiscent of that of D3h Fe3(CO)12 (0.18).[3][19]

Fluxional Behavior[edit]

Fluxional Ring rocking in Fe3(COT)3 depicted by DeKock et al.[13]

The V comformation of COT has an angle at 135° and is thought to be highly stabilized via bonding with the iron atoms (in free COT, this comformation is disfavored by approximately 36 kcal/mol.[13]

Fascinatingly, in benzene solution, 1H NMR reveals a single broadened resonance with a chemical shift at -3.15 ppm. This suggests that the cyclooctatetraene ligands are highly fluxional and some degree of paramagnetism.[1]

Fluxional Ring rotation in Fe3(COT)3 depicted by DeKock et al.[13]

COT is known to be a highly fluxional ligand in other compounds too, such compounds being deemed “ring-whizzers” (like the related (cyclooctatetraene)iron tricarbonyl).[20][21]

The conformational fluxionality is supported by computational studies which show very low barriers to COT rotation (on the scale of 1.4 kcal/mol) and rocking (0.1 kcal/mol). The transformation from the C3h singlet comformation to the triplet C2v comformation have been calculated to be nearly isoenergetic, driving the possibility of the singlet state existing in equilibrium with the triplet state- an explanation for the observation of paramagnetic NMR resonances at ambient temperatures.[13]

Reactivity and Applications[edit]

The compound is highly reactive and pyrophoric, self-igniting in air.[1] The lack of reactivity studies may be in part sourced from its very low solubility in organic solvents. However, it may be able to find use as a specialized source of reactive low-valent iron.

Computational studies have estimated COT dissociation from Fe3(COT)3 to be 57 kcal/mol uphill, which would not be readily accessible at room temperature in solution.[13]

With analogy to Fe3(CO)12,[22] the compound could potentially be susceptible to Fe-Fe bond homolysis via photoexcitation.

Analogies to Related Compounds[edit]

[23]

At the time of discovery, it was the only homoleptic trimetallic non-carbonyl cluster featuring hydrocarbon ligands. An isolobal analogy[23] can be made with the related Fe3(CO)10 (μ-CO)2[24] cluster first prepared by Jones and Dewar.[2][19][1]

Alternatively, the compound could be compared to ferrocene. In this interpretation, each iron fragment bears an η3 allyl and η5 pentadienyl ligand as depicted. Under Green and Parkin's covalent bond classification method, this would yield LX and L2X respectively. Then additionally, the iron centers donate a pair of electrons into an adjacent empty iron orbital as an adduct (donor-acceptor pairs L and Z type), to overall yield ML3X4 (18 electrons).[25] In fact, the bond orders in ferrocene and Fe3(COT)3 are reported to be very similar.[13]

In fact, Fe3(CO)12 and the group 5 analogues, Ru3(CO)12 and Os3(CO)12 are also known to feature highly fluxional CO ligands in solution (and even in the solid state).[26][27][28]

Hypoelectronic derivatives (M=Ti, Cr, V, Mn) of Fe3(C8H8)3 would be predicted to have metal multiple bonds. Some somewhat related compounds, notably, Ti2(C8H8)3[29][30][31] and Cr2(C8H8)3[29][32] have been experimentally isolated.

References[edit]

  1. ^ a b c d e f g h i Lavallo, Vincent; Grubbs, Robert H. (2009-10-23). "Carbenes As Catalysts for Transformations of Organometallic Iron Complexes". Science. 326 (5952): 559–562. Bibcode:2009Sci...326..559L. doi:10.1126/science.1178919. ISSN 0036-8075. PMC 2841742. PMID 19900894.
  2. ^ a b Dewar, J.; Jones, H. O. Proc. R. Soc. London, Ser. A 1907, 79, 66-80.
  3. ^ a b c d Wang, Hongyan; Sun, Zhonghua; Xie, Yaoming; King, R. Bruce; Schaefer, Henry F. (2011-09-06). "Analogues of the Lavallo–Grubbs Compound Fe3(C8H8)3: Equilateral, Isosceles, and Scalene Metal Triangles in Trinuclear Cyclooctatetraene Complexes M3(C8H8)3 of the First Row Transition Metals (M = Ti, V, Cr, Mn, Fe, Co, and Ni)". Inorganic Chemistry. 50 (19): 9256–9265. doi:10.1021/ic200337w. ISSN 0020-1669. PMID 21894922.
  4. ^ Albrecht, Martin (2009-10-23). "Carbenes in Action". Science. 326 (5952): 532–533. Bibcode:2009Sci...326..532A. doi:10.1126/science.1181553. hdl:10197/6755. ISSN 0036-8075. PMID 19900887.
  5. ^ a b Lavallo, Vincent; El-Batta, Amer; Bertrand, Guy; Grubbs, Robert H. (2010-11-25). "Insights Into the Carbene-Initiated Aggregation of [Fe(cot)2]". Angewandte Chemie. 123 (1): 282–285. doi:10.1002/ange.201005212. ISSN 0044-8249.
  6. ^ Knizia, Gerald (2013-10-17). "Intrinsic Atomic Orbitals: An Unbiased Bridge between Quantum Theory and Chemical Concepts". Journal of Chemical Theory and Computation. 9 (11): 4834–4843. arXiv:1306.6884. doi:10.1021/ct400687b. ISSN 1549-9618. PMID 26583402.
  7. ^ Grimme, Stefan (2006-09-05). "Semiempirical GGA-type density functional constructed with a long-range dispersion correction". Journal of Computational Chemistry. 27 (15): 1787–1799. doi:10.1002/jcc.20495. ISSN 0192-8651. PMID 16955487.
  8. ^ Becke, A. D. (1988-09-01). "Density-functional exchange-energy approximation with correct asymptotic behavior". Physical Review A. 38 (6): 3098–3100. Bibcode:1988PhRvA..38.3098B. doi:10.1103/physreva.38.3098. ISSN 0556-2791. PMID 9900728.
  9. ^ Lee, Chengteh; Yang, Weitao; Parr, Robert G. (1988-01-15). "Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density". Physical Review B. 37 (2): 785–789. Bibcode:1988PhRvB..37..785L. doi:10.1103/physrevb.37.785. ISSN 0163-1829. PMID 9944570.
  10. ^ Duncan Lyngdoh, Richard H.; Schaefer, Henry F.; King, R. Bruce (2018-12-26). "Metal–Metal (MM) Bond Distances and Bond Orders in Binuclear Metal Complexes of the First Row Transition Metals Titanium Through Zinc". Chemical Reviews. 118 (24): 11626–11706. doi:10.1021/acs.chemrev.8b00297. ISSN 0009-2665. PMID 30543419.
  11. ^ Mingos, D.M.P (1990). Introduction to Cluster Chemistry. Prentice Hall.
  12. ^ Dyson, P. J.; McIndoe, J. S. Transition Metal Carbonyl Cluster Chemistry; Taylor & Francis: Oxford, U.K., 2000.
  13. ^ a b c d e f g h Martinie, Ryan J.; Burkhart, Brandon J.; DeKock, Roger L. (2011-09-07). "Computational Investigations of Ligand Fluxionality in Fe3(C8H8)3". Organometallics. 30 (19): 5196–5201. doi:10.1021/om200561x. ISSN 0276-7333.
  14. ^ Mukhopadhyay, Tufan K.; Flores, Marco; Feller, Russell K.; Scott, Brian L.; Taylor, R. Dean; Paz-Pasternak, Moshe; Henson, Neil J.; Rein, Francisca N.; Smythe, Nathan C.; Trovitch, Ryan J.; Gordon, John C. (2014-12-22). "A New Spin on Cyclooctatetraene (COT) Redox Activity: Low-Spin Iron(I) Complexes That Exhibit Antiferromagnetic Coupling to a Singly Reduced η 4 -COT Ligand". Organometallics. 33 (24): 7101–7112. doi:10.1021/om500909h. hdl:2286/R.I.28117. ISSN 0276-7333.
  15. ^ Tereniak, Stephen J.; Lu, Connie C. (2015-03-23). "Group 8 Metal–Metal Bonds". Molecular Metal-Metal Bonds. pp. 225–278. doi:10.1002/9783527673353.ch8. ISBN 978-3-527-33541-1.
  16. ^ Allegra, G.; Colombo, A.; Immirzi, A.; Bassi, I. W. (1968). "The crystal structure of bis(cyclooctatetraene)iron". Journal of the American Chemical Society. 90 (16): 4455–4456. doi:10.1021/ja01018a046. ISSN 0002-7863.
  17. ^ Sokolov, Alexander Yu.; Magers, D. Brandon; Wu, Judy I.; Allen, Wesley D.; Schleyer, Paul v. R.; Schaefer, Henry F. (2013-10-08). "Free Cyclooctatetraene Dianion: Planarity, Aromaticity, and Theoretical Challenges". Journal of Chemical Theory and Computation. 9 (10): 4436–4443. doi:10.1021/ct400642y. ISSN 1549-9618. PMID 26589161.
  18. ^ Dominikowska, Justyna; Palusiak, Marcin (2010). "Cyclooctatetraene in metal complexes—planar does not mean aromatic". New Journal of Chemistry. 34 (9): 1855. doi:10.1039/c0nj00060d. ISSN 1144-0546.
  19. ^ a b c Wang, Hongyan; Xie, Yaoming; King, R. Bruce; Schaefer, Henry F. (2006-08-12). "Remarkable Aspects of Unsaturation in Trinuclear Metal Carbonyl Clusters: The Triiron Species Fe3(CO)n (n = 12, 11, 10, 9)". Journal of the American Chemical Society. 128 (35): 11376–11384. doi:10.1021/ja055223+. ISSN 0002-7863. PMID 16939260.
  20. ^ Grubbs, R.; Breslow, Ronald.; Herber, R.; Lippard, Stephen J. (1967). "Iron tricarbonyl cyclooctatetraene complexes". Journal of the American Chemical Society. 89 (26): 6864–6870. doi:10.1021/ja01002a010. ISSN 0002-7863.
  21. ^ Cotton, F. Albert; Hunter, Douglas L. (1976). "Carbon-13 nuclear magnetic resonance study of the fluxional behavior of cyclooctatetraenetricarbonyliron and -ruthenium". Journal of the American Chemical Society. 98 (6): 1413–1417. doi:10.1021/ja00422a022. ISSN 0002-7863.
  22. ^ Lomont, Justin P.; Harris, Charles B. (2015). "Primary photochemical dynamics of metal carbonyl dimers and clusters in solution: Insights into the results of metal–metal bond cleavage from ultrafast spectroscopic studies". Inorganica Chimica Acta. 424: 38–50. doi:10.1016/j.ica.2014.07.064. ISSN 0020-1693.
  23. ^ a b Hoffmann, Roald (1982). "Building Bridges Between Inorganic and Organic Chemistry (Nobel Lecture)". Angewandte Chemie International Edition in English. 21 (10): 711–724. doi:10.1002/anie.198207113. ISSN 0570-0833.
  24. ^ Wei, Chin Hsuan; Dahl, Lawrence F. (1966). "Molecular Structures of Triiron Dodecacarbonyl and Tetracobalt Dodecacarbonyl". Journal of the American Chemical Society. 88 (8): 1821–1822. doi:10.1021/ja00960a046. ISSN 0002-7863.
  25. ^ Green, Malcolm L. H.; Parkin, Gerard (2014-06-10). "Application of the Covalent Bond Classification Method for the Teaching of Inorganic Chemistry". Journal of Chemical Education. 91 (6): 807–816. Bibcode:2014JChEd..91..807G. doi:10.1021/ed400504f. ISSN 0021-9584.
  26. ^ Wei, Chin Hsuan; Dahl, Lawrence F. (1969). "Dodecacarbonyltriiron: analysis of its stereochemistry". Journal of the American Chemical Society. 91 (6): 1351–1361. doi:10.1021/ja01034a016. ISSN 0002-7863.
  27. ^ Campana, Charles F.; Guzei, Ilia A.; Mednikov, Evgueni G.; Dahl, Lawrence F. (2013-12-19). "Sixty-Year Saga (1952–2013) of the Solid-State Structure of Triiron Dodecacarbonyl". Journal of Cluster Science. 25 (1): 205–224. doi:10.1007/s10876-013-0667-z. ISSN 1040-7278.
  28. ^ Cotton, F. Albert; Troup, Jan M. (1974). "Further refinement of the molecular structure of triiron dodecacarbonyl". Journal of the American Chemical Society. 96 (13): 4155–4159. doi:10.1021/ja00820a016. ISSN 0002-7863.
  29. ^ a b Lin, Yuexia; Wang, Hongyan; Wang, Xiaoting; Wang, Hui; King, R. Bruce (2022-08-08). "Alternatives to Triple-Decker Sandwich Structures for Binuclear Cyclooctatetraene First-Row Transition Metal Complexes of the Type (C 8 H 8 ) 3 M 2". Organometallics. 41 (15): 1977–1987. doi:10.1021/acs.organomet.2c00181. ISSN 0276-7333.
  30. ^ Breil, H.; Wilke, G. (1966). "Di(cyclooctatetraene)titanium and Tri(cyclooctatetraene)dititanium". Angewandte Chemie International Edition in English. 5 (10): 898–899. doi:10.1002/anie.196608982. ISSN 0570-0833.
  31. ^ Lauher, Joseph W.; Elian, Mihai; Summerville, Richard H.; Hoffmann, Roald (1976). "Triple-decker sandwiches". Journal of the American Chemical Society. 98 (11): 3219–3224. doi:10.1021/ja00427a028. ISSN 0002-7863.
  32. ^ Brauer, D. J.; Krueger, C. (1976-10-01). "Stereochemistry of transition metal cyclooctatetraenyl complexes. Tris(cyclooctatetraene)dichromium, a novel stereochemistry for a compound containing a quadruple metal-metal bond". Inorganic Chemistry. 15 (10): 2511–2514. doi:10.1021/ic50164a038. ISSN 0020-1669.