Some propellanes. From left to right: [1.1.1]propellane, [2.2.2]propellane, and 1,3-dehydroadamantane (a methylene-bridged derivative of [3.3.1]propellane).

In organic chemistry, propellane is any member of a class of polycyclic hydrocarbons, whose carbon skeleton consists of three rings of carbon atoms sharing a common carbon–carbon covalent bond.[1][2] The concept was introduced in 1966 by D. Ginsburg [1][3] Propellanes with small cycles are highly strained and unstable, and are easily turned into polymers with interesting structures, such as staffanes. Partly for these reasons, they have been the object of much research.

Nomenclature

Lewis-Structure of an eneral carbocyclic propellane labelling bridge- and interbridgehead bond with x,y,z counters.
General nomenclature of carbocyclic Propellane.

The name derives from a supposed resemblance of the molecule to a propeller: namely, the rings would be the propeller's blades, and the shared C–C bond would be its axis. The bond shared by the three cycles is usually called the "bridge"; the shared carbon atoms are then the "bridgeheads".

The IUPAC nomenclature of the homologue series of all-carbon propellanes would be called tricyclo[x.y.z.01,(x+2)]alkane. More common in literature is the notation [x.y.z]propellane means the member of the family whose rings have x, y, and z carbons, not counting the two bridgeheads; or x + 2, y + 2, and z + 2 carbons, counting them. The chemical formula is therefore C
2+x+y+z
H
2(x+y+z)
. The minimum value for x, y, and z is 1, meaning three fused cyclopropyl-rings forming the [1.1.1]propellane. There is no structural ordering between the rings; for example, [1.3.2]propellane is the same substance as [3.2.1]propellane. Therefore, it is customary to sort the indices in decreasing order, xyz.

Further, heterosubstituted propellanes or structurally embedded propellane moieties exist and have been synthesised and follow a more complex nomenclature (see below).

General properties

Strain

Propellanes with small cycles, such as [1.1.1]propellane or [2.2.2]propellane, bear a high absolute strain energy. The two interbridgeheaded carbons and their bonds may even be described as an inverted tetrahedral geometry.

Computed Strain energies of Propellanes[4]
Propellane Strain energy
[1.1.1]Propellane 98 kcal mol−1
[3.1.1]Propellane 76 kcal mol−1
[2.1.1]Propellane 86 kcal mol−1
[2.2.1]Propellane 82 kcal mol−1
[3.2.1]Propellane 67 kcal mol−1

The resulting steric strain causes such compounds to be unstable and highly reactive. The interbridgehead C-C bond is easily broken (even spontaneously) to yield less-strained bicyclic or even monocyclic hydrocarbons. This so-called strain-release chemistry is used in strategies to access otherwise hard-to-obtain structures.

Surprisingly, the most strained member [1.1.1] is far more stable than the other small ring members ([2.1.1], [2.2.1], [2.2.2], [3.2.1], [3.1.1], and [4.1.1]),[5] which can be explained by special bonding situation of the interbridgehead bond.

Bonding properties

The bonding situation of small-ring propellanes, such as [n.1.1]propellanes, is topic of debate. Recent computational studies explain the interbridgehead bond as a Charge-shift bond possessing an unusual positive Laplace operator of the electron density .[6] Studies by Sterling et al. suggest delocalisation effects onto the three-membered bridges relaxing Pauli-repulsion and thus stabilising the propellane core.[7]

Reactivity

Propellanes, especially the synthetically studied [1.1.1]Propellane, is known to possess omniphilic reactivity. Anions and radicals add towards the interbridgehead bond resulting in bicyclo[1.1.1]pentyl-units. In contrary, cations and metals decompose the tricyclic core towards monocyclic systems by opening of the bridged bonds forming exo-methylene cyclobutanes.[8] For [3.1.1]propellane only radical addition is reported.[9][10] The reactivity of other propellanes is far less explored and their reactivity profile is less clear.

First line shows addition of radical and anions to [1.1.1]Propellane producing Bicyclo[1.1.1]pentane units. It also shows addition of Cations add to the bridgebond of [1.1.1]Propellane. The second line of the figure shows the general reactivity of [3.1.1]Propellane undergoing addition of radicals to produce bicyclo[3.1.1]heptyl-units (no cationic and anionic reactivity reported so far)
General reactivity profile of [1.1.1]Propellane and [3.1.1]Propellane

Polymerization

In principle, any propellane can be polymerized by breaking the axial C–C bond to yield a radical with two active centers, and then joining these radicals in a linear chain. For the propellanes with small cycles (such as [1.1.1], [3.2.1], or 1,3-dihydroadamantane), this process is easily achieved, yielding either simple polymers or alternating copolymers. For example, [1.1.1]propellane yields spontaneously an interesting rigid polymer called staffane;[11] and [3.2.1]propellane combines spontaneously with oxygen at room temperature to give a copolymer where the bridge-opened propellane units [–C8H12–] alternate with [–O–O–] groups.[12]

Synthesis

The smaller-cycle propellanes are difficult to synthesize because of their strain. Larger members are more easily obtained. Weber and Cook described in 1978 a general method which should yield [n.3.3]propellanes for any n ≥ 3.[13]

Members

True propellanes

Propellane derivatives

Propellane natural products

See also

References

  1. ^ a b Dilmaç, A. M.; Spuling, E.; de Meijere, A.; Bräse, S. (2017). "Propellanes—From a Chemical Curiosity to "Explosive" Materials and Natural Products". Angew. Chem. Int. Ed. 56 (21): 5684–5718. doi:10.1002/anie.201603951. PMID 27905166.
  2. ^ Osmont; et al. (2008). "Physicochemical Properties and Thermochemistry of Propellanes". Energy and Fuels. 22 (4): 2241–2257. doi:10.1021/ef8000423.
  3. ^ Altman, J.; Babad, E.; Itzchaki, J.; Ginsburg, D. (1966). "Propellanes—I". Tetrahedron. 22: 279–304. doi:10.1016/S0040-4020(01)82189-X.
  4. ^ Wiberg, Kenneth B. (1986). "The Concept of Strain in Organic Chemistry". Angew. Chem. Int. Ed. Engl. 25 (4): 312–322. doi:10.1002/anie.198603121.
  5. ^ a b Michl, Josef; Radziszewski, George J.; Downing, John W.; Wiberg, Kenneth B.; Walker, Frederick H.; Miller, Robert D.; Kovacic, Peter; Jawdosiuk, Mikolaj; Bonačić-Koutecký, Vlasta (1983). "Highly strained single and double bonds". Pure Appl. Chem. 55 (2): 315–321. doi:10.1351/pac198855020315.
  6. ^ Shaik, Sason; Danovich, David; Wu, Wei; Hiberty, Philippe C. (2009). "Charge-shift bonding and its manifestations in chemistry". Nature Chemistry. 1 (6): 443–449. doi:10.1038/nchem.327. PMID 21378912.
  7. ^ Sterling, Alistair J.; Dürr, Alexander B.; Smith, Russel C.; Anderson, Edward A.; Duarte, Fernanda (2020). "Rationalizing the diverse reactivity of [1.1.1]propellane through σ–π-delocalization". Chem. Sci. 11 (19): 4895–4903. doi:10.1039/D0SC01386B. PMC 8159217. PMID 34122945.
  8. ^ Wiberg, Kenneth B.; Waddell, Sherman T. (1990). "Reactions of [1.1.1]propellane". J. Am. Chem. Soc. 112 (6): 2194–2216. doi:10.1021/ja00162a022.
  9. ^ a b Fuchs, Josef; Szeimies, Günter (1992). "Synthese von [n.1.1]Propellanen (n = 2, 3, 4)". Chem. Ber. 125 (11): 2517–2522. doi:10.1002/cber.19921251126.
  10. ^ a b c Frank, Nils; Nugent, Jeremy; Shire, Bethany R.; Pickford, Helena D.; Rabe, Patrick; Sterling, Alistair J.; Zarganes-Tzitzikas, Tryfon; Grimes, Thomas; Thompson, Amber L.; Smith, Russell C.; Schofield, Christopher J.; Brennan, Paul E.; Duarte, Fernanda; Anderson, Edward A. (2022). "Synthesis of meta-substituted arene bioisosteres from [3.1.1]propellane". Nature. 611 (7937): 721–726. doi:10.1038/s41586-022-05290-z. PMID 36108675. S2CID 252310498.
  11. ^ a b Kaszynski, Piotr; Michl, Josef (1988). "[n]Staffanes: a molecular-size "Tinkertoy" construction set for nanotechnology. Preparation of end-functionalized telomers and a polymer of [1.1.1]propellane". J. Am. Chem. Soc. 110 (15): 5225–5226. doi:10.1021/ja00223a070.
  12. ^ a b Wiberg, Kenneth B.; Burgmaier, George J. (1972). "Tricyclo[3.2.1.01,5]octane. A 3,2,1-Propellane". J. Am. Chem. Soc. 94 (21): 7396–7401. doi:10.1021/ja00776a022.
  13. ^ a b c d e Weber, Robert W.; Cook, James M. (1978). "General method for the synthesis of [n.3.3]propellanes, n ≥ 3". Can. J. Chem. 56 (2): 189–192. doi:10.1139/v78-030.
  14. ^ Wiberg, Kenneth B.; Walker, Frederick H. (1982). "[1.1.1]Propellane". J. Am. Chem. Soc. 104 (19): 5239–5240. doi:10.1021/ja00383a046.
  15. ^ Stepan, Antonia F.; et al. (2012). "Application of the Bicyclo[1.1.1]pentane Motif as a Nonclassical Phenyl Ring Bioisostere in the Design of a Potent and Orally Active γ-Secretase Inhibitor". J. Med. Chem. 55 (7): 3414–3424. doi:10.1021/jm300094u. PMID 22420884.
  16. ^ Jarosch, Oliver; Szeimies, Günter (2003). "Thermal Behavior of [2.1.1]Propellane: A DFT/Ab Initio Study". J. Org. Chem. 68 (10): 3797–3801. doi:10.1021/jo020741d. PMID 12737556.
  17. ^ Walker, Frederick H.; Wiberg, Kenneth B.; Michl, Josef (1982). "[2.2.1]Propellane". J. Am. Chem. Soc. 104 (7): 2056. doi:10.1021/ja00371a059.
  18. ^ Gassman, P. G.; Proehl, G. S. (1980). "[3.1.1]Propellane". J. Am. Chem. Soc. 102 (22): 6862. doi:10.1021/ja00542a040.
  19. ^ Wiberg, Kenneth B.; Burgmaier, George J. (1969). "Tricyclo[3.2.1.01,5]octane". Tetrahedron Letters. 10 (5): 317–319. doi:10.1016/s0040-4039(01)87681-4.
  20. ^ Gassman, Paul G.; Topp, Alwin; Keller, John W. (1969). "Tricyclo[3.2.1.01,5]octane – a highly strained "propellerane"". Tetrahedron Letters. 10 (14): 1093–1095. doi:10.1016/s0040-4039(01)97748-2.
  21. ^ Aue, D. H.; Reynolds, R. N. (1974). "Reactions of a highly strained propellane. Tetracyclo[4.2.1.12,5.O1,6]decane". J. Org. Chem. 39 (15): 2315. doi:10.1021/jo00929a051.
  22. ^ a b Wiberg, Kenneth B.; Pratt, William E.; Bailey, William F. (1977). "Reaction of 1,4-diiodonorbornane, 1,4-diiodobicyclo[2.2.2]octane, and 1,5-diiodobicyclo[3.2.1]octane with butyllithium. Convenient preparative routes to the [2.2.2]- and [3.2.1]propellanes". J. Am. Chem. Soc. 99 (7): 2297–2302. doi:10.1021/ja00449a045.
  23. ^ Hamon, David P. G.; Trenerry, V. Craige (1981). "Carbenoid insertion reactions: formation of [4.1.1]propellane". J. Am. Chem. Soc. 103 (16): 4962–4965. doi:10.1021/ja00406a059.
  24. ^ Szeimies-Seebach, Ursula; Harnish, J.; Szeimies, Günter; Meerssche, M. V.; Germain, G.; Declerq, J. P. (1978). "Existence of a New C6H6 Isomer: Tricyclo[3.1.0.02,6]hex-1(6)-ene". Angew. Chem. Int. Ed. Engl. 17 (11): 848. doi:10.1002/anie.197808481.
  25. ^ Szeimies-Seebach, Ursula; Szeimies, Günter (1978). "A facile route to the [4.1.1]propellane system". J. Am. Chem. Soc. 100 (12): 3966–3967. doi:10.1021/ja00480a072.
  26. ^ Eaton, Philip E.; Temme, George H. (1973). "[2.2.2]Propellane system". J. Am. Chem. Soc. 95 (22): 7508–7510. doi:10.1021/ja00803a052.
  27. ^ Yang, S.; Cook, James M. (1976). "Reactions of dicarbonyl compounds with dimethyl β-ketoglutarate: II. Simple synthesis of compounds of the [10.3.3]- and [6.3.3]-propellane series". J. Org. Chem. 41 (11): 1903–1907. doi:10.1021/jo00873a004.
  28. ^ Pincock, Richard E.; Torupka, Edward J. (1969). "Tetracyclo[3.3.1.13,7.01,3]decane. Highly reactive 1,3-dehydro derivative of adamantane". J. Am. Chem. Soc. 91 (16): 4593. doi:10.1021/ja01044a072.
  29. ^ Mlinaric-Majerski, K.; Majerski, Z. (1980). "2,4-Methano-2,4-dehydroadamantane. A [3.1.1]propellane". J. Am. Chem. Soc. 102 (4): 1418. doi:10.1021/ja00524a033.
  30. ^ Tian, X; Li, L; Hu, Y; Zhang, H; Liu, Y; Chen, H; Ding, G; Zou, Z (2013). "Dichrocephones A and B, two cytotoxic sesquiterpenoids with the unique [3.3.3] propellane nucleus skeleton from Dichrocephala benthamii". RSC Adv. 3 (19): 7880–7883. Bibcode:2013RSCAd...3.7880T. doi:10.1039/C3RA23364B.
  31. ^ Schmiedel, V. M.; Hong, Y. J.; Lentz, D; Tantillo, D. J.; Christmann, M (2018). "Synthesis and Structure Revision of Dichrocephones A and B". Angew. Chem. Int. Ed. 57 (9): 2419–2422. doi:10.1002/anie.201711766. PMID 29251825.
  32. ^ Schneider, L. M.; Schmiedel, V. M.; Pecchioli, T; Lentz, T; Merten, C; Christmann, M (2017). "Asymmetric Synthesis of Carbocyclic Propellanes". Org. Lett. 19 (9): 2310–2313. doi:10.1021/acs.orglett.7b00836. PMID 28445060.