Dewar benzene

Last updated
Dewar benzene
Dewar benzene (edge on).svg
The conjoined cyclobutene rings of Dewar benzene form an obtuse angle.
Dewar-benzene-3D-balls.png
Names
Preferred IUPAC name
Bicyclo[2.2.0]hexa-2,5-diene
Identifiers
3D model (JSmol)
ChemSpider
PubChem CID
UNII
  • InChI=1S/C6H6/c1-2-6-4-3-5(1)6/h1-6H Yes check.svgY
    Key: CTLSARLLLBZBRV-UHFFFAOYSA-N Yes check.svgY
  • InChI=1/C6H6/c1-2-6-4-3-5(1)6/h1-6H
    Key: CTLSARLLLBZBRV-UHFFFAOYAO
  • C\1=C\C2/C=C\C/12
Properties
C6H6
Molar mass 78.1 g·mol1
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).

Dewar benzene (also spelled dewarbenzene) or bicyclo[2.2.0]hexa-2,5-diene is a bicyclic isomer of benzene with the molecular formula C6H6. The compound is named after James Dewar who included this structure in a list of possible C6H6 structures in 1869. [1] However, he did not propose it as the structure of benzene, and in fact he supported the correct structure previously proposed by August Kekulé in 1865. [2]

Contents

Structure and properties

Unlike benzene, Dewar benzene is not flat because the carbons where the rings join are bonded to four atoms rather than three. These carbons tend toward tetrahedral geometry, and the two cyclobutene rings make an angle where they are cis -fused to each other. The compound has nevertheless considerable strain energy and reverts to benzene with a chemical half-life of two days. This thermal conversion is relatively slow because it is symmetry forbidden based on orbital symmetry arguments. [3]

Synthesis

The compound itself was first synthesized in 1962 as a tert-butyl derivative [4] and then as the unsubstituted compound by Eugene van Tamelen in 1963 by photolysis of the cis-1,2-dihydro derivative of phthalic anhydride followed by oxidation with lead tetraacetate. [5] [6]

Dewar benzene synthesis.svg

"Dewar benzene" and benzene

Seven possible isomers proposed by Dewar, with "Dewar benzene" in second row, right. Historic Benzene Formula Dewar 1869 proposals (original).png
Seven possible isomers proposed by Dewar, with "Dewar benzene" in second row, right.

It is sometimes incorrectly claimed that Dewar proposed his structure as the true structure of benzene. In fact, Dewar merely wrote the structure as one of seven possible isomers [1] and believed that his experiments on benzene supported the (correct) structure that had been proposed by Kekulé. [2]

After the development of valence bond theory in 1928, benzene was described primarily using its two major resonance contributors, the two Kekulé structures. The three possible Dewar structures were considered as minor resonance contributors in the overall description of benzene, alongside other classic structures such as the isomers prismane, benzvalene and Claus' benzene. Prismane and benzvalene were synthesized in the 1970s; Claus' benzene is impossible to synthesize. [7]

Hexamethyl Dewar benzene

Hexamethyl Dewar benzene has been prepared by bicyclotrimerization of dimethylacetylene with aluminium chloride. [8] It undergoes a rearrangement reaction with hydrohalic acids to which the appropriate salt can be added to form the organometallic pentamethylcyclopentadienyl rhodium dichloride [9] [10] [11] [12] and pentamethylcyclopentadienyl iridium dichloride dimers; [13] consequently, it can be used as a starting material for synthesising some pentamethylcyclopentadienyl organometallic compounds [14] [15] including [Cp*Rh(CO)2]. [16] Attempting a similar reaction with potassium tetrachloroplatinate results in the formation of a pentamethylcyclopentadiene complex, [(η4-Cp*H)PtCl2], indicating that the rhodium and iridium metal centres are necessary for the step in which the aromatic anion is formed. [12]

Synthesis of the rhodium(III) dimer [Cp*RhCl2]2 from hexamethyl Dewar benzene Hexamethyl Dewar benzene reacting with rhodium chloride under acidic conditions.PNG
Synthesis of the rhodium(III) dimer [Cp*RhCl2]2 from hexamethyl Dewar benzene

One of the alkenes can be epoxidized using mCPBA, [17] peroxybenzoic acid, [18] or dimethyldioxirane (DMDO). [19] Using a peracid (mCPBA or peroxybenzoic acid), the epoxy product quickly rearranges, catalyzed by the acid byproduct of the epoxidation. [17]

Hexamethyl Dewar epoxidation-rearrangement.png

Using DMDO gives the epoxide as a stable product—the byproduct of the epoxidation is neutral acetone. By varying the amount of DMDO, either the mono- or diepoxide can be formed, with the oxygen atoms exo on the bicyclic carbon framework. [19]

Hexamethyl Dewar epoxidations.png

In 1973, the dication of hexamethylbenzene, C
6
(CH
3
)2+
6
, was produced by Hepke Hogeveen and Peter Kwant. [20] This can be done by dissolving the hexamethyl Dewar benzene monoepoxide in magic acid, which removes the oxygen as an anion. [21] NMR had previously hinted at a pentagonal pyramidal structure in a related cation [22] as had spectral data on the Hogeveen and Kwant dication. [23] [24] The pyramidal structure having an apex carbon bonding to six other carbon atoms was confirmed by X-ray crystallographic analysis of the hexafluoroantimonate salt published in 2016. [21]

C6(CH3)6 (SbF6)2 synthesis.png
Bachrach image of hexamethylbenzene dication.png
C6(CH3)6(2+) 3D skeletal.png
Left: Structure of C
6
(CH
3
)2+
6
, as drawn by Steven Bachrach [25]
Right: Three-dimensional representation of the dication's rearranged pentagonal-pyramid framework, from the crystal structure [21]

Computational organic chemist Steven Bachrach discussed the dication, noting that the weak bonds forming the upright edges of the pyramid, shown as dashed lines in the structure he drew, have a Wiberg bond order of about 0.54; it follows that the total bond order for the apical carbon is 5 × 0.54 + 1 = 3.7 < 4, and thus the species is not hypervalent, but it is hypercoordinate. [25] From the perspective of organometallic chemistry, the species can be viewed as having a carbon(IV) centre (C4+
) bound to an aromatic η5pentamethylcyclopentadienyl anion (six-electron donor) and a methyl anion (two-electron donor), thereby satisfying the octet rule [26] and being analogous to the gas-phase organozinc monomer [(η5
–C
5
(CH
3
)
5
)Zn(CH
3
)],
which has the same ligands bound to a zinc(II) centre (Zn2+
) and satisfies the 18 electron rule on the metal. [27] [28] Thus, while unprecedented, [21] and having attracted comment in Chemical & Engineering News , [29] New Scientist , [30] Science News , [31] and ZME Science, [32] the structure is consistent with the usual bonding rules of chemistry. Moritz Malischewski, who carried out the work with Konrad Seppelt, [21] commented that one the motivations for undertaking the work was to illustrate "the possibility to astonish chemists about what can be possible." [30]

Related Research Articles

Ferrocene is an organometallic compound with the formula Fe(C5H5)2. The molecule is a complex consisting of two cyclopentadienyl rings bound to a central iron atom. It is an orange solid with a camphor-like odor, that sublimes above room temperature, and is soluble in most organic solvents. It is remarkable for its stability: it is unaffected by air, water, strong bases, and can be heated to 400 °C without decomposition. In oxidizing conditions it can reversibly react with strong acids to form the ferrocenium cation Fe(C5H5)+2.

In organic chemistry, an alkyne trimerisation is a [2+2+2] cycloaddition reaction in which three alkyne units react to form a benzene ring. The reaction requires a metal catalyst. The process is of historic interest as well as being applicable to organic synthesis. Being a cycloaddition reaction, it has high atom economy. Many variations have been developed, including cyclisation of mixtures of alkynes and alkenes as well as alkynes and nitriles.

<span class="mw-page-title-main">Rhodium(III) chloride</span> Chemical compound

Rhodium(III) chloride refers to inorganic compounds with the formula RhCl3(H2O)n, where n varies from 0 to 3. These are diamagnetic solids featuring octahedral Rh(III) centres. Depending on the value of n, the material is either a dense brown solid or a soluble reddish salt. The soluble trihydrated (n = 3) salt is widely used to prepare compounds used in homogeneous catalysis, notably for the industrial production of acetic acid and hydroformylation.

<span class="mw-page-title-main">Pentamethylcyclopentadiene</span> Chemical compound

1,2,3,4,5-Pentamethylcyclopentadiene is a cyclic diene with the formula C5Me5H (Me = CH3). 1,2,3,4,5-Pentamethylcyclopentadiene is the precursor to the ligand 1,2,3,4,5-pentamethylcyclopentadienyl, which is often denoted Cp* (C5Me5) and read as "C P star", the "star" signifying the five methyl groups radiating from the core of the ligand. In contrast to less-substituted cyclopentadiene derivatives, Cp*H is not prone to dimerization.

<span class="mw-page-title-main">Homoaromaticity</span> Organic molecular structure

Homoaromaticity, in organic chemistry, refers to a special case of aromaticity in which conjugation is interrupted by a single sp3 hybridized carbon atom. Although this sp3 center disrupts the continuous overlap of p-orbitals, traditionally thought to be a requirement for aromaticity, considerable thermodynamic stability and many of the spectroscopic, magnetic, and chemical properties associated with aromatic compounds are still observed for such compounds. This formal discontinuity is apparently bridged by p-orbital overlap, maintaining a contiguous cycle of π electrons that is responsible for this preserved chemical stability.

<span class="mw-page-title-main">Hapticity</span> Number of contiguous atoms in a ligand that bond to the central atom in a coordination complex

In coordination chemistry, hapticity is the coordination of a ligand to a metal center via an uninterrupted and contiguous series of atoms. The hapticity of a ligand is described with the Greek letter η ('eta'). For example, η2 describes a ligand that coordinates through 2 contiguous atoms. In general the η-notation only applies when multiple atoms are coordinated. In addition, if the ligand coordinates through multiple atoms that are not contiguous then this is considered denticity, and the κ-notation is used once again. When naming complexes care should be taken not to confuse η with μ ('mu'), which relates to bridging ligands.

Carbon–hydrogen bond functionalization is a type of reaction in which a carbon–hydrogen bond is cleaved and replaced with a carbon–X bond. The term usually implies that a transition metal is involved in the C-H cleavage process. Reactions classified by the term typically involve the hydrocarbon first to react with a metal catalyst to create an organometallic complex in which the hydrocarbon is coordinated to the inner-sphere of a metal, either via an intermediate "alkane or arene complex" or as a transition state leading to a "M−C" intermediate. The intermediate of this first step can then undergo subsequent reactions to produce the functionalized product. Important to this definition is the requirement that during the C–H cleavage event, the hydrocarbyl species remains associated in the inner-sphere and under the influence of "M".

<i>sec</i>-Butyllithium Chemical compound

sec-Butyllithium is an organometallic compound with the formula CH3CHLiCH2CH3, abbreviated sec-BuLi or s-BuLi. This chiral organolithium reagent is used as a source of sec-butyl carbanion in organic synthesis.

<span class="mw-page-title-main">Organotitanium compound</span>

Organotitanium compounds in organometallic chemistry contain carbon-titanium chemical bonds. Organotitanium chemistry is the science of organotitanium compounds describing their physical properties, synthesis and reactions. They are reagents in organic chemistry and are involved in major industrial processes.

<span class="mw-page-title-main">Pentamethylcyclopentadienyl iridium dichloride dimer</span> Chemical compound

Pentamethylcyclopentadienyl iridium dichloride dimer is an organometallic compound with the formula [(C5(CH3)5IrCl2)]2, commonly abbreviated [Cp*IrCl2]2 This bright orange air-stable diamagnetic solid is a reagent in organometallic chemistry.

<span class="mw-page-title-main">Organocobalt chemistry</span> Chemistry of compounds with a carbon to cobalt bond

Organocobalt chemistry is the chemistry of organometallic compounds containing a carbon to cobalt chemical bond. Organocobalt compounds are involved in several organic reactions and the important biomolecule vitamin B12 has a cobalt-carbon bond. Many organocobalt compounds exhibit useful catalytic properties, the preeminent example being dicobalt octacarbonyl.

<span class="mw-page-title-main">Rhodocene</span> Organometallic chemical compound

Rhodocene is a chemical compound with the formula [Rh(C5H5)2]. Each molecule contains an atom of rhodium bound between two planar aromatic systems of five carbon atoms known as cyclopentadienyl rings in a sandwich arrangement. It is an organometallic compound as it has (haptic) covalent rhodium–carbon bonds. The [Rh(C5H5)2] radical is found above 150 °C (302 °F) or when trapped by cooling to liquid nitrogen temperatures (−196 °C [−321 °F]). At room temperature, pairs of these radicals join via their cyclopentadienyl rings to form a dimer, a yellow solid.

Transition metal carbyne complexes are organometallic compounds with a triple bond between carbon and the transition metal. This triple bond consists of a σ-bond and two π-bonds. The HOMO of the carbyne ligand interacts with the LUMO of the metal to create the σ-bond. The two π-bonds are formed when the two HOMO orbitals of the metal back-donate to the LUMO of the carbyne. They are also called metal alkylidynes—the carbon is a carbyne ligand. Such compounds are useful in organic synthesis of alkynes and nitriles. They have been the focus on much fundamental research.

Metal carbon dioxide complexes are coordination complexes that contain carbon dioxide ligands. Aside from the fundamental interest in the coordination chemistry of simple molecules, studies in this field are motivated by the possibility that transition metals might catalyze useful transformations of CO2. This research is relevant both to organic synthesis and to the production of "solar fuels" that would avoid the use of petroleum-based fuels.

Peter Michael Maitlis, FRS was a British organometallic chemist.

<span class="mw-page-title-main">Hexamethylbenzene</span> Chemical compound

Hexamethylbenzene, also known as mellitene, is a hydrocarbon with the molecular formula C12H18 and the condensed structural formula C6(CH3)6. It is an aromatic compound and a derivative of benzene, where benzene's six hydrogen atoms have each been replaced by a methyl group. In 1929 Kathleen Lonsdale reported the crystal structure of hexamethylbenzene, demonstrating that the central ring is hexagonal and flat and thereby ending an ongoing debate about the physical parameters of the benzene system. This was a historically significant result, both for the field of X-ray crystallography and for understanding aromaticity.

The Buchner ring expansion is a two-step organic C-C bond forming reaction used to access 7-membered rings. The first step involves formation of a carbene from ethyl diazoacetate, which cyclopropanates an aromatic ring. The ring expansion occurs in the second step, with an electrocyclic reaction opening the cyclopropane ring to form the 7-membered ring.

<span class="mw-page-title-main">Half sandwich compound</span> Class of coordination compounds

Half sandwich compounds, also known as piano stool complexes, are organometallic complexes that feature a cyclic polyhapto ligand bound to an MLn center, where L is a unidentate ligand. Thousands of such complexes are known. Well-known examples include cyclobutadieneiron tricarbonyl and (C5H5)TiCl3. Commercially useful examples include (C5H5)Co(CO)2, which is used in the synthesis of substituted pyridines, and methylcyclopentadienyl manganese tricarbonyl, an antiknock agent in petrol.

<span class="mw-page-title-main">Pyramidal carbocation</span>

A pyramidal carbocation is a type of carbocation with a specific configuration. This ion exists as a third class, besides the classical and non-classical ions. In these ions, a single carbon atom hovers over a four- or five-sided polygon, in effect forming a pyramid. The four-sided pyramidal ion will carry a charge of 1+, and the five-sided pyramid will carry 2+. In the images, the black spot on the vertical line represents the hovering carbon atom.

<span class="mw-page-title-main">Pentamethylcyclopentadienyl rhodium dichloride dimer</span> Chemical compound

Pentamethylcyclopentadienyl rhodium dichloride dimer is an organometallic compound with the formula [(C5(CH3)5RhCl2)]2, commonly abbreviated [Cp*RhCl2]2 This dark red air-stable diamagnetic solid is a reagent in organometallic chemistry.

References

  1. 1 2 Dewar, James (1869). "On the Oxidation af Phenyl Alcohol, and a Mechanical Arrangement adapted to illustrate Structure in the Non-saturated Hydrocarbons". Proc. R. Soc. Edinb. 6: 82–86. doi:10.1017/S0370164600045387.
  2. 1 2 Baker, Wilson; Rouvray, Dennis H. (1978). "Para-Bond or "Dewar" Benzene?". J. Chem. Educ. 55 (10): 645. Bibcode:1978JChEd..55..645B. doi:10.1021/ed055p645.
  3. Jensen, James O. (2004). "Vibrational Frequencies and Structural Determination of Dewar Benzene". J. Mol. Struct.:THEOCHEM . 680 (1–3): 227–236. doi:10.1016/j.theochem.2004.03.042.
  4. van Tamelen, Eugene E.; Pappas, S. P. (1962). "Chemistry of Dewar Benzene. 1,2,5-Tri-t-Butylbicyclo[2.2.0]Hexa-2,5-Diene". J. Am. Chem. Soc. 84 (19): 3789–3791. doi:10.1021/ja00878a054.
  5. van Tamelen, Eugene E.; Pappas, S. P. (1963). "Bicyclo [2.2.0]hexa-2,5-diene". J. Am. Chem. Soc. 85 (20): 3297–3298. doi:10.1021/ja00903a056.
  6. van Tamelen, Eugene E.; Pappas, S. P.; Kirk, K. L. (1971). "Valence Bond Isomers of Aromatic Systems. Bicyclo[2.2.0]hexa-2,5-dienes (Dewar benzenes)". J. Am. Chem. Soc. 93 (23): 6092–6101. doi:10.1021/ja00752a021.
  7. Hoffmann, Roald; Hopf, Henning (2008). "Learning from Molecules in Distress". Angew. Chem. Int. Ed. 47 (24): 4474–4481. doi:10.1002/anie.200705775. PMID   18418829.
  8. Shama, Sami A.; Wamser, Carl C. (1990). "Hexamethyl Dewar Benzene". Organic Syntheses . 61: 62. doi:10.15227/orgsyn.061.0062.; Collective Volume, vol. 7, p. 256
  9. Paquette, Leo A.; Krow, Grant R. (1968). "Electrophilic Additions to Hexamethyldewarbenzene". Tetrahedron Lett. 9 (17): 2139–2142. doi:10.1016/S0040-4039(00)89761-0.
  10. Criegee, Rudolf; Grüner, H. (1968). "Acid-catalyzed Rearrangements of Hexamethyl-prismane and Hexamethyl-Dewar-benzene". Angew. Chem. Int. Ed. 7 (6): 467–468. doi:10.1002/anie.196804672.
  11. Herrmann, Wolfgang A.; Zybill, Christian (1996). "Bis{(μ-chloro)[chloro(η-pentamethylcyclopentadienyl)rhodium]} {Rh(μ-Cl)Cl[η-C5(CH3)5]}2". In Herrmann, Wolfgang A.; Salzer, Albrecht (eds.). Synthetic Methods of Organometallic and Inorganic Chemistry Volume 1: Literature, Laboratory Techniques, and Common Starting Materials. Georg Thieme Verlag. pp. 148–149. ISBN   9783131791610.
  12. 1 2 Heck, Richard F. (1974). "Reactions of Dienes Trienes and Tetraenes with Transition Metal Compounds". Organotransition Metal Chemistry: A Mechanistic Approach. Academic Press. pp. 116–117. ISBN   9780323154703.
  13. Kang, Jung W.; Moseley, K.; Maitlis, Peter M. (1969). "Pentamethylcyclopentadienylrhodium and -iridium halides. I. Synthesis and properties". J. Am. Chem. Soc. 91 (22): 5970–5977. doi:10.1021/ja01050a008.
  14. Kang, J. W.; Mosley, K.; Maitlis, Peter M. (1968). "Mechanisms of Reactions of Dewar Hexamethylbenzene with Rhodium and Iridium Chlorides". Chem. Commun. (21): 1304–1305. doi:10.1039/C19680001304.
  15. Kang, J. W.; Maitlis, Peter M. (1968). "Conversion of Dewar Hexamethylbenzene to Pentamethylcyclopentadienylrhodium(III) Chloride". J. Am. Chem. Soc. 90 (12): 3259–3261. doi:10.1021/ja01014a063.
  16. Herrmann, Wolfgang A.; Zybill, Christian (1996). "Dicarbonyl(η-pentamethylcyclopentadienyl)rhodium Rh[η-C5(CH3)5](CO)2". In Herrmann, Wolfgang A.; Salzer, Albrecht (eds.). Synthetic Methods of Organometallic and Inorganic Chemistry Volume 1: Literature, Laboratory Techniques, and Common Starting Materials. Georg Thieme Verlag. pp. 147–148. ISBN   9783131791610.
  17. 1 2 King, R. B.; Douglas, W. M.; Efraty, A. (1977). "5-Acetyl-1,2,3,4,5-pentamethylcyclopentadiene". Organic Syntheses . 56: 1. doi:10.15227/orgsyn.056.0001.; Collective Volume, vol. 6, p. 39
  18. Junker, Hans-Nikolaus; Schäfer, Wolfgang; Niedenbrück, Hans (1967). "Oxydationsreaktionen mit Hexamethyl-bicyclo[2.2.0]-hexadien-(2.5) (= Hexamethyl-Dewar-Benzol)" [Oxidation reactions with hexamethylbicyclo[2.2.0]-hexa-2,5-diene (= Hexamethyl Dewar Benzene)]. Chem. Ber. (in German). 100 (8): 2508–2514. doi:10.1002/cber.19671000807.
  19. 1 2 Asouti, Amalia; Hadjiarapoglou, Lazaros P. (2000). "Regioselective and diastereoselective dimethyldioxirane epoxidation of substituted norbornenes and hexamethyl Dewar benzene". Tetrahedron Lett. 41 (4): 539–542. doi:10.1016/S0040-4039(99)02113-9.
  20. Hogeveen, Hepke; Kwant, Peter W. (1973). "Direct observation of a remarkably stable dication of unusual structure: (CCH3)62⊕". Tetrahedron Lett. 14 (19): 1665–1670. doi:10.1016/S0040-4039(01)96023-X.
  21. 1 2 3 4 5 Malischewski, Moritz; Seppelt, Konrad (2016). "Crystal Structure Determination of the Pentagonal-Pyramidal Hexamethylbenzene Dication C6(CH3)62+". Angew. Chem. Int. Ed. 56 (1): 368–370. doi:10.1002/anie.201608795. PMID   27885766.
  22. Paquette, Leo A.; Krow, Grant R.; Bollinger, J. Martin; Olah, George A. (1968). "Protonation of hexamethyl Dewar benzene and hexamethylprismane in fluorosulfuric acid – antimony pentafluoride – sulfur dioxide". J. Am. Chem. Soc. 90 (25): 7147–7149. doi:10.1021/ja01027a060.
  23. Hogeveen, Hepke; Kwant, Peter W.; Postma, J.; van Duynen, P. Th. (1974). "Electronic spectra of pyramidal dications, (CCH3)62+ and (CCH)62+". Tetrahedron Lett. 15 (49–50): 4351–4354. doi:10.1016/S0040-4039(01)92161-6.
  24. Hogeveen, Hepke; Kwant, Peter W. (1974). "Chemistry and spectroscopy in strongly acidic solutions. XL. (CCH3)62+, an unusual dication". J. Am. Chem. Soc. 96 (7): 2208–2214. doi:10.1021/ja00814a034.
  25. 1 2 Bachrach, Steven M. (January 17, 2017). "A six-coordinate carbon atom". comporgchem.com. Archived from the original on January 19, 2017. Retrieved January 18, 2017.
  26. Hogeveen, Hepke; Kwant, Peter W. (1975). "Pyramidal mono- and dications. Bridge between organic and organometallic chemistry". Acc. Chem. Res. 8 (12): 413–420. doi:10.1021/ar50096a004.
  27. Haaland, Arne; Samdal, Svein; Seip, Ragnhild (1978). "The molecular structure of monomeric methyl(cyclopentadienyl)zinc, (CH3)Zn(η-C5H5), determined by gas phase electron diffraction". J. Organomet. Chem. 153 (2): 187–192. doi:10.1016/S0022-328X(00)85041-X.
  28. Elschenbroich, Christoph (2006). "Organometallic Compounds of Groups 2 and 12". Organometallics (3rd ed.). John Wiley & Sons. pp. 59–85. ISBN   9783527805143.
  29. Ritter, Stephen K. (December 19, 2016). "Six bonds to carbon: Confirmed". Chem. Eng. News . 94 (49): 13. Archived from the original on January 9, 2017.
  30. 1 2 Boyle, Rebecca (January 14, 2017). "Carbon seen bonding with six other atoms for the first time". New Scientist (3108). Archived from the original on January 16, 2017. Retrieved January 14, 2017.
  31. Hamers, Laurel (December 24, 2016). "Carbon can exceed four-bond limit". Science News . 190 (13): 17. Archived from the original on February 3, 2017.
  32. Puiu, Tibi (January 5, 2017). "Exotic carbon molecule has six bonds, breaking the four-bond limit". zmescience.com. ZME Science. Archived from the original on January 16, 2017. Retrieved January 14, 2017.