Hexamethylbenzene

Last updated

Contents

Hexamethylbenzene
Hexamethylbenzene.svg
Hexamethylbenzene 3D ball.png
HMB2wBottle.jpg
Names
Preferred IUPAC name
Hexamethylbenzene
Other names
1,2,3,4,5,6-Hexamethylbenzene
Mellitene
Identifiers
3D model (JSmol)
ChEBI
ChemSpider
ECHA InfoCard 100.001.616 OOjs UI icon edit-ltr-progressive.svg
PubChem CID
UNII
  • InChI=1S/C12H18/c1-7-8(2)10(4)12(6)11(5)9(7)3/h1-6H3 X mark.svgN
    Key: YUWFEBAXEOLKSG-UHFFFAOYSA-N X mark.svgN
  • InChI=1/C12H18/c1-7-8(2)10(4)12(6)11(5)9(7)3/h1-6H3
    Key: YUWFEBAXEOLKSG-UHFFFAOYAF
  • c1(c(c(c(c(c1C)C)C)C)C)C
Properties
C12H18
Molar mass 162.276 g·mol−1
AppearanceWhite crystalline powder
Density 1.0630 g cm−3
Melting point 165.6 ± 0.7 °C
Boiling point 265.2 °C (509.4 °F; 538.3 K)
insoluble
Solubility acetic acid, acetone, benzene, chloroform, diethyl ether, ethanol
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).
X mark.svgN  verify  (what is  Yes check.svgYX mark.svgN ?)

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 It is an aromatic compound and a derivative of benzene, where benzene's six hydrogen atoms have each been replaced by a methyl group.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 [1] 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. [2] [3]

Hexamethylbenzene can be oxidised to mellitic acid, [4] which is found in nature as its aluminium salt in the rare mineral mellite. [5] Hexamethylbenzene can be used as a ligand in organometallic compounds. [6] An example from organoruthenium chemistry shows structural change in the ligand associated with changes in the oxidation state of the metal centre, [7] [8] though the same change is not observed in the analogous organoiron system. [7]

In 2016 the crystal structure of the hexamethylbenzene dication C
6
(CH
3
)2+
6
was reported in Angewandte Chemie International Edition , [9] showing a pyramidal structure in which a single carbon atom has a bonding interaction with six other carbon atoms. [10] [11] This structure was "unprecedented", [9] as the usual maximum valence of carbon is four, and it attracted attention from New Scientist , [10] Chemical & Engineering News , [11] and Science News . [12] The structure does not violate the octet rule since the carbon–carbon bonds formed are not two-electron bonds, and is pedagogically valuable for illustrating that a carbon atom "can [directly bond] with more than four atoms". [12] Steven Bachrach has demonstrated that the compound is hypercoordinated but not hypervalent, and also explained its aromaticity. [13] The idea of describing the chemical bonding in compounds and chemical species in this way through the lens of organometallic chemistry was proposed in 1975, [14] soon after the dication C
6
(CH
3
)2+
6
was first observed. [15] [16] [17]

Nomenclature and properties

According to the Blue Book , this chemical can be systematically named as 1,2,3,4,5,6-hexamethylbenzene.[ citation needed ] The locants (the numbers in front of the name) are superfluous, however, as the name hexamethylbenzene uniquely identifies a single substance and thus is the formal IUPAC name for the compound. [18] It is an aromatic compound, with six π electrons (satisfying Hückel's rule) delocalised over a cyclic planar system; each of the six ring carbon atoms is sp2 hybridised and displays trigonal planar geometry, while each methyl carbon is tetrahedral with sp3 hybridisation, consistent with the empirical description of its structure. [1] Solid hexamethylbenzene occurs as colourless to white crystalline orthorhombic prisms or needles [19] with a melting point of 165–166 °C, [20] a boiling point of 268 °C, and a density of 1.0630 g cm−3. [19] It is insoluble in water, but soluble in organic solvents including benzene and ethanol. [19]

Mellite-20130.jpg
Mellitic-acid.svg
The mineral mellite (left) is composed of a hydrated aluminium salt of mellitic acid (right)

Hexamethylbenzene is sometimes called mellitene, [19] a name derived from mellite, a rare honey-coloured mineral ( μέλι meli ( GEN μέλιτοςmelitos) is the Greek word for honey. [21] ) Mellite is composed of a hydrated aluminium salt of benzenehexacarboxylic acid (mellitic acid), with formula Al
2
[C
6
(CO
2
)
6
]16H
2
O
. [5] Mellitic acid itself can be derived from the mineral, [22] and subsequent reduction yields mellitene. Conversely, mellitene can be oxidised to form mellitic acid: [4]

Mellitic redox.png

Treatment of hexamethylbenzene with a superelectrophilic mixture of methyl chloride and aluminum trichloride (a source of Meδ⊕Cl---δ⊖AlCl3) gives heptamethylbenzenium cation, one of the first carbocations to be directly observed.

Heptamethylbenzenium2.png

Structure

In 1927 Kathleen Lonsdale determined the solid structure of hexamethylbenzene from crystals provided by Christopher Kelk Ingold. [3] Her X-ray diffraction analysis was published in Nature [23] and was subsequently described as "remarkable ... for that early date". [3] Lonsdale described the work in her book Crystals and X-Rays, [24] explaining that she recognised that, though the unit cell was triclinic, the diffraction pattern had pseudo-hexagonal symmetry that allowed the structural possibilities to be restricted sufficiently for a trial-and-error approach to produce a model. [3] This work definitively showed that hexamethylbenzene is flat and that the carbon-to-carbon distances within the ring are the same, [2] providing crucial evidence in understanding the nature of aromaticity.

Preparation

The compound can be prepared by reacting phenol with methanol at elevated temperatures over a suitable solid catalyst such as alumina. [25] [20] [26] The mechanism of the process has been studied extensively, [27] [28] [29] [30] with several intermediates having been identified. [26] [31] [32] Alkyne trimerisation of dimethylacetylene also yields hexamethylbenzene [33] in the presence of a suitable catalyst. [34] [35]

In 1880, Joseph Achille Le Bel and William H. Greene reported [36] what has been described as an "extraordinary" zinc chloride-catalysed one-pot synthesis of hexamethylbenzene from methanol. [37] At the catalyst's melting point (283 °C), the reaction has a Gibbs free energy (ΔG) of −1090 kJ mol−1 and can be idealised as: [37]

15 CH
3
OH
  C
6
(CH
3
)
6
  +   3 CH
4
  +   15 H
2
O

Le Bel and Greene rationalised the process as involving aromatisation by condensation of methylene units, formed by dehydration of methanol molecules, followed by complete Friedel–Crafts methylation of the resulting benzene ring with chloromethane generated in situ. [37] The major products were a mixture of saturated hydrocarbons, with hexamethylbenzene as a minor product. [38] Hexamethylbenzene is also produced as a minor product in the Friedel–Crafts alkylation synthesis of durene from p-xylene, and can be produced by alkylation in good yield from durene or pentamethylbenzene. [39]

Hexamethylbenzene is typically prepared in the gas phase at elevated temperatures over solid catalysts. An early approach to preparing hexamethylbenzene involved reacting a mixture of acetone and methanol vapours over an alumina catalyst at 400 °C. [40] Combining phenols with methanol over alumina in a dry carbon dioxide atmosphere at 410–440 °C also produces hexamethylbenzene, [25] though as part of a complex mixture of anisole (methoxybenzene), cresols (methylphenols), and other methylated phenols. [31] An Organic Syntheses preparation, using methanol and phenol with an alumina catalyst at 530 °C, gives approximately a 66% yield, [20] though synthesis under different conditions has also been reported. [26]

Hexamethylbenzene synthesis.png

The mechanisms of such surface-mediated reactions have been investigated, with an eye to achieving greater control over the outcome of the reaction, [28] [41] especially in search of selective and controlled ortho -methylation. [29] [30] [42] [43] Both anisole [31] and pentamethylbenzene [26] have been reported as intermediates in the process. Valentin Koptyug and co-workers found that both hexamethylcyclohexadienone isomers (2,3,4,4,5,6- and 2,3,4,5,6,6-) are intermediates in the process, undergoing methyl migration to form the 1,2,3,4,5,6-hexamethylbenzene carbon skeleton. [27] [32]

Trimerisation of three 2-butyne (dimethylacetylene) molecules yields hexamethylbenzene. [33] The reaction is catalyzed by triphenylchromium tri-tetrahydrofuranate [34] or by a complex of triisobutylaluminium and titanium tetrachloride. [35]

2-Butyne cyclotrimerization.png

Uses

Synthetic uses

Hexamethylbenzene can be used as a ligand in organometallic compounds.

Other uses

Hexamethylbenzene has no commercial or widespread uses. It is exclusively of interest for chemical research.

Reactions

It forms orange-yellow 1:1 adduct with picryl chloride, [44] probably due to π-stacking of the aromatic systems.

Oxidation with trifluoroperacetic acid or hydrogen peroxide gives 2,3,4,5,6,6-hexamethyl-2,4-cyclohexadienone: [45] [27] [32] )

Hexamethylbenzene oxidation.png

It has also been used as a solvent for 3He-NMR spectroscopy. [46]

Just as with benzene itself, the electron-rich aromatic system in hexamethylbenzene allows it to act as a ligand in organometallic chemistry. [6] The electron-donating nature of the methyl groups—both that there are six of them individually and that there are six meta pairs among them—enhance the basicity of the central ring by six to seven orders of magnitude relative to benzene. [47] Examples of such complexes have been reported for a variety of metal centres, including cobalt, [48] chromium, [34] iron, [7] rhenium, [49] rhodium, [48] ruthenium, [8] and titanium. [35] Known cations of sandwich complexes of cobalt and rhodium with hexamethylbenzene take the form [M(C
6
(CH
3
)
6
)
2
]
n+ (M = Co, Fe, Rh, Ru; n = 1, 2), where the metal centre is bound by the π electrons of the two arene moieties, and can easily be synthesised from appropriate metal salts by ligand exchange, for example: [48]

CoBr
2
  +   2 AlBr
3
  [Co(C
6
(CH
3
)
6
)
2
]2+
  +   2 AlBr
4

The complexes can undergo redox reactions. The rhodium and cobalt dications undergo a one-electron reduction with a suitable active metal (aluminium for the cobalt system, zinc for the rhodium), and the equations describing the reactions in the cobalt system are as follows: [48]

3 [Co(C
6
(CH
3
)
6
)
2
]2+
  +   Al    3 [Co(C
6
(CH
3
)
6
)
2
]+
  +  Al3+
The structure of the [Ru(C6(CH3)6)2] moiety changes with the oxidation state of the metal centre
Left: n = 2, [Ru (e -C6(CH3)6)2]
Right: n = 0, [Ru (e -C6(CH3)6)(e -C6(CH3)6)]
Methyl groups omitted for clarity. The electron-pairs involved with carbon-ruthenium bonding are in red. Ru(C6H6)2redox pi-highlight.png
The structure of the [Ru(C6(CH3)6)2] moiety changes with the oxidation state of the metal centre
Left: n = 2, [Ru (η -C6(CH3)6)2]
Right: n = 0, [Ru (η -C6(CH3)6)(η -C6(CH3)6)]
Methyl groups omitted for clarity. The electron-pairs involved with carbon–ruthenium bonding are in red.

In the field of organoruthenium chemistry, the redox interconversion of the analogous two-electron reduction of the dication and its neutral product occurs at −1.02 V in acetonitrile [7] and is accompanied by a structural change. [8] [50] The hapticity of one of the hexamethylbenzene ligands changes with the oxidation state of the ruthenium centre, the dication [Ru(η6-C6(CH3)6)2]2+ being reduced to [Ru(η4-C6(CH3)6)(η6-C6(CH3)6)], [8] with the structural change allowing each complex to comply with the 18-electron rule and maximise stability.

The equivalent iron(II) complex undergoes a reversible one-electron reduction (at −0.48 V in aqueous ethanol), but the two-electron reduction (at −1.46 V) is irreversible, [7] suggesting a change in structure different from that found in the ruthenium system.

Dication

Pyramidal carbocation with composition C
6(CH
3)
6 C6(CH3)6 2+.png
Pyramidal carbocation with composition C
6
(CH
3
)
6

The isolation of an ion with composition C
6
(CH
3
)
6
H+
was first reported from investigations of hexamethyl Dewar benzene in the 1960s; [51] a pyramidal structure was suggested based on NMR evidence [52] and subsequently supported by disordered [9] crystal structure data. [53] In the early 1970s theoretical work led by Hepke Hogeveen predicted the existence of a pyramidal dication C
6
(CH
3
)2+
6
, and the suggestion was soon supported by experimental evidence. [15] [16] [17] Spectroscopic investigation of the two-electron oxidation of benzene at very low temperatures (below 4 K) shows that a hexagonal dication forms and then rapidly rearranges into a pyramidal structure: [54]

Oxidation of benzene to its dication.jpg
Three-dimensional representation of C
6(CH
3)
6 having a rearranged pentagonal-pyramid framework C6(CH3)6(2+) 3D skeletal.png
Three-dimensional representation of C
6
(CH
3
)
6
having a rearranged pentagonal-pyramid framework

Two-electron oxidation of hexamethylbenzene would be expected to result in a near-identical rearrangement to a pyramidal carbocation, but attempts to synthesise it in bulk by this method have been unsuccessful. [9] However, a modification of the Hogeveen approach was reported in 2016, along with a high-quality crystal structure determination of [C
6
(CH
3
)
6
][SbF
6
]
2
HSO
3
F
. The pyramidal core is about 1.18  ångströms high, and each of the methyl groups on the ring is located slightly above that base plane [9] to give a somewhat inverted tetrahedral geometry for the carbons of the base of the pyramid. The preparation method involved treating the epoxide of hexamethyl Dewar benzene with magic acid, which formally abstracts an oxide anion (O2−
) to form the dication: [9]

C6(CH3)6 (SbF6)2 synthesis.png

Though indirect spectroscopic evidence and theoretical calculations previously pointed to their existence, the isolation and structural determination of a species with a hexacoordinate carbon bound only to other carbon atoms is unprecedented, [9] and has attracted comment in Chemical & Engineering News , [11] New Scientist , [10] Science News , [12] and ZME Science. [55] The carbon atom at the top of the pyramid is bonding with six other atoms, an unusual arrangement as the usual maximum valence for this element is four. [11] The molecule is aromatic and avoids exceeding the octet on carbon by having only a total of six electrons in the five bonds between the base of the pyramid and its apex. That is, each of the vertical edges of the pyramid is only a partial bond rather than a normal covalent bond that would have two electrons shared between two atoms. Although the top carbon does bond to six others, it does so using a total of no more than eight electrons. [14]

The dication, noting the weak bonds forming the upright edges of the pyramid, shown as dashed lines in the structure, have a Wiberg bond order of about 0.54; it follows that the total bond order is 5 × 0.54 + 1 = 3.7 < 4, and thus the species is not hypervalent, though it is hypercoordinate. [13] The differences in bonding in the dication—the ring having aromatic character and the vertical edges being weak partial bonds—are reflected in variations of the carbon–carbon bond lengths: the ring bonds are 1.439–1.445 Å,, the bonds to the methyl groups are 1.479–1.489 Å,, and the vertical edges are 1.694–1.715 Å. [9] Bachrach rationalised the three-dimensional aromaticity of the dication by considering it as comprising the ring C
5
(CH
3
)+
5
as a four-electron donor and topped by the CCH+
3
fragment, which provides two electrons, for a total of six electrons in the aromatic cage, in line with Hückel's rule for n = 1. [13] From the perspective of organometallic chemistry, the species can be viewed as [(η5
–C
5
(CH
3
)
5
)C(CH
3
)]
. [14]
This satisfies the octet rule by binding a carbon(IV) centre (C4+
) to an aromatic η5pentamethylcyclopentadienyl anion (six-electron donor) and methyl anion (two-electron donor), analogous to the way the gas-phase organozinc monomer [(η5
–C
5
(CH
3
)
5
)Zn(CH
3
)],
having the same ligands bound to a zinc(II) centre (Zn2+
) satisfies the 18 electron rule on the metal. [56] [57]

Bachrach image of hexamethylbenzene dication.png
Structure of organometallic zinc complex (C5Me5)Zn(Me).png
Left: Structure of C
6
(CH
3
)2+
6
, as drawn by Steven Bachrach [13]
Right: The analogous organometallic complex [(η5
–C
5
(CH
3
)
5
)Zn(CH
3
)] [56]

It has been commented that "[i]t's super important that people realize that, although we're taught carbon can only have four friends, carbon can be associated with more than four atoms" and added that the "carbon isn't making six bonds in the sense that we usually think of a carbon-carbon bond as a two-electron bond." [12] "It is all about the challenge and the possibility to astonish chemists about what can be possible." [10]

Related Research Articles

<span class="mw-page-title-main">Aromatic compound</span> Compound containing rings with delocalized pi electrons

Aromatic compounds or arenes usually refers to organic compounds "with a chemistry typified by benzene" and "cyclically conjugated." The word "aromatic" originates from the past grouping of molecules based on odor, before their general chemical properties were understood. The current definition of aromatic compounds does not have any relation to their odor. Aromatic compounds are now defined as cyclic compounds satisfying Hückel's Rule. Aromatic compounds have the following general properties:

Ferrocene is an organometallic compound with the formula Fe(C5H5)2. The molecule is a complex consisting of two cyclopentadienyl rings sandwiching 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. Ferrocene and the ferrocenium cation are sometimes abbreviated as Fc and Fc+ respectively.

<span class="mw-page-title-main">Aromaticity</span> Chemical property

In organic chemistry, aromaticity is a chemical property describing the way in which a conjugated ring of unsaturated bonds, lone pairs, or empty orbitals exhibits a stabilization stronger than would be expected by the stabilization of conjugation alone. The earliest use of the term was in an article by August Wilhelm Hofmann in 1855. There is no general relationship between aromaticity as a chemical property and the olfactory properties of such compounds.

<span class="mw-page-title-main">Organolithium reagent</span> Chemical compounds containing C–Li bonds

In organometallic chemistry, organolithium reagents are chemical compounds that contain carbon–lithium (C–Li) bonds. These reagents are important in organic synthesis, and are frequently used to transfer the organic group or the lithium atom to the substrates in synthetic steps, through nucleophilic addition or simple deprotonation. Organolithium reagents are used in industry as an initiator for anionic polymerization, which leads to the production of various elastomers. They have also been applied in asymmetric synthesis in the pharmaceutical industry. Due to the large difference in electronegativity between the carbon atom and the lithium atom, the C−Li bond is highly ionic. Owing to the polar nature of the C−Li bond, organolithium reagents are good nucleophiles and strong bases. For laboratory organic synthesis, many organolithium reagents are commercially available in solution form. These reagents are highly reactive, and are sometimes pyrophoric.

A sigmatropic reaction in organic chemistry is a pericyclic reaction wherein the net result is one σ-bond is changed to another σ-bond in an uncatalyzed intramolecular reaction. The name sigmatropic is the result of a compounding of the long-established sigma designation from single carbon–carbon bonds and the Greek word tropos, meaning turn. In this type of rearrangement reaction, a substituent moves from one part of a π-bonded system to another part in an intramolecular reaction with simultaneous rearrangement of the π system. True sigmatropic reactions are usually uncatalyzed, although Lewis acid catalysis is possible. Sigmatropic reactions often have transition-metal catalysts that form intermediates in analogous reactions. The most well-known of the sigmatropic rearrangements are the [3,3] Cope rearrangement, Claisen rearrangement, Carroll rearrangement, and the Fischer indole synthesis.

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.

In molecular geometry, bond length or bond distance is defined as the average distance between nuclei of two bonded atoms in a molecule. It is a transferable property of a bond between atoms of fixed types, relatively independent of the rest of the molecule.

<span class="mw-page-title-main">Carbenium ion</span> Class of ions

A carbenium ion is a positive ion with the structure RR′R″C+, that is, a chemical species with carbon atom having three covalent bonds, and it bears a +1 formal charge. But IUPAC confuses coordination number with valence, incorrectly considering carbon in carbenium as trivalent.

<span class="mw-page-title-main">Arenium ion</span> Forms during electrophilic substitution on benzene ring

An arenium ion in organic chemistry is a cyclohexadienyl cation that appears as a reactive intermediate in electrophilic aromatic substitution. For historic reasons this complex is also called a Wheland intermediate, after American chemist George Willard Wheland (1907–1976). They are also called sigma complexes. The smallest arenium ion is the benzenium ion, which is protonated benzene.

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

1,2,3,4,5-Pentamethylcyclopentadiene is a cyclic diene with the formula C5(CH3)5H, often written C5Me5H, where Me is CH3. It is a colorless liquid.

In organic chemistry and organometallic chemistry, carbon–hydrogen bond activation is a type of organic reaction in which a carbon–hydrogen bond is cleaved and replaced with a C−X bond. Some authors further restrict the term C–H activation to reactions in which a C–H bond, one that is typically considered to be "unreactive", interacts with a transition metal center M, resulting in its cleavage and the generation of an organometallic species with an M–C bond. The intermediate of this step could then undergo subsequent reactions with other reagents, either in situ or in a separate step, to produce the functionalized product.

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

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. 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.

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

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

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

Hexamethyltungsten is the chemical compound W(CH3)6 also written WMe6. Classified as a transition metal alkyl complex, hexamethyltungsten is an air-sensitive, red, crystalline solid at room temperature; however, it is extremely volatile and sublimes at −30 °C. Owing to its six methyl groups it is extremely soluble in petroleum, aromatic hydrocarbons, ethers, carbon disulfide, and carbon tetrachloride.

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

In organometallic chemistry, a metallacycle is a derivative of a carbocyclic compound wherein a metal has replaced at least one carbon center; this is to some extent similar to heterocycles. Metallacycles appear frequently as reactive intermediates in catalysis, e.g. olefin metathesis and alkyne trimerization. In organic synthesis, directed ortho metalation is widely used for the functionalization of arene rings via C-H activation. One main effect that metallic atom substitution on a cyclic carbon compound is distorting the geometry due to the large size of typical metals.

An insertion reaction is a chemical reaction where one chemical entity interposes itself into an existing bond of typically a second chemical entity e.g.:

<span class="mw-page-title-main">Peter Maitlis</span> British chemist (1933–2022)

Peter Michael Maitlis, FRS was a British organometallic chemist.

<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">Trifluoroperacetic acid</span> Chemical compound

Trifluoroperacetic acid is an organofluorine compound, the peroxy acid analog of trifluoroacetic acid, with the condensed structural formula CF
3
COOOH
. It is a strong oxidizing agent for organic oxidation reactions, such as in Baeyer–Villiger oxidations of ketones. It is the most reactive of the organic peroxy acids, allowing it to successfully oxidise relatively unreactive alkenes to epoxides where other peroxy acids are ineffective. It can also oxidise the chalcogens in some functional groups, such as by transforming selenoethers to selones. It is a potentially explosive material and is not commercially available, but it can be quickly prepared as needed. Its use as a laboratory reagent was pioneered and developed by William D. Emmons.

<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 Lonsdale, Kathleen (1929). "The Structure of the Benzene Ring in Hexamethylbenzene". Proc. R. Soc. A . 123 (792): 494–515. doi: 10.1098/rspa.1929.0081 .
  2. 1 2 Lydon, John (January 2006). "A Welcome to Leeds" (PDF). Newsletter of the History of Physics Group (19): 8–11.
  3. 1 2 3 4 Lydon, John (July 2006). "Letters" (PDF). Newsletter of the History of Physics Group (20): 34–35.
  4. 1 2 Wibaut, J. P.; Overhoff, J.; Jonker, E. W.; Gratama, K. (1941). "On the preparation of mellitic acid from hexa-methylbenzene and on the hexachloride of mellitic acid". Recl. Trav. Chim. Pays-Bas . 60 (10): 742–746. doi:10.1002/recl.19410601005.
  5. 1 2 Wenk, Hans-Rudolf; Bulakh, Andrey (2016). "Organic Minerals". Minerals – Their Constitution and Origin (2nd ed.). Cambridge University Press. ISBN   9781316423684.
  6. 1 2 Pampaloni, Guido (2010). "Aromatic hydrocarbons as ligands. Recent advances in the synthesis, the reactivity and the applications of bis(η6-arene) complexes". Coord. Chem. Rev. 254 (5–6): 402–419. doi:10.1016/j.ccr.2009.05.014.
  7. 1 2 3 4 5 Kotz, John C. (1986). "The Electrochemistry of Transition Metal Organometallic Compounds". In Fry, Albert J.; Britton, Wayne E. (eds.). Topics in Organic Electrochemistry. Springer Science & Business Media. pp. 83–176. ISBN   9781489920348.
  8. 1 2 3 4 5 Huttner, Gottfried; Lange, Siegfried; Fischer, Ernst O. (1971). "Molecular Structure of Bis(Hexamethylbenzene)Ruthenium(0)". Angew. Chem. Int. Ed. Engl. 10 (8): 556–557. doi:10.1002/anie.197105561.
  9. 1 2 3 4 5 6 7 8 Malischewski, Moritz; Seppelt, Konrad (2017). "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.
  10. 1 2 3 4 Boyle, Rebecca (14 January 2017). "Carbon seen bonding with six other atoms for the first time". New Scientist (3108). Archived from the original on 16 January 2017. Retrieved 14 January 2017.
  11. 1 2 3 4 Ritter, Stephen K. (19 December 2016). "Six bonds to carbon: Confirmed". Chem. Eng. News . 94 (49): 13. doi:10.1021/cen-09449-scicon007. Archived from the original on 9 January 2017.
  12. 1 2 3 4 Hamers, Laurel (24 December 2016). "Carbon can exceed four-bond limit". Science News . 190 (13): 17. Archived from the original on 3 February 2017.
  13. 1 2 3 4 Bachrach, Steven M. (17 January 2017). "A six-coordinate carbon atom". comporgchem.com. Archived from the original on 19 January 2017. Retrieved 18 January 2017.
  14. 1 2 3 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.
  15. 1 2 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.
  16. 1 2 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.
  17. 1 2 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.
  18. Favre, Henri A.; Powell, Warren H. (2013). Nomenclature of Organic Chemistry. IUPAC Recommendations and Preferred Name 2013. Royal Society of Chemistry. ISBN   9780854041824.
  19. 1 2 3 4 Haynes, William M., ed. (2016). CRC Handbook of Chemistry and Physics (93rd ed.). CRC Press. p. 3-296. ISBN   9781439880500.
  20. 1 2 3 Cullinane, N. M.; Chard, S. J.; Dawkins, C. W. C. (1955). "Hexamethylbenzene". Organic Syntheses . 35: 73. doi:10.15227/orgsyn.035.0073 ; Collected Volumes, vol. 4, p. 520.
  21. μέλι  in Liddell, Henry George ; Scott, Robert (1940) A Greek–English Lexicon, revised and augmented throughout by Jones, Sir Henry Stuart , with the assistance of McKenzie, Roderick. Oxford: Clarendon Press. In the Perseus Digital Library , Tufts University..
  22. Liebig, Justus (1844). "Lectures on organic chemistry: delivered during the winter session, 1844, in the University of Giessen". The Lancet. 2 (1106): 190–192. doi:10.1016/s0140-6736(02)64759-2.
  23. Lonsdale, Kathleen (1928). "The Structure of the Benzene Ring". Nature . 122 (810): 810. Bibcode:1928Natur.122..810L. doi: 10.1038/122810c0 . S2CID   4105837.
  24. Lonsdale, Kathleen (1948). Crystals and X-Rays. George Bell & Sons.
  25. 1 2 Briner, E.; Plüss, W.; Paillard, H. (1924). "Recherches sur la déshydration catalytique des systèmes phénols-alcools" [Research on the catalytic dehydration of phenol-alcohol systems]. Helv. Chim. Acta (in French). 7 (1): 1046–1056. doi:10.1002/hlca.192400701132.
  26. 1 2 3 4 Landis, Phillip S.; Haag, Werner O. (1963). "Formation of Hexamethylbenzene from Phenol and Methanol". J. Org. Chem. 28 (2): 585. doi:10.1021/jo01037a517.
  27. 1 2 3 Krysin, A. P.; Koptyug, V. A. (1969). "Reaction of phenols with alcohols on aluminum oxide II. The mechanism of hexamethylbenzene formation from phenol and methyl alcohol". Russ. Chem. Bull. 18 (7): 1479–1482. doi:10.1007/BF00908756.
  28. 1 2 Ipatiew, W.; Petrow, A. D. (1926). "Über die katalytische Kondensation von Aceton bei hohen Temperaturen und Drucken. (I. Mitteilung)" [On the catalytic condensation of acetone at high temperatures and pressures. (I. Communication)]. Ber. Dtsch. Chem. Ges. A/B (in German). 59 (8): 2035–2038. doi:10.1002/cber.19260590859.
  29. 1 2 Kotanigawa, Takeshi; Yamamoto, Mitsuyoshi; Shimokawa, Katsuyoshi; Yoshida, Yuji (1971). "Methylation of Phenol over Metallic Oxides". Bulletin of the Chemical Society of Japan. 44 (7): 1961–1964. doi: 10.1246/bcsj.44.1961 .
  30. 1 2 Kotanigawa, Takeshi (1974). "Mechanisms for the Reaction of Phenol with Methanol over the ZnO–Fe2O3 Catalyst". Bull. Chem. Soc. Jpn. 47 (4): 950–953. doi: 10.1246/bcsj.47.950 .
  31. 1 2 3 Cullinane, N. M.; Chard, S. J. (1945). "215. The action of methanol on phenol in the presence of alumina. Formation of anisole, methylated phenols, and hexamethylbenzene". J. Chem. Soc. : 821–823. doi:10.1039/JR9450000821. PMID   21008356.
  32. 1 2 3 Shubin, V. G.; Chzhu, V. P.; Korobeinicheva, I. K.; Rezvukhin, A. I.; Koptyug, V. A. (1970). "UV, IR, AND PMR spectra of hydroxyhexamethylbenzenonium ions". Russ. Chem. Bull. 19 (8): 1643–1648. doi:10.1007/BF00996497.
  33. 1 2 Weber, S. R.; Brintzinger, H. H. (1977). "Reactions of Bis(hexamethylbenzene)iron(0) with Carbon Monoxide and with Unsaturated Hydrocarbons". J. Organomet. Chem. 127 (1): 45–54. doi:10.1016/S0022-328X(00)84196-0. hdl: 2027.42/22975 .
  34. 1 2 3 Zeiss, H. H.; Herwig, W. (1958). "Acetylenic π-complexes of chromium in organic synthesis". J. Am. Chem. Soc. 80 (11): 2913. doi:10.1021/ja01544a091.
  35. 1 2 3 Franzus, B.; Canterino, P. J.; Wickliffe, R. A. (1959). "Titanium tetrachloride–trialkylaluminum complex—A cyclizing catalyst for acetylenic compounds". J. Am. Chem. Soc. 81 (6): 1514. doi:10.1021/ja01515a061.
  36. Le Bel, Joseph Achille; Greene, William H. (1880). "On the decomposition of alcohols, etc., by zinc chloride at high temperatures". American Chemical Journal. 2: 20–26.
  37. 1 2 3 Chang, Clarence D. (1983). "Hydrocarbons from Methanol". Catal. Rev. - Sci. Eng. 25 (1): 1–118. doi:10.1080/01614948308078874.
  38. Olah, George A.; Doggweiler, Hans; Felberg, Jeff D.; Frohlich, Stephan; Grdina, Mary Jo; Karpeles, Richard; Keumi, Takashi; Inaba, Shin-ichi; Ip, Wai M.; Lammertsma, Koop; Salem, George; Tabor, Derrick (1984). "Onium Ylide chemistry. 1. Bifunctional acid-base-catalyzed conversion of heterosubstituted methanes into ethylene and derived hydrocarbons. The onium ylide mechanism of the C1→C2 conversion". J. Am. Chem. Soc. 106 (7): 2143–2149. doi:10.1021/ja00319a039.
  39. Smith, Lee Irvin (1930). "Durene". Organic Syntheses . 10: 32. doi:10.15227/orgsyn.010.0032 ; Collected Volumes, vol. 2, p. 248.
  40. Reckleben, Hans; Scheiber, Johannes (1913). "Über eine einfache Darstellung des Hexamethyl-benzols" [A simple representation of hexamethylbenzene]. Ber. Dtsch. Chem. Ges. (in German). 46 (2): 2363–2365. doi:10.1002/cber.191304602168.
  41. Ipatiew, W. N.; Petrow, A. D. (1927). "Über die katalytische Kondensation des Acetons bei hohen Temperaturen und Drucken (II. Mitteilung)" [On the catalytic condensation of acetone at high temperatures and pressures (II. Communication)]. Ber. Dtsch. Chem. Ges. A/B (in German). 60 (3): 753–755. doi:10.1002/cber.19270600328.
  42. Kotanigawa, Takeshi; Shimokawa, Katsuyoshi (1974). "The Alkylation of Phenol over the ZnO–Fe2O3 Catalyst". Bull. Chem. Soc. Jpn. 47 (6): 1535–1536. doi: 10.1246/bcsj.47.1535 .
  43. Kotanigawa, Takeshi (1974). "The Methylation of Phenol and the Decomposition of Methanol on ZnO–Fe2O3 Catalyst". Bull. Chem. Soc. Jpn. 47 (10): 2466–2468. doi: 10.1246/bcsj.47.2466 .
  44. Ross, Sidney D.; Bassin, Morton; Finkelstein, Manuel; Leach, William A. (1954). "Molecular Compounds. I. Picryl Chloride-Hexamethylbenzene in Chloroform Solution". J. Am. Chem. Soc. 76 (1): 69–74. doi:10.1021/ja01630a018.
  45. Hart, Harold; Lange, Richard M.; Collins, Peter M. (1968). "2,3,4,5,6,6-Hexamethyl-2,4-cyclohexadien-1-one". Organic Syntheses . 48: 87. doi:10.15227/orgsyn.048.0087 ; Collected Volumes, vol. 5, p. 598.
  46. Saunders, Martin; Jiménez-Vázquez, Hugo A.; Khong, Anthony (1996). "NMR of 3He Dissolved in Organic Solids". J. Phys. Chem. 100 (39): 15968–15971. doi:10.1021/jp9617783.
  47. Earhart, H. W.; Komin, Andrew P. (2000). "Polymethylbenzenes". Kirk-Othmer Encyclopedia of Chemical Technology. New York: John Wiley. doi:10.1002/0471238961.1615122505011808.a01. ISBN   9780471238966.
  48. 1 2 3 4 Fischer, Ernst Otto; Lindner, Hans Hasso (1964). "Über Aromatenkomplexe von Metallen. LXXVI. Di-hexamethylbenzol-metall-π-komplexe des ein- und zweiwertigen Kobalts und Rhodiums" [About Aromatic Complexes of Metals. LXXVI. Di-hexamethylbenzene metal-π-complexes of mono- and bivalent cobalt and rhodium]. J. Organomet. Chem. (in German). 1 (4): 307–317. doi:10.1016/S0022-328X(00)80056-X.
  49. Fischer, Ernst Otto; Schmidt, Manfred W. (1966). "Über Aromatenkomplexe von Metallen, XCI. Über monomeres und dimeres Bis-hexamethylbenzol-rhenium". Chem. Ber. 99 (7): 2206–2212. doi:10.1002/cber.19660990719.
  50. Bennett, Martin A.; Huang, T.-N.; Matheson, T. W.; Smith, A. K. (1982). 16. (η6-Hexamethylbenzene)Ruthenium Complexes. Vol. 21. pp. 74–78. doi:10.1002/9780470132524.ch16. ISBN   9780470132524.{{cite book}}: |journal= ignored (help)
  51. Schäfer, W.; Hellmann, H. (1967). "Hexamethyl(Dewar Benzene) (Hexamethylbicyclo[2.2.0]hexa-2,5-diene)". Angew. Chem. Int. Ed. Engl. 6 (6): 518–525. doi:10.1002/anie.196705181.
  52. 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.
  53. Laube, Thomas; Lohse, Christian (1994). "X-ray Crystal Structures of Two (deloc-2,3,5)-1,2,3,4,5,6- Hexamethylbicyclo[2.1.1]hex-2-en-5-ylium Ions". J. Am. Chem. Soc. 116 (20): 9001–9008. doi:10.1021/ja00099a018.
  54. Jašík, Juraj; Gerlich, Dieter; Roithová, Jana (2014). "Probing Isomers of the Benzene Dication in a Low-Temperature Trap". J. Am. Chem. Soc. 136 (8): 2960–2962. doi:10.1021/ja412109h. PMID   24528384.
  55. Puiu, Tibi (5 January 2017). "Exotic carbon molecule has six bonds, breaking the four-bond limit". zmescience.com. ZME Science. Archived from the original on 16 January 2017. Retrieved 14 January 2017.
  56. 1 2 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.
  57. Elschenbroich, Christoph (2006). "Organometallic Compounds of Groups 2 and 12". Organometallics (3rd ed.). John Wiley & Sons. pp. 59–85. ISBN   9783527805143.