1-Methylcyclohexene

Last updated
1-Methylcyclohexene
1-methylcyclohexene.svg
Names
IUPAC name
1-Methylcyclohexene
Systematic IUPAC name
1-Methylcyclohexene
Other names
  • 2,3,4,5-Tetrahydrotoluene
  • 1-Methyl-1-cyclohexene
Identifiers
3D model (JSmol)
1304483
ChEBI
ChemSpider
ECHA InfoCard 100.008.836 OOjs UI icon edit-ltr-progressive.svg
EC Number
  • 209-718-0
PubChem CID
UNII
UN number 3295
  • InChI=1S/C7H12/c1-7-5-3-2-4-6-7/h5H,2-4,6H2,1H3
    Key: CTMHWPIWNRWQEG-UHFFFAOYSA-N
  • CC1=CCCCC1
Properties
C7H12
Molar mass 96.173 g·mol−1
Appearancecolorless liquid
Density 0.811 g/mL at 20 °C
Melting point −120.4 °C (−184.7 °F; 152.8 K)
Boiling point 110 °C (230 °F; 383 K)
0.052 g/kg for 1-methylcyclohexene
1.44
Hazards
GHS labelling:
GHS-pictogram-flamme.svg GHS-pictogram-exclam.svg GHS-pictogram-silhouette.svg
Warning
H225, H304, H315, H319, H335
P210, P233, P240, P241, P242, P243, P261, P264, P271, P280, P301+P310, P302+P352, P303+P361+P353, P304+P340, P305+P351+P338, P312, P321, P331, P332+P313, P337+P313, P362, P370+P378, P403+P233, P403+P235, P405, P501
Flash point −3 °C (27 °F; 270 K)
Safety data sheet (SDS) MSDS (1-methylcyclohexene)
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).

1-Methylcyclohexene an organic compound consisting of cyclohexene with a methyl group substituent attached to the alkene group. Two other structural isomers are known: 3-methylcyclohexene and 4-methylcyclohexene. All are colorless volatile liquids. They are specialized reagents. Methylcyclohexenes are a cyclic olefins.

Contents

Synthesis and reactions

Methylcyclohexenes are formed by the partial hydrogenation of toluene to methylcyclohexane over ruthenium catalyst. [1]

Ozonolysis of 1-methylcyclohexene results in ring-opening. [2]

Stereochemical probe

1-Methylcyclohexene is used as a probe of the stereochemistry of reactions involving alkenes because it is prochiral and the two sp2-carbon atoms differ.

"Hydrosilylation of Cyclohexene" Hydrosilylation of methylcyclohexene.jpg
"Hydrosilylation of Cyclohexene"

The regioselectivity and stereoselectivity of hydrosilylation of 1-methylcyclohexene with chloro(methyl)silanes depends on the number of chlorine atoms in the hydrosilylating agent. [3] Using chlorodimethylsilane produces a mixture of seven different products including cis- and trans-isomers of 2-, 3-, 4-chlorodimethyl(methylcyclohexyl)silanes and chlorodimethyl(cyclohexylmethyl)silane. The poor selectivity is due to the migration of the double bond in the cyclohexene ring. Reaction with dichloromethylsilane is more regioselective and stereoselective, only giving three of the seven products obtained from monochlorodimethylsilane. With trichloromethylsilane, trichlorocyclohexylmethylsilane is the only possible product and is obtained at 60 percent yield. All these products can be further reacted with Grignard reagents such as ethynylmagnesium bromide to synthesize ethynyl derivatives.

Oxidation of 1-methylcyclohexene catalyzed by cytochrome P450 yields a 2:1 mixture of hydroxylation to epoxidation products. [4]

The stereochemistry of hydroformylation has been examined using 1-methylcyclohexene. The main product has the formyl group on the less substituted alkene-carbon, trans with respect to the methyl substituent. [5]

Related Research Articles

<span class="mw-page-title-main">Alkene</span> Hydrocarbon compound containing one or more C=C bonds

In organic chemistry, an alkene, or olefin, is a hydrocarbon containing a carbon–carbon double bond. The double bond may be internal or in the terminal position. Terminal alkenes are also known as α-olefins.

In chemistry, a structural isomer of a compound is another compound whose molecule has the same number of atoms of each element, but with logically distinct bonds between them. The term metamer was formerly used for the same concept.

In organic chemistry, Markovnikov's rule or Markownikoff's rule describes the outcome of some addition reactions. The rule was formulated by Russian chemist Vladimir Markovnikov in 1870.

In organic chemistry, the oxymercuration reaction is an electrophilic addition reaction that transforms an alkene into a neutral alcohol. In oxymercuration, the alkene reacts with mercuric acetate in aqueous solution to yield the addition of an acetoxymercury group and a hydroxy group across the double bond. Carbocations are not formed in this process and thus rearrangements are not observed. The reaction follows Markovnikov's rule and it is an anti addition.

Hydroboration–oxidation reaction is a two-step hydration reaction that converts an alkene into an alcohol. The process results in the syn addition of a hydrogen and a hydroxyl group where the double bond had been. Hydroboration–oxidation is an anti-Markovnikov reaction, with the hydroxyl group attaching to the less-substituted carbon. The reaction thus provides a more stereospecific and complementary regiochemical alternative to other hydration reactions such as acid-catalyzed addition and the oxymercuration–reduction process. The reaction was first reported by Herbert C. Brown in the late 1950s and it was recognized in his receiving the Nobel Prize in Chemistry in 1979.

<span class="mw-page-title-main">Epoxide</span> Organic compounds with a carbon-carbon-oxygen ring

In organic chemistry, an epoxide is a cyclic ether, where the ether forms a three-atom ring: two atoms of carbon and one atom of oxygen. This triangular structure has substantial ring strain, making epoxides highly reactive, more so than other ethers. They are produced on a large scale for many applications. In general, low molecular weight epoxides are colourless and nonpolar, and often volatile.

In chemistry, an electrophile is a chemical species that forms bonds with nucleophiles by accepting an electron pair. Because electrophiles accept electrons, they are Lewis acids. Most electrophiles are positively charged, have an atom that carries a partial positive charge, or have an atom that does not have an octet of electrons.

<span class="mw-page-title-main">Regioselectivity</span> Preference of chemical bonding or breaking in one direction over others

In organic chemistry, regioselectivity is the preference of chemical bonding or breaking in one direction over all other possible directions. It can often apply to which of many possible positions a reagent will affect, such as which proton a strong base will abstract from an organic molecule, or where on a substituted benzene ring a further substituent will be added.

The 1,3-dipolar cycloaddition is a chemical reaction between a 1,3-dipole and a dipolarophile to form a five-membered ring. The earliest 1,3-dipolar cycloadditions were described in the late 19th century to the early 20th century, following the discovery of 1,3-dipoles. Mechanistic investigation and synthetic application were established in the 1960s, primarily through the work of Rolf Huisgen. Hence, the reaction is sometimes referred to as the Huisgen cycloaddition. 1,3-dipolar cycloaddition is an important route to the regio- and stereoselective synthesis of five-membered heterocycles and their ring-opened acyclic derivatives. The dipolarophile is typically an alkene or alkyne, but can be other pi systems. When the dipolarophile is an alkyne, aromatic rings are generally produced.

In chemistry, stereoselectivity is the property of a chemical reaction in which a single reactant forms an unequal mixture of stereoisomers during a non-stereospecific creation of a new stereocenter or during a non-stereospecific transformation of a pre-existing one. The selectivity arises from differences in steric and electronic effects in the mechanistic pathways leading to the different products. Stereoselectivity can vary in degree but it can never be total since the activation energy difference between the two pathways is finite: both products are at least possible and merely differ in amount. However, in favorable cases, the minor stereoisomer may not be detectable by the analytic methods used.

<span class="mw-page-title-main">Bamford–Stevens reaction</span> Synthesis of alkenes by base-catalysed decomposition of tosylhydrazones

The Bamford–Stevens reaction is a chemical reaction whereby treatment of tosylhydrazones with strong base gives alkenes. It is named for the British chemist William Randall Bamford and the Scottish chemist Thomas Stevens Stevens (1900–2000). The usage of aprotic solvents gives predominantly Z-alkenes, while protic solvent gives a mixture of E- and Z-alkenes. As an alkene-generating transformation, the Bamford–Stevens reaction has broad utility in synthetic methodology and complex molecule synthesis.

In organic chemistry, hydroboration refers to the addition of a hydrogen-boron bond to certain double and triple bonds involving carbon. This chemical reaction is useful in the organic synthesis of organic compounds.

The Shapiro reaction or tosylhydrazone decomposition is an organic reaction in which a ketone or aldehyde is converted to an alkene through an intermediate hydrazone in the presence of 2 equivalents of organolithium reagent. The reaction was discovered by Robert H. Shapiro in 1967. The Shapiro reaction was used in the Nicolaou Taxol total synthesis. This reaction is very similar to the Bamford–Stevens reaction, which also involves the basic decomposition of tosyl hydrazones.

Ring-closing metathesis (RCM) is a widely used variation of olefin metathesis in organic chemistry for the synthesis of various unsaturated rings via the intramolecular metathesis of two terminal alkenes, which forms the cycloalkene as the E- or Z- isomers and volatile ethylene.

<span class="mw-page-title-main">Asymmetric induction</span> Preferential formation of one chiral isomer over another in a chemical reaction

Asymmetric induction describes the preferential formation in a chemical reaction of one enantiomer or diastereoisomer over the other as a result of the influence of a chiral feature present in the substrate, reagent, catalyst or environment. Asymmetric induction is a key element in asymmetric synthesis.

The Nazarov cyclization reaction is a chemical reaction used in organic chemistry for the synthesis of cyclopentenones. The reaction is typically divided into classical and modern variants, depending on the reagents and substrates employed. It was originally discovered by Ivan Nikolaevich Nazarov (1906–1957) in 1941 while studying the rearrangements of allyl vinyl ketones.

<span class="mw-page-title-main">Allylic strain</span> Type of strain energy in organic chemistry

Allylic strain in organic chemistry is a type of strain energy resulting from the interaction between a substituent on one end of an olefin with an allylic substituent on the other end. If the substituents are large enough in size, they can sterically interfere with each other such that one conformer is greatly favored over the other. Allylic strain was first recognized in the literature in 1965 by Johnson and Malhotra. The authors were investigating cyclohexane conformations including endocyclic and exocylic double bonds when they noticed certain conformations were disfavored due to the geometry constraints caused by the double bond. Organic chemists capitalize on the rigidity resulting from allylic strain for use in asymmetric reactions.

In organic chemistry, enone–alkene cycloadditions are a version of the [2+2] cycloaddition. This reaction involves an enone and alkene as substrates. Although the concerted photochemical [2+2] cycloaddition is allowed, the reaction between enones and alkenes is stepwise and involves discrete diradical intermediates.

The Evelyn effect is defined as the phenomena in which the product ratios in a chemical reaction change as the reaction proceeds. This phenomenon contradicts the fundamental principle in organic chemistry by reactions always go by the lowest energy pathway. The favored product should remain so throughout a reaction run at constant conditions. However, the ratio of alkenes before the synthesis is complete shows that the favored product to is not the favored product. The basic idea here is that the proportions of the various alkene products changes as a function of time with a change in mechanism.

<span class="mw-page-title-main">4-Methylcyclohexene</span> Chemical compound

4-Methylcyclohexene is an organic compound consisting of cyclohexene with a methyl group substituent attached to carbon most distant from the alkene group. Two other structural isomers are known: 1-methylcyclohexene and 3-methylcyclohexene. All are colorless volatile liquids classified as a cyclic olefins. They are specialized reagents.

References

  1. Belohlav, H.; Kluson, P.; Cerveny, L. (1997). "Partial Hydrogenation of Toluene Over A Ruthenium Catalyst, A Model Treatment of A Deactivation Process". Res. Chem. Intermed. 32 (2): 161–168. doi:10.1163/156856797X00312. S2CID   95532469.
  2. Atkinson, Roger; Tuazon, Ernesto C.; Aschmann, Sara M. (1995). "Products". Environ. Sci. Technol. (29). doi: 10.1029/98JD00524 .
  3. Voronkov, M.; et al. (December 2004). "Hydrosilylation of Cyclohexene, 1-Methylcyclohexene, and Isopropylidenecyclohexane". Russian Journal of General Chemistry. 74 (12): 1895–1899. doi:10.1007/s11176-005-0114-4. S2CID   98097289.
  4. Khan, M. M. T.; Rao, A. P.; Bhatt, S. D.; Merchant, R. R. (1990). "Epoxidation of cyclohexene, methylcyclohexene and cis-cyclooctene by molecular oxygen using ruthenium(III) aquo ion as catalyst: A kinetic study". Journal of Molecular Catalysis. 62 (3): 265–276. doi:10.1016/0304-5102(90)85222-4.
  5. Molnár, Árpád; Olah, George A.; Surya Prakash, G. K. (2017). "Carbonylation and Carboxylation". Hydrocarbon Chemistry. Wiley. pp. 509–568. doi:10.1002/9781119390541.ch7. ISBN   978-1-119-39051-0.