Nitroalkene

Last updated

A nitroalkene, or nitro olefin, is a functional group combining the functionality of its constituent parts, an alkene and nitro group, while displaying its own chemical properties through alkene activation, making the functional group useful in specialty reactions such as the Michael reaction or Diels-Alder additions. [1]

Contents

Synthesis

Nitroalkenes are synthesized by various means, notable examples include:

Furfural nitroaldol condensation.png
Nitrogenation of a phenylisopropene.png
Direct nitration of styrene with alumina catalyst.svg
Direct nitration of styrene using FeNO3 on a Clayfen support.png
Dehydration of 2-nitroethanol to nitroethylene via phthalic anhydride.svg

Reactions

Nitroalkenes are useful intermediates for various chemical functionalities.

Michael acceptor intermediate in Lycoricidine Synthesis.svg
Nitroalkene dienophile in cycloaddition with butadiene.svg
Barton-Zard reaction.svg
Pericyclic reaction of a nitroalkene yielding an indole.svg
Partial hydrogenation of a nitrostyrene to an alkene hydroxylamine.svg
Hydrogenation of a nitrostyrene to a primary amine.svg
Asymmetric Stetter Reaction with Nitroalkenes.png

Related Research Articles

<span class="mw-page-title-main">Diels–Alder reaction</span> Chemical reaction

In organic chemistry, the Diels–Alder reaction is a chemical reaction between a conjugated diene and a substituted alkene, commonly termed the dienophile, to form a substituted cyclohexene derivative. It is the prototypical example of a pericyclic reaction with a concerted mechanism. More specifically, it is classified as a thermally-allowed [4+2] cycloaddition with Woodward–Hoffmann symbol [π4s + π2s]. It was first described by Otto Diels and Kurt Alder in 1928. For the discovery of this reaction, they were awarded the Nobel Prize in Chemistry in 1950. Through the simultaneous construction of two new carbon–carbon bonds, the Diels–Alder reaction provides a reliable way to form six-membered rings with good control over the regio- and stereochemical outcomes. Consequently, it has served as a powerful and widely applied tool for the introduction of chemical complexity in the synthesis of natural products and new materials. The underlying concept has also been applied to π-systems involving heteroatoms, such as carbonyls and imines, which furnish the corresponding heterocycles; this variant is known as the hetero-Diels–Alder reaction. The reaction has also been generalized to other ring sizes, although none of these generalizations have matched the formation of six-membered rings in terms of scope or versatility. Because of the negative values of ΔH° and ΔS° for a typical Diels–Alder reaction, the microscopic reverse of a Diels–Alder reaction becomes favorable at high temperatures, although this is of synthetic importance for only a limited range of Diels-Alder adducts, generally with some special structural features; this reverse reaction is known as the retro-Diels–Alder reaction.

<span class="mw-page-title-main">Nitro compound</span> Organic compound containing an −NO₂ group

In organic chemistry, nitro compounds are organic compounds that contain one or more nitro functional groups. The nitro group is one of the most common explosophores used globally. The nitro group is also strongly electron-withdrawing. Because of this property, C−H bonds alpha (adjacent) to the nitro group can be acidic. For similar reasons, the presence of nitro groups in aromatic compounds retards electrophilic aromatic substitution but facilitates nucleophilic aromatic substitution. Nitro groups are rarely found in nature. They are almost invariably produced by nitration reactions starting with nitric acid.

<span class="mw-page-title-main">Michael addition reaction</span> Reaction in organic chemistry

In organic chemistry, the Michael reaction or Michael 1,4 addition is a reaction between a Michael donor and a Michael acceptor to produce a Michael adduct by creating a carbon-carbon bond at the acceptor's β-carbon. It belongs to the larger class of conjugate additions and is widely used for the mild formation of carbon-carbon bonds.

The Simmons–Smith reaction is an organic cheletropic reaction involving an organozinc carbenoid that reacts with an alkene to form a cyclopropane. It is named after Howard Ensign Simmons, Jr. and Ronald D. Smith. It uses a methylene free radical intermediate that is delivered to both carbons of the alkene simultaneously, therefore the configuration of the double bond is preserved in the product and the reaction is stereospecific.

<span class="mw-page-title-main">Olefin metathesis</span> Organic reaction involving the breakup and reassembly of alkene double bonds

In organic chemistry, olefin metathesis is an organic reaction that entails the redistribution of fragments of alkenes (olefins) by the scission and regeneration of carbon-carbon double bonds. Because of the relative simplicity of olefin metathesis, it often creates fewer undesired by-products and hazardous wastes than alternative organic reactions. For their elucidation of the reaction mechanism and their discovery of a variety of highly active catalysts, Yves Chauvin, Robert H. Grubbs, and Richard R. Schrock were collectively awarded the 2005 Nobel Prize in Chemistry.

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

The Henry reaction is a classic carbon–carbon bond formation reaction in organic chemistry. Discovered in 1895 by the Belgian chemist Louis Henry (1834–1913), it is the combination of a nitroalkane and an aldehyde or ketone in the presence of a base to form β-nitro alcohols. This type of reaction is also referred to as a nitroaldol reaction. It is nearly analogous to the aldol reaction that had been discovered 23 years prior that couples two carbonyl compounds to form β-hydroxy carbonyl compounds known as "aldols". The Henry reaction is a useful technique in the area of organic chemistry due to the synthetic utility of its corresponding products, as they can be easily converted to other useful synthetic intermediates. These conversions include subsequent dehydration to yield nitroalkenes, oxidation of the secondary alcohol to yield α-nitro ketones, or reduction of the nitro group to yield β-amino alcohols.

<span class="mw-page-title-main">Danishefsky's diene</span> Chemical compound

Danishefsky's diene is an organosilicon compound and a diene with the formal name trans-1-methoxy-3-trimethylsilyloxy-buta-1,3-diene named after Samuel J. Danishefsky. Because the diene is very electron-rich it is a very reactive reagent in Diels-Alder reactions. This diene reacts rapidly with electrophilic alkenes, such as maleic anhydride. The methoxy group promotes highly regioselective additions. The diene is known to react with amines, aldehydes, alkenes and alkynes. Reactions with imines and nitro-olefins have been reported.

The reduction of nitro compounds are chemical reactions of wide interest in organic chemistry. The conversion can be effected by many reagents. The nitro group was one of the first functional groups to be reduced. Alkyl and aryl nitro compounds behave differently. Most useful is the reduction of aryl nitro compounds.

Selenoxide elimination is a method for the chemical synthesis of alkenes from selenoxides. It is most commonly used to synthesize α,β-unsaturated carbonyl compounds from the corresponding saturated analogues. It is mechanistically related to the Cope reaction.

The Barton–Zard reaction is a route to pyrrole derivatives via the reaction of a nitroalkene with an α-isocyanide under basic conditions. It is named after Derek Barton and Samir Zard who first reported it in 1985.

<span class="mw-page-title-main">Reductions with samarium(II) iodide</span>

Reductions with samarium(II) iodide involve the conversion of various classes of organic compounds into reduced products through the action of samarium(II) iodide, a mild one-electron reducing agent.

The Saegusa–Ito oxidation is a chemical reaction used in organic chemistry. It was discovered in 1978 by Takeo Saegusa and Yoshihiko Ito as a method to introduce α-β unsaturation in carbonyl compounds. The reaction as originally reported involved formation of a silyl enol ether followed by treatment with palladium(II) acetate and benzoquinone to yield the corresponding enone. The original publication noted its utility for regeneration of unsaturation following 1,4-addition with nucleophiles such as organocuprates.

The inverse electron demand Diels–Alder reaction, or DAINV or IEDDA is an organic chemical reaction, in which two new chemical bonds and a six-membered ring are formed. It is related to the Diels–Alder reaction, but unlike the Diels–Alder reaction, the DAINV is a cycloaddition between an electron-rich dienophile and an electron-poor diene. During a DAINV reaction, three pi-bonds are broken, and two sigma bonds and one new pi-bond are formed. A prototypical DAINV reaction is shown on the right.

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

Photoredox catalysis is a branch of photochemistry that uses single-electron transfer. Photoredox catalysts are generally drawn from three classes of materials: transition-metal complexes, organic dyes, and semiconductors. While organic photoredox catalysts were dominant throughout the 1990s and early 2000s, soluble transition-metal complexes are more commonly used today.

<i>beta</i>-Nitrostyrene Chemical compound

β-Nitrostyrene is an aromatic compound and a nitroalkene used in the synthesis of indigo dye and the slimicide bromo-nitrostyrene.

The Mukaiyama hydration is an organic reaction involving formal addition of an equivalent of water across an olefin by the action of catalytic bis(acetylacetonato)cobalt(II) complex, phenylsilane and atmospheric oxygen to produce an alcohol with Markovnikov selectivity.

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

Diethyl oxomalonate is the diethyl ester of mesoxalic acid (ketomalonic acid), the simplest oxodicarboxylic acid and thus the first member (n = 0) of a homologous series HOOC–CO–(CH2)n–COOH with the higher homologues oxalacetic acid (n = 1), α-ketoglutaric acid (n = 2) and α-ketoadipic acid (n = 3) (the latter a metabolite of the amino acid lysine). Diethyl oxomalonate reacts because of its highly polarized keto group as electrophile in addition reactions and is a highly active reactant in pericyclic reactions such as the Diels-Alder reactions, cycloadditions or ene reactions. At humid air, mesoxalic acid diethyl ester reacts with water to give diethyl mesoxalate hydrate and the green-yellow oil are spontaneously converted to white crystals.

The Cadogan–Sundberg indole synthesis, or simply Cadogan indole synthesis, is a name reaction in organic chemistry that allows for the generation of indoles from o-nitrostyrenes with the use of trialkyl phosphites, such as triethyl phosphite.

<span class="mw-page-title-main">Nitro-Mannich reaction</span>

The nitro-Mannich reaction is the nucleophilic addition of a nitroalkane to an imine, resulting in the formation of a beta-nitroamine. With the reaction involving the addition of an acidic carbon nucleophile to a carbon-heteroatom double bond, the nitro-Mannich reaction is related to some of the most fundamental carbon-carbon bond forming reactions in organic chemistry, including the aldol reaction, Henry reaction and Mannich reaction.

References

  1. 1 2 3 4 Furniss, Brian; Hannaford, Antony; Smith, Peter & Tatchell, Austin (1996). Vogel's Textbook of Practical Organic Chemistry 5th Ed. London: Longman Science & Technical. pp.  635, 768, 1035–1036, & 1121. ISBN   9780582462366.
  2. Ballini, Roberto; Castagnani, Roberto; Petrini, Marino (1992). "Chemoselective synthesis of functionalized conjugated nitroalkenes". The Journal of Organic Chemistry. 57 (7): 2160–2162. doi:10.1021/jo00033a045.
  3. Worrall, David E. (1929). "Nitrostyrene". Org. Synth. 9: 66. doi:10.15227/orgsyn.009.0066.
  4. Chandrasekhar, S.; Shrinidhi, A. (2014). "Useful Extensions of the Henry Reaction: Expeditious Routes to Nitroalkanes and Nitroalkenes in Aqueous Media". Synthetic Communications. 44 (20): 3008–3018. doi:10.1080/00397911.2014.926373. S2CID   98439096.
  5. Waldman, Steve; Monte, Aaron, Monte; Bracey, Ann & Nichols, David (1996). "One-pot Claisen rearrangement/O-methylation/alkene isomerization in the synthesis of ortho-methoxylated phenylisopropylamines". Tetrahedron Letters. 37 (44): 7889–7892. doi:10.1016/0040-4039(96)01807-2.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  6. Mukaiyama, T.; Hata E. & Yamada, T. (1995). "Convenient and Simple Preparation of Nitroolefins Nitration of Olefins with Nitric Oxide". Chemistry Letters. 24 (7): 505–506. doi:10.1246/cl.1995.505.
  7. Varma, Rajender; Naicker, Kannan; Liesen, Per (1998). "Selective nitration of styrenes with clayfen and clayan: A solvent-free synthesis of β-nitrostyrenes". Tetrahedron Letters. 39 (23): 3977–3980. doi:10.1016/S0040-4039(98)00740-0.
  8. Ranganathan, Darshan; Rao, Bhushan; Ranganathan, Subramania; Mehrotra, Ashok & Iyengar, Radha (1980). "Nitroethylene: a stable, clean, and reactive agent for organic synthesis". The Journal of Organic Chemistry. 45 (7): 1185–1189. doi:10.1021/jo01295a003.
  9. Jubert, Carole & Knochel, Paul (1992). "Preparation of polyfunctional nitro olefins and nitroalkanes using the copper-zinc reagents RCu(CN)ZnI". The Journal of Organic Chemistry. 57 (20): 5431–5438. doi:10.1021/jo00046a027.
  10. Noboru Ono; Hideyoshi Miyake; Akio Kamimura & Aritsune, Kaji (1987). "Regioselective Diels–Alder reactions. The nitro group as a regiochemical control element". Perkin Transactions. 1: 1929–1935. doi:10.1039/P19870001929.
  11. Jie Jack Li (2013). Heterocyclic Chemistry in Drug Discovery. New York: Wiley. ISBN   9781118354421. pp.43-4
  12. Novellino, Luisa; d'Ischia, Marco & Prota, Giuseppe (1999). "Expedient Synthesis of 5,6-Dihydroxyindole and Derivatives via an Improved Zn(II)-Assisted 2,β-Dinitrostyrene Approach". Synthesis. 1999 (5): 793–796. doi:10.1055/s-1999-3469.
  13. 1 2 Masahiko Kohno; Shigehiro Sasao & Shun-Ichi Murahashi (1990). "Synthesis of Phenethylamines by Hydrogenation of β-Nitrostyrenes". Bulletin of the Chemical Society of Japan. 63 (4): 1252–1254. doi: 10.1246/bcsj.63.1252 .
  14. Koch, Werner & Reichert, Benno (1935). "Über die katalytische Hydrierung substituierter ω-Nitrostyrole". Archiv der Pharmazie. 273 (18–20): 265–274. doi:10.1002/ardp.19352731802. S2CID   95731916.
  15. DiRocco, D. A.; Oberg, K. M.; Dalton, D. M.; Rovis, T. (2009). "Catalytic Asymmetric Intermolecular Stetter Reaction of Heterocyclic Aldehydes with Nitroalkenes: Backbone Fluorination Improves Selectivity". Journal of the American Chemical Society. 131 (31): 10872–10874. doi:10.1021/ja904375q. PMC   2747345 . PMID   19722669.