Birch reduction

Last updated
Birch reduction
Named after Arthur Birch
Reaction type Organic redox reaction
Identifiers
Organic Chemistry Portal birch-reduction
RSC ontology ID RXNO:0000042

The Birch reduction is an organic reaction that is used to convert arenes to 1,4-cyclohexadienes. The reaction is named after the Australian chemist Arthur Birch and involves the organic reduction of aromatic rings in an amine solvent (traditionally liquid ammonia) with an alkali metal (traditionally sodium) and a proton source (traditionally an alcohol). Unlike catalytic hydrogenation, Birch reduction does not reduce the aromatic ring all the way to a cyclohexane.

Contents

The Birch reduction BirchReductionScheme.svg
The Birch reduction

An example is the reduction of naphthalene in ammonia and ethanol:

naphthalene Birch Reduction Naphthalene Birch Reduction.png
naphthalene Birch Reduction

Reaction mechanism and regioselectivity

A solution of sodium in liquid ammonia consists of the intensely blue electride salt [Na(NH3)x]+ e. The solvated electrons add to the aromatic ring to give a radical anion, which then abstracts a proton from the alcohol. The process then repeats at either the ortho or para position (depending on substituents) to give the final diene. [1] The residual double bonds do not stabilize further radical additions. [2] [3]

Birch reduction of benzene, also available in animated form. BirchReactionMechanismmk2.tif
Birch reduction of benzene, also available in animated form.
Birch reduction of anisole. Birch Basic-Mech.svg
Birch reduction of anisole.

The reaction is known to be third order – first order in the aromatic, first order in the alkali metal, and first order in the alcohol. [4] This requires that the rate-limiting step be the conversion of radical anion B to the cyclohexadienyl radical C.

That step also determines the structure of the product. Although Arthur Birch originally argued that the protonation occurred at the meta position, [5] subsequent investigation has revealed that protonation occurs at either the ortho or para position. Electron donors tend to induce ortho protonation, as shown in the reduction of anisole (1). Electron-withdrawing substituents tend to induce para protonation, as shown in the reduction of benzoic acid (2). [6]

Birch-Anisole.svg


Birch-Benzoic.svg

Solvated electrons will preferentially reduce sufficiently electronegative functional groups, such as ketones or nitro groups, but do not attack alcohols, carboxylic acids, or ethers. [6]

Secondary protonation regioselectivity

The second reduction and protonation also poses mechanistic questions. Thus there are three resonance structures for the carbanion (labeled B, C and D in the picture).

Cyclohexadienyl Anion1.svg

Simple Hückel computations lead to equal electron densities at the three atoms 1, 3 and 5, but asymmetric bond orders. Modifying the exchange integrals to account for varying interatomic distances, produces maximum electron density at the central atom 1, [7] [8] [9] a result confirmed by more modern RHF computations. [10]

ApproximationDensity Atom 3Density Atom 2Density Atom 1Bond Order 2–3Bond Order 1–2
Hückel (1st approx)0.3330.000.3330.7880.578
2nd approx0.3170.000.3650.8020.564
3rd approx0.3160.000.3680.8020.562

The result is analogous to conjugated enolates. When those anions (but not the enol tautomer) kinetically protonate, they do so at the center to afford the β,γ-unsaturated carbonyl. [7] [11]

Modifications

Traditional Birch reduction requires cryogenic temperatures to liquify ammonia and pyrophoric alkali-metal electron donors. Variants have developed to reduce either inconvenience.

Many amines serve as alternative solvents: for example, THF [12] [13] or mixed n-propylamine and ethylenediamine. [14]

To avoid direct alkali, there are chemical alternatives, such as M-SG reducing agent. The reduction can also be powered by an external potential or sacrificial anode (magnesium or aluminum), but then alkali metal salts are necessary to colocate the reactants via complexation. [15]

Birch alkylation

In Birch alkylation the anion formed in the Birch reduction is trapped by a suitable electrophile such as a haloalkane, for example: [16]

Birch Alkylation Org Synth 1990 BirchAlkylationOrgSynth1990.svg
Birch Alkylation Org Synth 1990

In substituted aromatics, an electron-withdrawing substituent, such as a carboxylic acid, will stabilize the carbanion to generate the least-substituted olefin; [17] an electron-donating substituent has the opposite effect. [18]

Adding 1,4-dibromobutane to a Birch reduction of tert-butyl benzoate forms the 1,1-cyclohexadiene product. BirchAlkylation.png
Adding 1,4-dibromobutane to a Birch reduction of tert-butyl benzoate forms the 1,1-cyclohexadiene product.

Benkeser reduction

The Benkeser reduction is the hydrogenation of polycyclic aromatic hydrocarbons, especially naphthalenes using lithium or calcium metal in low molecular weight alkyl amines solvents. Unlike traditional Birch reduction, the reaction can be conducted at temperatures higher than the boiling point of ammonia (−33 °C). [20] [21]

For the reduction of naphthalene with lithium in a mixed ethylamine-dimethylamine solution, the principal products are bicyclo[3.3.0]dec-(1,9)-ene, bicyclo[3.3.0]dec-(1,2)-ene and bicyclo[3.3.0]decane. [22] [23]

The Benkeser reaction Benkeser Reduction(corrected).png
The Benkeser reaction
Modified Benkeser reduction BirchBenkeser small.png
Modified Benkeser reduction

The directing effects of naphthalene substituents remain relatively unstudied theoretically. Substituents adjacent to the bridge appear to direct reduction to the unsubstituted ring; β substituents (one bond further) tend to direct reduction to the substituted ring. [6]

History

Arthur Birch, building on earlier work by Wooster and Godfrey, [24] developed the reaction while working in the Dyson Perrins Laboratory at the University of Oxford. [25] Birch's original procedure used sodium and ethanol; [5] [26] [27] Alfred L. Wilds later discovered that lithium gives better yields. [28] [29]

The reaction was difficult to understand mechanistically, with controversy lasting into the 1990s.

The case with electron-withdrawing groups is obvious, because the Birch alkylation serves as a trap for the penultimate dianion D. This dianion appears even in alcohol-free reactions. Thus the initial protonation is para rather than ipso, as seen in the B-C transformation. [30] [31] [32]

Benzoic acid reduction, including possible alkylation Birch-Benzoic Mech.svg
Benzoic acid reduction, including possible alkylation

For electron-donating substituents, Birch initially proposed meta attack, corresponding to the location of greatest electron density in a neutral benzene ring, a position endorsed by Krapcho and Bothner-By. [4] [33] These conclusions were challenged by Zimmerman in 1961, who computed electron densities of the radical and diene anions, revealing that the ortho site which was most negative and thus most likely to protonate. [7] [9] But the situation remained uncertain, because computations remained highly sensitive to transition geometry. Worse, Hückel orbital and unrestricted Hartree-Fock computations gave conflicting answers. Burnham, in 1969, concluded that the trustworthiest computations supported meta attack; [34] Birch and Radom, in 1980, concluded that both ortho and meta substitutions would occur with a slight preference for ortho. [35]

In the earlier 1990s, Zimmerman and Wang developed an experiment technique to distinguish between ortho and meta protonation. The method began with the premise that carbanions are much more basic than the corresponding radical anions and thus protonate less selectively. Correspondingly, the two protonations in Birch reduction should exhibit an isotope effect: in a protium–deuterium medium, the radical anion should preferentially protonate and the carbanion deuterate. Indeed, a variety of methoxylated aromatics exhibited less ortho deuterium than meta (a 1:7 ratio). Moreover, modern electron density computations now firmly indicated ortho protonation; frontier orbital densities, most analogous to the traditional computations used in past studies, did not. [10]

Although Birch remained reluctant to concede that ortho protonation was preferred as late as 1996, [36] Zimmerman and Wang had won the day: modern textbooks unequivocally agree that electron-donating substituents promote ortho attack. [6]

Additional reading

See also

Related Research Articles

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

Aromatic compounds, also known as "mono- and polycyclic aromatic hydrocarbons", are organic compounds containing one or more aromatic rings. The word "aromatic" originates from the past grouping of molecules based on smell, before their general chemical properties were understood. The current definition of aromatic compounds does not have any relation with their smell.

In organic chemistry, a methyl group is an alkyl derived from methane, containing one carbon atom bonded to three hydrogen atoms, having chemical formula CH3. In formulas, the group is often abbreviated as Me. This hydrocarbon group occurs in many organic compounds. It is a very stable group in most molecules. While the methyl group is usually part of a larger molecule, bounded to the rest of the molecule by a single covalent bond, it can be found on its own in any of three forms: methanide anion, methylium cation or methyl radical. The anion has eight valence electrons, the radical seven and the cation six. All three forms are highly reactive and rarely observed.

Pyrrole is a heterocyclic, aromatic, organic compound, a five-membered ring with the formula C4H4NH. It is a colorless volatile liquid that darkens readily upon exposure to air. Substituted derivatives are also called pyrroles, e.g., N-methylpyrrole, C4H4NCH3. Porphobilinogen, a trisubstituted pyrrole, is the biosynthetic precursor to many natural products such as heme.

The Friedel–Crafts reactions are a set of reactions developed by Charles Friedel and James Crafts in 1877 to attach substituents to an aromatic ring. Friedel–Crafts reactions are of two main types: alkylation reactions and acylation reactions. Both proceed by electrophilic aromatic substitution.

In organic chemistry, a carbanion is an anion in which carbon is negatively charged.

<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.

<span class="mw-page-title-main">Phosphonium</span> Family of polyatomic cations containing phosphorus

In chemistry, the term phosphonium describes polyatomic cations with the chemical formula PR+
4
. These cations have tetrahedral structures. The salts are generally colorless or take the color of the anions.

<span class="mw-page-title-main">Radical anion</span> Free radical species

In organic chemistry, a radical anion is a free radical species that carries a negative charge. Radical anions are encountered in organic chemistry as reduced derivatives of polycyclic aromatic compounds, e.g. sodium naphthenide. An example of a non-carbon radical anion is the superoxide anion, formed by transfer of one electron to an oxygen molecule. Radical anions are typically indicated by .

A solvated electron is a free electron in a solution, and is the smallest possible anion. Solvated electrons occur widely. Often, discussions of solvated electrons focus on their solutions in ammonia, which are stable for days, but solvated electrons also occur in water and other solvents – in fact, in any solvent that mediates outer-sphere electron transfer. The solvated electron is responsible for a great deal of radiation chemistry.

<span class="mw-page-title-main">Directed ortho metalation</span> Chemical reaction

Directed ortho metalation (DoM) is an adaptation of electrophilic aromatic substitution in which electrophiles attach themselves exclusively to the ortho- position of a direct metalation group or DMG through the intermediary of an aryllithium compound. The DMG interacts with lithium through a hetero atom. Examples of DMG's are the methoxy group, a tertiary amine group and an amide group.The compound can be produced by directed lithiation of anisole.

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

The Stevens rearrangement in organic chemistry is an organic reaction converting quaternary ammonium salts and sulfonium salts to the corresponding amines or sulfides in presence of a strong base in a 1,2-rearrangement.

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

Borohydride refers to the anion [BH4], which is also called tetrahydroborate, and its salts. Borohydride or hydroborate is also the term used for compounds containing [BH4−nXn], where n is an integer from 0 to 3, for example cyanoborohydride or cyanotrihydroborate [BH3(CN)] and triethylborohydride or triethylhydroborate [BH(CH2CH3)3]. Borohydrides find wide use as reducing agents in organic synthesis. The most important borohydrides are lithium borohydride and sodium borohydride, but other salts are well known. Tetrahydroborates are also of academic and industrial interest in inorganic chemistry.

Radical-nucleophilic aromatic substitution or SRN1 in organic chemistry is a type of substitution reaction in which a certain substituent on an aromatic compound is replaced by a nucleophile through an intermediary free radical species:

<span class="mw-page-title-main">Smiles rearrangement</span> Organic reaction

In organic chemistry, the Smiles rearrangement is an organic reaction and a rearrangement reaction named after British chemist Samuel Smiles. It is an intramolecular, nucleophilic aromatic substitution of the type:

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

Cyclododecahexaene or [12]annulene is a member of the series of annulenes with some interest in organic chemistry with regard to the study of aromaticity. Cyclododecahexaene is non-aromatic due to the lack of planarity of the structure. On the other hand the dianion with 14 electrons is a Hückel aromat and more stable.

Nitrile anions is jargon from the organic product resulting from the deprotonation of alkylnitriles. The proton(s) α to the nitrile group are sufficiently acidic that they undergo deprotonation by strong bases, usually lithium-derived. The products are not anions but covalent organolithium complexes. Regardless, these organolithium compounds are reactive toward various electrophiles.

Desulfonylation reactions are chemical reactions leading to the removal of a sulfonyl group from organic compounds. As the sulfonyl functional group is electron-withdrawing, methods for cleaving the sulfur–carbon bonds of sulfones are typically reductive in nature. Olefination or replacement with hydrogen may be accomplished using reductive desulfonylation methods.

Benzylic activation and stereocontrol in tricarbonyl(arene)chromium complexes refers to the enhanced rates and stereoselectivities of reactions at the benzylic position of aromatic rings complexed to chromium(0) relative to uncomplexed arenes. Complexation of an aromatic ring to chromium stabilizes both anions and cations at the benzylic position and provides a steric blocking element for diastereoselective functionalization of the benzylic position. A large number of stereoselective methods for benzylic and homobenzylic functionalization have been developed based on this property.

Electrophilic aromatic substitution is an organic reaction in which an atom that is attached to an aromatic system is replaced by an electrophile. Some of the most important electrophilic aromatic substitutions are aromatic nitration, aromatic halogenation, aromatic sulfonation, alkylation and acylation Friedel–Crafts reaction.

Metal-catalyzed C–H borylation reactions are transition metal catalyzed organic reactions that produce an organoboron compound through functionalization of aliphatic and aromatic C–H bonds and are therefore useful reactions for carbon–hydrogen bond activation. Metal-catalyzed C–H borylation reactions utilize transition metals to directly convert a C–H bond into a C–B bond. This route can be advantageous compared to traditional borylation reactions by making use of cheap and abundant hydrocarbon starting material, limiting prefunctionalized organic compounds, reducing toxic byproducts, and streamlining the synthesis of biologically important molecules. Boronic acids, and boronic esters are common boryl groups incorporated into organic molecules through borylation reactions. Boronic acids are trivalent boron-containing organic compounds that possess one alkyl substituent and two hydroxyl groups. Similarly, boronic esters possess one alkyl substituent and two ester groups. Boronic acids and esters are classified depending on the type of carbon group (R) directly bonded to boron, for example alkyl-, alkenyl-, alkynyl-, and aryl-boronic esters. The most common type of starting materials that incorporate boronic esters into organic compounds for transition metal catalyzed borylation reactions have the general formula (RO)2B-B(OR)2. For example, bis(pinacolato)diboron (B2Pin2), and bis(catecholato)diborane (B2Cat2) are common boron sources of this general formula.

References

  1. March, Jerry (1985), Advanced Organic Chemistry: Reactions, Mechanisms, and Structure, 3rd edition, New York: Wiley, ISBN   9780471854722, OCLC   642506595
  2. Rabideau, P. W.; Marcinow, Z. (1992). "The Birch Reduction of Aromatic Compounds". Org. React. (review). 42: 1–334. doi:10.1002/0471264180.or042.01. ISBN   0471264180.
  3. Mander, L. N. (1991). "Partial Reduction of Aromatic Rings by Dissolving Metals and by Other Methods". Compr. Org. Synth. (review). 8: 489–521. doi:10.1016/B978-0-08-052349-1.00237-7. ISBN   978-0-08-052349-1.
  4. 1 2 Krapcho, A. P.; Bothner-By, A. A. (1959). "Kinetics of the Metal-Ammonia-Alcohol Reductions of Benzene and Substituted Benzenes1". J. Am. Chem. Soc. 81 (14): 3658–3666. doi:10.1021/ja01523a042.
  5. 1 2 Birch 1944.
  6. 1 2 3 4 Carey, Francis A.; Sundberg, Richard J. (2007). Advanced Organic Chemistry. Vol. B: Reactions and Synthesis (5th ed.). New York: Springer. pp. 437–439. ISBN   978-0-387-44899-2.
  7. 1 2 3 Zimmerman, H. E. (1961). "Orientation in Metal Ammonia Reductions". Tetrahedron. 16 (1–4): 169–176. doi:10.1016/0040-4020(61)80067-7.
  8. Zimmerman, Howard E (1975). Quantum Mechanics for Organic Chemists . New York: Academic Press. pp.  154–5. ISBN   0-12-781650-X.
  9. 1 2 Zimmerman, H. E. (1963). "Base-Catalyzed Rearrangements". In De Mayo, P. (ed.). Molecular Rearrangements. New York: Interscience. pp. 350–352.
  10. 1 2
    • Zimmerman, H. E.; Wang, P. A. (1990). "The Regioselectivity of the Birch Reduction". J. Am. Chem. Soc. 112 (3): 1280–1281. doi:10.1021/ja00159a078.
    • Zimmerman, H. E.; Wang, P. A. (1993). "Regioselectivity of the Birch Reduction". J. Am. Chem. Soc. 115 (6): 2205–2216. doi:10.1021/ja00059a015.
  11. Paufler, R. M. (1960) Ph.D. Thesis, Northwestern University, Evanston, IL.
  12. Ecsery, Zoltan & Muller, Miklos (1961). "Reduction vitamin D2 with alkaly metals". Magyar Kémiai Folyóirat . 67: 330–332.
  13. Donohoe, Timothy J. & House, David (2002). "Ammonia Free Partial Reduction of Aromatic Compounds Using Lithium Di-tert-butylbiphenyl (LiDBB)". Journal of Organic Chemistry. 67 (14): 5015–5018. doi:10.1021/jo0257593. PMID   12098328.
  14. Garst, Michael E.; Lloyd J.; Shervin; N. Andrew; Natalie C.; Alfred A.; et al. (2000). "Reductions with Lithium in Low Molecular Weight Amines and Ethylenediamine". Journal of Organic Chemistry. 65 (21): 7098–7104. doi:10.1021/jo0008136. PMID   11031034.
  15. Peters, Byron K.; Rodriguez, Kevin X.; Reisberg, Solomon H.; Beil, Sebastian B.; Hickey, David P.; Kawamata, Yu; Collins, Michael; Starr, Jeremy; Chen, Longrui; Udyavara, Sagar; Klunder, Kevin; Gorey, Timothy J.; Anderson, Scott L.; Neurock, Matthew; Minteer, Shelley D.; Baran, Phil S. (21 February 2019). "Scalable and safe synthetic organic electroreduction inspired by Li-ion battery chemistry". Science. 363 (6429): 838–845. Bibcode:2019Sci...363..838P. doi:10.1126/science.aav5606. PMC   7001862 . PMID   30792297.
  16. Taber, D. F.; Gunn, B. P.; Ching Chiu, I. (1983). "Alkylation of the anion from Birch reduction of o-Anisic acid: 2-Heptyl-2-cyclohexenone". Organic Syntheses .; Collective Volume, vol. 7, p. 249
  17. Kuehne, M. E.; Lambert, B. F. (1963). "1,4-Dihydrobenzoic acid". Organic Syntheses .; Collective Volume, vol. 5, p. 400
  18. Paquette, L. A.; Barrett, J. H. (1969). "2,7-Dimethyloxepin". Organic Syntheses .; Collective Volume, vol. 5, p. 467
  19. Clive, Derrick L. J. & Sunasee, Rajesh (2007). "Formation of Benzo-Fused Carbocycles by Formal Radical Cyclization onto an Aromatic Ring". Organic Letters. 9 (14): 2677–2680. doi:10.1021/ol070849l. PMID   17559217.
  20. Birch Reductions, Institute of Chemistry, Skopje, Macedonia
  21. Vogel, E.; Klug, W.; Breuer, A. (1974). "1,6-Methano[10]annulene". Organic Syntheses .; Collective Volume, vol. 6
  22. Edwin M. Kaiser and Robert A. Benkeser "Δ9,10-Octalin" Org. Synth. 1970, vol. 50, p. 88ff. doi : 10.15227/orgsyn.050.0088
  23. Merck Index , 13th Ed.
  24. Wooster, C. B.; Godfrey, K. L. (1937). "Mechanism of the Reduction of Unsaturated Compounds with Alkali Metals and Water". Journal of the American Chemical Society. 59 (3): 596. doi:10.1021/ja01282a504.
  25. Birch 1945.
  26. Birch 1946.
  27. Wilds, A. L.; Nelson, N. A. (1953). "A Superior Method for Reducing Phenol Ethers to Dihydro Derivatives and Unsaturated Ketones". J. Am. Chem. Soc. 75 (21): 5360–5365. doi:10.1021/ja01117a064.
  28. Birch, A. J.; Smith, H. (1958). "Reduction by metal–amine solutions: applications in synthesis and determination of structure". Quart. Rev. (review). 12 (1): 17. doi:10.1039/qr9581200017.
  29. Bachi, J. W.; Epstein, Y.; Herzberg-Minzly, H.; Loewnenthal, J. E. (1969). "Synthesis of compounds related to gibberellic acid. III. Analogs of ring a of the gibberellins". J. Org. Chem. 34: 126–135. doi:10.1021/jo00838a030.
  30. Taber, D. F.; Gunn, B.P; Ching Chiu, I (1983). "Alkylation of the Anion from Birch Reduction of o-Anisic Acid: 2-Heptyl-2-Cyclohexenone". Organic Syntheses . 61: 59.; Collective Volume, vol. 7, p. 249
  31. Guo, Z.; Schultz, A. G. (2001). "Organic synthesis methodology. Preparation and diastereoselective birch reduction-alkylation of 3-substituted 2-methyl-2,3-dihydroisoindol-1-ones". J. Org. Chem. 66 (6): 2154–2157. doi:10.1021/jo005693g. PMID   11300915.
  32. Birch, A. J.; Nasipuri, D. (1959). "Reaction mechanisms in reduction by metal-ammonia solutions". Tetrahedron. 6 (2): 148–153. doi:10.1016/0040-4020(59)85008-0.
  33. Burnham, D. R. (1969). "Orientation in the mechanism of the Birch reduction of anisole". Tetrahedron. 25 (4): 897–904. doi:10.1016/0040-4020(69)85023-4.
    • Birch, A. J.; Hinde, A. L.; Radom, L. (1980). "A theoretical approach to the Birch reduction. Structures and stabilities of the radical anions of substituted benzenes". J. Am. Chem. Soc. 102 (10): 3370–3376. doi:10.1021/ja00530a012.
    • Birch, A. J.; Radom, L. (1980). "A theoretical approach to the Birch reduction. Structures and stabilities of cyclohexadienyl radicals". J. Am. Chem. Soc. 102 (12): 4074–4080. doi:10.1021/ja00532a016.
  34. See diagrams in: