fbpx
Wikipedia

Birch reduction

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.

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

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

naphthalene Birch Reduction

Reaction mechanism and regioselectivity edit

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.

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.

 
Birch reduction of anisole.

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]

 


 

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 edit

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).

 

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]

Approximation Density Atom 3 Density Atom 2 Density Atom 1 Bond Order 2–3 Bond Order 1–2
Hückel (1st approx) 0.333 0.00 0.333 0.788 0.578
2nd approx 0.317 0.00 0.365 0.802 0.564
3rd approx 0.316 0.00 0.368 0.802 0.562

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 edit

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 edit

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

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.[19]

Benkeser reduction edit

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
 
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 edit

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

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 edit

  • Caine, D. (1976). "Reduction and Related Reactions of α,β-Unsaturated Carbonyl Compounds with Metals in Liquid Ammonia". Org. React. (review). 23: 1–258. doi:10.1002/0471264180.or023.01. ISBN 0471264180.

See also edit

References edit

  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. ^ a b 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. ^ a b Birch 1944.
  6. ^ a b c d 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. ^ a b c 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. ^ a b Zimmerman, H. E. (1963). "Base-Catalyzed Rearrangements". In De Mayo, P. (ed.). Molecular Rearrangements. New York: Interscience. pp. 350–352.
  10. ^ a b
    • 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; Collected Volumes, vol. 7, p. 249.
  17. ^ Kuehne, M. E.; Lambert, B. F. (1963). "1,4-Dihydrobenzoic acid". Organic Syntheses; Collected Volumes, vol. 5, p. 400.
  18. ^ Paquette, L. A.; Barrett, J. H. (1969). "2,7-Dimethyloxepin". Organic Syntheses; Collected Volumes, 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. ^ , Institute of Chemistry, Skopje, Macedonia
  21. ^ Vogel, E.; Klug, W.; Breuer, A. (1974). "1,6-Methano[10]annulene". Organic Syntheses; Collected Volumes, 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, A. J. (1944). "Reduction by dissolving metals. Part I". J. Chem. Soc.: 430. doi:10.1039/JR9440000430.
    • Birch, A. J. (1945). "Reduction by dissolving metals. Part II". J. Chem. Soc.: 809. doi:10.1039/jr9450000809.
    • Birch, A. J. (1946). "Reduction by dissolving metals. Part III". J. Chem. Soc.: 593. doi:10.1039/jr9460000593.
    • Birch, A. J. (1947). "Reduction by dissolving metals. Part IV". J. Chem. Soc.: 102. doi:10.1039/jr9470000102.
    • Birch, Arthur J. (1947). "Reduction by dissolving metals. Part V". J. Chem. Soc.: 1642. doi:10.1039/jr9470001642.
    • Birch, A. J.; Mukherji, S. M. (1949). "Reduction by dissolving metals. Part VI. Some applications in synthesis". J. Chem. Soc.: 2531. doi:10.1039/jr9490002531.
  26. ^ Birch 1945.
  27. ^ Birch 1946.
  28. ^ 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.
  29. ^ 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.
  30. ^ 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.
  31. ^ 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; Collected Volumes, vol. 7, p. 249.
  32. ^ 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.
  33. ^ 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.
  34. ^ 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.
  35. ^
    • 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.
  36. ^ See diagrams in:
    • Birch, A. J. (1992). "Steroid hormones and the Luftwaffe. A venture into fundamental strategic research and some of its consequences: The Birch reduction becomes a birth reduction". Steroids. 57 (8): 363–377. doi:10.1016/0039-128X(92)90080-S. PMID 1519267. S2CID 24827957.
    • Birch, A. J. (1996). "The Birch reduction in organic synthesis". Pure Appl. Chem. 68 (3): 553–556. doi:10.1351/pac199668030553. S2CID 41494178.

birch, reduction, organic, reaction, that, used, convert, arenes, cyclohexadienes, reaction, named, after, australian, chemist, arthur, birch, involves, organic, reduction, aromatic, rings, amine, solvent, traditionally, liquid, ammonia, with, alkali, metal, t. 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 Birch reduction Named after Arthur Birch Reaction type Organic redox reaction Identifiers Organic Chemistry Portal birch reduction RSC ontology ID RXNO 0000042 The Birch reductionAn example is the reduction of naphthalene in ammonia and ethanol naphthalene Birch Reduction Contents 1 Reaction mechanism and regioselectivity 1 1 Secondary protonation regioselectivity 2 Modifications 2 1 Birch alkylation 2 2 Benkeser reduction 3 History 4 Additional reading 5 See also 6 ReferencesReaction mechanism and regioselectivity editA 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 nbsp Birch reduction of benzene also available in animated form 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 nbsp Birch reduction of anisole 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 nbsp nbsp 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 editThe 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 nbsp Simple Huckel 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 Approximation Density Atom 3 Density Atom 2 Density Atom 1 Bond Order 2 3 Bond Order 1 2 Huckel 1st approx 0 333 0 00 0 333 0 788 0 578 2nd approx 0 317 0 00 0 365 0 802 0 564 3rd approx 0 316 0 00 0 368 0 802 0 562 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 b g unsaturated carbonyl 7 11 Modifications editTraditional 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 edit In Birch alkylation the anion formed in the Birch reduction is trapped by a suitable electrophile such as a haloalkane for example 16 nbsp 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 nbsp Adding 1 4 dibromobutane to a Birch reduction of tert butyl benzoate forms the 1 1 cyclohexadiene product 19 Benkeser reduction edit 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 nbsp The Benkeser reaction nbsp 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 b substituents one bond further tend to direct reduction to the substituted ring 6 History editArthur 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 nbsp Benzoic acid reduction including possible alkylationFor 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 Huckel 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 editCaine D 1976 Reduction and Related Reactions of a b Unsaturated Carbonyl Compounds with Metals in Liquid Ammonia Org React review 23 1 258 doi 10 1002 0471264180 or023 01 ISBN 0471264180 See also editSolvated electron the reducing agent Bouveault Blanc reduction another reaction using solvated electrons Synthesis of methamphetamine an applicationReferences edit March Jerry 1985 Advanced Organic Chemistry Reactions Mechanisms and Structure 3rd edition New York Wiley ISBN 9780471854722 OCLC 642506595 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 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 a b 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 a b Birch 1944 a b c d 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 a b c Zimmerman H E 1961 Orientation in Metal Ammonia Reductions Tetrahedron 16 1 4 169 176 doi 10 1016 0040 4020 61 80067 7 Zimmerman Howard E 1975 Quantum Mechanics for Organic Chemists New York Academic Press pp 154 5 ISBN 0 12 781650 X a b Zimmerman H E 1963 Base Catalyzed Rearrangements In De Mayo P ed Molecular Rearrangements New York Interscience pp 350 352 a b 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 Paufler R M 1960 Ph D Thesis Northwestern University Evanston IL Ecsery Zoltan amp Muller Miklos 1961 Reduction vitamin D2 with alkaly metals Magyar Kemiai Folyoirat 67 330 332 Donohoe Timothy J amp 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 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 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 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 Collected Volumes vol 7 p 249 Kuehne M E Lambert B F 1963 1 4 Dihydrobenzoic acid Organic Syntheses Collected Volumes vol 5 p 400 Paquette L A Barrett J H 1969 2 7 Dimethyloxepin Organic Syntheses Collected Volumes vol 5 p 467 Clive Derrick L J amp 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 Birch Reductions Institute of Chemistry Skopje Macedonia Vogel E Klug W Breuer A 1974 1 6 Methano 10 annulene Organic Syntheses Collected Volumes vol 6 Edwin M Kaiser and Robert A Benkeser D9 10 Octalin Org Synth 1970 vol 50 p 88ff doi 10 15227 orgsyn 050 0088 Merck Index 13th Ed 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 Birch A J 1944 Reduction by dissolving metals Part I J Chem Soc 430 doi 10 1039 JR9440000430 Birch A J 1945 Reduction by dissolving metals Part II J Chem Soc 809 doi 10 1039 jr9450000809 Birch A J 1946 Reduction by dissolving metals Part III J Chem Soc 593 doi 10 1039 jr9460000593 Birch A J 1947 Reduction by dissolving metals Part IV J Chem Soc 102 doi 10 1039 jr9470000102 Birch Arthur J 1947 Reduction by dissolving metals Part V J Chem Soc 1642 doi 10 1039 jr9470001642 Birch A J Mukherji S M 1949 Reduction by dissolving metals Part VI Some applications in synthesis J Chem Soc 2531 doi 10 1039 jr9490002531 Birch 1945 Birch 1946 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 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 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 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 Collected Volumes vol 7 p 249 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 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 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 See diagrams in Birch A J 1992 Steroid hormones and the Luftwaffe A venture into fundamental strategic research and some of its consequences The Birch reduction becomes a birth reduction Steroids 57 8 363 377 doi 10 1016 0039 128X 92 90080 S PMID 1519267 S2CID 24827957 Birch A J 1996 The Birch reduction in organic synthesis Pure Appl Chem 68 3 553 556 doi 10 1351 pac199668030553 S2CID 41494178 Retrieved from https en wikipedia org w index php title Birch reduction amp oldid 1189999415 Birch alkylation, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.