To install click the Add extension button. That's it.

The source code for the WIKI 2 extension is being checked by specialists of the Mozilla Foundation, Google, and Apple. You could also do it yourself at any point in time.

4,5
Kelly Slayton
Congratulations on this excellent venture… what a great idea!
Alexander Grigorievskiy
I use WIKI 2 every day and almost forgot how the original Wikipedia looks like.
What we do. Every page goes through several hundred of perfecting techniques; in live mode. Quite the same Wikipedia. Just better.
.
Leo
Newton
Brights
Milds

Birch reduction

From Wikipedia, the free encyclopedia

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.

The Birch reduction
The Birch reduction

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

naphthalene Birch Reduction
naphthalene Birch Reduction

YouTube Encyclopedic

  • 1/5
    Views:
    87 472
    629 867
    37 012
    43 025
    885
  • Birch reduction I | Aromatic Compounds | Organic chemistry | Khan Academy
  • The Birch reduction
  • Birch Reduction Reaction and Mechanism Benzene and Substituted Rings Leah Fisch
  • Birch reduction II | Aromatic Compounds | Organic chemistry | Khan Academy
  • 37.05 Birch Reduction

Transcription

In this video, we're going to look at the general mechanism for the Birch reduction. So we start with benzene and to it we add an alkaline metal like sodium and liquid ammonia and also an alcohol, and the end result is to reduce the benzene ring to form 1, 4-cyclohexadiene. Let's look at the mechanism for the Birch reduction. So we know that sodium is in group one of the periodic table and so it has one valence electron, which I will go ahead and color magenta there. And so the start of the mechanism is for sodium to donate its one valence electron to the benzene ring, and so we can show the movement of that one electron with a fish hook arrow or a half-headed arrow here where we show that one electron moving over here to this carbon, so this carbon right here. Now, we're also going to get some movement of electrons in our benzene ring. So when I think about these electrons in here, so these pi electrons in red, we know that there are two electrons there. So let me go ahead and show those two electrons like that. So those two electrons are also going to move. So this electron over here is going to come off on this carbon as well, and then this other electron in red is going to move over to here. So let's go ahead and show the start of the movement of those electrons. We're going to come back and move some more, but I just want to do this really slowly here so we can follow along. So we had our hydrogens attached to our rings so let me go ahead and sketch those in really fast here. So these pi electrons are going to stay put for our mechanism. And let me show the electron in magenta, the one that sodium donated, so it ends up being on this carbon right here. And then one of the electrons in red also moved on to that carbon. So let me show that electron in red right there. So that carbon gets a negative 1 formal charge so we form an anion here. One of those red electrons is going to move over to this position right here, and then we are also going to show these electrons moving around, so the electrons in green here. Let me just go ahead and highlight those. So these electrons, so there's one and there's two. So one of them is going to move to the same position that the red one did in there, and the other one is going to come off onto this carbon. So let me see if I can show that. So one of them moved in here like that, and the other one moved off onto this carbon like that. So the one in red and the one in green are, of course, now a pi bond. You could think about it that way, and so we've now generated what's called a radical anion here. So this is a radical anion. So it's a radical because you have that one unpaired electron in green, and it's an anion since you have a negative 1 formal charge over here on this topic carbon. And then I forgot to put in this hydrogen so let me go ahead and add that one in there like that. Second step of our mechanism, our alcohol comes along so we have our generic alcohol, which is going to function as an acid because the negative 1 formal charge, the anion here, the carbanion, is going to function as a base. And so these electrons here are going to pick up a proton from the alcohol so these electrons will kick off onto the oxygen here. And so let's go ahead and draw the result of that acid base reaction. So we have these pi electrons in here, and we now have two hydrogens on that top carbon. We have one hydrogen on each of our other carbons here. And we still have a radical. So let me go ahead and show this electron is still on that carbon. So now we have a radical instead of a radical anion. And I should point out that for our radical anion and for our radical, the electron density can be delocalized throughout the ring, but here we're just trying to show just moving around some electrons. And so let me go ahead and highlight to these two electrons here. So the electron in red and the electron in magenta are forming a bond with that proton right here on our ring. So next step in our mechanism, we get some more sodium so some more sodium comes along here. Let me go ahead and show that. And, of course, once again, sodium has one valence electron so here's sodium's one valence electron. The sodium can donate that valence electron to our benzene ring, and so it's going to donate it over here to this carbon, the carbon that had the green electron on it already. And so let's go ahead and show the result of that. We would have our ring, we would have our pi electrons. We had two hydrogens bonded to the top carbon. We had these hydrogens around my ring like that. The bottom carbon still has a hydrogen bonded to it. And we started with a green electron on that carbon, and now we're going to add a magenta electron, giving that carbon a negative 1 formal charge so we form an anion again. So now that we have an anion, the last step of the mechanism is another acid base reaction so our alcohol comes along, and the carbanion is going to function as a base and pick up a proton from our alcohol. So the same step that we saw before pretty much. And we go ahead and draw our final product. So we have those pi electrons. We had these hydrogens on our top carbon, these carbons all get hydrogens, and then finally, we have added on a proton to this bottom carbon here. So let me go ahead and highlight those. So these electrons here, those two electrons pick up a proton so we protonate our ring and we finish. So this is our 1, 4 cycylohexadiene product here. Now, a simple way of thinking about this mechanism for the Birch reduction is to break it into these four steps and to make those steps very simple. So in the first step, sodium is donating electrons an electron so you could think electron for step one. Second step, we know that the anion is picking up a proton from our alcohol so you could think proton for the second step. For the third step, once again, sodium is going to donate an electron so you could think electron. And then finally, once again the, anion is picking up a proton so you could think proton. So you could think electron, proton, electron, and proton is a simple way of thinking about the steps for a Birch reduction. In the next video, we're going to look at what happens with the Birch reduction when you get a substituted benzene ring. And I'll show you mechanisms for the two possibilities that you might see on an exam.

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]

Electron attacks a benzene ring, which then abstracts a proton from ROH; process then repeats in the para position.
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.

Reaction as with benzene, but protonation proceeds immediately ortho.
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

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

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
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]

Birch alkylation
Adding 1,4-dibromobutane to a Birch reduction of <i>tert</i>-butyl benzoate forms the 1,1-cyclohexadiene product.[19]

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
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

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

  • 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

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. ^ 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. ^ Birch Reductions, 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. ^
  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:
This page was last edited on 15 December 2023, at 08:56
Basis of this page is in Wikipedia. Text is available under the CC BY-SA 3.0 Unported License. Non-text media are available under their specified licenses. Wikipedia® is a registered trademark of the Wikimedia Foundation, Inc. WIKI 2 is an independent company and has no affiliation with Wikimedia Foundation.