Kulinkovich reaction

Last updated

Contents

Kulinkovich reaction
Named afterOleg Kulinkovich
Reaction type Ring forming reaction
Identifiers
Organic Chemistry Portal kulinkovich-reaction
RSC ontology ID RXNO:0000682

The Kulinkovich reaction describes the organic synthesis of substituted cyclopropanols through reaction of esters with dialkyl­dialkoxy­titanium reagents, which are generated in situ from Grignard reagents containing a hydrogen in beta-position and titanium(IV) alkoxides such as titanium isopropoxide. [1] This reaction was first reported by Oleg Kulinkovich and coworkers in 1989. [2]

Kulinkovich reaction Kulinkovich overview.png
Kulinkovich reaction

Titanium catalysts are ClTi(OiPr)3 or Ti(OiPr)4, ClTi(OtBu)3 or Ti(OtBu)4, Grignard reagents are EtMgX, PrMgX or BuMgX. Solvents can be Et2O, THF, toluene. Tolerated Functional Groups: Ethers R–O–R, R–S–R, Imines RN=CHR. Amides, primary and secondary amines. Carbamates typically do not tolerate the reaction conditions, but tert-butyl carbamates (N-Boc derivatives) survive the transformation.

An asymmetric version of this reaction is also known with a TADDOL-based catalyst. [1]

Reaction mechanism

The generally accepted reaction mechanism initially utilizes two successive stages of transmetallation of the committed Grignard reagent, leading to an intermediate dialkyldiisopropoxytitanium complex. This complex undergoes a dismutation to give an alkane molecule and a titanacyclopropane 1. The insertion of the carbonyl group of the ester in the weakest carbon-titanium bond, leads to an oxatitanacyclopentane 2 being rearranged to ketone 3. Lastly, the insertion of the carbonyl group of 3 in the residual carbon-titanium connection forms a cyclopropane ring. In the transition state of this elementary stage, which is the limiting stage of the reaction, an agostic interaction stabilizing between the beta hydrogen and the R2 group and the titanium atom was called upon to explain the diastereoselectivity observed. Complex 4 obtained is a tetraalkyloxytitanium compound able to play a part similar to that of the starting tetraisopropyloxytitanate, which closes the catalytic cycle. At the end of the reaction, the product is mainly in the shape of the magnesium alcoholate 5, giving the cyclopropanol after hydrolysis by the reaction medium.

The step leading to the titanacyclopropane has been scrutinized computationally. Although the dialkyldiisopropoxytitanium complex has been proposed to undergo β hydrogen elimination followed by C–H reductive elimination to give the alkane and 1, it was found that β hydrogen abstraction by the alkyl group, leading directly to products without the intermediate titanium hydride, is a more favorable process. [3]

In broad strokes, and in a formal retrosynthetic sense, titanacyclopropane 1 behaves like a 1,2-dianion which adds into the ester twice: after the first addition into the ester, the resultant tetrahedral intermediate 2 collapses to give β-titanio ketone 3, which undergoes a second intramolecular addition to give the titanium salt of the cyclopropanol (4). (This species then undergoes transmetalation with Grignard reagent to regenerate 1 and close the catalytic cycle and give the product in the form of the magnesium salt (5).)

ReactionKulinkovich2.png

The reaction mechanism has been the subject of theoretical analysis. [4] Certain points remain nevertheless obscure. Intermediate titanium complexes of the ate type have been proposed by Kulinkovich. [5]

Ligand exchange with olefins

In 1993, the team of Kulinkovich highlighted the aptitude of the titanacyclopropanes to undergo ligand exchange with olefins. [6] This discovery was important, because it gave access to cyclopropanols more functionalized by making economic use of the Grignard of which normally at least two equivalents should have been engaged to obtain good yields. Cha and its team introduced the use of cyclic Grignard reagents, particularly adapted for these reactions. [7]

ReactionKulinkovich3.png

The methodology has been extended to intramolecular reactions [8]

De Meijere variation

Kulinkovich–De Meijere reaction
Named afterOleg Kulinkovich
Armin de Meijere
Reaction type Ring forming reaction
Identifiers
Organic Chemistry Portal kulinkovich-demeijere-reaction
RSC ontology ID RXNO:0000683

With amides instead of esters the reaction product is an aminocyclopropane in the De Meijere variation [9] [10]

ReactionKdM1.png

The intramolecular reaction is also known: [11] [12] [13] [14] [15] [16] [17] [18] [19] [20]

ReactionKdM4.png

Szymoniak variation

Kulinkovich–Szymoniak reaction
Named afterOleg Kulinkovich
Jan Szymoniak
Reaction type Ring forming reaction
Identifiers
Organic Chemistry Portal kulinkovich-szymoniak-reaction
RSC ontology ID RXNO:0000684

In the Szymoniak variation the substrate is a nitrile and the reaction product a cyclopropane with a primary amine group. [21] [22]

ReactionKSz1.png

The reaction mechanism is akin the Kulinkovich reaction:

ReactionKSz2.png

Additional reading

Related Research Articles

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

In organic chemistry, the ene reaction is a chemical reaction between an alkene with an allylic hydrogen and a compound containing a multiple bond, in order to form a new σ-bond with migration of the ene double bond and 1,5 hydrogen shift. The product is a substituted alkene with the double bond shifted to the allylic position.

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.

A transition metal carbene complex is an organometallic compound featuring a divalent organic ligand. The divalent organic ligand coordinated to the metal center is called a carbene. Carbene complexes for almost all transition metals have been reported. Many methods for synthesizing them and reactions utilizing them have been reported. The term carbene ligand is a formalism since many are not derived from carbenes and almost none exhibit the reactivity characteristic of carbenes. Described often as M=CR2, they represent a class of organic ligands intermediate between alkyls (−CR3) and carbynes (≡CR). They feature in some catalytic reactions, especially alkene metathesis, and are of value in the preparation of some fine chemicals.

<span class="mw-page-title-main">Pauson–Khand reaction</span> Chemical reaction

The Pauson–Khand (PK) reaction is a chemical reaction, described as a [2+2+1] cycloaddition. In it, an alkyne, an alkene and carbon monoxide combine into a α,β-cyclopentenone in the presence of a metal-carbonyl catalyst.

The Reformatsky reaction is an organic reaction which condenses aldehydes or ketones with α-halo esters using metallic zinc to form β-hydroxy-esters:

<span class="mw-page-title-main">Chiral auxiliary</span> Stereogenic group placed on a molecule to encourage stereoselectivity in reactions

In stereochemistry, a chiral auxiliary is a stereogenic group or unit that is temporarily incorporated into an organic compound in order to control the stereochemical outcome of the synthesis. The chirality present in the auxiliary can bias the stereoselectivity of one or more subsequent reactions. The auxiliary can then be typically recovered for future use.

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

The McMurry reaction is an organic reaction in which two ketone or aldehyde groups are coupled to form an alkene using a titanium chloride compound such as titanium(III) chloride and a reducing agent. The reaction is named after its co-discoverer, John E. McMurry. The McMurry reaction originally involved the use of a mixture TiCl3 and LiAlH4, which produces the active reagents. Related species have been developed involving the combination of TiCl3 or TiCl4 with various other reducing agents, including potassium, zinc, and magnesium. This reaction is related to the Pinacol coupling reaction which also proceeds by reductive coupling of carbonyl compounds.

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

Titanocene dichloride is the organotitanium compound with the formula (η5-C5H5)2TiCl2, commonly abbreviated as Cp2TiCl2. This metallocene is a common reagent in organometallic and organic synthesis. It exists as a bright red solid that slowly hydrolyzes in air. It shows antitumour activity and was the first non-platinum complex to undergo clinical trials as a chemotherapy drug.

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

Tebbe's reagent is the organometallic compound with the formula (C5H5)2TiCH2ClAl(CH3)2. It is used in the methylidenation of carbonyl compounds, that is it converts organic compounds containing the R2C=O group into the related R2C=CH2 derivative. It is a red solid that is pyrophoric in the air, and thus is typically handled with air-free techniques. It was originally synthesized by Fred Tebbe at DuPont Central Research.

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

The Petasis reagent, named after Nicos A. Petasis, is an organotitanium compound with the formula Cp2Ti(CH3)2. It is an orange-colored solid.

<span class="mw-page-title-main">Cyclopropanation</span> Chemical process which generates cyclopropane rings

In organic chemistry, cyclopropanation refers to any chemical process which generates cyclopropane rings. It is an important process in modern chemistry as many useful compounds bear this motif; for example pyrethroid insecticides and a number of quinolone antibiotics. However, the high ring strain present in cyclopropanes makes them challenging to produce and generally requires the use of highly reactive species, such as carbenes, ylids and carbanions. Many of the reactions proceed in a cheletropic manner.

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

Organozinc chemistry is the study of the physical properties, synthesis, and reactions of organozinc compounds, which are organometallic compounds that contain carbon (C) to zinc (Zn) chemical bonds.

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

Organotitanium chemistry is the science of organotitanium compounds describing their physical properties, synthesis, and reactions. Organotitanium compounds in organometallic chemistry contain carbon-titanium chemical bonds. They are reagents in organic chemistry and are involved in major industrial processes.

A carbometallation is any reaction where a carbon-metal bond reacts with a carbon-carbon π-bond to produce a new carbon-carbon σ-bond and a carbon-metal σ-bond. The resulting carbon-metal bond can undergo further carbometallation reactions or it can be reacted with a variety of electrophiles including halogenating reagents, carbonyls, oxygen, and inorganic salts to produce different organometallic reagents. Carbometallations can be performed on alkynes and alkenes to form products with high geometric purity or enantioselectivity, respectively. Some metals prefer to give the anti-addition product with high selectivity and some yield the syn-addition product. The outcome of syn and anti- addition products is determined by the mechanism of the carbometallation.

Electrophilic amination is a chemical process involving the formation of a carbon–nitrogen bond through the reaction of a nucleophilic carbanion with an electrophilic source of nitrogen.

<span class="mw-page-title-main">Carbonyl reduction</span> Organic reduction of any carbonyl group by a reducing agent

In organic chemistry, carbonyl reduction is the conversion of any carbonyl group, usually to an alcohol. It is a common transformation that is practiced in many ways. Ketones, aldehydes, carboxylic acids, esters, amides, and acid halides - some of the most pervasive functional groups, -comprise carbonyl compounds. Carboxylic acids, esters, and acid halides can be reduced to either aldehydes or a step further to primary alcohols, depending on the strength of the reducing agent. Aldehydes and ketones can be reduced respectively to primary and secondary alcohols. In deoxygenation, the alcohol group can be further reduced and removed altogether by replacement with H.

<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.

An insertion reaction is a chemical reaction where one chemical entity interposes itself into an existing bond of typically a second chemical entity e.g.:

Carbonyl olefin metathesis is a type of metathesis reaction that entails, formally, the redistribution of fragments of an alkene and a carbonyl by the scission and regeneration of carbon-carbon and carbon-oxygen double bonds respectively. It is a powerful method in organic synthesis using simple carbonyls and olefins and converting them into less accessible products with higher structural complexity.

In organic chemistry, the Lombardo methylenation is a name reaction that allows for the methylenation of carbonyl compounds with the use of Lombardo's reagent, which is a mix of zinc, dibromomethane, and titanium tetrachloride.

References

  1. 1 2 Cha, Jin Kun; Kulinkovich, Oleg G. (2012). "The Kulinkovich cyclopropanation of carboxylic acid derivatives". Organic Reactions. 77: 1–159. doi:10.1002/0471264180.or077.01. ISBN   978-0471264187.
  2. Kulinkovich, O. G.; Sviridov, S. V.; Vasilevskii, D. A.; Pritytskaya, T. S. (1989). "Reaction of ethylmagnesium bromide with carboxylic acid esters in the presence of tetraisopropoxytitanium". Zh. Org. Khim. 25: 2244.
  3. Bertus, Philippe (11 November 2019). "From Dialkyltitanium Species to Titanacyclopropanes: An Ab Initio Study". Organometallics. 38 (21): 4171–4182. doi:10.1021/acs.organomet.9b00509. ISSN   0276-7333.
  4. Wu, Y.–D. and Yu, Z.-X. (2001). "A theoretical study on the mechanism and diastereoselectivity of the Kulinkovich hydroxycyclopropanation reaction". J. Am. Chem. Soc. 123 (24): 5777–86. doi:10.1021/ja010114q. PMID   11403612.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  5. Kulinkovich, O. G.; Kananovich, D. G. (2007). "Advanced Procedure for the Preparation ofcis-1,2-Dialkylcyclopropanols – Modified Ate Complex Mechanism for Titanium-Mediated Cyclopropanation of Carboxylic Esters with Grignard Reagents". Eur. J. Org. Chem. 2007 (13): 2121–32. doi:10.1002/ejoc.200601035.
  6. Kulinkovich, O. G.; Savchenko, A. I.; Sviridov, S. V.; Vasilevski, D. A. (1993). "Titanium(IV) Isopropoxide-catalysed Reaction of Ethylmagnesium Bromide with Ethyl Acetate in the Presence of Styrene". Mendeleev Commun. 3 (6): 230–31. doi:10.1070/MC1993v003n06ABEH000304.
  7. Lee, J.; Kim, H.; Cha, J. K. (1996). "A New Variant of the Kulinkovich Hydroxycyclopropanation. Reductive Coupling of Carboxylic Esters with Terminal Olefins". J. Am. Chem. Soc. 118 (17): 4198–99. doi:10.1021/ja954147f.
  8. Kasatkin, A.; Sato, F. (1995). "Diastereoselective synthesis of trans-1,2-disubstituted cyclopropanols from homoallyl or bis-homoallyl esters via tandem intramolecular nucleophilic acyl substitution and intramolecular carbonyl addition reactions mediated by Ti(OPr-i)4 / 2 i-PrMgBr reagent". Tetrahedron Lett. 36 (34): 6079–82. doi:10.1016/0040-4039(95)01208-Y.
  9. Chaplinski, V.; De Meijere, A. (1996). "A Versatile New Preparation of Cyclopropylamines from Acid Dialkylamides". Angew. Chem. Int. Ed. 35 (4): 413–14. doi:10.1002/anie.199604131.
  10. De Meijere, A.; Winsel, H. and Stecker, B. Organic Syntheses, Vol. 81, p. 14 Archived 18 September 2012 at the Wayback Machine
  11. Lee, J.; Cha, J. K. (1997). "Facile Preparation of Cyclopropylamines from Carboxamides". The Journal of Organic Chemistry. 62 (6): 1584. doi:10.1021/jo962368d.
  12. Chaplinski, V.; Winsel, H.; Kordes, M.; De Meijere, A. (1997). "A New Versatile Reagent for the Synthesis of Cyclopropylamines Including 4-Azaspiro[2.n]alkanes and Bicyclo[n.1.0]alkylamines". Synlett. 1997: 111–114. doi:10.1055/s-1997-17828.
  13. Cao, B.; Xiao, D.; Joullié, M. M. (1999). "Synthesis of Bicyclic Cyclopropylamines by Intramolecular Cyclopropanation of N-Allylamino Acid Dimethylamides". Organic Letters. 1 (11): 1799. doi:10.1021/ol9910520. PMID   10814442.
  14. Lee, H. B.; Sung, M. J.; Blackstock, S. C.; Cha, J. K. (2001). "Radical cation-mediated annulation. Stereoselective construction of bicyclo[5.3.0]decan-3-ones by aerobic oxidation of cyclopropylamines". Journal of the American Chemical Society. 123 (45): 11322–11324. doi:10.1021/ja017043f. PMID   11697988.
  15. Gensini, M.; Kozhushkov, S. I.; Yufit, D. S.; Howard, J. A. K.; Es-Sayed, M.; Meijere, A. D. (2002). "3-Azabicyclo[3.1.0]hex-1-ylamines by Ti-Mediated Intramolecular Reductive Cyclopropanation of α-(N-Allylamino)-Substituted N,N-Dialkylcarboxamides and Carbonitriles". European Journal of Organic Chemistry. 2002 (15): 2499. doi:10.1002/1099-0690(200208)2002:15<2499::AID-EJOC2499>3.0.CO;2-V.
  16. Tebben, G. D.; Rauch, K.; Stratmann, C.; Williams, C. M.; De Meijere, A. (2003). "Intramolecular Titanium-Mediated Aminocyclopropanation of Terminal Alkenes: Easy Access to Various Substituted Azabicyclo[n.1.0]alkanes1". Organic Letters. 5 (4): 483–485. doi:10.1021/ol027352q. PMID   12583749.
  17. Ouhamou, N.; Six, Y. (2003). "Studies on the intramolecular Kulinkovich?De Meijere reaction of disubstituted alkenes bearing carboxylic amide groups". Organic & Biomolecular Chemistry. 1 (17): 3007–9. doi:10.1039/b306719j. PMID   14518121.
  18. Gensini, M.; De Meijere, A. (2004). "Cyclopropane-Annelated Azaoligoheterocycles by Ti-Mediated Intramolecular Reductive Cyclopropanation of Cyclic Amino Acid Amides". Chemistry: A European Journal. 10 (3): 785–790. doi:10.1002/chem.200305068. PMID   14767944.
  19. Larquetoux, L.; Kowalska, J. A.; Six, Y. (2004). "The Formal [3+2+1] Cyclisation of Cyclopropylamines with Carboxylic Anhydrides: A Quick Access to Polysubstituted 2,3,3a,4-Tetrahydro6(5H)-indolone Ring Systems". European Journal of Organic Chemistry. 2004 (16): 3517. doi:10.1002/ejoc.200400291.
  20. Larquetoux, L.; Ouhamou, N.; Chiaroni, A. L.; Six, Y. (2005). "The Intramolecular Aromatic Electrophilic Substitution of Aminocyclopropanes Prepared by the Kulinkovich–De Meijere Reaction". European Journal of Organic Chemistry. 2005 (21): 4654. doi:10.1002/ejoc.200500428.
  21. Bertus, P.; Szymoniak, J. (2001). "New and easy route to primary cyclopropylamines from nitriles". Chemical Communications (18): 1792–1793. doi:10.1039/b105293b. PMID   12240317.
  22. Chaplinski, V.; De Meijere, A. (1996). "A Versatile New Preparation of Cyclopropylamines from Acid Dialkylamides". Angewandte Chemie International Edition in English. 35 (4): 413. doi:10.1002/anie.199604131.