Benzylic activation and stereocontrol in tricarbonyl(arene)chromium complexes

Last updated

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. [1] [2]

Contents

Introduction

The reaction between tricarbonylchromium complexes Cr(CO)3L3 and electron-rich or electron-neutral aromatic rings produces tricarbonyl(arene)chromium complexes (arene)Cr(CO)3. [3] Complexation to chromium(0) activates the side chain of the arene, facilitating dissociation of a benzylic proton, leaving group, or nucleophilic addition to the homobenzylic position of styrenes. Further transformations of the resulting conformationally restricted, benzylic anion or cation involve the approach of reagents exo to the chromium fragment. Thus, benzylic functionalization reactions of planar chiral chromium arene complexes are highly diastereoselective. Additionally, the chromium tri(carbonyl) fragment can be used as a blocking element in addition reactions to ortho-substituted aromatic aldehydes and alkenes. An ortho substituent is necessary in these reactions to restrict conformations available to the aldehyde or alkene. [4] Removal of the chromium fragment to afford the metal-free functionalized aromatic compound is possible photolytically [5] or with an oxidant. [6]

(1)

CrAreneGen.png

Mechanism and stereochemistry

The majority of benzylic functionalization reactions of tricarbonyl(arene)chromium complexes proceed by mechanisms analogous to those followed by the free arenes. The aromatic ring and benzylic position are activated towards solvolysis, deprotonation, and nucleophilic attack (at the ortho and para positions of the arene) upon complexation to chromium, which is able to stabilize developing charges in the arene ligand. As a result, these reactions of chromium arene complexes are often faster than analogous reactions of free arenes. [7]

(2)

CrAreneMech1.png

Second, in benzylic cations and anions of chromium arene complexes, rotation about the bond connecting the benzylic carbon and aromatic ring is severely restricted. This bond possesses a significant amount of double bond character due to the delocalization of charge into the aromatic ring (and the stabilization of that charge by chromium). [8]

(3)

CrAreneMech2.png

Finally, the chromium tri(carbonyl) moiety serves as a sterically bulky group in reactions of arene chromium complexes, preventing the approach of a reagent endo to chromium. In addition, ortho-substituted aromatic aldehydes and styrenes prefer to adopt a conformation in which the doubly bound oxygen or carbon is pointed away from the ortho substituent. As a result, only one face of the double bond is exposed on the exo face of the aromatic ring. If this were not the case, addition to styrenes and aromatic aldehydes would not be diastereoselective, despite the presence of the chromium tri(carbonyl) group. The ortho substituent is necessary for high stereoselectivity; meta-substituted arenes exhibit very low diastereoselectivity. [4]

(4)

CrAreneMech3.png

Enantioselective variants

Enantioselective benzylic functionalization methods use the complexed chromium tri(carbonyl) moiety essentially as a chiral auxiliary. [9] Approach of the functionalizing reagent anti to the ----chromium tri(carbonyl) fragment leads to a single diastereomer of the product complex. After removal of the chromium group with light [5] or an oxidizing agent such as iodine, [6] a nearly enantiopure product remains. See the Scope and Limitations section below for several methods for diastereoselective benzylic functionalization.

Scope and limitations

Enantioselective benzylic functionalization reactions depend on the use of enantiomerically pure, planar chiral chromium arene complexes. This section describes methods for the enantioselective synthesis of planar chiral chromium arene complexes, then outlines methods for functionalization of both sp2- and sp3-hybridized benzylic positions.

Preparation of enantiopure, planar chiral chromium complexes

Enantiopure, planar chiral chromium arene complexes can be synthesized using several strategies. Diastereoselective complexation of a chiral, non-racemic arene to chromium is one such strategy. In the example in equation (5), enantioselective Corey-Itsuno reduction [10] sets up a diastereoselective ligand substitution reaction. After complexation, the alcohol is reduced with triethylsilane. [11]

(5)

CrAreneScope1.png

A second strategy involves enantioselective ortho-lithiation and in situ quenching with an electrophile. Isolation of the lithium arene and subsequent treatment with TMSCl led to lower enantioselectivities. [12]

(6)

CrAreneScope2.png

Site-selective conjugate addition to chiral aryl hydrazone complexes can also be used for the enantioselective formation of planar chiral chromium arenes. Hydride abstraction neutralizes the addition product, and treatment with acid cleaves the hydrazone. [13]

(7)

CrAreneScope3.png

Benzylic functionalization reactions

ortho-Substituted aryl aldehyde complexes undergo diastereoselective nucleophilic addition with organometallic reagents [14] and other nucleophiles. Equation (8) is an example of a diastereoselective Morita-Baylis-Hillman reaction. [15]

(8)

CrAreneScope4.png

Pinacol coupling and the corresponding diamine coupling [16] are possible in the presence of a one-electron reducing agent such as samarium(II) iodide. [17]

(9)

CrAreneScope5.png

Benzylic cations of chromium arene complexes are conformationally stable, and undergo only exo attack to afford SN1 products stereospecifically, with retention of configuration. [8] Propargyl [18] and oxonium [19] cations undergo retentive substitution reactions, and even β carbocations react with a significant degree of retention. [20]

(10)

CrAreneScope6.png

Benzylic anions of chromium arene complexes exhibit similar reactivity to cations. They are also conformationally restricted and undergo substitution reactions with retention of stereochemistry at the benzylic carbon. In the example below, complexation of the pyridine nitrogen to lithium is essential for high stereoselectivity. [21]

(11)

CrAreneScope7.png

Nucleophilic addition to styrenes followed by quenching with an electrophile leads to cis products with essentially complete stereoselectivity. [22]

(12)

CrAreneScope8.png

Diastereoselective reduction of styrenes is possible with samarium(II) iodide. A distant alkene is untouched during this reaction, which provides the reduced alkylarene product in high yield. [23]

(13)

CrAreneScope9.png

Complexation of a haloarene to chromium increases its propensity to undergo oxidative addition. [24] Suzuki cross coupling of a planar chiral chromium haloarene complex with an aryl boronic acid is thus a viable method for the synthesis of axially chiral biaryls. In the example below, the syn isomer is formed in preference to the anti isomer; when R2 is the formyl group, the selectivity reverses. [25]

(14)

CrAreneScope10.png

Tetralones complexed to chromium may be deprotonated without side reactions. Alkylation of the resulting enolate proceeds with complete diastereoselectivity to afford the exo product. [26]

(15)

CrAreneScope11.png

Related Research Articles

The aldol reaction is a reaction that combines two carbonyl compounds to form a new β-hydroxy carbonyl compound.

<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.

In chemistry, stereoselectivity is the property of a chemical reaction in which a single reactant forms an unequal mixture of stereoisomers during a non-stereospecific creation of a new stereocenter or during a non-stereospecific transformation of a pre-existing one. The selectivity arises from differences in steric and electronic effects in the mechanistic pathways leading to the different products. Stereoselectivity can vary in degree but it can never be total since the activation energy difference between the two pathways is finite: both products are at least possible and merely differ in amount. However, in favorable cases, the minor stereoisomer may not be detectable by the analytic methods used.

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

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

A cascade reaction, also known as a domino reaction or tandem reaction, is a chemical process that comprises at least two consecutive reactions such that each subsequent reaction occurs only in virtue of the chemical functionality formed in the previous step. In cascade reactions, isolation of intermediates is not required, as each reaction composing the sequence occurs spontaneously. In the strictest definition of the term, the reaction conditions do not change among the consecutive steps of a cascade and no new reagents are added after the initial step. By contrast, one-pot procedures similarly allow at least two reactions to be carried out consecutively without any isolation of intermediates, but do not preclude the addition of new reagents or the change of conditions after the first reaction. Thus, any cascade reaction is also a one-pot procedure, while the reverse does not hold true. Although often composed solely of intramolecular transformations, cascade reactions can also occur intermolecularly, in which case they also fall under the category of multicomponent reactions.

<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">Petasis reaction</span>

The Petasis reaction is the multi-component reaction of an amine, a carbonyl, and a vinyl- or aryl-boronic acid to form substituted amines.

<span class="mw-page-title-main">Asymmetric induction</span> Preferential formation of one chiral isomer over another in a chemical reaction

Asymmetric induction describes the preferential formation in a chemical reaction of one enantiomer or diastereoisomer over the other as a result of the influence of a chiral feature present in the substrate, reagent, catalyst or environment. Asymmetric induction is a key element in asymmetric synthesis.

<span class="mw-page-title-main">Organoindium chemistry</span> Chemistry of compounds with a carbon to indium bond

Organoindium chemistry is the chemistry of compounds containing In-C bonds. The main application of organoindium chemistry is in the preparation of semiconducting components for microelectronic applications. The area is also of some interest in organic synthesis. Most organoindium compounds feature the In(III) oxidation state, akin to its lighter congeners Ga(III) and B(III).

Organostannane addition reactions comprise the nucleophilic addition of an allyl-, allenyl-, or propargylstannane to an aldehyde, imine, or, in rare cases, a ketone. The reaction is widely used for carbonyl allylation.

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

Electrophilic fluorination is the combination of a carbon-centered nucleophile with an electrophilic source of fluorine to afford organofluorine compounds. Although elemental fluorine and reagents incorporating an oxygen-fluorine bond can be used for this purpose, they have largely been replaced by reagents containing a nitrogen-fluorine bond.

The [2,3]-Wittig rearrangement is the transformation of an allylic ether into a homoallylic alcohol via a concerted, pericyclic process. Because the reaction is concerted, it exhibits a high degree of stereocontrol, and can be employed early in a synthetic route to establish stereochemistry. The Wittig rearrangement requires strongly basic conditions, however, as a carbanion intermediate is essential. [1,2]-Wittig rearrangement is a competitive process.

Nucleophilic epoxidation is the formation of epoxides from electron-deficient double bonds through the action of nucleophilic oxidants. Nucleophilic epoxidation methods represent a viable alternative to electrophilic methods, many of which do not epoxidize electron-poor double bonds efficiently.

Reductions with metal alkoxyaluminium hydrides are chemical reactions that involve either the net hydrogenation of an unsaturated compound or the replacement of a reducible functional group with hydrogen by metal alkoxyaluminium hydride reagents.

Metal-catalyzed cyclopropanations are chemical reactions that result in the formation of a cyclopropane ring from a metal carbenoid species and an alkene. In the Simmons–Smith reaction the metal involved is zinc. Metal carbenoid species can be generated through the reaction of a diazo compound with a transition metal). The intramolecular variant of this reaction was first reported in 1961. Rhodium carboxylate complexes, such as dirhodium tetraacetate, are common catalysts. Enantioselective cyclopropanations have been developed.

Reactions of organocopper reagents involve species containing copper-carbon bonds acting as nucleophiles in the presence of organic electrophiles. Organocopper reagents are now commonly used in organic synthesis as mild, selective nucleophiles for substitution and conjugate addition reactions.

Heteroatom-promoted lateral lithiation is the site-selective replacement of a benzylic hydrogen atom for lithium for the purpose of further functionalization. Heteroatom-containing substituents may direct metalation to the benzylic site closest to the heteroatom or increase the acidity of the ring carbons via an inductive effect.

In organic chemistry, the Baylis–Hillman, Morita–Baylis–Hillman, or MBH reaction is a carbon-carbon bond-forming reaction between an activated alkene and a carbon electrophile in the presence of a nucleophilic catalyst, such as a tertiary amine or phosphine. The product is densely functionalized, joining the alkene at the α-position to a reduced form of the electrophile.

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.

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

Ugi’s amine is a chemical compound named for the chemist who first reported its synthesis in 1970, Ivar Ugi. It is a ferrocene derivative. Since its first report, Ugi’s amine has found extensive use as the synthetic precursor to a large number of metal ligands that bear planar chirality. These ligands have since found extensive use in a variety of catalytic reactions. The compound may exist in either the 1S or 1R isomer, both of which have synthetic utility and are commercially available. Most notably, it is the synthetic precursor to the Josiphos class of ligands.

References

  1. Uemura, M. Org. React. 2006, 67, 217. doi : 10.1002/0471264180.or067.02
  2. E. Peter Kündig (2004). "Synthesis of Transition Metal η6-Arene Complexes". Topics Organomet Chem. Topics in Organometallic Chemistry. 7: 3–20. doi:10.1007/b94489. ISBN   978-3-540-01604-5.
  3. Mahaffy, C. A. L.; Pauson, P. Inorg. Synth.1990, 28, 137.
  4. 1 2 Besançon, J.; Tirouflet, J.; Card, A.; Dusausoy, Y. J. Organomet. Chem.1973, 59, 267.
  5. 1 2 Mishchenko, O. G.; Klementeva, S. V.; Maslennikov, S. V.; Artemov, A. N.; Spirina, I. V. Rus. J. Gen. Chem.2006, 76, 1907.
  6. 1 2 Semmelhack, M. F.; Zask, A. J. Am. Chem. Soc.1983, 105, 2034.
  7. Holmes, J. D.; Jones, D. A. K.; Pettit, R. J. Organomet. Chem.1965, 4, 324.
  8. 1 2 Davies, S. G.; Donohoe, T. J. Synlett1993, 323.
  9. Uemura, M.; Hayashi, Y.; Hayashi, Y. Tetrahedron: Asymmetry, 1993, 4, 2291.
  10. Itsuno, S. Org. React. 1998, 52, 395.
  11. Schmalz, H.-G.; Arnold, M.; Hollander, J.; Bats, J. W. Angew. Chem. Int. Ed. Engl.1994, 33, 109.
  12. Price, D. A.; Simpkins, N. S.; MacLeod, A. M.; Watt, A. P. J. Org. Chem.1994, 59, 1961.
  13. Kündig, E. P.; Liu, R.; Ripa, A. Helv. Chim. Acta1992, 75, 2657.
  14. Bitterwolf, T. E.; Dai, X. J. Organomet. Chem.1992, 440, 103.
  15. Kündig, E. P.; Xu, L. H.; Romanens, P.; Bernardinelli, G. Tetrahedron Lett.1993, 34, 7049.
  16. Davies, S. G.; Donohoe, T. J.; Williams, J. M. J. Pure Appl. Chem.1992, 64, 379.
  17. Taniguchi, N.; Kaneta, N.; Uemura, M. J. Org. Chem.1996, 61, 6088.
  18. Müller, T. J. J.; Netz, A. Tetrahedron Lett.1999, 40, 3145.
  19. Davies, S. G.; Newton, R. F.; Williams, J. M. J. Tetrahedron Lett.1989, 30, 2967.
  20. Merlic, C. A.; Miller, M. M. Organometallics2001, 20, 373.
  21. Davies, S. G.; Shipton, M. R. J. Chem. Soc., Chem. Commun.1990, 1780.
  22. Majdalani, A.; Schmalz, H.-G. Tetrahedron Lett.1997, 38, 4545.
  23. Schmalz, H.-G.; Siegel, S.; Bernicke, D. Tetrahedron Lett.1998, 39, 6683.
  24. Widdowson, D. A.; Wilhelm, R. Chem. Commun.1999, 2211.
  25. Kamikawa, K.; Watanabe, T.; Uemura, M. J. Org. Chem.1996, 61, 1375.
  26. Davies, S. G.; Coote, S. J.; Goodfellow, C. L. In Advances in Metal-Organic Chemistry; Liebeskind, L.S., Ed.; JAI Press: Greenwich, 1991; Vol. 2. pp. 1–48.