Intramolecular Heck reaction

Last updated

The intramolecular Heck reaction (IMHR) in chemistry is the coupling of an aryl or alkenyl halide with an alkene in the same molecule. The reaction may be used to produce carbocyclic or heterocyclic organic compounds with a variety of ring sizes. Chiral palladium complexes can be used to synthesize chiral intramolecular Heck reaction products in non-racemic form. [1]

Contents

Introduction

The Heck reaction is the palladium-catalyzed coupling of an aryl or alkenyl halide with an alkene to form a substituted alkene. [2] Intramolecular variants of the reaction may be used to generate cyclic products containing endo or exo double bonds. Ring sizes produced by the intramolecular Heck reaction range from four to twenty-seven atoms. Additionally, in the presence of a chiral palladium catalyst, the intramolecular Heck reaction may be used to establish tertiary or quaternary stereocenters with high enantioselectivity. [3] A number of tandem reactions, in which the intermediate alkylpalladium complex is intercepted either intra- or intermolecularly before β-hydride elimination, have also been developed. [4]

(1)

HeckGen.png

Mechanism and stereochemistry

The neutral pathway

As shown in Eq. 2, the neutral pathway of the Heck reaction begins with the oxidative addition of the aryl or alkenyl halide into a coordinatively unsaturated palladium(0) complex (typically bound to two phosphine ligands) to give complex I. Dissociation of a phosphine ligand followed by association of the alkene yields complex II, and migratory insertion of the alkene into the carbon-palladium bond establishes the key carbon-carbon bond.

Insertion takes place in a suprafacial fashion, but the dihedral angle between the alkene and palladium-carbon bond during insertion can vary from 0° to ~90°. After insertion, β-hydride elimination affords the product and a palladium(II)-hydrido complex IV, which is reduced by base back to palladium(0). [5]

(2)

HeckMech1.png

The cationic pathway

Most asymmetric Heck reactions employing chiral phosphines proceed by the cationic pathway, which does not require the dissociation of a phosphine ligand. Oxidative addition of an aryl perfluorosulfonate generates a cationic palladium aryl complex V. The mechanism then proceeds as in the neutral case, with the difference that an extra site of coordinative unsaturation exists on palladium throughout the process.

Thus, coordination of the alkene does not require ligand dissociation. Stoichiometric amounts of base are still required to reduce the palladium(II)-hydrido complex VIII back to palladium(0). [6] Silver salts may be used to initiate the cationic pathway in reactions of aryl halides. [7]

(3)

HeckMech2.png

The anionic pathway

Reactions involving palladium(II) acetate and phosphine ligands proceed by a third mechanism, the anionic pathway. [8] Base mediates the oxidation of a phosphine ligand by palladium(II) to a phosphine oxide. Oxidative addition then generates the anionic palladium complex IX. Loss of halide leads to neutral complex X, which undergoes steps analogous to the neutral pathway to regenerate anionic complex IX. A similar anionic pathway is also likely operative in reactions of bulky palladium tri(tert-butyl)phosphine complexes. [9]

(4)

HeckMech3.png

Establishing tertiary or quaternary stereocenters

Asymmetric Heck reactions establish quaternary or tertiary stereocenters. If migratory insertion generates a quaternary center adjacent to the palladium-carbon bond (as in reactions of trisubstituted or 1,1-disubstituted alkenes), β-hydride elimination toward that center is not possible and it is retained in the product. [3] Similarly, β-hydride elimination is not possible if a hydrogen syn to the palladium-carbon bond is not available. Thus, tertiary stereocenters can be established in conformationally restricted systems. [10]

(5)

HeckStereo.png

Scope and limitations

The intramolecular Heck reaction may be used to form rings of a variety of sizes and topologies. β-Hydride elimination need not be the final step of the reaction, and tandem methods have been developed that involve the interception of palladium alkyl intermediates formed after migratory insertion by an additional reactant. This section discusses the most common ring sizes formed by the intramolecular Heck reaction and some of its tandem and asymmetric variants.

5-Exo cyclization, which establishes a five-membered ring with an exocyclic alkene, is the most facile cyclization mode in intramolecular Heck reactions. In this and many other modes of intramolecular Heck cyclization, annulations typically produce a cis ring juncture. [11]

(6)

HeckScope1.png

6-Exo cyclization is also common. The high stability of Heck reaction catalysts permits the synthesis of highly strained compounds at elevated temperatures. In the example below, the arene and alkene must both be in energetically unfavorable axial positions in order to react. [12]

(7)

HeckScope2.png

Endo cyclization is observed most often when small or large rings are involved. For instance, 5-endo cyclization is generally preferred over 4-exo cyclization. [13] The yield of endo product increases with increasing ring size in the synthesis of cycloheptenes, -octenes, and -nonenes. [14]

(8)

HeckScope3.png

Tandem reactions initiated by IMHR have been extensively explored. Palladium alkyl intermediates generated after migratory insertion may undergo a second round of insertion in the presence of a second alkene (either intra- or intermolecular). [15] When dienes are involved in the intramolecular Heck reaction, insertion affords π-allylpalldium intermediates, which may be intercepted by nucleophiles. This idea was applied to a synthesis of (–)-morphine. [16]

(9)

HeckScope4.png

Asymmetric IMHR may establish tertiary or quaternary stereocenters. BINAP is the most commonly chiral ligand used in this context. An interesting application of IMHR is group-selective desymmetrization (enantiotopic group selection), in which the chiral palladium aryl intermediate undergoes insertion predominantly with one of the enantiotopic double bonds. [17]

(10)

HeckScope5.png

Synthetic applications

The high functional group tolerance of the intramolecular Heck reaction allows it to be used at a very late stage in synthetic routes. In a synthesis of (±)-FR900482, IMHR establishes a tricyclic ring system in high yield without disturbing any of the sensitive functionality nearby. [18]

(11)

HeckSynth1.png

Intramolecular Heck reactions have been employed for the construction of complex natural products. An example is the late-stage, macrocyclic ring closure in the total synthesis of the cytotoxic natural product (–)-Mandelalide A. [19] In another example a fully intramolecular tandem Heck reaction is used in a synthesis of (–)-scopadulcic acid. A 6-exo cyclization sets the quaternary center and provides a neopentyl σ-palladium intermediate, which undergoes a 5-exo reaction to provide the ring system. [20]

(12)

HeckSynth2.png

Comparison with other methods

The closest competing method to IMHR is radical cyclization. [21] Radical cyclizations are often reductive, which can cause undesired side reactions to occur if sensitive substrates are employed. The IMHR, on the other hand, can be run under reductive conditions if desired. [22] Unlike the IMHR, radical cyclization does not require the coupling of two sp2-hybridized carbons. In some cases, the results of radical cyclization and IMHR are complementary. [23]

Experimental conditions and procedure

Typical conditions

A variety of experimental concerns exist for IMHR reactions. Although most of the common Pd(0) catalysts are commercially available (Pd(PPh3)4, Pd2(dba)3, and derivatives), they may also be prepared by simple, high-yielding procedures. [24] Palladium(II) acetate is cheap and may be reduced in situ to palladium(0) with phosphine. Three equivalents of phosphine per equivalent of palladium acetate are commonly used; these conditions generate Pd(PR3)2 as the active catalyst. Bidentate phosphine ligands are common in asymmetric reactions to enhance stereoselectivity.

A wide variety of bases may be used, and the base is often employed in excess. Potassium carbonate is the most common base employed, and inorganic bases are generally used more often than organic bases. A number of additives have also been identified for the Heck reaction—silver salts may be used to drive the reaction down the cationic pathway, and halide salts may be used to convert aryl triflates via the neutral pathway. Alcohols have been shown to enhance catalyst stability in some cases, [25] and acetate salts are beneficial in reactions following the anionic pathway. [8]

Example procedure

(13)

HeckEx.png

A solution of the amide (0.365 g, 0.809 mmol), Pd(PPh3)4 (0.187 g, 0.162 mmol), and triethylamine (1.12 mL, 8.08 mmol) in MeCN (8 mL) in a sealed tube was heated slowly to 120°. After stirring for 4 hours, the reaction mixture was cooled to room temperature, and the solvent was evaporated. The residue was chromatographed (loaded with CH2Cl2) to give the title product 316 (0.270 g, 90%) as a colorless oil; Rf 0.42 (EtOAc/petroleum ether 10:1); [α]22D +14.9 (c, 1.0, CHCl3); IR 3027, 2930, 1712, 1673, 1608, 1492, 1343, 1248 cm−1; 1H NMR (400 MHz) δ 7.33–7.21 (m, 6 H), 7.07 (dd, J = 7.3, 16.4 Hz, 1 H), 7.00 (t, J = 7.5 Hz, 1 H), 6.77 (d, J = 7.7 Hz, 1 H), 6.30 (dd, J = 8.7, 11.4 Hz, 1 H), 5.32 (d, J = 15.7 Hz, 1 H), 5.04 (s, 1 H), 4.95 (s, 1 H), 4.93 (d, J = 11.1 Hz, 1 H), 4.17 (s, 1 H), 3.98 (d, J = 15.7 Hz, 1 H), 3.62 (d, J = 8.7 Hz, 1 H), 3.17 (s, 3 H), 2.56 (dd, J = 3.5, 15.5 Hz, 1 H), 2.06 (dd, J = 2.8, 15.5 Hz, 1 H); 13C NMR (100 MHz) δ 177.4, 172.9, 147.8, 142.2, 136.5, 132.2, 131.6, 128.8, 128.4, 128.2, 127.7, 127.1, 123.7, 122.9, 107.9, 105.9, 61.0, 54.7, 49.9, 44.4, 38.2, 26.4; HRMS Calcd. for C24H22N2O2: 370.1681. Found: 370.1692.

Related Research Articles

The Heck reaction is the chemical reaction of an unsaturated halide with an alkene in the presence of a base and a palladium catalyst to form a substituted alkene. It is named after Tsutomu Mizoroki and Richard F. Heck. Heck was awarded the 2010 Nobel Prize in Chemistry, which he shared with Ei-ichi Negishi and Akira Suzuki, for the discovery and development of this reaction. This reaction was the first example of a carbon-carbon bond-forming reaction that followed a Pd(0)/Pd(II) catalytic cycle, the same catalytic cycle that is seen in other Pd(0)-catalyzed cross-coupling reactions. The Heck reaction is a way to substitute alkenes.

The Stille reaction is a chemical reaction widely used in organic synthesis. The reaction involves the coupling of two organic groups, one of which is carried as an organotin compound. A variety of organic electrophiles provide the other coupling partner. The Stille reaction is one of many palladium-catalyzed coupling reactions.

The Sonogashira reaction is a cross-coupling reaction used in organic synthesis to form carbon–carbon bonds. It employs a palladium catalyst as well as copper co-catalyst to form a carbon–carbon bond between a terminal alkyne and an aryl or vinyl halide.

Organopalladium chemistry is a branch of organometallic chemistry that deals with organic palladium compounds and their reactions. Palladium is often used as a catalyst in the reduction of alkenes and alkynes with hydrogen. This process involves the formation of a palladium-carbon covalent bond. Palladium is also prominent in carbon-carbon coupling reactions, as demonstrated in tandem reactions.

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.

In organic chemistry, the Buchwald–Hartwig amination is a chemical reaction for the synthesis of carbon–nitrogen bonds via the palladium-catalyzed coupling reactions of amines with aryl halides. Although Pd-catalyzed C-N couplings were reported as early as 1983, Stephen L. Buchwald and John F. Hartwig have been credited, whose publications between 1994 and the late 2000s established the scope of the transformation. The reaction's synthetic utility stems primarily from the shortcomings of typical methods for the synthesis of aromatic C−N bonds, with most methods suffering from limited substrate scope and functional group tolerance. The development of the Buchwald–Hartwig reaction allowed for the facile synthesis of aryl amines, replacing to an extent harsher methods while significantly expanding the repertoire of possible C−N bond formation.

In organic chemistry, the Kumada coupling is a type of cross coupling reaction, useful for generating carbon–carbon bonds by the reaction of a Grignard reagent and an organic halide. The procedure uses transition metal catalysts, typically nickel or palladium, to couple a combination of two alkyl, aryl or vinyl groups. The groups of Robert Corriu and Makoto Kumada reported the reaction independently in 1972.

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

In organometallic chemistry, a metallacycle is a derivative of a carbocyclic compound wherein a metal has replaced at least one carbon center; this is to some extent similar to heterocycles. Metallacycles appear frequently as reactive intermediates in catalysis, e.g. olefin metathesis and alkyne trimerization. In organic synthesis, directed ortho metalation is widely used for the functionalization of arene rings via C-H activation. One main effect that metallic atom substitution on a cyclic carbon compound is distorting the geometry due to the large size of typical metals.

In organometallic chemistry, a migratory insertion is a type of reaction wherein two ligands on a metal complex combine. It is a subset of reactions that very closely resembles the insertion reactions, and both are differentiated by the mechanism that leads to the resulting stereochemistry of the products. However, often the two are used interchangeably because the mechanism is sometimes unknown. Therefore, migratory insertion reactions or insertion reactions, for short, are defined not by the mechanism but by the overall regiochemistry wherein one chemical entity interposes itself into an existing bond of typically a second chemical entity e.g.:

Hydroacylation is a type of organic reaction in which an alkene is inserted into the a formyl C-H bond. The product is a ketone. The reaction requires a metal catalyst. It is almost invariably practiced as an intramolecular reaction using homogeneous catalysts, often based on rhodium phosphines.

Radical cyclization reactions are organic chemical transformations that yield cyclic products through radical intermediates. They usually proceed in three basic steps: selective radical generation, radical cyclization, and conversion of the cyclized radical to product.

Trimethylenemethane cycloaddition is the formal [3+2] annulation of trimethylenemethane (TMM) derivatives to two-atom pi systems. Although TMM itself is too reactive and unstable to be stored, reagents which can generate TMM or TMM synthons in situ can be used to effect cycloaddition reactions with appropriate electron acceptors. Generally, electron-deficient pi bonds undergo cyclization with TMMs more easily than electron-rich pi bonds.

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.

<span class="mw-page-title-main">Heck–Matsuda reaction</span>

The Heck–Matsuda (HM) reaction is an organic reaction and a type of palladium catalysed arylation of olefins that uses arenediazonium salts as an alternative to aryl halides and triflates.

The Tsuji–Trost reaction is a palladium-catalysed substitution reaction involving a substrate that contains a leaving group in an allylic position. The palladium catalyst first coordinates with the allyl group and then undergoes oxidative addition, forming the π-allyl complex. This allyl complex can then be attacked by a nucleophile, resulting in the substituted product.

Decarboxylative cross coupling reactions are chemical reactions in which a carboxylic acid is reacted with an organic halide to form a new carbon-carbon bond, concomitant with loss of CO2. Aryl and alkyl halides participate. Metal catalyst, base, and oxidant are required.

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

Phosphinooxazolines are a class of chiral ligands used in asymmetric catalysis. Their complexes are particularly effective at generating single enatiomers in reactions involving highly symmetric transition states, such as allylic substitutions, which are typically difficult to perform stereoselectively. The ligands are bidentate and have been shown to be hemilabile with the softer P‑donor being more firmly bound than the harder N‑donor.

<span class="mw-page-title-main">John P. Wolfe</span>

John Perry Wolfe is an American chemist and a professor of chemistry at the University of Michigan. He is best known for palladium-catalyzed C-C and C-N bond formation reactions. He was also one of the key scientists in the development of Buchwald ligands, one of which is appropriately named "JohnPhos" after him. Wolfe has taught at the University of Michigan since 2002.

Heterobimetallic catalysis is an approach to catalysis that employs two different metals to promote a chemical reaction. Included in this definition are cases where: 1) each metal activates a different substrate, 2) both metals interact with the same substrate, and 3) only one metal directly interacts with the substrate(s), while the second metal interacts with the first.

<span class="mw-page-title-main">Mizoroki-Heck vs. Reductive Heck</span>

The Mizoroki−Heck coupling of aryl halides and alkenes to form C(sp2)–C(sp2) bonds has become a staple transformation in organic synthesis, owing to its broad functional group compatibility and varied scope. In stark contrast, the palladium-catalyzed reductive Heck reaction has received considerably less attention, despite the fact that early reports of this reaction date back almost half a century. From the perspective of retrosynthetic logic, this transformation is highly enabling because it can forge alkyl–aryl linkages from widely available alkenes, rather than from the less accessible and/or more expensive alkyl halide or organometallic C(sp3) synthons that are needed in a classical aryl/alkyl cross-coupling.

References

  1. Link, J. T. Org. React. 2002, 60, 157. doi : 10.1002/0471264180.or060.02
  2. Beletskaya, I.; Cheprakov, A. Chem. Rev.2000100, 3009.
  3. 1 2 Overman, L. E. Pure Appl. Chem.1994, 66, 1423.
  4. Larock, R. C.; Lee, N. H. J. Org. Chem.1991, 56, 6253.
  5. Amatore, C.; Azzabi, M.; Jutand, A. J. Am. Chem. Soc.1991, 113, 8375.
  6. Cabri, W.; Candiani, I.; DeBernardinis, S.; Francalanci, F.; Penco, S.; Santi, R. J. Org. Chem.1991, 56, 5796.
  7. Karabelas, K.; Westerlund, C.; Hallberg, A. J. Org. Chem.1985, 50, 3896.
  8. 1 2 Amatore, C.; Jutand, A. Acc. Chem. Res.2000, 33, 314.
  9. Carrow, B.; Hartwig, J. F. J. Am. Chem. Soc., 2010, 132, 79.
  10. Honzawa, H.; Mizutani, T.; Shibasaki, M. Tetrahedron Lett.1999, 40, 311.
  11. Larock, R. C.; Song, H.; Baker, B. E.; Gong, W. H. Tetrahedron Lett.1988, 29, 2919.
  12. Wünsch, B.; Diekmann, H.; Höfner, G. Tetrahedron: Asymmetry1995, 6, 1527.
  13. Koerber-Plé, K.; Massiott, G. Synlett1994, 759.
  14. Gibson, S. E.; Guillo, N.; Middleton, R. J.; Thuillez, A.; Tozer, M. J. J. Chem. Soc., Perkin Trans. 11997, 447.
  15. Grigg, R.; Sridharan, V. Tetrahedron Lett.1992, 33, 7965.
  16. Hong, C. Y.; Overman, L. E. Tetrahedron Lett.1994, 35, 3453.
  17. Sato, Y.; Mori, M.; Shibasaki, M. Tetrahedron Lett.1992, 33, 2593.
  18. Schkeryantz, J. M.; Danishefsky, S. J. J. Am. Chem. Soc.1995, 117, 4722.
  19. Nguyen, M. H.; Imanishi, M.; Kurogi, T.; Smith, A. B. III J. Am. Chem. Soc.2016, 138, 3675.
  20. Fox, M. E.; Li, C.; Marino, J. P.; Overman, L. E. J. Am. Chem. Soc.1999, 121, 5467.
  21. Giese, B.; Kopping, B.; Göbel, T.; Dickhaut, J.; Thoma, G.; Kulicke, K. j.; Trach, F. Org. React.1997, 48, 301.
  22. Burns, B.; Grigg, R.; Ratananukul, P.; Sridharan, V.; Stevenson, P.; Worakun, T. Tetrahedron Lett.1988, 29, 4329.
  23. Ishibashi, H.; Ito, K.; Hirano, T.; Tabuchi, M.; Ikeda, M. Tetrahedron1993, 49, 4173.
  24. Hegedus, L. S. In Organometallics in Synthesis; Schlosser, M., Ed.; John Wiley and Sons Ltd.: Chichester, England, 1994; p. 383.
  25. Ohrai, K.; Kondo, K.; Sodeoka, M.; Shibasaki, M. J. Am. Chem. Soc.1994, 116, 11737.