fbpx
Wikipedia

Stille reaction

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 (also known as organostannanes). A variety of organic electrophiles provide the other coupling partner. The Stille reaction is one of many palladium-catalyzed coupling reactions.[1][2][3]

Stille reaction
Named after John Kenneth Stille
Reaction type Coupling reaction
Identifiers
Organic Chemistry Portal stille-coupling
RSC ontology ID RXNO:0000035
  • : Allyl, alkenyl, aryl, benzyl, acyl
  • : halides (Cl, Br, I), pseudohalides (OTf, OPO(OR)2), OAc

The R1 group attached to the trialkyltin is normally sp2-hybridized, including vinyl, and aryl groups.

These organostannanes are also stable to both air and moisture, and many of these reagents either are commercially available or can be synthesized from literature precedent. However, these tin reagents tend to be highly toxic. X is typically a halide, such as Cl, Br, or I, yet pseudohalides such as triflates and sulfonates and phosphates can also be used.[4][5] Several reviews have been published.[6][2][7][8][9][10][11][12][13][14][15][excessive citations]

History edit

The first example of a palladium catalyzed coupling of aryl halides with organotin reagents was reported by Colin Eaborn in 1976.[16] This reaction yielded from 7% to 53% of diaryl product. This process was expanded to the coupling of acyl chlorides with alkyl-tin reagents in 1977 by Toshihiko Migita, yielding 53% to 87% ketone product.[17]

 
First reactions of organotin reagents

In 1977, Migita published further work on the coupling of allyl-tin reagents with both aryl (C) and acyl (D) halides. The greater ability of allyl groups to migrate to the palladium catalyst allowed the reactions to be performed at lower temperatures. Yields for aryl halides ranged from 4% to 100%, and for acyl halides from 27% to 86%.[18][19] Reflecting the early contributions of Migita and Kosugi, the Stille reaction is sometimes called the Migita–Kosugi–Stille coupling.

 
First reactions of organotin reagents

John Kenneth Stille subsequently reported the coupling of a variety of alkyl tin reagents in 1978 with numerous aryl and acyl halides under mild reaction conditions with much better yields (76%–99%).[18][20] Stille continued his work in the 1980s on the synthesis of a multitude of ketones using this broad and mild process and elucidated a mechanism for this transformation.[21][22]

 
First reactions of organotin reagents

By the mid-1980s, over 65 papers on the topic of coupling reactions involving tin had been published, continuing to explore the substrate scope of this reaction. While initial research in the field focused on the coupling of alkyl groups, most future work involved the much more synthetically useful coupling of vinyl, alkenyl, aryl, and allyl organostannanes to halides. Due to these organotin reagent's stability to air and their ease of synthesis, the Stille reaction became common in organic synthesis.[8]

Mechanism edit

The mechanism of the Stille reaction has been extensively studied.[11][23] The catalytic cycle involves an oxidative addition of a halide or pseudohalide (2) to a palladium catalyst (1), transmetalation of 3 with an organotin reagent (4), and reductive elimination of 5 to yield the coupled product (7) and the regenerated palladium catalyst (1).[24]

 
Catalytic cycle of the Stille Reaction

However, the detailed mechanism of the Stille coupling is extremely complex and can occur via numerous reaction pathways. Like other palladium-catalyzed coupling reactions, the active palladium catalyst is believed to be a 14-electron Pd(0) complex, which can be generated in a variety of ways. Use of an 18- or 16- electron Pd(0) source Pd(PPh3)4, Pd(dba)2 can undergo ligand dissociation to form the active species. Second, phosphines can be added to ligandless palladium(0). Finally, as pictured, reduction of a Pd(II) source (8) (Pd(OAc)2, PdCl2(MeCN)2, PdCl2(PPh3)2, BnPdCl(PPh3)2, etc.) by added phosphine ligands or organotin reagents is also common[6]

Oxidative addition edit

Oxidative addition to the 14-electron Pd(0) complex is proposed. This process gives a 16-electron Pd(II) species. It has been suggested that anionic ligands, such as OAc, accelerate this step by the formation of [Pd(OAc)(PR3)n], making the palladium species more nucleophillic.[11][25] In some cases, especially when an sp3-hybridized organohalide is used, an SN2 type mechanism tends to prevail, yet this is not as commonly seen in the literature.[11][25] However, despite normally forming a cis-intermediate after a concerted oxidative addition, this product is in rapid equilibrium with its trans-isomer.[26][27]

 
Cis/Trans Isomerization

There are multiple reasons why isomerization is favored here. First, a bulky ligand set is usually used in these processes, such as phosphines, and it is highly unfavorable for them to adopt a cis orientation relative to each other, resulting in isomerization to the more favorable trans product.[26][27] An alternative explanation for this phenomenon, dubbed antisymbiosis or transphobia, is by invocation of the sdn model.[24][28] Under this theory, palladium is a hypervalent species. Hence R1 and the trans ligand, being trans to each other, will compete with one palladium orbital for bonding. This 4-electron 3-center bond is weakest when two strong donating groups are present, which heavily compete for the palladium orbital. Relative to any ligand normally used, the C-donor R1 ligand has a much higher trans effect. This trans influence is a measure of how competitive ligands trans to each other will compete for palladium's orbital. The usual ligand set, phosphines, and C-donors (R1) are both soft ligands, meaning that they will form strong bonds to palladium, and heavily compete with each other for bonding.[29][30] Since halides or pseudohalides are significantly more electronegative, their bonding with palladium will be highly polarized, with most of the electron density on the X group, making them low trans effect ligands. Hence, it will be highly favorable for R1 to be trans to X, since the R1 group will be able to form a stronger bond to the palladium.[24][28][30]

 
sd^n model for cis/trans isomers

Transmetallation edit

The transmetallation of the trans intermediate from the oxidative addition step is believed to proceed via a variety of mechanisms depending on the substrates and conditions. The most common type of transmetallation for the Stille coupling involves an associative mechanism. This pathway implies that the organostannane, normally a tin atom bonded to an allyl, alkenyl, or aryl group, can coordinate to the palladium via one of these double bonds. This produces a fleeting pentavalent, 18-electron species, which can then undergo ligand detachment to form a square planar complex again. Despite the organostannane being coordinated to the palladium through the R2 group, R2 must be formally transferred to the palladium (the R2-Sn bond must be broken), and the X group must leave with the tin, completing the transmetalation. This is believed to occur through two mechanisms.[31]

First, when the organostannane initially adds to the trans metal complex, the X group can coordinate to the tin, in addition to the palladium, producing a cyclic transition state. Breakdown of this adduct results in the loss of R3Sn-X and a trivalent palladium complex with R1 and R2 present in a cis relationship. Another commonly seen mechanism involves the same initial addition of the organostannane to the trans palladium complex as seen above; however, in this case, the X group does not coordinate to the tin, producing an open transition state. After the α-carbon relative to tin attacks the palladium, the tin complex will leave with a net positive charge. In the scheme below, please note that the double bond coordinating to tin denotes R2, so any alkenyl, allyl, or aryl group. Furthermore, the X group can dissociate at any time during the mechanism and bind to the Sn+ complex at the end. Density functional theory calculations predict that an open mechanism will prevail if the 2 ligands remain attached to the palladium and the X group leaves, while the cyclic mechanism is more probable if a ligand dissociates prior to the transmetalation. Hence, good leaving groups such as triflates in polar solvents favor the cyclic transition state, while bulky phosphine ligands will favor the open transition state.[31]

 
The two mechanisms, cyclic and open, of transmetallation in the Stille reaction

A less common pathway for transmetalation is through a dissociative or solvent assisted mechanism. Here, a ligand from the tetravalent palladium species dissociates, and a coordinating solvent can add onto the palladium. When the solvent detaches, to form a 14-electron trivalent intermediate, the organostannane can add to the palladium, undergoing an open or cyclic type process as above.[31]

Reductive elimination step edit

In order for R1-R2 to reductively eliminate, these groups must occupy mutually cis coordination sites. Any trans-adducts must therefore isomerize to the cis intermediate or the coupling will be frustrated. A variety of mechanisms exist for reductive elimination and these are usually considered to be concerted.[11][32][33]

First, the 16-electron tetravalent intermediate from the transmetalation step can undergo unassisted reductive elimination from a square planar complex. This reaction occurs in two steps: first, the reductive elimination is followed by coordination of the newly formed sigma bond between R1 and R2 to the metal, with ultimate dissociation yielding the coupled product.[11][32][33]

 
Concerted reductive elimination for the Stille reaction

The previous process, however, is sometimes slow and can be greatly accelerated by dissociation of a ligand to yield a 14-electron T shaped intermediate. This intermediate can then rearrange to form a Y-shaped adduct, which can undergo faster reductive elimination.[11][32][33]

 
Dissociative reductive elimination for the Stille reaction

Finally, an extra ligand can associate to the palladium to form an 18-electron trigonal bipyramidal structure, with R1 and R2 cis to each other in equatorial positions. The geometry of this intermediate makes it similar to the Y-shaped above.[11][32][33]

 
Associative reductive elimination for the Stille reaction

The presence of bulky ligands can also increase the rate of elimination. Ligands such as phosphines with large bite angles cause steric repulsion between L and R1 and R2, resulting in the angle between L and the R groups to increase and the angle between R1 and R2 to hence decrease, allowing for quicker reductive elimination.[11][24]

 
Cis-reductive elimination in the Stille reaction

Kinetics edit

The rate at which organostannanes transmetalate with palladium catalysts is shown below. Sp2-hybridized carbon groups attached to tin are the most commonly used coupling partners, and sp3-hybridized carbons require harsher conditions and terminal alkynes may be coupled via a C-H bond through the Sonogashira reaction.

 
Relative rates of the Stille reaction

As the organic tin compound, a trimethylstannyl or tributylstannyl compound is normally used. Although trimethylstannyl compounds show higher reactivity compared with tributylstannyl compounds and have much simpler 1H-NMR spectra, the toxicity of the former is much larger.[34]

Optimizing which ligands are best at carrying out the reaction with high yield and turnover rate can be difficult. This is because the oxidative addition requires an electron rich metal, hence favoring electron donating ligands. However, an electron deficient metal is more favorable for the transmetalation and reductive elimination steps, making electron withdrawing ligands the best here. Therefore, the optimal ligand set heavily depends on the individual substrates and conditions used. These can change the rate determining step, as well as the mechanism for the transmetalation step.[35]

Normally, ligands of intermediate donicity, such as phosphines, are utilized. Rate enhancements can be seen when moderately electron-poor ligands, such as tri-2-furylphosphine or triphenylarsenine are used. Likewise, ligands of high donor number can slow down or inhibit coupling reactions.[35][36]

These observations imply that normally, the rate-determining step for the Stille reaction is transmetalation.[36]

Additives edit

The most common additive to the Stille reaction is stoichiometric or co-catalytic copper(I), specifically copper iodide, which can enhance rates up by >103 fold. It has been theorized that in polar solvents copper transmetalate with the organostannane. The resulting organocuprate reagent could then transmetalate with the palladium catalyst. Furthermore, in ethereal solvents, the copper could also facilitate the removal of a phosphine ligand, activating the Pd center.[9][37][38][39][40]

Lithium chloride has been found to be a powerful rate accelerant in cases where the X group dissociates from palladium (i.e. the open mechanism). The chloride ion is believed to either displace the X group on the palladium making the catalyst more active for transmetalation or by coordination to the Pd(0) adduct to accelerate the oxidative addition. Also, LiCl salt enhances the polarity of the solvent, making it easier for this normally anionic ligand (–Cl, –Br, –OTf, etc.) to leave. This additive is necessary when a solvent like THF is used; however, utilization of a more polar solvent, such as NMP, can replace the need for this salt additive. However, when the coupling's transmetalation step proceeds via the cyclic mechanism, addition of lithium chloride can actually decrease the rate. As in the cyclic mechanism, a neutral ligand, such as phosphine, must dissociate instead of the anionic X group.[10][41]

Finally, sources of fluoride ions, such as cesium fluoride, also effect on the catalytic cycle. First, fluoride can increase the rates of reactions of organotriflates, possibly by the same effect as lithium chloride. Furthermore, fluoride ions can act as scavengers for tin byproducts, making them easier to remove via filtration.[39]

Competing side reactions edit

The most common side reactivity associated with the Stille reaction is homocoupling of the stannane reagents to form an R2-R2 dimer. It is believed to proceed through two possible mechanisms. First, reaction of two equivalents of organostannane with the Pd(II) precatalyst will yield the homocoupled product after reductive elimination. Second, the Pd(0) catalyst can undergo a radical process to yield the dimer. The organostannane reagent used is traditionally tetravalent at tin, normally consisting of the sp2-hybridized group to be transferred and three "non-transferable" alkyl groups. As seen above, alkyl groups are normally the slowest at migrating onto the palladium catalyst.[10]

 
Homocoupling and transfer of "inert" ligands

It has also been found that at temperatures as low as 50 °C, aryl groups on both palladium and a coordinated phosphine can exchange. While normally not detected, they can be a potential minor product in many cases.[10]

 
Aryl transfer through phosphines

Finally, a rather rare and exotic side reaction is known as cine substitution. Here, after initial oxidative addition of an aryl halide, this Pd-Ar species can insert across a vinyl tin double bond. After β-hydride elimination, migratory insertion, and protodestannylation, a 1,2-disubstituted olefin can be synthesized.[10]

 
Cine substitution

Numerous other side reactions can occur, and these include E/Z isomerization, which can potentially be a problem when an alkenylstannane is utilized. The mechanism of this transformation is currently unknown. Normally, organostannanes are quite stable to hydrolysis, yet when very electron-rich aryl stannanes are used, this can become a significant side reaction.[10]

Scope edit

Electrophile edit

Vinyl halides are common coupling partners in the Stille reaction, and reactions of this type are found in numerous natural product total syntheses. Normally, vinyl iodides and bromides are used. Vinyl chlorides are insufficiently reactive toward oxidative addition to Pd(0). Iodides are normally preferred: they will typically react faster and under milder conditions than will bromides. This difference is demonstrated below by the selective coupling of a vinyl iodide in the presence of a vinyl bromide.[10]

 
Vinyl Iodide reacts faster than vinyl bromide

Normally, the stereochemistry of the alkene is retained throughout the reaction, except under harsh reaction conditions. A variety of alkenes may be used, and these include both α- and β-halo-α,β unsaturated ketones, esters, and sulfoxides (which normally need a copper (I) additive to proceed), and more (see example below).[42] Vinyl triflates are also sometimes used. Some reactions require the addition of LiCl and others are slowed down, implying that two mechanistic pathways are present.[10]

 
Addition to an alpha, beta unsaturated alkene

Another class of common electrophiles are aryl and heterocyclic halides. As for the vinyl substrates, bromides and iodides are more common despite their greater expense. A multitude of aryl groups can be chosen, including rings substituted with electron donating substituents, biaryl rings, and more. Halogen-substituted heterocycles have also been used as coupling partners, including pyridines, furans, thiophenes, thiazoles, indoles, imidazoles, purines, uracil, cytosines, pyrimidines, and more (See below for table of heterocycles; halogens can be substituted at a variety of positions on each).[10]

 
Variety of heterocycles which can undergo addition

Below is an example of the use of Stille coupling to build complexity on heterocycles of nucleosides, such as purines.[43]

 
Addition to a heterocycle

Aryl triflates and sulfonates are also couple to a wide variety of organostannane reagents. Triflates tend to react comparably to bromides in the Stille reaction.[10]

Acyl chlorides are also used as coupling partners and can be used with a large range of organostannane, even alkyl-tin reagents, to produce ketones (see example below).[44] However, it is sometimes difficult to introduce acyl chloride functional groups into large molecules with sensitive functional groups. An alternative developed to this process is the Stille-carbonylative cross-coupling reaction, which introduces the carbonyl group via carbon monoxide insertion.[10]

 
Acyl chlorides can be used as well

Allylic, benzylic, and propargylic halides can also be coupled. While commonly employed, allylic halides proceed via an η3 transition state, allowing for coupling with the organostannane at either the α or γ position, occurring predominantly at the least substituted carbon (see example below).[45] Alkenyl epoxides (adjacent epoxides and alkenes) can also undergo this same coupling through an η3 transition state as, opening the epoxide to an alcohol. While allylic and benzylic acetates are commonly used, propargylic acetates are unreactive with organostannanes.[10]

 
allylic bromides will form an heta-3 complex.

Stannane edit

Organostannane reagents are common. Several are commercially available.[46] Stannane reagents can be synthesized by the reaction of a Grignard or organolithium reagent with trialkyltin chlorides. For example, vinyltributyltin is prepared by the reaction of vinylmagnesium bromide with tributyltin chloride.[47] Hydrostannylation of alkynes or alkenes provides many derivatives. Organotin reagents are air and moisture stable. Some reactions can even take place in water.[48] They can be purified by chromatography. They are tolerant to most functional groups. Some organotin compounds are heavily toxic, especially trimethylstannyl derivatives.[10]

The use of vinylstannane, or alkenylstannane reagents is widespread.[10] In regards to limitations, both very bulky stannane reagents and stannanes with substitution on the α-carbon tend to react sluggishly or require optimization. For example, in the case below, the α-substituted vinylstannane only reacts with a terminal iodide due to steric hindrance.[49]

 
Stannae 1

Arylstannane reagents are also common and both electron donating and electron withdrawing groups actually increase the rate of the transmetalation. This again implies that two mechanisms of transmetalation can occur. The only limitation to these reagents are substituents at the ortho-position as small as methyl groups can decrease the rate of reaction. A wide variety of heterocycles (see Electrophile section) can also be used as coupling partners (see example with a thiazole ring below).[10][50]

 
Regioselective coupling of a heterocyclic-stannae with an aryl bromide
 
Coupling of stannane to acyl chloride

Alkynylstannanes, the most reactive of stannanes, have also been used in Stille couplings. They are not usually needed as terminal alkynes can couple directly to palladium catalysts through their C-H bond via Sonogashira coupling. Allylstannanes have been reported to have worked, yet difficulties arise, like with allylic halides, with the difficulty in control regioselectivity for α and γ addition. Distannane and acyl stannane reagents have also been used in Stille couplings.[10]

Applications edit

The Stille reaction has been used in the synthesis of a variety of polymers.[51][52][53] However, the most widespread use of the Stille reaction is its use in organic syntheses, and specifically, in the synthesis of natural products.

Natural product total synthesis edit

Larry Overman's 19-step enantioselective total synthesis of quadrigemine C involves a double Stille cross metathesis reaction.[6][54] The complex organostannane is coupled onto two aryl iodide groups. After a double Heck cyclization, the product is achieved.

 
Total Synthesis of Quadrigemine C

Panek's 32 step enantioselective total synthesis of ansamycin antibiotic (+)-mycotrienol makes use of a late stage tandem Stille type macrocycle coupling. Here, the organostannane has two terminal tributyl tin groups attacked to an alkene. This organostannane "stitches" the two ends of the linear starting material into a macrocycle, adding the missing two methylene units in the process. After oxidation of the aromatic core with ceric ammonium nitrate (CAN) and deprotection with hydrofluoric acid yields the natural product in 54% yield for the 3 steps.[6][55]

 
Total synthesis of mycotrienol

Stephen F. Martin and coworkers' 21 step enantioselective total synthesis of the manzamine antitumor alkaloid Ircinal A makes use of a tandem one-pot Stille/Diels-Alder reaction. An alkene group is added to vinyl bromide, followed by an in situ Diels-Alder cycloaddition between the added alkene and the alkene in the pyrrolidine ring.[6][56]

 
Total synthesis of ircinal A

Numerous other total syntheses utilize the Stille reaction, including those of oxazolomycin,[57] lankacidin C,[58] onamide A,[59] calyculin A,[60] lepicidin A,[61] ripostatin A,[62] and lucilactaene.[6][63] The image below displays the final natural product, the organohalide (blue), the organostannane (red), and the bond being formed (green and circled). From these examples, it is clear that the Stille reaction can be used both at the early stages of the synthesis (oxazolomycin and calyculin A), at the end of a convergent route (onamide A, lankacidin C, ripostatin A), or in the middle (lepicidin A and lucilactaene). The synthesis of ripostatin A features two concurrent Stille couplings followed by a ring-closing metathesis. The synthesis of lucilactaene features a middle subunit, having a borane on one side and a stannane on the other, allowing for Stille reactionfollowed by a subsequent Suzuki coupling.

 
A variety of total syntheses which make use of the Stille reaction

Variations edit

In addition to performing the reaction in a variety of organic solvents, conditions have been devised which allow for a broad range of Stille couplings in aqueous solvent.[14]

In the presence of Cu(I) salts, palladium-on-carbon has been shown to be an effective catalyst.[64][65]

In the realm of green chemistry a Stille reaction is reported taking place in a low melting and highly polar mixture of a sugar such as mannitol, a urea such as dimethylurea and a salt such as ammonium chloride[66] .[67] The catalyst system is tris(dibenzylideneacetone)dipalladium(0) with triphenylarsine:

 
A Stille reaction variation: coupling of phenyliodide and tetramethyltin

Stille–carbonylative cross-coupling edit

A common alteration to the Stille coupling is the incorporation of a carbonyl group between R1 and R2, serving as an efficient method to form ketones. This process is extremely similar to the initial exploration by Migita and Stille (see History) of coupling organostannane to acyl chlorides. However, these moieties are not always readily available and can be difficult to form, especially in the presence of sensitive functional groups. Furthermore, controlling their high reactivity can be challenging. The Stille-carbonylative cross-coupling employs the same conditions as the Stille coupling, except with an atmosphere of carbon monoxide (CO) being used. The CO can coordinate to the palladium catalyst (9) after initial oxidative addition, followed by CO insertion into the Pd-R1 bond (10), resulting in subsequent reductive elimination to the ketone (12). The transmetalation step is normally the rate-determining step.[6]

 
Catalytic cycle of the Stille-carbonylative cross-coupling

Larry Overman and coworkers make use of the Stille-carbonylative cross-coupling in their 20-step enantioselective total synthesis of strychnine. The added carbonyl is later converted to a terminal alkene via a Wittig reaction, allowing for the key tertiary nitrogen and the pentacyclic core to be formed via an aza-Cope-Mannich reaction.[6][68]

 
Total synthesis of strychnine

Giorgio Ortar et al. explored how the Stille-carbonylative cross-coupling could be used to synthesize benzophenone phosphores. These were embedded into 4-benzoyl-L-phenylalanine peptides and used for their photoaffinity labelling properties to explore various peptide-protein interactions.[6][69]

 
Synthesis of phosphores

Louis Hegedus' 16-step racemic total synthesis of Jatraphone involved a Stille-carbonylative cross-coupling as its final step to form the 11-membered macrocycle. Instead of a halide, a vinyl triflate is used there as the coupling partner.[6][70]

 
Total synthesis of Jatraphone

Stille–Kelly coupling edit

Using the seminal publication by Eaborn in 1976, which forms arylstannanes from arylhalides and distannanes, T. Ross Kelly applied this process to the intramolecular coupling of arylhalides. This tandem stannylation/aryl halide coupling was used for the syntheses of a variety of dihydrophenanthrenes. Most of the internal rings formed are limited to 5 or 6 members, however some cases of macrocyclization have been reported. Unlike a normal Stille coupling, chlorine does not work as a halogen, possibly due to its lower reactivity in the halogen sequence (its shorter bond length and stronger bond dissociation energy makes it more difficult to break via oxidative addition). Starting in the middle of the scheme below and going clockwise, the palladium catalyst (1) oxidatively adds to the most reactive C-X bond (13) to form 14, followed by transmetalation with distannane (15) to yield 16 and reductive elimination to yield an arylstannane (18). The regenerated palladium catalyst (1) can oxidative add to the second C-X bond of 18 to form 19, followed by intramolecular transmetalation to yield 20, followed by reductive elimination to yield the coupled product (22).[6]

 
Catalytic cycle of the Stille-Kelly reaction

Jie Jack Lie et al. made use of the Stille-Kelly coupling in their synthesis of a variety of benzo[4,5]furopyridines ring systems. They invoke a three-step process, involving a Buchwald-Hartwig amination, another palladium-catalyzed coupling reaction, followed by an intramolecular Stille-Kelly coupling. Note that the aryl-iodide bond will oxidatively add to the palladium faster than either of the aryl-bromide bonds.[6][71]

 
Synthesis of benzo[4,5]furopyridines

See also edit

References edit

  1. ^ Hartwig, J. F. Organotransition Metal Chemistry, from Bonding to Catalysis; University Science Books: New York, 2010. ISBN 189138953X
  2. ^ a b Stille, J. K. Angew. Chem. Int. Ed. Engl. 1986, 25, 508–524. (Review)
  3. ^ Farina, V.; Krishnamurthy, V.; Scott, W. J. Org. React. 1998, 50, 1–652. (Review)
  4. ^ Scott, W. J.; Crisp, G. T.; Stille, J. K. Organic Syntheses, Coll. Vol. 8, p. 97 (1993); Vol. 68, p. 116 (1990). (Article)
  5. ^ Stille, J. K.; Echavarren, A. M.; Williams, R. M.; Hendrix, J. A. Organic Syntheses, Coll. Vol. 9, p. 553 (1998); Vol. 71, p. 97 (1993). (Article)
  6. ^ a b c d e f g h i j k l Kurti, L.; Czako, B. Strategic Applications of Named Reactions in Organic Synthesis; Elsevier: Burlington, 2005.
  7. ^ Mitchell, T. N. J. Organomet. Chem., 1986, 304, 1–16.
  8. ^ a b Mitchell, T. N. Synthesis, 1992, 803–815. (doi:10.1055/s-1992-26230)
  9. ^ a b Farina, V. Pure Appl. Chem., 1996, 68, 73–78. (doi:10.1351/pac199668010073).
  10. ^ a b c d e f g h i j k l m n o p Farina, V.; Krishnamurthy, V.; Scott, W. J. The Stille Reaction; Wiley: Online, 2004. (doi:10.1002/0471264180.or050.01).
  11. ^ a b c d e f g h i Espinet, P.; Echavarren, A. M. Angew. Chem. Int. Ed., 2004, 43, 4704–4734.(doi:10.1002/anie.200300638)
  12. ^ Pattenden, G.; Sinclair, D. J. J.Organomet. Chem., 2002, 653, 261–268.
  13. ^ Kosugi, M.; Fugami, K. J. Organomet. Chem., 2002, 19, 10–16.
  14. ^ a b Pierre Genet, J.; Savignac, M. J. Organomet. Chem., 1999, 576, 305–317.
  15. ^ Cordova, C.; Bartolomé, C.; Martínez-Ilarduya, J.M..; Espinet, P. ACS Catal., 2015, 5, 3040–3053.(doi:10.1021/acscatal.5b00448).
  16. ^ Azarian, D.; Dua, S. S.; Eaborn, C.; Walton, D. R. M. J. Organomet. Chem., 1976, 117, C55-C57. (doi:10.1016/S0022-328X(00)91902-8)
  17. ^ Kosugi, M.; Shimizu, Y.; Migita, T. Chem. Lett., 1977, 6, 1423–1424. (doi:10.1246/cl.1977.1423)
  18. ^ a b Kosugi, M.; Sasazawa, K.; Shikizu, Y.; Migita, T. Chem. Lett., 1977, 6, 301–302. (doi:10.1246/cl.1977.301)
  19. ^ Kosugi, M.; Shimizu, Y.; Migita, T. J. Organomet. Chem., 1977, 129, C36-C38. (doi:10.1016/S0022-328X(00)92505-1)
  20. ^ Milstein, D.; Stille, J. K. Journal of the American Chemical Society, 1978, 100, 3636–3638. (doi:10.1021/ja00479a077)
  21. ^ Milstein, D.; Stille, J. K. Journal of the American Chemical Society, 1979, 101, 4992–4998. (doi:10.1021/ja00511a032)
  22. ^ Milstein, D.; Stille, J. K. J. Org. Chem., 1979, 44, 1613–1618. (doi:10.1021/jo01324a006)
  23. ^ Casado, A. L.; Espinet, P.; Gallego, A. M. J. Am, Chem. Soc., 2000, 122, 11771-11782. (doi:10.1021/ja001511o)
  24. ^ a b c d Crabtree, R. H. The Organometallic Chemistry of the Transition Metals, 5th ed.; Wiley: New York, 2009.
  25. ^ a b Perez-Temprano, M. H.; Gallego, A. M.; Casares, J. A.; Espinet, P. Organometallics, 2011, 30, 611–617. (doi:10.1021/om100978w).
  26. ^ a b Minniti, D. Inorg. Chem, 1994, 33, 2631–2634.(doi:10.1021/ic00090a025).
  27. ^ a b Casado, A. L.; Espinet, P. Organometallics, 1998, 17, 954–959. (doi:10.1021/om9709502).
  28. ^ a b Landis, C. R.; Firman, T. K.' Root, D. M.; Cleveland, T. Journal of the American Chemical Society, 1998, 120, 1842–1854. (doi:10.1021/ja9710114).
  29. ^ Vicente, J.; Arcas, A.; Bautista, D. Organometallics, 1997, 16, 2127–2138. (doi:10.1021/om961094h).
  30. ^ a b Pearson, R. G. Inorg. Chem, 1973, 12, 712–713.(doi:10.1021/ic50121a052).
  31. ^ a b c Garcia-Melchor, M.; Braga, A. A. C.; Lledos, A.; Ujaque, G.; Maseras, F. Acc. Chem. Res., 2013, 46, 2626–2634. (doi:10.1021/ar400080r)
  32. ^ a b c d Gillie, A.; Stille, J. K. Journal of the American Chemical Society, 1980, 102, 4933–4941. (doi:10.1021/ja00535a018).
  33. ^ a b c d Brown, J. M.; Cooley, N. A. Chem. Rev., 1988, 88, 1031–1046. (doi:10.1021/cr00089a003).
  34. ^ McKillop, A.; Abel, E. W.; Stone, F. G. A.; Wilkinson, G. Comprehensive Organometallic Chemistry II, Elsevier Scientific: Oxford, 1995.
  35. ^ a b Farina, V.; Journal of the American Chemical Society, 1991, 113, 9585–9595. (doi:10.1021/ja00025a025).
  36. ^ a b "The Stille Reaction" (PDF). hwpi.harvard.edu.
  37. ^ Liebeskind, L. S.; Fengl, R. W. J. Org. Chem., 1990, 55, 5359–5364. (doi:10.1021/jo00306a012).
  38. ^ Farina, V.; Kapadia, S.; Brishnan, B.; Wang, C.; Liebeskind, L. S. J, Org. Chem, 1994, 59, 5905–5911. (doi:10.1021/jo00099a018).
  39. ^ a b Mee, S. P. H.; Lee, V.; Baldwin, J. E. Angew. Chem. Int. Ed., 2004, 43, 1132–1136.
  40. ^ Liebeskind, L. S.; Peña-Cabrera, E. Organic Syntheses, Coll. Vol. 10, p. 9 (2004); Vol. 77, p. 135 (2000). (Article)
  41. ^ Scott, W. J.; Stille, J. K. Journal of the American Chemical Society, 1986, 108, 3033–3040. (doi:10.1021/ja00271a037).
  42. ^ Johnson, C. R.; Adams, J. P.; Braun, M.P.; Senanayake, C. B. W. Tetrahedron Lett., 1992, 33, 919–922. (doi:10.1016/S0040-4039(00)91576-4)
  43. ^ Nair, V.; Turner, G. A.; Chamberlain, S. D. Journal of the American Chemical Society, 1987, 109, 7223–7224. (doi:10.1021/ja00257a071).
  44. ^ Jousseaume, B.; Kwon, W.; Verlhac, J. B.; Denat, F.; Dubac, J. Synlett, 1993, 117–118. (doi:10.1055/s-1993-22368)
  45. ^ Sheffy, F. K.; Godschalx, J. P.; Stille, J. K. Journal of the American Chemical Society, 1984, 106, 4833–4840. (doi:10.1021/ja00329a032)
  46. ^ "Organotin Reagents".
  47. ^ Dietmar Seyferth (1959). "Di-n-butyldivinyltin". Org. Synth. 39: 10. doi:10.15227/orgsyn.039.0010.
  48. ^ Wolf, C.; Lerebours, R. J. Org. Chem., 2003,68 7551–7554. (doi:10.1021/jo0347056).
  49. ^ Crisp, G.T.; Glink, P. T. Tetrahedron, 1994, 50, 2623. (doi:10.1016/S0040-4020(01)86978-7)
  50. ^ Bailey, T. R. Tetrahedron Lett., 1986, 27, 4407. (doi:10.1016/S0040-4039(00)84964-3).
  51. ^ Bao, Z.; Chan, W.; Yu, L. Chem. Mater., 1993, 5, 2–3. (doi:10.1021/cm00025a001).
  52. ^ Bao, Z.; Chan, W. K.; Yu, L. Journal of the American Chemical Society, 1995, 117, 12426-12435. (doi:10.1021/ja00155a007).
  53. ^ Sun, S. S.; Lewis, J. E.; Zhang, J.; Jiang, X.; Zhang, C.; Matos, T.; Li, R.; Polym. Chem., 2010, 1, 663–669. (doi:10.1039/B9PY00324J)
  54. ^ Lebsack, A. D.; Link, J. T.; Overman, L. E.; Stearns, B. A. Journal of the American Chemical Society, 2002, 124, 9008–9009. (doi:10.1021/ja0267425)
  55. ^ Masse, C. E.; Yang, M.; Solomon, J.; Panek, J. S. Journal of the American Chemical Society, 1998, 120, 4123–4134. (doi:10.1021/ja9743194)
  56. ^ Martin, S. F.; Humphrey, J. M.; Ali, A.; Hillier, M. C. Journal of the American Chemical Society, 1999, 121, 866–867. (doi:10.1021/ja9829259)
  57. ^ Kende, A. S.; Kawamura, K.; DeVita, R. J. Journal of the American Chemical Society, 1990, 112 4070–4072. (doi:10.1021/ja00166a072).
  58. ^ Kende, A. S., Koch, K.; Dorey, G.; Kaldor, I.; Liu, K. Journal of the American Chemical Society, 1993, 115, 9842–9843. (doi:10.1021/ja00074a078).
  59. ^ Hong, C. Y, Kishi, Y. Journal of the American Chemical Society, 1991, 113, 9693–9694. (doi:10.1021/ja00025a056).
  60. ^ Tanimoto, N.; Gerritz, S. W.; Sawabe, A.; Noda, T.; Filla, S. A.; Masamune, S. Angew. Chem. Int. Ed., 2003, 33, 673–675. (doi:10.1002/anie.199406731).
  61. ^ Evans, D. A.; Black, W. C. Journal of the American Chemical Society, 1993, 115, 4497–4513. (doi:10.1021/ja00064a011).
  62. ^ Tang, W.; Prusov, E. V. Org. Lett., 2012, 14 4690–4693. (doi:10.1021/ol302219x).
  63. ^ Coleman, R. S.; Walczak, M. C.; Campbell, E. L. Journal of the American Chemical Society, 2005, 127, 16036-16039. (doi:10.1021/ja056217g).
  64. ^ Roth, G. P.; Farina, V.; Liebeskind, L. S.; Peña-Cabrera, E. Tetrahedron Lett. 1995, 36, 2191.
  65. ^ Renaldo, A. F.; Labadie, J. W.; Stille, J. K. Organic Syntheses, Coll. Vol. 8, p. 268 (1993); Vol. 67, p. 86 (1989). (Article)
  66. ^ Stille Reactions with Tetraalkylstannanes and Phenyltrialkylstannanes in Low Melting Sugar-Urea-Salt MixturesGiovanni Imperato, Rudolf Vasold, Burkhard König Advanced Synthesis & Catalysis Volume 348, Issue 15 , Pages 2243–47 2006 doi:10.1002/adsc.2006
  67. ^ P. Espinet, A. M. Echavarren (2004). "The Mechanisms of the Stille Reaction". Angewandte Chemie International Edition. 43 (36): 4704–4734. doi:10.1002/anie.200300638. PMID 15366073.
  68. ^ Knight, S. D.; Overman, L. E.; Pairaudeau, G. Journal of the American Chemical Society, 1993, 115, 9293–9294. (doi:10.1021/ja00073a057)
  69. ^ Monera, E.; Ortar, G. Biorg. Med. Chem. Lett., 2000, 10, 1815–1818. (doi:10.1016/S0960-894X(00)00344-9).
  70. ^ Gyorkos, A. C.; Stille, J. K.; Hegedus, L. S. Journal of the American Chemical Society, 1990, 112, 8465–8472. (doi:10.1021/ja00179a035).
  71. ^ Yue, W. S.; Li, J. J. Org. Lett., 2002, 4, 2201–2203. (doi:10.1021/ol0260425)

External links edit

  • Stille reaction handout from the Myers group.
  • Stille reaction at organic-chemistry.org

stille, reaction, chemical, reaction, widely, used, organic, synthesis, reaction, involves, coupling, organic, groups, which, carried, organotin, compound, also, known, organostannanes, variety, organic, electrophiles, provide, other, coupling, partner, many, . 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 also known as organostannanes A variety of organic electrophiles provide the other coupling partner The Stille reaction is one of many palladium catalyzed coupling reactions 1 2 3 Stille reaction Named after John Kenneth Stille Reaction type Coupling reaction Identifiers Organic Chemistry Portal stille coupling RSC ontology ID RXNO 0000035 R 1 Sn Alkyl 3 R 2 X ligand set Pd 0 catalytic R 1 R 2 c o u p l e d p r o d u c t X Sn Alkyl 3 displaystyle color Blue ce R 1 Sn Alkyl 3 color Red ce R 2 X ce gt color Green ce Pd 0 text catalytic text ligand set overbrace color Blue ce R 1 color Red ce R 2 coupled product color Red ce X color Blue ce Sn Alkyl 3 R 1 R 2 displaystyle color Blue ce R 1 color Red ce R 2 Allyl alkenyl aryl benzyl acyl X displaystyle color Red ce X halides Cl Br I pseudohalides OTf OPO OR 2 OAc The R1 group attached to the trialkyltin is normally sp2 hybridized including vinyl and aryl groups These organostannanes are also stable to both air and moisture and many of these reagents either are commercially available or can be synthesized from literature precedent However these tin reagents tend to be highly toxic X is typically a halide such as Cl Br or I yet pseudohalides such as triflates and sulfonates and phosphates can also be used 4 5 Several reviews have been published 6 2 7 8 9 10 11 12 13 14 15 excessive citations Contents 1 History 2 Mechanism 2 1 Oxidative addition 2 2 Transmetallation 2 3 Reductive elimination step 3 Kinetics 3 1 Additives 3 2 Competing side reactions 4 Scope 4 1 Electrophile 4 2 Stannane 5 Applications 5 1 Natural product total synthesis 6 Variations 6 1 Stille carbonylative cross coupling 6 2 Stille Kelly coupling 7 See also 8 References 9 External linksHistory editThe first example of a palladium catalyzed coupling of aryl halides with organotin reagents was reported by Colin Eaborn in 1976 16 This reaction yielded from 7 to 53 of diaryl product This process was expanded to the coupling of acyl chlorides with alkyl tin reagents in 1977 by Toshihiko Migita yielding 53 to 87 ketone product 17 nbsp First reactions of organotin reagents In 1977 Migita published further work on the coupling of allyl tin reagents with both aryl C and acyl D halides The greater ability of allyl groups to migrate to the palladium catalyst allowed the reactions to be performed at lower temperatures Yields for aryl halides ranged from 4 to 100 and for acyl halides from 27 to 86 18 19 Reflecting the early contributions of Migita and Kosugi the Stille reaction is sometimes called the Migita Kosugi Stille coupling nbsp First reactions of organotin reagents John Kenneth Stille subsequently reported the coupling of a variety of alkyl tin reagents in 1978 with numerous aryl and acyl halides under mild reaction conditions with much better yields 76 99 18 20 Stille continued his work in the 1980s on the synthesis of a multitude of ketones using this broad and mild process and elucidated a mechanism for this transformation 21 22 nbsp First reactions of organotin reagents By the mid 1980s over 65 papers on the topic of coupling reactions involving tin had been published continuing to explore the substrate scope of this reaction While initial research in the field focused on the coupling of alkyl groups most future work involved the much more synthetically useful coupling of vinyl alkenyl aryl and allyl organostannanes to halides Due to these organotin reagent s stability to air and their ease of synthesis the Stille reaction became common in organic synthesis 8 Mechanism editThe mechanism of the Stille reaction has been extensively studied 11 23 The catalytic cycle involves an oxidative addition of a halide or pseudohalide 2 to a palladium catalyst 1 transmetalation of 3 with an organotin reagent 4 and reductive elimination of 5 to yield the coupled product 7 and the regenerated palladium catalyst 1 24 nbsp Catalytic cycle of the Stille Reaction However the detailed mechanism of the Stille coupling is extremely complex and can occur via numerous reaction pathways Like other palladium catalyzed coupling reactions the active palladium catalyst is believed to be a 14 electron Pd 0 complex which can be generated in a variety of ways Use of an 18 or 16 electron Pd 0 source Pd PPh3 4 Pd dba 2 can undergo ligand dissociation to form the active species Second phosphines can be added to ligandless palladium 0 Finally as pictured reduction of a Pd II source 8 Pd OAc 2 PdCl2 MeCN 2 PdCl2 PPh3 2 BnPdCl PPh3 2 etc by added phosphine ligands or organotin reagents is also common 6 Oxidative addition edit Oxidative addition to the 14 electron Pd 0 complex is proposed This process gives a 16 electron Pd II species It has been suggested that anionic ligands such as OAc accelerate this step by the formation of Pd OAc PR3 n making the palladium species more nucleophillic 11 25 In some cases especially when an sp3 hybridized organohalide is used an SN2 type mechanism tends to prevail yet this is not as commonly seen in the literature 11 25 However despite normally forming a cis intermediate after a concerted oxidative addition this product is in rapid equilibrium with its trans isomer 26 27 nbsp Cis Trans Isomerization There are multiple reasons why isomerization is favored here First a bulky ligand set is usually used in these processes such as phosphines and it is highly unfavorable for them to adopt a cis orientation relative to each other resulting in isomerization to the more favorable trans product 26 27 An alternative explanation for this phenomenon dubbed antisymbiosis or transphobia is by invocation of the sdn model 24 28 Under this theory palladium is a hypervalent species Hence R1 and the trans ligand being trans to each other will compete with one palladium orbital for bonding This 4 electron 3 center bond is weakest when two strong donating groups are present which heavily compete for the palladium orbital Relative to any ligand normally used the C donor R1 ligand has a much higher trans effect This trans influence is a measure of how competitive ligands trans to each other will compete for palladium s orbital The usual ligand set phosphines and C donors R1 are both soft ligands meaning that they will form strong bonds to palladium and heavily compete with each other for bonding 29 30 Since halides or pseudohalides are significantly more electronegative their bonding with palladium will be highly polarized with most of the electron density on the X group making them low trans effect ligands Hence it will be highly favorable for R1 to be trans to X since the R1 group will be able to form a stronger bond to the palladium 24 28 30 nbsp sd n model for cis trans isomers Transmetallation edit The transmetallation of the trans intermediate from the oxidative addition step is believed to proceed via a variety of mechanisms depending on the substrates and conditions The most common type of transmetallation for the Stille coupling involves an associative mechanism This pathway implies that the organostannane normally a tin atom bonded to an allyl alkenyl or aryl group can coordinate to the palladium via one of these double bonds This produces a fleeting pentavalent 18 electron species which can then undergo ligand detachment to form a square planar complex again Despite the organostannane being coordinated to the palladium through the R2 group R2 must be formally transferred to the palladium the R2 Sn bond must be broken and the X group must leave with the tin completing the transmetalation This is believed to occur through two mechanisms 31 First when the organostannane initially adds to the trans metal complex the X group can coordinate to the tin in addition to the palladium producing a cyclic transition state Breakdown of this adduct results in the loss of R3Sn X and a trivalent palladium complex with R1 and R2 present in a cis relationship Another commonly seen mechanism involves the same initial addition of the organostannane to the trans palladium complex as seen above however in this case the X group does not coordinate to the tin producing an open transition state After the a carbon relative to tin attacks the palladium the tin complex will leave with a net positive charge In the scheme below please note that the double bond coordinating to tin denotes R2 so any alkenyl allyl or aryl group Furthermore the X group can dissociate at any time during the mechanism and bind to the Sn complex at the end Density functional theory calculations predict that an open mechanism will prevail if the 2 ligands remain attached to the palladium and the X group leaves while the cyclic mechanism is more probable if a ligand dissociates prior to the transmetalation Hence good leaving groups such as triflates in polar solvents favor the cyclic transition state while bulky phosphine ligands will favor the open transition state 31 nbsp The two mechanisms cyclic and open of transmetallation in the Stille reaction A less common pathway for transmetalation is through a dissociative or solvent assisted mechanism Here a ligand from the tetravalent palladium species dissociates and a coordinating solvent can add onto the palladium When the solvent detaches to form a 14 electron trivalent intermediate the organostannane can add to the palladium undergoing an open or cyclic type process as above 31 Reductive elimination step edit In order for R1 R2 to reductively eliminate these groups must occupy mutually cis coordination sites Any trans adducts must therefore isomerize to the cis intermediate or the coupling will be frustrated A variety of mechanisms exist for reductive elimination and these are usually considered to be concerted 11 32 33 First the 16 electron tetravalent intermediate from the transmetalation step can undergo unassisted reductive elimination from a square planar complex This reaction occurs in two steps first the reductive elimination is followed by coordination of the newly formed sigma bond between R1 and R2 to the metal with ultimate dissociation yielding the coupled product 11 32 33 nbsp Concerted reductive elimination for the Stille reaction The previous process however is sometimes slow and can be greatly accelerated by dissociation of a ligand to yield a 14 electron T shaped intermediate This intermediate can then rearrange to form a Y shaped adduct which can undergo faster reductive elimination 11 32 33 nbsp Dissociative reductive elimination for the Stille reaction Finally an extra ligand can associate to the palladium to form an 18 electron trigonal bipyramidal structure with R1 and R2 cis to each other in equatorial positions The geometry of this intermediate makes it similar to the Y shaped above 11 32 33 nbsp Associative reductive elimination for the Stille reaction The presence of bulky ligands can also increase the rate of elimination Ligands such as phosphines with large bite angles cause steric repulsion between L and R1 and R2 resulting in the angle between L and the R groups to increase and the angle between R1 and R2 to hence decrease allowing for quicker reductive elimination 11 24 nbsp Cis reductive elimination in the Stille reactionKinetics editThe rate at which organostannanes transmetalate with palladium catalysts is shown below Sp2 hybridized carbon groups attached to tin are the most commonly used coupling partners and sp3 hybridized carbons require harsher conditions and terminal alkynes may be coupled via a C H bond through the Sonogashira reaction nbsp Relative rates of the Stille reaction As the organic tin compound a trimethylstannyl or tributylstannyl compound is normally used Although trimethylstannyl compounds show higher reactivity compared with tributylstannyl compounds and have much simpler 1H NMR spectra the toxicity of the former is much larger 34 Optimizing which ligands are best at carrying out the reaction with high yield and turnover rate can be difficult This is because the oxidative addition requires an electron rich metal hence favoring electron donating ligands However an electron deficient metal is more favorable for the transmetalation and reductive elimination steps making electron withdrawing ligands the best here Therefore the optimal ligand set heavily depends on the individual substrates and conditions used These can change the rate determining step as well as the mechanism for the transmetalation step 35 Normally ligands of intermediate donicity such as phosphines are utilized Rate enhancements can be seen when moderately electron poor ligands such as tri 2 furylphosphine or triphenylarsenine are used Likewise ligands of high donor number can slow down or inhibit coupling reactions 35 36 These observations imply that normally the rate determining step for the Stille reaction is transmetalation 36 Additives edit The most common additive to the Stille reaction is stoichiometric or co catalytic copper I specifically copper iodide which can enhance rates up by gt 103 fold It has been theorized that in polar solvents copper transmetalate with the organostannane The resulting organocuprate reagent could then transmetalate with the palladium catalyst Furthermore in ethereal solvents the copper could also facilitate the removal of a phosphine ligand activating the Pd center 9 37 38 39 40 Lithium chloride has been found to be a powerful rate accelerant in cases where the X group dissociates from palladium i e the open mechanism The chloride ion is believed to either displace the X group on the palladium making the catalyst more active for transmetalation or by coordination to the Pd 0 adduct to accelerate the oxidative addition Also LiCl salt enhances the polarity of the solvent making it easier for this normally anionic ligand Cl Br OTf etc to leave This additive is necessary when a solvent like THF is used however utilization of a more polar solvent such as NMP can replace the need for this salt additive However when the coupling s transmetalation step proceeds via the cyclic mechanism addition of lithium chloride can actually decrease the rate As in the cyclic mechanism a neutral ligand such as phosphine must dissociate instead of the anionic X group 10 41 Finally sources of fluoride ions such as cesium fluoride also effect on the catalytic cycle First fluoride can increase the rates of reactions of organotriflates possibly by the same effect as lithium chloride Furthermore fluoride ions can act as scavengers for tin byproducts making them easier to remove via filtration 39 Competing side reactions edit The most common side reactivity associated with the Stille reaction is homocoupling of the stannane reagents to form an R2 R2 dimer It is believed to proceed through two possible mechanisms First reaction of two equivalents of organostannane with the Pd II precatalyst will yield the homocoupled product after reductive elimination Second the Pd 0 catalyst can undergo a radical process to yield the dimer The organostannane reagent used is traditionally tetravalent at tin normally consisting of the sp2 hybridized group to be transferred and three non transferable alkyl groups As seen above alkyl groups are normally the slowest at migrating onto the palladium catalyst 10 nbsp Homocoupling and transfer of inert ligands It has also been found that at temperatures as low as 50 C aryl groups on both palladium and a coordinated phosphine can exchange While normally not detected they can be a potential minor product in many cases 10 nbsp Aryl transfer through phosphines Finally a rather rare and exotic side reaction is known as cine substitution Here after initial oxidative addition of an aryl halide this Pd Ar species can insert across a vinyl tin double bond After b hydride elimination migratory insertion and protodestannylation a 1 2 disubstituted olefin can be synthesized 10 nbsp Cine substitution Numerous other side reactions can occur and these include E Z isomerization which can potentially be a problem when an alkenylstannane is utilized The mechanism of this transformation is currently unknown Normally organostannanes are quite stable to hydrolysis yet when very electron rich aryl stannanes are used this can become a significant side reaction 10 Scope editElectrophile edit Vinyl halides are common coupling partners in the Stille reaction and reactions of this type are found in numerous natural product total syntheses Normally vinyl iodides and bromides are used Vinyl chlorides are insufficiently reactive toward oxidative addition to Pd 0 Iodides are normally preferred they will typically react faster and under milder conditions than will bromides This difference is demonstrated below by the selective coupling of a vinyl iodide in the presence of a vinyl bromide 10 nbsp Vinyl Iodide reacts faster than vinyl bromide Normally the stereochemistry of the alkene is retained throughout the reaction except under harsh reaction conditions A variety of alkenes may be used and these include both a and b halo a b unsaturated ketones esters and sulfoxides which normally need a copper I additive to proceed and more see example below 42 Vinyl triflates are also sometimes used Some reactions require the addition of LiCl and others are slowed down implying that two mechanistic pathways are present 10 nbsp Addition to an alpha beta unsaturated alkene Another class of common electrophiles are aryl and heterocyclic halides As for the vinyl substrates bromides and iodides are more common despite their greater expense A multitude of aryl groups can be chosen including rings substituted with electron donating substituents biaryl rings and more Halogen substituted heterocycles have also been used as coupling partners including pyridines furans thiophenes thiazoles indoles imidazoles purines uracil cytosines pyrimidines and more See below for table of heterocycles halogens can be substituted at a variety of positions on each 10 nbsp Variety of heterocycles which can undergo addition Below is an example of the use of Stille coupling to build complexity on heterocycles of nucleosides such as purines 43 nbsp Addition to a heterocycle Aryl triflates and sulfonates are also couple to a wide variety of organostannane reagents Triflates tend to react comparably to bromides in the Stille reaction 10 Acyl chlorides are also used as coupling partners and can be used with a large range of organostannane even alkyl tin reagents to produce ketones see example below 44 However it is sometimes difficult to introduce acyl chloride functional groups into large molecules with sensitive functional groups An alternative developed to this process is the Stille carbonylative cross coupling reaction which introduces the carbonyl group via carbon monoxide insertion 10 nbsp Acyl chlorides can be used as well Allylic benzylic and propargylic halides can also be coupled While commonly employed allylic halides proceed via an h3 transition state allowing for coupling with the organostannane at either the a or g position occurring predominantly at the least substituted carbon see example below 45 Alkenyl epoxides adjacent epoxides and alkenes can also undergo this same coupling through an h3 transition state as opening the epoxide to an alcohol While allylic and benzylic acetates are commonly used propargylic acetates are unreactive with organostannanes 10 nbsp allylic bromides will form an heta 3 complex Stannane edit Organostannane reagents are common Several are commercially available 46 Stannane reagents can be synthesized by the reaction of a Grignard or organolithium reagent with trialkyltin chlorides For example vinyltributyltin is prepared by the reaction of vinylmagnesium bromide with tributyltin chloride 47 Hydrostannylation of alkynes or alkenes provides many derivatives Organotin reagents are air and moisture stable Some reactions can even take place in water 48 They can be purified by chromatography They are tolerant to most functional groups Some organotin compounds are heavily toxic especially trimethylstannyl derivatives 10 The use of vinylstannane or alkenylstannane reagents is widespread 10 In regards to limitations both very bulky stannane reagents and stannanes with substitution on the a carbon tend to react sluggishly or require optimization For example in the case below the a substituted vinylstannane only reacts with a terminal iodide due to steric hindrance 49 nbsp Stannae 1 Arylstannane reagents are also common and both electron donating and electron withdrawing groups actually increase the rate of the transmetalation This again implies that two mechanisms of transmetalation can occur The only limitation to these reagents are substituents at the ortho position as small as methyl groups can decrease the rate of reaction A wide variety of heterocycles see Electrophile section can also be used as coupling partners see example with a thiazole ring below 10 50 nbsp Regioselective coupling of a heterocyclic stannae with an aryl bromide nbsp Coupling of stannane to acyl chloride Alkynylstannanes the most reactive of stannanes have also been used in Stille couplings They are not usually needed as terminal alkynes can couple directly to palladium catalysts through their C H bond via Sonogashira coupling Allylstannanes have been reported to have worked yet difficulties arise like with allylic halides with the difficulty in control regioselectivity for a and g addition Distannane and acyl stannane reagents have also been used in Stille couplings 10 Applications editThe Stille reaction has been used in the synthesis of a variety of polymers 51 52 53 However the most widespread use of the Stille reaction is its use in organic syntheses and specifically in the synthesis of natural products Natural product total synthesis edit Larry Overman s 19 step enantioselective total synthesis of quadrigemine C involves a double Stille cross metathesis reaction 6 54 The complex organostannane is coupled onto two aryl iodide groups After a double Heck cyclization the product is achieved nbsp Total Synthesis of Quadrigemine C Panek s 32 step enantioselective total synthesis of ansamycin antibiotic mycotrienol makes use of a late stage tandem Stille type macrocycle coupling Here the organostannane has two terminal tributyl tin groups attacked to an alkene This organostannane stitches the two ends of the linear starting material into a macrocycle adding the missing two methylene units in the process After oxidation of the aromatic core with ceric ammonium nitrate CAN and deprotection with hydrofluoric acid yields the natural product in 54 yield for the 3 steps 6 55 nbsp Total synthesis of mycotrienol Stephen F Martin and coworkers 21 step enantioselective total synthesis of the manzamine antitumor alkaloid Ircinal A makes use of a tandem one pot Stille Diels Alder reaction An alkene group is added to vinyl bromide followed by an in situ Diels Alder cycloaddition between the added alkene and the alkene in the pyrrolidine ring 6 56 nbsp Total synthesis of ircinal ANumerous other total syntheses utilize the Stille reaction including those of oxazolomycin 57 lankacidin C 58 onamide A 59 calyculin A 60 lepicidin A 61 ripostatin A 62 and lucilactaene 6 63 The image below displays the final natural product the organohalide blue the organostannane red and the bond being formed green and circled From these examples it is clear that the Stille reaction can be used both at the early stages of the synthesis oxazolomycin and calyculin A at the end of a convergent route onamide A lankacidin C ripostatin A or in the middle lepicidin A and lucilactaene The synthesis of ripostatin A features two concurrent Stille couplings followed by a ring closing metathesis The synthesis of lucilactaene features a middle subunit having a borane on one side and a stannane on the other allowing for Stille reactionfollowed by a subsequent Suzuki coupling nbsp A variety of total syntheses which make use of the Stille reactionVariations editIn addition to performing the reaction in a variety of organic solvents conditions have been devised which allow for a broad range of Stille couplings in aqueous solvent 14 In the presence of Cu I salts palladium on carbon has been shown to be an effective catalyst 64 65 In the realm of green chemistry a Stille reaction is reported taking place in a low melting and highly polar mixture of a sugar such as mannitol a urea such as dimethylurea and a salt such as ammonium chloride 66 67 The catalyst system is tris dibenzylideneacetone dipalladium 0 with triphenylarsine nbsp A Stille reaction variation coupling of phenyliodide and tetramethyltin Stille carbonylative cross coupling edit A common alteration to the Stille coupling is the incorporation of a carbonyl group between R1 and R2 serving as an efficient method to form ketones This process is extremely similar to the initial exploration by Migita and Stille see History of coupling organostannane to acyl chlorides However these moieties are not always readily available and can be difficult to form especially in the presence of sensitive functional groups Furthermore controlling their high reactivity can be challenging The Stille carbonylative cross coupling employs the same conditions as the Stille coupling except with an atmosphere of carbon monoxide CO being used The CO can coordinate to the palladium catalyst 9 after initial oxidative addition followed by CO insertion into the Pd R1 bond 10 resulting in subsequent reductive elimination to the ketone 12 The transmetalation step is normally the rate determining step 6 nbsp Catalytic cycle of the Stille carbonylative cross coupling Larry Overman and coworkers make use of the Stille carbonylative cross coupling in their 20 step enantioselective total synthesis of strychnine The added carbonyl is later converted to a terminal alkene via a Wittig reaction allowing for the key tertiary nitrogen and the pentacyclic core to be formed via an aza Cope Mannich reaction 6 68 nbsp Total synthesis of strychnine Giorgio Ortar et al explored how the Stille carbonylative cross coupling could be used to synthesize benzophenone phosphores These were embedded into 4 benzoyl L phenylalanine peptides and used for their photoaffinity labelling properties to explore various peptide protein interactions 6 69 nbsp Synthesis of phosphores Louis Hegedus 16 step racemic total synthesis of Jatraphone involved a Stille carbonylative cross coupling as its final step to form the 11 membered macrocycle Instead of a halide a vinyl triflate is used there as the coupling partner 6 70 nbsp Total synthesis of Jatraphone Stille Kelly coupling edit Using the seminal publication by Eaborn in 1976 which forms arylstannanes from arylhalides and distannanes T Ross Kelly applied this process to the intramolecular coupling of arylhalides This tandem stannylation aryl halide coupling was used for the syntheses of a variety of dihydrophenanthrenes Most of the internal rings formed are limited to 5 or 6 members however some cases of macrocyclization have been reported Unlike a normal Stille coupling chlorine does not work as a halogen possibly due to its lower reactivity in the halogen sequence its shorter bond length and stronger bond dissociation energy makes it more difficult to break via oxidative addition Starting in the middle of the scheme below and going clockwise the palladium catalyst 1 oxidatively adds to the most reactive C X bond 13 to form 14 followed by transmetalation with distannane 15 to yield 16 and reductive elimination to yield an arylstannane 18 The regenerated palladium catalyst 1 can oxidative add to the second C X bond of 18 to form 19 followed by intramolecular transmetalation to yield 20 followed by reductive elimination to yield the coupled product 22 6 nbsp Catalytic cycle of the Stille Kelly reaction Jie Jack Lie et al made use of the Stille Kelly coupling in their synthesis of a variety of benzo 4 5 furopyridines ring systems They invoke a three step process involving a Buchwald Hartwig amination another palladium catalyzed coupling reaction followed by an intramolecular Stille Kelly coupling Note that the aryl iodide bond will oxidatively add to the palladium faster than either of the aryl bromide bonds 6 71 nbsp Synthesis of benzo 4 5 furopyridinesSee also editOrganotin chemistry Organostannane addition Palladium catalyzed coupling reactions Suzuki reaction Negishi coupling Heck reaction Hiyama couplingReferences edit Hartwig J F Organotransition Metal Chemistry from Bonding to Catalysis University Science Books New York 2010 ISBN 189138953X a b Stille J K Angew Chem Int Ed Engl 1986 25 508 524 Review Farina V Krishnamurthy V Scott W J Org React 1998 50 1 652 Review Scott W J Crisp G T Stille J K Organic Syntheses Coll Vol 8 p 97 1993 Vol 68 p 116 1990 Article Stille J K Echavarren A M Williams R M Hendrix J A Organic Syntheses Coll Vol 9 p 553 1998 Vol 71 p 97 1993 Article a b c d e f g h i j k l Kurti L Czako B Strategic Applications of Named Reactions in Organic Synthesis Elsevier Burlington 2005 Mitchell T N J Organomet Chem 1986 304 1 16 a b Mitchell T N Synthesis 1992 803 815 doi 10 1055 s 1992 26230 a b Farina V Pure Appl Chem 1996 68 73 78 doi 10 1351 pac199668010073 a b c d e f g h i j k l m n o p Farina V Krishnamurthy V Scott W J The Stille Reaction Wiley Online 2004 doi 10 1002 0471264180 or050 01 a b c d e f g h i Espinet P Echavarren A M Angew Chem Int Ed 2004 43 4704 4734 doi 10 1002 anie 200300638 Pattenden G Sinclair D J J Organomet Chem 2002 653 261 268 Kosugi M Fugami K J Organomet Chem 2002 19 10 16 a b Pierre Genet J Savignac M J Organomet Chem 1999 576 305 317 Cordova C Bartolome C Martinez Ilarduya J M Espinet P ACS Catal 2015 5 3040 3053 doi 10 1021 acscatal 5b00448 Azarian D Dua S S Eaborn C Walton D R M J Organomet Chem 1976 117 C55 C57 doi 10 1016 S0022 328X 00 91902 8 Kosugi M Shimizu Y Migita T Chem Lett 1977 6 1423 1424 doi 10 1246 cl 1977 1423 a b Kosugi M Sasazawa K Shikizu Y Migita T Chem Lett 1977 6 301 302 doi 10 1246 cl 1977 301 Kosugi M Shimizu Y Migita T J Organomet Chem 1977 129 C36 C38 doi 10 1016 S0022 328X 00 92505 1 Milstein D Stille J K Journal of the American Chemical Society 1978 100 3636 3638 doi 10 1021 ja00479a077 Milstein D Stille J K Journal of the American Chemical Society 1979 101 4992 4998 doi 10 1021 ja00511a032 Milstein D Stille J K J Org Chem 1979 44 1613 1618 doi 10 1021 jo01324a006 Casado A L Espinet P Gallego A M J Am Chem Soc 2000 122 11771 11782 doi 10 1021 ja001511o a b c d Crabtree R H The Organometallic Chemistry of the Transition Metals 5th ed Wiley New York 2009 a b Perez Temprano M H Gallego A M Casares J A Espinet P Organometallics 2011 30 611 617 doi 10 1021 om100978w a b Minniti D Inorg Chem 1994 33 2631 2634 doi 10 1021 ic00090a025 a b Casado A L Espinet P Organometallics 1998 17 954 959 doi 10 1021 om9709502 a b Landis C R Firman T K Root D M Cleveland T Journal of the American Chemical Society 1998 120 1842 1854 doi 10 1021 ja9710114 Vicente J Arcas A Bautista D Organometallics 1997 16 2127 2138 doi 10 1021 om961094h a b Pearson R G Inorg Chem 1973 12 712 713 doi 10 1021 ic50121a052 a b c Garcia Melchor M Braga A A C Lledos A Ujaque G Maseras F Acc Chem Res 2013 46 2626 2634 doi 10 1021 ar400080r a b c d Gillie A Stille J K Journal of the American Chemical Society 1980 102 4933 4941 doi 10 1021 ja00535a018 a b c d Brown J M Cooley N A Chem Rev 1988 88 1031 1046 doi 10 1021 cr00089a003 McKillop A Abel E W Stone F G A Wilkinson G Comprehensive Organometallic Chemistry II Elsevier Scientific Oxford 1995 a b Farina V Journal of the American Chemical Society 1991 113 9585 9595 doi 10 1021 ja00025a025 a b The Stille Reaction PDF hwpi harvard edu Liebeskind L S Fengl R W J Org Chem 1990 55 5359 5364 doi 10 1021 jo00306a012 Farina V Kapadia S Brishnan B Wang C Liebeskind L S J Org Chem 1994 59 5905 5911 doi 10 1021 jo00099a018 a b Mee S P H Lee V Baldwin J E Angew Chem Int Ed 2004 43 1132 1136 Liebeskind L S Pena Cabrera E Organic Syntheses Coll Vol 10 p 9 2004 Vol 77 p 135 2000 Article Scott W J Stille J K Journal of the American Chemical Society 1986 108 3033 3040 doi 10 1021 ja00271a037 Johnson C R Adams J P Braun M P Senanayake C B W Tetrahedron Lett 1992 33 919 922 doi 10 1016 S0040 4039 00 91576 4 Nair V Turner G A Chamberlain S D Journal of the American Chemical Society 1987 109 7223 7224 doi 10 1021 ja00257a071 Jousseaume B Kwon W Verlhac J B Denat F Dubac J Synlett 1993 117 118 doi 10 1055 s 1993 22368 Sheffy F K Godschalx J P Stille J K Journal of the American Chemical Society 1984 106 4833 4840 doi 10 1021 ja00329a032 Organotin Reagents Dietmar Seyferth 1959 Di n butyldivinyltin Org Synth 39 10 doi 10 15227 orgsyn 039 0010 Wolf C Lerebours R J Org Chem 2003 68 7551 7554 doi 10 1021 jo0347056 Crisp G T Glink P T Tetrahedron 1994 50 2623 doi 10 1016 S0040 4020 01 86978 7 Bailey T R Tetrahedron Lett 1986 27 4407 doi 10 1016 S0040 4039 00 84964 3 Bao Z Chan W Yu L Chem Mater 1993 5 2 3 doi 10 1021 cm00025a001 Bao Z Chan W K Yu L Journal of the American Chemical Society 1995 117 12426 12435 doi 10 1021 ja00155a007 Sun S S Lewis J E Zhang J Jiang X Zhang C Matos T Li R Polym Chem 2010 1 663 669 doi 10 1039 B9PY00324J Lebsack A D Link J T Overman L E Stearns B A Journal of the American Chemical Society 2002 124 9008 9009 doi 10 1021 ja0267425 Masse C E Yang M Solomon J Panek J S Journal of the American Chemical Society 1998 120 4123 4134 doi 10 1021 ja9743194 Martin S F Humphrey J M Ali A Hillier M C Journal of the American Chemical Society 1999 121 866 867 doi 10 1021 ja9829259 Kende A S Kawamura K DeVita R J Journal of the American Chemical Society 1990 112 4070 4072 doi 10 1021 ja00166a072 Kende A S Koch K Dorey G Kaldor I Liu K Journal of the American Chemical Society 1993 115 9842 9843 doi 10 1021 ja00074a078 Hong C Y Kishi Y Journal of the American Chemical Society 1991 113 9693 9694 doi 10 1021 ja00025a056 Tanimoto N Gerritz S W Sawabe A Noda T Filla S A Masamune S Angew Chem Int Ed 2003 33 673 675 doi 10 1002 anie 199406731 Evans D A Black W C Journal of the American Chemical Society 1993 115 4497 4513 doi 10 1021 ja00064a011 Tang W Prusov E V Org Lett 2012 14 4690 4693 doi 10 1021 ol302219x Coleman R S Walczak M C Campbell E L Journal of the American Chemical Society 2005 127 16036 16039 doi 10 1021 ja056217g Roth G P Farina V Liebeskind L S Pena Cabrera E Tetrahedron Lett 1995 36 2191 Renaldo A F Labadie J W Stille J K Organic Syntheses Coll Vol 8 p 268 1993 Vol 67 p 86 1989 Article Stille Reactions with Tetraalkylstannanes and Phenyltrialkylstannanes in Low Melting Sugar Urea Salt MixturesGiovanni Imperato Rudolf Vasold Burkhard Konig Advanced Synthesis amp Catalysis Volume 348 Issue 15 Pages 2243 47 2006 doi 10 1002 adsc 2006 P Espinet A M Echavarren 2004 The Mechanisms of the Stille Reaction Angewandte Chemie International Edition 43 36 4704 4734 doi 10 1002 anie 200300638 PMID 15366073 Knight S D Overman L E Pairaudeau G Journal of the American Chemical Society 1993 115 9293 9294 doi 10 1021 ja00073a057 Monera E Ortar G Biorg Med Chem Lett 2000 10 1815 1818 doi 10 1016 S0960 894X 00 00344 9 Gyorkos A C Stille J K Hegedus L S Journal of the American Chemical Society 1990 112 8465 8472 doi 10 1021 ja00179a035 Yue W S Li J J Org Lett 2002 4 2201 2203 doi 10 1021 ol0260425 External links editStille reaction handout from the Myers group Stille reaction at organic chemistry org Retrieved from https en wikipedia org w index php title Stille reaction amp oldid 1209554688, 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.