Stannatrane

Last updated
The general structure of a stannatrane reagent, where variants have been synthesized with X=C (carbastannatrane) and X=N (azastannatrane) and R is an alkyl or aryl substituent. General stannatrane.jpg
The general structure of a stannatrane reagent, where variants have been synthesized with X=C (carbastannatrane) and X=N (azastannatrane) and R is an alkyl or aryl substituent.

A stannatrane (IUPAC: 1-aza-5-stannabicyclo[3.3.3]undecane) is a tin-based atrane belonging to the larger class of organostannanes. Though the term stannatrane is often used to refer to the more commonly employed carbastannatrane, azastannatranes have also been synthesized (prefix refers to the identity of the atom bound directly to tin center). [1] Stannatrane reagents offer highly selective methods for the incorporation of "R" substituents in complex molecules for late-stage diversification. These reagents differ from their tetraalkyl organostannane analogues in that there is no participation of dummy ligands in the transmetalation step, offering selective alkyl transfer in Stille Coupling reactions. [2] [3] [4] These transmetalating agents are known to be air- and moisture-stable, as well as generally less toxic than their tetraalkyl counterparts.

Contents

History and structural properties

The first carbastannatrane was reported in 1984 by Jurkschat and Tzschach. [5] By reaction of an amino-triGrignard reagent with tin(IV) chloride to yield the stannatrane chloride, which was treated with methyl lithium to yield the corresponding methyl stannatrane. Based on a very small methyl J(119Sn–13C) coupling constant of 171 Hz, it was determined that the tin center of methyl stannatrane was indeed pentacoordinate, indicating nitrogen coordination.

Synthesis of stannatrane chloride and methyl stannatrane by Jurkschat and colleagues. Stannatranesynthesis.jpg
Synthesis of stannatrane chloride and methyl stannatrane by Jurkschat and colleagues.
Stereoscopic representation of methyl stannatrane crystal structure. MethylStannatrane Crystal Structure.png
Stereoscopic representation of methyl stannatrane crystal structure.

The crystal and molecular structure was explained by X-ray crystallography. [6] The X-ray diffraction study confirmed the tricyclic ring structure and gave insight toward the geometry of the complex. With a tin-nitrogen distance of 2.624 Å, the formal bond order was calculated to be about 0.46. The presence of the tin-nitrogen interaction, albeit weaker than anticipated, led to a few key discoveries: (1) the distortion from ideal trigonal bipyramidal toward monocapped tetrahedron geometry; [7] (2) the lengthening of the apical tin-methyl bond by ~ 0.1 Å (largest known value for any existing tetraorganotin compounds); (3) the observation of unusual hybridization at the apical tin-methyl bond.

Syntheses of alkyl stannatranes

A modified synthesis of atrane tricycle utilizes Schwartz's Reagent, triallylamine, and tin(IV) chloride in a one-pot method. [8]

(1) Hydrozirconation method for synthesizing stannatrane chloride. (2) Synthesis of lithium carbastannatrane and subsequent mesylate displacement. Stannatrane Lithiation.jpg
(1) Hydrozirconation method for synthesizing stannatrane chloride. (2) Synthesis of lithium carbastannatrane and subsequent mesylate displacement.

Though the latter step is still commonly used for functionalization of stannatrane chloride to simple alkyl derivatives via transmetalation, Biscoe and coworkers have developed a lithiation method that provides access to a variety of enantioenriched alkyl substituents from optically active mesylates (2). [9] [10] After treatment of stannatrane chloride with lithium napthalide, a lithium carbastannatrane was quenched with the corresponding enantiopure mesylate to yield the desired enantioenriched alkyl carbastannatrane in moderate yield with high enantiomeric excess.

Applications in cross-coupling

The earliest reported use of carbastannatranes in palladium-catalyzed Stille coupling reactions in 1992 compared the efficiency of methyl stannatrane with tetramethyltin in the presence of aryl bromides and alkenyl iodides. [8] Tetramethyltin only resulted in less than five percent conversion, whereas methyl stannatrane resulted in 67% yield under the same conditions. This difference was attributed to the nitrogen lone pair lengthening the tin-methyl bond, increasing its lability toward transmetalation. A method was developed for Stille couplings of aziridinyl stannatranes with aryl electrophiles. [11]

A stereoretentive Stille coupling using alkyl carbastannatranes. Stereoretentive Stille Coupling Stannatrane.jpg
A stereoretentive Stille coupling using alkyl carbastannatranes.

Palladium also catalyzes Stille coupling of secondary alkyl carbastannatranes and aryl electrophiles. [12] [13] This report serves as the first example of employing chiral alkyl carbastannatrane reagents in enantioselective synthesis. Related methodology enable selective acyl substitution using enantioenriched stannatranes as an alternative to classical enolate chemistry. [9]

A stannatrane-mediated Stille coupling was utilized for the synthesis of an anti-methicillin-resistant carbapenem to incorporate an entire side-chain in a single step. [14]

Related Research Articles

<span class="mw-page-title-main">Organolithium reagent</span> Chemical compounds containing C–Li bonds

In organometallic chemistry, organolithium reagents are chemical compounds that contain carbon–lithium (C–Li) bonds. These reagents are important in organic synthesis, and are frequently used to transfer the organic group or the lithium atom to the substrates in synthetic steps, through nucleophilic addition or simple deprotonation. Organolithium reagents are used in industry as an initiator for anionic polymerization, which leads to the production of various elastomers. They have also been applied in asymmetric synthesis in the pharmaceutical industry. Due to the large difference in electronegativity between the carbon atom and the lithium atom, the C−Li bond is highly ionic. Owing to the polar nature of the C−Li bond, organolithium reagents are good nucleophiles and strong bases. For laboratory organic synthesis, many organolithium reagents are commercially available in solution form. These reagents are highly reactive, and are sometimes pyrophoric.

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 Suzuki reaction is an organic reaction, classified as a cross-coupling reaction, where the coupling partners are a boronic acid and an organohalide and the catalyst is a palladium(0) complex. It was first published in 1979 by Akira Suzuki, and he shared the 2010 Nobel Prize in Chemistry with Richard F. Heck and Ei-ichi Negishi for their contribution to the discovery and development of palladium-catalyzed cross-couplings in organic synthesis. This reaction is also known as the Suzuki–Miyaura reaction or simply as the Suzuki coupling. It is widely used to synthesize polyolefins, styrenes, and substituted biphenyls. Several reviews have been published describing advancements and the development of the Suzuki reaction. The general scheme for the Suzuki reaction is shown below, where a carbon-carbon single bond is formed by coupling a halide (R1-X) with an organoboron species (R2-BY2) using a palladium catalyst and a base.

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.

The Hiyama coupling is a palladium-catalyzed cross-coupling reaction of organosilanes with organic halides used in organic chemistry to form carbon–carbon bonds. This reaction was discovered in 1988 by Tamejiro Hiyama and Yasuo Hatanaka as a method to form carbon-carbon bonds synthetically with chemo- and regioselectivity. The Hiyama coupling has been applied to the synthesis of various natural products.

The Corey–House synthesis is an organic reaction that involves the reaction of a lithium diorganylcuprate with an organic pseudohalide to form a new alkane, as well as an ill-defined organocopper species and lithium halide as byproducts.

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

Schwartz's reagent is the common name for the organozirconium compound with the formula (C5H5)2ZrHCl, sometimes called zirconocene hydrochloride or zirconocene chloride hydride, and is named after Jeffrey Schwartz, a chemistry professor at Princeton University.This metallocene is used in organic synthesis for various transformations of alkenes and alkynes.

The Negishi coupling is a widely employed transition metal catalyzed cross-coupling reaction. The reaction couples organic halides or triflates with organozinc compounds, forming carbon-carbon bonds (C-C) in the process. A palladium (0) species is generally utilized as the metal catalyst, though nickel is sometimes used. A variety of nickel catalysts in either Ni0 or NiII oxidation state can be employed in Negishi cross couplings such as Ni(PPh3)4, Ni(acac)2, Ni(COD)2 etc.

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

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

Transmetalation (alt. spelling: transmetallation) is a type of organometallic reaction that involves the transfer of ligands from one metal to another. It has the general form:

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.

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

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">Tetramethyltin</span> Chemical compound

Tetramethyltin is an organometallic compound with the formula (CH3)4Sn. This liquid, one of the simplest organotin compounds, is useful for transition-metal mediated conversion of acid chlorides to methyl ketones and aryl halides to aryl methyl ketones. It is volatile and toxic, so care should be taken when using it in the laboratory.

Organomanganese chemistry is the chemistry of organometallic compounds containing a carbon to manganese chemical bond. In a 2009 review, Cahiez et al. argued that as manganese is cheap and benign, organomanganese compounds have potential as chemical reagents, although currently they are not widely used as such despite extensive research.

<span class="mw-page-title-main">Perfluorobutanesulfonyl fluoride</span> Chemical compound

Perfluorobutanesulfonyl fluoride (nonafluorobutanesulfonyl fluoride, NfF) is a colorless, volatile liquid that is immiscible with water but soluble in common organic solvents. It is prepared by the electrochemical fluorination of sulfolane. NfF serves as an entry point to nonafluorobutanesulfonates (nonaflates), which are valuable as electrophiles in palladium catalyzed cross coupling reactions. As a perfluoroalkylsulfonylating agent, NfF offers the advantages of lower cost and greater stability over the more frequently used triflic anhydride. The fluoride leaving group is readily substituted by nucleophiles such as amines, phenoxides, and enolates, giving sulfonamides, aryl nonaflates, and alkenyl nonaflates, respectively. However, it is not attacked by water (in which it is stable at pH<12). Hydrolysis by barium hydroxide gives Ba(ONf)2, which upon treatment with sulfuric acid gives perfluorobutanesulfonic acid and insoluble barium sulfate.

<span class="mw-page-title-main">Palladium–NHC complex</span>

In organometallic chemistry, palladium-NHC complexes are a family of organopalladium compounds in which palladium forms a coordination complex with N-heterocyclic carbenes (NHCs). They have been investigated for applications in homogeneous catalysis, particularly cross-coupling reactions.

Cross electrophile coupling is a type of cross-coupling reaction that occurs between two electrophiles often catalyzed by transition metal catalyst(s). Unlike conventional cross-coupling reactions of an electrophile with an organometallic reagent, the coupling partners in cross electrophile coupling reactions are both electrophiles. Generally, additional reductant to regenerate active catalyst is needed in this reaction.

Miyaura borylation, also known as the Miyaura borylation reaction, is a named reaction in organic chemistry that allows for the generation of boronates from vinyl or aryl halides with the cross-coupling of bis(pinacolato)diboron in basic conditions with a catalyst such as PdCl2(dppf). The resulting borylated products can be used as coupling partners for the Suzuki reaction.

References

  1. Plass, Winfried; Verkade, John G. (1993-11-01). "Azastannatranes: synthesis and structural characterization". Inorganic Chemistry. 32 (23): 5145–5152. doi:10.1021/ic00075a034. ISSN   0020-1669.
  2. Stille, John K. (1986-06-01). "The Palladium-Catalyzed Cross-Coupling Reactions of Organotin Reagents with Organic Electrophiles [New Synthetic Methods (58)]". Angewandte Chemie International Edition in English. 25 (6): 508–524. doi:10.1002/anie.198605081. ISSN   1521-3773.
  3. Hartwig, John Frederick (2010-01-01). Organotransition metal chemistry : from bonding to catalysis. University Science Books. ISBN   9781891389535. OCLC   781082054.
  4. "Column: Totally Synthetic". Chemistry World. Retrieved 2017-04-21.
  5. Jurkschat, K.; Tzschach, A. (1984). "1-Aza-5-stanna-5,5-dimethylbicyclo[3.3.01,5] octan und 1-aza-5-stanna-5-methyltricyclo[3.3.3.01,5] undecan, pentakoordinierte tetraorganozinnverbindungen". Journal of Organometallic Chemistry. 272: C13–C16. doi:10.1016/0022-328X(84)80450-7.
  6. Jurkschat, K.; Tszchach, A.; Meunier-Piret, J. (1986). "Crystal and molecular structure of 1-AZA-5-STANNA-5-methyltricyclo[3.3.3.01,5]undecane. Evidence for a transannular donor−acceptor interaction in a tetraorganotin compound". Journal of Organometallic Chemistry. 315: 45–49. doi:10.1016/0022-328X(86)80409-0.
  7. Crabtree, Robert H. (2014-04-21). The organometallic chemistry of the transition metals. ISBN   9781118138076. OCLC   930855497.
  8. 1 2 Vedejs, Edwin; Haight, Anthony R.; Moss, William O. (1992). "Internal coordination at tin promotes selective alkyl transfer in the Stille coupling reaction". Journal of the American Chemical Society. 114 (16): 6556–6558. doi:10.1021/ja00042a044.
  9. 1 2 Wang, Chao-Yuan; Ralph, Glenn; Derosa, Joseph; Biscoe, Mark R. (2017-01-16). "Stereospecific Palladium-Catalyzed Acylation of Enantioenriched Alkylcarbastannatranes: A General Alternative to Asymmetric Enolate Reactions". Angewandte Chemie International Edition. 56 (3): 856–860. doi:10.1002/anie.201609930. ISSN   1521-3773. PMC   6078499 . PMID   27981696.
  10. Wang, Dong-Yu; Wang, Chao; Uchiyama, Masanobu (2015-08-26). "Stannyl-Lithium: A Facile and Efficient Synthesis Facilitating Further Applications". Journal of the American Chemical Society. 137 (33): 10488–10491. doi:10.1021/jacs.5b06587. ISSN   0002-7863. PMID   26217966.
  11. Theddu, Naresh; Vedejs, Edwin (2013). "Stille Coupling of an Aziridinyl Stannatrane". The Journal of Organic Chemistry. 78 (10): 5061–5066. doi:10.1021/jo4005052. PMID   23581391.
  12. Li, Ling; Wang, Chao-Yuan; Huang, Rongcai; Biscoe, Mark R. (2013). "Stereoretentive Pd-catalysed Stille cross-coupling reactions of secondary alkyl azastannatranes and aryl halides". Nature Chemistry. 5 (7): 607–612. Bibcode:2013NatCh...5..607L. doi:10.1038/nchem.1652. PMC   8048766 . PMID   23787752.
  13. Wang, Chao-Yuan; Derosa, Joseph; Biscoe, Mark R. (2015-08-10). "Configurationally stable, enantioenriched organometallic nucleophiles in stereospecific Pd-catalyzed cross-coupling reactions: an alternative approach to asymmetric synthesis". Chem. Sci. 6 (9): 5105–5113. doi:10.1039/c5sc01710f. ISSN   2041-6539. PMC   4571484 . PMID   26388985.
  14. Jensen, Mark S.; Yang, Chunhua; Hsiao, Yi; Rivera, Nelo; Wells, Kenneth M.; Chung, John Y. L.; Yasuda, Nobuyoshi; Hughes, David L.; Reider, Paul J. (2000-04-01). "Synthesis of an Anti-Methicillin-Resistant Staphylococcus aureus (MRSA) Carbapenem via Stannatrane-Mediated Stille Coupling". Organic Letters. 2 (8): 1081–1084. doi:10.1021/ol005641d. ISSN   1523-7060. PMID   10804559.