Krische allylation

Last updated

The Krische allylation involves the enantioselective iridium-catalyzed addition of an allyl group to an aldehyde or an alcohol, resulting in the formation of a secondary homoallylic alcohol. [1] [2] The mechanism of the Krische allylation involves primary alcohol dehydrogenation or, when using aldehyde reactants, hydrogen transfer from 2-propanol. Unlike other allylation methods, the Krische allylation avoids the use of preformed allyl metal reagents and enables the direct conversion of primary alcohols to secondary homoallylic alcohols (precluding alcohol to aldehyde oxidation). [1] [3]

Contents

insert a caption here General scheme of Krische Allylation.png
insert a caption here

Background

Enantioselective carbonyl allylations are frequently applied to the synthesis of polyketide natural products. [3] In 1978, Hoffmann reported the first asymmetric carbonyl allylation using a chiral allylmetal reagent, an allylborane derived from camphor. [4] [5] Subsequently, other chiral allylmetal reagents were developed by Kumada, Roush, Brown, Leighton, and others. [6] [7] [8] [9] [10] [11] These methods utilize preformed allyl metal reagents and generate stoichiometric quantities of metal byproducts.

In 1991, Yamamoto disclosed the first catalytic enantioselective method for carbonyl allylation, which employed a chiral boron Lewis acid-catalyst in combination with allyltrimethylsilane. [12] Numerous catalytic enantioselective methods for carbonyl allylation followed, including work by Umani-Ronchi [13] and Keck. [14] While these methods had a significant impact, they do not circumvent the use of preformed allylmetal reagents. Catalytic variants of the Nozaki-Hiyama-Kishi reaction represent an alternative method for asymmetric carbonyl allylation, but stoichiometric metallic reductants are required. [15]

Whereas the allylmetal reagents used in these first-generation technologies are often difficult to prepare and handle, the Krische allylation exploits highly tractable allylic acetates. Additionally, the Krische allylation avoids the use of preformed allyl metal reagents or metallic reductants and chiral auxiliaries, significantly reducing waste generation.

Reaction features

The Krische allylation involves “transfer hydrogenative” carbon-carbon bond formations. [16] In a series of papers published in the early 2000s, Krische and coworkers demonstrated that allenes, dienes, and allyl acetates could be converted to transient allylmetal nucleophiles via hydrogenation, transfer hydrogenation or hydrogen auto-transfer. [17] This strategy for enantioselective carbonyl allylation avoids preformed organometallic reagents or metallic reductants. A remarkable feature of these reactions is the ability to conduct carbonyl allylation from the alcohol oxidation state. Due to a kinetic preference for primary alcohol dehydrogenation, diols containing both primary and secondary alcohols undergo site-selective carbonyl allylation at the primary alcohol without the need for protecting groups. [18] Additionally, by using alcohol reactants, the use of chiral α-stereogenic aldehydes, which are prone to racemization, can be avoided. [19]

insert a caption here Scheme 2 MJK.png
insert a caption here

The excellent functional group compatibility of the Krische allylation combined with the tractability of the allyl acetate pronucleophiles enables the use of allyl donors bearing highly complex nitrogen-rich substituents. [20]

insert a caption here Scheme 3 MJK.png
insert a caption here

The figure below shows some of the different allyl donors that have been used in the Krische allylation. These methods are summarized in the review literature. [16] [17]

insert a caption here Scheme 4 MJK.png
insert a caption here

Mechanism

The active catalyst in the Krische allylation is a cyclometallated π-allyliridium C,O-benzoate complex. This complex can be generated in situ or can be isolated via precipitation or conventional chromatography on silica gel.

Catalytic cycle 1-2 Scheme 5 MJK.png
Catalytic cycle 1-2

The mechanism of the Krische allylation has been corroborated by DFT calculations. [21] Entry into the catalytic cycle involves protonation of the cyclometallated π-allyliridium precatalyst to generate the iridium alkoxide I. β-Hydride elimination of alkoxide I generates the aldehyde, which dissociates to form the iridium hydride III. Deprotonation of the iridium hydride III provides an anionic iridium(I) species IV, which upon oxidative addition to the allyl donor forms the π-allyliridium complex V. Association of the aldehyde to the σ-allyliridium species VI triggers carbonyl addition by way of the six-centered transition structure VII to form the homoallylic alkoxide VIII. The homoallylic alkoxide VIII is stable with respect to beta-hydride elimination due to coordination of the double bond with the metal. Exchange with the primary alcohol reactant regenerates the iridium alkoxide I and releases the reaction product.

Catalytic cycle 1-2 Scheme 6 MJK.png
Catalytic cycle 1-2

Applications in synthesis

Iridium-catalyzed transfer-hydrogenative carbonyl allylation method has been applied to the synthesis of polyketide natural products. [3] Some examples are shown below. In every case, the target compound was prepared in significantly fewer steps than was previously achieved. For example, total syntheses of roxaticin, bryostatin and cryptocaryol were accomplished via double Krische allylation of 1,3-propane diol. [22] [23] [24] This method was also used in the synthesis of mandelalide A. [25]

Krische allylation in the synthesis of bryostatin 7, Scheme 7 MJK.png
Krische allylation in the synthesis of bryostatin 7,

The Krische bisallylation has been applied to the synthesis of psymberin in 17 LLS and 32 total steps. [26] Through the use of the Krische allylation, this synthesis was accomplished via a much shorter route than previous syntheses. The Krische allylation to his synthesis of callyspongiolide using the chiral SEGPHOS catalyst complex. [27] ] In 2018, Harran also prepared callyspongiolide using the Krische allylation as a convergent method for fragment union. [28] Double crotylation was used by Krische to prepare 6-deoxyerythronolide B and swinholide A. [29] [30]

Synthesis of Psymberin Scheme 8 MJK.png
Synthesis of Psymberin

Related Research Articles

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

The Pauson–Khand reaction is a chemical reaction described as a [2+2+1] cycloaddition between an alkyne, an alkene and carbon monoxide to form a α,β-cyclopentenone. Ihsan Ullah Khand (1935-1980) discovered the reaction around 1970, while working as a postdoctoral associate with Peter Ludwig Pauson (1925–2013) at the University of Strathclyde in Glasgow. Pauson and Khand's initial findings were intermolecular in nature, but starting a decade after the reaction's discovery, many intramolecular examples have been highlighted in both synthesis and methodology reports. This reaction was originally mediated by stoichiometric amounts of dicobalt octacarbonyl, but newer versions are both more efficient, enhancing reactivity and yield via utilizing different chiral auxiliaries for stereo induction, main group transition-metals, and additives.

<span class="mw-page-title-main">Nucleophilic conjugate addition</span>

Nucleophilic conjugate addition is a type of organic reaction. Ordinary nucleophilic additions or 1,2-nucleophilic additions deal mostly with additions to carbonyl compounds. Simple alkene compounds do not show 1,2 reactivity due to lack of polarity, unless the alkene is activated with special substituents. With α,β-unsaturated carbonyl compounds such as cyclohexenone it can be deduced from resonance structures that the β position is an electrophilic site which can react with a nucleophile. The negative charge in these structures is stored as an alkoxide anion. Such a nucleophilic addition is called a nucleophilic conjugate addition or 1,4-nucleophilic addition. The most important active alkenes are the aforementioned conjugated carbonyls and acrylonitriles.

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

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

<span class="mw-page-title-main">Organocatalysis</span> Method in organic chemistry

In organic chemistry, organocatalysis is a form of catalysis in which the rate of a chemical reaction is increased by an organic catalyst. This "organocatalyst" consists of carbon, hydrogen, sulfur and other nonmetal elements found in organic compounds. Because of their similarity in composition and description, they are often mistaken as a misnomer for enzymes due to their comparable effects on reaction rates and forms of catalysis involved.

Asymmetric hydrogenation is a chemical reaction that adds two atoms of hydrogen to a target (substrate) molecule with three-dimensional spatial selectivity. Critically, this selectivity does not come from the target molecule itself, but from other reagents or catalysts present in the reaction. This allows spatial information to transfer from one molecule to the target, forming the product as a single enantiomer. The chiral information is most commonly contained in a catalyst and, in this case, the information in a single molecule of catalyst may be transferred to many substrate molecules, amplifying the amount of chiral information present. Similar processes occur in nature, where a chiral molecule like an enzyme can catalyse the introduction of a chiral centre to give a product as a single enantiomer, such as amino acids, that a cell needs to function. By imitating this process, chemists can generate many novel synthetic molecules that interact with biological systems in specific ways, leading to new pharmaceutical agents and agrochemicals. The importance of asymmetric hydrogenation in both academia and industry contributed to two of its pioneers — William Standish Knowles and Ryōji Noyori — being awarded one half of the 2001 Nobel Prize in Chemistry.

Chiral Lewis acids (CLAs) are a type of Lewis acid catalyst. These acids affect the chirality of the substrate as they react with it. In such reactions, synthesis favors the formation of a specific enantiomer or diastereomer. The method is an enantioselective asymmetric synthesis reaction. Since they affect chirality, they produce optically active products from optically inactive or mixed starting materials. This type of preferential formation of one enantiomer or diastereomer over the other is formally known as asymmetric induction. In this kind of Lewis acid, the electron-accepting atom is typically a metal, such as indium, zinc, lithium, aluminium, titanium, or boron. The chiral-altering ligands employed for synthesizing these acids often have multiple Lewis basic sites that allow the formation of a ring structure involving the metal atom.

Within the area of organocatalysis, (thio)urea organocatalysis describes the use of ureas and thioureas to accelerate and stereochemically alter organic transformations. The effects arise through hydrogen-bonding interactions between the substrate and the (thio)urea. Unlike classical catalysts, these organocatalysts interact by non-covalent interactions, especially hydrogen bonding. The scope of these small-molecule H-bond donors termed (thio)urea organocatalysis covers both non-stereoselective and stereoselective applications.

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

In organic chemistry, carbonyl reduction is the organic reduction of any carbonyl group by a reducing agent.

<span class="mw-page-title-main">Hydrogen auto-transfer</span>

Hydrogen auto-transfer, also known as borrowing hydrogen, is the activation of a chemical reaction by temporary transfer of two hydrogen atoms from the reactant to a catalyst and return of those hydrogen atoms back to a reaction intermediate to form the final product. Two major classes of borrowing hydrogen reactions exist: (a) those that result in hydroxyl substitution, and (b) those that result in carbonyl addition. In the former case, alcohol dehydrogenation generates a transient carbonyl compound that is subject to condensation followed by the return of hydrogen. In the latter case, alcohol dehydrogenation is followed by reductive generation of a nucleophile, which triggers carbonyl addition. As borrowing hydrogen processes avoid manipulations otherwise required for discrete alcohol oxidation and the use of stoichiometric organometallic reagents, they typically display high levels of atom-economy and, hence, are viewed as examples of Green chemistry.

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

<span class="mw-page-title-main">Hydrogen-bond catalysis</span>

Hydrogen-bond catalysis is a type of organocatalysis that relies on use of hydrogen bonding interactions to accelerate and control organic reactions. In biological systems, hydrogen bonding plays a key role in many enzymatic reactions, both in orienting the substrate molecules and lowering barriers to reaction. However, chemists have only recently attempted to harness the power of using hydrogen bonds to perform catalysis, and the field is relatively undeveloped compared to research in Lewis acid catalysis.

Metal-catalyzed C–H borylation reactions are transition metal catalyzed organic reactions that produce an organoboron compound through functionalization of aliphatic and aromatic C–H bonds and are therefore useful reactions for carbon–hydrogen bond activation. Metal-catalyzed C–H borylation reactions utilize transition metals to directly convert a C–H bond into a C–B bond. This route can be advantageous compared to traditional borylation reactions by making use of cheap and abundant hydrocarbon starting material, limiting prefunctionalized organic compounds, reducing toxic byproducts, and streamlining the synthesis of biologically important molecules. Boronic acids, and boronic esters are common boryl groups incorporated into organic molecules through borylation reactions. Boronic acids are trivalent boron-containing organic compounds that possess one alkyl substituent and two hydroxyl groups. Similarly, boronic esters possess one alkyl substituent and two ester groups. Boronic acids and esters are classified depending on the type of carbon group (R) directly bonded to boron, for example alkyl-, alkenyl-, alkynyl-, and aryl-boronic esters. The most common type of starting materials that incorporate boronic esters into organic compounds for transition metal catalyzed borylation reactions have the general formula (RO)2B-B(OR)2. For example, bis(pinacolato)diboron (B2Pin2), and bis(catecholato)diborane (B2Cat2) are common boron sources of this general formula.

In organic chemistry, the Keck asymmetric allylation is a chemical reaction that involves the nucleophilic addition of an allyl group to an aldehyde. The catalyst is a chiral complex that contains titanium as a Lewis acid. The chirality of the catalyst induces a stereoselective addition, so the secondary alcohol of the product has a predictable absolute stereochemistry based on the choice of catalyst. This name reaction is named for Gary Keck.

In organic chemistry, carbonyl allylation describes methods for adding an allyl anion to an aldehyde or ketone to produce a homoallylic alcohol. The carbonyl allylation was first reported in 1876 by Alexander Zaitsev and employed an allylzinc reagent.

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

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

<span class="mw-page-title-main">Organotantalum chemistry</span> Chemistry of compounds containing a carbon-to-tantalum bond

Organotantalum chemistry is the chemistry of chemical compounds containing a carbon-to-tantalum chemical bond. A wide variety of compound have been reported, initially with cyclopentadienyl and CO ligands. Oxidation states vary from Ta(V) to Ta(-I).

Cobalt(II)–porphyrin catalysis is a process in which a Co(II) porphyrin complex acts as a catalyst, inducing and accelerating a chemical reaction.

The Roskamp reaction was first discovered by Eric J. Roskamp and co-workers in 1989. This reaction is very useful in synthesizing β-keto esters from aldehydes and diazoacetate, using various Lewis acids as catalysts (such as BF3, SnCl2, GeCl2).

<span class="mw-page-title-main">Michael J. Krische</span> American chemist and Robert A (born 1966)

Michael J. Krische is an American chemist and Robert A. Welch Chair in Science at the Department of Chemistry, University of Texas at Austin. Krische has pioneered a broad, new family of catalytic C-C bond formations that occur through the addition or redistribution of hydrogen. These processes merge the characteristics of catalytic hydrogenation and carbonyl addition, contributing to a departure from the use of stoichiometric organometallic reagents in chemical synthesis.

T.V. (Babu) RajanBabu is an organic chemist who holds the position of Distinguished Professor of Chemistry in the College of Arts and Sciences at the Ohio State University. His laboratory traditionally focuses on developing transition metal-catalyzed reactions. RajanBabu is known for helping develop the Nugent-RajanBabu reagent, a chemical reagent used in synthetic organic chemistry as a single electron reductant.

References

  1. 1 2 Kim, In Su; Ngai, Ming-Yu; Krische, Michael J. (2008-11-05). "Enantioselective Iridium-Catalyzed Carbonyl Allylation from the Alcohol or Aldehyde Oxidation Level via Transfer Hydrogenative Coupling of Allyl Acetate: Departure from Chirally Modified Allyl Metal Reagents in Carbonyl Addition". Journal of the American Chemical Society. 130 (44): 14891–14899. doi:10.1021/ja805722e. ISSN   0002-7863. PMC   2890235 . PMID   18841896.
  2. Strategies and Tactics in Organic Synthesis, Volume 10 Michael Harmata Ed.
  3. 1 2 3 Feng, Jiajie; Kasun, Zachary A.; Krische, Michael J. (2016-05-04). "Enantioselective Alcohol C–H Functionalization for Polyketide Construction: Unlocking Redox-Economy and Site-Selectivity for Ideal Chemical Synthesis". Journal of the American Chemical Society. 138 (17): 5467–5478. doi:10.1021/jacs.6b02019. ISSN   0002-7863. PMC   4871165 . PMID   27113543.
  4. Herold, Thomas; Hoffmann, Reinhard W. (October 1978). "Enantioselective Synthesis of Homoallyl Alcoholsvia Chiral Allylboronic Esters". Angewandte Chemie International Edition in English. 17 (10): 768–769. doi:10.1002/anie.197807682. ISSN   0570-0833.
  5. Hoffmann, Reinhard W.; Herold, Thomas (January 1981). "Stereoselektive Synthese von Alkoholen, VII1) Optisch aktive Homoallylalkohole durch Addition chiraler Boronsäureester an Aldehyde". Chemische Berichte (in German). 114 (1): 375–383. doi:10.1002/cber.19811140139.
  6. Hayashi, Tamio; Konishi, Mitsuo; Kumada, Makoto (September 1982). "Optically active allylsilanes. 2. High stereoselectivity in asymmetric reaction with aldehydes producing homoallylic alcohols". Journal of the American Chemical Society. 104 (18): 4963–4965. doi:10.1021/ja00382a046. ISSN   0002-7863.
  7. Brown, Herbert C.; Jadhav, Prabhakar K. (April 1983). "Asymmetric carbon-carbon bond formation via .beta.-allyldiisopinocampheylborane. Simple synthesis of secondary homoallylic alcohols with excellent enantiomeric purities". Journal of the American Chemical Society. 105 (7): 2092–2093. doi:10.1021/ja00345a085. ISSN   0002-7863.
  8. Roush, William R.; Walts, Alan E.; Hoong, Lee K. (December 1985). "Diastereo- and enantioselective aldehyde addition reactions of 2-allyl-1,3,2-dioxaborolane-4,5-dicarboxylic esters, a useful class of tartrate ester modified allylboronates". Journal of the American Chemical Society. 107 (26): 8186–8190. doi:10.1021/ja00312a062. ISSN   0002-7863.
  9. Kinnaird, James W. A.; Ng, Pui Yee; Kubota, Katsumi; Wang, Xiaolun; Leighton, James L. (2002-07-01). "Strained Silacycles in Organic Synthesis: A New Reagent for the Enantioselective Allylation of Aldehydes". Journal of the American Chemical Society. 124 (27): 7920–7921. doi:10.1021/ja0264908. ISSN   0002-7863. PMID   12095334.
  10. Short, Robert P.; Masamune, Satoru (March 1989). "Asymmetric allylboration with B-allyl-2-(trimethylsilyl)borolane". Journal of the American Chemical Society. 111 (5): 1892–1894. doi:10.1021/ja00187a061. ISSN   0002-7863.
  11. Corey, E. J.; Yu, Chan Mo; Kim, Sung Soo (July 1989). "A practical and efficient method for enantioselective allylation of aldehydes". Journal of the American Chemical Society. 111 (14): 5495–5496. doi:10.1021/ja00196a082. ISSN   0002-7863.
  12. Furuta, Kyoji; Mouri, Makoto; Yamamoto, Hisashi (1991). "Chiral (Acyloxy)borane Catalyzed Asymmetric Allylation of Aldehydes". Synlett. 1991 (8): 561–562. doi:10.1055/s-1991-20797. ISSN   0936-5214.
  13. Costa, Anna Luisa; Piazza, Maria Giulia; Tagliavini, Emilio; Trombini, Claudio; Umani-Ronchi, Achille (July 1993). "Catalytic asymmetric synthesis of homoallylic alcohols". Journal of the American Chemical Society. 115 (15): 7001–7002. doi:10.1021/ja00068a079. ISSN   0002-7863.
  14. Keck, Gary E.; Tarbet, Kenneth H.; Geraci, Leo S. (September 1993). "Catalytic asymmetric allylation of aldehydes". Journal of the American Chemical Society. 115 (18): 8467–8468. doi:10.1021/ja00071a074. ISSN   0002-7863.
  15. Hargaden, Gráinne C.; Guiry, Patrick J. (2007-11-05). "The Development of the Asymmetric Nozaki–Hiyama–Kishi Reaction". Advanced Synthesis & Catalysis. 349 (16): 2407–2424. doi:10.1002/adsc.200700324.
  16. 1 2 Santana, Catherine Gazolla; Krische, Michael J. (2021-05-07). "From Hydrogenation to Transfer Hydrogenation to Hydrogen Auto-Transfer in Enantioselective Metal-Catalyzed Carbonyl Reductive Coupling: Past, Present, and Future". ACS Catalysis. 11 (9): 5572–5585. doi:10.1021/acscatal.1c01109. ISSN   2155-5435. PMC   8302072 . PMID   34306816.
  17. 1 2 Kim, Seung Wook; Zhang, Wandi; Krische, Michael J. (2017-09-19). "Catalytic Enantioselective Carbonyl Allylation and Propargylation via Alcohol-Mediated Hydrogen Transfer: Merging the Chemistry of Grignard and Sabatier". Accounts of Chemical Research. 50 (9): 2371–2380. doi:10.1021/acs.accounts.7b00308. ISSN   0001-4842. PMC   5641472 . PMID   28792731.
  18. Dechert-Schmitt, Anne-Marie R.; Schmitt, Daniel C.; Krische, Michael J. (2013-03-11). "Protecting-Group-Free Diastereoselective CC Coupling of 1,3-Glycols and Allyl Acetate through Site-Selective Primary Alcohol Dehydrogenation". Angewandte Chemie International Edition. 52 (11): 3195–3198. doi:10.1002/anie.201209863. PMC   3711384 . PMID   23364927.
  19. Schmitt, Daniel C.; Dechert-Schmitt, Anne-Marie R.; Krische, Michael J. (2012-12-21). "Iridium-Catalyzed Allylation of Chiral β-Stereogenic Alcohols: Bypassing Discrete Formation of Epimerizable Aldehydes". Organic Letters. 14 (24): 6302–6305. doi:10.1021/ol3030692. ISSN   1523-7060. PMC   3529126 . PMID   23231774.
  20. Meyer, Cole C.; Stafford, Nicholas P.; Cheng, Melinda J.; Krische, Michael J. (2021-05-03). "Ethanol: Unlocking an Abundant Renewable C 2 ‐Feedstock for Catalytic Enantioselective C−C Coupling". Angewandte Chemie International Edition. 60 (19): 10542–10546. doi:10.1002/anie.202102694. ISSN   1433-7851. PMC   8085048 . PMID   33689214.
  21. Kim, Seung Wook; Meyer, Cole C.; Mai, Binh Khanh; Liu, Peng; Krische, Michael J. (2019-10-04). "Inversion of Enantioselectivity in Allene Gas versus Allyl Acetate Reductive Aldehyde Allylation Guided by Metal-Centered Stereogenicity: An Experimental and Computational Study". ACS Catalysis. 9 (10): 9158–9163. doi:10.1021/acscatal.9b03695. ISSN   2155-5435. PMC   6921087 . PMID   31857913.
  22. Han, Soo Bong; Hassan, Abbas; Kim, In Su; Krische, Michael J. (2010-11-10). "Total Synthesis of (+)-Roxaticin via C−C Bond Forming Transfer Hydrogenation: A Departure from Stoichiometric Chiral Reagents, Auxiliaries, and Premetalated Nucleophiles in Polyketide Construction". Journal of the American Chemical Society. 132 (44): 15559–15561. doi:10.1021/ja1082798. ISSN   0002-7863. PMC   2975273 . PMID   20961111.
  23. Lu, Yu; Woo, Sang Kook; Krische, Michael J. (2011-09-07). "Total Synthesis of Bryostatin 7 via C–C Bond-Forming Hydrogenation". Journal of the American Chemical Society. 133 (35): 13876–13879. doi:10.1021/ja205673e. ISSN   0002-7863. PMC   3164899 . PMID   21780806.
  24. Perez, Felix; Waldeck, Andrew R.; Krische, Michael J. (2016-04-11). "Total Synthesis of Cryptocaryol A by Enantioselective Iridium-Catalyzed Alcohol C−H Allylation". Angewandte Chemie International Edition. 55 (16): 5049–5052. doi:10.1002/anie.201600591. PMC   4834877 . PMID   27079820.
  25. Willwacher, Jens; Fürstner, Alois (2014-04-14). "Catalysis-Based Total Synthesis of Putative Mandelalide A". Angewandte Chemie International Edition. 53 (16): 4217–4221. doi:10.1002/anie.201400605. PMID   24623640.
  26. Feng, Yu; Jiang, Xin; De Brabander, Jef K. (2012-10-17). "Studies toward the Unique Pederin Family Member Psymberin: Full Structure Elucidation, Two Alternative Total Syntheses, and Analogs". Journal of the American Chemical Society. 134 (41): 17083–17093. doi:10.1021/ja3057612. ISSN   0002-7863. PMC   3482988 . PMID   23004238.
  27. Zhou, Jingjing; Gao, Bowen; Xu, Zhengshuang; Ye, Tao (2016-06-08). "Total Synthesis and Stereochemical Assignment of Callyspongiolide". Journal of the American Chemical Society. 138 (22): 6948–6951. doi:10.1021/jacs.6b03533. ISSN   0002-7863. PMID   27227371.
  28. Manoni, Francesco; Rumo, Corentin; Li, Liubo; Harran, Patrick G. (2018-01-31). "Unconventional Fragment Usage Enables a Concise Total Synthesis of (−)-Callyspongiolide". Journal of the American Chemical Society. 140 (4): 1280–1284. doi:10.1021/jacs.7b13591. ISSN   0002-7863. PMID   29332397.
  29. Gao, Xin; Woo, Sang Kook; Krische, Michael J. (2013-03-20). "Total Synthesis of 6-Deoxyerythronolide B via C–C Bond-Forming Transfer Hydrogenation". Journal of the American Chemical Society. 135 (11): 4223–4226. doi:10.1021/ja4008722. ISSN   0002-7863. PMC   3625983 . PMID   23464668.
  30. Shin, Inji; Hong, Suckchang; Krische, Michael J. (2016-11-02). "Total Synthesis of Swinholide A: An Exposition in Hydrogen-Mediated C–C Bond Formation". Journal of the American Chemical Society. 138 (43): 14246–14249. doi:10.1021/jacs.6b10645. ISSN   0002-7863. PMC   5096380 . PMID   27779393.