Liebeskind–Srogl coupling

Last updated

Contents

Liebeskind–Srogl coupling
Named afterLanny S. Liebeskind
Jiri Srogl
Reaction type Coupling reaction
Identifiers
RSC ontology ID RXNO:0000604

The Liebeskind–Srogl coupling reaction is an organic reaction forming a new carbon–carbon bond from a thioester and a boronic acid using a metal catalyst. It is a cross-coupling reaction. [1] This reaction was invented by and named after Jiri Srogl from the Academy of Sciences, Czech Republic, and Lanny S. Liebeskind from Emory University, Atlanta, Georgia, USA. There are three generations of this reaction, with the first generation shown below. The original transformation used catalytic Pd(0), TFP = tris(2-furyl)phosphine as an additional ligand and stoichiometric CuTC = copper(I) thiophene-2-carboxylate as a co-metal catalyst. The overall reaction scheme is shown below.

The Liebeskind-Srogl coupling reaction Liebeskind-Strogl coupling2.png
The Liebeskind-Srogl coupling reaction

Liebeskind-Srogl reaction is most commonly seen with sulfide or thioester electrophiles and boronic acid or stannane nucleophiles but many other coupling partners are viable. In addition to alkyl and aryl thioesters; (hetero)aryl sulfides, thioamides, sulfanyl alkynes, and thiocyanates are competent electrophiles. [2] Virtually any metal-R bond capable of transmetalation has been demonstrated. [2] Indium derived nucleophiles require no copper or base. Note that this scope is applicable for the first generation coupling as the second and third generations are mechanistically distinct and have only been demonstrated with thioesters capable of forming the six-membered metallocycle, boronic acids, and stannanes.

ScopeofLSC.jpg

The first-generation approach to cross coupling is run under anaerobic conditions using stoichiometric copper and catalytic palladium. [1]

Second generation approach renders the reactions catalytic in copper by using an extra equivalent of boronic acid under aerobic, palladium free conditions. [3] The additional equivalent liberates the copper from the sulfur auxiliary and allows it to turn over. This chemistry is limited to thioesters and sulfides and could also be limited by the cost and availability of the organoboron reagent.

The third generation renders the reaction catalytic in copper while using only one equivalent of boronic acid. [4]

Mechanism

Generation 1

The proposed reaction mechanism for the first generation is shown below. [5] [6] The thioester 1 complexes with copper complex 3 to form compound 4. With the oxidative insertion of [Pd] into the carbon–sulfur bond, compound 5 is formed, and with transmetallation, organopalladium species 8 is formed. The transmetallation proceeds via the transfer of R2 to the palladium metal center with concomitant transfer of the sulfur atom to the copper complex. Reductive elimination gives ketone 3 with the regeneration of the active catalyst 9.

The Liebeskind-Srogl coupling mechanism Liebeskind-Strogl mechanism.png
The Liebeskind–Srogl coupling mechanism

Generation 2

The mechanism for the second generation is shown below. [3] The mechanism does not follow a traditional oxidative addition-transmetelation-reductive elimination pathway like the first generation. In parallel to studies of Cu(I)-dioxygen reactions, a higher oxidation state, Cu-templated coupling is proposed. [7] [8] [9] [10] [11] Coordination of copper(I) to the thioester undergoes oxidation by air to give a copper (II/III) intermediate. Metal templating by Cu(II/III) acts as a Lewis acid to both activate the thiol ester and deliver R2 (from either boron directly or via an intermediate Cu-R2 species), which produces the ketone and a Cu-thiolate. A second equivalent of boronic acid is needed to break the copper sulfur bond and liberate copper back into the catalytic cycle.

Gen2bitch.jpg

Generation 3

The third generation renders the reaction catalytic in copper and uses only one equivalent of boronic acid by mimicking the metallothionein (MT) system that sponges metals from biological systems. [4] The thio-auxilary features an N-O motif that mimics the S-S motif in the MT biosystem, that is necessary to break the copper sulfur bond and turn over the catalyst. This generation is palladium free and under microwave conditions. The mechanism is expected to follow that of the second generation (shown as an active Cu(I)-R2 species but R2 could be delivered directly from the coordinated boronic acid) but includes the auxiliary releasing copper back into the catalytic cycle instead of additional boronic acid.

Gen3.0.jpg

Applications in synthesis

The Liebeskind–Srogl coupling has been used as a key retrosynthetic disconnection in several natural product total synthesis.

For example, in the synthesis of Goniodomin A, the Sasakki lab utilized this chemistry to rapidly access the northern half of the natural product. [12]                      

Goniodomin.jpg

The Guerrero lab used the Liebeskind–Srogl coupling to construct the entire carbon skeleton of viridin in high yield on multi-gram scale. [13]

Viridin.jpg

The lab of Figadere used the Liebeskind–Srogl coupling early in their synthesis of amphidinolide F [14] by employing this reaction to construct the north eastern fragment of the macrocycle and the terpene chain.

Amphidinolide.jpg

Other

Directed difunctionalization

The Yu lab has demonstrated that in the presence of two sulfide bonds, one can be selectively functionalized in the presence of one equivalent of nucleophile if directed by a carbonyl oxygen. [15] This reaction proceeds through a five-membered palladacycle with oxidative addition taking place on this cis-thioether. Additional equivalence of nucleophile will functionalize the trans-position.

Yu.directed.lsc.jpg

Related Research Articles

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

The Wacker process or the Hoechst-Wacker process refers to the oxidation of ethylene to acetaldehyde in the presence of palladium(II) chloride as the catalyst. This chemical reaction was one of the first homogeneous catalysis with organopalladium chemistry applied on an industrial scale.

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.

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">Boronic acid</span> Organic compound of the form R–B(OH)2

A boronic acid is an organic compound related to boric acid in which one of the three hydroxyl groups is replaced by an alkyl or aryl group. As a compound containing a carbon–boron bond, members of this class thus belong to the larger class of organoboranes.

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.

<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.

Metal carbon dioxide complexes are coordination complexes that contain carbon dioxide ligands. Aside from the fundamental interest in the coordination chemistry of simple molecules, studies in this field are motivated by the possibility that transition metals might catalyze useful transformations of CO2. This research is relevant both to organic synthesis and to the production of "solar fuels" that would avoid the use of petroleum-based fuels.

<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.

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

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

The Catellani reaction was discovered by Marta Catellani and co-workers in 1997. The reaction uses aryl iodides to perform bi- or tri-functionalization, including C-H functionalization of the unsubstituted ortho position(s), followed a terminating cross-coupling reaction at the ipso position. This cross-coupling cascade reaction depends on the ortho-directing transient mediator, norbornene.

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">Krische allylation</span>

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. 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.

In organic chemistry, the Fujiwara–Moritani reaction is a type of cross coupling reaction where an aromatic C-H bond is directly coupled to an olefinic C-H bond, generating a new C-C bond. This reaction is performed in the presence of a transition metal, typically palladium. The reaction was discovered by Yuzo Fujiwara and Ichiro Moritani in 1967. An external oxidant is required to this reaction to be run catalytically. Thus, this reaction can be classified as a C-H activation reaction, an oxidative Heck reaction, and a C-H olefination. Surprisingly, the Fujiwara–Moritani reaction was discovered before the Heck reaction.

Dialkylbiaryl phosphine ligands are phosphine ligands that are used in homogeneous catalysis. They have proved useful in Buchwald-Hartwig amination and etherification reactions as well as Negishi cross-coupling, Suzuki-Miyaura cross-coupling, and related reactions. In addition to these Pd-based processes, their use has also been extended to transformations catalyzed by nickel, gold, silver, copper, rhodium, and ruthenium, among other transition metals.

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

Sukbok Chang is a South Korean organic chemist. He is a distinguished professor in the Department of Chemistry at Korea Advanced Institute of Science and Technology (KAIST). He is also the director of the Institute for Basic Science (IBS) Center for Catalytic Hydrocarbon Functionalizations (CCHF). He was an associate editor on ACS Catalysis and has served on the editorial advisory boards of The Journal of Organic Chemistry, Journal of the American Chemical Society, and Accounts of Chemical Research. His major research interest is transition metal catalyzed C-H bond functionalization for the carbon-carbon bond and carbon-heteroatom bond formation.

F. Dean Toste is the Gerald E. K. Branch Distinguished Professor of Chemistry at the University of California, Berkeley and Faculty Scientist at the Chemical Sciences Division of Lawrence Berkeley National Lab. He is a prominent figure in the field of organic chemistry and is best known for his contributions to gold chemistry and asymmetric ion-pairing catalysis. Toste was elected a member of the National Academy of Sciences in 2020, and a member of the American Academy of Arts and Sciences in 2018.

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

The nitro-Mannich reaction is the nucleophilic addition of a nitroalkane to an imine, resulting in the formation of a beta-nitroamine. With the reaction involving the addition of an acidic carbon nucleophile to a carbon-heteroatom double bond, the nitro-Mannich reaction is related to some of the most fundamental carbon-carbon bond forming reactions in organic chemistry, including the aldol reaction, Henry reaction and Mannich reaction.

<span class="mw-page-title-main">Strained phosphorus heterocycles</span> Class of main group element compounds

Strained phosphorus containing heterocycles are a class of main group element compounds that uniquely demonstrates chemical reactivities such as oxidative addition (OA), reductive elimination (RE), and O-atom transfer, which are in general associated with transition metals. Key to their success in "main group as transition metal" are two factors: (1) The center atom phosphorus shows moderate tendency to both oxidative addition and reductive elimination. In contrast, while OA is facile with low valent group 13/14 elements such as Al(I), carbene and silylene, RE is considered very difficult. Similarly, while RE on hypervalent group 16/17 compounds such as S(IV) and I(III) is facile, the OA is challenging. Group 15 pnictogens show advantage over their neighbors in terms of their comparative ease in both OA and RE.(2) Geometry distortion due to ring strain leads to the non-degeneracy of the e set of orbitals (LUMO) in C3v symmetry. The decreased HOMO-LUMO gap enables better interaction between the molecules, leading to easier activation and novel chemistry opportunities. Additionally, the HOMO can be raised in diphosphetanes due to interaction between the lone pairs.

References

  1. 1 2 Liebeskind, L.; Srogl, Jiri (2000). "Thiol Ester−Boronic Acid Coupling. A Mechanistically Unprecedented and General Ketone Synthesis". J. Am. Chem. Soc. 122 (45): 11260–11261. doi:10.1021/ja005613q.
  2. 1 2 Cheng, Hong-Gang; Chen, Han; Liu, Yue; Zhou, Qianghui (March 2018). "The Liebeskind-Srogl Cross-Coupling Reaction and its Synthetic Applications". Asian Journal of Organic Chemistry. 7 (3): 490–508. doi: 10.1002/ajoc.201700651 .
  3. 1 2 Villalobos, Janette M.; Srogl, Jiri; Liebeskind, Lanny S. (December 2007). "A New Paradigm for Carbon−Carbon Bond Formation: Aerobic, Copper-Templated Cross-Coupling". Journal of the American Chemical Society. 129 (51): 15734–15735. doi:10.1021/ja074931n. ISSN   0002-7863. PMC   2561227 . PMID   18047333.
  4. 1 2 Zhang, Zhihui; Lindale, Matthew G.; Liebeskind, Lanny S. (27 April 2011). "Mobilizing Cu(I) for Carbon−Carbon Bond Forming Catalysis in the Presence of Thiolate. Chemical Mimicking of Metallothioneins". Journal of the American Chemical Society. 133 (16): 6403–6410. doi:10.1021/ja200792m. ISSN   0002-7863. PMC   3128984 . PMID   21449537.
  5. Yu, Y.; Liebeskind, L. S. (2004). "Copper-mediated, palladium-catalyzed coupling of thiol esters with aliphatic organoboron reagents". J. Org. Chem. 69 (10): 3554–3557. doi:10.1021/jo049964p. PMID   15132570.
  6. ^ Villalobos, J. M.; Srogl, J.; Liebeskind, L. S. (2007). "A new paradigm for carbon–carbon bond formation: aerobic, copper-templated cross-coupling". J. Am. Chem. Soc. 129 (51): 15734–15735. doi:10.1021/ja074931n. PMC   2561227 . PMID   18047333.
  7. Hatcher, Lanying Q.; Vance, Michael A.; Narducci Sarjeant, Amy A.; Solomon, Edward I.; Karlin, Kenneth D. (April 2006). "Copper−Dioxygen Adducts and the Side-on Peroxo Dicopper(II)/Bis(μ-oxo) Dicopper(III) Equilibrium: Significant Ligand Electronic Effects". Inorganic Chemistry. 45 (7): 3004–3013. doi:10.1021/ic052185m. ISSN   0020-1669. PMID   16562956.
  8. Mirica, Liviu M.; Rudd, Deanne Jackson; Vance, Michael A.; Solomon, Edward I.; Hodgson, Keith O.; Hedman, Britt; Stack, T. Daniel P. (March 2006). "μ-η2:η2-Peroxodicopper(II) Complex with a Secondary Diamine Ligand: A Functional Model of Tyrosinase". Journal of the American Chemical Society. 128 (8): 2654–2665. doi:10.1021/ja056740v. ISSN   0002-7863. PMID   16492052.
  9. Matsumoto, Takahiro; Furutachi, Hideki; Kobino, Masashi; Tomii, Masato; Nagatomo, Shigenori; Tosha, Takehiko; Osako, Takao; Fujinami, Shuhei; Itoh, Shinobu (March 2006). "Intramolecular Arene Hydroxylation versus Intermolecular Olefin Epoxidation by (μ-η2:η2-Peroxo)dicopper(II) Complex Supported by Dinucleating Ligand". Journal of the American Chemical Society. 128 (12): 3874–3875. doi:10.1021/ja058117g. ISSN   0002-7863. PMID   16551071.
  10. Lewis, Elizabeth A.; Tolman, William B. (February 2004). "Reactivity of Dioxygen−Copper Systems". Chemical Reviews. 104 (2): 1047–1076. doi:10.1021/cr020633r. ISSN   0009-2665. PMID   14871149.
  11. Chemical Reviews. 104 (8): 6. 11 August 2004. doi:10.1021/cr040141+. ISSN   0009-2665.{{cite journal}}: Missing or empty |title= (help)
  12. Saito, Tomoyuki; Fuwa, Haruhiko; Sasaki, Makoto (19 November 2009). "Toward the Total Synthesis of Goniodomin A, An Actin-Targeting Marine Polyether Macrolide: Convergent Synthesis of the C15−C36 Segment". Organic Letters. 11 (22): 5274–5277. doi:10.1021/ol902217q. ISSN   1523-7060. PMID   19905029.
  13. Del Bel, Matthew; Abela, Alexander R.; Ng, Jeffrey D.; Guerrero, Carlos A. (24 May 2017). "Enantioselective Chemical Syntheses of the Furanosteroids (−)-Viridin and (−)-Viridiol". Journal of the American Chemical Society. 139 (20): 6819–6822. doi:10.1021/jacs.7b02829. ISSN   0002-7863. PMID   28463562.
  14. Ferrié, Laurent; Fenneteau, Johan; Figadère, Bruno (June 2018). "Total Synthesis of the Marine Macrolide Amphidinolide F" (PDF). Organic Letters. 20 (11): 3192–3196. doi:10.1021/acs.orglett.8b01020. ISSN   1523-7060. PMID   29762038.
  15. Jin, Weiwei; Du, Wangming; Yang, Qin; Yu, Haifeng; Chen, Jiping; Yu, Zhengkun (19 August 2011). "Regio- and Stereoselective Synthesis of Multisubstituted Olefins and Conjugate Dienes by Using α-Oxo Ketene Dithioacetals as the Building Blocks". Organic Letters. 13 (16): 4272–4275. doi:10.1021/ol201620g. ISSN   1523-7060. PMID   21761823.