Next Article in Journal
Variation of Vitamin D in Cow’s Milk and Interaction with β-Lactoglobulin
Next Article in Special Issue
Palladium-Catalyzed Synthesis of Natural and Unnatural 2-, 5-, and 7-Oxygenated Carbazole Alkaloids from N-Arylcyclohexane Enaminones
Previous Article in Journal
Enhancement of Cunninghamella elegans UCP/WFCC 0542 Biomass and Chitosan with Amino Acid Supply
Previous Article in Special Issue
Enhanced Stability of CaO and/or La2O3 Promoted Pd/Al2O3 Egg-Shell Catalysts in Partial Oxidation of Methane to Syngas
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Palladium and Organocatalysis: An Excellent Recipe for Asymmetric Synthesis

1
Departamento de Química Orgánica, Facultad de Ciencias, Universidad Autónoma de Madrid, Cantoblanco, Madrid 28049, Spain
2
Division of Chemistry and Environmental Science, Faculty of Science and Engineering, Manchester Metropolitan University, John Dalton Extension, Oxford Road, Manchester M1 5GD, UK
3
Departamento de Química Orgánica, Facultad de Ciencias and Instituto de Síntesis Orgánica (ISO), Universidad de Alicante, Apdo. 99, Alicante 03080, Spain
*
Authors to whom correspondence should be addressed.
Molecules 2013, 18(9), 10108-10121; https://doi.org/10.3390/molecules180910108
Submission received: 16 July 2013 / Revised: 9 August 2013 / Accepted: 15 August 2013 / Published: 22 August 2013
(This article belongs to the Special Issue Palladium Catalysts)

Abstract

:
The dual activation of simple substrates by the combination of organocatalysis and palladium catalysis has been successfully applied in a variety of different asymmetric transformations. Thus, the asymmetric α-allylation of carbonyl compounds, α-fluorination of acyl derivatives, decarboxylative protonation of β-dicarbonyl compounds, cyclization reactions of alkynyl carbonyl compounds and β-functionalization of aldehydes have been efficiently achieved employing this double-catalytic methodology.

1. Introduction

Organocatalysis has represented, in this new century, a breakthrough in the field of asymmetric organic synthesis [1], becoming an attractive complementary alternative for the well established transition-metal catalysis [2,3]. It was just a matter of time before both strategies were successfully combined for the improvement of catalytic processes [4,5,6]. The combination of organocatalysis and transition-metal catalysis allows not only the development of novel transformations, but also the enhancement of the stereochemistry, efficiency and/or scope of well known reactions [7]. One of the main challenges to overcome when merging these two systems, is tuning their compatibility to avoid negative interferences between them. Amongst transition metal catalysis, palladium-catalyzed reactions have gained a predominant place in the arsenal of synthetic chemists [2,8,9,10,11,12] and their combination with organocatalysts has resulted into very efficient cooperative dual systems. This review compiles the contributions of this cooperative dual catalytic system on asymmetric transformations.

2. Asymmetric α-Allylic Alkylation

The α-alkylation to carbonyl moieties is a classical reaction for the formation of carbon-carbon bonds in organic synthesis [13]. In particular, the asymmetric α-allylic alkylation stands out as a versatile methodology [14] where palladium complexes are of paramount importance, due to the easy generation of π-allyl intermediates as electrophiles [15,16]. Recently, α-allylation of carbonyl compounds has been achieved by merging the formation of π-allyl-palladium and the organocatalytic generation of nucleophilic species. In 2001, the asymmetric allylation of tert-butyl N-(diphenylmethylene)glycinate by the combination of the palladium complex [Pd(allyl)Cl]2, in the presence of PPh3 as ligand and salt 1 as chiral phase transfer catalyst (PTC) (Figure 1) was achieved, with a maximum ee of 61% [17]. The authors describe that the use of chiral bidentate ligands, such as BINAP, does not improve the outcome of the reaction, whilst the use of molecular sieves is crucial to obtain high enantioselectivities.
Figure 1. Chiral organocatalysts for α-allylation reactions.
Figure 1. Chiral organocatalysts for α-allylation reactions.
Molecules 18 10108 g001
Scheme 1. Allylic alkylation catalyzed by 2 and [Pd(C3H5)Cl]2/P(OPh)3.
Scheme 1. Allylic alkylation catalyzed by 2 and [Pd(C3H5)Cl]2/P(OPh)3.
Molecules 18 10108 g007
Soon after, Takemoto and co-workers reported the allylation of glycinate derivatives with allylic acetates under similar reaction conditions: cinchonidinium salt 2 (10 mol%, Figure 1), [Pd(C3H5)Cl]2 (9 mol%), and P(OPh)3 (20 mol%) as ligand (Scheme 1) [18,19]. In this case, the use of a phosphite ligand is essential to achieve good enantioselectivities, which fall in the range 91%–96% ee. Indeed, the authors claim that more σ-donating ligands, such as phosphines, form a more reactive allyl-palladium complex which favors the reaction with the enolate in the absence of the chiral ion pair, and yields the corresponding product with poor enantioselectivity.
The α-allylation of aldehydes by combination of an amine organocatalyst and a palladium complex, which generate a nucleophilic enamine and an electrophilic π-allyl palladium intermediate, respectively, has been described by the group of Córdova [20,21]. The co-catalyzed asymmetric allylic alkylations are carried out with the chiral pyrrolidine 3 (Figure 1) and [Pd(PPh3)4]. Thus, the corresponding alcohols, after in situ reduction of the allylated aldehydes, are obtained in high yields (50%–85%) and enantioselectivities ranging from 88% to 96% (Scheme 2). Interestingly, the sequence α-allylation of aldehydes and reductive amination allows the preparation of chiral amines in 41%–66% yield [21].
Scheme 2. Allylic alkylation of aldehydes catalyzed by 3 and [Pd(PPh3)4].
Scheme 2. Allylic alkylation of aldehydes catalyzed by 3 and [Pd(PPh3)4].
Molecules 18 10108 g008
More recently, the intramolecular version of this transformation has been reported with the formation of 5 and 6-membered rings starting from the corresponding 6- or 7-alkenal derivatives [22,23]; albeit an excess of an achiral secondary amine has to be employed [24] in combination with a chiral palladium complex, in order to obtain certain enantioselectivity. Contrary to this, the intramolecular cyclization of analogous allenic aldehydes, i.e., octa-6,7-dienal derivatives, was achieved using catalytic amounts of the chiral secondary amine 4 and Pd(OAc)2, without any phosphine ligand. This methodology provides the corresponding cyclopentanecarbaldehydes in good trans-diastereoselection (up to 20:1) and enantioselectivities up to 82% ee [25] (Scheme 3).
Scheme 3. Amine/palladium co-catalyzed asymmetric cyclization.
Scheme 3. Amine/palladium co-catalyzed asymmetric cyclization.
Molecules 18 10108 g009
The group of List has reported the combination of a Pd(0) catalyst with the chiral phosphoric acid 5 as an effective formula for the enantioselective α-allylation of branched aldehydes with allylic amines [26] or allylic alcohols [27]. Thus, different aldehydes were reacted with N-benzhydrylallylamines in the presence of phosphoric acid 5 (1.5 mol%), Pd(PPh3)4 (3 mol%) and molecular sieves in tert-butyl methyl ether, to form the corresponding allylated aldehydes with α-quaternary stereogenic centers (Scheme 4) [26].
Scheme 4. Enantioselective Pd/chiral phosphoric acid catalyzed α-allylation of aldehydes.
Scheme 4. Enantioselective Pd/chiral phosphoric acid catalyzed α-allylation of aldehydes.
Molecules 18 10108 g010
The authors suggest that the condensation of the aldehyde and the secondary amine in the presence of 5 produces the enamonium salt 5A, which affords the intermediate 5B in the presence of a Pd(0) complex (Scheme 5).
Scheme 5. Proposed mechanism for the α-allylation of aldehydes in the presence of a chiral phosphoric acid and Pd(0) complex.
Scheme 5. Proposed mechanism for the α-allylation of aldehydes in the presence of a chiral phosphoric acid and Pd(0) complex.
Molecules 18 10108 g011
Finally, 5B leads to the allylated iminium salt 5C via nucleophilic attack of the enamine onto the π-allyl-palladium complex (Scheme 5). Regarding the use of allylic alcohols as allylating agents under the same reaction conditions, the results are not as satisfactory, and the expected products are obtained with lower enantiomeric excess (c.a. 10% ee), probably due to the formation of both E- and Z-enol isomers [27]. This problem can be overcome by using a secondary amine, such as benzhydryl amine (40 mol%), to form the corresponding enamine in situ. In this way, the α-allylation of aldehydes with allylic alcohols reaches similar levels of enantioselectivities as for the allylic amines [27].

3. Asymmetric α-Fluorination

The catalytic asymmetric α-fluorination of acid chlorides has been shown by Lectka and co-workers to be a powerful method to synthesize a wide range of α-fluorocarboxylic acid derivatives in good yields and excellent enantioselectivities through electrophilic fluorination of a ketene enolate intermediate [28,29,30]. Lectka’s bifunctional catalytic system is based on the combination of the chiral nucleophiles benzoylquinine (6) or benzoylquinidine (7) and a transition metal Lewis acid cocatalyst, such as (PPh3)2PdCl2, in the presence of Hünig’s base. Under these conditions, the chiral ketene enolate intermediate reacts with N-fluorodibenzenesulfonamide (NFSi) to provide the corresponding α-fluoro bis(sulfonimide) intermediate, which can be quenched with a variety of nucleophiles to furnish α-fluorocarboxylic acids, amides, esters and even peptides (Scheme 6) [28]. The methodology is particularly attractive since both enantiomers of the product are available in similar selectivity by the selection of either benzoylquinine (6) or its pseudoenantiomer benzoylquinidine (7) (Figure 2).
Figure 2. Benzoylquinine (BQ) (6) and benzoylquinidine (BQd) (7).
Figure 2. Benzoylquinine (BQ) (6) and benzoylquinidine (BQd) (7).
Molecules 18 10108 g002
Scheme 6. Catalytic enantioselective α-fluorination of acyl chlorides.
Scheme 6. Catalytic enantioselective α-fluorination of acyl chlorides.
Molecules 18 10108 g012
In addition, this dual activation strategy proved to be very efficient, in terms of yield and enantioselectivity, for 2-substituted aromatic and heteroaromatic acid chlorides with a variety of nucleophiles (Figure 3). However, the asymmetric α-fluorination of aliphatic acyl chlorides needs of a third catalyst, the alkali metal Lewis acid LiClO4 (10 mol%), to work cooperatively with the chiral nucleophile and the Pd catalyst and afford optically pure products in good yields (for a selected example, Scheme 7) [30]. Mechanistic studies have shown that the lithium salt activates the fluorinated agent; this input being necessary in the case of the less reactive aliphatic acyl chlorides (Figure 4).
Figure 3. Representative examples of the asymmetric α-fluorination through a dual activation mechanism.
Figure 3. Representative examples of the asymmetric α-fluorination through a dual activation mechanism.
Molecules 18 10108 g003
Scheme 7. Catalytic enantioselective α-fluorination of an aliphatic acyl chloride.
Scheme 7. Catalytic enantioselective α-fluorination of an aliphatic acyl chloride.
Molecules 18 10108 g013
Figure 4. Proposed activation mode of the trifunctional catalytic system.
Figure 4. Proposed activation mode of the trifunctional catalytic system.
Molecules 18 10108 g004
The α-fluorination of acid chlorides developed by Lectka et al. has been successfully employed for the site-specific functionalization of natural products and biologically active molecules with excellent diastereoselectivity [28,29]. For instance, quenching the fluorination reaction of 3-phthalimidopropionyl chloride with the lactol of the anti-malarian agent artemisinin, affords compound 8 in 75% yield and 81% ee (Figure 5) [29]. Also, the chemotherapeutic drug taxol, which has three distinct and potentially nucleophilic hydroxyl groups, reacted selectively by the secondary alcohol with the fluorinated intermediate of p-methoxyphenylacetyl chloride, affording compound 9 in 43% yield and 99% de (Figure 5) [29]. Using this strategy, Leckta has reported the synthesis of biologically relevant α-fluorinated carbonyl derivatives from a selection of chemotherapeutics, antibiotics, and other pharmaceuticals.
Figure 5. Fluorinated artemisinin derivative 8 and fluorinated taxol derivative 9.
Figure 5. Fluorinated artemisinin derivative 8 and fluorinated taxol derivative 9.
Molecules 18 10108 g005

4. Enantioselective Decarboxylative Protonation

Enantioselective protonation is a very common process in biosynthetic transformations. Recently, several chemical methods for achieving enantioselective protonation have been developed, employing a variety of precursors and protocols [31]. Clearly, the generation of tertiary carbon stereocenters by using the very small proton atom is a very interesting and attractive process; however, it is also very challenging, because it normally involves undesired racemization reactions. The enantioselective protonation of enolates, using the combination of palladium and organocatalysis, has mainly focused on the generation of the enolate via decarboxylation of β-ketoesters [32,33]. An enantioselective decarboxylative protonation can be achieved either generating a chiral Brønsted base (i.e., a chiral metal enolate), or by protonation of an achiral enolate by a chiral proton source. The formation of the enolate in the case of Pd-catalyzed reactions is typically achieved by decarboxylation of benzyl or allyl β-ketoesters.
Hénin, Muzart, and co-workers have studied the enantioselective decarboxylative protonation of benzyl β-ketoesters to generate an achiral enol that is then protonated by a chiral amino alcohol [34,35,36,37,38]. Usually, moderate enantioselectivities (17%–75%) are obtained for the protonation of both cyclic (which generate stereodefined proton acceptors) and acyclic benzyl β-ketoesters under the optimized conditions (Scheme 8). With respect to the reaction mechanism, Baiker et al. have demonstrated that only the debenzylation reaction to afford the corresponding β-ketoacid takes place on the Pd surface, whilst the decarboxylation is a homogeneous process, organocatalyzed by the chiral amino alcohol [39].
Scheme 8. Enantioselective protonation of a linear achiral enolate by a chiral amino alcohol.
Scheme 8. Enantioselective protonation of a linear achiral enolate by a chiral amino alcohol.
Molecules 18 10108 g014
The enantioselective protonation by chiral Brønsted acids is strongly dependent on the reaction conditions, thus limiting the scope and applicability of the process. Recently, Stoltz et al. have studied the decarboxylative enantioselective protonation of cyclic allyl β-ketoesters [40,41]. In this case, the corresponding chiral enolates are protonated with an achiral proton source, such as Meldrum’s acid [41]. As depicted in Scheme 9, a wide range of tertiary-substituted cyclic ketones can be prepared in very good yields and enantioselectivities using the chiral phosphinooxazoline (PHOX) as ligand for the palladium catalyst [Pd(OAc)2].
Scheme 9. Enantioselective protonation of a chiral enolate by Meldrum’s acid.
Scheme 9. Enantioselective protonation of a chiral enolate by Meldrum’s acid.
Molecules 18 10108 g015

5. Cascade Iminium/Enamine-Palladium Cooperative Catalysis

Lately, remarkable efforts have been made into developing stereocontrolled sequential, cascade or domino reactions, since they allow the syntheses of complex chiral molecules from relatively simple starting materials through multiple consecutive catalytic cycles in an atom-economy fashion [42,43,44]. In this regard, the combination of iminium/enamine and palladium catalysis has been used by different authors to prepare optically active highly functionalized carbo- and heterocycles [45,46,47,48]. After Dixon’s decisive work using copper catalysis [45], Córdova et al. reported a simple and highly enantioselective dynamic kinetic asymmetric transformation (DYKAT) involving catalytic iminium activation, enamine activation, and palladium-catalyzed enyne cycloisomerization [46]. Córdova’s methodology allows the preparation of functionalized optically active cyclopentenes with all-carbon quaternary stereocenters in good diastereo- and high enantioselectivities from simple starting materials such as α,β-unsaturated aldehydes (aromatic and aliphatic) and activated propargylated methylenes (Scheme 10). Both terminal and internal alkynes are tolerated under the optimized reaction conditions.
The reaction mechanism involves an initial Michael addition of the activated propargylated methylene derivative to the chiral iminium salt formed between the Jørgensen-Hayashi catalyst and the enal. This reversible addition generates a diastereomeric mixture of Michael adducts which undergoes, at different rates, a Pd(0)/enamine catalyzed cycloisomerization to afford the corresponding products (Scheme 11).
Scheme 10. Amine- and palladium-catalyzed enantioselective cycloisomerization.
Scheme 10. Amine- and palladium-catalyzed enantioselective cycloisomerization.
Molecules 18 10108 g016
Scheme 11. Proposed mechanism for the amine/palladium-catalyzed enantioselective cycloisomerization.
Scheme 11. Proposed mechanism for the amine/palladium-catalyzed enantioselective cycloisomerization.
Molecules 18 10108 g017
In parallel, chiral dihydrofurans and dihydropyrroles have been prepared using a similar iminium/enamine-palladium cooperative catalytic strategy [47,48]. Thus, Zhao and Córdova have reported a highly enantioselective domino oxa-Michael/carbocyclization between propargyl alcohols and enals to afford the corresponding chiral dihydrofurans in good to high yields and very high enantioselectivities (Figure 6) [47]. Similarly, the reaction between N-tosyl propargylamines and enals provides highly functionalized optically active pyrrolines, as depicted in Figure 6, for selected examples [48]. In these two studies, aromatic enals lead to higher enantioselectivities than aliphatic ones. Also, it is worthy to note the ability of the catalytic system towards the synthesis of 2,3,4,5-tetrasubstituted dihydrofurans and pyrrolines, which can be prepared in moderate yield and diastereoselectivity but with high enantioselectivity, by using the corresponding 2-secondary and tertiary propargylic nucleophile (Figure 6). On the contrary, very low yields have been reported when using internal alkynes as nucleophiles.
Figure 6. Representative examples of chiral dihydrofurans and pyrrolines prepared by amine- and palladium-catalyzed enantioselective cycloisomerization.
Figure 6. Representative examples of chiral dihydrofurans and pyrrolines prepared by amine- and palladium-catalyzed enantioselective cycloisomerization.
Molecules 18 10108 g006
Finally, the enantioselective β-functionalization of aldehydes has been recently achieved by co-catalysis, employing pyrrolidine 3 and palladium acetate. The process involves the formation of the corresponding enamine by reaction of the aldehyde with 3, followed by the palladium-catalyzed oxidation to the α,β-unsaturated iminium intermediate, which undergoes a Michael addition with diethyl malonate (Scheme 12) [49]. The final products are obtained in moderate to good yields (45%–72%) and with high enantioselectivities (87%–99%).
Scheme 12. Asymmetric β-functionalization of aldehydes by combination of organocatalysis and palladium catalytic oxidation.
Scheme 12. Asymmetric β-functionalization of aldehydes by combination of organocatalysis and palladium catalytic oxidation.
Molecules 18 10108 g018

6. Conclusions

The dual activation of simple substrates using the combination of organocatalysis and palladium catalysis has been recently applied to the synthesis of chiral molecules. This distinctive combination brings new opportunities in terms of both selectivity and reactivity, by allowing novel transformations that cannot be performed by using metal catalysis or organocatalysis independently. Pioneering work has been recently reported on the asymmetric α-allylation of carbonyl compounds [17,18,19,20,21,22,23,24,25,26,27], α-fluorination of acyl derivatives [28,29,30], decarboxylative protonation of β-dicarbonyls [32,33,34,35,36,37,38,39,40,41], cyclization reactions of alkynyl carbonyl compounds [46,47,48] and β-functionalization of aldehydes [49], employing this double-catalytic methodology. These reports evidence the versatility of this emerging strategy; however, there is still room for improvement and expansion on the areas of application, as well as on the better understanding of the processes.

Acknowledgments

MAFI thanks financial support from the Ministerio de Economía y Competitividad (MINECO, CTQ 2012-35790) and the Consejería de Educación de la Comunidad de Madrid (programme AVANCAT, S2009/PPQ-1634). BM thanks the European Commission for a Marie Curie Integration Grant (FP7-PEOPLE-2012-CIG). IMP thanks financial support from the Ministerio de Ciencia e Innovación (MICINN) of Spain (CTQ2011-24165) and from the University of Alicante. DAA thanks financial support for the European Union (ORCA action CM0905) and from the University of Alicante.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bertelsen, S.; Jørgensen, K.A. Organocatalysis—After the gold rush. Chem. Soc. Rev. 2009, 38, 2178–2189. [Google Scholar] [CrossRef]
  2. Beller, M.; Bolm, C. Transition Metals for Organic Synthesis: Building Blocks and Fine Chemicals, 2nd ed; Wiley-VCH: New York, NY, USA, 2004; Volume 1–2. [Google Scholar]
  3. Jacobsen, E.N.; Pfaltz, A.; Yamamoto, H. Comprehensive Asymmetric Catalysis I-III; Springer: Berlin, Germany, 2000. [Google Scholar]
  4. Shao, Z.; Zhang, H. Combining transition metal catalysis and organocatalysis: A broad new concept for catalysis. Chem. Soc. Rev. 2009, 38, 2745–2755. [Google Scholar] [CrossRef]
  5. Zhong, C.; Shi, X. When organocatalysis meets transition-metal catalysis. Eur. J. Org. Chem. 2010, 2010, 2999–3025. [Google Scholar] [CrossRef]
  6. Zhou, J. Recent advances in multicatalyst promoted asymmetric tandem reactions. Chem. Asian J. 2010, 5, 422–434. [Google Scholar] [CrossRef]
  7. Du, Z.; Shao, Z. Combining transition metal catalysis and organocatalysis—An update. Chem. Soc. Rev. 2013, 42, 1337–1378. [Google Scholar] [CrossRef]
  8. Diederich, F.; Stang, P.J. Metal-Catalyzed Cross-Coupling Reactions; Wiley-VCH: New York, NY, USA, 1998. [Google Scholar]
  9. Negishi, E. Handbook of Organopalladium Chemistry for Organic Synthesis; Wiley-Interscience: New York, NY, USA, 2002. [Google Scholar]
  10. Miyaura, N. Cross-Coupling Reactions. A Practical Guide; Springer: Berlin, Germany, 2002. [Google Scholar]
  11. Tsuji, J. Palladium Reagents and Catalysts: Innovations in Organic Synthesis; Wiley: New York, NY, USA, 2004. [Google Scholar]
  12. Diederich, F.; de Meijere, A. Metal-Catalyzed Cross-Coupling Reactions, 2nd ed; Wiley-VCH: New York, NY, USA, 2004. [Google Scholar]
  13. Otera, J. Modern Carbonyl Chemistry; Wiley-VCH: New York, NY, USA, 2000. [Google Scholar]
  14. Trost, B.M.; Crawley, M.L. Asymmetric transition-metal-catalyzed allylic alkylations: Applications in total synthesis. Chem. Rev. 2003, 103, 2921–2943. [Google Scholar] [CrossRef]
  15. Diéguez, M.; Pàmies, O. Biaryl phosphites: New efficient adaptative ligands for Pd-catalyzed asymmetric allylic substitution reactions. Acc. Chem. Res. 2010, 43, 312–322. [Google Scholar] [CrossRef]
  16. Guerrero-Rios, I.; Rosas-Hernández, A.; Martin, E. Recent advances in the application of chiral phosphine ligands in Pd-catalysed asymmetric allylic alkylation. Molecules 2011, 16, 970–1010. [Google Scholar] [CrossRef]
  17. Chen, G.; Deng, Y.; Gong, L.; Mi, A.; Cui, X.; Jiang, Y.; Choi, M.C.K.; Chan, A.S.C. Palladium-catalyzed allylic alkylation of tert-butyl(depenylmethylene)-glycinate with simple allyl esters under chiral phase transfer conditions. Tetrahedron Asymmetry 2001, 12, 1567–1571. [Google Scholar] [CrossRef]
  18. Nakoji, M.; Kanayama, T.; Okino, T.; Takemoto, Y. Chiral phosphine-free Pd-mediated asymmetric allylation of prochiral enolate with a chiral phase-transfer catalyst. Org. Lett. 2001, 3, 3329–3331. [Google Scholar]
  19. Nakoji, M.; Kanayama, T.; Okino, T.; Takemoto, Y. Pd-catalyzed asymmetric allylic alkylation of glycine imino ester using a chiral phase-transfer catalyst. J. Org. Chem. 2002, 67, 7418–7423. [Google Scholar] [CrossRef]
  20. Ibrahem, I.; Córdova, A. Direct catalytic intermolecular α-allylic alkylation of aldehydes by combination of transition-metal and organocatalysis. Angew. Chem. Int. Ed. 2006, 45, 1952–1956. [Google Scholar] [CrossRef]
  21. Afewerki, S.; Ibrahem, I.; Rydfjord, J.; Breistein, P.; Córdova, A. Direct regiospecific and highly enantioselective intermolecular α-allylic alkylation of aldehydes by a combination of transition-metal and chiral amine catalysts. Chem. Eur. J. 2012, 18, 2972–2977. [Google Scholar] [CrossRef]
  22. Bihelovic, F.; Matovic, R.; Vulovic, B.; Saicic, R.N. Organocatalyzed cyclizations of π-allylpalladium complexes: A new method for the construction of five- and six-membered rings. Org. Lett. 2007, 9, 5063–5066. [Google Scholar] [CrossRef]
  23. Vulovic, B.; Bihelovic, F.; Matovic, R.; Saicic, R.N. Organocatalyzed tsuji—Trost reaction: A new method for the closure of five- and six-membered rings. Tetrahedron 2009, 65, 10485–10494. [Google Scholar] [CrossRef]
  24. The excess of amine needed to perform this particular cyclization strictly disqualifies this methodology as organocatalytic.
  25. Li, M.; Datta, S.; Barber, D.M.; Dixon, D.J. Dual amine and palladium catalysis in diastereo- and enantioselective allene carbocyclization reactions. Org. Lett. 2012, 14, 6350–6353. [Google Scholar]
  26. Mukherjee, S.; List, B. Chiral counteranions in asymmetric transition-Metal catalysis: Highly enantioselective Pd/Brønsted acid-catalyzed direct α-allylation of aldehydes. J. Am. Chem. Soc. 2007, 129, 11336–11337. [Google Scholar] [CrossRef]
  27. Jiang, G.; List, B. Direct asymmetric α-allylation of aldehydes with simple allylic alcohols enabled by the concerted action of three different catalysts. Angew. Chem. Int. Ed. 2011, 50, 9471–9474. [Google Scholar] [CrossRef]
  28. Paull, D.H.; Scerba, M.T.; Alden-Danforth, E.; Widger, L.R.; Lectka, T. Catalytic, asymmetric α-fluorination of acid chlorides: Dual metal-ketene enolate activation. J. Am. Chem. Soc. 2008, 130, 17260–17261. [Google Scholar]
  29. Erb, J.; Alden-Danforth, E.; Kopf, N.; Scerba, M.T.; Lectka, T. Combining asymmetric catalysis with natural product functionalization through enantioselective α-fluorination. J. Org. Chem. 2010, 75, 969–971. [Google Scholar] [CrossRef]
  30. Erb, J.; Paull, D.H.; Dudding, T.; Belding, L.; Lectka, T. From bifucntional to trifunctional (tricomponent nucleophile-transition metal-lewis acid) catalysis: The catalytic, enantioselective α-fluorination of acid chlorides. J. Am. Chem. Soc. 2011, 133, 7536–7546. [Google Scholar]
  31. Mohr, J.T.; Hong, A.Y.; Stoltz, B.M. Enantioselective protonation. Nat. Chem. 2009, 1, 359–369. [Google Scholar] [CrossRef]
  32. Blanchet, J.; Baudoux, J.; Amere, M.; Lasne, M.C.; Rouden, J. Asymmetric malonic and acetoacetic acid syntheses—A century of enantioselective decarboxylative protonations. Eur. J. Org. Chem. 2008, 2008, 5493–5596. [Google Scholar] [CrossRef]
  33. Claraz, A.; Oudeyer, S.; Levacher, V. Deracemization of α-substituted carbonyl compounds via catalytic enantioselective protonation of their corresponding enolates. Curr. Org. Chem. 2012, 16, 2192–2205. [Google Scholar] [CrossRef]
  34. Aboulhoda, S.J.; Hénin, F.; Muzart, J.; Thorey, C.; Behnen, W.; Martens, J.; Mehler, T. Production of optically active ketones by a palladium-induced cascade reaction from racemic β-ketoesters. Tetrahedron Asymmetry 1994, 5, 1321–1326. [Google Scholar] [CrossRef]
  35. Muzart, J.; Hénin, F.; Aboulhoda, S.J. Asymmetric protonation of enolic species: Dramatic increase in the selectivity with temperature and unexpected eyring diagram. Tetrahedron Asymmetry 1997, 8, 381–389. [Google Scholar] [CrossRef]
  36. Detalle, J.F.; Riahi, A.; Steinmetz, V.; Hénin, F.; Muzart, J. Mechanistic insights into the palladium-induced domino reaction leading to ketones from benzyl β-ketoesters:  First characterization of the enol as an intermediate. J. Org. Chem. 2004, 69, 6528–6532. [Google Scholar] [CrossRef]
  37. Roy, O.; Diekmann, M.; Riahi, A.; Hénin, F.; Muzart, J. Access to optically active linear ketones by one-pot catalytic deprotection, decarboxylation, asymmetric tautomerization from racemic benzyl β-ketoesters. Chem. Commun. 2001, 2001, 533–534. [Google Scholar]
  38. Roy, O.; Riahi, A.; Hénin, F.; Muzart, J. Catalysed asymmetric protonation of simple linear keto-enolic species—A route to chiral α-arylpropionic acids. Eur. J. Org. Chem. 2002, 2002, 3986–3994. [Google Scholar] [CrossRef]
  39. Kukula, P.; Matoušek, V.; Mallat, T.; Baiker, A. Enantioselective decarboxylation of β-keto esters with Pd/amino alcohol systems: Successive metal catalysis and organocatalysis. Chem. Eur. J. 2008, 14, 2699–2708. [Google Scholar] [CrossRef]
  40. Mohr, J.T.; Nishimata, T.; Behenna, D.C.; Stoltz, B.M. Catalytic enantioselective decarboxylative protonation. J. Am. Chem. Soc. 2006, 128, 11348–11349. [Google Scholar] [CrossRef]
  41. Marinescu, S.C.; Nishimata, T.; Mohr, J.T.; Stoltz, B.M. Homogeneous Pd-catalyzed enantioselective decarboxylative protonation. Org. Lett. 2008, 8, 1039–1042. [Google Scholar]
  42. Pellissier, H. Stereocontrolled domino reactions. Chem. Rev. 2013, 113, 442–524. [Google Scholar] [CrossRef]
  43. Patil, N.T.; Shinde, V.S.; Gajula, B. A one-pot catalysis: The strategic classification with some recent examples. Org. Biomol. Chem. 2012, 10, 211–224. [Google Scholar] [CrossRef]
  44. Lu, L.Q.; Chen, J.R.; Xiao, W.J. Development of cascade reactions for the concise construction of diverse heterocyclic architectures. Acc. Chem. Res. 2012, 45, 1278–1293. [Google Scholar] [CrossRef]
  45. Yang, T.; Ferrali, A.; Campbell, L.; Dixon, D.J. Combination iminium, enamine and copper(I) cascade catalysis: A carboannulation for the synthesis of cyclopentenes. Chem. Commun. 2008, 2008, 2923–2925. [Google Scholar]
  46. Zhao, G.L.; Ullah, F.; Deiana, L.; Lin, S.; Zhang, Q.; Sun, J.; Ibrahem, I.; Dziedzic, P.; Córdova, A. Dynamic kinetoc asymmetric transformation (DYKAT) by combined amine- and transition-metal-catalyzed enantioselective cycloisomerization. Chem. Eur. J. 2010, 16, 1585–1591. [Google Scholar] [CrossRef]
  47. Lin, S.; Zhao, G.L.; Deiana, L.; Sun, J.; Zhang, Q.; Leijonmarck, H.; Córdova, A. Dynamic kinetic asymmetric domino oxa-michael/carbocyclization by combination of transition-metal and amine catalysis: Catalytic enantioselective synthesis of dihydrofurans. Chem. Eur. J. 2010, 16, 13930–13934. [Google Scholar]
  48. Sun, W.; Zhu, G.; Hong, L.; Wang, R. The marriage of organocatalysis with metal catalysis: access to multisubstituted chiral 2,5-dihydropyrroles by cascade iminium/enamine-metal cooperative catalysis. Chem. Eur. J. 2011, 17, 13958–13962. [Google Scholar] [CrossRef]
  49. Zhao, Y.L.; Wang, Y.; Hu, X.Q.; Xu, P.F. Merging organocatalysis with transition metal catalysis and using O2 as the oxidant for enantioselective C-H functionalization of aldehydes. Chem. Commun. 2013, 49, 7555–7557. [Google Scholar] [CrossRef]

Share and Cite

MDPI and ACS Style

Fernández-Ibañez, M.Á.; Maciá, B.; Alonso, D.A.; Pastor, I.M. Palladium and Organocatalysis: An Excellent Recipe for Asymmetric Synthesis. Molecules 2013, 18, 10108-10121. https://doi.org/10.3390/molecules180910108

AMA Style

Fernández-Ibañez MÁ, Maciá B, Alonso DA, Pastor IM. Palladium and Organocatalysis: An Excellent Recipe for Asymmetric Synthesis. Molecules. 2013; 18(9):10108-10121. https://doi.org/10.3390/molecules180910108

Chicago/Turabian Style

Fernández-Ibañez, M. Ángeles, Beatriz Maciá, Diego A. Alonso, and Isidro M. Pastor. 2013. "Palladium and Organocatalysis: An Excellent Recipe for Asymmetric Synthesis" Molecules 18, no. 9: 10108-10121. https://doi.org/10.3390/molecules180910108

Article Metrics

Back to TopTop