• Aucun résultat trouvé

Enantioselective palladium(0)-catalyzed C-H arylation strategy for chiral heterocycles

N/A
N/A
Protected

Academic year: 2021

Partager "Enantioselective palladium(0)-catalyzed C-H arylation strategy for chiral heterocycles"

Copied!
8
0
0

Texte intégral

(1)

Conference paper

Tanguy Saget and Nicolai Cramer*

Enantioselective palladium(0)-catalyzed C–H arylation

strategy for chiral heterocycles

1

Abstract: Transition-metal-catalyzed C–H functionalizations have emerged as complementary and powerful

tools to access molecular complexity from widely available starting materials. Herein, we present a strategy for asymmetric intramolecular Pd(0)-catalyzed C–H functionalizations. The outlined reactivity is based on the cooperative effect between a chiral phosphorous ligand and a carboxylate base acting as a relay of chi-rality during the enantio-discriminating concerted metalation deprotonation step. This approach allows the enantioselective construction of a range of important semi-saturated chiral nitrogen-containing heterocycles such as indolines, tetrahydroquinolines, and dibenzazepinones.

Keywords: asymmetric catalysis; C–H functionalization; cooperative effects; heterocycles; OMCOS-17;

palladium.

1A collection of invited papers based on presentations at the 17th International IUPAC Conference on Organometallic Chemistry Directed Towards Organic Synthesis (OMCOS-17), Fort Collins, Colorado, USA, 28 July–1 August 2013.

*Corresponding author: Nicolai Cramer, Laboratory of Asymmetric Catalysis and Synthesis, Institute of Chemical Sciences and

Engineering, School of Basic Sciences, Ecole Polytechnique Fédérale de Lausanne (EPFL), 1015 Lausanne, Switzerland, e-mail: nicolai.cramer@epfl.ch

Tanguy Saget: Laboratory of Asymmetric Catalysis and Synthesis, Institute of Chemical Sciences and Engineering, School of

Basic Sciences, Ecole Polytechnique Fédérale de Lausanne (EPFL), 1015 Lausanne, Switzerland

Introduction

Heterocycles are privileged structures in organic synthesis. In particular, nitrogen-containing heterocycles are ubiquitous in natural products and extensively used in drug discovery, crop science, organic materials, and dyes [1].For simple and large heterocycle families like indoles, quinolines, isoquinolines, and benzaz-epines, numerous syntheses are reported. Their partially saturated analogs – indolines, tetrahydroquino-lines, tetrahydroisoquinotetrahydroquino-lines, and tetrahydrobenzazepines – are as well important building blocks covering a broad structural space. Moreover, the enhanced number of sp3-hybridized carbon atoms can have additional

benefits (Fig. 1) [2]. However, the semi-saturated congeners which often contain stereogenic centers are more challenging to synthesize selectively, and, therefore, complementary and efficient routes are highly valuable.

Over the past decade, the direct functionalization of unactivated C–H bonds gained significant impor-tance because of its potential to streamline the synthesis of molecules by a rapid increase in molecular com-plexity [3]. The refined understanding of the concerted metalation–deprotonation (CMD) mechanism [4] was critical for the large extension of the scope of Pd(0)-catalyzed C–H functionalizations, leading to direct aryla-tions of C(sp3)–H bonds [3b, 5] as well as to the development of intermolecular processes [6]. A carboxylate

ligand bound to the transition metal is an integral part of the catalyst system. Moreover, this additive allows for milder reaction conditions and enhanced functional group tolerance. In the current mechanistic picture of Pd(0)-catalyzed C–H functionalizations following the CMD mechanism, the reaction is initiated by an oxi-dative addition (Scheme 1). Subsequently, the counterion is exchanged by the carboxylate co-catalyst, setting the stage for the CMD step. A three-coordinated Pd species renders the C–H bond more acidic through agostic interaction, and the carboxylate acts as an internal base deprotonating the C–H bond via a six-membered transition state. In turn, the carboxylic acid dissociates and is deprotonated by an external insoluble base.

(2)

N H N NH R1 R2 R1 R2 R3 R4 R3 R2 R1 N R1 R2 R3 N H NH NH R1 R2 R1 R2 R3 R4 R3 R2 R1 NH R1 R2 R3

Expanded structural space and isomers

Partial saturation * * * * * * * * * * * * Key heterocycles

Fig. 1 Semi-saturation expands structural space of common nitrogen-containing heterocycles.

Y R OH O Y Y Y Pd L X [Pd0-L] Pd L R OM O Base · M MX Base · H+ Y Pd L O O R Concerted metalation deprotonation X H H H Y Pd H O O R L ( )n ( )n n( ) ( )n ( ) n n( ) Oxidative addition Reductive elimination Carboxylate coordination

Scheme 1 Catalytic cycle for Pd(0)-catalyzed C–H functionalization with aryl halides.

Reductive elimination of the cyclometalated Pd(II)-species releases the product and regenerates the Pd(0) catalyst.

Asymmetric catalysis is an efficient way to access chiral molecules and its implementation towards the synthesis of nitrogen-containing heterocycles is a well exploited strategy [7]. However, only limited enanti-oselective C–H functionalizations have been reported so far [8],often because the harsh conditions required for such reaction interfere with selectivity. For Pd(0)-catalyzed C–H functionalizations calling for monoden-tate ligands, the reduced availability of efficient chiral monodenmonoden-tate phosphine families compared to the ubiquitous bidentate ligands represents also a limitation. Moreover, the single point coordination of the ligand on the metal results in free rotation across the metal-ligand bond rendering the imprint of chirality to the substrate more difficult compared to rigid bidentate counterparts. For asymmetric reaction of substrates possessing two enantiotopic C–H bonds, the CMD step is the enantio-discriminating event of the overall reac-tion. During this step, the chiral ligand of the Pd complex is remote from the C–H bond to be activated. This makes a good chiral transfer from the ligand to the substrate challenging. However, the carboxylate base, which is directly involved in the deprotonation event, is both close to the phosphine ligand and the substrate and therefore likely to be well suited to relay the chirality (Scheme 2). A key aspect of our approach was to evaluate and exploit this cooperative effect between the two different ligand types for enantioselective pro-cesses in Pd(0)-catalyzed C–H functionalizations.

(3)

Results and discussion

Palladium(0)-catalyzed enantioselective C–H alkenylation: An initial proof of principle

When we first became interested in enantioselective C–H functionalizations, enantioselective activation of C–H bonds was illusive with Pd(0)-catalyst systems [9].Several reports show that achiral transformations operating by the CMD-mechanism require phosphine or carbene ligands to proceed [4b]. These findings prompted us to believe that the enantiocontrol of such transformation could be feasible by suitable ancil-lary phosphine ligands. We selected a simple model having an alkenyl triflate and two enantiotopic phenyl groups to establish the initial proof of principle of our hypothesis [10]. This substrate class would be able to undergo enantioselective arylation of aromatic C(sp2)–H bonds via a favored six-membered palladacycle.

The use of a triflate in contrast to a halide bears the benefits of a faster exchange step of the counterion for the carboxylate. After an exhaustive screen of a large variety of different phosphorous-based ligands, taddol-based phosphoramidite ligands crystallized as the most promising family. However, the reaction turned out to be very sensitive to the substitution pattern of the taddol ligand and tailored ligand L1 provided the desired high enantioselectivities of  > 90 % ee. The reaction conditions are extraordinarily mild, proceeding at ambient temperature under virtually neutral conditions with sodium bicarbonate as base. Indanes having an all- carbon quaternary center were formed in excellent enantioselectivities, proving the general feasibility of our envisioned enantioselective C–H activation concept (Scheme 3).

C(sp

3

)–H arylations: Synthesis of functionalized indolines

In contrast to aromatic C(sp2)–H bonds, alkyl C(sp3)–H bonds are less acidic and no π electrons are

avail-able for catalyst pre-coordination making C(sp3)–H functionalization a significantly more challenging task.

enantiotopic C-H f unctionalization N R R N Pd Relay of chirality H O O R* L* H H N R R' N O R' R1 Ph

indolines tetrahydroquinolines benzazepinones

one strategy, multiple heterocycles high yields and enantioselectivities [Pd] NR' H H R' R' H H

Scheme 2 Ligand and carboxylate cooperative effects enable high enantioselectivities in Pd(0)-catalyzed C–H

functionalizations. X OTf X R R R R O O O O P NBu2 tBu tBu tBu tBu 5 mol% Pd(OAc)2 12 mol%L1 NaHCO3, DMA, 23 °C L1 N CO2Et S S Cl Cl 93 %, 93 % ee 96 %, 93 % ee 98 %, 93 % ee 98 %, 85 % ee

(4)

Fagnou and co-workers reported the synthesis of dihydrobenzofurans via Pd-catalyzed arylation of C(sp3)–H

bonds showing the importance of catalytic amounts of pivalic acid as a key additive for the CMD step [5a]. Subsequently, Ohno and co-workers demonstrated the synthesis of indolines with the same set of conditions [11]. Given the synthetic importance and value of indolines, we aimed to develop an asymmetric variant of this reaction. During the course of our studies, Kündig et al. [12] and Kagan et al. [13] independently devel-oped enantioselective methods for this transformation. Kündig uses his bulky chiral carbene ligands [14] in combination with the standard pivalic acid additive to achieve high enantioselectivities. To investigate our hypothesis of relaying the chirality to the carboxylic acid, we explored the different combination of chiral carboxylic acids and phosphine ligands and found a strong matched/mismatched effect (Scheme 4).

The need for a bulky electron-rich ligand for this transformation combined with the lack of chiral monodentate phosphines forced us to develop a family of bulky chiral biaryl monophospholanes. Among them, Sagephos L3 turned out to be highly selective for the enantioselective functionalization of methyl as well as methylene C–H bonds starting from aryl triflates. As optimal carboxylic acid partner for L3, our screening revealed the bulky 9H-xanthene-9-carboxylic acid providing robust and high selectivities [15].The reaction has a broad substrate scope and allows for the simultaneous construction of up to three stereocent-ers (Scheme 5). L2 OiPr iPrO P N Tf OTf Na3PO4, cumene, 135 °C 10 mol% [Pd0] 20 mol%L2 10 mol%acid 81 %, 42 % ee N Tf H H 75 %, 64 % ee N Tf H H O O tBu CO2H O O tBu CO2H

Scheme 4 Strong cooperative effects with chiral carboxylic acids.

O CO2H acid (R,R)-Sagephos L3 OiPr iPrO P N Tf OTf Na3PO4, cumene, 135 °C N Tf R R R R 10 mol% [(η3-cinnamyl)PdCp] 20 mol%Sagephos 10 mol%acid 66 %, 95 % ee N Tf H H 75 %, 94 % ee N Tf H H O 64 %, 94 % ee N Tf H H Ph 60 %, 95 % ee N Tf 70 %, 62 % ee N Tf H H 55 %, 88 % ee N NTf 71 %, 95 % ee N Tf MeO 75 %, 96 % ee N Tf F

(5)

Similar aryl bromide substrates react sluggishly with L3. However, ligand L4 based on the C2-symmetric

phospholane architecture mimics the profile of PCy3 and PtBu3 and displays with aryl bromides excellent reactivity and promising selectivity (Scheme 6) [16].

During the development of enantioselective cyclopropane C–H arylations (vide infra), we discovered that substrates possessing cyclopropane substituents with a methine C–H underwent exclusive methine C–H bond activation rather than methylene C–H bond activation leading to larger palladacycle.No destruction of the cyclopropane ring occurs, and the method affords spiroindolines in high yields. Moreover, this reaction can be conducted as a tandem process coupled with intermolecular Suzuki couplings or direct arylation of heteroaromaticsto access more complex spiroindolines in a single operation (Scheme 7) [17].

Enantioselective C–H arylation of cyclopropanes

Next, we aimed to extend the scope of the arylation methodology to larger rings such as tetrahydroquino-lines. This would require less favorable seven-membered palladacycle intermediates, which were elusive for C(sp3)–H arylations. We concentrate our efforts on cyclopropane C–H functionalizations that would be a

complementary route to the many cyclopropane construction methods [18].Such a reaction would not only deliver synthetically valuable products,but as well facilitate the activation process due to the higher s-char-acter of the cyclopropane C–H bonds rendering them more acidic [19] and susceptible towards activation. A re-optimization of the previously successful reaction conditions was necessary, and taddol-based phospho-ramidite L5 gave the best selectivities. Again, the carboxylic acid was crucial as co-catalyst for both reactivity and selectivity, and simple pivalic acid was among the most efficient for this purpose. This new catalytic system is well tolerant towards functional groups and enables the enantioselective synthesis of functional-ized tetrahydroquinolines possessing a cyclopropane ring with a quaternary stereocenter (Scheme 8) [20].

Submitting the tetrahydroquinoline product to mild heterogeneous reduction conditions with H2 and Pd/C, a regioselective hydrogenation of the inner bis-benzylic C–C bond of the cyclopropane moiety took

L4 P N Tf Br Cs2CO3, mesitylene, 130 °C N Tf 5 mol% Pd(dba)2, 12 mol%L4 · HBF4

30 mol%PivOH

96 %, 71 % ee

Scheme 6 Pd(0)-catalyzed enantioselective arylation starting from aryl bromides

R1 R1 N Tf Br Br N Tf Y X H R2 H Y X R2

10-20 mol% Pd(OAc)2/PCy3

50 mol%PivOH Cs2CO3, toluene, 110 °C or (HO)2B R1 N Tf or R3 R3 + N Tf N S N Tf R N Tf Ph N Tf O N Tf S CO2Me R = H, 77 % R = CO2Me, 56 % R = OMe, 61 % 60 % 84 % 58 % 56 % N Tf S CO2Et 56 %

(6)

O O O O P NMe2 L5

2 mol% Pd(dba)2, 3 mol%L5

30 mol%PivOH Cs2CO3, p-xylene, 130 °C N Tf R1 R2 N Tf R1 R2 Br N Tf iPr N Tf Ph N Tf SiMe3 N Tf OPiv N Tf O2N N Tf F F N Tf NTf MeO F3C 90 %, 94 % ee 89 %, 94 % ee 80 %, 92 % ee 90 %, 91 % ee 95 %, 84 % ee 93 %, 91 % ee 90 %, 92 % ee 89 %, 86 % ee N Tf 98 %, 92 % ee

Scheme 8 Intramolecular enantioselective direct arylation of cyclopropanes.

1 atm.H2, 10 % Pd/C, EtOAc, 45 °C

N Tf

Ph

Regioselective and enantiospecific cyclopropane cleavage N Tf Ph H 74 %, 93 % ee 94 % ee benzazepine scaffold Scheme 9 Access to benzazepines via enantiospecific cyclopropane reduction.

place (Scheme 9). Intriguingly, this reduction is completely enantiospecific and afforded the benzazepine scaffold in good yields.

Eight-membered palladacycle intermediates leading to chiral benzazepinones

While with the illustrated regioselective and enantiospecific opening of cyclopropane-substituted tetrahyd-roquinolines chiral benzazepines could be obtained indirectly, a direct access to this scaffold was more desir-able. Intramolecular C–H arylations are by and large limited to the synthesis of four-, five-, and six-membered rings. In contrast, scarce examples of seven-membered ring formations are reported which all require high catalyst loadings (10 mol%) and/or harsh reaction temperatures ( > 130 °C) [21]. This gap can be attributed to the more difficult formation of the eight-membered palladacycle intermediates. To address this issue, we focused on aromatic C(sp2)–H bonds embedded in a substrate with a rigid amide tether [22].We expected

that this would conformationally predispose the aromatic groups towards the aryl Pd(II) species resulting in a facilitated metalation. The anticipated concept worked successfully delivering highly substituted benzaz-epinones at reaction temperatures as low as 80 °C (Scheme 10). The archetypical and simplest congener of the taddol-phosphoramidite family L6 [23] proved to be together with pivalic acid as co-catalyst the most efficient ligand for this transformation. The process is highly regioselective and functional group tolerant, affording complex chiral benzazepinones in high enantioselectivities.

Conclusion

In conclusion, we presented an enantioselective intramolecular C–H arylation strategy to efficiently access functionalized heterocycles. The method is based on chiral phosphine and phosphoramidite ligands working in cooperation with a matching carboxylate, which is required to assist the concerted metalation

(7)

deprotona-tion step. A broad spectrum of C–H bonds can be successfully addressed ranging from aromatic C(sp2)–H

bonds to methyl, methylene, and methine C(sp3)–H bonds. Furthermore, the size of the palladacyclic

inter-mediate can be varied from six-, seven-, and eight-membered palladacycles, allowing access of some of the most important semi-saturated chiral nitrogen-containing heterocycles in an enantioselective fashion.

References

[1] (a) A. F. Pozharskii, A. R. Katritsky, A. T. Soldatenkov (Eds.). Heterocycles in Life and Society: An Introduction to Heterocy-clic Chemistry and Biochemistry and the Role of Heterocycles in Science Technology, Medicine and Agriculture, Wiley-VCH, Weinheim (1997); (b) T. Eicher, S. Hauptmann (Eds.). The Chemistry of Heterocycles: Structure, Reactions, Syntheses, and Applications, 2nd ed., Wiley-VCH, Weinheim (2003); (c) L. D. Quin, J. Tyrell (Eds.). Fundamentals of Heterocyclic Chemistry: Importance in Nature and in the Synthesis of Pharmaceuticals, John Wiley, Hoboken, NJ (2010).

[2] F. Lovering, J. Bikker, C. Humblet. J. Med. Chem. 52, 6752 (2009).

[3] (a) X. Chen, K. M. Engle, D. H. Wang, J. Q. Yu. Angew. Chem. Int. Ed. 48, 5094 (2009); (b) R. Jazzar, J. Hitce, A. Renaudat, J. Sofack-Kreutzer, O. Baudoin. Chem. Eur. J. 16, 2654 (2010); (c) T. W. Lyons, M. S. Sanford. Chem. Rev. 110, 1147 (2010); (d) D. A. Colby, R. G. Bergman, J. A. Ellman. Chem. Rev. 110, 624 (2010); (e) J. Yamaguchi, A. D. Yamaguchi, K. Itami. Angew. Chem. Int. Ed. 51, 8960 (2012); (f) N. Kuhl, M. N. Hopkinson, J. Wencel-Delord, F. Glorius. Angew. Chem. Int. Ed. 51, 10236 (2012); (g) J. J. Mousseau, A. B. Charette. Acc. Chem. Res. 46, 412 (2013).

[4] (a) D. Garcia-Cuadrado, P. de Mendoza, A. A. C. Braga, F. Maseras, A. M. Echavarren. J. Am. Chem. Soc. 129, 6880 (2007); (b) D. Lapointe, K. Fagnou. Chem. Lett. 39, 1118 (2010); (c) L. Ackermann. Chem. Rev. 111, 1315 (2011).

[5] (a) M. Lafrance, S. I. Gorelsky, K. Fagnou. J. Am. Chem. Soc. 129, 14570 (2007); (b) M. Chaumontet, R. Piccardi, N. Audic, J. Hitce, J.-L. Peglion, E. Clot, O. Baudoin. J. Am. Chem. Soc. 130, 15157 (2008); (c) S. Rousseaux, S. I. Gorelsky, B. K. W. Chung, K. Fagnou. J. Am. Chem. Soc. 132, 10706 (2010); (d) S. Rousseaux, B. Liegault, K. Fagnou. Chem. Sci. 3, 244 (2012).

[6] (a) M. Lafrance, K. Fagnou. J. Am. Chem. Soc. 128, 16496 (2006); (b) M. Livendahl, A. M. Echavarren. Isr. J. Chem. 50, 630 (2010).

[7] For selected recent examples, see: (a) A. B. Dounay, L. E. Overman, A. D. Wrobleski. J. Am. Chem. Soc. 127, 10186 (2005); (b) A. T. Parsons, A. G. Smith, A. J. Neel, J. S. Johnson. J. Am. Chem. Soc. 132, 9688 (2010); (c) S. B. Jones, B. Simmons, A. Mastracchio, D. W. C. MacMillan. Nature 475, 183 (2011); (d) R. H. Snell, R. L. Woodward, M. C. Willis. Angew. Chem. Int. Ed.

50, 9116 (2011); (e) D. Katayev, Y.-X. Jia, D. Banerjee, C. Besnard, E. P. Kundig. Chem. Eur. J. 19, 11916 (2013); (f) N. Ortega,

D.-T. D. Tang, S. Urban, D. Zhao, F. Glorius. Angew. Chem. Int. Ed. 52, 9500 (2013); (g) M. Amatore, D. Leboeuf, M. Malacria, V. Gandon, C. Aubert. J. Am. Chem. Soc. 135, 4576 (2013).

N N O R3 Br O 5 mol% Pd(dba)2 10 mol%L6 30 mol%PivOH Cs2CO3, mesitylene, 80 °C R3 R2 R2 R4 R4 R4 R4 R1 R1 84 %, 86 % ee N O Ph 78 %, 93 % ee N O Ph F Cl N O S S 93 %, 94 % ee N O Ph N CO2Me 87 %, 91 % ee N O Ph OMe 97 %, 94 % ee N O Ph 86 %, 87 % ee O O O O Ph Ph Ph Ph P NMe2 L6 R = H: 95 %, 93 % ee R = OMe: 95 %, 93 % ee R = Cl: 98 %, 91 % ee N O R R N O NMe MeN 87 %, 73 % ee

(8)

[8] For a review on enantioselective C–H activation, see: (a) R. Giri, B.-F. Shi, K. M. Engle, N. Maugel, J.-Q. Yu. Chem. Soc. Rev. 38, 3242 (2009); for enantioselective Pd-catalyzed C–H functionalizations, see: (b) B.-F. Shi, N. Maugel, Y.-H. Zhang, J.-Q. Yu. Angew. Chem. Int. Ed. 47, 4882 (2008); (c) B.-F. Shi, Y.-H. Zhang, J. K. Lam, D.-H. Wang, J.-Q. Yu. J. Am. Chem. Soc.

132, 460 (2009); (d) M. Wasa, K. M. Engle, D. W. Lin, E. J. Yoo, J.-Q. Yu. J. Am. Chem. Soc. 133, 19598 (2011); (e) N. Martin,

C. Pierre, M. Davi, R. Jazzar, O. Baudoin. Chem. Eur. J. 18, 4480 (2012); (f) R. Shintani, H. Otomo, K. Ota, T. Hayashi. J. Am. Chem. Soc. 134, 7305 (2012); (g) D.-W. Gao, Y.-C. Shi, Q. Gu, Z.-L. Zhao, S.-L. You. J. Am. Chem. Soc. 135, 86 (2013); (h) C. Pi, Y. Li, X. Cui, H. Zhang, Y. Han, Y. Wu. Chem. Sci. 4, 2675 (2013); (i) X.-F. Cheng, Y. Li, Y.-M. Su, F. Yin, J.-Y. Wang, J. Sheng, H. U. Vora, X.-S. Wang, J.-Q. Yu. J. Am. Chem. Soc. 135, 1236 (2013); for selected examples of enantioselective C–H function-alization from our group, see: (j) D. N. Tran, N. Cramer. Angew. Chem. Int. Ed. 50, 11098 (2011); (k) B. Ye, N. Cramer. Science

338, 504 (2012); (l) P. A. Donets, N. Cramer. J. Am. Chem. Soc. 135, 11772 (2013).

[9] Only one enantioselective Pd(II)-catalyzed C–H functionalization was reported at this time: see ref. [8b]. [10] M. Albicker, N. Cramer. Angew. Chem. Int. Ed. 48, 9139 (2009).

[11] T. Watanabe, S. Oishi, N. Fujii, H. Ohno. Org. Lett. 10, 1759 (2008).

[12] M. Nakanishi, D. Katayev, C. Besnard, E. P. Kundig. Angew. Chem. Int. Ed. 50, 7438 (2011). [13] S. Anas, A. Cordi, H. B. Kagan. Chem. Commun. 47, 11483 (2011).

[14] (a) E. P. Kundig, Y. Jia, D. Katayev, M. Nakanishi. Pure Appl. Chem. 84, 1741 (2012); (b) D. Katayev, M. Nakanishi, T. Burgi, E. P. Kundig. Chem. Sci. 3, 1422 (2012); (c) E. Larionov, M. Nakanishi, D. Katayev, C. Besnard, E. P. Kundig. Chem. Sci. 4, 1995 (2013).

[15] T. Saget, S. J. Lemouzy, N. Cramer. Angew. Chem. Int. Ed. 51, 2238 (2012). [16] P. A. Donets, T. Saget, N. Cramer. Organometallics 31, 8040 (2012). [17] T. Saget, D. Perez, N. Cramer. Org. Lett. 15, 1354 (2013).

[18] H. Lebel, J.-F. Marcoux, C. Molinaro, A. B. Charette. Chem. Rev. 103, 977 (2003). [19] A. de Meijere. Angew. Chem. Int. Ed. 18, 809 (1979).

[20] T. Saget, N. Cramer. Angew. Chem. Int. Ed. 51, 13014 (2012).

[21] (a) M. Burwood, B. Davies, I. Diaz, R. Grigg, P. Molina, V. Sridharan, M. Hughes. Tetrahedron Lett. 36, 9053 (1995); (b) T. Harayama, H. Akamatsu, K. Okamura, T. Miyagoe, T. Akiyama, H. Abe, Y. Takeuchi. J. Chem. Soc. Perkin Trans. 1 523 (2001); (c) L. C. Campeau, M. Parisien, M. Leblanc, K. Fagnou. J. Am. Chem. Soc. 126, 9186 (2004); (d) M. Leblanc, K. Fagnou. Org. Lett. 7, 2849 (2005); (e) L. C. Campeau, M. Parisien, A. Jean, K. Fagnou. J. Am. Chem. Soc. 128, 581 (2006); (f) C. Huang, V. Gevorgyan. Org. Lett. 12, 2442 (2010); (g) D. G. Pintori, M. F. Greaney. J. Am. Chem. Soc. 133, 1209 (2011); (h) C. B. Bheeter, J. K. Bera, H. Doucet. Adv. Synth. Catal. 354, 3533 (2012).

[22] T. Saget, N. Cramer. Angew. Chem. Int. Ed. 52, 7865 (2013). [23] H. W. Lam. Synthesis 43, 2011 (2011).

Figure

Fig. 1 Semi-saturation expands structural space of common nitrogen-containing heterocycles.

Références

Documents relatifs

In summary, we have demonstrated that palladium-catalyzed direct arylation of electron-deficient arenes such as 3,5- difluorobenzene, 2,4-difluorobenzene or 3,5-dichlorobenzene

29 Herein, we describe a readily available copper(I) catalytic system generated from bisoxazoline ligands bearing a chiral dihydroethanoanthracene backbone (L*) (Fig. 1) and CuOTf

Filipa Siopa, Valérie-Anne Ramis Cladera, Carlos Alberto Mateus Afonso, Julie Oble, Giovanni Poli.. To cite

Selected palladium homogeneous catalysts based on chiral diphosphine ligands were the single effective species allowing the hydrogenation of 5-methylenhydantoin in

This result suggested that under these conditions only electron-deficient aryl bromide can be used as coupling partner to allow the C–H bond activation of the fluorobenzene

The same year, Yamaguchi and co-workers achieved the synthesis of cyclic thiazole tetramers with a head-to-tail connection using palladium-catalyzed C–H bond annulation reaction

2 Despite the recent interest of selenophenes as building blocks for the preparation of optoelectronic organic materials,3 only scarce examples of planar π-extended molecules

25 Owing the importance of (hetero)biaryl units in pharmaceuticals, we decided to reinvestigate the Pd-catalyzed C–H bond arylation of antipyrine to find greener conditions