COMMUNICATION DOI: 10.1002/asia.201400049

Asymmetric Decarboxylative 1,4-Addition of Malonic Acid Half Thioesters to Vinyl Sulfones: Highly Enantioselective Synthesis of 3-Monofluoromethyl3-Arylpropanoic Esters Baokun Qiao,[a] Qian Liu,[b] Hongjun Liu,[b] Lin Yan,[a] and Zhiyong Jiang*[a, b]

Abstract: An asymmetric decarboxylative 1,4-addition of malonic acid half thioesters (MAHTs) to 2-aryl-substituted vinyl sulfones has been developed, yielding adducts with excellent enantioselectivity (up to 97 % ee). In view of tuning pKa values, a quinine-based benzyl-substituted thiourea was designed and demonstrated as the most efficient catalyst. The enantioselective synthesis of 3-monofluorinated analogues of 3-methyl indanone and (+)-turmerone has been accomplished from decarboxylative 1,4-addition adducts with satisfactory results.

fumes, flavors, and cosmetics. Florhydral is a prominent representative in the family of modern synthetic lily-of-thevalley odorants that shaped todays perfumery. In particular, chiral 3-methyl-substituted indanones are key intermediates to readily access various bioactive indane derivatives and are thus of tremendous commercial interest. We recently presented several organocatalytic asymmetric methods to construct monofluorinated bioactive molecules.[7] In this context, we were keen to develop new approaches for the synthesis of monofluorinated 3-arylbutanoic esters, namely 3-monofluoromethyl-3-arylpropanoic esters, to provide an efficient protocol for generating the monofluorinated analogues of bisbolane sesquiterpenes, (+)-flohydral, 3-methylsubstituted indanones, and other derivatives (Figure 1). A number of excellent methods have been devised for the synthesis of chiral 3-arylbutanoic esters, including transitionmetal-catalyzed enantioselective hydrogenation of a,b-unsa-

Monofluorinated analogues of biologically active compounds have received increasing attention, particularly in pharmaceutical research as bioisosteres of parent molecules, as they frequently have enhanced desirable properties without requiring significant changes in the molecular structures.[1] The development of an efficient method for stereoselective installation of monofluoromethyl groups has thus attracted considerable interest in organic synthesis and medicinal chemistry in recent years.[2] Chiral 3-arylbutanoic esters are significant building blocks for the synthesis of a number of natural products, fragrances, and medicinally important agents, such as bisabolane sesquiterpenes,[3] for example, (+)-curcumene, (+)-dehydrocurcumene, (+)-turmerone, ()-curcuhydroquinone, (+)-heliannuol D and (+)-erogorgiaene, (+)-florhydral,[4] 3-methylsubstituted indanones,[5] and so on.[6] Bisabolane sesquiterpenes exhibit anticancer, antimicrobial, and phytotoxic allelopathic activities and can be utilized as additives in per-

[a] B. Qiao,+ Dr. L. Yan, Prof. Z. Jiang Institute of Chemical Biology Henan University Kaifeng, Henan, 475004 (P. R. China) Fax: (+ 86) 378-2864-665 E-mail: [email protected] [b] Q. Liu,+ Dr. H. Liu, Prof. Z. Jiang Key Laboratory of Natural Medicine and Immuno-Engineering of Henan Province Henan University Kaifeng, Henan, 475004 (P. R. China) [+] These authors contributed equally to this work.

Figure 1. Monofluorinated analogues of biologically active compounds derived from 3-monofluoromethyl-3-arylpropanoic esters (top) and the retrosynthetic analysis (bottom).

Supporting information for this article is available on the WWW under http://dx.doi.org/10.1002/asia.201400049.

Chem. Asian J. 2014, 9, 1252 – 1256

1252

 2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Zhiyong Jiang et al.

www.chemasianj.org

turated esters,[8] as well as transition-metal-catalyzed asymmetric 1,4-addition of aryltrialkoxysilanes,[9a–b] arylboronic acids,[9c] potassium trifluoroACHTUNGRE(organo)borates,[9d] Grignard reagents,[9e] or trimethylaluminium[9f] to a,b-unsaturated esters. However, these methods are not suitable to prepare monofluorinated 3-arylbutanoic ester derivatives. To date, the well-established protocols to access 3-monofluoromethyl-3arylpropanoic esters are still limited to the asymmetric addition of fluorobis(phenylsufonyl)methane (FBSM) to a,b-unsaturated aldehydes.[10] Undoubtedly, to attain these attractive compounds, the development of novel and efficient asymmetric protocols is highly desirable. From the retrosynthetic analysis (Figure 1) we conceived that the Michael addition of acetates to aryl-substituted vinyl sulfones is a reasonable choice, but might be subjected to the poor activity of acetates. In the past decade, the biomimetic decarboxylative reactions of malonic acid half thioesters (MAHTs) have been recognized as the most efficient protocol to introduce acetates onto the chiral centers of the molecules.[11] Therefore, we proposed an unprecedented decarboxylative 1,4-addition of MAHTs to 2-aryl-substituted vinyl sulfones, which represents a formidable task due to the rare example on hydrogen-bond catalysis with 2-aryl-substituted vinyl sulfones as electrophiles.[12] To achieve stereoselective decarboxylative 1,4-addition of MAHTs to 2-aryl-substituted vinyl sulfones, we envisioned that a bifunctional tertiary amine catalyst with a properly installed Brønsted acid moiety could be employed. A number of thiourea bifunctional catalysts[13] derived from commercially available Cinchona alkaloids were thus examined in the decarboxylative 1,4-addition of MAHT 1 a to 2-phenylsubstituted vinyl sulfone 2 a in THF at 25 8C (Figure 2, Table 1). First, in the presence of 10 mol % quinine-derived aryl-substituted thiourea A, the reaction proceeded smoothly to afford the desired adduct 3 a in 86 % yield and 59 % ee in 12 h (Table 1, entry 1). Catalysts B–D, derived from quinidine, cinchonine, and cinchonidine, respectively, were found to be less efficient (Table 1, entries 2–4). Further optimization of the reaction conditions was carried out by screening

Table 1. Screening studies of organocatalytic decarboxylative 1,4-addition of MAHT 1 a to 2-phenyl-substituted vinyl sulfone 2 a.[a]

Catalyst

Solvent

T [8C]

t [h]

Yield [%][b]

ee [%][c]

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15[d] 16[f]

A B C D A A A A A A A A E F F G

THF THF THF THF Toluene CH2Cl2 MTBE 1,4-Dioxane EVE CPME CPME CPME CPME CPME CPME CPME

25 25 25 25 25 25 25 25 25 25 10 0 0 0 0 0

12 12 12 12 24 24 48 48 35 12 50 72 72 120 96 96

86 87 88 86 N.R. N.R. 38 74 23 84 88 85 87 81 96 95

59 49 45 55 N.A. N.A. 50 45 64 64 77 81 86 89 92 86

[a] The reaction was carried out with 0.075 mmol of 1 a, 0.05 mmol of 2 a, and 0.005 mmol of catalyst in 0.5 mL solvent. [b] Isolated yield. N.R., no reaction. [c] Determined by HPLC. N.A., not applicable.[d] Carried out on 3.0 mmol scale; 1.49 g of 3 a (96 % yield) was obtained with 92 % ee. [f] The reaction was carried out with 0.15 mmol of 1 a, 0.10 mmol of 2 a, and 0.02 mmol of G in 1.0 mL CPME. MTBE = tert-butyl methyl ether, EVE = ethyl vinyl ether, CPME = cyclopentyl methyl ether.

solvents with A as catalyst at 25 8C (Table 1, entries 5–10). Cyclopentyl methyl ether (CPME) proved to be the best solvent with respect to catalytic reactivity and enantioselectivity (Table 1, entries 9–10). When the temperature was decreased to 0 8C, the enantiomeric excess of 3 a improved to 81 % (Table 1, entry 12). In the asymmetric catalysis of bifunctional tertiary amine/ thioureas, recent physical-organic studies on the structure– activity–enantioselectivity relationships have revealed that the tunable nature of substituents of thioureas could provide a crucial handle for the optimization of catalytic enantioselectivity.[14a,b] We thus sought to examine quinine-derived alkyl-substituted thiourea E as a catalyst in the reaction of 1 a with 2 a in CPME at 0 8C (Table 1, entry 13). The ee value of 3 a was found to be increased to 86 % while the reactivity was similar (Table 1, entry 13). We then continued to test catalyst F bearing a benzyl-substituted thiourea[15] to introduce a steric factor. To our delight, the ee value of 3 a was further increased to 89 % (Table 1, entry 14). Higher selectivity and reactivity were observed when the reaction was conducted in the presence of 20 mol % F (Table 1, entry 15). Moreover, the enantiomer of 3 a could be obtained in 95 % yield with 86 % ee in the presence of 20 mol % quinidine-derived G, which is the pseudoenantiomer of F (Table 1, entry 16). With a set of optimized reaction conditions in hand (20 mol % F as catalyst, CPME as solvent, 0 8C), we found that the decarboxylative 1,4-addition reaction could be extended to a variety of aryl-substituted vinyl sulfones to

Figure 2. Structures of catalysts A–G.

Chem. Asian J. 2014, 9, 1252 – 1256

Entry

1253

 2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Zhiyong Jiang et al.

www.chemasianj.org

isopropyl was used as the ester group, the enantioselectivity was decreased (Table 3, entry 3). Only decarboxylative product 4 was obtained when MAHT 1 d with a b-substituted methyl group was used (Table 3, entry 4). The results show that nucleophilic addition possibly precedes decarboxylation.[11l] A plausible transition-state model is proposed, although the mechanism of the reaction remains to be clarified (Scheme 1). We suggested that the b-hydrogen of MAHT is first abstracted by the Cinchona alkaloid.[ 11l] The catalyst

Table 2. Highly enantioselective decarboxylative 1,4-addition of MAHT 1 a to 2-benzyl-substituted vinyl sulfones 2 b–m.[a]

Entry

2, [Ar]

3

t [h]

Yield [%][b]

ee [%][c]

1 2 3 4 5 6 7 8[d] 9 10 11[e] 12

2 b, [4-FC6H4] 2 c, [4-ClC6H4] 2 d, [4-BrC6H4] 2 e, [3-FC6H4] 2 f, [3-ClC6H4] 2 g, [3,5-F2C6H3] 2 h, [2-ClC6H4] 2 i, [4-MeC6H4] 2 j, [4-MeOC6H4] 2 k, [3-MeOC6H4] 2 l, [2-naphthyl] 2 m, [2-thienyl]

3b 3c 3d 3e 3f 3g 3h 3i 3j 3k 3l 3m

48 72 60 48 72 72 72 96 96 72 192 120

98 92 88 99 89 88 82 90 86 86 72 82

94 92 92 95 94 92 97 94 94 93 95 90

[a] The reaction was carried out with 0.15 mmol of 1 a, 0.10 mmol of 2 a, and 0.02 mmol of F in 1.0 mL CPME. [b] Isolated yield. [c] Determined by HPLC. [d] Carried out on 3.0 mmol scale; 1.43 g of 3 i (90 % yield) was obtained with 94 % ee. [e] Conducted at 5 8C.

Scheme 1. Plausible reaction mechanism.

would bind to the vinyl sulfone substrate with the R3NH + afford the desired adducts 3 b–m with 90–97 % ee and 72– arm and interact with the deprotonated MAHT from its thi99 % yield (Table 2). Our studies showed that the introducourea moiety through multiple hydrogen bondings to attain tion of various substituents onto the phenyl group of vinyl the adducts with S-configuration.[17] An extra non-classical sulfones did not affect the enantioselectivity (Table 2, eninteraction between the deprotonated MAHT and the hytries 1–11). A high ee value was also achieved when the drogen of the methylene linker of the benzyl group should phenyl group of vinyl sulfones was replaced with heteroaroprobably exist and lead to a more stabilized intermediate, matic groups such as thienyl (Table 2, entry 12). We also which endows stronger stereocontrollability to the benzylfound that vinyl sulfones 2 b–h with electron-withdrawing substituted thioureas (F and G) than the phenyl-substituted groups appended on the aromatic rings were more reactive thioureas (A and B).[18] than those with electron-neutral and -donating groups (2 j– To reveal the utility of the established decarboxylative m). The absolute configuration of the decarboxylative 1,41,4-addition protocol, we attempted to prepare the monoaddition products was assigned based on single-crystal X-ray fluorinated analogues of 3-methyl indanone and (+)-turmercrystallographic analysis of product 3 m.[16] one from the decarboxylative 1,4-addition adducts. Upon treatment with Selectfluor, the decarboxylative 1,4-addition Next, MAHTs 1 b–d were evaluated as carboxylate enoadduct 3 a could be conveniently fluorinated to achieve 5, late surrogates under the established protocol (Table 3). The which was then subjected to desulfonation and hydrolyzasurvey revealed that MAHT 1 b with a bulkier ester moiety tion to give b-monofluoromethyl-substituted carboxylic acid than 1 a had a similar reactivity and led to comparable enan6. After intramolecular Friedel–Crafts acylation of 6, 3-montiomeric excess (Table 3, entries 1–2). When less hindered ofluoromethyl-substituted indanone 7 was obtained in 45 % overall yield from 3 a with 91 % ee Table 3. Decarboxylative 1,4-addition of MAHTs 1 b–d to 2-phenyl-substituted vinyl [Scheme 2, Eq. (1)]. Notably, this is the first exam[a] sulfones 2. ple on the organocatalytic asymmetric synthesis of chiral 3-substituted indanones.[5e, 19] As the key intermediate of monofluorinated (+)-turmerone [F(+)-turmerone, 11], 3-monofluoromethyl-3-(p-tolyl)propanal 9 was subsequently prepared. Through Entry 1, [R1, R2] 2, [Ar] Product t [h] Yield [%][b] ee [%][c] fluorination, desulfonation, reduction, and oxida1b tion of adduct 3 i, aldehyde 9 could be obtained 1 2 a, [C6H5] 3n 72 86 91 with excellent yield [Scheme 2, Eq. (2)]. Treatment 1b of 9 with Grignard reagent 10 followed by oxida2 e, [3-FC6H4] 3o 71 81 94 2 tion afforded the desired F-(+)-turmerone (11) in 3p 73 88 82 3 1 c [iPr, H] 2 e, [3-FC6H4] 86 % yield with 93 % enantiomeric excess. 4 1 d [tBu, CH3] 2 a, [C6H5] 4 72 90 N.A.[d] In summary, we report a highly enantioselective [a] The reaction was carried out with 0.15 mmol of 1 a, 0.10 mmol of 2 a, and reaction, in which 2-aryl-substituted vinyl sulfones 0.02 mmol of F in 1.0 mL CPME. [b] Isolated yield. [c] Determined by HPLC. are used as electrophiles via the challenging hydro[d] N.A., not applicable.

Chem. Asian J. 2014, 9, 1252 – 1256

1254

 2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Zhiyong Jiang et al.

www.chemasianj.org

1.0 equiv) was added. The reaction mixture was stirred and maintained at 0 8C, and the reaction progress was monitored by thin-layer chromatography (TLC). Upon complete consumption of 2 a, the reaction mixture was purified by silica gel chromatography (petroleum ether/EtOAc, 10:1! 4:1) to afford product 3 a as a white solid (49.5 mg, 96 % yield, 92 % ee).

Acknowledgements We are grateful for financial support from the NSFC (nos. 21072044, U1204207, 21202034), Excellent Youth Foundation of Henan Scientific Committee (114100510003), and the Program for New Century Excellent Talents in University of Ministry of Education (NCET-11-0938).

Keywords: asymmetric synthesis · decarboxylative 1,4addition · monofluoromethylation · organocatalysis · vinyl sulfones

[1] a) Organofluorine Compounds in Medicinal Chemistry and Biomedical Applications (Eds.: R. Filler, Y. Kobayashi, L. M. Yagupolskii), Elsevier, 1993; b) T. Hiyama in Organofluorine Compounds: Chemistry and Applications, Springer, Berlin, 2000; c) K. Uneyama in Organofluorine Chemistry, Blackwell, Oxford, 2006; d) D. O’Hagan, Chem. Soc. Rev. 2008, 37, 308; e) Y. Wang, Y.-C. Luo, X.Q. Hu, P.-F. Xu, Org. Lett. 2011, 13, 5346; f) T. Furuya, A. S. Kamlet, T. Ritter, Nature 2011, 473, 470. [2] For selected reviews, see: a) J.-A. Ma, D. Cahard, Chem. Rev. 2004, 104, 6119; b) H. Ibrahim, A. Togni, Chem. Commun. 2004, 1147; c) P. M. Pihko, Angew. Chem. Int. Ed. 2006, 45, 544; Angew. Chem. 2006, 118, 558; d) V. A. Brunet, D. OHagan, Angew. Chem. Int. Ed. 2008, 47, 1179; Angew. Chem. 2008, 120, 1198; e) S. Lectard, Y. Hamashima, M. Sodeoka, Adv. Synth. Catal. 2010, 352, 2708; f) D. Cahard, X. Xu, S. Couve-Bonnaire, X. Pannecoucke, Chem. Soc. Rev. 2010, 39, 558; g) C. Hollingworth, V. Gouverneur, Chem. Commun. 2012, 48, 2929. [3] a) V. K. Honwad, A. S. Rao, Tetrahedron 1965, 21, 2593; b) A. I. Meyers, D. Stoianova, J. Org. Chem. 1997, 62, 5219; c) C. Fuganti, S. Serra, A. Dulio, J. Chem. Soc. Perkin Trans. 1 1999, 279; d) C. Fuganti, S. Serra, J. Chem. Soc. Perkin Trans. 1 2000, 3758; e) M. Osaka, M. Kanematsu, M. Yoshida, K. Shishido, Tetrahedron: Asymmetry 2010, 21, 2319. [4] a) A. Abate, E. Brenna, C. D. Negri, C. Fuganti, S. Serra, Tetrahedron: Asymmetry 2002, 13, 899; b) P. Kraft, J. Chem. Educ. 2010, 87, 598. [5] a) J. Almy, D. J. Cram, J. Am. Chem. Soc. 1969, 91, 4459; b) W. M. Welch, A. R. Kraska, R. Sarges, B. K. Koe, J. Med. Chem. 1984, 27, 1508; c) J. A. Lowe, D. L. Hageman, S. E. Drozda, S. McLean, D. K. Bryce, R. T. Crawford, S. Zorn, J. Morrone, J. Bordner, J. Med. Chem. 1994, 37, 3789; d) J. Yun, S. L. Buchwald, J. Org. Chem. 2000, 65, 767; e) F. O. Arp, G. C. Fu, J. Am. Chem. Soc. 2005, 127, 10482. [6] P. Ghorai, A. Kraus, T. Birnkammer, R. Geyer, G. Bernhardt, S. Dove, R. Seifert, S. Elz, A. Buschauer, Bioorg. Med. Chem. Lett. 2010, 20, 3173. [7] a) Y. Pan, Y. Zhao, T. Ma, Y. Yang, H. Liu, Z. Jiang, C.-H. Tan, Chem. Eur. J. 2010, 16, 779; b) Y. Zhao, Y. Pan, H. Liu, Y. Yang, Z. Jiang, C.-H. Tan, Chem. Eur. J. 2011, 17, 3571; c) W. Yang, X. Wei, Y. Pan, R. Lee, B. Zhu, H. Liu, L. Yan, K.-W. Huang, Z. Jiang, C.H. Tan, Chem. Eur. J. 2011, 17, 8066; d) L. Yan, Z. Han, B. Zhu, C. Yang, C.-H. Tan, Z. Jiang, Beilstein J. Org. Chem. 2013, 9, 1853. [8] a) D. Liu, W. Tang, X. Zhang, Org. Lett. 2004, 6, 513; b) D. liu, Q. Dai, X. Zhang, Tetrahedron 2005, 61, 6460; c) C.-J. Hou, W.-L. Guo, X.-P. Hu, J. Deng, Z. Zheng, Tetrahedron: Asymmetry 2011, 22, 195; d) J. Mazuela, P.-O. Norrby, P. G. Andersson, O. Pmies, M. Di-

Scheme 2. Transformations of decarboxylative 1,4-addition adducts. Reagents and conditions: a) Selectfluor (1.3 equiv), NaH (1.3 equiv), THF, 0 8C to rt, 2 h, 89% yield, b) Mg (30 equiv), I2, MeOH/THF (5:1), rt, 10 h; c) KOH (5.0 equiv), H2O/MeOH (5:1), rt, 12 h, 59% yield over two steps; d) (COCl)2 (2.0 equiv), CH2Cl2, 0 8C to rt, 2 h, e) AlCl3 (1.5 equiv), CH2Cl2, 0 8C to rt, 2 h, 85% yield over two steps; f) Selectfluor (1.3 equiv), NaH (1.3 equiv), THF, 0 8C to rt, 2 h, 94% yield; g) Mg (30 equiv), I2, MeOH/THF (5:1), rt, 12 h, 63% yield; h) LiAlH4 (2.5 equiv), THF, 50 8C, 6 h; i) DMP (1.5 equiv), CH2Cl2, 0 8C to rt, 2 h, 85% yield over two steps; j) 10 (4.0 equiv), THF, 0 8C to rt, 8 h; k) DMP (2.0 equiv), CH2Cl2, 0 8C to rt, 2 h, 75% yield over two steps.

gen-bond catalysis. Cinchona alkaloid-derived benzyl-substituted thiourea catalysts furnished an unprecedented asymmetric decarboxylative 1,4-addition of MAHTs to 2-arylsubstituted vinyl sulfones (up to 97 % ee). A series of 3monofluoromethyl-3-arylpropanoic esters could be attained from the decarboxylative 1,4-addition adducts after convenient transformations. From 3-monofluoromethyl-3-arylpropanoic esters, numerous monofluorinated natural and nonnatural products, such as 3-methyl-substituted indanones and turmerone, could be readily obtained, with potentially improved properties. Investigations to fully understand the role of the methylene link in the new catalysts and applications of the catalysts in other asymmetric reactions are ongoing and will be reported in due course.

Experimental Section Representative procedure for the decarboxylative 1,4-addition of MAHT (1 a) to 2-phenyl-substituted vinyl sulfone (2 a) catalyzed by Cinchona alkaloid-derived benzyl-substituted thiourea (F) MAHT 1 a (26.4 mg, 0.15 mmol, 1.5 equiv) and F (12.2 mg, 0.02 mmol, 0.2 equiv) were dissolved in CPME (1.0 mL) at 0 8C for 30 min. Subsequently, 2-phenyl-substituted vinyl sulfone 2 a (38.4 mg, 0.10 mmol,

Chem. Asian J. 2014, 9, 1252 – 1256

1255

 2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemasianj.org

[9]

[10]

[11]

[12]

guez, J. Am. Chem. Soc. 2011, 133, 13634; e) J.-Q. Li, X. Quan, P. G. Andersson, Chem. Eur. J. 2012, 18, 10609. a) S. Oi, A. Taira, Y. Honma, Y. Inoue, Org. Lett. 2003, 5, 97; b) S. Oi, A. Taira, Y. Honma, T. Sato, Y. Inoue, Tetrahedron: Asymmetry 2006, 17, 598; c) R. Shintani, T. Kimura, T. Hayashi, Chem. Commun. 2005, 3213; d) L. Navarre, M. Pucheault, S. Darses, J.-P. Genet, Tetrahedron Lett. 2005, 46, 4247; e) S.-Y. Wang, T.-P. Loh, Chem. Commun. 2010, 46, 8694; f) L. Gremaud, A. Alexakis, Angew. Chem. Int. Ed. 2012, 51, 794; Angew. Chem. 2012, 124, 818. a) A.-N. Alba, X. Company, A. Moyano, R. Rios, Chem. Eur. J. 2009, 15, 7035; b) S. Zhang, Y. Zhang, Y. Ji, H. Li, W. Wang, Chem. Commun. 2009, 4886; c) F. Ullah, G.-L. Zhao, L. Deiana, M. Zhu, P. Dziedzic, I. Ibrahem, P. Hammar, J. Sun, A. Crdova, Chem. Eur. J. 2009, 15, 10013. For selected reviews, see: a) A. Hill, Nat. Prod. Rep. 2006, 23, 256; b) Y. Pan, C.-H. Tan, Synthesis 2011, 2044; For selected asymmetric decarboxylative reactions, see: c) S. Orlandi, M. Benaglia, F. Cozzi, Tetrahedron Lett. 2004, 45, 1747; d) D. Magdziak, G. Lalic, H. M. Lee, K. C. Fortner, A. D. Aloise, M. D. Shair, J. Am. Chem. Soc. 2005, 127, 7284; e) F. Berru, S. Antoniotti, O. P. Thomas, P. Amade, Eur. J. Org. Chem. 2007, 1743; f) A. Ricci, D. Pettersen, L. Bernardi, F. Fini, M. Fochi, R. P. Herrera, V. Sgarzani, Adv. Synth. Catal. 2007, 349, 1037; g) J. Lubkoll, H. Wennemers, Angew. Chem. Int. Ed. 2007, 46, 6841; Angew. Chem. 2007, 119, 6965; h) M. Furutachi, S. Mouri, S. Matsunaga, M. Shibasaki, Chem. Asian J. 2010, 5, 2351; i) H. Y. Bae, S. Some, J. H. Lee, J.-Y. Kim, M. J. Song, S. Lee, Y. J. Zhang, C. E. Song, Adv. Synth. Catal. 2011, 353, 3196; j) N. Hara, S. Nakamura, Y. Funahashi, N. Shibata, Adv. Synth. Catal. 2011, 353, 2976; k) N. Hara, S. Nakamura, M. Sano, R. Tamura, Y. Funahashi, N. Shibata, Chem. Eur. J. 2012, 18, 9276; l) Y. Pan, C. W. Kee, Z. Jiang, T. Ma, Y. Zhao, Y. Yang, H. Xue, C.-H. Tan, Chem. Eur. J. 2011, 17, 8363; m) H. Y. Bae, H. J. Sim, J.-W. Lee, B. List, C. E. Song, Angew. Chem. Int. Ed. 2013, 52, 12143; Angew. Chem. 2013, 125, 12365. For the examples through the formation of a covalent intermediate of the catalyst and the substrate, see: a) Q. Zhu, Y. Lu, Org. Lett. 2008, 10, 4803; b) A. Landa, M. Maestro, C. Masdeu, . Puente, S. Vera, M. Oiarbide, C. Palomo, Chem. Eur. J. 2009, 15, 1562; the first

Chem. Asian J. 2014, 9, 1252 – 1256

[13]

[14]

[15]

[16]

[17] [18]

[19]

Zhiyong Jiang et al.

example of hydrogen-bond catalysis with 2-aryl-substituted vinyl sulfones as electrophiles was reported by Lu group: c) X. Dou, Y. Lu, Org. Biomol. Chem. 2013, 11, 5217. For selected reviews, see: a) P. R. Schreiner, Chem. Soc. Rev. 2003, 32, 289; b) T. Akiyama, J. Itoh, K. Fuchibe, Adv. Synth. Catal. 2006, 348, 999; c) M. S. Taylor, E. N. Jacobsen, Angew. Chem. Int. Ed. 2006, 45, 1520; Angew. Chem. 2006, 118, 1550; d) A. G. Doyle, E. N. Jacobsen, Chem. Rev. 2007, 107, 5713; e) S. J. Connon, Chem. Commun. 2008, 2499; f) X. Yu, W. Wang, Chem. Asian J. 2008, 3, 516. a) X. Li, H. Deng, B. Zhang, J. Li, L. Zhang, S. Luo, J.-P. Cheng, Chem. Eur. J. 2010, 16, 450; b) X. Li, B. Zhang, Z.-G. Xi, S. Luo, J.P. Cheng, Adv. Synth. Catal. 2010, 352, 416; c) P. Li, G. Zhao, S. Zhu, Chin. J. Chem. 2011, 29, 2749. a) J. P. Malerich, K. Hagihara, V. H. Rawal, J. Am. Chem. Soc. 2008, 130, 14416; b) W.-H. Wang, T. Abe, X.-B. Wang, K. Kodama, T. Hirose, G.-Y. Zhang, Tetrahedron: Asymmetry 2010, 21, 2925; c) H.P. Deng, Y. Wei, M. Shi, Eur. J. Org. Chem. 2011, 1956; d) H. J. Lee, S. H. Kang, D. Y. Kim, Synlett 2011, 1559. CCDC 980639 (3 m) contains the supplementary crystallographic data for this paper. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif. Q. Zhu, Y. Lu, Angew. Chem. Int. Ed. 2010, 49, 7753; Angew. Chem. 2010, 122, 7919. a) W. Zhang, D. Tan, R. Lee, G. Tong, W. Chen, B. Qi, K.-W. Huang, C.-H. Tan, Z. Jiang, Angew. Chem. 2012, 124, 10216; Angew. Chem. Int. Ed. 2012, 51, 10069; b) B. Zhu, W. Zhang, R. Lee, Z. Han, W. Yang, D. Tan, K.-W. Huang, Z. Jiang, Angew. Chem. Int. Ed. 2013, 52, 6666; Angew. Chem. 2013, 125, 6798. a) K. Kundu, J. V. McCullagh, A. T. Morehead, Jr., J. Am. Chem. Soc. 2005, 127, 16042; b) Y. Oonishi, J. Ogura, Y. Sato, Tetrahedron Lett. 2007, 48, 7505; c) M. M. Knudsen, N. Kalashnyk, F. Masini, J. R. Cramer, E. Lægsgaard, F. Besenbacher, T. R. Linderoth, K. V. Gothelf, J. Am. Chem. Soc. 2011, 133, 4896. Received: January 9, 2014 Published online: March 3, 2014

1256

 2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Asymmetric decarboxylative 1,4-addition of malonic acid half thioesters to vinyl sulfones: highly enantioselective synthesis of 3-monofluoromethyl-3-arylpropanoic esters.

An asymmetric decarboxylative 1,4-addition of malonic acid half thioesters (MAHTs) to 2-aryl-substituted vinyl sulfones has been developed, yielding a...
270KB Sizes 0 Downloads 0 Views

Recommend Documents