Abstract
Highly congested vicinal stereocenters comprised of tertiary and all-carbon quaternary centers were generated via Ir-catalyzed asymmetric allylic alkylation of β-ketoesters. These catalytic reactions proceed in excellent yields with a broad scope on either reaction partner, and with outstanding regio-, diastereo-, and enantiocontrol. Implementation of a subsequent Pd-catalyzed alkylation affords dialkylated products with pinpoint stereochemical control of both chiral centers.
The asymmetric construction of sterically encumbered vicinal stereogenic centers is of great interest to synthetic chemists due to the prevalence of such structural arrangements in natural products and bioactive compounds.1 The limited number of methods that provide selective access to vicinal tertiary and all-carbon quaternary stereocenters highlights the challenging nature of this task. Enantioselective approaches for accessing this structural dyad have generally relied on asymmetric Michael additions2 and Claisen rearrangements.3 Among the methods available for forging this motif, only a relatively small number have been reported to do so by employing transition metals in a catalytic, asymmetric fashion.4,5 Thus, further investigations into the development of metal-catalyzed methods to directly and selectively generate such stereochemical arrays should prove valuable.
Allylic alkylation chemistry represents a successful strategy for the assembly of highly congested chemical architectures6 and, within this domain, Ir-catalyzed processes are among the most selective and highest yielding.7,8 Initial reports from the Helmchen9 and Hartwig10 groups demonstrated the utility of Ir-catalyzed allylic substitutions for the synthesis of enantioenriched 3,3-disubstituted (branched) allyl compounds.8 As this research area has developed, Ir/phosphoramidite catalysts (i.e. [Ir(cod)Cl]2/L1, Figure 1)11 have emerged as privileged scaffolds for the regio- and enantioselective allylic alkylation of achiral nucleophiles, such as malonate derivatives and ketone enolates.12 However, methods for Ir-catalyzed intermolecular allylic alkylation that employ prochiral nucleophiles and display high (1) regio-, (2) diastereo-, and (3) enantioselectively remain elusive (Scheme 1).13 To date, only two reports,14 from the laboratories of Takemoto and Hartwig, detail success in attaining all three of these goals; however, in these accounts, the nucleophiles investigated were limited to amino acid derivatives and azlactones.15 Herein, we report the development of a highly regio-, diastereo-, and enantioselective Ir-catalyzed α-allylic alkylation of cyclic β-ketoesters that forges vicinal tertiary and all-carbon quaternary centers in one step and in excellent yields. Moreover, we describe the deployment of a novel 2-(trimethylsilyl)-ethyl β-ketoester, which serves as an oxycarbonyl-protected enolate, enabling sequential catalyst-controlled α-allylic alkylations and, in turn, the ability to select the diastereomer produced within the nascent stereochemical dyad.
Figure 1.
Selected Phosphoramidite and PHOX Ligands.
Scheme 1.
Ir-Catalyzed Allylic Substitutions.
Our preliminary studies focused on probing the effects of different ligands, bases, additives, and solvents on the efficiency and selectivity of the reaction. Cyclic β-ketoester 1a, cinnamyl carbonate 2a, and [Ir(cod)Cl]2/phosphoramidite complexes16 were chosen as standard reaction components at the outset of our investigations.17 Selected results of these experiments are summarized in Table 1.18 Our investigations commenced with commonly used phosphoramidite ligand L119 and we were pleased to find that the proposed reaction proceeded smoothly under the conditions described (Table 1), delivering α-quaternary β-ketoester 3aa in >95% conversion and in 96% ee. Unfortunately, no diastereoselectivity was observed in this case (Table 1, entry 1). Use of ligand L2, a diastereoisomer of L1, again produced a high yielding reaction, but in significantly diminished ee (32%) and modest 1:2 dr (entry 2). Inspired by the You group’s use of Ir-N-arylphosphoramidite complexes (derived from [Ir(cod)Cl]2 and L3)20 to effect the diastereo- and enantioselective intramolecular allylation of indoles and pyrroles,21 we envisioned that analogous Ir complexes may prove valuable for the generation of all-carbon quaternary stereocenters. We were delighted to discover that the use of N-aryl-phosphoramidite ligand L3 furnished the desired product in 98% ee, >20:1 dr, and 95:5 branched to linear ratio (entry 3).
Table 1.
Optimization of Reaction Parameters.a
![]() | |||||
---|---|---|---|---|---|
entry | L | base or additive (equiv) | 3aa:4aab | dr of 3aab | ee of 3aa (%)c |
1 | L1 | NaH (2) | >95:5 | 1:1 | 96 (99)d |
2 | L2 | NaH (2) | >95:5 | 1:2 | 32 (3)e |
3 | L3 | NaH (2) | 95:5 | >20:1 | 98 |
4 | L3 | – | 80:20 | 11:1 | 96 |
5 | L3 | Et3N (2) | 77:23 | 11:1 | 97 |
6 | L3 | Cs2CO3 (2) | 63:37 | 6:1 | 93 |
7 | L3 | K3PO4 (2) | 63:37 | 4:1 | 90 |
8 | L3 | LiOt-Bu (2) | 95:5 | >20:1 | 99 |
9 | L3 | LiCl (1) | 88:12 | 14:1 | 98 |
10 | L3 | LiBr (1) | 95:5 | >20:1 | >99 |
11 | L4 | LiBr (1) | 80:20 | 12:1 | 96 |
12f | L5 | LiBr (1) | 12:88 | – | – |
13g | L3 | LiBr (1) | 95:5 | >20:1 | 99 |
Reactions performed with 0.1 mmol of 2a, 0.2 mmol of 1a at 0.1 M in THF at 20 °C and allowed to proceed to complete consumption of 2a.
Determined by 1H NMR and UHPLC-MS analysis of the crude mixture.
Determined by chiral HPLC analysis of the major diastereomer.
(ee) of the alternate diastereomer.
(ee) of the major diastereomer.
Measured after 60 h at 60% conversion.
1 mol % [Ir(cod)Cl]2 and 2 mol % L3 were used.
Extensive exploration of various bases, including organic and inorganic bases, revealed that the use of LiOt-Bu afforded the desired product in comparable selectivities as NaH (entries 4–8). Previous reports demonstrating the marked effect of LiCl over the regioselectivity12b,12j,13b in Ir-catalyzed allylic alkylations prompted us to investigate this and related additives. As a result of these efforts (entries 9–10), the combination of LiBr and THF, at 25 °C, was found to provide 3aa in >20:1 dr, 95:5 branched:linear ratio, and with >99% ee (Table 1, entry 10).
Under these superior conditions, several more ligands were examined. Use of ligand L420b afforded 3aa in 96% ee and 12:1 dr (entry 11). Employment of N-aryl-phosphoramidite scaffolds proved critical to maintaining high diastereoselectivity in the reaction. Phosphinooxazoline (PHOX) type ligands (e.g. L5), first used in Ir-catalyzed allylation by Helmchen,9 were also examined but we found these to be poorly suited for our reaction (entry 12). Finally, we found that the catalyst loading could be reduced to 1 mol % (entry 13) without loss of selectivity.
With optimized conditions in hand, the scope of substrates tolerated in the reaction was explored. We found that cinnamyl derived carbonates bearing either electron-donating (–OMe) or electron-withdrawing (–Br, –CF3) groups on the aromatic ring gave remarkably high dr, ee and yields (99% ee and 20:1 dr, Table 2, entries 1–4). The branched to linear ratio (i.e., 3:4) tended to decrease as the electron deficiency of the aryl substituents increased (from 95:5 to 71:29, with 4-OMe-C6H4 to 4-CF3-C6H4, respectively, entries 2–4). Heteroaryl substituents, such as 3-pyridyl, 2-thienyl and 2-furanyl, were also installed with uniformly excellent enantioselectivities and high diastereoselectivities (95–98% ee and 10:1–17:1 dr, entries 5–7).
Table 2.
Substrate Scope of Ir-Catalyzed Allylic Alkylation of β-Ketoesters.a
![]() | ||||||||
---|---|---|---|---|---|---|---|---|
entry | product (3) | t (h) | 3:4b | dr of 3b | yield (%)c | ee of 3 (%)d | ||
1 |
![]() 3aa–3ah |
3aa: | R2 = Ph | 1 | 95:5 | >20:1 | 98 | >99 |
2 | 3ab: | R2 = p-MeO-C6H4 | 1 | 95:5 | 20:1 | 99 | >99 | |
3 | 3ac: | R2 = p-Br-C6H4 | 8 | 90:10 | >20:1 | 90 | 99 | |
4 | 3ad: | R2 = p-CF3-C6H4 | 12 | 71:29 | >20:1 | 99 | >99 | |
5 | 3ae: | R2 = 3-pyridyl | 12 | 50:50 | 11:1 | 98 | 98 | |
6 |
![]() 3ba |
3af: | R2 = 2-thienyl | 12 | 92:8 | 17:1 | 97 | 95 |
7 | 3ag: | R2 = 2-furanyl | 12 | 95:5 | 10:1 | 90 | 95 | |
8 | 3ah: | R2 = (E)-MeCH=CH | 10 | 95:5 | 8:1 | 92 | 90 | |
9 | 3ba: | R1 = Et | 2 | 95:5 | >20:1 | 98 | >99 | |
| ||||||||
10 |
![]() |
3ca: | n = 1, R = Et | 3 | 93:7 | 8:1 | 85 | 99 |
11 | 3da: | n = 2, R = Et | 12 | 90:10 | 20:1 | 88 | 98 | |
12 | 3ea: | n = 2 R = CH2CH2TMS |
12 | 90:10 | 20:1 | 87 | >99 | |
13 |
![]() |
3fa: | R = Oi-Bu | 10 | 91:9 | 20:1 | 99 | >99 |
14 |
![]() |
3ga: | X = O | 10 | 90:10 | 16:1 | 88 | 98 |
15 | 3ha: | X = NBn | 12 | 85:15 | 13:1 | 85 | 97 |
Reactions performed under the conditions of Table 1, entry 10.
Determined by 1H NMR analysis of the crude mixture.
Isolated yield of 3 and 4.
Determined by chiral HPLC or SFC analysis of the major diastereomer.
In addition to aromatic substituents, methyl sorbyl carbonate was also well tolerated in the chemistry providing diene 3ah, although with a slight decrease in dr and ee (8:1 dr and 90% ee, entry 8).22 Moreover, the reaction proceeded smoothly with ethyl β-ketoester 1b, providing the α-allyl β-ketoester 3ba with excellent yield and selectivity (entry 9). Gratifyingly, aliphatic cyclic ketones also proved to be viable participants in the reaction. Cyclopentanone and cyclohexanone based substrates delivered the products 3ca and 3da in 98–99% ee and 8:1–20:1 dr, respectively (entries 10–11). Vinylogous ester, tetrahydropyran-4-one, and 4-piperidinone derivatives furnished the corresponding products (3fa–3ha) in high yields (85–99%), good diastereoselectivites (13:1–20:1), and enantioselectivities (97–99%, entries 13–15, Table 2). The absolute stereochemistry of the product 3af (>99% ee) was determined as (R,R) by single-crystal X-ray analysis.18
During the course of our investigations we became intrigued by the possibility of developing a sequential allylic alkylation reaction, in which allylation of a dicarbonyl-stabilized enolate would be followed by decarboxylative allylic alkylation and, thus, engender the ability to select among all four possible stereochemical outcomes.23,24 In order to realize such a consecutive allylic alkylation, we devised a novel oxycarbonyl-protected enolate, which we hypothesized would successfully undergo Ir-catalyzed allylic alkylation and be poised to subsequently participate in Pd catalysis. Specifically, we envisioned that treatment of 2-(trimethylsilyl)ethyl β-ketoester 5 with fluoride would trigger the release of fluorotrimethylsilane, ethylene, and CO2 to reveal prochiral enolate 6 (Scheme 2). Enolate 6 could then be intercepted and engage in Pd-catalyzed allylic alkylation to deliver α-quaternary ketone 7. In the case at hand, where β-ketoester 5 contains a chiral branched R group at the α position, we anticipated that with careful choice of catalyst, we could control the newly generated stereocenter independent of the absolute stereochemistry of the side chain.
Scheme 2.
Strategic Enolate Protection.
We were pleased to find that 2-(trimethylsilyl)ethyl β-ketoester 1e is a highly competent substrate for Ir-catalyzed allylic alkylation and, under standard conditions, gave the desired product (3ea) with excellent yield and selectivity (Table 2, entry 12). Moreover, exposure of 3ea to catalytic Pd2(dba)3/L6 (Table 3) in the presence of allyl methylcarbonate and tetrabutylammonium difluorotriphenylsilicate (TBAT) generated the desired diallylated α-quaternary ketone 8a in good yield. The use of achiral PHOX ligand L6 revealed that substrate 3ea displays inherent selectivity under Pd catalysis, furnishing 8a25 as the major diastereomer (Table 3, entry 1). Use of (S)-t-BuPHOX ligand (S)-L5 resulted in modest reversal of the inherent diastereoselectivity to generate 9a predominantly (entry 2). Furthermore, we were interested to find that use of ligand (S)-L7, possessing both an electronically modified phosphine and a smaller i-Pr substituent on the oxazoline ring in contrast to the more standard t-Bu, produced 9a with improved diastereoselectivity (8a:9a, 1:8 dr) and 91% yield (entry 3). Alternatively, through judicious choice of ligand (e.g., (R)-L8), the inherent selectivity of the system could be enhanced to afford 8a with up to 18:1 dr (entries 4–5). Cursory investigation revealed that 2-aryl and 2-alkyl substitutions at the allyl- carbonate are well tolerated: allylic alkylation products 8b and 8c were obtained in good yields and with excellent diastereoselectivities.
Table 3.
Development of Pd-Catalyzed Diastereoselective Decarboxylative Allylic Alkylation of TMSE-β-Ketoesters.a
![]() | ||||
---|---|---|---|---|
entry | product (8/9) | ligand | %yieldb | dr (8:9)c |
1 | 8a/9a: R = H | L6 | 91 | 2:1 |
2 | 8a/9a: R = H | (S)-L5 | 79 | 1:2.2 |
3 | 8a/9a: R = H | (S)-L7d | 91 | 1:8 |
4 | 8a/9a: R = H | (R)-L5 | 73 | 12:1 |
5 | 8a/9a: R = H | (R)-L8 | 85 | 18:1 |
6 | 8b/9b: R = Ph | (R)-L8 | 87 | 12:1 |
7 | 8c/9c: R = Me | (R)-L8 | 70 | 13:1 |
Reactions performed with 1.2 equiv of TBAT, and 1.2 equiv of allyl methylcarbonate at 0.03 M.18
Isolated yield of 8 and 9.
Determined by 1H NMR analysis of the crude mixture and confirmed by GC analysis.
10 mol % ligand was used.
In summary, a highly regio-, diastereo-, and enantio-selective method for the synthesis of vicinal tertiary and all-carbon quaternary centers was realized through the use of an [Ir(cod)Cl]2/N-aryl-phosphoramidite (L3) catalyst system. Varied substitutions were well tolerated on both the β-ketoester and allyl carbonate fragments. A sequential Ir/Pd-catalyzed dialkylation protocol was also established to deliver bis-allylated α-quaternary ketones with excellent stereoselectivity, while affording access to either product diastereomer with catalyst control. Further studies exploring the mechanisms of these reactions and exploiting their applications in the total synthesis of complex natural products are underway in our laboratory.
Supplementary Material
Acknowledgments
The authors wish to thank NIH-NIGMS (R01GM080269), Amgen, the Gordon and Betty Moore Foundation, and Caltech for financial support. Shanghai Institute of Organic Chemistry (SIOC) is thanked for a postdoctoral fellowship to W.-B. L. Dr. Mona Shahgholi and Naseem Torian are acknowledged for mass spectrometry assistance. Lawrence Henling is acknowledged for X-ray crystallographic structural determination. The authors thank Mr. Jeff Holder for the preparation of [Ir(cod)Cl]2. The authors are thankful to Professor Shu-Li You and Mr. Robert Craig for L3 and (R)-L8, respectively, and for helpful discussions.
Footnotes
Experimental procedures, characterization data, single crystal X-ray analysis. This material is available free of charge via the Internet at http://pubs.acs.org.
References
- 1.For selected reviews, see: Douglas CJ, Overman LE. Proc Natl Acad Sci USA. 2004;101:5363. doi: 10.1073/pnas.0307113101.Christoffers J, Baro A. Quaternary Stereocenters: Challenges and Solutions for Organic Synthesis. Wiley-VCH; Weinheim: 2005. Trost BM, Jiang C. Synthesis. 2006:369.
- 2.For reviews, see: Krause N, Hoffmann-Röder A. Synthesis. 2001:171.Christoffers J, Koripelly G, Rosiak A, Rössle M. Synthesis. 2007:1279.Selected examples, see: Barnes DM, Ji J, Fickes MG, Fitzgerald MA, King SA, Morton HE, Plagge FA, Preskill M, Wagaw SH, Wittenberger SJ, Zang J. J Am Chem Soc. 2002;124:13097. doi: 10.1021/ja026788y.Taylor MS, Jacobsen EN. J Am Chem Soc. 2003;125:11204. doi: 10.1021/ja037177o.Mase N, Thayumanavan R, Tanaka F, Barbas CF., III Org Lett. 2004;6:2527. doi: 10.1021/ol049196o.Li H, Wang Y, Tang L, Wu F, Liu X, Guo C, Foxman BM, Deng L. Angew Chem, Int Ed. 2005;44:105. doi: 10.1002/anie.200461923.Bencivenni G, Wu LY, Mazzanti A, Giannichi B, Pesciaioli F, Song MP, Bartoli G, Melchiorre P. Angew Chem, Int Ed. 2009;48:7200. doi: 10.1002/anie.200903192.Ding M, Zhou F, Liu YL, Wang CH, Zhao XL, Zhou J. Chem Sci. 2011;2:2035.Chauhan P, Chimni SS. Adv Synth Catal. 2011;353:3203.Mailhol D, Duque MdMS, Raimondi W, Bonne D, Constantieux T, Coquerel Y, Rodriguez J. Adv Synth Catal. 2012;354:3523.
- 3.For reviews, see: Ito H, Taguchi T. Chem Soc Rev. 1999;28:43.Martín Castro AM. Chem Rev. 2004;104:2939. doi: 10.1021/cr020703u.Ilardi EA, Stivala CE, Zakarian A. Chem Soc Rev. 2009;38:3133. doi: 10.1039/b901177n.For selected communication, see: Yoon TP, MacMillan DWC. J Am Chem Soc. 2001;123:2911. doi: 10.1021/ja015612d.Helmboldt H, Hiersemann M. Tetrahedron. 2003;59:4031.Abraham L, Körner M, Schwab P, Hiersemann M. Adv Synth Catal. 2004;346:1281.Linton EC, Kozlowski MC. J Am Chem Soc. 2008;130:16162. doi: 10.1021/ja807026z.Uyeda C, Jacobsen EN. J Am Chem Soc. 2008;130:9228. doi: 10.1021/ja803370x.Uyeda C, Rötheli AR, Jacobsen EN. Angew Chem, Int Ed. 2010;49:9753. doi: 10.1002/anie.201005183.
- 4.Streuff J, White DE, Virgil SC, Stoltz BM. Nature Chem. 2010;2:192. doi: 10.1038/nchem.518. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.(a) Doyle AG, Jacobsen EN. Angew Chem, Int Ed. 2007;46:3701. doi: 10.1002/anie.200604901. [DOI] [PubMed] [Google Scholar]; (b) Trost BM, Cramer N, Silverman SM. J Am Chem Soc. 2007;129:12396. doi: 10.1021/ja075335w. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Trost BM, Zhang Y. J Am Chem Soc. 2007;129:14548. doi: 10.1021/ja0755717. [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) Trost BM, Miller JR, Hoffman CM., Jr J Am Chem Soc. 2011;133:8165. doi: 10.1021/ja2029602. [DOI] [PubMed] [Google Scholar]; (e) Trost BM, Silverman SM, Stambuli JP. J Am Chem Soc. 2011;133:19483. doi: 10.1021/ja207550t. [DOI] [PMC free article] [PubMed] [Google Scholar]; (f) Tan J, Cheon CH, Yamamoto H. Angew Chem, Int Ed. 2012;51:8264. doi: 10.1002/anie.201203092. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.(a) Trost BM, Crawley ML. Chem Rev. 2003;103:2921. doi: 10.1021/cr020027w. [DOI] [PubMed] [Google Scholar]; (b) Helmchen G, Kazmaier U, Fröster S. In: Catalytic Asymmetric Synthesis. 3. Ojima I, editor. Wiley; New Jersey: 2010. p. 497. [Google Scholar]
- 7.For the first example: Takeuchi R, Kashio M. Angew Chem, Int Ed Engl. 1997;36:263.
- 8.For recent reviews, see: Helmchen G, Dahnz A, Dübon P, Schelwies M, Weihofen R. Chem Commun. 2007:675. doi: 10.1039/b614169b.Helmchen G. In: In Iridium Complexes in Organic Synthesis. Oro LA, Claver C, editors. Wiley-VCH; Weinheim, Germany: 2009. p. 211.Hartwig JF, Pouy MJ. Top Organomet Chem. 2011;34:169.Tosatti P, Nelson A, Marsden SP. Org Biomol Chem. 2012;10:3147. doi: 10.1039/c2ob07086c.Liu WB, Xia JB, You SL. Top Organomet Chem. 2012;38:155.
- 9.Janssen JP, Helmchen G. Tetrahedron Lett. 1997;38:8025. [Google Scholar]
- 10.Ohmura T, Hartwig JF. J Am Chem Soc. 2002;124:15164. doi: 10.1021/ja028614m. [DOI] [PubMed] [Google Scholar]
- 11.For mechanistic studies, see: Kiener CA, Shu C, Incarvito C, Hartwig JF. J Am Chem Soc. 2003;125:14272. doi: 10.1021/ja038319h.Madrahimov ST, Markovic D, Hartwig JF. J Am Chem Soc. 2009;131:7228. doi: 10.1021/ja902609g.Spiess S, Raskatov JA, Gnamm C, Brödner K, Helmchen G. Chem Eur J. 2009;15:11087. doi: 10.1002/chem.200902065.Raskatov JA, Spiess S, Gnamm C, Brödner K, Rominger F, Helmchen G. Chem Eur J. 2010;16:6601. doi: 10.1002/chem.200903465.Madrahimov ST, Hartwig JF. J Am Chem Soc. 2012;134:8136. doi: 10.1021/ja212217j.Raskatov JA, Jäkel M, Straub BF, Rominger F, Helmchen G. Chem Eur J. 2012;18:14314. doi: 10.1002/chem.201201772.For a review: Hartwig JF, Stanley LM. Acc Chem Res. 2010;43:1461. doi: 10.1021/ar100047x.
- 12.Selected examples, see: Lipowsky G, Miller N, Helmchen G. Angew Chem, Int Ed. 2004;43:4595. doi: 10.1002/anie.200460016.Alexakis A, Polet D. Org Lett. 2004;6:3529. doi: 10.1021/ol048607y.Graening T, Hartwig JF. J Am Chem Soc. 2005;127:17192. doi: 10.1021/ja0566275.Weix DJ, Hartwig JF. J Am Chem Soc. 2007;129:7720. doi: 10.1021/ja071455s.He H, Zheng XJ, Li Y, Dai LX, You SL. Org Lett. 2007;9:4339. doi: 10.1021/ol7019394.Spiess S, Welter C, Franck G, Taquet JP, Helmchen G. Angew Chem, Int Ed. 2008;47:7652. doi: 10.1002/anie.200802480.Liu WB, He H, Dai LX, You SL. Org Lett. 2008;10:1815. doi: 10.1021/ol800409d.Chen W, Hartwig JF. J Am Chem Soc. 2012;134:15249. doi: 10.1021/ja306850b.Hamilton JY, Sarlah D, Carreira EM. J Am Chem Soc. 2013;135:994. doi: 10.1021/ja311422z.Bartels B, Helmchen G. Chem Commun. 1999:741. and references therein.
- 13.(a) Dahnz A, Helmchen G. Synlett. 2006:697. [Google Scholar]; (b) Polet D, Alexakis A, Tissot-Croset K, Corminboeuf C, Ditrich K. Chem Eur J. 2006;12:3596. doi: 10.1002/chem.200501180. [DOI] [PubMed] [Google Scholar]; (c) Xu QL, Dai LX, You SL. Adv Synth Catal. 2012;354:2275. [Google Scholar]
- 14.(a) Kanayama T, Yoshida K, Miyabe H, Takemoto Y. Angew Chem, Int Ed. 2003;42:2054. doi: 10.1002/anie.200250654. [DOI] [PubMed] [Google Scholar]; (b) (a) Chen W, Hartwig JF. J Am Chem Soc. 2013;135:2068. doi: 10.1021/ja311363a. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Following submission of this manuscript, a diastereoselective allylic alkylation of aldehydes was reported: Krautwald S, Sarlah D, Schafroth MA, Carreira EM. Science. 2013;340:1065. doi: 10.1126/science.1237068.
- 16.The Ir-phosphoramidite complex was generated in situ by treatment of [Ir(cod)Cl]2 and ligand with TBD, see: Welter C, Koch O, Lipowsky G, Helmchen G. Chem Commun. 2004:896. doi: 10.1039/b400023d.Shu C, Leitner A, Hartwig JF. Angew Chem, Int Ed. 2004;43:4797. doi: 10.1002/anie.200460276. and reference 20b below.
- 17.Methyl 2-oxocyclohexanecarboxylate was previously employed in Ir-catalyzed allylation with 95% ee and 1:1 dr, see ref 13b for detail.
- 18.See the Supporting Information for details.
- 19.For a review of the use of phosphoramidite ligands in asymmetric catalysis, see: Teichert JF, Feringa BL. Angew Chem, Int Ed. 2010;49:2486. doi: 10.1002/anie.200904948.
- 20.(a) Liu WB, He H, Dai LX, You SL. Synthesis. 2009:2076. [Google Scholar]; (b) Liu WB, Zheng C, Zhuo CX, Dai LX, You SL. J Am Chem Soc. 2012;134:4812. doi: 10.1021/ja210923k. [DOI] [PubMed] [Google Scholar]
- 21.Wu QF, He H, Liu WB, You SL. J Am Chem Soc. 2010;132:11418. doi: 10.1021/ja105111n.Zhuo CX, Liu WB, Wu QF, You SL. Chem Sci. 2012;3:205.Wu QF, Zheng C, You SL. Angew Chem, Int Ed. 2012;51:1680. doi: 10.1002/anie.201107677.For an example using Ir-complex/Lewis acid: Schafroth MA, Sarlah D, Krautwald S, Carreira EM. J Am Chem Soc. 2012;134:20276. doi: 10.1021/ja310386m.
- 22.Poor regioselectivity was observed when aliphatic allyl carbonates, such as crotyl carbonate, were employed.
- 23.We and others have extensively investigated the enantioselective decarboxylative allylic alkylation of cyclic systems, see: Mohr JT, Stoltz BM. Chem Asian J. 2007;2:1476. doi: 10.1002/asia.200700183.Behenna DC, Stoltz BM. J Am Chem Soc. 2004;126:15044. doi: 10.1021/ja044812x.Mohr JT, Behenna DC, Harned AM, Stoltz BM. Angew Chem, Int Ed. 2005;44:6924. doi: 10.1002/anie.200502018.Stoltz BM, et al. Chem Eur J. 2011;17:14199.Trost BM, Xu J. J Am Chem Soc. 2005;127:2846. doi: 10.1021/ja043472c.Nakamura M, Hajra A, Endo K, Nakamura E. Angew Chem, Int Ed. 2005;44:7248. doi: 10.1002/anie.200502703.Burger EC, Barron BR, Tunge JA. Synlett. 2006:2824.Trost BM, Bream RN, Xu J. Angew Chem, Int Ed. 2006;45:3109. doi: 10.1002/anie.200504421.
- 24.Attempts to establish a single-vessel diallylation protocol have, thus far, proved unsuccessful. Subjecting allyl 2-oxocyclohexanecarboxylate to the optimized reaction conditions resulted in nonstereoselective decarboxylative allylic alkylation. Sequential addition of all reaction components to substrate 3ea also failed.
- 25.See SI for the determination of the relative configuration.18
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.