Next Article in Journal
Hybrid Chelator-Based PSMA Radiopharmaceuticals: Translational Approach
Next Article in Special Issue
Azetidinium Lead Halide Ruddlesden–Popper Phases
Previous Article in Journal
Oxy-Polybrominated Diphenyl Ethers from the Indonesian Marine Sponge, Lamellodysidea herbacea: X-ray, SAR, and Computational Studies
Previous Article in Special Issue
Synthesis of Indoloquinolines: An Intramolecular Cyclization Leading to Advanced Perophoramidine-Relevant Intermediates
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Isothiourea-Catalyzed Enantioselective α-Alkylation of Esters via 1,6-Conjugate Addition to para-Quinone Methides

1
EaStCHEM, School of Chemistry, University of St Andrews, North Haugh, St Andrews KY16 9ST, UK
2
Department of Chemistry, University of Warwick, Coventry CV4 7AL, UK
3
AstraZeneca, Oncology R&D, Research & Early Development, Darwin Building, 310, Cambridge Science Park, Milton Road, Cambridge CB4 0WG, UK
4
Syngenta, Jealott’s Hill International Research Centre, Bracknell RG42 6EY, UK
*
Author to whom correspondence should be addressed.
Molecules 2021, 26(21), 6333; https://doi.org/10.3390/molecules26216333
Submission received: 31 August 2021 / Revised: 12 October 2021 / Accepted: 14 October 2021 / Published: 20 October 2021

Abstract

:
The isothiourea-catalyzed enantioselective 1,6-conjugate addition of para-nitrophenyl esters to 2,6-disubstituted para-quinone methides is reported. para-Nitrophenoxide, generated in situ from initial N-acylation of the isothiourea by the para-nitrophenyl ester, is proposed to facilitate catalyst turnover in this transformation. A range of para-nitrophenyl ester products can be isolated, or derivatized in situ by addition of benzylamine to give amides at up to 99% yield. Although low diastereocontrol is observed, the diastereoisomeric ester products are separable and formed with high enantiocontrol (up to 94:6 er).

Graphical Abstract

1. Introduction

Quinone methides (QMs) are electrophilic compounds composed of a cyclohexadiene core bearing a carbonyl either ortho or para to an exocyclic alkylidene unit [1,2]. Due to their electrophilicity [3,4,5,6], QMs have been used in a variety of biological and medicinal processes [1,2,7], are present within natural products and pharmaceuticals [1,2,8,9,10,11], and have been applied as electrophiles in a variety of synthetic reactions [1,2,12,13,14,15,16,17,18]. While ortho-QMs have been used extensively in enantioselective catalysis [19], particularly as components in formal [4+2] cycloaddition reactions, the use of para-QMs has only recently received increased attention [19,20,21,22,23]. The majority of enantioselective organocatalytic methods that involve para-QMs have utilized Brønsted acid [24,25,26,27,28,29,30,31,32,33] or hydrogen bonding catalysts [34,35,36,37,38], with only a relatively small number of examples using Lewis base catalysis [39,40,41,42,43,44,45,46,47,48,49,50,51].
C(1)-Ammonium enolate intermediates [52,53,54,55], generated by the reaction of a tertiary amine Lewis base catalyst with a ketene, anhydride or acyl imidazole [56], have found widespread application for the synthesis of heterocyclic scaffolds in high yield and with excellent enantiocontrol. Traditionally these approaches have been limited by the requirement for the electrophilic reaction partner to contain a latent nucleophilic site to facilitate catalyst turnover. This conceptual obstacle has resulted in catalysis via C(1)-ammonium enolates being mostly applied for formal cycloaddition reactions. More recently, aryl esters have emerged as alternative C(1)-ammonium enolate precursors [55,57,58]. Significantly, following acylation of the tertiary amine catalyst by the aryl ester, a nucleophilic aryloxide is liberated, which may be exploited again in the catalytic cycle to facilitate ammonium enolate formation and catalyst turnover (Scheme 1a) [55,59]. This strategy offers a potentially general solution to allow the expansion of electrophile scope within catalytic processes using C(1)-ammonium enolate intermediates. In 2014, we applied this concept for the isothiourea-catalyzed[2,3]-rearrangement of allylic ammonium ylides (Scheme 1b) [60,61,62,63,64]. More recently, this approach has been used by Snaddon (Scheme 1c) [65,66,67,68,69,70,71,72], Hartwig (Scheme 1d) [73] and Gong [74,75] for co-operative isothiourea/transition metal-catalyzed α-functionalization of pentafluorophenyl esters. In both cases, an isothiourea-derived C(1)-ammonium enolate is intercepted by an electrophilic transition metal complex to affect an allylation or benzylation reaction. We have expanded the scope of electrophiles applicable within this catalyst turnover strategy to include iminium ions generated under either photoredox conditions or Brønsted acid catalysis, as well as bis-sulfone Michael acceptors, and pyridinium salts (Scheme 1e) [76,77,78,79]. Recently, Waser also reported an elegant example of this turnover strategy for the enantioselective α-chlorination of pentafluorophenyl esters [80], while Zheng and co-workers reported a related approach using diphenyl methanol as an external turnover reagent for the fluorination of carboxylic acids [81,82]. A significant challenge within this area is the identification of electrophilic reaction partners that react with the catalytically-generated C(1)-ammonium enolate, but are compatible with the nucleophilic tertiary amine catalyst and aryloxide, which is essential for catalyst turnover. Building upon this conceptual platform, it was envisaged that para-QMs may be suitable electrophiles to apply in formal 1,6-conjugate additions.

2. Results

2.1. Reaction Optimization

Initial studies focused on the isothiourea-catalyzed 1,6-conjugate addition of para-nitrophenyl (PNP) ester 1 to 2,6-di-tert-butyl para-QM 5 (Table 1). Benzylamine was added at the end of the reaction to convert the PNP ester product to the corresponding amide. Based on our previous experience, the amide was expected to be more stable to chromatographic purification. Using tetramisole HCl 6 (20 mol%) as the catalyst, i-Pr2NEt as the base and MeCN as the solvent gave a 55:45 ratio of diastereoisomeric amide products 7 and 8 in high yield (82%) and with excellent enantioselectivity (7: 97:3 er; 8: 94:6 er) (Entry 1). Chromatographic separation of the diastereoisomers was not possible, however, the enantioenrichment of both 7 and 8 could be reliably determined by chiral stationary phase (CSP)-HPLC analysis of the mixture. A control reaction in the absence of the catalyst showed no conversion (Entry 2). The use of six alternative solvents was investigated (PhMe, CH2Cl2, CHCl3, THF, 1,4-dioxane and DMF), (see the Supporting Information for details) however, only the use of DMF provided any conversion to the product, indicating that solvent polarity may be significant for the success of this transformation. A control reaction in DMF in the absence of the catalyst, however, also led to comparable conversion to the product, consistent with the operation of a competitive Brønsted base-promoted reaction (see the Supporting Information for details). Taking MeCN as the optimal solvent, the use of eight different organic and inorganic bases was investigated (see the Supporting Information for details). Of those tested, Et3N provided an improved yield of 98%, whilst maintaining comparable diastereo- and enantioselectivity (Entry 3). The use of alternative aryl esters was next probed, with pentafluorophenyl ester 2 and bis(trifluoromethyl)phenyl ester 3 giving amide products 7 and 8 in high yield, but with lower enantioselectivity than when using PNP ester 1 (Entries 4 and 5). The use of 2,4,6-trichlorophenyl ester 4 resulted in only 31% yield (Entry 6), which is consistent with previous studies in this field [65,72,76,79], and most likely reflects the increased steric hindrance of the aryloxide attenuating its nucleophilicity. Finally, using PNP ester 1, the catalyst loading could be reduced to 5 mol% with only a small drop in stereoselectivity (Entry 7), while heating the reaction to 40 °C provided a slight improvement in yield (Entry 8). The reaction could also be performed in the absence of a base (Entry 9), however, slightly lower yield was obtained and therefore, during investigation of the substrate scope, Et3N was routinely used as an auxiliary base.

2.2. Reaction Scope and Limitations

Due to the low diastereoselectivity observed using 2,6-di-tert-butyl para-QM 5, the alternative use of 2,6-disubstituted para-QMs were investigated (Scheme 2). 2,6-Dimethyl, dibromo and diphenyl para-QMs 911 bearing a phenyl substituent at the exocyclic olefin were applied under optimized conditions. In all cases significantly lower conversion was observed (≤43%), and the amide products 1416 were difficult to isolate. Based on analysis of the crude reaction mixture by 1H NMR spectroscopy, the 2,6-dimethyl and dibromo-substituted analogues 14 and 15 were obtained with marginally improved diastereoselectivity (~70:30 dr), indicating that alternative substituents in these positions could prove beneficial if the products were isolable. Next, variation of the exocyclic substituent was probed. Incorporation of a methyl group at this position provided no improvement in dr, but the amide product 17 was isolated in a 50% yield and with moderate enantioenrichment for both diastereoisomers. Finally, incorporation of a 2-naphthyl substituent at this position was well tolerated, with amide 18 obtained in an 83% yield, 70:30 dr and excellent enantiocontrol (94:6 er) for the major diastereoisomer.
Although structural variation of the para-QM provided marginal improvements in dr, the use of 2,6-di-tert-butyl para-QM 5 was considered most convenient for further investigations due to its stability and ease of synthesis, and the higher yields of product obtained from catalysis. To investigate if the dr obtained in these reactions was a manifestation of a kinetic or thermodynamic preference, isolation of diastereoisomeric PNP esters 19 and 20 and resubjection to catalysis conditions was attempted. Although the diastereoisomeric amides 7 and 8 had proved difficult to separate, the corresponding PNP esters 19 and 20 were chromatographically separable and displayed high stability. Epimerization studies were conducted using each diastereoisomer through sequential treatment with i-Pr2NEt, (S)-TM HCl 6, para-nitrophenoxide and benzylamine (Scheme 3). These experiments were followed by in situ 1H NMR spectroscopic analysis and revealed no epimerization in either case. This indicates the dr obtained in the catalytic reaction most likely reflects the inherent diastereoselectivity of the transformation. Following separation of the diastereoisomers, the absolute configuration of the major diastereoisomer could also be confirmed as (2S,1′R) by single crystal X-ray crystallographic analysis [83,84]. Based on literature precedent, the (S)-configuration at C(2) was expected to be generated under catalyst-control, and therefore the absolute configuration of the minor diastereoisomer was predicted to be (2S,1′S).
Having established the stability of the PNP ester products and demonstrated the potential to separate the diastereoisomers by column chromatography, the scope of the catalytic transformation was investigated through variation of the PNP ester substrate. In each case, the PNP ester product diastereoisomers were at least partially separable, enabling unambiguous characterization. To test the applicability of the procedure, p-tolyl-substituted PNP ester product 19 was prepared on a larger scale (1.25 mmol) (Scheme 4). A combined 85% yield of both diastereoisomers was obtained, with comparable stereoselectivity to that observed when the reaction was conducted on an analytical scale (Table 1, Entry 7). The generality of the procedure was further probed using five electronically- and sterically-differentiated PNP esters. Introduction of an electron-donating 4-methoxy substituent was well tolerated, with PNP ester 26 obtained in quantitative yield, 60:40 dr and with high enantioselectivity for both diastereoisomers. Under the optimized conditions, the introduction of an electron-withdrawing 4-trifluoromethyl group resulted in low enantioselectivity (27: 63:37 er), which was attributed to a competitive Brønsted base-promoted background reaction. Consistent with this hypothesis, repeating the reaction in the absence of Et3N, and using the free base of the isothiourea catalyst 6, provided PNP ester product 27 in 94% yield and significantly improved enantioselectivity (88:12 ermaj; 79:21 ermin). A similar effect was observed when using a 2-naphthyl-substituted PNP ester, with optimal enantioselectivity obtained in the absence of an auxiliary base (28: 91:9 ermaj; 85:15 ermin). Introduction of a sterically-imposing 1-naphthyl or a heteroaromatic thienyl substituent was also tolerated, with 29 and 30 obtained in excellent yield and with high enantioselectivity.

2.3. Proposed Mechanism

The mechanism of this transformation is proposed to begin with N-acylation of the free base isothiourea catalyst 6 by PNP ester 31 to generate the corresponding acyl ammonium para-nitrophenoxide ion pair 32 (Scheme 5). Subsequent deprotonation leads to (Z)-ammonium enolate 33. Based on previous mechanistic studies [78,85], and the catalytic activity observed in the absence of an auxiliary base (Table 1, Entry 9), deprotonation is likely to be affected by the para-nitrophenoxide counterion. 1,6-Conjugate addition of ammonium enolate 33 to para-QM electrophile 34, followed by protonation, gives acyl ammonium intermediate 35. Finally, regeneration of catalyst 6, and concurrent release of product 37, is proposed to be facilitated by para-nitrophenoxide [59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,86,87]. Although not essential for reactivity, the addition of Et3N as an auxiliary base may be beneficial as a proton shuttle, and to maintain the isothiourea catalyst in its non-protonated form 6 [77,86]. The enantioselectivity of the transformation indicates the C–C bond forming event takes place on the Si-face of the ammonium enolate. This selectivity can be rationalized through preferential formation of the (Z)-ammonium enolate [76,77,78,79,85], which is conformationally-restricted by an intramolecular 1,5-O⋯S interaction [61,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104] and results in the phenyl stereodirecting group of the catalyst blocking the enolate Re-face. The observed poor diastereoselectivity can be tentatively rationalized by a simple stereochemical model that assumes a favored, open pre-transition state assembly where steric interactions are minimized about the forming C–C bond. Minimal differentiation between the aryl- and quinone substituents of the para-QM quinone leads to the two transition state assemblies 38 and 39 that give the major and minor diastereoisomers, respectively.

3. Materials and Methods

3.1. General Procedure for the Enantioselective 1,6-Addition

In a flame-dried vial, the requisite para-quinone methide (1.0 equiv.), aryl ester (1.5 equiv.), (S)-TM HCl (5 mol%), Et3N (1.0 equiv.) and anhydrous MeCN (0.6 M) was added and stirred at r.t. for 24 h. The reaction was then quenched with benzylamine (5.0 equiv.) and stirred at r.t. for a further 12 h before being concentrated in vacuo. The residue was diluted with EtOAc (20 mL) and washed successively with 10% citric acid (20 mL × 1), aqueous NaOH (20 mL × 3) and brine (20 mL × 1). The organic layer was extracted, dried over MgSO4 and the filtrate was concentrated in vacuo. The crude material was purified by flash silica column chromatography to give the desired product.

3.2. Representative Synthesis and Characterization of Compounds 7 and 8 (Entry 8)

Following the general procedure above, 4-benzylidene-2,6-di-tert-butylcyclohexa-2,5-dien-1-one 5 (74 mg, 0.25 mmol), 4’-nitrophenyl 2-(p-tolyl)acetate 1 (102 mg, 0.375 mmol), (S)-TM HCl 6 (3 mg, 5 mol%) and Et3N (35 μL, 0.25 mmol) were dissolved in anhydrous MeCN (0.42 mL). The reaction mixture was stirred at 40 °C for 24 h before being quenched with benzylamine (137 μL, 1.25 mmol) at r.t. to give a crude mixture containing the title compound in 60:40 dr. The mixture was purified by flash silica column chromatography (petroleum ether/EtOAc, 85:15) to afford diastereoisomers 7 and 8 (60:40 dr) (132 mg, 99%) as a pale yellow solid.
mp 126–128 °C; [ α ] D 20 +10.0 (c 1.0, CHCl3); IR νmax (film)/cm−1 3638 (O−H) 3304 (N−H), 2955 (C−H), 1645 (C=O); HRMS (ESI+) C37H43NO2 ([M + H]+), found 534.3359, requires 534.3367 (−1.4 ppm).
Data for major diastereoisomer (7): Chiral HPLC analysis, Chiralpak AD-H (10% i-PrOH/hexane, flow rate 1.5 mLmin−1, 211 nm, 40 °C), tR 8.5 min and 29.3 min, 92:8 er; 1H NMR (500 MHz, CDCl3) δH: 1.42 (18H, s, (C(3″″)C(CH3)3, C(5″″)C(CH3)3), 2.24 (3H, s, C(4″)CH3), 3.90–4.05 (2H, m, C(2)H, CHAPh), 4.44 (1H, dd, J 15.0, 6.8, CHBPh), 4.82 (1H, d, J 11.7, C(1’)H), 5.14 (1H, s, OH), 5.55 (1H, t, J 5.6, NH), 6.71–6.76 (2H, m, Ar), 6.96–7.05 (2H, m, Ar), 7.09–7.15 (2H, m, Ar), 7.15–7.22 (4H, m, Ar), 7.23–7.30 (4H, m, C(4‴)H, C(2″″)H, C(6″″)H, Ar), 7.34 (1H, t, J 7.6, Ar), 7.43–7.51 (1H, m, Ar); 13C{1H} NMR (126 MHz, CDCl3) δC: 21.0 (C(4″)CH3), 30.4 (C(3″″)C(CH3)3, C(5″″)C(CH3)3), 34.4 (C(3″″)C(CH3)3, C(5″″)C(CH3)3), 43.6 (CH2Ph), 54.1 (C(1’)), 59.5 (C(2)), 124.7 (C(2″″), C(6″″)), 125.8 (C(4‴)), 128.2 (Ar), 127.2 (Ar), 127.3 (Ar), 128.0 (Ar), 128.4 (Ar), 128.5 (Ar), 128.6 (Ar), 129.0 (Ar), 133.8 (C(1‴)), 135.2 (C(1″)), 135.6 (C(3″″), C(5″″)), 136.4 (C(4″)), 137.9 (i-Ph), 142.4 (C(1″″)), 152.4 (C(4″″)), 172.1 (C(1)).
Selected data for minor diastereoisomer (8): Chiral HPLC analysis, Chiralpak AD-H (10% i-PrOH/hexane, flow rate 1.5 mLmin−1, 211 nm, 40 °C), tR 3.8 min and 18.6 min, 85:15 er; 1H NMR (500 MHz, CDCl3) δH: 1.27 (18H, s, (C(3″″)C(CH3)3, C(5″″)C(CH3)3), 2.27 (3H, s, C(4″)CH3), 3.90–4.05 (2H, m, C(2)H, CHAPh), 4.50 (1H, dd, J 15.0, 6.8, CHBPh), 4.69 (1H, d, J 11.7, C(1’)H), 4.89 (1H, s, OH), 5.69 (1H, t, J 5.6, NH), 6.81–6.84 (2H, m, Ar); 13C{1H} NMR (126 MHz, CDCl3) δC: 21.0 (C(4″)CH3), 30.2 (C(3″″)C(CH3)3, C(5″″)C(CH3)3), 34.2 (C(3″″)C(CH3)3, C(5″″)C(CH3)3), 43.4 (CH2Ph), 54.7 (C(1’)), 59.5 (C(2)), 125.3 (C(2″″), C(6″″)), 126.3 (C(4‴)), 127.5 (Ar), 128.2 (Ar), 128.4 (Ar), 128.8 (Ar), 132.1 (C(1‴)), 134.9 (C(3″″), C(5″″)), 135.5 (C(1″)), 136.4 (C(4″)), 138.2 (i-Ph), 143.6 (C(1″″), 151.7 (C(4″″)), 172.0 (C(1)).

4. Conclusions

An isothiourea-catalyzed enantioselective 1,6-conjugate addition of para-nitrophenyl (PNP) esters to para-quinone methides (QMs) has been developed. Variation of the arylacetic ester and para-QM substrates has provided a range of functionalized products in generally excellent yields and high enantiocontrol (up to 94:6 er). An inherent limitation of the method is that the products were routinely obtained in ~60:40 dr. This diastereoselectivity was shown to arise from kinetic control, but was relatively insensitive to changes in reaction conditions and structural variation of the substrates. Although the dr could not be improved, the diastereoisomeric PNP ester products could be separated by column chromatography. The success of this catalytic methodology is proposed to rely upon the para-nitrophenoxide, expelled during N-acylation of the catalyst, to facilitate catalyst turnover and release the product. Current work in our laboratory is focused on further applications of using in situ-generated aryloxides to promote catalyst turnover in Lewis base catalysis.

Supplementary Materials

Full experimental procedures, characterization data, NMR spectra and HPLC chromatograms for all new compounds, as well as crystallographic data for product 19 (CCDC 1992504) are available online.

Author Contributions

Conceptualization, A.D.S. and J.N.A.; investigation, J.N.A.; W.C.H.; C.M.; X-ray crystallographic analysis, A.M.Z.S.; writing—original draft preparation, M.D.G.; writing—review and editing, A.D.S. and all authors; supervision, D.S., S.N. and A.D.S.; funding acquisition, A.D.S. All authors have read and agreed to the published version of the manuscript.

Funding

We thank the ERC under the European Union’s Seventh Framework Programme (FP7/2007-2013)/E.R.C. grant agreement n° 279850, AstraZeneca and EPSRC (EP/M506631/1 (J.N.A.)), Syngenta and the EPSRC Centre for Doctoral Training in Critical Resource Catalysis (CRITICAT, EP/L016419/1 (W.C.H.)), and EPSRC (EP/M508214/1 (C.M.)) for funding. A.D.S. thanks the Royal Society for a Wolfson Research Merit Award. We thank the EPSRC UK National Mass Spectrometry Facility at Swansea University.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The research data underpinning this publication can be found at DOI: 10.17630/f6cf6c80-483d-4f16-bd79-80e1537513b2.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are available from the authors on request.

References and Notes

  1. Rokita, S.E. Quinone Methides; Wiley: Hoboken, NJ, USA, 2009. [Google Scholar]
  2. Toteva, M.M.; Richard, J.P. The generation and reactions of quinone methides. Adv. Phys. Org. Chem. 2011, 45, 39–91. [Google Scholar] [PubMed] [Green Version]
  3. Turner, A.B. Quinone methides. Q. Rev. Chem. Soc. 1964, 18, 347–360. [Google Scholar] [CrossRef]
  4. Wagner, H.-U.R. Gompper. In The Chemistry of Quinonoid Compounds; Wiley: London, UK, 1974; pp. 1145–1178. [Google Scholar]
  5. Richter, D.; Hampel, N.; Singer, T.; Ofial, A.R.; Mayr, H. Synthesis and Characterization of Novel Quinone Methides: Reference Electrophiles for the Construction of Nucleophilicity Scales. Eur. J. Org. Chem. 2009, 3203–3211. [Google Scholar] [CrossRef]
  6. Singh, M.S. Reactive Intermediates in Organic Chemistry: Structure and Mechanism; Wiley-VCH: Weinheim, Germany, 2014. [Google Scholar]
  7. Freccero, M. Quinone methides as alkylating and cross-linking agents. Mini Rev. Org. Chem. 2004, 1, 403–415. [Google Scholar] [CrossRef]
  8. Peters, M.G. Chemical modifications of biopolymers by quinones and quinone methides. Angew. Chem. Int. Ed. 1989, 28, 555–570, Angew. Chem.1989, 101, 572–587. [Google Scholar] [CrossRef]
  9. Martin, H.J.; Magauer, T.; Mulzer, J. In Pursuit of a Competitive Target: Total Synthesis of the Antibiotic Kendomycin. Angew. Chem. Int. Ed. 2010, 49, 5614–5626, Angew. Chem.2010, 122, 5746–5758. [Google Scholar] [CrossRef]
  10. Jansen, R.; Gerth, K.; Steinmetz, H.; Reinecke, S.; Kessler, W.; Kirschning, A.; Müller, R. Elansolid A3, a Unique p-Quinone Methide Antibiotic from Chitinophaga sancti. Chem. Eur. J. 2011, 17, 7739–7744. [Google Scholar] [CrossRef]
  11. Dehn, R.; Katsuyama, Y.; Weber, A.; Gerth, K.; Jansen, R.; Steinmetz, H.; Höfle, G.; Müller, R.; Kirschning, A. Molecular Basis of Elansolid Biosynthesis: Evidence for an Unprecedented Quinone Methide Initiated Intramolecular Diels–Alder Cycloaddition/Macrolactonization. Angew. Chem. Int. Ed. 2011, 50, 3882–3887, Angew. Chem.2011, 123, 3968–3973. [Google Scholar] [CrossRef]
  12. Gai, K.; Fang, X.; Li, X.; Xu, J.; Wu, X.; Lin, A.; Yao, H. Synthesis of spiro[2.5]octa-4,7-dien-6-one with consecutive quaternary centers via 1,6-conjugate addition induced dearomatization of para-quinone methides. Chem. Commun. 2015, 51, 15831–15834. [Google Scholar] [CrossRef]
  13. Lόpez, A.; Parra, A.; Jarava-Barrera, C.; Tortosa, M. Copper-catalyzed silylation of p-quinone methides: New entry to dibenzylic silanes. Chem. Commun. 2015, 51, 17684–17687. [Google Scholar] [CrossRef] [Green Version]
  14. Ramanjaneyulu, B.T.; Mahesh, S.; Anand, R.V. Bis(amino)cyclopropenylidene-Catalyzed 1,6-Conjugate Addition of Aromatic Aldehydes to para-Quinone Methides: Expedient Access to α,α′-Diarylated Ketones. Org. Lett. 2015, 17, 3952–3955. [Google Scholar] [CrossRef]
  15. Yuan, Z.; Fang, X.; Li, X.; Wu, J.; Yao, H.; Lin, A. 1,6-Conjugated Addition-Mediated [2+1] Annulation: Approach to Spiro[2.5]octa-4,7-dien-6-one. J. Org. Chem. 2015, 80, 11123–11130. [Google Scholar] [CrossRef] [PubMed]
  16. Reddy, V.; Anand, R.V. Expedient Access to Unsymmetrical Diarylindolylmethanes through Palladium-Catalyzed Domino Electrophilic Cyclization–Extended Conjugate Addition Approach. Org. Lett. 2015, 17, 3390–3393. [Google Scholar] [CrossRef]
  17. Roiser, L.; Zielke, K.; Waser, M. Enantioselective Spirocyclopropanation of para-Quinone Methides using Ammonium Ylides. Org. Lett. 2017, 19, 2338–2341. [Google Scholar] [CrossRef]
  18. Roiser, L.; Zielke, K.; Waser, M. Formal (4+1) Cyclization of Ammonium Ylides with Vinylogous para-Quinone Methides. Synthesis 2018, 50, 4047–4054. [Google Scholar]
  19. Caruana, L.; Fochi, M.; Bernardi, L. The Emergence of Quinone Methides in Asymmetric Organocatalysis. Molecules 2015, 20, 11733–11764. [Google Scholar] [CrossRef]
  20. Parra, A.; Tortosa, M. para-Quinone Methide: A New Player in Asymmetric Catalysis. ChemCatChem 2015, 7, 1524–1526. [Google Scholar] [CrossRef] [Green Version]
  21. Chauhan, P.; Kaya, U.; Enders, D. Advances in Organocatalytic 1,6-Addition Reactions: Enantioselective Construction of Remote Stereogenic Centers. Adv. Synth. Catal. 2017, 359, 888–912. [Google Scholar] [CrossRef]
  22. Li, W.; Xu, X.; Zhang, P.; Li, P. Recent Advances in the Catalytic Enantioselective Reactions of para-Quinone Methides. Chem. Asian J. 2018, 13, 2350–2359. [Google Scholar] [CrossRef]
  23. Wang, J.-Y.; Hao, W.-J.; Tu, S.-J.; Jiang, B. Recent developments in 1,6-addition reactions of para-quinone methides (p-QMs). Org. Chem. Front. 2020, 7, 1743–1778. [Google Scholar] [CrossRef]
  24. Wang, Z.; Wong, Y.F.; Sun, J. Catalytic Asymmetric 1,6-Conjugate Addition of para-Quinone Methides: Formation of All-Carbon Quaternary Stereocenters. Angew. Chem. Int. Ed. 2015, 54, 13711–13714, Angew. Chem.2015, 127, 13915–13918. [Google Scholar] [CrossRef]
  25. Dong, N.; Zhang, Z.-P.; Xue, X.-S.; Li, X.; Cheng, J.-P. Phosphoric Acid Catalyzed Asymmetric 1,6-Conjugate Addition of Thioacetic Acid to para-Quinone Methides. Angew. Chem. Int. Ed. 2016, 55, 1460–1464, Angew. Chem.2016, 128, 1482–1486. [Google Scholar] [CrossRef]
  26. Wong, Y.F.; Wang, Z.; Sun, J. Chiral phosphoric acid catalyzed asymmetric addition of naphthols to para-quinone methides. Org. Biomol. Chem. 2016, 14, 5751–5754. [Google Scholar] [CrossRef]
  27. Chen, M.; Sun, J. How understanding the role of an additive can lead to an improved synthetic protocol without an additive: Organocatalytic synthesis of chiral diarylmethyl alkynes. Angew. Chem. 2017, 129, 12128–12132, Angew. Chem. Int. Ed.2017, 56, 11966–11970. [Google Scholar] [CrossRef]
  28. Yan, J.; Chen, M.; Sung, H.H.-Y.; Williams, I.D.; Sun, J. An Organocatalytic Asymmetric Synthesis of Chiral β,β-Diaryl-α-amino Acids via Addition of Azlactones to In Situ Generated para-Quinone Methides. Chem. Asian J. 2018, 13, 2440–2444. [Google Scholar] [CrossRef]
  29. Rahman, A.; Zhou, Q.; Lin, X. Asymmetric organocatalytic synthesis of chiral 3,3-disubstituted oxindoles via a 1, 6-conjugate addition reaction. Org. Biomol. Chem. 2018, 16, 5301–5309. [Google Scholar] [CrossRef]
  30. Li, W.; Xu, X.; Liu, Y.; Gao, H.; Cheng, Y.; Li, P. Enantioselective Organocatalytic 1,6-Addition of Azlactones to para-Quinone Methides: An Access to α,α-Disubstituted and β,β-Diaryl-α-amino acid Esters. Org. Lett. 2018, 20, 1142–1145. [Google Scholar] [CrossRef] [PubMed]
  31. Wang, J.-R.; Jiang, X.-L.; Hang, Q.-Q.; Zhang, S.; Mei, G.-J.; Shi, F. Catalytic Asymmetric Conjugate Addition of Indoles to para-Quinone Methide Derivatives. J. Org. Chem. 2019, 84, 7829–7839. [Google Scholar] [CrossRef]
  32. Wang, Z.; Zhu, Y.; Pan, X.; Wang, G.; Liu, L. Synthesis of Chiral Triarylmethanes Bearing All-Carbon Quaternary Stereocenters: Catalytic Asymmetric Oxidative Cross-Coupling of 2,2-Diarylacetonitriles and (Hetero)arenes. Angew. Chem. 2020, 132, 3077–3081, Angew. Chem.2020, 59, 3053–3057. [Google Scholar] [CrossRef]
  33. Niu, J.-P.; Nie, J.; Li, S.; Ma, J.-A. Organocatalytic asymmetric synthesis of β,β-diaryl ketones via one-pot tandem dehydration/1,6-addition/decarboxylation transformation of β-keto acids and 4-hydroxybenzyl alcohols. Chem. Commun. 2020, 56, 8687–8690. [Google Scholar] [CrossRef] [PubMed]
  34. Zhao, K.; Zhi, Y.; Wang, A.; Enders, D. Asymmetric Organocatalytic Synthesis of 3-Diarylmethine-Substituted Oxindoles Bearing a Quaternary Stereocenter via 1,6-Conjugate Addition to para-Quinone Methides. ACS Catal. 2016, 6, 657–660. [Google Scholar] [CrossRef]
  35. Li, X.; Xu, X.; Wei, W.; Lin, A.; Yao, H. Organocatalyzed Asymmetric 1,6-Conjugate Addition of para-Quinone Methides with Dicyanoolefins. Org. Lett. 2016, 18, 428–431. [Google Scholar] [CrossRef]
  36. Deng, Y.-H.; Zhang, X.-Z.; Yu, K.-Y.; Yan, X.; Du, J.-Y.; Huang, H.; Fan, C.-A. Bifunctional tertiary amine-squaramide catalyzed asymmetric catalytic 1,6-conjugate addition/aromatization of para-quinone methides with oxindoles. Chem. Commun. 2016, 52, 4183–4186. [Google Scholar] [CrossRef] [PubMed]
  37. Toràn, R.; Vila, C.; Sanz-Marco, A.; Muñoz, M.C.; Pedro, J.R.; Blay, G. Organocatalytic Enantioselective 1,6-aza-Michael Addition of Isoxazolin-5-ones to p-Quinone Methides. Eur. J. Org. Chem. 2020, 2020, 627–630. [Google Scholar] [CrossRef]
  38. Wang, L.; Yang, F.; Xua, X.; Jiang, J. Organocatalytic 1,6-hydrophosphination of para-quinone methides: Enantioselective access to chiral 3-phosphoxindoles bearing phosphorus-substituted quaternary carbon stereocenters. Org. Chem. Front. 2021, 8, 2002–2008. [Google Scholar] [CrossRef]
  39. Caruana, L.; Kniep, F.; Johansen, T.K.; Poulsen, H.P.; Jørgensen, K.A. A New Organocatalytic Concept for Asymmetric α-Alkylation of Aldehydes. J. Am. Chem. Soc. 2014, 136, 15929–15932. [Google Scholar] [CrossRef]
  40. Li, S.; Liu, Y.; Huang, B.; Zhou, T.; Tao, H.; Xiao, Y.; Liu, L.; Zhang, J. Phosphine-Catalyzed Asymmetric Intermolecular Cross-Vinylogous Rauhut–Currier Reactions of Vinyl Ketones with para-Quinone Methides. ACS Catal. 2017, 7, 2805–2809. [Google Scholar] [CrossRef]
  41. Kang, T.-C.; Wu, L.-P.; Yu, Q.-W.; Wu, X.-Y. Enantioselective Rauhut–Currier-Type 1,6-Conjugate Addition of Methyl Vinyl Ketone to para-Quinone Methides. Chem. Eur. J. 2017, 23, 6509–6513. [Google Scholar] [CrossRef]
  42. Wang, D.; Song, Z.-F.; Wang, W.-J.; Xu, T. Highly Regio- and Enantioselective Dienylation of p-Quinone Methides Enabled by an Organocatalyzed Isomerization/Addition Cascade of Allenoates. Org. Lett. 2019, 21, 3963–3967. [Google Scholar] [CrossRef]
  43. Li, W.; Yuan, H.; Liu, Z.; Zhang, Z.; Cheng, Y.; Li, P. NHC-Catalyzed Enantioselective [4+3] Cycloaddition of Ortho-Hydroxyphenyl Substituted Para-Quinone Methides with Isatin-Derived Enals. Adv. Synth. Catal. 2018, 360, 2460–2464. [Google Scholar] [CrossRef]
  44. Liu, Q.; Chen, X.-Y.; Rissanen, K.; Ender, D. Asymmetric synthesis of spiro-oxindole-ε-lactones through N-heterocyclic carbene catalysis. Org. Lett. 2018, 20, 3622–3626. [Google Scholar] [CrossRef]
  45. Zhao, M.-X.; Xiang, J.; Zhao, Z.-Q.; Zhao, X.-L.; Shi, M. Asymmetric synthesis of dihydrocoumarins via catalytic sequential 1,6-addition/transesterification of α-isocyanoacetates with para-quinone methides. Org. Biomol. Chem. 2020, 18, 1637–1646. [Google Scholar] [CrossRef] [PubMed]
  46. Roy, S.; Pradhan, S.; Kumar, K.; Chatterjee, I. Asymmetric organocatalytic double 1,6-addition: Rapid access to chiral chromans with molecular complexity. Org. Chem. Front. 2020, 7, 1388–1394. [Google Scholar] [CrossRef]
  47. Chu, W.D.; Zhang, L.F.; Bao, X.; Zhao, X.H.; Zeng, C.; Du, J.Y.; Zhang, G.B.; Wang, F.X.; Ma, X.Y.; Fan, C.A. Asymmetric Catalytic 1,6-Conjugate Addition/Aromatization of para-Quinone Methides: Enantioselective Introduction of Functionalized Diarylmethine Stereogenic Centers. Angew. Chem. Int. Ed. 2013, 52, 9229–9233, Angew. Chem.2013, 125, 9399–9403. [Google Scholar] [CrossRef] [PubMed]
  48. Zhang, X.-Z.; Deng, Y.-H.; Yan, X.; Yu, K.-Y.; Wang, F.-X.; Ma, X.-Y.; Fan, C.-A. Diastereoselective and Enantioselective Synthesis of Unsymmetric β,β-Diaryl-α-Amino Acid Esters via Organocatalytic 1,6-Conjugate Addition of para-Quinone Methides. J. Org. Chem. 2016, 81, 5655–5662. [Google Scholar] [CrossRef]
  49. Ge, L.; Lu, X.; Cheng, C.; Chen, J.; Cao, W.; Wu, X.; Zhao, G. Amide-Phosphonium Salt as Bifunctional Phase Transfer Catalyst for Asymmetric 1,6-Addition of Malonate Esters to para-Quinone Methides. J. Org. Chem. 2016, 81, 9315–9325. [Google Scholar] [CrossRef]
  50. Santra, S.; Porey, A.; Jana, B.; Guin, J. N-Heterocyclic carbenes as chiral Brønsted base catalysts: A highly diastereo- and enantioselective 1,6-addition reaction. Chem. Sci. 2018, 9, 6446–6450. [Google Scholar] [CrossRef] [Green Version]
  51. Eitzinger, A.; Winter, M.; Schörgenhumer, J.; Waser, M. Quaternary β 2, 2-amino acid derivatives by asymmetric addition of isoxazolidin-5-ones to para-quinone methides. Chem. Commun. 2020, 56, 579–582. [Google Scholar] [CrossRef] [Green Version]
  52. Van, K.N.; Morrill, L.C.; Smith, A.D.; Romo, D. Lewis Base Catalysis in Organic Synthesis; Vedejs, E., Denmark, S.E., Eds.; Wiley-VCH: Weinheim, Germany, 2016; Chapter 13; pp. 527–653. [Google Scholar]
  53. Morrill, L.C.; Smith, A.D. Organocatalytic Lewis base functionalisation of carboxylic acids, esters and anhydrides via C1-ammonium or azolium enolates. Chem. Soc. Rev. 2014, 43, 6214–6226. [Google Scholar] [CrossRef] [Green Version]
  54. Gaunt, M.J.; Johansson, C.C.C. Recent developments in the use of catalytic asymmetric ammonium enolates in chemical synthesis. Chem. Rev. 2007, 107, 5596–5605. [Google Scholar] [CrossRef]
  55. McLaughlin, C.; Smith, A.D. Generation and Reactivity of C (1)-Ammonium Enolates by Using Isothiourea Catalysis. Chem. Eur. J. 2021, 27, 1533–1555. [Google Scholar] [CrossRef] [PubMed]
  56. Young, C.M.; Stark, D.G.; West, T.H.; Taylor, J.E.; Smith, A.D. Exploiting the Imidazolium Effect in Base-free Ammonium Enolate Generation: Synthetic and Mechanistic Studies. Angew. Chem. 2016, 128, 14606–14611, Angew. Chem.2016, 55, 14394–14399. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Hao, L.; Chen, X.; Chen, S.; Jiang, K.; Torres, J.; Chi, Y.R. Access to pyridines via DMAP-catalyzed activation of α-chloro acetic ester to react with unsaturated imines. Org. Chem. Front. 2014, 1, 148–150. [Google Scholar] [CrossRef]
  58. Young, C.M.; Taylor, J.E.; Smith, A.D. Evaluating aryl esters as bench-stable C (1)-ammonium enolate precursors in catalytic, enantioselective Michael addition–lactonisations. Org. Biomol. Chem. 2019, 17, 4747–4752. [Google Scholar] [CrossRef] [PubMed]
  59. Hartley, W.C.; O’Riordan, T.J.C.; Smith, A.D. Aryloxide-Promoted Catalyst Turnover in Lewis Base Organocatalysis. Synthesis 2017, 49, 3303–3310. [Google Scholar]
  60. West, T.H.; Daniels, D.S.B.; Slawin, A.M.Z.; Smith, A.D. An isothiourea-catalyzed asymmetric [2, 3]-rearrangement of allylic ammonium ylides. J. Am. Chem. Soc. 2014, 136, 4476–4479. [Google Scholar] [CrossRef] [Green Version]
  61. West, T.H.; Walden, D.M.; Taylor, J.E.; Brueckner, A.C.; Johnston, R.C.; Cheong, P.H.-Y.; Lloyd-Jones, G.C.; Smith, A.D. Catalytic Enantioselective [2, 3]-Rearrangements of Allylic Ammonium Ylides: A Mechanistic and Computational Study. J. Am. Chem. Soc. 2017, 139, 4366–4375. [Google Scholar] [CrossRef] [Green Version]
  62. Spoehrle, S.S.M.; West, T.H.; Taylor, J.E.; Slawin, A.M.Z.; Smith, A.D. Tandem Palladium and Isothiourea Relay Catalysis: Enantioselective Synthesis of α-Amino Acid Derivatives via Allylic Amination and [2, 3]-Sigmatropic Rearrangement. J. Am. Chem. Soc. 2017, 139, 11895–11902. [Google Scholar] [CrossRef] [Green Version]
  63. Kasten, K.; Slawin, A.M.Z.; Smith, A.D. Enantioselective Synthesis of β-Fluoro-β-aryl-α-aminopentenamides by Organocatalytic [2, 3]-Sigmatropic Rearrangement. Org. Lett. 2017, 19, 5182–5185. [Google Scholar] [CrossRef] [Green Version]
  64. West, T.H.; Spoehrle, S.S.M.; Smith, A.D. Isothiourea-catalysed chemo-and enantioselective [2, 3]-sigmatropic rearrangements of N, N-diallyl allylic ammonium ylides. Tetrahedron 2017, 73, 4138–4149. [Google Scholar] [CrossRef] [Green Version]
  65. Schwarz, K.J.; Amos, J.L.; Klein, J.C.; Do, D.T.; Snaddon, T.N. Uniting C1-ammonium enolates and transition metal electrophiles via cooperative catalysis: The direct asymmetric α-allylation of aryl acetic acid esters. J. Am. Chem. Soc. 2016, 138, 5214–5217. [Google Scholar] [CrossRef]
  66. Schwarz, K.J.; Pearson, C.M.; Cintron-Rosado, G.A.; Liu, P.; Snaddon, T.N. Traversing Steric Limitations by Cooperative Lewis Base/Palladium Catalysis: An Enantioselective Synthesis of α-Branched Esters Using 2-Substituted Allyl Electrophiles. Angew. Chem. 2018, 130, 7926–7929, Angew. Chem.2018, 57, 7800–7803. [Google Scholar] [CrossRef]
  67. Scaggs, W.R.; Snaddon, T.N. Enantioselective α-Allylation of Acyclic Esters using B(pin)-Substituted Electrophiles: Independent Regulation of Stereocontrol Elements via Cooperative Pd/Lewis Base Catalysis. Chem. Eur. J. 2018, 24, 14378–14381. [Google Scholar] [CrossRef]
  68. Fyfe, J.W.B.; Kabia, O.M.; Pearson, C.M.; Snaddon, T.N. Si-directed regiocontrol in asymmetric Pd-catalyzed allylic alkylations using C1-ammonium enolate nucleophiles. Tetrahedron 2018, 74, 5383–5391. [Google Scholar] [CrossRef]
  69. Hutchings-Goetz, J.; Yang, C.; Snaddon, T.N. Enantioselective Syntheses of Strychnos and Chelidonium Alkaloids through Regio- and Stereocontrolled Cooperative Catalysis. ACS Catal. 2018, 8, 10537–10544. [Google Scholar] [CrossRef] [Green Version]
  70. Pearson, C.M.; Fyfe, J.W.B.; Snaddon, T.N. A Regio- and Stereodivergent Synthesis of Homoallylic Amines by a One-Pot Cooperative-Catalysis-Based Allylic Alkylation/Hofmann Rearrangement Strategy. Angew. Chem. Int. Ed. 2019, 58, 10521–10527, Angew. Chem.2019, 131, 10631–10637. [Google Scholar] [CrossRef]
  71. Scaggs, W.R.; Scaggs, T.D.; Snaddon, T.N. An enantioselective synthesis of α-alkylated pyrroles via cooperative isothiourea/palladium catalysis. Org. Biomol. Chem. 2019, 17, 1787–1790. [Google Scholar] [CrossRef] [Green Version]
  72. Schwarz, K.J.; Yang, C.; Fyfe, J.W.B.; Snaddon, T.N. Enantioselective α-Benzylation of Acyclic Esters Using π-Extended Electrophiles. Angew. Chem. Int. Ed. 2018, 57, 12102–12105, Angew. Chem.2018, 130, 12278–12281. [Google Scholar] [CrossRef] [PubMed]
  73. Jiang, X.; Beiger, J.J.; Hartwig, J.F. Stereodivergent allylic substitutions with aryl acetic acid esters by synergistic iridium and Lewis base catalysis. J. Am. Chem. Soc. 2017, 139, 87–90. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Song, J.; Zhang, Z.J.; Gong, L.Z. Asymmetric [4+2] Annulation of C1 Ammonium Enolates with Copper-Allenylidenes. Angew. Chem. Int. Ed. 2017, 56, 5212–5216, Angew. Chem.2017, 129, 5296–5300. [Google Scholar] [CrossRef] [PubMed]
  75. Song, J.; Zhang, Z.J.; Chen, S.S.; Fan, T.; Gong, L.Z. Lewis base/copper cooperatively catalyzed asymmetric α-amination of esters with diaziridinone. J. Am. Chem. Soc. 2018, 140, 3177–3180. [Google Scholar] [CrossRef] [PubMed]
  76. Arokianathar, J.N.; Frost, A.B.; Slawin, A.M.Z.; Stead, D.; Smith, A.D. Isothiourea-Catalyzed Enantioselective Addition of 4-Nitrophenyl Esters to Iminium Ions. ACS Catal. 2018, 8, 1067–1075. [Google Scholar] [CrossRef]
  77. Zhao, F.; Shu, C.; Young, C.M.; Carpenter-Warren, C.; Slawin, A.M.Z.; Smith, A.D. Enantioselective Synthesis of α-Aryl-β2-Amino-Esters by Cooperative Isothiourea and Brønsted Acid Catalysis. Angew. Chem. Int. Ed. 2021, 60, 11892–11900, Angew. Chem.2021, 133, 11999–12007. [Google Scholar] [CrossRef]
  78. McLaughlin, C.; Slawin, A.M.Z.; Smith, A.D. Base-free Enantioselective C (1)-Ammonium Enolate Catalysis Exploiting Aryloxides: A Synthetic and Mechanistic Study. Angew. Chem. Int. Ed. 2019, 131, 15255–15263, Angew. Chem.2019, 58, 15111–15119. [Google Scholar] [CrossRef]
  79. McLaughlin, C.; Bitai, J.; Barber, L.; Slawin, A.M.Z.; Smith, A.D. Catalytic enantioselective synthesis of 1, 4-dihydropyridines via the addition of C (1)-ammonium enolates to pyridinium salts. Chem. Sci. 2021. accepted for publication. [Google Scholar] [CrossRef]
  80. Stockhammer, L.; Weinzierl, D.; Bögl, T.; Waser, M. Enantioselective α-Chlorination Reactions of in Situ Generated C1 Ammonium Enolates under Base-Free Conditions. Org. Lett. 2021, 23, 6143–6147. [Google Scholar] [CrossRef]
  81. Kim, B.; Kim, Y.; Lee, S.Y. Stereodivergent Carbon–Carbon Bond Formation between Iminium and Enolate Intermediates by Synergistic Organocatalysis. J. Am. Chem. Soc. 2021, 143, 73–79. [Google Scholar]
  82. Yuan, S.; Liao, C.; Zheng, W.-H. Paracyclophane-Based Isothiourea-Catalyzed Highly Enantioselective α-Fluorination of Carboxylic Acids. Org. Lett. 2021, 23, 4142–4146. [Google Scholar] [CrossRef]
  83. Crystallographic data for (2S,1′R)-19 (CCDC 1992504), Cambridge Crystallographic Data Centre.
  84. The relative and absolute configuration of all other products was assigned by analogy, with consistent diagnostic 1H NMR signals of the proton at C(2) and C(1ʹ) stereocentres, indicating the same major diastereoisomer in each case.
  85. Morrill, L.C.; Douglas, J.; Lebl, T.; Slawin, A.M.Z.; Fox, D.J.; Smith, A.D. Isothiourea-mediated asymmetric Michael-lactonisation of trifluoromethylenones: A synthetic and mechanistic study. Chem. Sci. 2013, 4, 4146–4155. [Google Scholar] [CrossRef]
  86. Matviistsuk, A.; Greenhalgh, M.D.; Antúnez, D.-J.B.; Slawin, A.M.Z.; Smith, A.D. Aryloxide-Facilitated Catalyst Turnover in Enantioselective α, β-Unsaturated Acyl Ammonium Catalysis. Angew. Chem. Int. Ed. 2017, 56, 12282–12287, Angew. Chem.2017, 129, 12450–12455. [Google Scholar] [CrossRef]
  87. Catalyst turnover could also be envisaged through intramolecular nucleophilic displacement by the pendant phenol to give a cyclobutanone intermediate via a formal [2+2]-cycloaddition, which may then be ring-opened by para-nitrophenoxide to give the final product. Based on the high steric congestion of this proposed cyclobutanone intermediate, the mechanism given in Scheme 5 is considered more probable.
  88. Liu, P.; Yang, X.; Birman, V.B.; Houk, K.N. Origin of enantioselectivity in benzotetramisole-catalyzed dynamic kinetic resolution of azlactones. Org. Lett. 2012, 14, 3288–3291. [Google Scholar] [CrossRef] [PubMed]
  89. Abbasov, M.E.; Hudson, B.M.; Tantillo, D.J.; Romo, D. Acylammonium salts as dienophiles in Diels–Alder/lactonization organocascades. J. Am. Chem. Soc. 2014, 136, 4492–4495. [Google Scholar] [CrossRef] [PubMed]
  90. Robinson, E.R.T.; Walden, D.M.; Fallan, C.; Greenhalgh, M.D.; Cheong, P.H.-Y.; Smith, A.D. Non-bonding 1,5-S⋯O interactions govern chemo- and enantioselectivity in isothiourea-catalyzed annulations of benzazoles. Chem. Sci. 2016, 7, 6919–6927. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Greenhalgh, M.D.; Smith, S.M.; Walden, D.M.; Taylor, J.E.; Brice, Z.; Robinson, E.R.T.; Fallan, C.; Cordes, D.B.; Slawin, A.M.Z.; Richardson, H.C.; et al. AC= O⋅⋅⋅ Isothiouronium Interaction Dictates Enantiodiscrimination in Acylative Kinetic Resolutions of Tertiary Heterocyclic Alcohols. Angew. Chem. Int. Ed. 2018, 57, 3200–3206, Angew. Chem.2018, 130, 3254–3260. [Google Scholar] [CrossRef] [Green Version]
  92. Young, C.M.; Elmi, A.; Pascoe, D.J.; Morris, R.K.; McLaughlin, C.; Woods, A.M.; Frost, A.B.; de la Houpliere, A.; Ling, K.B.; Smith, T.K.; et al. The Importance of 1, 5-Oxygen Chalcogen Interactions in Enantioselective Isochalcogenourea Catalysis. Angew. Chem. 2020, 132, 3734–3739, Angew. Chem.2020, 59, 3705–3710. [Google Scholar] [CrossRef]
  93. Pascoe, D.J.; Ling, K.B.; Cockroft, S.L. The origin of chalcogen-bonding interactions. J. Am. Chem. Soc. 2017, 139, 15160–15167. [Google Scholar] [CrossRef] [Green Version]
  94. Nagao, Y.; Miyamoto, S.; Miyamoto, M.; Takeshige, H.; Hayashi, K.; Sano, S.; Shiro, M.; Yamaguchi, K.; Sei, Y. Highly Stereoselective Asymmetric Pummerer Reactions That Incorporate Intermolecular and Intramolecular Nonbonded S⋯O Interactions. J. Am. Chem. Soc. 2006, 128, 9722–9729. [Google Scholar] [CrossRef]
  95. Beno, B.R.; Yeung, K.-S.; Bartberger, M.D.; Pennington, L.D.; Meanwell, N.A. A survey of the role of noncovalent sulfur interactions in drug design. J. Med. Chem. 2015, 58, 4383–4438. [Google Scholar] [CrossRef]
  96. Breugst, M.; Koenig, J.J. σ-Hole Interactions in Catalysis. Eur. J. Org. Chem. 2020, 34, 5473–5487. [Google Scholar] [CrossRef]
  97. Bleiholder, C.; Gleiter, R.; Werz, D.B.; Köppel, H. Theoretical Investigations on Heteronuclear Chalcogen—Chalcogen Interactions: On the Nature of Weak Bonds between Chalcogen Centers. Inorg. Chem. 2007, 46, 2249–2260. [Google Scholar] [CrossRef]
  98. Gleiter, R.; Haberhauer, G.; Werz, D.B.; Rominger, F. From noncovalent chalcogen–chalcogen interactions to supramolecular aggregates: Experiments and calculations. Bleiholder. Chem. Rev. 2018, 118, 2010–2041. [Google Scholar] [CrossRef] [PubMed]
  99. Benz, S.; López-Andarias, J.; Mareda, J.; Sakai, N.; Matile, S. Catalysis with chalcogen bonds. Angew. Chem. 2017, 129, 830–833, Angew. Chem.2017, 56, 812–815. [Google Scholar] [CrossRef] [Green Version]
  100. Wonner, P.; Vogel, L.; Düser, M.; Gomes, L.; Kniep, F.; Mallick, B.; Werz, D.B.; Huber, S.M. Carbon–Halogen Bond Activation by Selenium-Based Chalcogen Bonding. Angew. Chem. Int. Ed. 2017, 56, 12009–12012, Angew. Chem.2017, 129, 12172–12176. [Google Scholar] [CrossRef] [PubMed]
  101. Wonner, P.; Vogel, L.; Kniep, F.; Huber, S.M. Catalytic Carbon–Chlorine Bond Activation by Selenium-Based Chalcogen Bond Donors. Chem. Eur. J. 2017, 23, 16972–16975. [Google Scholar] [CrossRef]
  102. Wonner, P.; Dreger, A.; Vogel, L.; Engelage, E.; Huber, S.M. Chalcogen Bonding Catalysis of a Nitro-Michael Reaction. Angew. Chem. Int. Ed. 2019, 58, 16923–16927, Angew. Chem.2019, 131, 17079–17083. [Google Scholar] [CrossRef] [Green Version]
  103. Wang, W.; Zhu, H.; Liu, S.; Zhao, Z.; Zhang, L.; Hao, J.; Wang, Y. Chalcogen–chalcogen bonding catalysis enables assembly of discrete molecules. J. Am. Chem. Soc. 2019, 141, 9175–9179. [Google Scholar] [CrossRef] [Green Version]
  104. Wang, W.; Zhu, H.; Feng, L.; Yu, Q.; Hao, J.; Zhu, R.; Wang, Y. Dual Chalcogen–Chalcogen Bonding Catalysis. J. Am. Chem. Soc. 2020, 142, 3117–3124. [Google Scholar] [CrossRef]
Scheme 1. Isothiourea-catalyzed enantioselective processes using aryloxide-facilitated catalyst turnover.
Scheme 1. Isothiourea-catalyzed enantioselective processes using aryloxide-facilitated catalyst turnover.
Molecules 26 06333 sch001
Scheme 2. Scope: Variation of para-quinone methide—0.25 mmol scale; only the structure of the major diastereoisomer is shown; isolated yield is a mixture of diastereoisomers; dr determined by 1H NMR spectroscopic analysis of the crude reaction product mixture; er determined by CSP-HPLC analysis. [a] Product not isolable: conversion based on 1H NMR analysis of the crude reaction product mixture; ers could not be determined.
Scheme 2. Scope: Variation of para-quinone methide—0.25 mmol scale; only the structure of the major diastereoisomer is shown; isolated yield is a mixture of diastereoisomers; dr determined by 1H NMR spectroscopic analysis of the crude reaction product mixture; er determined by CSP-HPLC analysis. [a] Product not isolable: conversion based on 1H NMR analysis of the crude reaction product mixture; ers could not be determined.
Molecules 26 06333 sch002
Scheme 3. Control studies and confirmation of absolute configuration of products. The majority of hydrogen atoms are omitted for clarity within X-ray crystal structure representation of 19.
Scheme 3. Control studies and confirmation of absolute configuration of products. The majority of hydrogen atoms are omitted for clarity within X-ray crystal structure representation of 19.
Molecules 26 06333 sch003
Scheme 4. Scope: Variation of para-nitrophenyl ester—0.25 mmol scale; only the structure of the major diastereoisomer is shown; isolated yields are given for the combination of diastereoisomers; dr determined by 1H NMR spectroscopic analysis of the crude reaction product mixture; er determined by CSP-HPLC analysis. [a] 1.25 mmol scale. [b] Conducted at r.t., in the absence of Et3N, and using free base of 6.
Scheme 4. Scope: Variation of para-nitrophenyl ester—0.25 mmol scale; only the structure of the major diastereoisomer is shown; isolated yields are given for the combination of diastereoisomers; dr determined by 1H NMR spectroscopic analysis of the crude reaction product mixture; er determined by CSP-HPLC analysis. [a] 1.25 mmol scale. [b] Conducted at r.t., in the absence of Et3N, and using free base of 6.
Molecules 26 06333 sch004
Scheme 5. Proposed mechanism (only the pathway for the formation of the major diastereoisomer is shown) and transition state assemblies. ArOH corresponds to either PNPOH or the reaction product.
Scheme 5. Proposed mechanism (only the pathway for the formation of the major diastereoisomer is shown) and transition state assemblies. ArOH corresponds to either PNPOH or the reaction product.
Molecules 26 06333 sch005
Table 1. Reaction optimization.
Table 1. Reaction optimization.
Molecules 26 06333 i001
Entry6 (mol%)Aryl EsterBaseYield (%)dr (7:8)er (7)er (8)
1201i-Pr2NEt8255:4597:394:6
201i-Pr2NEt0---
3201Et3N9860:4097:394:6
4202Et3N9965:3581:1994:6
5203Et3N8860:4094:691:9
6204Et3N3155:4584:1671:29
751Et3N9555:4593:788:12
8 [a]51Et3N9960:4092:885:15
9 [a]51none9060:4093:789:11
All reactions were carried out on a 0.25 mmol scale; isolated yields are a mixture of diastereoisomers 7 and 8; dr was determined by 1H NMR spectroscopic analysis of the crude reaction mixture; er was determined by CSP-HPLC analysis: 7 (2S,1′R:2R,1′S) and 8 (2S,1′S:2R,1′R). [a] Reaction performed at 40 °C.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Arokianathar, J.N.; Hartley, W.C.; McLaughlin, C.; Greenhalgh, M.D.; Stead, D.; Ng, S.; Slawin, A.M.Z.; Smith, A.D. Isothiourea-Catalyzed Enantioselective α-Alkylation of Esters via 1,6-Conjugate Addition to para-Quinone Methides. Molecules 2021, 26, 6333. https://doi.org/10.3390/molecules26216333

AMA Style

Arokianathar JN, Hartley WC, McLaughlin C, Greenhalgh MD, Stead D, Ng S, Slawin AMZ, Smith AD. Isothiourea-Catalyzed Enantioselective α-Alkylation of Esters via 1,6-Conjugate Addition to para-Quinone Methides. Molecules. 2021; 26(21):6333. https://doi.org/10.3390/molecules26216333

Chicago/Turabian Style

Arokianathar, Jude N., Will C. Hartley, Calum McLaughlin, Mark D. Greenhalgh, Darren Stead, Sean Ng, Alexandra M. Z. Slawin, and Andrew D. Smith. 2021. "Isothiourea-Catalyzed Enantioselective α-Alkylation of Esters via 1,6-Conjugate Addition to para-Quinone Methides" Molecules 26, no. 21: 6333. https://doi.org/10.3390/molecules26216333

Article Metrics

Back to TopTop