Next Article in Journal
Bioactive Properties of Kakadu Plum-Blended Products
Next Article in Special Issue
Palladium-Catalyzed Stereoselective Construction of 1,3-Stereocenters Displaying Axial and Central Chirality via Asymmetric Alkylations
Previous Article in Journal
Correlation between Magnetic and Dielectric Response of CoFe2O4:Li1+/Zn2+ Nanopowders Having Improved Structural and Morphological Properties
Previous Article in Special Issue
Copper-Catalyzed Asymmetric Dearomative [3+2] Cycloaddition of Nitroheteroarenes with Azomethines
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Chiral Phosphine Catalyzed Allylic Alkylation of Benzylidene Succinimides with Morita–Baylis–Hillman Carbonates

1
Shenzhen Grubbs Institute, Department of Chemistry, Guangdong Provincial Key Laboratory of Catalysis, College of Science, Southern University of Science and Technology (SUSTech), Shenzhen 518055, China
2
Department of Chemistry, The Hong Kong University of Science and Technology, Clear Water Bay, Kowloon, Hong Kong SAR, China
3
Southern University of Science and Technology Guangming Advanced Research Institute, Southern University of Science and Technology, Shenzhen 518055, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(6), 2825; https://doi.org/10.3390/molecules28062825
Submission received: 28 February 2023 / Revised: 16 March 2023 / Accepted: 20 March 2023 / Published: 21 March 2023
(This article belongs to the Special Issue Recent Advances of Catalytic Asymmetric Synthesis)

Abstract

:
Owing to their unique chemical properties, α-alkylidene succinimides generally act as versatile synthons in organic synthesis. Compared with well-established annulations, nucleophilic alkylations of α-alkylidene succinimides are very limited. Accordingly, an organocatalytic allylic alkylation of α-benzylidene succinimides with Morita–Baylis–Hillman (MBH) carbonates was established. In the presence of a chiral phosphine catalyst, α-benzylidene succinimides reacted smoothly with MBH carbonates under mild conditions to furnish a series of optical active succinimides in high yields and enantioselectivities. Different from the reported results, the organocatalytic enantioselective construction of pyrrolidine-2,5-dione frameworks bearing contiguous chiral tertiary carbon centers was achieved via this synthetic strategy. Scaling up the reaction indicated that it is a practical strategy for the organocatalytic enantioselective alkylation of α-alkylidene succinimides. A possible reaction mechanism was also proposed.

1. Introduction

The pyrrolidine-2,5-dione scaffold is the core unit of numerous bioactive compounds with anticonvulsant [1], antimycobacterial [2], and antidepressant properties [3]. The chemistry of succinimides is fascinating and is receiving much attention. Particularly, α-alkylidene succinimides bearing multiple electron-withdrawing groups and nucleophilic and electrophilic sites are versatile building blocks in diversified organic synthesis (Scheme 1A). Unsurprisingly, α-alkylidene succinimides could be used as Michael acceptors to afford enantioenriched succinimide derivatives (Scheme 1B) [4,5]. Notably, as shown in Scheme 1C, α-alkylidene succinimides have also been successfully employed not only as C2-synthons, opening robust access to spiropyrrolizidines [6,7,8,9,10,11,12,13] and pyrones [14], but also as C3-synthons, making way for complex fused rings [15,16,17,18]. However, in stark contrast, reports on the catalytic nucleophilic alkylation of α-alkylidene succinimides (Scheme 1D) that result in the formation of either maleimides or succinimide derivatives are very limited. Up to now, only a handful of reports have been disclosed on the organocatalytic enantioselective nucleophilic alkylation of α-alkylidene succinimides. In 2010, Jiang and Tan et al. reported a bicyclic guanidine-catalyzed Mannich-type allylic addition reaction of N-aryl alkylidene-succinimides with imines (Scheme 1E) [19]. Importantly, the reactions of both 2-methylidene-N-aryl succinimides and N-aryl esterlidene-succinimides led to the formation of enantioenriched maleimides. Furthermore, the reaction of N-aryl benzylidene-succinimides enabled the formation of optically active succinimides. In 2016, Du’s group presented a bifunctional squaramide-catalyzed asymmetric Michael addition reaction of α-alkylidene succinimides with nitroalkenes to afford chiral functionalized succinimides (Scheme 1F) [20]. Differently, as established by Wang and Yuan et al. in 2020, a bifunctional thiourea-catalyzed asymmetric Michael addition of benzylidene succinimides to β-trifluoromethyl enones followed by a 1,3-proton shift furnished a series of F3C-containing chiral Rauhut–Currier-type products (Scheme 1G) [21]. Recently, Wan and Wang et al. developed a cinchona thiourea-catalyzed enantioselective Mannich reaction between benzothiazolimines and α-benzylidene succinimides for the synthesis of chiral benzothiazol succinimides (Scheme 1H) [22].
Notably, Morita–Baylis–Hillman (MBH) carbonates were successfully employed as versatile allylic alkylation reagents to undergo asymmetric allylic substitutions with a wide range of nucleophiles with the aid of a chiral nucleophilic catalyst [23]. In 2012, Huang, Tan, and Jiang et al. reported a hydroquinidine-catalyzed allylic alkylation of Morita–Baylis–Hillman (MBH) carbonates with N-itaconimides as nucleophiles to afford a variety of multifunctional chiral α-methylene-β-maleimide esters (Scheme 1I) [24]. However, over the past decade, the organocatalytic enantioselective nucleophilic alkylation of α-alkylidene succinimides with MBH carbonates as allylic alkylation reagents has remained silent [25,26]. Notably, as shown in Scheme 1J, we successfully established an organocatalytic enantioselective [1 + 4]-annulation of MBH carbonates with a series of electron-deficient alkenes, such as β,γ-unsaturated α-keto esters and chalcones [27], ortho-quinone methides [28], α,β-unsaturated imines [29], 2-enoylpyridine N-oxides [30], and modified enones [31]. Particularly, we realized an organocatalytic regio- and enantioselective allylic alkylation of indolin-2-imines with MBH carbonates [32]. To fill this gap and as part of our ongoing interest in the field of organocatalytic asymmetric transformation of MBH carbonates, here we disclose a chiral phosphine-catalyzed allylic alkylation of α-benzylidene succinimides with MBH carbonates (Scheme 1K). Different from the previous work, this strategy features nucleophilic phosphine catalysis and mild conditions without additives, and it works well over a broad substrate scope to furnish succinimide derivatives bearing contiguous chiral tertiary carbon centers in high yields with high asymmetric induction.

2. Results and Discussion

We started our investigation with the model reaction between 3-benzylidene-1-phenylpyrrolidine-2,5-dione 1a and MBH carbonate 2a in CH2Cl2 at room temperature for 16 h (Table 1). Initially, the reaction proceeded smoothly in the presence of phosphine C1 to furnish the desired succinimide 3aa at a 76% yield with 29% ee and 10:1 dr (Table 1, entry 1). To improve the efficiency and stereoselectivity, the chiral nucleophilic phosphine catalyst was carefully screened (Table 1, entries 2–8). Pleasingly, the C3-catalyzed reaction afforded the desired product 3aa at an 85% yield with 56% ee and 10:1 dr (Table 1, entry 3). The use of C4 as a catalyst enabled the formation of product 3aa at a 91% yield with 62% ee and 8:1 dr (Table 1, entry 4). Particularly, the desired product 3aa was obtained at a 95% yield with 85% ee and 11:1 dr when phosphine C6 was employed (Table 1, entry 6). Further modifying the catalyst structure did not enhance the efficiency or the asymmetric induction (Table 1, entries 7 and 8). The effect of the substituent on the nitrogen atom of α-alkylidene succinimides was also surveyed. The C6-catalyzed reaction of 3-benzylidene-1-methylpyrrolidine-2,5-dione 1b furnished the corresponding product 3ba at a 94% yield with 77% ee and 9:1 dr (Table 1, entry 9). Notably, product 3ca was obtained at a 90% yield with 92% ee and 10:1 dr from the C6-catalyzed reaction of 1-benzyl-3-benzylidenepyrrolidine-2,5-dione 1c (Table 1, entry 10). With these encouraging data in hand, we then further optimized reaction conditions to obtain better results. The screening of reaction media disclosed that CH2Cl2 was suitable (Table 1, entries 11–14). The concentration was found to have a large effect on the reaction (Table 1, entries 15–18), enabling the formation of product 3ca at a 92% yield with 92% ee and 11:1 dr (Table 1, entry 17). Neither increasing nor lowing the temperature could achieve better results (Table 1, entries 19–20). Prolonging reaction time to 36 h, the yield was increased to 96% without compromising stereoselectivity (Table 1, entry 21). However, further prolonging reaction time decreased the enantioselectivity (Table 1, entry 22). As a result, we identified the optimal reaction conditions: when 1-benzyl-3-benzylidenepyrrolidine-2,5-dione 1c (0.05 mmol) was treated with MBH carbonate 2a (0.06 mmol) in the presence of catalyst C6 (10 mol%) in CH2Cl2 (0.75 mL) at room temperature for 36 h, the desired succinimide 3ca was obtained at a 96% yield with 92% ee and 11:1 dr.
With the optimized reaction condition in hand, we then examined the substrate scope. As shown in Scheme 2, the scope of α-benzylidene succinimide 1 was investigated with the C6-catalyzed reaction of MBH carbonate 2a. Succinimides with different substitute groups (R1) on the nitrogen atom were tested. Under standard conditions, product 3aa was obtained at a 96% yield with 81% ee and 11:1 dr, and 3ba was obtained at a 91% yield with 81% ee and 5:1 dr, respectively. Furthermore, the succinimide with n-Bu group 3da was isolated at a 91% yield with 82% ee and 11:1 dr. In particular, the C6-catalyzed reaction of 3-benzylidene-1-(t-butyl)pyrrolidine-2,5-dione 1e generated the desired product 3ea at a 94% yield with 92% ee and 14:1 dr. Moreover, the α-benzylidene succinimide with naphthalen-1-ylmethyl residue 1f was also compatible to afford the corresponding product 3fa at a 91% yield with 81% ee and 13:1 dr. The effect of aromatic group Ar1 was also surveyed. In general, a wide range of α-benzylidene succinimides 1gn reacted smoothly with MBH carbonate 2a under standard conditions to give the corresponding succinimides 3gana at high yields (86–94%) and stereoselectivities (76–98% ee, 9:1–14:1 dr). Both electron-donating (Me, MeS, and MeO) and electron-withdrawing (F, Cl, and Br) groups could be introduced into the aromatic ring of residue Ar1 with a slight effect on the efficiency and asymmetric induction. With one exception, the C6-catalyzed reaction of 1-benzyl-3-(2-methoxybenzylidene)pyrrolidine-2,5-dione 1k resulted in the formation of product 3ka at a 94% yield with 28% ee and 3:1 dr. It was found that 1-benzyl-3-(naphthalen-2-ylmethylene)pyrrolidine-2,5-dione 1o was also tolerated to give the desired product 3oa at a 93% yield with 85% ee and 11:1 dr.
Subsequently, the scope of MBH carbonates was also investigated with the C6-catalyzed reaction of α-benzylidene succinimide 1e (Scheme 3). To our delight, various MBH carbonates 2bf with different substituents (Ar2) were found to be compatible under standard conditions to afford the corresponding succinimides 3ebef at a 90–94% yield with 92–95% ee and 10:1–15:1 dr. No significant electronic effect on the aromatic moiety was observed. Particularly, the MBH carbonate bearing thiophen-2-yl group 2g reacted smoothly with α-benzylidene succinimide 1e to afford the desired product 3eg at a 91% yield with 92% ee and >20:1 dr. Moreover, the ester groups of MBH carbonate 2 had a slight effect on the efficiency and stereoselectivity, furnishing products 3ehek at an 88–94% yield with 91–93% ee and 7:1–15:1 dr. Taken together, these encouraging results indicated that the chiral phosphine-catalyzed allylic alkylation of α-benzylidene succinimides with MBH carbonates was achieved, furnishing succinimide derivatives bearing contiguous chiral tertiary carbon centers at high yields with asymmetric induction.
To demonstrate the synthetic potential, the C6-catalyzed reaction was scaled up to 0.5 mmol of the starting material under standard reaction conditions. The corresponding product 3ea was obtained at a 92% yield with 94% ee and 14:1 dr (Scheme 4A). The absolute configuration of 3la was unambiguously confirmed by X-ray crystallography (CCDC 2244711 (3la) 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. For details concerning the crystal structure of 3la see the Supplementary Materials as well). The stereochemistry of other products was assumed by analogy. On the basis of the reported results and our previous work [33], a proposed reaction mechanism was shown in Scheme 4B. The nucleophilic attack of organocatalyst C6 to MBH carbonate 2a formed the intermediate IM-I and released the basic t-BuO ion to deprotonate α-benzylidene succinimide 1e to afford the intermediate IM-II. The intermediate IM-II could isomerize into the intermediate IM-III. Via IMIV, the reaction of intermediate IM-I with intermediate IM-III led to the formation of the desired product 3ea and regenerated the catalyst C6.

3. Materials and Methods

All chemicals were used without further purification as commercially available unless otherwise noted. Thin-layer chromatography (TLC) was performed on silica gel plates (60F-254) using UV light (254 and 365 nm). Flash chromatography was conducted on silica gel (300–400 mesh). NMR (400, 500, or 600 MHz for 1H NMR; 100 or 126 MHz for 13C NMR; 376 MHz for 19F NMR) spectra were recorded in CDCl3 with TMS as the internal standard. Chemical shifts are reported in ppm, and coupling constants are given in Hz. Data for 1H NMR are recorded as follows: chemical shift (ppm), multiplicity (s, singlet; d, doublet; t, triplet; q, quartet; m, multiplet; dd, doublet–doublet), coupling constant (Hz), and integration. Data for 13C NMR are reported in terms of chemical shift (δ, ppm). Data for 19F NMR are reported in terms of chemical shift (δ, ppm). High-resolution mass spectral (HRMS) analyses were recorded on a Thermo Scientific Q Exactive Orbitrap mass spectrometer (Bremen, Germany) with ESI source. The crystal structure and data were recorded on a Rigaku HomeLab diffractometer. More details can be found in the Supplementary Materials. In addition, the X-ray of 3la, copies of NMR, and chiral HPLC analysis can be found in the Supplementary Materials.

3.1. General Procedure for the Synthesis of α-Benzylidene Succinimide 1

Triphenylphosphine (10.5 mmol) was added to a solution of substituted 1-R1-1H-pyrrole-2,5-dione (10 mmol) and aldehyde (11 mmol) in EtOH (100 mL) at room temperature. The reaction mixture was stirred at room temperature overnight. When the reaction was completed (monitored by TLC), the reaction mixture was filtered, and the precipitation was washed with ethanol and dried to afford α-benzylidene succinimide 1.

3.2. General Procedure for the Synthesis of MBH Carbonate 2

1,4-Diazabicyclo[2.2.2]octane (DABCO, 10.5 mmol) was added to a solution of aldehyde (10 mmol) in acrylate (20 mL) at room temperature. The reaction mixture was stirred at room temperature for 3–7 days. When the reaction was completed (monitored by TLC), the reaction mixture was purified by silica gel column chromatography to afford MBH alcohol.
4-(Dimethylamino)pyridine (DMAP, 2.08 mmol) was added to a solution of MBH alcohol and Boc-anhydride (15 mmol) in CH2Cl2 (30 mL) in batches. When the reaction was complete (monitored by TLC), the organic phase was washed with distilled water (2 × 20 mL) and dried over anhydrous Na2SO4, and the solvent was removed under reduced pressure. The residue was purified by silica gel column chromatography, affording MBH carbonate 2.

3.3. General Procedure for the Phosphine-Catalyzed Allylic Alkylation

α-Benzylidene succinimide 1 (0.05 mmol), MBH carbonate 2 (0.06 mmol), and phosphine C6 (10 mol%) were added to a solution of CH2Cl2 (0.75 mL). The reaction mixture was stirred at room temperature for 36 h. After the removal of the solvent, the crude residue was purified by preparative TLC (petroleum/ethyl acetate = 2:1) to obtain the desired product 3.

3.4. Scale-Up of the Allylic Alkylation

MBH carbonate 2a (175 mg, 0.6 mmol) and catalyst C6 (25 mg, 0.05 mmol) were added to a solution of 3-benzylidene-1-(tert-butyl)pyrrolidine-2,5-dione 1e (122 mg, 0.5 mmol) in CH2Cl2 (7.5 mL). The reaction mixture was stirred at room temperature for 36 h. Then, the mixture was purified by silica gel column chromatography (eluent: petroleum ether/EtOAc = 2:1) to yield the desired product 3ea at a 92% yield (196 mg, 94% ee, 14:1 dr).

4. Conclusions

In conclusion, we developed an organocatalytic enantioselective allylic alkylation of α-benzylidene succinimides with MBH carbonates. With the aid of a chiral nucleophilic phosphine, a broad scope of α-benzylidene succinimides reacted smoothly with MBH carbonate to furnish functionalized α-benzylidene succinimides at high yields with high stereoselectivities. Importantly, this synthetic strategy not only achieved nucleophilic phosphine catalysis but also realized the asymmetric construction of enantioenriched succinimide derivatives bearing contiguous chiral tertiary carbon centers.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28062825/s1.

Author Contributions

J.S. and P.L. were responsible for conceptualization, data validation, and writing—review and editing; C.L. was responsible for the catalysis experiments and the spectroscopic and analytical analysis. All authors have read and agreed to the published version of the manuscript.

Funding

The authors acknowledge the financial support from the National Natural Science Foundation of China (21871128), Guangdong Innovative Program (2019BT02Y335).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in the Supplementary Materials.

Acknowledgments

We gratefully thank the assistance of SUSTech Core Research Facilities, Yang Yu (HRMS, SUSTech), and Xiaoyong Chang (X-ray, SUSTech).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fredenhagen, A.; Tamura, S.Y.; Kenny, P.T.M.; Komura, H.; Naya, Y.; Nakanishi, K.; Nishiyama, K.; Sugiura, M.; Kita, H. Andrimid, a new peptide antibiotic produced by an intracellular bacterial symbiont isolated from a brown planthopper. J. Am. Chem. Soc. 1987, 109, 4409–4411. [Google Scholar] [CrossRef]
  2. Isaka, M.; Rugseree, N.; Maithip, P.; Kongsaeree, P.; Prabpai, S.; Thebtaranonth, Y. Hirsutellones A–E, antimycobacterial alkaloids from the insect pathogenic fungus Hirsutella nivea BCC 2594. Tetrahedron 2005, 61, 5577–5583. [Google Scholar] [CrossRef]
  3. Wróbel, M.Z.; Chodkowski, A.; Herold, F.; Gomoółka, A.; Kleps, J.; Mazurek, A.P.; Plucinński, F.; Mazurek, A.; Nowak, G.; Siwek, A.; et al. Synthesis and biological evaluation of novel pyrrolidine-2,5-dione derivatives as potential antidepressant agents. Part 1. Eur. J. Med. Chem. 2013, 63, 484–500. [Google Scholar] [CrossRef]
  4. Lin, S.; Leow, D.; Huang, K.-W.; Tan, C.-H. Enantioselective protonation of itaconimides with thiols and the rotational kinetics of the axially chiral C-N bond. Chem. Asian J. 2009, 4, 1741–1744. [Google Scholar] [CrossRef]
  5. Liu, Y.; Zhang, W. Iridium-catalyzed asymmetric hydrogenation of α-alkylidene succinimides. Angew. Chem. Int. Ed. 2013, 52, 2203–2206. [Google Scholar] [CrossRef] [PubMed]
  6. Haddad, S.; Boudriga, S.; Porzio, F.; Soldera, A.; Askri, M.; Sriram, D.; Yogeeswari, P.; Knorr, M.; Rousselin, Y.; Kubicki, M.M. Synthesis of novel dispiropyrrolothiazoles by three component 1,3-dipolar cycloaddition and evaluation of their antimycobacterial activity. RSC Adv. 2014, 4, 59462–59471. [Google Scholar] [CrossRef]
  7. Haddad, S.; Boudriga, S.; Akhaja, T.N.; Raval, J.P.; Porzio, F.; Soldera, A.; Askri, M.; Knorr, M.; Rousselin, Y.; Kubicki, M.M.; et al. A strategic approach to the synthesis of functionalized spirooxindole pyrrolidine derivatives: In Vitro antibacterial, antifungal, antimalarial and antitubercular studies. New J. Chem. 2015, 39, 520–528. [Google Scholar] [CrossRef]
  8. Haddad, S.; Boudriga, S.; Porzio, F.; Soldera, A.; Askri, M.; Knorr, M.; Rousselin, Y.; Kubicki, M.M.; Golz, C.; Strohmann, C. Regio- and stereoselective synthesis of spiropyrrolizidines and piperazines through azomethine ylide cycloaddition reaction. J. Org. Chem. 2015, 80, 9064–9075. [Google Scholar] [CrossRef]
  9. Yang, W.-L.; Liu, Y.-Z.; Luo, S.; Yu, X.; Fossey, J.S.; Deng, W.-P. The copper-catalyzed asymmetric construction of a dispiropyrrolidine skeleton via 1,3-dipolar cycloaddition of azomethine ylides to α-alkylidene succinimides. Chem. Commun. 2015, 51, 9212–9215. [Google Scholar] [CrossRef]
  10. Luo, W.; Hu, H.; Nian, S.; Qi, L.; Ling, F.; Zhong, W. Phosphine-catalyzed [3 + 2] annulation reaction: Highly regio- and diastereoselective synthesis of 2-azaspiro[4.4]nonene-1,3-diones. Org. Biomol. Chem. 2017, 15, 7523–7526. [Google Scholar] [CrossRef] [Green Version]
  11. Boudriga, S.; Haddad, S.; Askri, M.; Soldera, A.; Knorr, M.; Strohmann, C.; Golz, C. Highly diastereoselective construction of novel dispiropyrrolo[2,1-a]isoquinoline derivatives via multicomponent 1,3-dipolar cycloaddition of cyclic diketones-based tetrahydroisoquinolinium N-ylides. RSC Adv. 2019, 9, 11082–11091. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Gan, Z.; Zhi, M.; Han, R.; Li, E.-Q.; Duan, Z.; Mathey, F. P-Stereogenic phosphines directed copper(I)-catalyzed enantioselective 1,3-dipolar cycloadditions. Org. Lett. 2019, 21, 2782–2785. [Google Scholar] [CrossRef] [PubMed]
  13. Zhao, M.-X.; Liu, Q.; Yu, K.-M.; Zhao, X.-L.; Shi, M. Organocatalyzed asymmetric formal [3+2] cycloaddition of isocyanoacetates with N-itaconimides: Facile access to optically active spiropyrroline succinimide derivatives. Org. Chem. Front. 2019, 6, 3879–3884. [Google Scholar] [CrossRef]
  14. Yuan, C.; Yang, H.; Gong, Q.; Luo, S.; Gu, J.; Ge, Y.; Guo, H.; Xie, L. Brønsted base-catalyzed tandem [2+4] annulation/tautomerization/aromatization reaction of α-alkylidene succinimides with 5-alkenyl thiazolones. Adv. Synth. Catal. 2021, 363, 3336–3347. [Google Scholar] [CrossRef]
  15. Zhao, B.-L.; Du, D.-M. Organocatalytic cascade Michael/Michael reaction for the asymmetric synthesis of spirooxindoles containing five contiguous stereocenters. Chem. Commun. 2016, 52, 6162–6165. [Google Scholar] [CrossRef]
  16. Tehri, P.; Peddinti, R.K. DBU-catalyzed [3+2] cycloaddition and Michael addition reactions of 3-benzylidene succinimides with 3-ylidene oxindoles and chalcones. Org. Biomol. Chem. 2019, 17, 3964–3970. [Google Scholar] [CrossRef]
  17. Li, C.-Y.; Xiang, M.; Song, X.-J.; Zou, Y.; Huang, Z.-C.; Li, X.; Tian, F.; Wang, L.-X. A base-catalyzed domino reaction between isoindigos and α-alkylidene succinimides—Convenient preparation of highly steric bispirooxindoles. Org. Biomol. Chem. 2020, 18, 9511–9515. [Google Scholar] [CrossRef]
  18. Zhang, X.-Y.; Dou, P.-H.; Lu, W.-Y.; You, Y.; Zhao, J.-Q.; Wang, Z.-H.; Yuan, W.-C. Benzylidene succinimides as 3C synthons for the asymmetric tandem Mannich reaction/ transamidation of cyclic trifluoromethyl ketimines to obtain F3C-containing polycyclic dihydroquinazolinones. Chem. Commun. 2021, 57, 2927–2930. [Google Scholar] [CrossRef]
  19. Wang, J.; Liu, H.; Fan, Y.; Yang, Y.; Jiang, Z.; Tan, C.-H. Bicyclic guanidine-catalyzed direct asymmetric allylic addition of N-aryl alkylidene-succinimides. Chem. Eur. J. 2010, 16, 12534–12537. [Google Scholar] [CrossRef]
  20. Zhao, B.-L.; Zhang, D.; Liu, L.; Du, D.-M. Organocatalytic asymmetric Michael addition of α-alkylidene succinimides to nitrostyrenes. Org. Biomol. Chem. 2016, 14, 6337–6345. [Google Scholar] [CrossRef]
  21. Zhang, X.-Y.; You, Y.; Wang, Z.-H.; Zhao, J.-Q.; Yuan, W.-C. Organocatalytic asymmetric tandem α-functionalization/1,3-proton shift reaction of benzylidene succinimides with β-trifluoromethyl enones. Chem. Commun. 2020, 56, 13449–13452. [Google Scholar] [CrossRef] [PubMed]
  22. Li, C.-Y.; Xiang, M.; Zhang, J.; Li, W.-S.; Zou, Y.; Wan, W.-J.; Wang, L.-X. Enantioselective organocatalyzed Mannich reaction between benzothiazolimines and α-benzylidene succinimides for the preparation of chiral benzothiazol succinimides. Asian J. Org. Chem. 2022, 11, e202200107. [Google Scholar] [CrossRef]
  23. Liu, T.-Y.; Xie, M.; Chen, Y.-C. Organocatalytic asymmetric transformations of modified Morita-Baylis-Hillman adducts. Chem. Soc. Rev. 2012, 41, 4101–4112. [Google Scholar] [CrossRef] [PubMed]
  24. Yang, W.; Tan, D.; Li, L.; Han, Z.; Yan, L.; Huang, K.-W.; Tan, C.-H.; Jiang, Z. Direct asymmetric allylic alkenylation of N-itaconimides with Morita-Baylis-Hillman carbonates. J. Org. Chem. 2012, 77, 6600–6607. [Google Scholar] [CrossRef]
  25. Kamlar, M.; Hybelbauerová, S.; Císařová, I.; Veselý, J. Organocatalytic enantioselective allylic alkylation of MBH carbonates with β-keto esters. Org. Biomol. Chem. 2014, 12, 5071–5076. [Google Scholar] [CrossRef] [PubMed]
  26. Ma, G.; Sibi, M.P. Enantioselective allylic amination of MBH carbonates catalyzed by novel chiral 4-dialkylaminopyridine catalysts. Org. Chem. Front. 2014, 1, 1152–1156. [Google Scholar] [CrossRef]
  27. Cheng, Y.; Han, Y.; Li, P. Organocatalytic enantioselective [1+4] annulation of Morita-Baylis-Hillman carbonates with electron-deficient olefins: Access to chiral 2,3-dihydrofuran derivatives. Org. Lett. 2017, 19, 4774–4777. [Google Scholar] [CrossRef]
  28. Cheng, Y.; Fang, Z.; Li, W.; Li, P. Phosphine-mediated enantioselective [4+1]-annulations between ortho-quinone methides and Morita-Baylis-Hillman carbonates. Org. Chem. Front. 2018, 5, 2728–2733. [Google Scholar] [CrossRef]
  29. Qian, C.; Zhang, P.; Li, W.; Li, P. Phosphine-catalyzed enantioselective [1+4]-annulation of Morita-Baylis-Hillman carbonates with α,β-unsaturated imines. Asian J. Org. Chem. 2019, 8, 242–245. [Google Scholar]
  30. Zhang, P.; Guo, X.; Liu, C.; Li, W.; Li, P. Enantioselective construction of pyridine N-oxides featuring 2,3-dihydrofuran motifs via phosphine-catalyzed [4+1]-annulations of 2-enoylpyridine N-oxides with Morita-Baylis-Hillman carbonates. Org. Lett. 2019, 21, 152–155. [Google Scholar] [CrossRef]
  31. Guo, X.; Shen, B.; Liu, C.; Zhao, H.; Li, X.; Yu, P.; Li, P. Rational design and organocatalytic enantioselective [1+4]-annulations of MBH carbonates with modified enones. Org. Chem. Front. 2023, 10, 150–156. [Google Scholar] [CrossRef]
  32. Chen, L.; Li, P. Organocatalytic regio- and enantioselective allylic alkylation of indolin-2-imines with MBH carbonates toward 3-allylindoles. J. Org. Chem. 2023. [Google Scholar] [CrossRef] [PubMed]
  33. Li, H.; Shi, W.; Wang, C.; Liu, H.; Wang, W.; Wu, Y.; Guo, H. Phosphine-catalyzed cascade annulation of MBH carbonates and diazenes: Synthesis of hexahydrocyclopenta[c]pyrazole derivatives. Org. Lett. 2021, 23, 5571–5575. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Chemistry of α-alkylidene succinimides.
Scheme 1. Chemistry of α-alkylidene succinimides.
Molecules 28 02825 sch001
Scheme 2. Scope of benzylidene succinimides.
Scheme 2. Scope of benzylidene succinimides.
Molecules 28 02825 sch002
Scheme 3. Scope of MBH carbonates.
Scheme 3. Scope of MBH carbonates.
Molecules 28 02825 sch003
Scheme 4. Scale-up of reaction and proposed mechanism.
Scheme 4. Scale-up of reaction and proposed mechanism.
Molecules 28 02825 sch004
Table 1. Condition optimization.
Table 1. Condition optimization.
Molecules 28 02825 i001
Entry aCR1SolventT/°CTime/hYield/% bee/% cdr d
1C1PhCH2Cl2rt163aa, 762910:1
2C2PhCH2Cl2rt163aa, 92611:1
3C3PhCH2Cl2rt163aa, 855610:1
4C4PhCH2Cl2rt163aa, 91628:1
5C5PhCH2Cl2rt163aa, 93912:1
6C6PhCH2Cl2rt163aa, 958511:1
7C7PhCH2Cl2rt163aa, 937910:1
8C8PhCH2Cl2rt163aa, 78879:1
9C6MeCH2Cl2rt163ba, 94779:1
10C6BnCH2Cl2rt163ca, 909210:1
11C6BnClCH2CH2Clrt163ca, 869311:1
12C6BnCHCl3rt163ca, 79809:1
13C6BnEtOAcrt163ca, 828811:1
14C6BnPhClrt163ca, 828112:1
15C6BnCH2Cl2 (0.25 mL)rt163ca, 689210:1
16C6BnCH2Cl2 (0.50 mL)rt163ca, 829110:1
17C6BnCH2Cl2 (0.75 mL)rt163ca, 929211:1
18C6BnCH2Cl2 (1.25 mL)rt163ca, 909011:1
19C6BnCH2Cl2 (0.75 mL)30163ca, 928810:1
20C6BnCH2Cl2 (0.75 mL)0163ca, 569512:1
21C6BnCH2Cl2 (0.75 mL)rt363ca, 969211:1
22C6BnCH2Cl2 (0.75 mL)rt483ca, 968711:1
a Unless noted, a mixture of 1 (0.05 mmol), 2a (0.06 mmol), and C (10 mol%) in the solvent (1.0 mL) was stirred at room temperature (rt) for the time given. b Isolated yield. c Determined by chiral-HPLC analysis. d For major enantiomers, determined by 1H NMR.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Liu, C.; Sun, J.; Li, P. Chiral Phosphine Catalyzed Allylic Alkylation of Benzylidene Succinimides with Morita–Baylis–Hillman Carbonates. Molecules 2023, 28, 2825. https://doi.org/10.3390/molecules28062825

AMA Style

Liu C, Sun J, Li P. Chiral Phosphine Catalyzed Allylic Alkylation of Benzylidene Succinimides with Morita–Baylis–Hillman Carbonates. Molecules. 2023; 28(6):2825. https://doi.org/10.3390/molecules28062825

Chicago/Turabian Style

Liu, Chang, Jianwei Sun, and Pengfei Li. 2023. "Chiral Phosphine Catalyzed Allylic Alkylation of Benzylidene Succinimides with Morita–Baylis–Hillman Carbonates" Molecules 28, no. 6: 2825. https://doi.org/10.3390/molecules28062825

Article Metrics

Back to TopTop