DOI:
10.1039/D4QO01424C
(Research Article)
Org. Chem. Front., 2024,
11, 6666-6671
Asymmetric synthesis of bis-spiro cyclopropane skeletons via bifunctional phosphonium salt-catalyzed [2 + 1] annulation†
Received
2nd August 2024
, Accepted 18th September 2024
First published on 23rd September 2024
Abstract
A new approach for constructing enantiopure spiro[cyclopropane-oxindole] and bispiro[oxindole-cyclopropane-cyclohexone] skeletons featuring three vicinal stereocenters was developed. In this approach, 3-alkenyl-oxindoles and α-bromoketones served as substrates for an asymmetric [2 + 1] cyclopropanation using a chiral bifunctional phosphonium salt catalyst. The reaction afforded the desired products in high yields (up to 97%) and with excellent stereoselectivities (up to 97% ee and >20:1 dr).
Introduction
The cyclopropane ring, a unique motif among carbocycles, has garnered significant scientific interest within the organic chemistry community due to its distinctive bonding and inherent ring strain.1,2 It is also prevalent in biologically active compounds and natural products.3 Spirocyclopropanes represent a significant class of derivatives within the cyclopropane framework, showing considerable promise for drug discovery applications.4 For instance, CFI-400945 and its derivatives containing the spirocyclopropane-oxindole skeleton are pioneering selective inhibitors of polo-like kinase 4 (PLK4) in advanced solid tumors.5 Illudin M and Illudin S, which bear the spiro-cyclopropane-cyclohexanone motif, have been reported to exhibit preferential cytotoxicity in vitro against various human tumor cell lines (Scheme 1a).6 The frequent occurrence of chiral spirocyclopropyl ring systems in drug molecules has driven advancements in sophisticated synthetic methodologies to prepare highly enantiopure complex spirocyclopropane skeletons. Asymmetric catalytic synthesis of chiral monospirocyclopropanes, such as chiral spiro[cyclopropane-oxindoles], has been extensively documented,7 whereas reports on the synthesis of chiral bispirocyclopropanes remain relatively scarce.8 Du et al. reported squaramide-catalyzed asymmetric [2 + 1] cyclopropanation for synthesizing bispiro[oxindole-cyclopropane-thiazolone] and bispiro[oxindole-cyclopropane-pyrazolone] derivatives.8e,f Lei and co-workers developed the diastereo- and enantioselective synthesis of novel bispiro[indanedione-oxindole-cyclopropanes].8g In 2020, our group devised an efficient method for constructing optically active bispiro-cyclopropane-pyrazolone derivatives.8h However, existing reports primarily focus on five-membered heterocycles with fused bispirocyclopropane units, whereas reports on chiral bispirocyclopropane frameworks incorporating six-membered rings, which hold considerable promise for future drug development, are very limited (Scheme 1b).8a Therefore, developing an efficient method to prepare novel chiral bispirocyclopropane frameworks, especially those featuring a spiro six-membered ring, is highly desirable.
|
| Scheme 1 Asymmetric synthesis of (bis)spirocyclopropane skeletons and our strategy. | |
Bifunctional phosphonium salt (BPS) catalysts constitute a pivotal category in ion-pairing catalysis,9 demonstrating excellent catalytic activity in various asymmetric reactions since their inception.10 By harnessing this catalytic system, we have achieved substantial advancements in synthesizing diverse chiral heterocycles and bridged ring frameworks via annulation reactions.11 Our objective is to achieve the selective synthesis of six-membered rings containing bispirocyclopropane frameworks through systematic substrate design and optimal reaction conditions. Here, we describe a highly efficient formal [2 + 1] annulation reaction between 3-alkenyl-oxindoles and α-bromoketones, yielding multi-substituted spirocyclopropane and bispirocyclopropane frameworks in high yields with exceptional diastereo- and enantioselectivities (Scheme 1c). Furthermore, mechanistic studies have revealed that hydrogen bonding interactions between the catalyst and substrate critically influence stereocontrol.
Results and discussion
Initially, the model reaction employed N-Boc-3-alkenyl-oxindole 1a and α-bromoketone 2a in CHCl3 at room temperature with cesium carbonate using bifunctional phosphonium salts as catalysts (Table 1). We observed that BPS catalysts with different backbones efficiently facilitated the cyclopropanation reaction, yielding the desired [2 + 1] adducts exclusively. As detailed in Table 1, when L-Val-derived phosphonium iodides (P1–P4) featuring amide, thiourea, or dipeptide scaffolds were used as chiral organocatalysts, cycloaddition adduct 3a was obtained in good yields. Notably, bifunctional phosphonium salt P4, incorporating an amide moiety, emerged as the optimal scaffold for further enhancement of both diastereoselectivity and stereoselectivity (entries 1–4). Subsequent investigations focused on modifying the electronic and steric effects of amide-functionalized PPS catalysts (P5–P8) by varying amino acid residues or adjusting benzyl groups at the active P-center (entries 5–8). Ultimately, P8 was identified as the most effective organocatalyst, achieving outstanding results in this cyclization process with high efficiency (88% yield), excellent diastereoselectivity (92:8 dr), and good enantioselectivity (87% ee) at room temperature. Further optimization included evaluating the influence of solvents on reaction conditions (entries 9 and 10), revealing that CHCl3 provided superior catalytic performance to other solvents. Fine-tuning of other reaction parameters, such as base equivalents, chiral catalyst loading, and temperature, pinpointed the optimal conditions (entry 14: −20 °C, 10 mol% of chiral catalyst P8, and 1.5 equivalents of Cs2CO3 in CHCl3 for 24 hours), resulting in the best outcome (96% yield, 98:2 dr, and 90% ee; additional details can be found in Tables S1–S6 in the ESI†).
Table 1 Optimization of the reaction conditionsa
|
Entry |
Catalyst |
Solvent |
dr |
Yieldb (%) |
eemajor, eeminor (%) |
Reaction conditions: 1a (0.1 mmol), 2a (0.15 mmol), catalyst (10 mol%) and Cs2CO3 (4.0 equiv.) in solvent (1.5 mL) were stirred at room temperature for 24 h, and the dr values were determined by 1H NMR.
Isolated total yields, and the ee values were determined by chiral HPLC.
Cs2CO3 (1.5 equiv.).
Cs2CO3 (6.0 equiv.).
The catalyst loading was 5 mol%.
At −20 °C.
|
1 |
P1
|
CHCl3 |
85:15 |
85 |
38, 16 |
2 |
P2
|
CHCl3 |
88:12 |
87 |
11, 8 |
3 |
P3
|
CHCl3 |
15:85 |
73 |
6, 20 |
4 |
P4
|
CHCl3 |
90:10 |
85 |
55, 30 |
5 |
P5
|
CHCl3 |
67:33 |
81 |
58, 42 |
6 |
P6
|
CHCl3 |
75:25 |
80 |
58, 52 |
7 |
P7
|
CHCl3 |
61:39 |
88 |
75, 71 |
8 |
P8
|
CHCl3 |
92:8 |
88 |
86, 80 |
9 |
P8
|
Toluene |
88:12 |
85 |
86, 81 |
10 |
P8
|
Et2O |
87:13 |
72 |
68, 53 |
11c |
P8
|
CHCl3 |
94:6 |
90 |
86, 80 |
12d |
P8
|
CHCl3 |
50:50 |
69 |
80, 76 |
13c,e |
P8
|
CHCl3 |
97:3 |
84 |
82, — |
14c,f |
P8
|
CHCl3 |
98:2 |
96 |
90, — |
Having established the optimized conditions for the [2 + 1] cyclopropanation of 3-alkenyl-oxindoles and α-bromoketones, we explored the scope of the reaction to construct structurally diverse chiral (bis)spirocyclopropanes. Initially, we investigated the tolerance of 3-alkenyl-oxindoles 1 under the optimized conditions using α-bromoketone 2a as the cyclization partner (Table 2). Remarkably, substrates with electron-donating, neutral, or electron-withdrawing substituents on the phenyl rings of 3-alkenyl-oxindoles afforded the desired products 3a–3g in good yields with excellent diastereoselectivities (mostly >20:1 dr) and good to excellent enantioselectivities (81–95% ee). Subsequently, we explored the compatibility of non-cyclic α-bromoketone 2. As shown in Table 2, α-bromoketones bearing various substituted aryl groups and heteroaryl groups reacted under standard conditions, yielding the desired monospirocyclopropanes with moderate diastereoselectivities and enantioselectivities (3h–3l). Notably, replacing the oxindole's ester group with a benzoyl group resulted in the formation of spirocyclic product 3m with high diastereoselectivity (>20:1 dr) and moderate enantioselectivity (51% ee). The absolute configuration of product 3m was determined by comparing its optical rotation and retention time in HPLC with literature values, while other spiro products were assigned by analogy (see Tables S1–S3 in the ESI† for details).7e
Table 2 Synthesis of spiro[cyclopropane-oxindole] scaffoldsa,b
Standard conditions I: 1 (0.2 mmol), 2 (0.3 mmol), P8 (10 mol%) and Cs2CO3 (1.5 equiv.) in CHCl3 (3.0 mL) were stirred at −20 °C for 36 h.
Isolated yields; the dr values were determined by 1H NMR and the ee values were determined by chiral HPLC.
In toluene (3.0 mL) and at room temperature for 48 h.
At 0 °C.
|
|
Given the promising potential of novel spirocyclopropane scaffolds in drug development, we undertook the challenge of constructing bispiro[oxindole-cyclopropane-cyclohexanone] compounds featuring two quaternary centers using 3-alkenyl-oxindoles and cyclic α-bromoketones (Table 3). Gratifyingly, a series of target molecules bearing these novel bispirocyclopropane scaffolds (3n–3v) were readily synthesized under such a BPS catalysis system, consistently achieving excellent enantio- and diastereo-selectivities (up to 97% ee and >20:1 dr). Additionally, compounds incorporating heterocyclohexanone groups (3w–3x) could also be efficiently prepared with notable ee values. It is worth noting that the reaction can still proceed smoothly by changing the size of the ring system in cyclic bromoketone, affording novel bispiro compounds (3y and 3z) in moderate yields with 67–75% ee. The observed non-ideal diastereoselectivities in some cases may be attributed to significant steric hindrance around the cyclohexyl group. However, it is noteworthy that the diastereomers could be effectively separated to obtain optically pure products through straightforward column chromatography. Besides, when alkyl bromoketones are used as reaction partners, this cyclization reaction is unable to proceed for providing the expected products (3aa and 3ab). Moreover, to assess the synthetic applicability of this methodology, we conducted a scale-up reaction of 3c. Using substrates 1c and 2a under standard conditions, we obtained the desired product 3c in 91% yield (233 mg), with a diastereomeric ratio of 98:2 and 92% enantiomeric excess (ee). Furthermore, treatment of 3c with boron trifluoride diethyl ether facilitated a hydrolysis and reduction process, yielding compound 4 in a moderate isolated yield with a slight decrease in enantiomeric purity (Scheme 2a).
|
| Scheme 2 Synthetic applications and mechanistic studies. | |
Table 3 Synthesis of bis-spiro[cyclopropane-oxindole] scaffoldsa,b
Standard conditions II: 1 (0.2 mmol), 2 (0.3 mmol), P8 (20 mol%) and Cs2CO3 (1.5 equiv.) in toluene (3.0 mL) were stirred at room temperature for 48 h.
Isolated yields; the dr values were determined by 1H NMR and the ee values were determined by chiral HPLC.
|
|
Building on our prior investigations into bi-/multifunctional phosphonium salt-catalyzed transformations, it is evident that hydrogen-bonding interactions, coupled with ion-pair activation, significantly contribute to asymmetric induction (Scheme 2b). Consequently, catalysts P8-1, featuring an amide-NH group protected by a methyl group, were synthesized and employed in the catalytic reaction under standard conditions, replacing P8. Predictably, both the yield and enantiomeric excess (ee) dropped markedly with P8-1 (entries 1 and 2). Of note, substituting polar methanol for CHCl3 as the solvent resulted in no product formation. Furthermore, replacing the Boc group on the indole with a methyl group led to a dramatic reduction in enantioselectivity in the catalytic reaction, indicating that the steric effect of the protecting group on the indole moiety positively influences stereocontrol (entries 3 and 4). Accordingly, a proposed transition state model, crucial for understanding the asymmetric annulation process, was developed based on these comprehensive mechanistic experiments (Scheme 2c).
Conclusions
In conclusion, we have developed an efficient chiral bifunctional phosphonium salt-catalyzed asymmetric [2 + 1] annulation of 3-alkenyl-oxindoles with two types of α-bromoketones, yielding a series of multi-substituted spirocyclopropyloxindoles and bispiro[oxindole-cyclopropane-cyclohexanones] with excellent stereoselectivities and high diastereoselectivities (up to 97% ee and up to 99:1 dr). These novel bispirocyclopropane scaffolds are expected to facilitate the discovery and development of new chiral drugs. Further exploration of the biological activities of these products is currently underway in our laboratory.
Data availability
The data supporting this article have been included as part of the ESI.†
All data included in this study are available upon request by contact with the corresponding author (i.e. Prof. Tianli Wang, email: wangtl@scu.edu.cn).
Conflicts of interest
There are no conflicts to declare.
Acknowledgements
Financial support was provided by the National Natural Science Foundation of China (22222109, 21921002 and 22101189), the Science and Technology Strategic Cooperation Project of Luzhou Municipal People's Government-Southwest Medical University (2020LZXNYDJ49), the National Key R&D Program of China (2018YFA0903500), the Beijing National Laboratory for Molecular Sciences (BNLMS202101), the Sichuan Science Foundation for Distinguished Young Scholars (2023NSFSC1921), the Sichuan Provincial Natural Science Foundation (2022NSFSC1181 and 24NSFSC6590), Fundamental Research Funds from Sichuan University (2020SCUNL108) and Fundamental Research Funds for the Central Universities.
References
- L. A. Wssjohann and W. Brandt, Biosynthesis and metabolism of cyclopropane rings in natural compounds, Chem. Rev., 2003, 103, 1625–1647 CrossRef PubMed.
-
O. G. Kulinkovich, in Cyclopropanes in Organic Synthesis, John Wiley & Sons, Hoboken NJ, 2015 Search PubMed.
- For selected reviews:
(a) J. Salaun, Optically active cyclopropanes, Chem. Rev., 1989, 89, 1247–1270 CrossRef CAS;
(b) A. H. Li and L. X. Dai, Asymmetric ylide reactions: epoxidation, cyclopropanation, aziridination, olefination, and rearrangement, Chem. Rev., 1997, 97, 2341–2372 CrossRef CAS PubMed;
(c)
A. Pfaltz, in Comprehensive Asymmetric Catalysis II, ed. E. N. Jacobsen, A. Pfaltz and H. Yammamoto, Springer, Berlin, 1999, pp. 513–538 Search PubMed;
(d) J. Xie and G. Dong, Cyclopropylcarbinyl cation chemistry in synthetic method development and natural product synthesis: cyclopropane formation and skeletal rearrangement, Org. Chem. Front., 2023, 10, 2346–2358 RSC;
(e) R. Moorthy, W. Bio-Sawe, S. S. Thorat and M. P. Sibi, Unveiling the beauty of cyclopropane formation: a comprehensive survey of enantioselective Michael initiated ring closure (MIRC) reactions, Org. Chem. Front., 2024, 11, 4560–4601 RSC.
- V. P. Sandanayaka, A. S. Prashad, Y. Yang, R. T. Williamson, Y. I. Lin and T. S. Mansour, Spirocyclopropyl β-Lactams as Mechanism-Based Inhibitors of Serine β-Lactamases. Synthesis by Rhodium-Catalyzed Cyclopropanation of 6-Diazopenicillanate Sulfone, J. Med. Chem., 2003, 46, 2569–2571 CrossRef CAS PubMed.
- P. B. Sampson, Y. Liu, N. K. Patel, M. Feher, B. Forrest, S. W. Li, L. Edwards, R. Laufer, Y. Lang, F. Ban, D. E. Awrey, G. Mao, O. Plotnikova, G. Leung, R. Hodgson, J. M. Mason, X. Wei, R. Kiarash, E. Green, W. Qiu, N. Y. Chirgadze, T. W. Mak, G. Pan and H. W. Pauls, The discovery of Polo-like Kinase 4 inhibitors: design and optimization of spiro[cyclopropane-1,3′[3H]indol]-2′(1′H)-ones as orally bioavailable antitumor agents, J. Med. Chem., 2015, 58, 130–146 CrossRef CAS PubMed.
- L. K. Smith and I. R. Baxendale, Total syntheses of natural products containing spirocarbocycles, Org. Biomol. Chem., 2015, 13, 9907–9933 RSC.
- For selected examples of the asymmetric synthesis of monospirocyclopropane:
(a) H. Lebel, J. F. Marcoux, C. Molinaro and A. B. Charette, Stereoselective cyclopropanation reactions, Chem. Rev., 2003, 103, 977–1050 CrossRef CAS PubMed;
(b) Z. Y. Cao, X. Wang, C. Tan, X. L. Zhao, J. Zhou and K. Ding, Highly stereoselective olefin cyclopropanation of diazooxindoles catalyzed by a C2-symmetric spiroketal bisphosphine/Au(I) complex, J. Am. Chem. Soc., 2013, 135, 8197–8200 CrossRef CAS PubMed;
(c) Y. Kuang, B. Shen, L. Dai, Q. Yao, X. Liu, L. Lin and X. Feng, Diastereodivergent asymmetric Michael-alkylation reactions using chiral N,N′-dioxide/metal complexes, Chem. Sci., 2018, 9, 688–692 RSC;
(d) H. Mei, G. Pan, X. Zhang, L. Lin, X. Liu and X. Feng, Catalytic asymmetric ring-opening/cyclopropanation of cyclic sulfur ylides: construction of sulfur-containing spirocyclopropyloxindoles with three vicinal stereocenters, Org. Lett., 2018, 20, 7794–7797 CrossRef CAS PubMed;
(e) L. Wang, W. Cao, H. Mei, L. Hu and X. Feng, Catalytic asymmetric synthesis of chiral spiro-cyclopropyl oxindoles from 3-alkenyl-oxindoles and sulfoxonium ylides, Adv. Synth. Catal., 2018, 360, 4089–4093 CrossRef CAS;
(f) X. Dou, W. Yao and Y. Lu, Asymmetric synthesis of 3-spirocyclopropyl-2-oxindoles via intramolecular trapping of chiral aza-ortho-xylylene, Chem. Commun., 2013, 49, 9224–9226 RSC;
(g) B. L. Zhao and D. M. Du, Chiral Squaramide-catalyzed Michael/alkylation
cascade reaction for the asymmetric synthesis of nitro-spirocyclopropanes, Eur. J. Org. Chem., 2015, 5350–5359 CrossRef CAS;
(h) A. Noole, N. S. Sucman, M. A. Kabeshov, T. Kanger, F. Z. Macaev and A. V. Malkov, Highly enantio- and diastereoselective generation of two quaternary centers in spirocyclopropanation of oxindole derivatives, Chem. – Eur. J., 2012, 18, 14929–14933 CrossRef CAS PubMed;
(i) T. Chen, Y. Zheng, Z. Fu and W. Huang, Cyclopropanation of fluorinated sulfur ylides with 1-azadienes: facile synthesis of CF3-substituted spiro scaffolds, Asian J. Org. Chem., 2019, 8, 2175–2179 CrossRef CAS;
(j) H. Wang, Y. Li, L. Huang, H. Xu, Y. Jiao, Z. Zhou, Z. Tang, F. Fang, X. Zhang, K. Ding, W. Yi, H. Liu, X. Wu and Y. Zhou, The synthesis of spirocyclopropane skeletons enabled by Rh(III)-catalyzed enantioselective C-H activation/[4 + 2] annulation, Chem. Catal., 2023, 3, 100822 CrossRef CAS;
(k) Y. Gong, X. C. Xu, Z. X. Yang and Y. L. Zhao, Palladium/copper-cocatalyzed three-component tandem cyclization of o-alkenyl arylisocyanides and sulfur ylides with alcohols: direct synthesis of spiro 3,3′-cyclopropyl oxindoles, Org. Chem. Front., 2024, 11, 3369–3375 RSC;
(l) J. H. Zhang, W. J. Yang, N. Li, Y. Tian, M. S. Xie and H. M. Guo, Reversal of enantioselectivity in cobalt(II)-catalyzed asymmetric Michael–alkylation reactions: synthesis of spiro-cyclopropane-oxindoles, Org. Chem. Front., 2024, 11, 4007–4013 RSC;
(m) N. Huang, L. Zou and Y. Peng, Enantioselective 1,3-dipolar cycloaddition of methyleneindolinones with α-diazomethylphosphonate to access chiral spirophosphonylpyrazoline-oxindoles catalyzed by tertiary amine thiourea and 1,5-diazabicyclo [4.3.0] non-5-ene, Org. Lett., 2017, 19, 5806–5809 CrossRef CAS PubMed.
- For selected examples of the asymmetric synthesis of bispirocyclopropane:
(a) J. R. Zhang, H. S. Jin, J. Sun, J. Wang and L. M. Zhao, Time-economical synthesis of bis-spiro cyclopropanes via cascade 1,6-conjugate addition/dearomatization reaction of para-quinone methides with 3-chlorooxindoles, Eur. J. Org. Chem., 2020, 4988–4994 CrossRef CAS;
(b) T. R. Penjarla, A. K. Shukla, R. Hazra, D. Roy, M. Kundarapu, M. Dixit and A. Bhattacharya, Copper acetate catalysed C–C bond formation en route to the synthesis of spiro indanedione cyclopropylpyrazolones, Org. Biomol. Chem., 2022, 20, 3779–3784 RSC;
(c) A. Noole, M. Oseka, T. Pehk, M. Oeren, I. Jarving, M. R. J. Elsegood, A. V. Malkov, M. Lopp and T. Kanger, 3-Chlorooxindoles: versatile starting materials for asymmetric organocatalytic synthesis of spirooxindoles, Adv. Synth. Catal., 2013, 355, 829–835 CrossRef CAS;
(d) S. Wang, Z. Guo, Y. Wu, W. Liu, X. Liu, S. Zhang and C. Sheng, Organocatalytic asymmetric synthesis of highly functionalized spiro-thiazolone–cyclopropaneoxindoles bearing two vicinal spiro quaternary centers, Org. Chem. Front., 2019, 6, 1442–1447 RSC;
(e) J. H. Li, T. F. Feng and D. M. Du, Construction of spirocyclopropane-linked heterocycles containing both pyrazolones and oxindoles through Michael/alkylation cascade reactions, J. Org. Chem., 2015, 80, 11369–11377 CrossRef CAS PubMed;
(f) Y. X. Song and D. M. Du, Asymmetric synthesis of spirooxindole-fused spirothiazolones via squaramide-catalysed reaction of 3-chlorooxindoles with 5-alkenyl thiazolones, Org. Biomol. Chem., 2019, 17, 5375–5380 RSC;
(g) Z. F. Hao, S. J. Zhu, Y. J. Hao, W. H. Zhang, Y. P. Tian and C. W. Lei, Enantioselective synthesis of bispiro[indanedione-oxindole-cyclopropane]s through organocatalytic [2+1] cycloaddition, J. Org. Chem., 2023, 88, 7641–7650 CrossRef CAS PubMed;
(h) D. Lu, X. Liu, J. H. Wu, S. Zhang, J. P. Tan, X. Yu and T. Wang, Asymmetric construction of bispiro-cyclopropane-pyrazolones via a [2+1] cyclization reaction by dipeptide-based
phosphonium salt catalysis, Adv. Synth. Catal., 2020, 362, 1966–1971 CrossRef CAS.
- For selected reviews on phase transfer catalysts, see:
(a)
E. V. Dehmlow and S. S. Dehmlow, Phase Transfer Catalysis, Wiley-VCH, Weinheim, 3rd edn, 1993 Search PubMed;
(b) A. Nelson, Asymmetric phase-transfer catalysis, Angew. Chem., Int. Ed., 1999, 38, 1583–1585 CrossRef CAS;
(c) M. J. O'Donnell, The enantioselective synthesis of α-amino acids by phase-transfer catalysis with achiral schiff base esters, Acc. Chem. Res., 2004, 37, 506–517 CrossRef PubMed;
(d) B. Lygo and B. I. Andrews, Asymmetric phase-transfer catalysis utilizing chiral quaternary ammonium salts: asymmetric alkylation of glycine imines, Acc. Chem. Res., 2004, 37, 518–525 CrossRef CAS PubMed;
(e) T. Ooi and K. Maruoka, Recent advances in asymmetric phase-transfer catalysis, Angew. Chem., Int. Ed., 2007, 46, 4222–4266 CrossRef CAS PubMed;
(f) T. Hashimoto and K. Maruoka, Recent development and application of chiral phase-transfer catalysts, Chem. Rev., 2007, 107, 5656–5682 CrossRef CAS PubMed;
(g) S. S. Jew and H. G. Park, Cinchona-based phase-transfer catalysts for asymmetric synthesis, Chem. Commun., 2009, 7090–7103 RSC.
- For selected reviews on asymmetric phosphonium salt catalysis, see:
(a) T. Werner, Phosphonium salt organocatalysis, Adv. Synth. Catal., 2009, 351, 1469–1481 CrossRef CAS;
(b) S. Liu, Y. Kumatabara and S. Shirakawa, Chiral quaternary phosphonium salts as phase-transfer catalysts for environmentally benign asymmetric transformations, Green Chem., 2016, 18, 331–341 RSC;
(c) A. Golandaj, A. Ahmad and D. Ramjugernath, Phosphonium salts in asymmetric catalysis: a journey in a decade's extensive research work, Adv. Synth. Catal., 2017, 359, 3676–3706 CrossRef CAS;
(d) M. Selva, M. Noè, A. Perosa and M. Gottardo, Carbonate, acetate and phenolate phosphonium salts as catalysts in transesterification reactions for the synthesis of non-symmetric dialkyl carbonates, Org. Biomol. Chem., 2012, 10, 6569–6578 RSC;
(e) S. Fang, Z. Liu and T. Wang, Design and application of peptide-mimic phosphonium salt catalysts in asymmetric synthesis, Angew. Chem., Int. Ed., 2023, 62, e202307258 CrossRef CAS PubMed;
(f) L. Chen, Y. Deng, T. Li, D. Hu, X. Ren and T. Wang, Asymmetric nucleophilic additions promoted by quaternary phosphonium ion-pair catalysts, CCS Chem., 2024, 6, 2110–2130 CrossRef.
- For selected examples, see:
(a) J. Pan, J. H. Wu, H. Zhang, X. Ren, J. P. Tan, L. Zhu, H. S. Zhang, C. Jiang and T. Wang, Highly enantioselective synthesis of fused tri- and tetrasubstituted aziridines: aza-Darzens reaction of cyclic imines with α-halogenated ketones catalyzed by bifunctional phosphonium salt, Angew. Chem., Int. Ed., 2019, 58, 7425–7430 CrossRef CAS PubMed;
(b) L. Zhu, X. Ren, Z. Liao, J. Pan, C. Jiang and T. Wang, Asymmetric three-component cyclizations toward structurally spiro pyrrolidines via bifunctional phosphonium salt catalysis, Org. Lett., 2019, 21, 8667–8672 CrossRef CAS PubMed;
(c) S. Zhang, X. Yu, J. Pan, C. Jiang, H. S. Zhang and T. Wang, Asymmetric synthesis of spiro-structural 2,3-dihydrobenzofurans via the bifunctional phosphonium salt-promoted [4 + 1] cyclization of ortho-quinone methides with α-bromoketones, Org. Chem. Front., 2019, 6, 3799–3803 RSC;
(d) H. Zhang, J. He, Y. Chen, C. Zhuang, C. Jiang, K. Xiao, Z. Su, X. Ren and T. Wang, Regio- and stereoselective cascade of β,γ-unsaturated ketones by dipeptide phosphonium salt catalysis: stereospecific construction of dihydrofuro-fused [2,3-b] skeletons, Angew. Chem., Int. Ed., 2021, 60, 19860–19870 CrossRef CAS PubMed;
(e) J. P. Tan, K. Li, B. Shen, C. Zhuang, Z. Liu, K. Xiao, P. Yu, B. Yi, X. Ren and T. Wang, Asymmetric synthesis
of N-bridged [3.3.1] ring systems by phosphonium salt/Lewis acid relay catalysis, Nat. Commun., 2022, 13, 357 CrossRef CAS PubMed;
(f) Y. Chen, J. He, C. Zhuang, Z. Liu, K. Xiao, Z. Su, X. Ren and T. Wang, Synergistic catalysis between a dipeptide phosphonium salt and a metal-based lewis acid for asymmetric synthesis of N-bridged [3.2.1] ring systems, Angew. Chem., Int. Ed., 2022, 61, e202207334 CrossRef CAS PubMed;
(g) J. H. Wu, J. P. Tan, J. Y. Zheng, J. He, Z. Song, Z. Su and T. Wang, Towards axially chiral pyrazole-based phosphorus scaffolds by dipeptide-phosphonium salt catalysis, Angew. Chem., Int. Ed., 2023, 62, e202215720 CrossRef CAS PubMed;
(h) Z. Liu, S. Fang, H. Li, C. Xiao, K. Xiao, Z. Su and T. Wang, Organocatalytic skeletal reorganization for enantioselective synthesis of S-stereogenic sulfinamides, Nat. Commun., 2024, 15, 4348 CrossRef CAS PubMed.
|
This journal is © the Partner Organisations 2024 |
Click here to see how this site uses Cookies. View our privacy policy here.